Vous êtes sur la page 1sur 429

Advances in

ATOMIC, MOLECULAR, AND OPTICAL PHYSICS

VOLUME 48
Editors
BENJAMIN BEDERSON
New York University
New York, New York
HERBERT WALTHER
Max-Plank-Institut j~ir Quantenoptik
Garching bei Miinchen
Germany

Editorial Board
ER. BERMAN
University of Michigan
Ann Arbor, Michigan
M. GAVRILA
EO.M. Instituut voor Atoom- en Molecuulfysica
Amsterdam, The Netherlands
M. INOKUTI
Argonne National Laboratory
Argonne, Illinois
CHUN C. LIN
University of Wisconsin
Madison, Wisconsin

Founding Editor
SIR DAVID BATES

Supplements
1. Atoms in Intense Laser Fields, Mihai Gavrila, Ed.
2. Cavity Quantum Electrodynamics, Paul R. Berman, Ed.
3. Cross Section Data, Mitio Inokuti, Ed.
A D VANCES IN

ATOMIC
MOLECOLAR
AND OPTICAL
PHYSICS

Edited by

Benjamin Bederson
DEPARTMENT OF PHYSICS
NEW YORK UNIVERSITY
NEW YORK, NEW YORK

Herbert Walther
UNIVERSITY OF MUNICH AND
MAX-PLANK INSTITUT FOR QUANTENOPTIK
MUNICH, GERMANY

Volume 4 8

ACADEMIC PRESS
An imprint of Elsevier Science

A m s t e r d a m . Boston- London. New York- Oxford. Paris


San D i e g o . San Francisco. Singapore. S y d n e y - T o k y o
This book is printed on acid-free paper.

9 2002 Elsevier Science (USA)


All rights reserved.

No part of this publication may be reproduced or transmitted in any form or by any means,
electronic or mechanical, including photocopy, recording, or any information storage and
retrieval system, without permission in writing from the Publisher.

The appearance of the code at the bottom of the first page of a chapter in this book
indicates the Publisher's consent that copies of the chapter may be made for personal
or internal use of specific clients. This consent is given on the condition, however, that
the copier pay the stated per copy fee through the Copyright Clearance Center, Inc.
(222 Rosewood Drive, Danvers, Massachusetts 01923), for copying beyond that permitted
by Sections 107 or 108 of the U.S. Copyright Law. This consent does not extend to other
kinds of copying, such as copying for general distribution, for advertising or promotional
purposes, for creating new collective works, or for resale. Copy fees for pre-2002 chapters
are as shown on the title pages. If no fee code appears on the title page, the copy fee is
the same as for current chapters.
1049-250X/02 $35.00

Explicit permission from Academic Press is not required to reproduce a maximum of


two figures or tables from an Academic Press chapter in another scientific or research
publication provided that the material has not been credited to another source and that
full credit to the Academic Press chapter is given.

A c a d e m i c Press
An Elsevier Science Imprint
525 B Street, Suite 1900, San Diego, California 92101-4495, USA
http://www.academicpress.com

International Standard Book Number: 0-12-003848-X

Printed and bound in Great Britain by MPG Books Ltd, Cornwall, UK

02 03 04 05 06 07 MP 9 8 7 6 5 4 3 2 1
Contents

CONTRIBUTORS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

Multiple Ionization in Strong Laser Fields


R. D6rner, Th. Weber, M. Weckenbrock, A. Staudte, M. Hattass,
R. Moshammer, J. Ullrich, H. Schmidt-B6cking
Io I n t r o d u c t i o n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
II. C O L T R I M S - A C l o u d C h a m b e r for A t o m i c P h y s i c s . . . . . . . 3
III. Single I o n i z a t i o n and the T w o - s t e p M o d e l . . . . . . . . . . . . . . . 6
IV. Mechanisms of Double Ionization ..................... 9
V. Recoil Ion M o m e n t a . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
VI. Electron Energies ................................ 19
VII. Correlated Electron Momenta ........................ 20
VIII. Outlook ...................................... 30
IX. Acknowledgments ............................... 30
X. References .................................... 31

Above-Threshold Ionization" From Classical Features to


Quantum Effects
W. Becker, E Grasbon, R. Kopold, D.B. MilodeviO, G.G. Paulus and
H. Walther
I~ I n t r o d u c t i o n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
II. Direct I o n i z a t i o n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
III. R e s c a t t e r i n g : T h e Classical T h e o r y . . . . . . . . . . . . . . . . . . . . 50
IV. Rescattering: Quantum-mechanical Description . . . . . . . . . . . . 53
V. ATI in the Relativistic R e g i m e . . . . . . . . . . . . . . . . . . . . . . . 73
VI. Q u a n t u m Orbits in H i g h - o r d e r H a r m o n i c G e n e r a t i o n . . . . . . . . 76
VII. A p p l i c a t i o n s o f ATI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
VIII. Acknowledgments ............................... 92
IX. References .................................... 92

Dark Optical Traps for Cold Atoms

Nir Friedman, Ariel Kaplan and Nir Davidson


I. I n t r o d u c t i o n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
II. B a c k g r o u n d . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
III. M u l t i p l e - L a s e r - B e a m s D a r k O p t i c a l Traps ............... 106
vi Contents

IV. S i n g l e - B e a m D a r k Optical Traps . . . . . . . . . . . . . . . . . . . . . 113


V. Applications ................................... 127
VI. Conclusions .................................... 147
VII. References .................................... 148

Manipulation of Cold Atoms in Hollow Laser Beams

Heung-Ryoul Noh, Xinye Xu and Wonho Jhe


I. Introduction ..................................... 153
II. T h e o r e t i c a l M o d e l s for C o l d A t o m s in H o l l o w L a s e r B e a m s .... 154
III. G e n e r a t i o n M e t h o d s for H o l l o w L a s e r B e a m s . . . . . . . . . . . . . . 160
IV. C o l d A t o m M a n i p u l a t i o n in H o l l o w L a s e r B e a m s . . . . . . . . . . . 170
V. Acknowledgment ................................. 188
VI. References ...................................... 188

Continuous Stern-Gerlach Effect on Atomic Ions

Giinther Werth, Hartmut Hdffner and Wolfgang Quint


I. I n t r o d u c t i o n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
II. A Single Ion in a P e n n i n g Trap . . . . . . . . . . . . . . . . . . . . . . 195
IIl. C o n t i n u o u s S t e r n - G e r l a c h Effect . . . . . . . . . . . . . . . . . . . . . 206
IV. Double-Trap Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
V. C o r r e c t i o n s and S y s t e m a t i c Line Shifts . . . . . . . . . . . . . . . . . 212
VI. Conclusions .................................... 213
VII. Outlook ...................................... 214
VIII. Acknowledgements ............................... 216
IX. References .................................... 216

The Chirality of Biomolecules

Robert N. Compton and Richard M. Pagni


I. I n t r o d u c t i o n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
II.F u n d a m e n t a l N a t u r e o f Chirality . . . . . . . . . . . . . . . . . . . . . . 219
III.True and False Chirality . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
IV. G a l a x i e s , Plants, and P h a r m a c e u t i c a l s . . . . . . . . . . . . . . . . . . 233
V. Plausible O r i g i n s o f H o m o c h i r a l i t y . . . . . . . . . . . . . . . . . . . . 236
VI. A s y m m e t r y in B e t a R a d i o l y s i s . . . . . . . . . . . . . . . . . . . . . . . 243
VII. Possible Effects o f the P a r i t y - V i o l a t i n g E n e r g y D i f f e r e n c e ( P V E D )
in E x t e n d e d M o l e c u l a r S y s t e m s . . . . . . . . . . . . . . . . . . . . . . 252
VIII. C o n c l u s i o n s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
IX. A c k n o w l e d g m e n t . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
X. R e f e r e n c e s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
Contents vii

Microscopic Atom Optics" From Wires to an Atom Chip

Ron Folman, Peter Kriiger, J6rg Schmiedmayer, Johannes Denschlag


and Carsten Henkel
I. Introduction .................................... 263
II. Designing Microscopic Atom Optics ................... 265
III. E x p e r i m e n t s with F r e e - S t a n d i n g S t r u c t u r e s . . . . . . . . . . . . . . . 292
IV. S u r f a c e - M o u n t e d Structures: T h e A t o m Chip . . . . . . . . . . . . . 303
V. Loss, H e a t i n g and D e c o h e r e n c e . . . . . . . . . . . . . . . . . . . . . . 324
VI. Vision and O u t l o o k . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
VII. Conclusion .................................... 351
VIII. Acknowledgement ............................... 351
IX. References .................................... 352

Methods of Measuring Electron-Atom Collision Cross Sections with


an Atom Trap

R.S. Schappe, M.L. Keeler, T.A. Zimmerman, M. Larsen, P. Feng,


R.C. Nesnidal, J.B. Boffard, T.G. Walker, L. W. Anderson and C.C. Lin
I. Introduction .................................... 357
II. General Experiment Overview ........................ 359
III. M e t h o d s for M e a s u r i n g C r o s s S e c t i o n s . . . . . . . . . . . . . . . . . . 367
IV. Conclusions .................................... 386
V. Acknowledgments ................................ 387
VI. A p p e n d i x . N u m e r i c a l M o d e l for R e s i d u a l Polarization ....... 387
VII. References ..................................... 389

INDEX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 391

CONTENTS OF VOLUMES IN THIS SERIAL . . . . . . . . . . . . . . . . . . . . . . . 405


This Page Intentionally Left Blank
Contributors

Numbers in parentheses indicate the pages on which the authors' contributions begin.
L.W. ANDERSON(357), Department of Physics, University of Wisconsin, Madi-
son, Wisconsin 53706
W. BECI~R (35), Max-Born-Institut, Max-Born-Str. 2A, 12489 Berlin, Germany
J.B. BOFFARD(357), Department of Physics, University of Wisconsin, Madison,
Wisconsin 53706
ROBERT N. COMPTON (219), Department of Chemistry, and Department of
Physics, University of Tennessee, Knoxville, Tennessee 37996
NIR DAVIDSON(99), Weizmann Institute of Science, Department of Physics of
Complex Systems, Rehovot, Israel
JOHANNES DENSCHLAG(263), Institut f'tir Experimentalphysik, Universitfit Inns-
bruck, 6020 Innsbruck, Austria
R. DORNER (1), Institut f'tir Kernphysik, August Euler Str. 6, 60486 Frankfurt,
Germany
P. FEN6 (357), Department of Physics, University of St. Thomas, St. Paul,
Minnesota 55105
RON FOEMAN (263), Physikalisches Institut, Universitfit Heidelberg, 69120
Heidelberg, Germany
NIR FRIEDMAN(99), Weizmann Institute of Science, Department of Physics of
Complex Systems, Rehovot, Israel
E G~SBON (35), Max-Planck-Institut f'tir Quantenoptik, Hans-Kopfermann-Str. 1,
85748 Garching, Germany
HARTMUT H~FFNER (191), Johannes Gutenberg University, Department of
Physics, 55099 Mainz, Germany
M. HATTASS(1), Institut f'tir Kernphysik, August Euler Str. 6, 60486 Frankfurt,
Germany
CARSTENHENKEL(263), Institut ffir Physik, Universit~it Potsdam, 14469 Potsdam,
Germany
WONHO JIqE (153), School of Physics and Center for Near-field Atom-photon
Technology, Seoul National University, Seoul 151-742, South Korea
x Contributors

ARIEL KAPLAN (99), Weizmann Institute of Science, Department of Physics of


Complex Systems, Rehovot, Israel
M.L. KELLER(357), Department of Physics, University of Wisconsin, Superior,
Wisconsin 54880
R. KOPOLD(35), Max-Born-Institut, Max-Born-Str. 2A, 12489 Berlin, Germany
PETER KRUGER (263), Physikalisches Institut, Universitfit Heidelberg, 69120
Heidelberg, Germany
M. LARSEN (357), Department of Physics, University of Wisconsin, Madison,
Wisconsin 53706
C.C. LIN (357), Department of Physics, University of Wisconsin, Madison,
Wisconsin 53706
D.B. MILO~EVI~(35), Faculty of Science, University of Sarajevo, Zmaja od Bosne
35, 71000 Sarajevo, Bosnia and Hercegovina
R. MOSHAMMER(1), Max-Planck-Institut ftir Kernphysik, Saupfercheckweg 1,
69117 Heidelberg, Germany
R.C. NESNIDAL(357), New Focus, Inc., Middleton, Wisconsin 53562
HEUNG-Ru NOH (153), School of Physics and Center for Near-field Atom-
photon Technology, Seoul National University, Seoul 151-742, South
Korea
RICHARD M. PAGNI (219), Department of Chemistry, University of Tennessee,
Knoxville, Tennessee 37996
G.G. PAULUS(35), Max-Planck-Institut ftir Quantenoptik, Hans-Kopfermann-Str.
1, 85748 Garching, Germany
WOLFGANG QUINT (191), Gesellschaft f'tir Schwerionenforschung, 64291 Darm-
stadt, Germany
R.S. SCHAPPE(357), Department of Physics, Lake Forest College, Lake Forest,
Illinois 60045
H. SCHMIDT-BOCKING(1), Institut ffir Kernphysik, August Euler Str. 6, 60486
Frankfurt, Germany
J6RG SCHMIEDMAYER(263), Physikalisches Institut, Universitfit Heidelberg,
69120 Heidelberg, Germany
A. STAUDTE(1), Institut for Kernphysik, August Euler Str. 6, 60486 Frankfurt,
Germany
J. ULLRICH(1), Max-Planck-Institut for Kernphysik, Saupfercheckweg 1, 69117
Heidelberg, Germany
Contributors xi

T.G. WALKER(357), Department of Physics, University of Wisconsin, Madison,


Wisconsin 53706
H. WALXHER(35), Ludwig-Maximilians-Universit~it Mfinchen, Germany
TH. WEBER (1), Institut for Kernphysik, August Euler Str. 6, 60486 Frankfurt,
Germany
M. WECKENBROCK(1), Institut fiir Kernphysik, August Euler Str. 6, 60486
Frankfurt, Germany
GONTHERWERTH(191), Johannes Gutenberg University, Department of Physics,
55099 Mainz, Germany
XINYE Xu (153), School of Physics and Center for Near-field Atom-photon
Technology, Seoul National University, Seoul 151-742, South Korea
T.A. ZIMMERMAN (357), Department of Physics, University of Wisconsin,
Madison, Wisconsin 53706
This Page Intentionally Left Blank
A D V A N C E S IN A T O M I C , M O L E C U L A R , A N D O P T I C A L P H Y S I C S , VOL. 48

M UL TIPL E IONIZATION IN
S TR ONG LASER FIELD S
R. DORNER*, Th. WEBER, M. WECKENBROCK, A. STAUDTE,
M. HATTASS and H. SCHMIDT-BOCKING
Institut fiir Kernphysik, August Euler Str. 6, 60486 Frankfurt, Germany

R. MOSHAMMER and J. ULLRICH


Max-Planck-Institut fiir Kernphysik, Saupfercheckweg 1, 69117 Heidelberg, Germany

I. I n t r o d u c t i o n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
II. C O L T R I M S - A C l o u d C h a m b e r for A t o m i c Physics ................... 3
III. Single I o n i z a t i o n and the Two-step M o d e l . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
IV. M e c h a n i s m s o f D o u b l e Ionization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
V. Recoil Ion M o m e n t a . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
A. F r o m N o n s e q u e n t i a l to Sequential D o u b l e I o n i z a t i o n ................. 11
B. The Origin o f the Double Peak Structure . . . . . . . . . . . . . . . . . . . . . . . . . . 13
VI. Electron E n e r g i e s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
VII. C o r r e l a t e d Electron M o m e n t a . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
A. E x p e r i m e n t a l Findings ...................................... 20
B. C o m p a r i s o n to S i n g l e - P h o t o n and C h a r g e d Particle I m p a c t Double Ionization . 23
C. Interpretation within the R e s c a t t e r i n g M o d e l ....................... 25
D. S-Matrix C a l c u l a t i o n s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
E. T i m e - d e p e n d e n t Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
VIII. Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
IX. A c k n o w l e d g m e n t s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
X. R e f e r e n c e s ................................................ 31

I. I n t r o d u c t i o n

70 years ago Maria G6ppert-Mayer [ 1] showed that the energy of many photons
can be combined to achieve ionization in cases where the energy of one photon is
not sufficient to overcome the binding. Modern short-pulse Ti:Sa lasers (800 nm,
1.5 eV) routinely provide intensities of more than 1016 W/cm 2 and pulses shorter
than 100 femtoseconds. Under these conditions the ionization probability of most
atoms is close to unity. 1016 W/cm 2 corresponds to about 10 l~ coherent photons
in a box of the size of the wavelength (800nm). This extreme photon density

* E-mail: d o e r n e r @ h s b . u n i - f r a n k f u r t . d e

Copyright 9 2002 Elsevier Science (USA)


All rights reserved
ISBN 0-12-003848-X/ISSN 1049-250X/02 $35.00
2 R. D6rner et al. [I

allows highly nonlinear multiphoton processes such as multiple ionization, where


typically more than 50 photons can be absorbed from the laser field.
Such densities of coherent photons in the laser pulse also suggests a change
from the "photon perspective" to the "field perspective": The laser field can be
described as a classical electromagnetic field, neglecting the quantum nature of
the photons. From this point of view the relevant quantities are the field strength
and its frequency. 1016 W/cm 2 at 800 nm corresponds to a field of 3 x 1011 V/m,
comparable to the field experienced by the electron in a Bohr orbit in atomic
hydrogen (5 • 1011 V/m).
Single ionization in such strong fields has been intensively studied for many
years now. The experimental observables are the ionization rates as function of
the laser intensity and wavelength, the electron energy and angular distribution
as well as the emission of higher harmonic light. We refer the reader to
several review articles covering this broad field[2-4]. Also the generation
of femtosecond laser pulses has been described in a number of detailed
reviews [5-8].
The present article focuses on some recent advances in unveiling the
mechanism of double and multiple ionization in strong fields. Since more
particles are involved, the number of observables and the challenge to the
experimental as well as to the theoretical techniques increases. Early studies
measured the rate of multiply charged ions as a function of laser intensity.
The work reviewed here employs mainly COLTRIMS (Cold Target Recoil Ion
Momentum Spectroscopy)[9] to detect not only the charge state but also the
momentum vector of the ion and of one of the electrons in coincidence. Today
such highly differential measurements are standard in the fields of ion-atom,
electron-atom and high-energy single-photon-atom collision studies.
The main question discussed in the context of strong fields as well as in the
above-mentioned areas of current research is the role of electron correlation
in the multiple ionization process. Do the electrons escape from the atom
"sequentially" or "nonsequentially," i.e. does each electron absorb the photons
independently, or does one electron absorb the energy from the field and then
share it with the second electron via electron-electron correlation?
Despite its long history the underlying question of the dynamics of electron
correlation is still one of the fundamental puzzles in quantum physics. Its
importance lies not only in the intellectual challenge of the few-body problem,
but also in its wide-ranging impact to many fields of science and technology.
It is the correlated motion of electrons that is responsible for the structure
and the evolution of large parts of our macroscopic world. It drives chemical
reactions, it is the ultimate reason for superconductivity and many other effects
in the condensed phase. In atomic processes few-body correlation effects can
be studied in a particularly clear manner. This, for example, was the motivation
for studying theoretically and experimentally the question of double ionization
II] MULTIPLE IONIZATION IN STRONG LASER FIELDS 3

by charged-particle (see ref. [10] for a review) or single-photon [1 l, 12] impact


in great detail. As soon as lasers became strong enough to eject two or more
electrons from an atom, electron correlation in strong light fields became subject
of increased attention, too. As we will show below, in comparison with some
of the latest results on double ionization by ion and single-photon impact, the
laser field generates new correlation mechanisms, thereby raising more exciting
new questions than settling old ones.

II. C O L T R I M S - A Cloud Chamber for Atomic Physics

For a long time the experimental study of electron correlation in ionization


processes of atoms, molecules and solids has suffered from the technical
challenge to observe more than one electron emerging from a multiple ionization
event. The main problem lies in performing coincidence studies employing
conventional electron spectrometers, which usually cover only a small part
of the total solid angle. COLTRIMS (Cold Target Recoil Ion Momentum
Spectroscopy) is an imaging technique that solves this fundamental problem
in atomic and molecular coincidence experiments. Like the cloud chamber and
its modern successors in nuclear and high-energy physics, it delivers complete
images of the momentum vectors of all charged fragments from an atomic or
molecular fragmentation process. The key feature of this technique is to provide
a 4:r collection solid angle for low-energy electrons (up to a few hundred eV) in
combination with 4:r solid angle and high resolution for the coincident imaging
of the ion momenta.
As we will show below, the ion momenta in most atomic reactions with
photons or charged particles are of the same order of magnitude as the electron
momenta. Due to their mass, however, this corresponds to ion energies in
the range of ~teV to meV. These energies are below thermal motion at room
temperature. Thus, the atoms have to be cooled substantially before the reaction.
In the experiments discussed here this is achieved by using a supersonic gas jet
as a target. More recently, atoms in magneto-optical traps have been used to
further increase the resolution [ 13-16].
A typical setup as used for the experiments discussed here is shown in
Fig. 1. The laser pulse is focused by a lens of 5 cm focal length or a parabolic
mirror into a supersonic gas jet providing target atoms with very small initial
momentum spread of under 0.1 au (atomic units are used throughout this chapter)
in the direction of the laser polarization (along the z-axis in Fig. 1). For
experiments in ion-atom collisions or with synchrotron radiation the ionization
probability is very small: That is why one aims at a target density in the range
of up to 10-4 mbar local pressure in the gas jet. Accordingly, a background
pressure in the chamber in the range of 10-8 mbar is sufficient. In contrast, for
multiple ionization by femtosecond laser pulses the single ionization probability
4 R. D 6 r n e r et al. [II

FIG. 1. Experimental setup. Electrons and ions are created in the supersonic gas-jet target. The thin
copper rings create a homogeneous electric field and the large Helmholtz coils an additional magnetic
field. These fields guide the charged particles onto fast time- and position-sensitive channel plate
detectors (Roentdek, www.roentdek.com). The time-of-flight (TOF) and the position of impact of
each electron-ion pair is recorded in list mode. From this the three-dimensional momentum vector
of each particle can be calculated.

easily reaches unity. Thus, within the reaction volume defined by the laser focus
of typically (10~m) 2 • 100~tm all atoms are ionized. Since for coincidence
experiments it is essential that much less than one atom is ionized per laser
shot, a background pressure of less than 10 -1~ mbar is required. The gas jet
has to be adjusted accordingly to reach single-collision conditions at the desired
laser peak power. With standard supersonic gas jets this can only be achieved
by tightly skimming the atomic beam, since a lower driving pressure for the
expansion would result in an increase of the internal temperature of the jet
along its direction of propagation. Single ionization (see Sect. III) allows for an
efficient monitoring of the resolution as well as on-line control of single-collision
conditions.
The ions created in the laser focus are guided by a weak electric field towards
a position-sensitive channel plate detector. From the position of impact and the
time-of-flight (TOF) of the ion all three components of the momentum vector and
the charge state are obtained. A typical ion TOF spectrum from the experiment
reported in ref. [17] is shown in Fig. 2.
The electric field also guides the electrons towards a second position-sensitive
channel plate detector. To collect electrons with large energies transverse to
the electric field a homogeneous magnetic field is superimposed parallel to
the electric field. This guides the electrons on cyclotron trajectories towards
the detector. Depending on their time-of-flight the electrons perform several
II] MULTIPLE IONIZATION IN STRONG LASER FIELDS 5

FIG. 2. Time-of-flight distribution of ions produced by a 6.6 x 1014 W/cm 2 laser pulse. The gas
target was 3He; the residual gas pressure in the chamber was about 2 x 10-l~ mbar. The double peak
structure in the 3He2+ peak can be seen. The total count rate was about 0.1 ion per laser shot.

FIG. 3. Horizontal axis: Electron time-of-flight. Vertical axis: radial distance from a central
trajectory with zero transverse momentum on electron detector, see text.

full turns on their w a y to the detector. F i g u r e 3 s h o w s the e l e c t r o n T O F


versus the radial distance o f the p o s i t i o n f r o m a central t r a j e c t o r y w i t h z e r o
transverse m o m e n t u m o f the electron. W h e n the T O F is an i n t e g e r m u l t i p l e o f the
c y c l o t r o n f r e q u e n c y the electrons hit the d e t e c t o r at this position, i n d e p e n d e n t l y
o f their m o m e n t u m t r a n s v e r s e to the field. T h e s e T O F s r e p r e s e n t p o i n t s in
p h a s e space w h e r e the s p e c t r o m e t e r has no r e s o l u t i o n in the t r a n s v e r s e direction.
6 R. D6rner et al. [III

For all other TOFs the initial momentum can be uniquely calculated from the
measured positions of impact and the TOE Using a magnetic field of 10 Gauss,
4:r solid angle collection is achieved for electrons up to about 30 eV. The typical
detection probability of an electron is in the range o f 3 0 - 4 0 % . Thus, even for
double ionization in most cases only one electron is detected. The positions o f
impact and the times-of-flight are stored for each event in list mode. Thus the
whole experiment can be replayed in the off-line analysis. A detailed description
of the integrated multi-electron-ion m o m e n t u m spectrometer can be found in
ref. [ 18].

III. Single Ionization and the Two-step Model


The m o m e n t u m distribution o f singly charged helium ions produced by absorp-
tion of one 85-eV photon (synchrotron radiation) and by multiphoton absorption
at 800 nm and 1.5 • 1015 W/cm 2 is shown in Fig. 4. In both cases the m o m e n t u m
of the photon is negligible compared to the electron momentum. Therefore,
electron and He ~+ ion are essentially emitted back-to-back compensating each
others momentum (The exact kinematics including the photon m o m e n t u m can be
found in section 2.3.1 of ref. [9]). Hence, for single ionization the spectroscopy
of the ion momentum is equivalent to electron spectroscopy. This can be directly
confirmed by looking at the coincidence between the ions and electrons in Fig. 5.
All true coincidence events are located on the diagonal with equal momenta Pz
in the TOF direction. The width of this diagonal gives the combined resolution

FIG. 4. Momentum distribution of He l+ ions. Left: For 85 eV single-photon absorption. Right:


1.5 eV (800nm), 220 fs, 1.5• 1015W/cm2. The polarization vector of the light is horizontal. The
photon momentum is perpendicular to the (ky,kz) plane. In the left-hand panel the momentum
component in the third dimension out of the plane of the figure is restricted to • au. The right-hand
panel is integrated over the momenta in the direction out of the plane of the figure.
III] MULTIPLE IONIZATION IN S T R O N G L A S E R FIELDS 7

FIG. 5. Single ionization of argon by 3.8 • 1014W/cm2. The horizontal axis shows the momentum
component of the recoil ion parallel to the polarization. The vertical axis represents the momentum
of the coincident electron in the same direction. By momentum conservation all true coincidences
are located on the diagonal. Along the diagonal ATI peaks can be seen. The z-axis is plotted in
linear scale.

of the electron and ion m o m e n t u m measurement for the pz component (in this
case 0.25 au full width at half maximum). All events off the diagonal result from
false coincidences in which the electron and ion were created in the same pulse
but did not emerge from the same atom. This allows a continuous monitoring o f
the fraction o f false coincidences during the experiment. Knowing this number
the false coincidences can also be subtracted for double-ionization events.
For single-photon absorption the electron energy is uniquely determined by
the photon energy Ev and the binding energy plus a possible internal excitation
energy of the ion. The resulting narrow lines in the photoelectron energy
spectrum correspond to spheres in m o m e n t u m space. The left-hand panel of
Fig. 4 shows a slice through this m o m e n t u m sphere. The outer ring corresponds
to He 1+ ions in the ground state, the inner rings to the excited states. The
photons are linearly polarized with the polarization direction horizontal in the
figure. The angular distribution of the outer ring shows an almost pure dipole
distribution according to the absorption of one single photon. On the contrary,
in the laser field any number of photons can be absorbed, leading to an almost
continuous energy distribution o f the electrons (right-hand panel in Fig. 4).
Structure o f individual ATI (above threshold ionization) peaks spaced by the
8 R. D6rner et al. [III

photon energy (1.5 eV) is not seen here. This is in agreement with electron
spectra, where at comparable laser intensities ATI structure is not observed either.
The electrons and ions are emitted in narrow jets along the polarization axis.
Such high-angular-momentum states, needed to produce this kind of distribution,
are accessible due to the large number of photons absorbed.
How do the ions and electrons get their momenta? For the case of single-
photon absorption the light field is so weak that there is no acceleration. Also, the
photon carries no significant momentum into the reaction. The photon cuts the tie
between nucleus and electron by providing the energy. The momenta observed in
the final state thus have to be present already in the initial-state Compton profile
of the atom. Single-photon absorption is therefore linked to a particular fraction
of the initial-state wave function, which in momentum representation coincides
with the final-state momentum. The scaling of the photo ionization cross section
at high energies follows, besides a phase space factor, the initial-state momentum
space Compton profile, i.e. the probability to find an electron-ion pair with the
appropriate momentum in the initial state.
In the strong-field case the situation changes completely. The field is strong
enough to accelerate the ions and electrons substantially after the electron is set
free. The momentum balance, however, is still the same as in the single-photon
limit: The laser field accelerates electron and ion to the opposite directions
resulting again in their back-to-back emission (see Fig. 5). This changes only if
the laser pulse is long enough that the electron can escape from the focus during
the pulse. In that case, which we do not consider here, the momenta are balanced
by a huge amount of elastically scattered photons. In the regime of wavelength
and binding energies under consideration here, a simple two-step picture has
been proven useful. In the first step the electron is set free by tunneling through
the potential barrier created by the superposition of the Coulomb potential of
the atom and the electric field of the laser. This process promotes electrons and
ions with zero momentum to the continuum. Then they are accelerated in the
laser field and perform a quiver motion. In this model the net momentum in the
polarization direction, which is observed after a pulse with an envelope of the
electric field strength E(t) being long compared to the laser frequency, is purely
a function of the phase of the field at the instant of tunneling (tunneling time to):

PzHeI+(too) = ft0 t~ E(t) sin ~ot dt. (1)

Tunneling at the field maximum thus leads to electrons and ions with zero
momentum. The maximum momentum corresponding to the zero crossing of
the laser field is x/~Up, where Up = I/4~o 2 is the ponderomotive potential
at intensity I and photon frequency ~o (Up = 39.4eV at 6.6x1014W/cm2).
Within this simple model the ion and electron momentum detection provides
a measurement of the phase of the field at the instant of tunneling. We will
generalize this idea below for the case of double ionization.
IV] MULTIPLE IONIZATION IN STRONG LASER FIELDS 9

Single ionization is shown here mainly for illustration. Much more detailed
experiments have been reported using conventional TOF spectrometers (see ref.
[4] and references therein) and photoelectron imaging [19,20].

IV. M e c h a n i s m s of Double Ionization

What are the "mechanisms" leading to double ionization? This seemingly clear-
cut question does not necessarily have a quantum-mechanical answer. The word
"mechanism" mostly refers to an intuitive mechanistical picture. It is not always
clear how this intuition can be translated into theory, and even if one finds
such a translation the contributions from different mechanisms have to be added
coherently to obtain the measurable final state of the reaction [21,22]. Thus, only
in some cases mechanisms are experimentally accessible. This is only the case
if different mechanisms occur at different strengths of the perturbation (such
as laser power or projectile charge) or if they predominantly populate different
regions of the final-state phase space. In these cases situations can be found
where one mechanism dominates such that interference becomes negligible. With
these words of caution in mind, we list the most discussed mechanisms leading
to double ionization:
(1) TS2 or Sequential Ionization: Here the two electrons are emitted
sequentially by two independent interactions of the laser field with the atom.
From a photon perspective one could say that each of the electrons absorbs
photons independently. From the field perspective one would say that each
electron tunnels independently at different times during the laser pulse.
This is equivalent to the TS2 (two-step-two) mechanism in ion-atom and
electron-atom collisions. In this approximation the probability of the double
ejection can be estimated in an independent-particle model. Most simply one
calculates double ionization as two independent steps of single ionization.
A somewhat more refined approach uses an independent-event model, which
takes into account the different binding energies for the ejection of the
first and the second electron (see, e.g., ref. [23] for ion impact, ref. [24] for
laser impact).
(2) Shake-Off: If one electron is removed rapidly (sudden approximation) from
an atom or a molecule, the wave function of the remaining electron has to
relax to the new eigenstates of the altered potential. Parts of these states
are in the continuum, so that a second electron can be "shaken off" in
this relaxation process. This is known for example from beta decay, where
the nuclear charge is changed. Shake-off is also known to be one of the
mechanisms for double ionization by absorption or Compton scattering of
a single photon (see the discussion in ref. [25] and references therein).
However, only for very high photon energies (in the keV range) it is
the dominating mechanism. For helium it leads to a ratio of double to
10 R. D 6 r n e r et al. [IV

single ionization of 1.66% [26,27] for photoabsorption (emission of the


first electron from close to the nucleus) and 0.86% for Compton scattering
(averaged over the initial-state Compton profile) [28].
(3) Two-Step-One (TS1): For single-photon absorption at lower photon energies
(threshold to several 100eV [22]) TS1 is known to dominate by far over
the shake-off contribution. A simplified picture of TS1 is that one electron
absorbs the photon and knocks out the second one via an electron-electron
collision on its way through the atom [29]. A close connection between the
electron impact ionization cross section and the ratio of double to single
ionization by single-photon absorption as function of the energy is seen
experimentally [29] and theoretically [22], supporting this simple picture. For
the TS 1 mechanism the electron correlation is on a very short time scale (a
few attoseconds) and confined to a small region of space (the size of the
electron cloud).
(4) Rescattering: Rescattering is a version of the TS1 mechanism which is
induced only by the laser field. The mechanism was proposed originally by
Kuchiev [30] under the name "antenna model." He suggested that one of the
electrons is driven in the laser field acting as an antenna absorbing the energy
which it then shares with the other electron via correlation. Corkum [31]
and Schafer [32] extended this basic idea and interpreted the process in
the two-step model: First one electron is set free by tunneling. Then it is
accelerated by the laser field and is driven back to its parent ion with about
50% probability. Upon recollision with the ion the electron can recombine
and emit higher harmonic radiation. Besides that it could be elastically
scattered and further accelerated or it could be inelastically scattered with
simultaneous excitation or ionization of the ion. In contrast to TS1 in this
case there is a femtosecond time delay between the first and the second step.
Also the wave function of the rescattered electron explores a larger region
of space than in the case of TS1 [33-35].
Strong experimental evidence favoring the rescattering process to be domi-
nantly responsible for double ionization by strong laser fields was later provided
by the observation that double ejection is strongly suppressed in ionization with
circularly polarized light [36,37] (see also Fig. 19 of ref. [3]). The rescattering
mechanism is inhibited by the circular polarization since the rotating electric
field does not drive the electrons back to their origin. The other mechanisms, in
contrast, are expected to be polarization independent.
Further insight in the double ionization process clearly necessitates differential
measurements beyond the ion yield. Two types of such experiments have been
reported recently: Electron time-of-flight measurements in coincidence with the
ion charge state [38,39] and those using COLTRIMS, where at first only the
ion momenta [40-42] and later the ion momenta in coincidence with one electron
[ 17,43-45] have been measured.
V] MULTIPLE IONIZATION IN S T R O N G L A S E R FIELDS 11

V. Recoil Ion M o m e n t a

A. FROM NONSEQUENTIAL TO SEQUENTIAL DOUBLE IONIZATION

Recoil ion m o m e n t u m distributions have been measured for helium (He 1+,
He2+)[40], neon (Ne 1+, Ne 2+, Ne3+)[41] and argon (Ar l+, Ar2+)[45,46].
Figure 6 summarizes some of the results for neon. The m o m e n t u m distribution
of the singly charged ion is strongly peaked at the origin as in the case o f helium
(Fig. 4), reflecting the fact that tunnel ionization is most likely at the m a x i m u m
of the field (see Eq. 1). The structure o f the m o m e n t u m distribution of the doubly
charged ions changes strongly with the peak intensity. In the region where the
rates suggest the dominance of nonsequential ionization the ion momenta show
a distinct double peak structure (Fig. 6(2)). At higher intensities, where rates can
be described by assuming sequential ionization, the momenta o f the Ne 2+ ions
are peaked at the origin as for single ionization. The studies for helium show a
similar double peak structure at 6.6 x 1014 W/cm 2 (see Fig. 9).
The evolution of the ion momentum distributions with laser peak power has
been studied in detail for argon [46], too, confirming the fact that at the transition
to the nonsequential regime an increase in laser power results in colder ions. The
argon data, however, show no distinct double peak structure (see Fig. 7), where
the sequential ionization already sets in at about 6.6x 1014 W/cm 2. The reason
might be that the sequential contribution fills "the valley" in the m o m e n t u m

FIG. 6. Neon double ionization by 800 nm, 25 fs laser pulses. Left-hand panel: Rate of single and
double ionization as a function of the laser power (from ref. [47]). The solid line shows the rate
calculated in an independent event model. Right-hand panel: Recoil-ion momentum distributions
at intensities marked in the left-hand panel. A projection of the double-peaked distribution (2) is
shown in Fig. 10. Horizontal axis: Momentum component parallel to the electric field. Vertical axis:
One momentum component perpendicular to the field (data partially from ref. [41]).
12 R. D 6 r n e r et al. [V

(a)

~3

9;, /1: - \\
9 //" " \\ o,~

....... , ........ . , .........

kl'/7,,.~,,~ .... t ; , ....... J ......... ~. . . . . . . . . J ....... La, . . . . 5a.aP~i


-4 -3 -2 -1 0 1 2 3 4
Prz (a.u.)

FIG. 7. Momentum distribution of Ar 2+ ions created in the focus of a 220 fs, 800 nm laser
pulse at peak intensities of (a) 3.75• l0 TMW/cm 2 and (b) 12• 1014 W/cm 2 in the direction of the
polarization. The distributions are integrated over the directions perpendicular to the polarization.
Solid circles: distribution of Ar 2+ ions; dotted line: distribution of Ar 1§ ions; dashed line: results of
the independent electron model of convoluting the Ar l+ distribution with itself; solid line: results
of the independent-electron ADK model (see text); open circles in (a): distribution of He 2+ ions at
3.8• l014 W/cm 2 (figure from ref. [46], helium data from ref. [40]).

distribution at the origin before a double peak structure has developed. In


ref. [45] it has been argued based on classical kinematics that excitation of a
second electron during recollision followed by tunneling ionization of the excited
electron might be responsible for "filling the valley." Due to the open 3d shell
in Ar, excitation cross sections are much larger than in Ne. At the highest
intensity the single peak distribution can be at least qualitatively understood in
an independent two-step picture (see Fig. 7). The dash-dotted lines in Fig. 7
show the measured momentum distributions of Ar 1+ ions, the dashed line is this
distribution convoluted with itself. Such a convolution models two sequential and
totally uncorrelated steps of single ionization spaced in time by a random number
of optical cycles. Figure 7 shows that for argon at 12• W/cm 2 (which,
V] MULTIPLE IONIZATION IN STRONG LASER FIELDS 13

judging from the rates, is in the sequential regime), this very simple approach
describes the ion momentum distributions in double ionization rather well. One
obvious oversimplification of this convolution procedure is that it implicitly
assumes that the momentum distributions do not change with binding energy.
A more refined independent-event approach would use different binding energies
for both steps. As an alternative simple model, the momentum distribution for
removal of the first electron and the second electron have been calculated in
the ADK (Ammosov-Delone-Krainov) model (see, e.g., ref. [48], Eq. 10) using
the correct binding energies for both steps. The result of convoluting these
two calculated distributions is shown by the solid lines in Fig. 7. Clearly such
modeling fails in the regime where sequential ionization dominates (Fig. 7a).

B. THE ORIGIN OF THE DOUBLE PEAK STRUCTURE

The recoil ion is an important messenger carrying detailed information on


the time evolution of the ionization process. It allows not only to distinguish
between sequential and nonsequential ionization but also to rule out some of the
nonsequential mechanisms as we will show now.
Analogous to the situation for single ionization discussed above one can
estimate the net momentum accumulated by the doubly charged ion from the

~tlt~
laser pulse as

He2+ ~t, 2
Pz (t~) = E ( t ) sin tot dt + 2 E ( t ) sin tot dt. (2)
2

The first electron is removed at time tl and the ion switches its charge from 1+
to 2+ at time tl 2. It is assumed that there is no momentum transfer to the ion
from the first emitted electron during double ionization. Thus, as in the case of
single ionization the phase of the field at the instant of the emission of the first
and of the second electron is encoded in the ion momentum.
Shake-off and TS2 will both lead to a momentum distribution peaked at zero,
similar to single ionization. In both cases the emission of the second electron
follows the first with a time delay, which is orders of magnitude shorter than
the laser period. Hence tl 2 = t~ in Eq. (2), and since the first electron is emitted
He2+
most likely at the field maximum Pz would also peak at zero for shake-off and
TS1. Consequently, the observed double peak structure for He and Ne directly
rules out these mechanisms.
For the rescattering there is a significant time delay between the emission of
the first electron and the return to its parent ion. Estimating tl 2 for a rescattering
trajectory which has sufficient energy to ionize leads to ion momenta close
to the measured peak positions [40,41,49]. The high momenta of the doubly
and triply charged ions are direct proof of the time delay introduced by the
rescattering trajectory. It is this time delay with respect to the field maximum
14 R. D 6 r n e r et al. [V

time

kb ka

{.

2 1

FIG. 8. Feynman diagram describing the rescattering and TS1 mechanism (from ref. [50]). See
text.

that is responsible for multiple ionization and allows an effective net momentum
transfer to the ion by accelerating the parent ion. Within the classical rescattering
model the final momentum of the doubly charged ion will be the momentum
received from the field (as given by Eq. 2) plus the momentum transfer from the
recolliding electron to the ion.
Soon after the measurement of the first ion momentum distributions Becker
and Faisal succeeded in the first theoretical prediction of this quantity. They
calculated double ionization of helium using (time-independent) S-matrix theory.
They evaluated the Feynman diagram shown in Fig. 8. Time progresses from
bottom to top. Starting with 2 electrons in the helium ground state at time ti, the
laser field couples once at tl to electron 1 (VATI). Electron 1 is then propagated
in a Volkov state (k) in the presence of the laser field, while electron 2 is
in the unperturbed He l+ ground state (j). Physically the Volkov electron does
not have a fixed energy but can pick up energy from the field. This describes
e.g. an acceleration of the electron in the field and its return to the ion. At
time t2 one interaction of the two electrons via the full Coulomb interaction
is included. This allows for an energy transfer from the Volkov electron to the
bound electron. Finally, both electrons are propagated independently in Volkov
states, describing their quiver motion in the field. By evaluating this diagram
Becker and Faisal obtained excellent agreement with the observed ion yields (see
ref. [51,52] for helium and ref. [53] for an approximated rate calculation on other
rare gases). The ion momentum distribution calculated as the sum momentum of
the two electrons predicted by this diagram is shown in Fig. 9b. The calculation
correctly predicts the double peak structure and the position of the maxima.
The minimum at momentum zero is more pronounced in the calculation than
in the data. The major approximations which might be responsible for this are:
Only one step of electron-electron energy transfer is taken into account (see
ref. [57] for a discussion of the importance of multiple steps); no intermediate
excited states are considered; and the laser field is neglected for all bound states
as in turn the Coulomb field is neglected in the continuum states. To unveil
the physical mechanism producing the double hump structure Becker and Faisal
V] MULTIPLE IONIZATION IN STRONG LASER FIELDS 15

I = 6 . 6 . 1 0 '4 W / c m 2
1.0 ' I ' I ' I ' I ' I ' I ' I ' ' I ' I ' I ' I ' I ' I ' i '

0.8 - a] i .. d]

o4 ,, J
o,~ ; ;
_.,.' ........ :. . . . .
9 ! 9 i 9 | - , - , 9 | 9 , - 9 | 9 | 9 | 9 | 9 | 9 | 9 ! 9

~" 0.8- b] _ I

-~ 0.6 , ,

~z 0,4 i
"- 0.2
_ , ~ . , . , . , .~,m .
' I ! I ' I ' i | i , I ' i ' 9 | ' | ' i 9 | 9 ! 9 ! 9 | 9

c] .. f]
0,8-

0,6-
.. ;% ;I

0,4-

L"
0 , 2 -/

0,0 l , a , I , , I ___ , I , I , I , L . I , i

-8 -6 -4 -2 0 2 4 6 -6 -4 -2 0 2 4 6 8
Prz [a.u.]

FIG. 9. Momentum distribution of He 2+ ions at an intensity of 6 . 6 • 1014 W / c m 2 for all panels.


Prz is the component parallel to the laser polarization. (a) Experiment (from ref. [40]); (b) results of
the S-matrix calculation (from Becker and Faisal [ 5 0 ] ) ; (c) S-matrix with additional saddle-point
approximation (from Goreslavskii and P o p r u z h e n k o [ 5 4 ] ) ; (d) solution of the one-dimensional
Schr6dinger equation (from Lein et al. [34]); (e) Classical Trajectory Monte Carlo calculations
(from Chen e t al. [55]); (f) Wannier-type calculation (from Sacha and Eckhardt [56]).

have evaluated the diagram also by replacing the final Volkov states by plane
waves. Physically this corresponds to switching off the laser field after both
electrons are in the continuum. In the calculation this led to a collapse of the
double peak structure to a single peak similar to single ionization. This confirms
our interpretation given above, that it is the acceleration of the ion in the field
after the rescattering (starting at tl 2 in Eq. 2) that leads to the high momenta.
The S-matrix theory also yielded good agreement with the observed narrow
momentum distribution in the direction perpendicular to the laser field.
Later, different approximations in the evaluation of the diagram (Fig. 8) have
been introduced. First, Kopold and coworkers [58] replaced the electron-electron
interaction by a contact potential and additionally used a zero-range potential
for the initial state. This simplified the computation considerably while still
yielding the observed double peak structure, not only for helium but also for neon
16 R. D 6 r n e r et al. [V
1.0

0,8 ~ .%, ,,X


0.6
..-9 ~9~ 9
..
9
-
13.10~4 W/cm2
0.4

0.2

O.8

-- 0.6
8 - 1 0 ~4 W / c ~
0,4
o
0,2

c)
0,8

0.6
13" 10~4W/crn:
0,4

0,2

0.0
-8 -6 -4 -2 0 2 4 6 8
Prz (a.u.)

FIc. 10. Momentum distribution ofNe2+: (a) projection of data in Fig. 6 (2) at 13• 1014 W/cm 2
(from Moshammer et al. [41]); (b) S-matrix calculation evaluating the diagram in Fig. 8 with
contact potentials at 8• W/cm 2 (from Kopold et al. [58]); (c) Wannier-type calculation at
13• 1014 W/cm 2 (from Sacha and Eckhardt [56]).

(Fig. 10b) and other rare gases. They found that the inclusion of intermediate
excited states of the singly charged ion yields a filling of the minimum at zero
momentum.
Goreslavskii and Popruzhenko [54,59] used the saddle-point approximation
for the intermediate step. This additional approximation did not change the
calculated ion momenta strongly (see Fig. 9c) but simplified the computation,
allowing to investigate also the correlated electron emission discussed in the next
section.
A conceptionally very different approach was used by Sacha and Eck-
hardt[56]. They argued that the rescattering will produce a highly excited
intermediate complex, which will then decay in the presence of the field.
This decay process will not have any memory of how it was created. They
assumed a certain excitation energy as free parameter in the calculations and then
propagated both electrons in the classical laser field semiclassically in reduced
dimensions. Therefore they analyzed this decay by a Wannier-type analysis.
Wannier theory is known to reproduce the electron angular dependence as well
as the recoil ion momenta for the case of single-photon double ionization [60-
62,25]. In this case the Wannier configuration would be the emission of both
V] MULTIPLE IONIZATION IN STRONG LASER FIELDS 17

electrons back-to-back, leaving the recoil ion at rest on the saddle of the electron-
electron potential. For single-photon absorption from an S state this configuration
is forbidden by selection rules; it would allow, however, the absorption of an even
number of photons.
In the multiphoton case the external field has to be included in addition
to the Coulomb potential among the particles. This leads to a saddle in the
potential, which is not at rest at the center between the electrons but at
momenta which correspond to the observed peaks. Sacha and Eckhardt analyzed
classical trajectories in the saddle potential created by the field and the Coulomb
potentials. At a given laser field the decay of the excited complex in the field is
characterized only by two parameters: The time when the complex is created and
the total energy. Interestingly the recoil ion momentum obtained this way exhibits
a double peak structure, which does not depend strongly on the creation time but
on the energy. They find parallel and perpendicular momentum distributions,
which for helium (Fig. 9f) and for neon (Fig. 10c) are in reasonable agreement
with the experiment. This argument of a time-independent intermediate complex
seems to contradict the claim that the high recoil momenta and the double peak
result from the time delay due to the rescattering. One has to keep in mind,
however, that within the rescattering model the recollision energy and hence
the total energy of the complex analyzed by Sacha and Eckhardt is uniquely
determined by the recollision time. In a recent work they extended this model
to examine the decay of highly excited three-electron atoms [63].
The S-matrix approaches discussed above are based on the time-independent
Schr6dinger equation. One of the advantages of such approaches is that they
allow a precise definition of a mechanism (see, e.g., ref. [ 10]). Each particular
diagram represents one mechanism. The price that has to be paid is the loss
of information on the time evolution of the system. The diagram contains the
time order of interactions, but not the real time between them. Starting from the
time-dependent Schr6dinger equation in contrast gives the full information on
the time evolution of the many-body wave function in momentum or coordinate
space. In these coordinate space density distributions it is, however, often difficult
to clearly define what one means with a mechanism. Lein and coworkers
found a very elegant way to solve this problem [34,35]. Instead of plotting the
density in coordinate space they calculated the Wigner transform of the wave
function, depending on momentum and position. Integrated over the momentum
coordinate it is the density in coordinate space and integrated over the position
it is the distribution in momentum space. The Wigner transform can be read
as a density in phase space. Lein and coworkers plotted for example the phase
space evolution of the recoil ion in the polarization direction. This presentation
of a quantum-mechanical wave function is very close to the presentation of the
classical phase space trajectories. The rescattering mechanism can be seen very
clearly in this presentation.
18 R. D 6 r n e r et al. [V

Computation of the time-dependent Schr6dinger equation for three particles


in three dimensions is extremely challenging. Even though great progress has
been made in this field (see e.g. refs. [64-69]), there are no predictions of recoil
ion momenta or other differential information based on the solution of the time-
dependent Schr6dinger equation in three dimensions for the "long" wavelength
regime of presently available high-intensity lasers.
To allow for a practical calculation of the time evolution of the three-body
system two rather different approximations have been made: (a) Reducing the
dimensions from three for each particle to only one along the laser polarization
and (b) keeping the full dimensionality but using classical mechanics instead of
the Schr6dinger equation [70].
Lein et al. [34] reported the first results on recoil ion momenta based on
an integration of the one-dimensional Schr6dinger equation (see Fig. 9d). The
momentum distribution of the He 2+ ions at 6.6 x 1014 W/cm 2 in Fig. 9d) peaks at
zero momentum in contrast to all other results. It will become clear in Sect. VII
that there is evidence in Lein et al. 's calculation for a correlated emission of both
electrons into the same hemisphere. A well-known problem of one-dimensional
calculations is that the effect of electron repulsion is overemphasized. This might
be partially responsible for "filling the valley" in these calculations. For further
discussion see Sect. VII.
Chen et al. [55] have performed a Classical Trajectory Monte Carlo calculation
(CTMC) in which they solved the classical Hamilton equations of motion for all
three particles in the field. Instead of a full classical simulation of the process
(see, e.g., ref. [71] for CTMC calculation for single ionization and ref. [70] for
refined classical calculation of double ionization rates) they have initialized one
electron by tunneling and then propagated all particles classically. This also
yields the observed double peak structure. Such CTMC calculations have proven
to be extremely successful in predicting the highly differential cross sections
from ion impact single and multiple ionization (see refs. [72-79] for some
examples). One of the virtues of this approach is that the output comprises the
momenta for each of the particles for each individual ionizing event, exactly like
in a COLTRIMS experiment. In addition, however, each particle can be followed
in time, shedding light on the mechanism. Such detailed studies would be highly
desirable for the strong field case, too.
All theoretical analyses of the observed double peak structure in the recoil
ion distribution confirm the first conclusion from both experimental teams
reporting these structures: (a) It is an indication of the nonsequential process
and (b) it is consistent with the rescattering mechanism, which is included in
one or the other way in the various theoretical models.
In the direction perpendicular to the polarization the observed and all
calculated distributions are very narrow and peak at zero. Since there is no
acceleration by the laser field in this direction the transverse momentum of the
VI] MULTIPLE IONIZATION IN STRONG LASER FIELDS 19

ion is purely either from the ground state or from the momentum transfer in the
recollision process. All theories which are not confined to one dimension agree
roughly with the experimental width of the distribution. This direction should
be most sensitive to the details of the recollision process since that is where
the parallel momentum acquired from the field is scattered to the transverse
direction. Hence, a closer inspection of the transverse momentum transfer is of
great interest for future experimental and theoretical studies.

VI. Electron Energies

Electron energy distributions for double ionization have been reported for
helium [39], argon [80], neon [81] and xenon [38]. All these experiments find in
the sequential regime that the electron energies from double ionization are much
higher than those generated in single ionization. This is in full agreement with the
recoil ion momenta discussed above, since the mechanism being predominantly
responsible for producing high-energy electrons is exactly the same: It is a fact
that due to the rescattering the electrons from double ionization are not promoted
to the continuum at the field maximum but at a later time. Depending on the
actual time delay an energy of up to 2Up (see Eq. 1) can be acquired. The
work for helium (Fig. 11) and neon shows that the electron spectra extend well
above this value. Energies beyond 2Up are only obtainable if the recolliding
electron is backscattered during the (e,2e) collision. In this case the momentum
they have after the recollision adds to the momentum acquired in the field.
A large amount of elastically backward-scattered electrons has been observed for
single ionization where a plateau in the energy distribution is found extending
to energies of up to l OUp.

Electron energy (Up) Electron energy (Up)


1 2 3 4 5 1
,
2
,
3
,
4
,
5
,

> 1000
9 ! 9 , 9 , , , 9 . , . .

N
C
100
oO 10

~N

E 0.1
k._
z
O
0.01 (a) (b)

2's 5'o 7; loo


, , , ,

50 100 150 200 250


Electron energy (eV) Electron energy (eV)

FIG. 11. Electron energy spectra from single ionization (solid line) and double ionization (dots)
of helium at (a) 8• W/cm 2 and (b) 4x 1014 W/cm 2 (from ref. [39]).
20 R. D 6 r n e r et al. [VII

VII. Correlated Electron M o m e n t a

More information can be obtained from the momentum correlation between


the two electrons. In an experiment one possible choice would be to observe
the momenta of both electrons in coincidence. In this case the recoil ion
momentum could be calculated employing momentum conservation. From an
experimental point of view however, it is easier to detect the ion and one of the
electrons, in which case the momentum of the second electron can be inferred
from momentum conservation. It is experimentally simpler since the additional
knowledge of the ion charge state allows for an effective suppression of random
coincidences. Moreover, electron and ion are detected by opposite detectors
circumventing possible problems of multihit detection. Many successful studies
for single-photon double ionization have been performed this way [25,62,82,
83]. Up to present, however, no fully differential experiment has been reported
for multiphoton double ionization. Weber et al. [ 17] and Feuerstein et al. [45]
reported measurements observing only the momentum component parallel to
the field of electron and ion integrating over all other momentum components.
Weckenbrock et al. [43] and Moshammer[41] have detected the transverse
momentum of one of the electrons in addition to the parallel momenta. In
these experiments, however, the transverse momentum of the ion could not be
measured with sufficient resolution, mainly due to the internal temperature of
the gas jet for argon and neon targets. Experiments on helium have not yet been
reported but are in preparation in several laboratories.

A. EXPERIMENTALFINDINGS

The correlation between the momentum components parallel to the polarization


is shown in Fig. 12. The electron momenta are integrated over all momentum
components perpendicular to the field direction. Events in the first and third
quadrants are those where both electrons are emitted to the same hemisphere,
the second and fourth quadrants correspond to emission to opposite half
spheres. The upper panel shows the electron momenta at an intensity of
3.6• 2, which is in the regime where nonsequential ionization is
expected. The distribution shows a strong correlation between the two electrons,
they are most likely emitted to the same hemisphere with a similar momentum
of about 1 au. At higher intensity, where double ionization proceeds sequentially,
this correlation is lost (lower panel in Fig. 12).
To interpret the correlation pattern it is helpful to consider the relationship
between the electron and the recoil ion momenta. We define the Jacobi
momentum coordinates kz+ and kz:

k+z = kezl + ke~2, (3)


k z = kez~ -- kez2, (4)
VII] M U L T I P L E I O N I Z A T I O N IN S T R O N G L A S E R F I E L D S 21

FIG. 12. Momentum correlation between the two electrons emitted when an Ar 2+ ion is
produced in the focus of a 220 fs, 800nm laser pulse at peak intensities of 3.8x 1014 W/cm 2 and
15x 1014 W/cm2. The horizontal axis shows the momentum component of one electron along the
polarization of the laser field; the vertical axis represents the same momentum component of the
corresponding second electron. Same sign of the momenta for both electrons represents an emission
to the same half sphere. The data are integrated over the momentum components in the direction
perpendicular to the polarization direction. The gray shading shows the differential rate in arbitrary
units on a linear scale (adapted from ref. [ 17]). Also compare this figure to Fig. 17.

with kzion = -kz+. T h e s e coordinates are a l o n g the d i a g o n a l s o f Fig. 12. H e n c e


the recoil ion m o m e n t u m distribution is s i m p l y a p r o j e c t i o n o f Fig. 12 onto
the diagonal kz+. T h e coordinates k] and k z are helpful to illustrate the relative
i m p o r t a n c e o f the two c o u n t e r a c t i n g effects o f e l e c t r o n - e l e c t r o n r e p u l s i o n and
acceleration o f particles by the optical field. B o t h influence the final-state
m o m e n t a in different ways. E l e c t r o n r e p u l s i o n (and t w o - b o d y e l e c t r o n - e l e c t r o n
scattering) does not c h a n g e kz+ but contributes to the m o m e n t u m kz. O n the
other hand, once both electrons are set free, the m o m e n t u m transfer r e c e i v e d
from the field is identical for both. T h e r e f o r e , this part o f the a c c e l e r a t i o n does
not change kz b u t adds to kz+. The o b s e r v e d wide kz+ and n a r r o w k z distributions
22 R. D 6 r n e r et al. [VII

FiG. 13. Momentum correlation between the two electrons emitted when an Ar 2+ ion is produced
in the focus of a 150 fs, 780nm laser pulse at peak intensities of 4.7• 1014 W/cm 2. Axis as in
Fig. 12. Each panel panel represents a part of the final state for a fixed transverse momentum
(p• of one of the electrons. (a) One of the electrons has a transverse momentum of p • < 0.1 au;
(b) 0.1 < p • < 0.2au; (c) 0.2 < p • < 0.3 au; (d) 0.3 < p • < 0.4au. The gray scale shows the
differential rate in arbitrary units and linear scale (from ref. [43]).

thus indicate that the joint acceleration of the electrons in the laser field clearly
dominates over the influence of electron repulsion.
For argon double ionization Weckenbrock et al. [43] and Moshammer et al.
[84] measured in addition to the momentum parallel to the field also the
transverse momentum of the detected electron. Both find that the correlation
pattern strongly depends on this transverse momentum (see Fig. 13). If one
electron is emitted with any transverse momentum larger than 0.1 au (i.e. at
some angle to the polarization axis) one mostly finds both electrons with a
similar momentum component in the field direction. It is this configuration that
dominates the integrated spectrum in Fig. 12. If, however, one electron is emitted
parallel to the polarization with a very small transverse momentum window of
p• < 0.1 au one finds that the parallel momentum distribution does no longer
peak on the diagonal. In this case most likely one electron is fast and the other
slow. This might be due to the fact that the 1/rl 2 potential forces the electrons
into different regions in the three-dimensional phase space. Consequently, for
electrons to have equal parallel momentum some angle between them is required.
VII] MULTIPLE IONIZATION IN STRONG LASER FIELDS 23

Accordingly, the peak at Pezl = P e z 2 = 1 au is found to be most pronounced if at


least one of the electrons has considerable transverse momentum. In tendency,
this feature can be explained by (e,2e) kinematics as discussed in ref. [84]:
Unequal momentum sharing is known to be most likely in field-free (e,2e)
reactions. The rescattered electron is only little deflected, losing only a little of its
longitudinal momentum during recollision. At the same time, the ionized electron
is low-energetic, resulting in very different start momenta of both electrons at
recollision time tl 2. At intensities not too close to the threshold this scenario
leads to asymmetric longitudinal energy sharing as calculated in refs. [54,85].

B. COMPARISON TO SINGLE-PHOTON AND CHARGED PARTICLE IMPACT


DOUBLE IONIZATION

One might expect that the pure effect of electron repulsion could be studied in
double ionization by single-photon absorption with synchrotron radiation. In this
case there is no external field in the final state that could accelerate the electrons.
Many studies have shown however, that the measured momentum distribution is
not only governed by the Coulomb forces in the final state, but also by selection
rules resulting from the absorption of one unit of angular momentum and the
accompanying change in parity. For helium for example the two-electron contin-
uum wave function has to have ~p0 character. Since these symmetry restrictions
on the final state are severe it is misleading to compare distributions of kezl versus
kez2 a s in Fig. 12 directly to those from single-photon absorption (this distribution
can be found in ref. [86]). The effect of electron repulsion can be more clearly
displayed in a slightly different geometry as shown in Fig. 14. Here one electron
is emitted along the positive x-direction and the momentum distribution of the
second electron is shown. The data are integrated over all directions of this
internal plane of the three-body system relative to the laboratory. Clearly electron
repulsion dominates the formation of this final state distribution: there is almost
no intensity for emission to the same half sphere. There is also a node for
emission of both electrons back-to-back. This is a result of the odd symmetry
of the final state. In the multiphoton case this node is expected for those events
where an odd number of photons is absorbed from the field (see e.g. [88]).
Another instructive comparison is the process of double ionization by charged
particle impact. Experiments have been reported for electron impact [89-91]
and fast highly charged ion impact [79,92]. The latter is of particular interest
from the strong field perspective since the potential "shock" induced at a target
atom by a fast highly charged projectile is in many aspects comparable to a
half cycle laser pulse. The time scale however is much shorter than that accessible
with lasers today. For their experiment colliding 1 GeV/u U 92+ projectiles on
helium for example Moshammer and coworkers [93] estimated a power density
of > 1019 W/cm 2 and a time of sub attoseconds. Under such conditions ion-
24 R. D 6 r n e r et al. [VII

Fie. 14. Single-photon double ionization of He at l eV and 20eV above threshold by linearly
polarized light (synchrotron radiation). Shown is the momentum distribution of electron 2 for fixed
direction of electron 1 as indicated. The plane of the figure is the internal momentum plane of the
two particles. The data are integrated over all orientations of the polarization axis with respect to
this plane. The figure thus samples the full cross section and all angular and energy distributions
of the fragments. The outer circle corresponds to the maximum possible electron momentum; the
inner one represents the case of equal energy sharing (from ref. [87]; compare also ref. [83]).

atom collisions can be successfully described by the Weizs/icker-Williams


formalism [94,95,93,96], which replaces the ion by a flash of virtual photons (for
a detailed discussion on the validity and limitations of this method see ref. [97]).
Since such an extremely short "photon field" also has contributions from very
high frequencies, i.e. virtual photon energies, the ionization is dominated by the
absorption of one photon per electron. This is contrary to the femtosecond laser
case discussed here. Multiple ionization in fast ion-atom collisions is dominated
by either the TS2 or the TS 1 process (with only a small amount of shake-off)
depending on the strength of the perturbation, i.e. the intensity of the virtual
photon field. The ratio of the projectile charge to the projectile velocity is usually
taken as a measure of the perturbation. Figure 15 shows the electron m o m e n t u m
correlation of double ionization of helium by 100 MeV/u C 6+ impact parallel
to the direction of the projectile. The dominant double ionization mechanism at
these small perturbations is TS1 [98], or, in a virtual photon picture, one photon
is absorbed during a collision by either one of the electrons and the second
is taken to the continuum due to electron-electron correlation. Under these
conditions the electron repulsion in the final state drives the electrons to opposite
half spheres, whereas the projectile itself passes so fast that during this short time
essentially no momentum is transferred to the system. Similar studies have been
performed with slower and more highly charged projectiles [79]. In this case the
dominant double ionization mechanism is TS2. The experiments show a joint
forward emission of both electrons. This effect has been interpreted in a two-step
picture: First the initial-state momentum distribution is lifted to the continuum by
absorption of two virtual photons, then in a second step the strong potential of the
VII] MULTIPLE IONIZATION IN STRONG LASER FIELDS 25

FIG. 15. Double ionization of helium by 100 MeV/u C 6+ impact. The horizontal and vertical axes
(Pill and P211) show the momentum components of electrons 1 and 2 parallel to the direction of the
projectile. The dashed curves demarcate the region of the two-electron momentum space which is
not accessed by the spectrometer. The gray scale is linear (adapted from ref. [92]).

projectile accelerates both electrons into the forward direction (see also ref. [96]
for a theoretical interpretation of the double ionization process; see ref. [99] for
another experiment showing directed multiple electron emission; see ref. [78] for
an analysis of the acceleration of an electron in the field of the projectile).

C. INTERPRETATION WITHIN THE RESCATTERING MODEL

The data shown in Fig. 12 can be qualitatively understood by estimating the


momentum transfer in the rescattering model. From this one obtains kinematical
boundaries of the momenta for different scenarios. For simplicity we restrict
ourselves to a single return of the electron.
If the electron recollides with an energy above the ionization threshold clearly
double ionization is possible. The electron will lose the energy (and hence the
momentum) necessary to overcome the binding of the second electron. The
remaining excess energy can be freely distributed among the two electrons in the
continuum. From electron impact ionization studies it is known that the electron
energy distribution is asymmetric, i.e. one fast, one slow electron is most likely,
for excess energies above 10 to 20 eV. The momentum vector of the electrons
can point in all directions, but forward scattering of one electron is most likely
(see ref. [100] for a review of electron impact ionization). After recollision the
electrons are further accelerated in the field yielding a net momentum transfer
at the end of the pulse, which is equal for both electrons and given by Eq. (1)
(replace to by the rescattering time tl 2). For each recollision energy this leads
to a classically allowed region of phase space, which is a circle centered on the
diagonal in Fig. 12. An example is shown in Fig. 16.
26 R. D 6 r n e r et al. [VII

5" 1
d

-2

-3

-4
-3 -2 -1 0 1 2 3 4
P z, e l ( a . u . ]

FIG. 16. Classically allowed region of phase space within the rescattering model for double
ionization of argon by 4.7 x 1014W/cm2, 800 nm light. Each circle corresponds to a fixed recollision
energy. Axis as in Fig. 12 (adapted from ref. [44]).

If the recollision energy is below the field-flee ionization threshold it is still


possible that double ionization occurs. Any detailed scenario for this case without
an explicit calculation is rather speculative since one deals with an excitation
process in a very strong field environment for which no experiment exist so
far. Already the levels of excited states are strongly modified compared to the
field-free case. The same will certainly be true for the cross sections. We can
however distinguish two extreme cases: An excited intermediate complex is
formed, which either is quenched immediately by the field or may survive at
least half a cycle of the field and will be quenched close to the next field
maximum. The probability of such survival will depend on the field at the time
of the return. For 3.8 or 4.7x1014 W/cm 2 the field at times corresponding to
a return energy sufficient to reach the first excited states of an Ar l+ ion (at
about 16-17 eV field flee) is so high that such a state would be above the
barrier and hence would not be bound. A scenario which leads to the observed
momentum of a 0.9-1 au for 3.8 and 4.7x 1014 W/cm 2 (data of figures 12, 13)
is the following [17,43]: Electron 1 has a return energy of about 17 eV, which
corresponds to the first exited states of the Ar l+ ion. Electron 1 is stopped,
electron 2 is excited and immediately field ionized. Both electrons thus start
with momentum zero at the time of the recollision. They are accelerated in the
field and, hence, end up with the same m o m e n t u m of about 0.9-1 au after the
pulse. This is in good agreement with the experimental observations for electrons
emitted in the same hemisphere. It does not explain, however, a considerable
number of events ejected into opposite hemispheres along the laser polarization.
Feuerstein et al. [45] performed the same experiment in argon at a lower laser
intensity of 2.5 x 1014 W/cm 2 (see Fig. 17). In this case sufficiently high return
energies for excitation correspond to a return time close to the zero crossing
VII] MULTIPLE IONIZATION IN STRONG LASER FIELDS 27

I I I I I I

- I I
I

_ I
I 9 ~

I
. , - m I I
II
I
I --O . . . . n l i ~ l l l I l i - . / I
9= , m , emInm, / I
I ii ~ I
I m l i m i I I
o

i
I I I l m , I
nImn I
Q_ I I I I

-I -- I W
I m I
I
I
_

-- I I . . . . . . I
I
I
I
I
I
I. . . . . . . . . . . I

I 1 I I I I

-3 -2 -1 0 1 2 3
pl I [a.u.]

FIG. 17. Correlated electron momentum spectrum of two electrons emitted from argon atoms
ii

at 0.25x 1015 W/cm 2. plI is the electron momentum component along the light polarization axis
of electron 1. Dashed line: kinematical constraints for recollision with excitation, assuming the
excited state is not immediately quenched. Solid line: kinematical constraints for recollision with
(e,2e) ionization (from ref. [45]).

of the field. Therefore one can expect that the excited state survives at least
until the next field maximum. Feuerstein et al. estimated an expected region in
phase space for excitation as shown in Fig. 17. For recollision events where the
second electron is lifted into the continuum the allowed region of phase space is
somewhat smaller than in Fig. 16 and confined to the two circles on the diagonal.
Feuerstein et al. used this argument to separate events in which the recollision
leads to an excited state and those which involve electron impact ionization.
Supporting this notion of an intermediate excited complex Peterson and
Bucksbaum [80] reported an enhanced production of low-energy electrons in the
ATI electron spectrum of argon previously unobserved which can be interpreted
in terms of inelastic excitation of Ar + or of multiple returns of the first electron.
Electrons from excited states field ionized at the field maximum will be detected
with very little momentum as they receive almost no drift velocity in the laser
field.

D. S-MATRIX CALCULATIONS

The full diagram shown in Fig. 8 has not yet been evaluated to obtain the
correlated electron momentum distribution. Goreslavskii and Popruzhenko suc-
ceeded, however, in calculating those distributions by making use of the saddle-
point approximation in the integration (see Fig. 18). The calculations shown
in Fig. 18 are restricted to zero transverse momentum; similar distributions for
28 R. D 6 r n e r et al. [VII

FIG. 18. Two-electron momentum distributions for double ionization of argon (similar to
Fig. 12), calculated by evaluating diagram 8 in the saddle-point approximation at an intensity of
3.8 x 10TMW/cm2. Contrary to the experimentthe calculations are not integrated over all momentum
components transverse to the field but restricted to electrons with no transverse momentum. The
right-hand panel presents the same distribution with the classically forbidden region of phase space
shown in white (compare Fig. 16) (adapted from ref. [54]).

neon and argon integrated over all transverse momenta can be found in ref. [85].
These calculations do not include intermediate excited states but only the direct
(e,2e) process. The calculations do not show a maximum on the diagonal as
seen in the experiments. To the contrary, they favor the situation where one
electron is slow and the other is fast. The authors of ref. [54] point out that
this is a direct consequence of the sharing of the excess energy in the (e,2e)
collision; the long-range Coulomb potential favors small momentum transfer
in the collision. By replacing the Coulomb potential with a contact potential
Goreslavskii and coworkers find a distribution which peaks on the diagonal,
much like the experimental results. The main reason is that a contact potential
does not emphasize small momentum transfers. It has to remain open at present
how well justified such a modification of the interaction potential is.
These calculations have been restricted to electrons with zero transverse
momentum. The trend seen in these calculations is in agreement with the
observation by Weckenbrock et al. [43] shown in Fig. 13a, where one electron
was confined to small transverse momenta. The calculations do not include
intermediate excited states but only direct electron impact ionization during
rescattering. Therefore the theoretical results are not too surprising since
electron impact ionization favors unequal energy sharing at the return energies
dominating here.

E. TIME-DEPENDENT CALCULATIONS

Calculations by the Taylor group solving the time-dependent Schr6dinger


equation in three dimensions predicted the emission of both electrons to the
same side prior to the experimental observation [64]. Similar conclusions have
been drawn from one-dimensional calculations [101]. In the low-field, short-
wavelength regime the full calculations have proven to be able to predict
VII] MULTIPLE IONIZATION IN STRONG LASER FIELDS 29

FIG. 19. Two-electron momentum distributions for double ionization of helium (similar to Fig. 12)
calculated by solving the one-dimensional time-dependent Schr6dinger equation at the following
intensities: (a) 1 x 1014 W/cm 2, (b) 3x 1014 W/cm 2, (c) 6.6x 1014 W/cm 2, (d) 10x 1014 W/cm 2,
(e) 13x 1014 W/cm 2, (f) 20x1014 W/cm 2 (adapted from [34]).

electron-electron angular distributions and the energy sharing among the


electrons as well as the total double ionization cross section [69,102]. In the
strong field case at 800 nm, however, the calculations are extremely demanding.
No electron-electron momentum space distributions have been reported up to
now. The total double ionization rates however are in good agreement with the
observations at 380 nm [66].
Several one-dimensional calculations have been performed at 800nm. All
calculations show the majority of electrons emitted to the same side [34,103,
104]. From the calculated electron densities in coordinate space Lein and
coworkers have obtained momentum distributions (Fig. 19). The enhanced
emission probability in the first and third quadrants at intermediate in panels
(d) and (e) is clearly visible. Different from the experiment, however, a strongly
reduced probability is observed along the diagonal, which is most likely an
artifact of the one-dimensional model. While in three dimensions electron
30 R. D 6 r n e r et al. [IX

repulsion can lead to an opening angle between the electrons having the same
momentum component in the polarization direction, this is impossible in one
dimension. Here the electron repulsion necessarily leads to a node on the
diagonal for electrons emitted at the same instant in the field.
For 400nm radiation these calculations have also shown clear rings corre-
sponding to ATI peaks in the sum energy of both electrons [ 105]. Analogous to
ATI peaks in single ionization they are spaced by the photon energy. Similar rings
have been seen also in three-dimensional calculations at shorter wavelength [65].

VIII. Outlook

The application of COLTRIMS yielded the first differential data for double
ionization in strong laser fields. Compared to the experimental situation in double
ionization by single-photon absorption, however, the experiments are still in
their infancy. So far correlated electron momenta have been measured only for
argon and neon. Clearly experiments on helium are highly desirable since this
is where theory is most tractable. Also, mainly the momentum component in
the polarization direction has been investigated so far, resulting in a big step
forward in the understanding of multiple ionization in strong laser fields. None
of the experiments up to now has provided fully differential data since not all
six momentum components of the two electrons were analyzed. Therefore, no
coincident angular distributions as for single-photon absorption are available at
this point (see ref. [88] for a theoretical prediction of these distributions). Most
important for such future studies is a high resolution of the sum energy of the two
electrons, which would allow to count the number of photons absorbed. From
single-photon absorption it is known that angular distributions are prominently
governed by selection rules resulting from angular momentum and parity, hence,
from the even or oddness of the number of absorbed photons.
Another important future direction is a study of the wavelength dependence of
double ionization. The two cases of single and multiphoton absorption discussed
here are only the two extremes. The region of two- and few-photon double
ionization is experimentally completely unexplored. Experiments for two-photon
double ionization of helium will become feasible in the near future at the
VUV FEL facilities such as the TESLA Test facility in Hamburg.

IX. Acknowledgments

The Frankfurt coauthors would like to thank H. Giessen, G. Urbasch, H. Roskos,


T. L6ttter and M. Thomson for collaboration on some of the experiments
described here and C. Freudenberger for preparation of many of the figures.
The Heidelberg coauthors are indebted to the Max-Born-Institute in Berlin,
X] MULTIPLE IONIZATION IN STRONG LASER FIELDS 31

providing the laser facilities for the experiments. Moreover, H. Rottke, C. Trump,
M. Wittmann, G. Korn and W. Sandner made decisive contributions to the
experimental setup, helping in the realization of the experiments during beam-
times and contributed strongly in the evaluation and interpretation of the data.
We thank A. Becker, E Faisal and W. Becker for many helpful discussions
and for educating us on S-matrix theory. We have also profited tremendously
from discussions with K. Taylor, D. Dundas, M. Lein, V. Engel, J. Feagin,
L. DiMauro and P. Corkum. This work is supported by DFG, BMBF, GSI. R.D.
acknowledges supported by the Heisenberg-Programm of the DFG. R.M., B.E
and J.U. acknowledge support by the Leibniz-Programm of the DFG. T.W. is
grateful for financial support of the Graduiertenf'6rderung des Landes Hessen.

X. R e f e r e n c e s

1. G6ppert-Mayer, M. (1931). Ann. d. Phys. 9, 273.


2. Burnett, K., Reed, V., and Knight, R (1993). J. Phys. B 26, 561.
3. DiMauro, L., and Agostini, R (1995). "Advances in Atomic and Molecular Physics." Academic
Press, New York.
4. Protopapas, M., Keitel, C., and Knight, R (1997). Phys. Rep. 60, 389.
5. Backus, S., Durfee, C., Murnane, M., and Kapteyn, H. (1998). Rev. Sci. Instrum. 69, 1207.
6. Brabec, T., and Krausz, E (2000). Rev. Mod. Phys. 72, 545.
7. Diels, J.-C., and Rudolph, W. (1995). "Ultrashort Laser Pulse Phenomena." Academic Press,
New York.
8. Rulliere, C. (1998). "Femtosecond Laser Pulses. Principles and Experiments." Springer Verlag,
New York.
9. D6rner, R., Mergel, V., Jagutzki, O., Spielberger, L., Ullrich, J., Moshammer, R., and Schmidt-
B6cking, H. (2000). Phys. Rep. 330, 96.
10. McGuire, J. (1997). "Electron Correlation Dynamics in Atomic Collisions." Cambridge
University Press, Cambridge.
11. McGuire, J., Berrah, N., Bartlett, R., Samson, J., Tanis, J., Cocke, C., and Schlachter, A. (1995).
J. Phys. B 28, 913.
12. Briggs, J., and Schmidt, V. (2000). J. Phys. 33, R1.
13. Wolf, S., and Helm, H. (1997). Phys. Rev. A 56, R4385.
14. van der Poel, M., Nielsen, C.V., Gearba, M.-A., and Andersen, N. (2001). Phys. Rev. Lett.
87, 123201.
15. Turkstra, J.W., Hoekstra, R., Knoop, S., Meyer, D., Morgenstern, R., and Olson, R.E. (2001).
Phys. Rev. Lett. 87, 123202.
16. Flechard, X., Nguyen, H., Wells, E., Ben-Itzhak, I., and DePaola, B.D. (2001). Phys. Rev. Lett.
87, 123203.
17. Weber, T., Giessen, H., Weckenbrock, M., Staudte, A., Spielberger, L., Jagutzki, O., Mergel, V.,
Urbasch, G., Vollmer, M., and D6rner, R. (2000a). Nature 404, 608.
18. Moshammer, R., Unverzagt, M., Schmitt, W., Ullrich, J., and Schmidt-B6cking, H. (1996a).
Nucl. Instr. Meth. B 108, 425.
19. Schyja, V., Lang, T., and Helm, H. (1998). Phys. Rev. 57, 3692.
20. Helm, H., Bjerre, N., Dyer, M., Huestis, D., and Saeed, M. (1993). Phys. Rev. Lett 70, 3221.
21. Hino, K., Ishihara, T., Shimizu, E, Toshima, N., and McGuire, J.H. (1993). Phys. Rev. A
48, 1271.
32 R. D 6 r n e r et al. [X

22. Kheifets, A., Bray, I., Soejima, K., Danjo, A., Okuno, K., and Yagishita, A. (2001). J. Phys. B
34, L247.
23. Shingal, R., and Lin, C. (1991). J. Phys. B 24, 251.
24. Lambropoulos, P., Maragakis, E, and Zhang, J. (1998). Phys. Rep. 305, 203.
25. D6rner, R., Br~iuning, H., Feagin, J., Mergel, V., Jagutzki, O., Spielberger, L., Vogt, T.,
Khemliche, H., Prior, M., Ullrich, J., et al. (1998a). Phys. Rev. A 57, 1074.
26. Byron, E, and Joachain, C. (1967). Phys. Rev. 164, 1.
27. Spielberger, L., Jagutzki, O., D6rner, R., Ullrich, J., Meyer, U., Mergel, V., Unverzagt, M.,
Damrau, M., Vogt, T., Ali, I., et al. (1995). Phys. Rev. Lett. 74, 4615.
28. Spielberger, L., Br~iuning, H., Muthig, A., Tang, J., Wang, J., Qui, Y., D6rner, R., Jagutzki, O.,
Tschentscher, T., Honkim~iki, V., et al. (1999). Phys. Rev. 59, 371.
29. Samson, J. (1990). Phys. Rev. Lett. 65, 2863.
30. Kuchiev, M.Y. (1987). Soy. Phys.-JETP Lett. 45, 404.
31. Corkum, P. (1993). Phys. Rev. Lett. 71, 1994.
32. Schafer, K., Yang, B., DiMauro, L., and Kulander, K. (1993). Phys. Rev. Lett. 70, 1599.
33. Watson, J., Sanpera, A., Lappas, D., Knight, P., and Burnett, K. (1997). Phys. Rev. Lett.
78, 1884.
34. Lein, M., Gross, E., and Engel, V. (2000a). Phys. Rev. Lett. 85, 4707.
35. Lein, M., Gross, E., and Engel, V. (2001a). Opt. Express 8, 441. On-line: http://
www.opticsexpress.org/oearchive/source/30744.htm.
36. Fittinghoff, D., Bolton, P., Chang, B., and Kulander, K. (1994). Phys. Rev. A 49, 2174.
37. Dietrich, P., Burnett, N.H., Ivanov, M., and Corkum, P.B. (1994). Phys. Rev. A 50, R3585.
38. Witzel, B., Papadogiannis, N.A., and Charalambidis, D. (2000). Phys. Rev. Lett. 85, 2268.
39. Lafon, R., Chaloupka, J.L., Sheehy, B., Paul, P.M., Agostini, P., Kulander, K.C., and
DiMauro, L.E (2001). Phys. Rev. Lett. 86, 2762.
40. Weber, T., Weckenbrock, M., Staudte, A., Spielberger, L., Jagutzki, O., Mergel, V., Urbasch, G.,
Vollmer, M., Giessen, H., and D6rner, R. (2000b). Phys. Rev. Lett. 84, 443.
41. Moshammer, R., Feuerstein, B., Schmitt, W., Dorn, A., Schr6ter, C., Ullrich, J., Rottke, H.,
Trump, C., Wittmann, M., Korn, G., et al. (2000). Phys. Rev. Lett. 84, 447.
42. Weber, T., Jagutzki, O., Hattass, M., Staudte, A., Nauert, A., Schmidt, L., Prior, M., Landers, A.,
Br/iuning-Demian, A., Br~iuning, H., et al. (2001a). J. Phys. B 34, 3669.
43. Weckenbrock, M., Hattass, M., Czasch, A., Jagutzki, O., Schmidt, L., Weber, T., Roskos, H.,
L6ffler, T., Thomson, M., and D6rner, R. (2001). J. Phys. B 34, L449.
44. Weckenbrock, M. (2001). Diploma Thesis. J.W. Goethe University, Frankfurt/Main. On-line:
http ://hsbpc 1.ikf. physik, uni- frankfurt, de/publications/Diplom_Doktor.html.
45. Feuerstein, B., Moshammer, R., Fischer, D., Dorn, A., Schr6ter, C.D., Deipenwisch, J., Lopez-
Urrutia, J., H6hr, C., Neumayer, P., Ullrich, J., et al. (2001). Phys. Rev. Lett. 87, 043003.
46. Weber, T., Weckenbrock, M., Staudte, A., Spielberger, L., Jagutzki, O., Mergel, V., Urbasch, G.,
Vollmer, M., Giessen, H., and D6rner, R. (2000). J. Phys. B 33, L127.
47. Larochelle, S., Talebpour, A., and Chin, S.L. (1998). J. Phys. B 31, 1201.
48. Delone, N., and Krainov, V. (1998). Phys. Usp. 41,469.
49. Feuerstein, B., Moshammer, R., and Ullrich, J. (2000). J. Phys. B 33, L823.
50. Becker, A., and Faisal, E (2000). Phys. Rev. Lett. 84, 3546.
51. Becker, A., and Faisal, E (1996). J. Phys. B 29, L 197.
52. Becker, A., and Faisal, E (1999a). Phys. Rev. A 59, R1742.
53. Becker, A., and Faisal, E (1999b). J. Phys. B 32, L335.
54. Goreslavskii, S., and Popruzhenko, S. (2001). Opt. Express 8, 395. On-line: http://
www.opticsexpress.org/oearchive/source/30694.htm.
55. Chen, J., Liu, J., Fu, L., and Zheng, W. (2000). Phys. Rev. 63, 011404R.
56. Sacha, K., and Eckhardt, B. (2001a). Phys. Rev. 63, 043414.
X] M U L T I P L E I O N I Z A T I O N IN S T R O N G L A S E R F I E L D S 33

57. Bhardwaj, V.R., Aseyev, S.A., Mehendale, M., Yudin, G.L., Villeneuve, D.M., Rayner, D.M.,
Ivanov, M.Y., and Corkum, P.B. (2001). Phys. Rev. Lett. 86, 3522.
58. Kopold, R., Becker, W., Rottke, H., and Sandner, W. (2000). Phys. Rev. Lett. 85, 3781.
59. Goreslavskii, S., and Popruzhenko, S. (2001). J. Phys. B 34, L239.
60. Feagin, J. (1995). J. Phys. B 28, 1495.
61. Feagin, J. (1996). J. Phys. B 29, L551.
62. D6rner, R., Feagin, J., Cocke, C., Br/iuning, H., Jagutzki, O., Jung, M., Kanter, E.,
Khemliche, H., Kravis, S., Mergel, V., et al. (1996). Phys. Rev. Lett. 77, 1024. Erratum:
(1997). Phys. Rev. Lett. 78, 2031.
63. Sacha, K., and Eckhardt, B. (2001b). Phys. Rev. 64, 053401.
64. Taylor, K., Parker, J., Dundas, D., Smyth, E., and Vitirito, S. (1999). Laser Physics 9, 98.
65. Parker, J.S., Moore, L.R., Meharg, K.J., Dundas, D., and Taylor, K.T. (2001). J. Phys. B
34, L69.
66. Parker, J.S., Moore, L.R., Dundas, D., and Taylor, K.T. (2000). J. Phys. B 33, L691.
67. Parker, J., Taylor, K., Clark, C., and Blodgett-Ford, S. (1996). J. Phys. B 29, L33.
68. Dundas, D., Taylor, K., Parker, J., and Smyth, E. (1999). J. Phys. B 32, L231.
69. Colgan, J., Pindzola, M.S., and Robicheaux, E (2001). J. Phys. 34, L457.
70. LaGattuta, K., and Cohen, J. (1998). J. Phys. B 31, 5281.
71. Feeler, C., and Olson, R. (2000). J. Phys. 33, 1997.
72. Cassimi, A., Duponchel, S., Flechard, X., Jardin, P., Sortais, P., Hennecart, D., and Olson, R.
(1996). Phys. Rev. Lett. 76, 3679.
73. D/Srner, R., Ullrich, J., Schmidt-B6cking, H., and Olson, R. (1989). Phys. Rev. Lett. 63, 147.
74. D6rner, R., Mergel, V., Ali, R., Buck, U., Cocke, C., Froschauer, K., Jagutzki, O., Lencinas, S.,
Meyerhof, W., Nfittgens, S., et al. (1994). Phys. Rev. Lett. 72, 3166.
75. D6rner, R., Khemliche, H., Prior, M., Cocke, C., Gary, J., Olson, R., Mergel, V., Ullrich, J.,
and Schmidt-B6cking, H. (1996b). Phys. Rev. Lett. 77, 4520.
76. Frohne, V., Cheng, S., Ali, R., Raphaelian, M., Cocke, C., and Olson, R. (1993). Phys. Rev.
Lett. 71,696.
77. Mergel, V., D6rner, R., Achier, M., Khayyat, K., Lencinas, S., Euler, J., Jagutzki, O.,
Nfittgens, S., Unverzagt, M., Spielberger, L., et al. (1997). Phys. Rev. Lett. 79, 387.
78. Moshammer, R., Ullrich, J., Unverzagt, M., Schmidt, W., Jardin, P., Olson, R., Mann, R.,
D6rner, R., Mergel, V., Buck, U., et al. (1994). Phys. Rev. Lett. 73, 3371.
79. Moshammer, R., Ullrich, J., Kollmus, H., Schmitt, W., Unverzagt, M., Jagutzki, O., Mergel, V.,
Schmidt-B6cking, H., Mann, R., Woods, C., et al. (1996b). Phys. Rev. Lett. 77, 1242.
80. Peterson, E., and Bucksbaum, P. (2001). Phys. Rev. 64, 053405.
81. Moshammer, R., Feuerstein, B., Fischer, D., Dorn, A., Schr6ter, C., Deipenwisch, J., Lopez-
Urrutia, J., H6hr, C., Neumayer, P., Ullrich, J., et al. (2001). Opt. Express 8, 358. On-line:
http://www.opticsexpress.org/oearchive/source/30944.htm.
82. Mergel, V., Achier, M., D6rner, R., Khayyat, K., Kambara, T., Awaya, Y., Zoran, V., Nystr6m, B.,
Spielberger, L., McGuire, J., et al. (1998). Phys. Rev. Lett. 80, 5301.
83. Achler, M., Mergel, V, Spielberger, L., R. D6rner, Y.A., and Schmidt-B6cking, H. (2001).
J. Phys. B 34, L965.
84. Moshammer, R., Feuerstein, B., Urrutin, I.C.L., Dorn, A., Fischer, D., Schr6ter, C., Schmitt, W.,
Ullrich, J., Rottke, H., Trump, C., et al. (2002). Phys. Rev. A 65, 035401.
85. Goreslavskii, S., Popruzhenko, S., Kopold, R., and Becker, W. (2001). Phys. Rev. 64, 053402.
86. Weber, T., Weckenbrock, M., Staudte, A., Hattass, M., Spielberger, L., Jagutzki, O., Mergel, V.,
H. Schmidt-B6cking, G.U., Giessen, H., Br/iuning, H., et al. (2001b). Opt. Express 7(9), 368.
On-line: http://www.opticsexpress.org/oearchive/source/30623.htm.
87. D6rner, R., Mergel, V., Br/iuning, H., Achier, M., Weber, T., Khayyat, K., Jagutzki, O.,
Spielberger, L., Ullrich, J., Moshammer, R., et al. (1998). In "Atomic processes in Plasmas"
(E. Oks, M. Pindzola, Eds.). AlP Conf. Proc. 443, 137.
34 R. D 6 r n e r et al. [X

88. Becker, A., and Faisal, E (1994). Phys. Rev. A 50, 3256.
89. Dorn, A., Moshammer, R., Schr6ter, C., Zouros, T., Schmitt, W., Kollmus, H., Mann, R., and
Ullrich, J. (1999). Phys. Rev. Lett 82, 2496.
90. Dorn, A., Kheifets, A., Schr6ter, C.D., Najjari, B., H6hr, C., Moshammer, R., and Ullrich, J.
(2001). Phys. Rev. Lett 86, 3755.
91. Taouil, I., Lahmam-Bennani, A., Duguet, A., Duguet, A., and Avaldi, L. (1998). Phys. Rev.
Lett 81, 4600.
92. Bapat, B., Keller, S., Moshammer, R., Mann, R., and Ullrich, J. (2000). J. Phys. B 33, 1437.
93. Moshammer, R., Ullrich, J., Schmitt, W., Kollmus, H., Cassimi, A., D6rner, R., Dreizler, R.,
Jagutzki, O., Keller, S., Lfidde, H.-J., et al. (1997). Phys. Rev. Lett. 79, 3621.
94. von Weiz/icker, C. (1934). Z. Phys. 88, 612.
95. Williams, E. (1934). Phys. Rev. 45, 729.
96. Keller, S., Lfidde, H., and Dreizler, R. (1997). Phys. Rev. A 55, 4215.
97. Voitkiv, A., and Ullrich, J. (2001). J. Phys 43, 1673.
98. Keller, S. (2000). J. Phys. B 33, L513.
99. Unverzagt, M., Moshammer, R., Schmitt, W., Olson, R., Jardin, R, Mergel, V., Ullrich, J., and
Schmidt-B6cking, H. (1996). Phys. Rec. Lett. 76, 1043.
100. Coplan, M.A., et al. (1994). Rev. Mod. Phys. 66, 985, and references therein.
101. Lein, M., Gross, E., and Engel, V. (2000b). J. Phys. B 33, 433.
102. Pindzola, M., and Robicheaux, E (1998). Phys. Rec. A 57, 318.
103. Haan, S.L., Hoekema, N., Poniatowski, S., Liu, W.-C., and Eberly, J.H. (2000). Opt. Express
7, 29. On-line: http://www.opticsexpress.org/oearchive/21863.htm.
104. Muller, H. (2001). Opt. Express 8, 417. On-line: http'//www.opticsexpress.org/oearchive/source/
30932.htm.
105. Lein, M., Gross, E., and Engel, V. (2001b). Phys. Rev. A 64, 023406.
A D V A N C E S IN ATOMIC, M O L E C U L A R , A N D O P T I C A L PHYSICS, VOL. 48

A B 0 VE- THRESHOLD IONIZATION:


FROM CLA SSICA L FEATURES TO
QUA N T UM EFFE C TS
W. BECKER l, E GRASBON 2, R. KOPOLD 1, D.B. MILOSEVIC 3,
G. G. PAUL US 2 and H. WAL THER 2,4
1Max-Born-Institut, Max-Born-Str. 2A, 12489 Berlin, Germany," 2Max-Planck-Institut fffr
Quantenoptik, Hans-Kopfermann-Str. 1, 85748 Garching, Germany; 3Faculty of Science,
University of Sarajevo, Zmaja od Bosne 35, 71000 Sarajevo, Bosnia and Hercegovina;
4Ludwig-Maximilians-Universitdt Miinchen, Germany

I. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
A. Experimental Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
B. Theoretical Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
II. Direct Ionization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
A. The Classical Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
B. Quantum-mechanical Description of Direct Electrons . . . . . . . . . . . . . . . . . 44
C. Interferences of Direct Electrons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
III. Rescattering: The Classical Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5O
IV. Rescattering: Quantum-mechanical Description . . . . . . . . . . . . . . . . . . . . . . . . 53
A. Saddle-point methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
B. Connection with Feynman's path integral . . . . . . . . . . . . . . . . . . . . . . . . . . 57
C. Connection with closed-orbit theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
D. The role of the binding potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
E. A homogeneous integral equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
E Quantum orbits for linear polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
G. Enhancements in ATI spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
H. Quantum orbits for elliptical polarization . . . . . . . . . . . . . . . . . . . . . . . . . 68
I. Interference between direct and rescattered electrons . . . . . . . . . . . . . . . . . . 71
V. ATI in the Relativistic Regime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
A. Basic Relativistic Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
B. Rescattering in the Relativistic Regime . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
VI. Quantum Orbits in High-order Harmonic Generation . . . . . . . . . . . . . . . . . . . . 76
A. The Lewenstein Model of High-order Harmonic Generation . . . . . . . . . . . . . 77
B. Elliptically Polarized Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
C. H H G by a Two-color Bicircular Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
D. H H G in the Relativistic Regime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
VII. Applications of ATI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
A. Characterization of High Harmonics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
B. The "Absolute Phase" of Few-cycle Laser Pulses . . . . . . . . . . . . . . . . . . . . 90
VIII. Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
IX. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

35 Copyright 9 2002 Elsevier Science (USA)


All rights reserved
ISBN 0-12-003848-X/ISSN 1049-250X/02 $35.00
36 W. Becker et al. [I

I. I n t r o d u c t i o n

With the discovery of above-threshold ionization (ATI) by Agostini et al.


(1979) intense-laser atom physics entered the nonperturbative regime. These
experiments recorded the photoelectron kinetic-energy spectra generated by laser
irradiation of atoms. Earlier experiments had measured total ionization rates
by way of counting ions, and the data were well described by lowest-order
perturbation theory (LOPT) with respect to the electron-field interaction. This
LOPT regime was already highly nonlinear (see, e.g., Mainfray and Manus,
1991), the lowest order being the minimal number N of photons necessary for
ionization. An ATI spectrum consists of a series of peaks separated by the photon
energy, see Fig. 1. They reveal that an atom may absorb many more photons than
the minimum number N, which corresponds to LOPT.
In the 1980s, the photon spectra emitted by laser-irradiated gaseous media
were investigated at comparable laser intensities and were found to exhibit peaks
at odd harmonics of the laser frequency (McPherson et al., 1987; Wildenauer,
1987). The spectra of this high-order harmonic generation (HHG) display a
plateau (Ferray et al., 1988), i.e., the initial decrease of the harmonic yield
with increasing harmonic order is followed by a flat region where the harmonic
intensity is more or less independent of its order. This plateau region terminates
at some well-defined order, the so-called cutoff.

FIG. 1. Photoelectron spectrumin the above-threshold-ionization(ATI) intensityregime. The series


of peaks corresponds to the absorption of photons in excess of the minimumrequired for ionization.
The figure shows the result of a numerical solution of the SchrSdinger equation (Paulus, 1996).
I] ATI: CLASSICAL TO QUANTUM 37

A simple semiclassical model of HHG was furnished by Kulander et al. (1993)


and by Corkum (1993): At some time, an electron enters the continuum due to
ionization. Thereafter, the laser's linearly polarized electric field accelerates the
electron away from the atom. However, when the field changes direction, then,
depending on the initial time of ionization, it may drive the electron back to
its parent ion, where it may recombine into the ground state, emitting its entire
energy - the sum of the kinetic energy that it acquired along its orbit plus the
binding energy- in the form of one single photon. This simple model beautifully
explains the cutoff energy of the plateau, as well as the fact that the yield of HHG
strongly decreases when the laser field is elliptically polarized. In this event, the
electron misses the ion. This model is often referred to as the simple-man model.
The model suggests (Corkum, 1993) that the electron, when it recollides with
the ion, may very well scatter off it, either elastically or inelastically. Elastic
scattering should contribute to ATI. Indeed, the corresponding characteristic
features in the angular distributions were observed by Yang et al. (1993), and
an extended plateau in the energy spectra due to this mechanism, much like
the plateau of HHG, was identified by Paulus et al. (1994c). Under the same
conditions, a surprisingly large yield of doubly charged ions was recorded
(l'Huillier et al., 1983; Fittinghoff et al., 1992) that was incompatible with a
sequential ionization process. A potential mechanism causing this nonsequential
ionization (NSDI) is inelastic scattering. It was only recently, however, that
this inelastic-scattering scenario emerged as the dominant mechanism of NSDI,
through analysis of measurements of the momentum distribution of the doubly
charged ions (Weber et al., 2000a,b; Moshammer et al., 2000).
The semiclassical rescattering model sketched above has proved invaluable
in providing intuitive understanding and predictive power. It was embedded in
fully quantum-mechanical descriptions of HHG (Lewenstein et al., 1994; Becker
et al., 1994b) and ATI (Becker et al. 1994a; Lewenstein et al., 1995a). This
work has led to the concept of "quantum orbits," a fully quantum-mechanical
generalization of the classical orbits of the simple-man model that retains the
intuitive appeal of the former, but allows for interference and incorporates
quantum-mechanical tunneling. The quantum orbits arise naturally in the context
of Feynman's path integral (Sali6res et al., 2001).
This review will concentrate on ATI and the various formulations of the
rescattering model, from the simplest classical model to the quantum orbits for
elliptical polarization. Alongside with theory, we will provide a review of the
experimental status of ATI. We also give a brief survey of recent applications of
ATI. High-order harmonic generation is considered only insofar as it provides
further illustrations of the concept and application of quantum orbits. We do
not deal with the important collective aspects of HHG, and no attempt is made
to represent the vast literature on HHG. For this purpose, we refer to the
recent reviews by Sali~res et al. (1999) and Brabec and Krausz (2000). Earlier
38 W. Becker et al. [I

reviews pertinent to ATI have been given by Eberly et al. (1991), Mainfray and
Manus (1991), DiMauro and Agostini (1995), and Protopapas et al. (1997a). The
entire field of laser-atom physics has been succinctly surveyed by Kulander and
Lewenstein (1996) and, recently, by Joachain et al. (2000). Both of these reviews
concentrate on the theory. Nonsequential double ionization is well covered in a
recent focus issue of Optics Express, Vol. 8.

A. EXPERIMENTAL METHODS

ATI is observed in the intensity regime 1012W/cm 2 to 1016W/cm 2. At such


intensities, atoms may ionize so quickly that complete ionization has taken
place before the laser pulse has reached its maximum. This calls, on the
one hand, for atoms with high ionization potential (i.e. the rare gases) and,
on the other, for ultrashort laser pulses. Owing to the rapid progress in
femtosecond laser technology, in particular since the invention of titanium-
sapphire (Ti:Sa) femtosecond lasers (Spence et al., 1991), generation of laser
fields with strengths comparable to inner atomic fields has become routine. The
prerequisite of detailed investigations of ATI, however, has been the development
of femtosecond laser systems with high repetition rate. Owing to the latter,
the detection of faint but qualitatively important features of ATI spectra with
low statistical noise has become possible. This holds, in particular, if multiply
differential ATI spectra are to be studied, such as angle-resolved energy spectra,
or spectra that are very weak, such as for elliptical polarization or outside the
classically allowed regions. State-of-the-art pulses are as short as 5 fs (Nisoli
et al., 1997) and repetition rates reach 100 kHz (Lindner et al., 2001).
The most widespread method of analyzing ATI electrons is time-of-flight
spectroscopy. When the laser pulse creates a photoelectron, it simultaneously
triggers a high-resolution clock. The electrons drift in a field-free flight tube
of known length towards an electron detector, which then gives the respective
stop pulses to the clock. Now, their kinetic energy can easily be calculated from
their time of flight. This approach has by far the highest energy resolution and
is comparatively simple. However, the higher the laser repetition rate, the more
demanding becomes the data aquisition.
Other approaches include photoelectron imaging spectroscopy (Bordas et al.,
1996), which is able to record angle-resolved ATI spectra, and so-called cold-
target recoil-ion-momentum spectroscopy (COLTRIMS) technology (D6rner
et al., 2000), which is capable of providing complete kinematic determination of
the fragments of photoionization, i.e. the electrons and ions. It requires, however,
conditions such that no more than one atom is ionized per laser shot. Therefore,
it can take particular advantage of high laser repetition rates. The disadvantage
of COLTRIMS is the poor energy resolution for the electrons and the extracting
technology.
I] ATI: CLASSICAL TO QUANTUM 39

~ ~ ~LA b) Calculation, 66. I

.~. .L

,,,,l,,,,l,,,,iiiiii-,,,,l,,,, i,,,,I,,,,I,,,,
I0 20 KE (eV) :30 40

FIG. 2. (a) Measured and (b) calculated photoelectron spectrum in argon for 800 nm, 120 fs pulses
at the intensities given in TW/cm 2 in the figure (10Up = 39eV). From Nandor et al. (1999).

B. THEORETICAL METHODS

The single-active-electron approximation (SAE) replaces the atom in the laser


field by a single electron that interacts with the laser field and is bound by an
effective potential so optimized as to reproduce the ground state and singly
excited states. Up to now, in single ionization no qualitative effect has been
identified that would reveal electron-electron correlation. The SAE has found
its most impressive support in the comparison of experimental ATI spectra in
argon with spectra calculated by numerical solution of the three-dimensional
time-dependent Schr6dinger equation (TDSE) (Nandor et al., 1999); see Fig. 2.
The agreement between theory and experiment is equally remarkable as it has
been achieved for low-order ATI in hydrogen; cf. D6rr et al. (1990) for the
Sturmian-Floquet calculation and Rottke et al. (1990) for the experiment. For
helium, a comparison of total ionization rates with and without the SAE in the
above-barrier regime has lent further support to the SAE (Scrinzi et al., 1999).
Numerical solution of the one-particle TDSE in one dimension was instru-
mental for the understanding of ATI in its early days; for a review, see Eberly
et al. (1992). For the various methods of solving the TDSE in more than one
dimension we refer to Joachain et al. (2000). Comparatively few papers have
dealt with high-order ATI in three (that is, in effect, two) dimensions. This is
particularly challenging since the emission of plateau electrons is caused by
very small changes in the wave function, and the large excursion amplitudes
of free-electron motion in high-intensity low-frequency fields necessitate a large
spatial grid. This is exacerbated for energies above the cutoff and for elliptical
polarization. Expansion of the radial wave function in terms of a set of B-spline
functions was used by Paulus (1996), by Cormier and Lambropoulos (1997),
and by Lambropoulos et al. (1998). Matrix-iterative methods were employed by
40 W. Becker et al. [II

Nurhuda and Faisal (1999). The most detailed calculations have been carried
out by Nandor et al. (1999) and by Muller (1999a,b, 2001a,b). The techniques
are detailed by Muller (1999c). To our knowledge, no results for high-order ATI
for elliptical polarization based on numerical solution of the TDSE have been
published to this day.
Recently, numerical solution of the TDSE for a two-dimensional model
atom by means of the split-operator method has been widely used in order to
investigate various problems such as elliptical polarization (Protopapas et al.,
1997b), stabilization (Patel et al., 1998; Kylstra et al., 2000), magnetic-drift
effects (V~izquez de Aldana and Roso, 1999; V~izquez de Aldana et al., 2001)
and various low-order relativistic effects (Hu and Keitel, 2001).
Efforts to deal with the two-electron TDSE and, in particular, to compute
double-electron ATI spectra are under way (Smyth et al., 1998; Parker et al.,
2001; Muller, 2001c). In one dimension for each electron, such spectra have
been obtained by Lein et al. (2001).
An approach that is almost complementary to the solution of the TDSE starts
from the analytic solution for a free electron in a plane-wave laser field, the so-
called Volkov solution (Volkov, 1935), which is available for the Schr6dinger
equation as well as for relativistic wave equations, and considers the binding
potential as a perturbation. The stronger the laser field, the lower its frequency,
and the longer the pulse becomes, the more demanding is the solution of the
TDSE, and the more the Volkov-based methods play out their strengths.
This review concentrates on methods of the latter variety.

II. Direct Ionization

A. THE CLASSICAL MODEL

The classical model of strong-field effects divides the ionization process into
several steps (van Linden van den Heuvell and Muller, 1988; Kulander et al.,
1993; Corkum, 1993; Paulus et al., 1994a, 1995). In a first step, an electron
enters the continuum at some time to. If this is caused by tunneling (Chin et al.,
1985; Yergeau et al., 1987; Walsh et al., 1994), the corresponding rate is a highly
nonlinear function of the laser electric field ,~(t0). For example, the quasistatic
Ammosov-Delone-Krainov (ADK) tunneling rate (Perelomov et al., 1966a,b;
Ammosov et al., 1986) is given by (in atomic units)

2n*-Iml-I
r'(t) = A E w [~7(t)l exp ' 3[E(t) I (1)

where ,~(t) is the instantaneous electric field, EIp > 0 is the ionization potential of
the atom, n* = Z~ 2x/2E~p is the effective principal quantum number, Z is the charge
II] ATI: CLASSICAL TO QUANTUM 41

of the nucleus, and m is the projection of the angular momentum on the


direction of the laser polarization. The constant A depends on the actual and
the effective quantum numbers. The rate F(t) was derived on the assumption that
the laser frequency is low, excited states play no role, and the Keldysh parameter
y = v/Exp/2Up is small compared with unity [Up is the ponderomotive potential
of the laser field; see Eq. (3) below]. Instantaneous rates that hold for arbitrary
values of the Keldysh parameter have been presented by Yudin and Ivanov
(2001b). For the discussion below, the important feature of the instantaneous
ionization rate F(t) is that it develops a sharp maximum at times when the field
,~(t) reaches a maximum.
The classical model considers the orbits of electrons that are released into the
laser-field environment at some time to. The contribution of such an orbit will
be weighted according to the value of the rate F(t0). Classically, an electron born
by tunneling will start its orbit with a velocity of zero at the classical "exit of the
tunnel" at r ~ Eip/[eF~ l, which, for strong fields, is a few atomic units away from
the position of the ion. We will, usually, ignore this small offset and have the
electronic orbit start at x(t = to) = 0 (the position of the ion) with x(t = to) = 0.
If, after the ionization process, the interaction of the electron with the ion
is negligible, we speak of a "direct" electron, in contrast to the case, to be
considered below in Sects. III and II.B, where the electron is driven back to the
ion and rescatters. An unambiguous distinction between direct and rescattered
electrons, in particular for low energy, is possible only in theoretical models.

A.1. Basic kinematics


The second step of the classical model is the evolution of the electron trajectory
in the strong laser field. During this step, the influence of the atomic potential
is neglected. For an intense laser field, the electron's oscillation amplitude is
much larger than the atomic diameter, and so this is well justified. For a vector
potential A(t) that is chosen so that its cycle average (A(t))r is zero, the
electron's velocity is

mv(t) = e(A(t0) - A(t)) -- p - eA(t), (2)

where e = - l e l is the electron's charge. The velocity consists of a constant term


p - eA(t0), which is the drift momentum measured at the detector, and a term
that oscillates in phase with the vector potential A(t). The kinetic energy of this
electron, averaged over a cycle T of the laser field, is
p2 e2
E = - 2 m (v(t)Z)v = ~mm + ~mm (A(t)2)T -- Edrifl + Up. (3)

The ponderomotive energy


e2
Up = z-- (A(t)z)v, (4)
zm- -
42 W. Becker et al. [II

viz. the cycle-averaged kinetic energy of the electron's wiggling motion, is


frequently employed to characterize the laser intensity. A useful formula is

Up [eV] = 0.09337I[W/cm 2]/~2[m]

for a laser with intensity I and wavelength/l. If the electron is to have a nonzero
velocity v0 at time to, one has to replace eA(t0) by eA(t0)+ mv0 = p in the
velocity (2).
Most of the time, we will be concerned with the monochromatic elliptically
polarized laser field (-1 ~< ~ <~ 1)

coA
F_,(t) = ~ (~ sin cot - ~S' cos cot) (5)
V/1 +~2

with vector potential

A
A(t) = ~ (~ cos cot + ~S' sin cot) (6)
V/1 +~2

and ponderomotive energy Up = (eA)e/4m. The drift energy Edrif t -- (eA(to))2/2m


is restricted to the interval

2~ 2 2
1+ ~2 Up ~ Edrift <~ 1 + ~2 Up. (7)

For linear polarization, it can acquire any value between 0 and 2Up, while
for circular polarization it is restricted to the value Up. Quantum mechanics
considerably softens these classical bounds. However, these bounds are useful as
benchmarks in the analysis of experimental spectra (Bucksbaum et al., 1986),
in particular for high intensity (Mohideen et al., 1993; Reiss, 1996) or low
frequency (Gallagher and Scholz, 1989).
In general, it is important to recall that the ionization probability depends
on the electric field, while the drift momentum p = eA(t0) is proportional
to the vector potential, both at the time to of ionization. The probability of
a certain drift momentum is weighted with the ionization rate at time to. The
electron is preferably ionized when the absolute value of the electric field is
near its maximum. Then, for linear polarization, the vector potential and, hence,
the drift momentum are near zero. In order to reach the maximal drift energy
of 2Up, the electron must be ionized when the electric field is zero and, hence,
the ionization rate is very low. This explains the pronounced drop of the ATI
electron spectrum for increasing energy, see Fig. 1. Sometimes, this interplay
between the instantaneous ionization rate and the drift momentum has surprising
consequences, notably for fields where the connection between the two is less
straightforward than for a linearly polarized sinusoidal field, e.g. for a two-color
II] ATI: CLASSICAL TO QUANTUM 43

100 ' ' ' ~ ' =104~. . . . . lin.'pollt


--~ 16.1eV f k~ E= 103[ ~ ]
o I \ 8
o,o, 1
._~ 101
=>" f \ -~1ooI . . . . . . . . ,~
_O9 p \ 0 10 20 30 40 50
o v]
_..m
(D
"O
(D
.N_
E
t-
r
O

' -olso ' o.oo ' o.5o ' .oo


ellipticity

FIG. 3. Dependence of the photoelectron yield as a function of the ellipticity ~ of the elliptically
polarized laser field (5) for electrons with an energy of 16.1 eV. Only electrons emitted parallel to
the major axis of the polarization ellipse are recorded. The ATI spectrum corresponding to linear
polarization (~ = 0) is shown in the inset. The laser intensity was 0.8x 1014 W/cm 2 at a wavelength
of 630 nm. The figure illustrates the dodging effect mentioned in Sect. II.A.I: ionization primarily
takes place when the electric field is near an extremum. For elliptical polarization, the electric field
then points in the direction of the major axis of the polarization ellipse, and the vector potential in
the direction of the minor axis. Hence, the electron's drift momentum p = eA(t0) is in the direction of
the minor axis. It is the larger, the larger the ellipticity ~ is. Consequently, emission in the direction
of the large component of the field decreases with increasing ellipticity: the electron dodges the
strong component of the field. The effect vanishes when circular polarization is approached and the
distinction between the major and minor axes disappears. From Paulus et al. (1998).

field (Paulus et al., 1995; Chelkowski and Bandrauk, 2000; Ehlotzky, 2001).
Another illustration is the dodging phenomenon for the direct ATI electrons
in an elliptically polarized laser field (Paulus et al., 1998; Goreslavskii and
Popruzhenko, 1996; Mur et al., 2001), see Fig. 3.
We have tacitly assumed that pulses are short enough to pass over the electron
before it has a chance to experience the transverse spatial gradient of the focused
pulse. In this event, the spatial dependence of the vector potential A(t) can
truly be neglected. Hence the drift momentum p is conserved and is indeed
the momentum recorded at the detector outside the field (Kibble, 1966; Becker
et al., 1987). The wiggling energy Up is lost or, in a self-consistent description,
returned to the field when the electron is left behind by the trailing edge of
the pulse. In the opposite case, where the electron escapes from the pulse
perpendicularly to its direction of propagation, the wiggling energy is converted
into drift energy (Kruit et al., 1983, Muller et al., 1983). This limit is also
realized in ATI by microwave ionization of Rydberg states (Gallagher and
44 W. Becker et al. [II

Scholz, 1989). The effects of a space-dependent ponderomotive potential Up


were observed in the "surfing" experiments of Bucksbaum et al. (1987).

B. QUANTUM-MECHANICAL DESCRIPTION OF DIRECT ELECTRONS

There is an enormous body of work on the quantum-mechanical description of


laser-induced ionization. For reviews, we refer to Delone and Krainov (1994,
1998). Here we want to concentrate on the analytical approach dating back to
Keldysh (1964) and Perelomov, Popov and Terent'ev (Perelomov et al. 1966a,b;
Perelomov and Popov, 1967). The goal is to find a suitable approximation to the
probability amplitude for detecting an ATI electron with drift momentum p that
originates from laser irradiation of an atom that was in its ground state ]~P0)
before the laser pulse arrived:

Mp =
t ~ e~,
lim
t t ~ --(X3
(lpp(t) Iu(t, t')l ~p0(t')). (8)

Here, U(t, t') is the time-evolution operator of the Hamiltonian (h = 1)

1 ~r2
H ( t ) = -2--mm - er . s + V(r), (9)

which includes the atomic binding potential V(r) and the interaction - e r . g ( t )
with the laser field. Furthermore, we introduce the Hamiltonians for the atom
without the field and for a free electron in the laser field without the atom,

Ha - Hatom -- 1 i~72 + V(r), (10)


2m
1 V2
H f ( t ) - Hfield(t) = -~mm - e r . g(t). (1 1)

The corresponding time-evolution operators are denoted by Ua and Uf, re-


spectively. In Eq. (8), [~pp) and [~P0) are a scattering state with asymptotic
momentum p and the ground state, respectively, of the atomic Hamiltonian Ha.
The eigenstates of the time-dependent Schr6dinger equation with the Hamilto-
nian H f ( t ) are known as the Volkov states and are of compact analytical form.
In the length gauge, one has

Iw~vv)(t)) - IP - eA(t)) e-isp(t), (12)

with ]p - eA(t)) a plane-wave state [(rip - eA(t)) = (2:r) -3/2 exp i[(p - eA(t)) 9r]
and
Sp(t) = ~ dr [p - eA(r)] 2. (13)

The lower limit of the integral is immaterial. It introduces a phase that does not
contribute to any observable.
II] ATI: CLASSICAL TO QUANTUM 45

The time-evolution operator U(t, t') satisfies integral equations (Dyson equa-
tions), which are convenient if one wants to generate perturbation expansions
with respect to either the interaction H t ( t ) = - e r . g(t) with the laser field,
t

U(t, t') = Ua(t, t') - i / dr U(t, r) Hi(r) U a ( r , t'), (14)


tt

or the binding potential V(r),


t

U(t, t') = Uf(t, t') - i / dr Us(t,r) V U(r, t'). (1 5)


t /
tt

Equation (14) also holds if U and Ua in the second term on the right-hand side
are interchanged. The equivalent is true of Eq. (15). With the help of the integral
equation (14), using the orthogonality of the eigenstates of Ha, we rewrite Eq. (8)
in the form

Mp = - i lim dr (~pp(t)IU(t, r)H~(r)l ~0(r)), (16)


t ----+ (N~ 0<3

which is still exact. A crucial simplification occurs if we now introduce the


strong-field approximation. That is, we make the substitutions I~pp) ---. I~ptpvv))
and U ~ Uy, with the result

Alp = - i dto (~(pVV)(t0)IHt(t0)l ~0(t0)) 9 (17)


oo

The physical content of this substitution is that, after the electron has been
promoted into the continuum at time to due to the interaction Ht(to) = - e r . F-.(to)
with the laser field, it no longer feels the atomic potential. This satisfies the above
definition of a "direct electron." Amplitudes of the type (17) are called Keldysh-
Faisal-Reiss (KFR) amplitudes (Keldysh, 1964; Perelomov et al., 1966a,b;
Faisal, 1973; Reiss, 1980); for a comparison of the various forms that exist,
see Reiss (1992).
In the amplitude (17), one may write

Y Y
- e r . e(t0) = H/(tol-Ha + V(r) = - i ~ 0 - i ~ 0 + V(r). (is)
Via integration by parts, the amplitude (17) can then be rewritten as

Mp : -i dto (w~vv)(t0)I V(r)l Wo(to)). (19)


oo

This form is particularly useful for a short-range or zero-range potential, since


these restrict the range of the spatial integration in the matrix element.
46 W. Becker et al. [II

Further evaluation of the amplitudes (17) or (19) leads to expansions


in terms of Bessel functions. For sufficiently high intensity (small Keldysh
parameter ~,), the saddle-point method (method of steepest descent) can be
invoked (Dykhne, 1960). This consists in expanding the phase of the integrand
about the points where the phase is stationary. Given the form of the Volkov
wave functions (12) and the time dependence of the ground-state wave function,
]~P0(t)) = exp(iEipt)]~po), this amounts to determining the solutions of

d
[Eipt + Sp(t)] = EIp + ~ [ p - eA(t)] 2 = 0. (20)
dt
Let us consider a periodic (not necessarily monochromatic) vector potential with
period T = 23:/co. In terms of the solutions ts of Eq. (20), the amplitude (19)
can then be written as (Gribakin and Kuchiev, 1997; Paulus et al., 1998)

/lip oc Z 6 (p2
-~-mm+ Eip + mp - n oJ
)
n
(21)
2~i
• ~s S~'(ts) ei[Ewts+Sp(ts)](p-eA(ts)lV[lP~

where S~~ denotes the second derivative of the action (13) with respect to time.
The sum over s extends over those solutions of Eq. (20) within one period of
the field (e.g. such that 0 ~< Re t, < T) that have a positive imaginary part.
Obviously, the saddle points are complex unless Eip = 0. For Eip - 0, we retrieve
the classical drift momentum (2) provided p is such that p = eA(t) at some time t.
The imaginary part of to can be related to a tunneling time (Hauge and Stovneng,
1989).
In Eq. (21), the ionization amplitude is represented as the coherent sum over all
saddle points within one period of the field. The fact that the spectrum consists
of the discrete energies
p2
Ep - 2m - n m - U p - Eip (22)

can be attributed to interference of the contributions from different periods.


This interference is destructive, unless the energy corresponds to one of the
discrete peaks (22). Depending on the shape of the vector potential A(t) and
its symmetries, there will be several solutions ts (for a sinusoidal field, there are
two in the upper half plane and two in the lower, which are complex conjugate
to the former) within one period of the field. Their interference creates a beat
pattern in the calculated spectrum. This is, however, difficult to observe due to
its sensitive dependence on the laser intensity, which is not very well controlled
in an experiment so that the interference effects are usually washed out.
II] ATI: C L A S S I C A L TO Q U A N T U M 47

Interferences also exist for an elliptically polarized laser field for fixed electron
momentum as a function of the ellipticity. Since, in experiments, the ellipticity is
better defined than the intensity, these interferences have been observed (Paulus
et al., 1998); see next subsection.
The amplitude (19) admits a vector potential A(t) of arbitrary shape; it is
by no means restricted to a monochromatic field of infinite extent. For a pulse
of finite extent, the saddle points are still determined by Eq. (20). They have,
however, no longer any periodicity. Hence, the discreteness of the spectrum is
lost. Interference from different parts o f the pulse may lead to unexpected effects
(Raczyfiski and Zaremba, 1997).
While for an infinitely long monochromatic pulse the spectrum is symmetric
with respect to p ~ - p , this forward-backward symmetry no longer holds for
a finite pulse. Analysis of the spatial asymmetry of the spectrum may aid in
determination o f the pulse length or the absolute carrier phase (Dietrich et al.,
2000; Hansen et al., 2001; Paulus et al., 2001b); see Sec. VII.B.

C. INTERFERENCESOF DIRECT ELECTRONS

For linear polarization and a drift m o m e n t u m p = p~ with [p] <~ eA, there
are two possible ionization times cot01 = Jr/2 + 6 and O)to2 = 3 3 r / 2 - 6. The
corresponding classical orbits are illustrated in Fig. 4. As discussed above, while
A(t01) = A(t02), the field satisfies s = -•(t02). Hence, electrons ionized at
t01 and t02 depart in opposite directions right after the instant of ionization. As
illustrated in Fig. 4, the electric field changes sign soon after t01. Hence, the
electron ionized at this time turns around at a later time and acquires the same
drift momentum as the electron ionized at time t02, which keeps its original
direction. We expect quantum-mechanical interference of the contribution of
these two ionization channels.

to~ = Jr/2 + 6

0 2 4 6 8 0 2 4 6 8
rot~Jr ~ot # r

FIG. 4. Classical trajectories (dashed lines) of electrons having the same drift momentum. The
solid line is the effective potential V(x)- exg(t) at times t01 (left) and t02 (right). The electron
ionized at to] is turned around by the field shortly after ionization. In contrast, the electron ionized
at t02 maintains its original direction. This is a strongly simplified picture of the physics underlying
the interferences of direct electrons.
48 W. B e c k e r et al. [II

3.0 a)
E
l ~=0.78
.,qo
0.0 0.2 0.4 0.6 0.8 1.0
~2.0 ~=o.7~..-.--T---....~=o.7
E
,,....=

1.0
, ,

rt/2 1"1; 3rd2 0.25 0.50 0.75


Re[o~ts]

F1G. 5. (a) Positions of the saddle points Ogts in the upper half of the complex ~ot plane in the
interval 89 ~<Rewt <~ 3Jr, calculated from Eq. (23) forE= 17~o-Ew-Up, where EIp = 15.76 eV,
Up = 3.68 eV, and h w - 1.96eV. The arrows indicate the motion of the saddle points for increasing
ellipticity ~. The two branches meet at ~0 = 0.755. For several values of the ellipticity, insets depict
the ellipse traced out by the electric-field vector, and the positions of the latter at the emission
times Re ts are marked by solid dots. (b) The function Re ~, which determines the magnitude of the
amplitude Mp. The existence of the valley near the ellipticity ~0 is related to the effect of dodging,
illustrated in Fig. 3. (c) The function cosZ(Im 9 + ~p), whose oscillations are caused by constructive
and destructive interference. The essential physics behind this interference is sketched in Fig. 4.
From Paulus et al. (1998).

For elliptical polarization, classically, there is at m o s t one ionization time


for given drift m o m e n t u m . However, Eq. (20) for the complex saddle points o f
the quantum-mechanical amplitude always has more than one solution. For the
field (6) and p = ~ i , the solutions are

cos Ots = v/1 + I-v7 +


v / c ~ 2 - ~ip(1 - ~2) - ~2~] , (23)

where ~ = ( 1 - ~2)/(1 + ~2), ~ = E / 2 U p , and eip = E i p / 2 U p . Obviously, the


solutions ts come in complex conjugate pairs. Those in the upper half-plane
enter the amplitude (21). The solutions are plotted in Fig. 5. F o r - ~ 0 < ~ ~< ~0
[with ~0 given by the zero of the square root in Eq. (23)], the second square
root on the right-hand side o f Eq. (23) is imaginary, and the solutions are
symmetric with respect to Re tOts = Jr. For I~l ~> ~0, this square root is
real, and all solutions have Re cot~ - Jr. This has important consequences for
the saddle-point amplitude (21). In the first case, both solutions contribute to
the amplitude and an interference pattern results. This corresponds to the case
o f linear polarization discussed above. In the second case, inspection o f the
integration contour in the complex plane shows that it has to be routed only
through the one solution that is closest to the real axis (Leubner, 1981). Hence,
II] ATI: CLASSICAL TO QUANTUM 49

10 0 , , ,

f - o - 8.23eV ~104
~ 10.03eV
11.38eV
~ ,,d %, ~
,o~2~ 4~a ~%,
o=o103
'-

t 48.88eV

"~"~ -~102 i"


1~ 10 20 30 40 50 60 70

c--

ln-2 , , ,
"-1 .00 -0.50 0.00 0.50 1.00
ellipticity

FIG. 6. ATI spectra in xenon for an intensity of 1.2x 1014 W/cm 2 for various energies in the
direction of the large component of the elliptically polarized field as a function of the ellipticity.
The inset shows the energy spectrum for linear polarization. The three traces for the lower energies
display the interference phenomenon of the direct electrons discussed in Sect. II.C; the one for the
highest energy belongs to a plateau electron. The interference dips are related to Fig. 5c. From
Paulus et al. (1998).

there is no interference. In the first case, the amplitude can be written in the
form Mp ~ exp(Re ~)cos(Im 9 + ~p), in the second case the cosine is absent.
The two arguments are also plotted in Fig. 5.
The corresponding interferences have been observed by Paulus et al. (1998);
see Fig. 6. They are responsible for the undulating pattern in the ellipticity
distribution, which moves to smaller ellipticity for increasing energy. The same
tendency can be observed in the numerical evaluation of the amplitude (21); see
Paulus et al. (1998) for an example.
For elliptical polarization, the KFR amplitude (19) must be applied with
due caution: it predicts fourfold symmetry of the angular distribution, while
the experimental distributions only display inversion symmetry (Bashkansky
et al., 1988). Mending this deficiency requires improved treatment of the binding
potential (Krsti6 and Mittleman, 1991). More discussion of this point has been
provided elsewhere by Becker et al. (1998), who also give further references.
The spatial dependence introduced by Coulomb-Volkov solutions in place of the
usual Volkov solutions (12) already suffices to destroy the fourfold symmetry,
and angular distributions have been calculated with their help by Jarofi et al.
(1999). However, even for a zero-range potential the fourfold symmetry is
50 W. B e c k e r et al. [III

broken provided the effects of the finite binding energy are treated beyond the
KFR approximation (Borca et al., 2001).
Very similar interferences have been seen by Bryant et al. (1987) in the
photodetachment of H- in a constant electric field. Here the electron, once
detached, has the choice of starting its subsequent travel either against or with
the direction of the electric field, by close analogy with the opposite directions
of initial travel for the ionization times t01 and/02 in the present case; see Fig. 4.
A spatial resolution of the same effect is observed by the photodetachment
microscope of Blondel et al. (1999). The theoretical description reproduces the
observed patterns. Additional bottle-neck structures develop when a magnetic
field is applied parallel to the electric field (Kramer et al., 2001).

III. Rescattering: The Classical Theory

Thus far, we have dealt with "direct" electrons, which after the first step of
ionization leave the laser focus without any additional interaction with the ion.
In this and the next section, we will consider the consequences of one such
additional encounter.
The classical model becomes much richer if rescattering effects are taken
into account. To this end, we integrate the electron's velocity (2) to obtain its
trajectory
x(t) = -- (t - to) A(t0) - dr A(r) . (24)
m to

The condition that the electron return to the ion at some time tl > to is
x(tl) - 0. For linear polarization in the x-direction, this implies x ( t l ) = 0, and
y ( t ) = z ( t ) - O. This yields tl as a function of to. We defer discussion of elliptical
polarization to Sect. IV.H.
When the electron returns, one of the following can happen (Corkum, 1993):
(1) The electron may recombine with the ion, emitting its energy plus the
ionization energy in the form of one photon. This process is responsible
for the plateau of high-order harmonic generation.
(2) The electron may scatter inelastically off the ion. In particular, it may dis-
lodge a second electron (or more) from the ionic ground state. This process is
now believed to constitute the dominant contribution to nonsequential double
ionization.
(3) The electron may scatter elastically. In this process, it can acquire drift
energies much higher than otherwise.
In the following, we will concentrate on this high-order above-threshold
ionization (HATI). We will, however, also briefly discuss high-order harmonic
generation in Sect. VI.
III] ATI" CLASSICAL TO QUANTUM 51

From Eq. (2), the kinetic energy of the electron at the time of its return is

e2
Eret - ~m [A(tl) - A(t0)] 2 . (25)

Maximizing this energy with respect to to under the condition that X(tl ) = 0 yields
Eret, m a x = 3.17Up for oJt0 = 108 ~ and ootl = 342 ~ (Corkum, 1993; Kulander
et al., 1993). It is easy to see that after rescattering the electron can attain a
much higher energy: Suppose that at t = tm the electron backscatters by 180 ~ so
that m v ( t l -- O) = e[A(to) - A ( t l )] j u s t before and mv(tl + O) = - e [ A ( t o ) - A(tl )]
just after the event of backscattering. Then, for t > t l , the electron's velocity is
again given by Eq. (2), but with Px = e [ 2 A ( t m ) - A(t0)] so that

e2
Ebs - ~m [2A(tl) - A(t0)] 2. (26)

Maximizing Ebs under the same condition as above yields Ebs, max = 10.007Up
(Paulus et al., 1994a) for tot0 = 105 ~ and totl = 352 ~ These values are very
close to those that afford the maximal return energy.
It is important to keep in mind that for maximal return energy or backscat-
tering energy, the electron has to start its orbit shortly after a maximum of the
electric field strength. As a consequence, it returns or rescatters near a zero of
the field, see Fig. 7. This also provides an intuitive explanation of the energy
gain through backscattering: if the electron returns near a zero of the field and
backscatters by 180 ~ then it will be accelerated by another half-cycle of the
field.
In general, the equation X(tl) = 0 for fixed to may have any number of
solutions. This becomes evident from the graphical solution presented in Fig. 7.
If the electron starts at a time to just past an extremum of the field, it returns
to the ion many times. These solutions having long "travel times" t l - to are
very important for the intensity-dependent quantum-mechanical enhancements
of the ATI plateau to be discussed in Sect. IV.G. Here we will be satisfied with
mentioning another property of the classical orbits" obviously, the return energy
will have extrema, e.g. the maximum of Ebs, max -- 10.007Up mentioned above,
which is assumed for a certain time t0,max (t0,max = 108 ~ in the example). If
we are interested in a fixed energy Ebs < Ebs, max, there are two start times that
will lead to this energy: one earlier than t0, max, the other one later. From the
graphical construction of Fig. 7 it is easy to see that the former has a longer
travel time than the latter. In the closely related case of HHG, these correspond
to the "long" and the "short" orbit (Lewenstein et al., 1995b). The cutoffs of the
solutions with longer and longer travel times are depicted in Fig. 8.
If we consider rescattering into an arbitrary angle 0 with respect to the
direction of the linearly polarized laser field, we expect a lower maximal energy
since part of the maximal energy 3.17Up of the returning electron will go into the
52 W. Becker et al. [III

[ ' I ' [ ' [

-- ....-.......
9 ~149
9 ~
.

t~ t',

- "" "'............. ......'" "'"

. . . . . .

3:/4 3:/2 33:/4 23:

FIG. 7. Graphical solution of the return time t 1 for given start time to; cf. Paulus et al. (1995):
The return condition X(tl) = 0 can be written in the form F ( t l ) - F(to) + (tl - to)Fl(to), where the
function F(t) = f t dr A ( r ) ~ sin ~ot (solid curve) is an integral of the vector potential A ( r ) ~ cos w r
(dotted curve) 9 The thick solid straight line, which is the tangent to F ( r ) at r = to, intersects F ( r )
for the first time at r = tl. The start (ionization) time to was chosen such that the kinetic energy Ere t
(Eq. 25) at the return time tl is maximal and equal to Eret,ma x = 3.17Up. The two adjacent straight
lines both yield the same kinetic energy Ere t < Eret,ma x. The figure shows that one starts earlier and
returns later while the other one starts later and returns earlier. Obviously, there can be many more
intersections with larger values of ti provided the start times are near the extrema of F ( r ) . They
correspond to the orbits with longer travel times.

transverse motion. This implies that, for fixed energy E b s , there is a cutoff in the
angular distribution; in other words, rescattering events will only be recorded for
angles such that 0 ~< 0 ~< 0max(Ebs). This is a manifestation of rainbow scattering
(Lewenstein et al., 1995a). All of this kinematics is contained in the following
equations (Paulus et al., 1994a):

Ebs = ~1 [A(t0)2 + 2A(tl) [A(tl) - A(t0)] (1 + cos 00)], (27)


A(tl)
cot 0 = cot 00 - (28)
sin 00 IA(t0) - A(h)["

Here 00 is the scattering angle at the instant of rescattering, which may have
any value between 0 and Jr, as opposed to the observed scattering angle 0 at
the detector (outside the field). In Eq. (27), the upper (lower) sign holds for
A(to) > A(h) (A(to) < A(tl)).
Pronounced lobes in the angular distributions off the polarization direction
were first observed by Yang et al. (1993), while the rescattering plateau in the
energy spectrum with its cutoff at 10Up was identified by Paulus et al. (1994b,c).
IV] ATI: C L A S S I C A L TO Q U A N T U M 53

Fie. 8. Maximum drift energy after rescattering (ATI plateau cutoff) upon the mth return to the ion
core during the ionization process. Electrons with the shortest orbits (m = 1) can acquire the highest
energy, whereas electrons that pass the ion core once before rescattering at the second return (m = 2)
have a rather low energy. Each return corresponds to two quantum orbits: the mth return corresponds
to the quantum orbits 2m + 1 and 2m + 2.

These spectra prominently display the classical cutoffs at 0max and Ebs,max.
The classical features become the better developed the higher the intensity is.
Hence, they are particularly conspicuous in the strong-field tunneling limit.
This has been shown theoretically by comparison with numerical solutions of
the Schr6dinger equation (Paulus et al., 1995) and experimentally for He at
intensities around 1015 W/cm 2. Indeed, the latter spectra show an extended
plateau for energies between 2Up and 10Up (Walker et al., 1996; Sheehy
et al., 1998). For comparatively low intensities, angular distributions have been
recorded in xenon with very high precision by Nandor et al. (1998). They also
show the effects just discussed, but with much additional structure that appears to
be attributable to quantum-mechanical interference and to multiphoton resonance
with ponderomotively upshifted Rydberg states (Freeman resonances; Freeman
et al., 1987).

IV. Rescattering: Quantum-mechanical Description


In order to incorporate the possibility of rescattering into the quantum-
mechanical description, we have to allow the freed electron once again to interact
with the ion (Lohr et al., 1997). To this end, we return to the exact equation (16)
and insert the Dyson integral equation (15). This yields two terms. Next, as we
did in Sect. II.B, we replace the exact scattering state ]~pp) by a Volkov state and
54 W. Becker et al. [IV

the exact time-evolution operator U by the Volkov-time evolution operator UU.


In other words, we disregard the interaction with the binding potential V(r),
except for the one single interaction that is explicit in the Dyson equation. This
procedure corresponds to adopting the Born approximation for the rescattering
process.
Of the two terms, the first is identical with the "direct" amplitude (17) or (19).
The second describes rescattering. Via integration by parts similar to that
explained in Eq. (18) the two terms can be combined into one,

Mp = - i dtl dto (lp;Vv)(tl)IVUf(tl,to)V I ~P0(t0)), (29)


O<3

which now describes both the direct and the rescattered electrons. The physical
content of the amplitude (29) corresponds to the recollision scenario: The
electron is promoted into the continuum at some time to; it propagates in the
continuum subject to the laser field until at the later time tl it returns to within the
range of the binding potential, whereupon it scatters into its final Volkov state.
Exact numerical evaluation of the amplitude (29) for a finite-range binding
potential is very cumbersome. For a zero-range potential, however, the spatial
integrations in the matrix element become trivial, and the computation is rather
straightforward. If the field dependence of the Volkov wave function and the
Volkov time-evolution operator is expanded in terms of Bessel functions, one of
the temporal integrations in the amplitude (29) can be carried out analytically
and yields the same 6 function as in Eq. (21), specifying the peak energies. The
remaining quadrature with respect to the travel time tl - t o has to be carried
out numerically; see Lohr et al. (1997) and Milo~evi6 and Ehlotzky (1998a),
where explicit formulas can be found; for elliptical polarization see Becker et al.
(1995) and Kopold (2001). Alternatively, the integral over the travel time may
be done first, and the integral over the return time tl is then evaluated by Fourier
transformation (Milo~evi6 and Ehlotzky, 1998b).
The relevance of the rescattering mechanism to ATI and multiple ionization
was suggested early by Kuchiev (1987) and by Beigman and Chichkov (1987).
Improvements of the customary KFR theory by including further interactions
with the binding potential were already discussed by Reiss (1980). The first
explicit calculations of angular-resolved energy spectra were carried out by
Becker et al. (1994a, 1995) and by Bao et al. (1996). Closely related rescattering
models were presented by Smirnov and Krainov (1998) and by Goreslavskii and
Popruzhenko (1999a,b, 2000).
The physics of high-order ATI is related to electron scattering at atoms in the
presence of a strong laser field. For high-order ATI, the initial state of the electron
is a wave packet created by tunneling, while for electron-atom scattering it is
a plane-wave state. This latter problem was studied theoretically by Bunkin and
Fedorov (1966) and by Kroll and Watson (1973). Corresponding experiments
IV] ATI: CLASSICAL TO QUANTUM 55

were done by Weingartshofer et al. (1977, 1983). Some quantum features of


electron scattering in intense laser fields are remarkably similar to HATI; see
Kull et al. (2000) and G6rlinger et al. (2000).

A. SADDLE-POINT METHODS

For sufficiently high intensity, the temporal integrations in the amplitude (29)
can be carried out by the saddle-point method, as in the case of the direct
amplitude (19). This procedure provides much more physical insight than
Bessel-function expansions, and establishes the connection with Feynman's path
integral, to be discussed below.
In this context, rather than taking advantage of the explicit form of the Volkov
time-evolution operator, we expand it in terms of the Volkov states (12),

S,'k I, (30.
so that the amplitude Mp is represented by the five-dimensional integral

Mp ~

with the function


f o~ dtl
/'f
dto d3k exp[iSp(tl , to, k)] mp(tl, to, k) (31 )

mp(tl, to, k) = (p - eA(t,)[ VI k - eA(tl)) (k - eA(to) lV[ ~Po). (32)


For ATI, the action

Sp(tl,to, k) - 1
2m d r [p - e A ( r ) ] 2
ftl ~
(33)
dr [k - eA(T)] 2 + dr EIp
2m
in the exponent consists of three parts, according to the three stages discussed
above.
As above in Eq. (20), we approximate the amplitude (31) by expanding the
phase (33) of the integrand about its stationary points. In this process, we assume
that the function mp(tl, to, k) depends only weakly on its arguments. Indeed, for
a zero-range potential, it is a constant. We now have to determine the stationary
points with respect to the five variables tl, to and k. They are given by the
solutions of the three conditions (Lewenstein et al., 1995a)
[k - eA(t0)] 2 = -2mEip, (34)

( t l - t0) k = dr eA(r), (35)

[k - eA(tl)]2 = [p _ eA(tl)]2. (36)


The first condition (34) attempts to enforce energy conservation at the time of
tunneling. The second condition (35) ensures that the electron returns to its
56 W. Becker et al. [IV

parent ion, and the third one (36) expresses that, on this occasion, it rescatters
elastically into its final state. In general, the saddle-point equations have several
solutions (tls, t0s, ks), (s = 1,2,...), of which only those are relevant for which
Re tls > Re t0s, such that the recollision is later than ionization, cf. the limits of
the integral in Eq. (31). The matrix element can be written as

k15)
1/2

eiSp(t's't~ tos, ks), (37)


det( O2So/Oq~S)Oq~S))j,

where ql s) (i = 1 , . . . , 5 ) runs over the five variables tls, tos and ks. As we
noted already for the direct electrons in the context of Eq. (21), the sum has
to be extended only over a subset of the solutions of the saddle-point equations
(34)-(36). However, in the present case, determining this subset may be tricky
(Kopold et al., 2000a). For a periodic field, the sum over the periods in Eq. (37)
can be carried out by Poisson's formula. This leaves a sum over the saddle points
within one period and produces a 6 function as in Eq. (21).
The computation of ATI now consists of two separate tasks. First, the solutions
of the saddle-point equations (34)-(36) have to be determined and, second,
the appropriate subset has to be inserted into expression (37). Note that we
apply the saddle-point approximation to the probability amplitude for given final
momentum p, and not to the complete wave function of the final state. This is the
reason why only few solutions contribute, while a semiclassical computation of
the wave function, which contains all possible outcomes, requires consideration
of a very large number of trajectories (van de Sand and Rost, 2000).
Since EIp > 0, the condition (34) of "energy conservation" at the time of
ionization cannot be satisfied for any real time to. As a consequence, all solutions
(tls, tos, ks) become complex. If the ionization potential EIp is zero, then, for
a linearly polarized field, the first saddle-point equation (34) implies that the
electron starts on its orbit with a speed of zero. Provided the final momentum p
is classically accessible, the resulting solutions are entirely real. They correspond
to the so-called "simple-man model" (van Linden van den Heuvell and Muller,
1988; Kulander et al., 1993; Corkum, 1993). For EIp r 0, so long as the Keldysh
parameter 72= Eip/(2Up) is small compared with unity, the imaginary parts of
the solutions of Eqs. (34)-(36) are still not too large, and the real parts are still
close to these simple-man solutions. In this case, approximate analytical solutions
to the saddle-point equations can be written down, which yield an analytical
approximation to the amplitude (31) (Goreslavskii and Popruzhenko, 2000). On
the other hand, for elliptical polarization, the solutions are always complex, even
when EIp = 0. This reflects the fact that, for any polarization other than linear,
an electron set free at any time during the optical cycle with velocity zero will
never return to the point where it was released. Equation (34) then only implies
that k - eA(t0) is a complex null vector.
IV] ATI: CLASSICAL TO QUANTUM 57

With the solutions (tls, to,, ks) (s = 1,2 .... ) of Eqs. (34)-(36), the sth quantum
orbit has the form

(t - tos)ks - fttos dr eA(r) (Re tos <~ t <~ Re tls),


rex(t) = (38)
(t I

tls)P -ftts dr eA(r) (t ~> Re tls).

We regard the orbit as a function of the real time t. The conditions x(t0) = 0
and X(tl) = 0, however, are satisfied for the complex times to and tl. As a
consequence, the quantum orbit (38) as a function of real time does not depart
from the origin but, rather, from the "exit of the tunnel." This is clearly visible in
Figs. 14, 15, 17 and 20 below. In contrast to the start time to, the return time tl is
real to a good approximation, see Fig. 10 (below). Accordingly, the orbits return
almost exactly to the origin.

B. CONNECTION WITH FEYNMAN'S PATH INTEGRAL

Any quantum-mechanical transition amplitude, such as the ionization ampli-


tude (8), can also be represented in terms of Feynman's path integral. To this
end, we recall the path-integral representation of the complete time-evolution
operator of the system atom + field,

U(rt, r't') = f(rt) ~ ~r't')


7)[r(r)]e is(''t'), (39)

where S(t, t') = ft t, dr L[r(r)], r] is the action calculated along a system path, and
the integral measure D[r(r)] mandates summation over all paths that connect
(rt) and (r' t') (see, e.g., Schulman, 1977).
The path integral (39) sums over the functional set of all continuous
paths. In the quasi-classical limit, this can be reduced to a sum over all
classical paths, which are those for which the action S(t,t') is stationary.
For quadratic Hamiltonians, this WKB approximation is exact. In our case,
motivated by the success of the classical three-step model of Sect. III, we
have reduced the exact transition amplitude to the form (31). In implementing
the strong-field approximation, we have approximated the exact action of the
system appropriately at the various stages of the process: before the initial
ionization, in between ionization and rescattering, and after rescattering, as in
the decomposition (33) of the action. This still left us with a five-dimensional
variety of paths. Out of those, finally, the saddle-point approximation (37) selects
the handful of "relevant paths" (Antoine et al., 1997; Kopold et al., 2000a;
Sali~}res et al., 2001). These are essentially the orbits of the classical model, yet
quantum mechanics is fully present: Their coherent superposition as expressed
in the form (37) allows for interference of the contributions of different orbits,
and the fact that they are complex accounts for their origin via tunneling.
58 W. Becker et al. [IV

C. CONNECTION WITH CLOSED-ORBIT THEORY

There appears to be a close similarity to the concepts of periodic-closed-orbit


theory, see, e.g., Du and Delos (1988), Gutzwiller (1990), and Delande and
Buchleitner (1994). The photoabsorption cross section o(E) of an atom in the
state ]~pi) with energy Ei can be expressed in the form (Du and Delos, 1988)

o(E) = 4~ hcc Re dt e iEt (lpi DU(t, 0)D I 1])i ) ] , (40)

where D = r . e is the dipole operator responsible for photoabsorption of the


field with polarization e and E ,.~ Ei + ]'tO). The quantity U(t, 0) is the time-
evolution operator in the presence of the binding potential as well as additional
static external electric and magnetic fields that may be present. In effect, the time-
evolution operator propagates wave packets at constant energy that emanate from
the atom and are reflected by the caustics of the potential back to the atom where
they interfere with each other and with the starting wave packets. This leads to
oscillations in the photoabsorption spectrum. In a semiclassical approximation,
the time-evolution operator can be expanded in terms of classical closed orbits
that start from and return to the vicinity of the atom, defined by the spatial range
of the wave function ]~Pi). Since the classical problem is chaotic, there are more
and more such orbits when the energy nears zero. Fourier transformation of the
photoabsorption spectrum reveals the recurrence times of the classical orbits.
There are several differences to the quantum orbits we are considering here.
In our case, the role of the binding potential is, in effect, reduced to acting as
a coherent source of electrons and to causing rescattering, while in closed-orbit
theory the interplay of its spatial shape with the external static fields generates
the rich structure of the closed orbits. In our case, closed orbits are entirely due
to the time dependence of the laser field. The most important difference is that
closed-orbit theory is concerned with total photoabsorption rates as a function
of frequency, while we consider differential electron spectra for a laser field
with fixed frequency. In other words, our orbits depend on the final state of
the electron.
From Eq. (19), in view of the completeness of the Volkov states, the total
ionization probability due to direct electrons is

d 3p IM, i2 = 2e 2 [ f~ ft,d,o (~or F-,(t,)u(Vv)(t,,to)r 9 s ~'o(to)).1


(41)
L a --OO d --OO

This differs from the photoabsorption cross section (40) only by the presence
of the Volkov time-evolution operator u(Vv)(tl,t0), which reflects the strong-
field approximation, instead of the exact time-evolution operator U(t, 0) of the
time-independent problem [for which the time-evolution operator U(t, t ~) only
IV] ATI: CLASSICAL TO QUANTUM 59

depends on the time difference t - t~]. In the total ionization probability (41),
via the same partial integration (18) as above, the electron-field interaction
r. s can be replaced by V(r). The result then looks like the differential
HAIl amplitude (29) except that it is sandwiched by the ground state. This
correspondence is a manifestation of the optical theorem.

D. THE ROLE OF THE BINDING POTENTIAL

The improved Keldysh approximation (29) has been written down for an
arbitrary binding potential V(r). The expansion in terms of the binding potential,
introduced via the Dyson equation (15), is a strong-field approximation (SFA),
which is valid when the electron's quiver amplitude is so large that most of its
orbit is outside the range of the binding potential. This is trivially guaranteed
for the three-dimensional binding potential of zero range,

V(r) : mtr 6(r)Or" (42)

This potential supports a single (s-wave) bound state at the energy -tc2/2m
and a continuum that is undistorted from the free continuum except for the
s wave, as required by completeness (Demkov and Ostrovskii, 1989). Without the
regularization operator (O/Or)r, which acts on the subsequent state, the potential
does not admit any bound state. There are several possibilities to adjust the one
parameter tr to an individual atom or ion. In most cases one will determine it
so as to reproduce the ionization potential; see, however, Sect. IV.G.
The zero-range potential (42) underlies many of the explicit results exhibited
in this chapter. However, we emphasize that the amplitude (29), as well as its
saddle-point approximation (37), hold for a much wider class of potentials.
Regardless of the potential, the saddle-point equations (34)-(36) have the
electron start from and return to the center of the binding potential, which is
the origin, and do not depend on its shape. The potential only enters via the
form factors in Eq. (32). For the SFA to be applicable, they must depend on
time only weakly. The procedure corresponds to the Born approximation. It will
be the better justified, the shorter the range of the potential is, so that the form
factor depends only weakly on the momenta. Excited bound states do not enter
the amplitude (29) regardless of the potential used.
For a comparison of a high-order ATI spectrum calculated for the zero-range
potential (42) with the same spectrum extracted from a solution of the three-
dimensional TDSE for hydrogen, see Cormier and Lambropoulos (1997) for the
latter and Kopold and Becker (1999) for the former. There is good qualitative
agreement within the ATI plateau; in particular, the positions of the dips in the
spectrum that are due to destructive interference agree within a few percent.
The comparison confirms that the detailed shape of the potential has only minor
significance for the HATI spectrum. This holds for hydrogen and the rare gases,
60 W. Becker et al. [IV

but not for the alkali-metal atoms (Gaarde et al., 2000). The height of the
rescattering plateau with respect to the direct electrons does depend on the atomic
species; for the theoretical modeling, see Goreslavskii and Popruzhenko (1999a,
2000). Clearly, however, the real physical systems best described by a zero-
range potential are negative ions with a s-wave ground state. Angular-resolved
photoelectron spectra in H- have been recorded recently by Reichle et al. (2001).
Attempts have been made at including the effects of the binding potential
of the residual ion beyond the first-order Born approximation. Kamifiski et al.
(1996) and Milogevi6 and Ehlotzky (1998c) have employed the so-called
Coulomb-Volkov (CV) states, which are obtained from the ordinary Volkov
states (12) by replacing the plane wave [ p - eA(t)) by a Coulomb outgoing-wave
scattering state with momentum p (ordinary CV state) or p - eA(t) (improved
CV state). These are then substituted in the transition amplitudes (17) or (29)
for the ordinary Volkov states. A systematic assessment of the merits of this
approach appears not to have been made; recently, however, it has been compared
with the solution of the TDSE and was shown to work well for ionization by
ultrashort pulses having a duration shorter than the orbital period of the initial
bound state (Duchateau et al., 2001).

E. A HOMOGENEOUS INTEGRAL EQUATION


An alternative route to the standard KFR matrix element (19) and its improved

/'
version (29) starts from the homogeneous integral equation

IvP(t)) = - i dr Uf(t, r) VIW(r)), (43)


OO

which holds for the state that develops out of the unperturbed ground state
due to its interaction with the laser field. This integral equation can be derived
immediately from the Dyson equation (15) if one applies both sides of the latter
to the atomic ground state [~P0(t~)) in the limit where t ~ ---, -oo. By inspection,
one may convince oneself that the term Uf(t,t')[~Po(t')) makes no contribution
for t - t ~ --, cx~ so that Eq. (43) is left. This equation was first introduced in the
context of the quasi-energy formalism by Berson (1975) and by Manakov and
Rapoport (1975) for circular polarization and Manakov and Fainshtein (1980)
for arbitrary polarization.
Inserting on the right-hand side of the integral equation (43) the expansion (30)
of Uf in terms of Volkov states and replacing [W(r)) by the unperturbed
atomic ground state [~P0(~')), one can read off the matrix element (19) for direct
ionization. Iterating Eq. (43) one gets

[W(t)) = - dr dr' Uf(t, r) VUf(T, .gt) V klJ(Tjt)), (44)


s (X)

which yields the improved KFR amplitude (29) in the same fashion.
The integral equation (43) is particularly useful for the zero-range poten-
tial (42), since in this case it allows one to calculate the wave function in all space
IV] ATI: CLASSICAL TO QUANTUM 61

provided it is known at the origin. For the latter, to a first approximation, one
may employ the unperturbed wave function. Better approximations are obtained
by using more accurate expressions. These incorporate the possibility that the
ionized electron revisits the core, as illustrated by Eq. (44).
For the zero-range potential and a monochromatic plane wave with circular
polarization, it can be shown that the wave function near the origin exactly obeys
qJ(r,t) o( ( 1 / r - t c ) e x p ( - i E t ) for all times. The complex quasi-energy E has
to be determined as the eigenvalue of a nonlinear integral equation (Berson,
1975; Manakov and Rapoport, 1975). For any polarization other than circular,
the time dependence at the origin is given by a Floquet expansion (Manakov and
Fainshtein, 1980; Manakov et al., 2000). The interaction with a laser field for a
finite period of time was considered along similar lines by Faisal et al. (1990)
and Filipowicz et al. (1991); see also Gottlieb et al. (1991) and Robustelli et al.
(1997). The integral equation (43) was also used for two-center potentials in
order to model molecular ions (Krstid et al., 1991; Kopold et al., 1998).

F. QUANTUM ORBITS FOR LINEAR POLARIZATION


For linear polarization, Fig. 9 presents a calculated ATI spectrum that is typical
of a high laser intensity, cf. the data of Walker et al. (1996). The solid circles
that make up the topmost curve of the upper panel were calculated from the
amplitude (29) by means of a zero-range potential, while the other curves give the
results of including an increasing number of quantum orbits in the saddle-point
approximation (37). The spectra that result from just the sth pair (which com-
prises the orbits 2 s - 1 and 2s) are displayed in the lower panel. Quantitatively,
the first pair dominates the entire spectrum, but the contribution of the second
pair comes close, in particular near its cutoff around 7 Up. The contribution of the
third pair is already weaker by almost one order of magnitude, and the subsequent
pairs hardly play a role anymore. Indeed, in the upper panel, already the third
curve from bottom virtually agrees with the result of the exact calculation.
The dependence of the parameters tls, t l s - tos (the travel time), and ks on the
electron energy E v is illustrated in Fig. 10 for the two orbits (s = 1,2) having the
shortest travel times. These parameters uniquely specify the quantum orbits in
space and time. Their behavior is very different for energies below and above the
classical cutoff at 10Up. Below the cutoff, the imaginary parts of the parameters
are only weakly dependent on the energy. Both orbits have to be included in
the sum (37), and their interference leads to the beat pattern, which is visible
in the spectrum of Fig. 9. Notice that the imaginary parts of both the return
times tls and the momenta kx, are small. In contrast, the imaginary part of the
travel times tls - t0s, which are related via t0s to the tunneling rate, is substantial;
see Fig. 5, where the ionization time ts is plotted for the direct electrons. Hence,
after rescattering, the orbits are real for all practical purposes: the electron has
forgotten its origin via tunneling. The parameter values of the two orbits (s = 1,2)
approach each other closely near the cutoff. At some point, one of the two orbits
62 W. B e c k e r et al. [IV

-8 "1""I 1'[ E~~ '/,, integration . . . . . .

"E - 1 2

-o -16 10

~-~ -22

~ -24
!
-26

-28 0 | !

1
J

"N,
P,

2 3 4 g t5 "7 8 9 10 11 12
electron energy [Up]
FIG. 9. Upper panel: ATI spectrum in the direction of the laser field for linear polarization for
1015 W/cm 2, hto = 0.0584 a.u., and a binding energy of Eip = 0.9 a.u. The electron energy is given
in multiples of Up. The curve at the top (solid circles) is the exact result from Eq. (29). The other
curves were calculated from the saddle-point approximation (37). From bottom to top, more and
more quantum orbits are taken into account; the results are displaced with respect to each other for
visual convenience. The curve at the bottom incorporates just the pair of orbits with the shortest
travel times, the next one up includes in addition the pair with the next-to-shortest travel times, and
so on. The occasional small spikes are artifacts of the saddle-point approximation, cf. Goreslavskii
and Popruzhenko (2000) and Kopold et al. (2000a). Lower panel: The envelopes of the contributions
of the individual pairs are shown all on the same scale so that the quantitative relevance of the various
pairs is put in perspective. The cutoffs of the various orbits agree with those displayed in Fig. 8.
From Kopold et al. (2000a).
IV] ATI: C L A S S I C A L TO Q U A N T U M 63

Re cot Re c o ( t - t ' )
2 3 4 3 4 5
i ' i ' i 0
, 1 , i [!11.51 , -
_ 11.5~ _
0.5
-0.5

2.5
_a o
-1
~,, 9 -o- - o- - -iBt~ 11.5 s
tl E
-1.5
I
--
-0.5 e91 1 5 9
t11.5
i ~ J , I -2
, _
' ' ' I ' ' ' '
i , i , i 1,15 I
0.2 i

m
-0.1
< 0.1 - 11.5 ~ ~ 2 5 -
<
i1) _ ~ 0 1 "" m

0
_.,. ~-.z-.:_ .. ,, 2.54 -0.2 x "

t
_ /,o-~ 11.5 -
E -0.1 - ~ ~ E m

-
i , i , "
~
9
11.5
, I , i i
...2.,s i 1 i i
v

-0.8 -0.6 -0.4 -0.2 -0.1 0


Re kx/leAI Reky/leAI

FIG. 10. Saddle points (ts, t~,ks) for the orbits (s = 1,2) having the two shortest travel times. In
this figure, ts is the return time (elsewhere denoted by t]s), and ts~ the start time (elsewhere denoted
by tos). The figure shows a comparison of elliptical polarization (~ --0.5, solid circles) and linear
polarization (open squares). The values of the other parameters are those of Fig. 9 (eA = 2.04 a.u.).
The symbols identify electron energies of 11.5, 10.4, 8.92, 6.01, and 2.49, all in multiples of Up.
The dashed orbits have to be dropped from the sum (37) after the cutoff. With the scaling of k
given on the ordinate, the saddle points depend only on the Keldysh parameter y = v/IEoI/2Up.

(drawn dashed in the figure) has to be dropped from the sum (37). This causes the
artifact of the small spikes visible in Fig. 9. For energies above the cutoff, just one
orbit contributes and, as a consequence, the spectrum smoothly decreases without
any trace o f interferences. The real part o f the parameters stays approximately
constant, while the imaginary part increases strongly with increasing energy. This
is responsible for the steep drop o f the spectrum after the cutoff. Similar behavior,
as a function o f ellipticity, occurs in Fig. 5.
The procedure o f dropping one o f the orbits o f each pair after its cutoff
can be replaced by a more rigorous method. In the vicinity o f the cutoffs,
an approximation in terms of Airy functions was used by Goreslavskii and
Popruzhenko (2000). A uniform approximation was described in a different
context by Schomerus and Sieber (1997). It reproduces the spectra o f Fig. 9
without the spikes (Figueira de Morisson Faria et al., 2002b).

G. ENHANCEMENTS IN ATI SPECTRA


In several experiments, pronounced e n h a n c e m e n t s o f groups o f ATI peaks in
64 W. B e c k e r et al. [IV

105

103 . . . . . . . . . .
==
8 lo 2

101
I

0 10 20 30 40
electron energy [eV]

FIc. 11. ATI spectra in argon at 800 nm recorded in the direction of the linearly polarized field for
various intensities rising by increments of 0.110 from 0.5 I0 (bottom curve) to 1.010 (top curve). The
horizontal lines mark the maxima of the ATI plateaus for each intensity. For intensities I > 0.8 I0 a
group of ATI peaks between 15 eV and 25 eV rears up quickly. (The spectra shown here represent
only a fraction of those actually measured.) From Paulus et al. (2001a).

the plateau region (by up to an order of magnitude) have been observed upon
a change of the laser intensity by just a few percent (Hertlein et al., 1997;
Hansch et al., 1997; Nandor et al., 1999). This behavior suggests a resonant
process. Near the resonances, the contrast of the spectra is remarkably reduced
(Cormier et al., 2001). For the experiments reported so far, the effect is most
pronounced for argon. This holds not only for a laser wavelength of 800nm
but also for 630nm (Paulus et al., 1994c). The enhancements are so strong
that in experiments implying significant focal averaging the observed spectral
intensity may well be dominated by these enhancements, regardless of the actual
peak intensity. In this sense, ATI in toto has been called a resonant process
(Muller, 1999b). A big step towards understanding the physical origin of the
enhancements was made in theoretical studies that reproduced the enhancements
in the single-active-electron approximation by numerical solution of the one-
particle time-dependent Schr6dinger equation in three dimensions (Muller and
Kooiman, 1998; Muller, 1999a,b; Nandor et al., 1999), thereby ruling out any
mechanism that invokes electron-electron correlation.
In Fig. 11 we show results of a measurement of the same effect, but for
a shorter pulse length of 50 fs (Paulus et al., 2001a). Spectra in an intensity
interval of 0.5 to 1.0 • I0 in steps of 0.1 • I0 are displayed. The maximum
intensity I0 was calibrated by using the cutoff energy of 10Up. This leads to
I0 ~ 8• 1013 W/cm 2. There is a striking difference between the spectra for
I ~< 0.810 and those for higher intensity: within a small intensity interval a group
IV] ATI: CLASSICAL TO QUANTUM 65

of ATI peaks corresponding to energies between about 15 eV and 25 eV grows


very quickly. In the figure this is emphasized by horizontal lines drawn at the
maximal heights of the plateaus. Increasing the intensity above 0.910 leads to a
smaller growth rate of these peaks. The plateau, however, preserves its shape.
For an interpretation, it should be kept in mind that a measured ATI spectrum is
made up of contributions from all intensities I ~< I0 that are contained within the
spatio-temporal pulse profile. This means that a spectrum for a fixed intensity
would show the enhanced group of ATI peaks only at that intensity where it first
appears in our measurement, namely at I ~ 0.8510 = 7• 1013 W/cm 2. In other
words, the enhancement happens at a well-defined intensity or at least within a
very narrow intensity interval.
Analyzing the wave function of the atom in the laser field, Muller (1999a)
suggested that the enhancements are related to multiphoton resonances with
ponderomotively upshifted Rydberg states. In some cases, in particular for
electrons with rather low energy, one particular Rydberg state could be definitely
identified as responsible. In others, notably for the strong enhancement that for
appropriate intensities dominates the middle of the plateau, this was not possible
(Muller, 2001 a).
A closer look at the data of the measurement shown in Fig. 11 reveals that
under the conditions of this experiment (i.e. a pulse duration of 50 fs as compared
with more than 100fs in the other measurements mentioned) resonantly
enhanced multiphoton ionization does not play an essential role. This can be
deduced from the different intensity dependence of the enhanced ATI peaks in
the plateau and of the low-energy ATI peaks, see Fig. 12. The latter are known to
originate from atomic resonances (Freeman et al., 1987; Agostini et al., 1989).
In Muller's numerical simulations, the existence of excited bound states
appears to be instrumental for the enhancements. Yet, the modified KFR
matrix element (29), which does not incorporate any excited states, produces
much the same enhancements (Paulus et al., 2001a; Kopold et al., 2002). An
example is shown in Fig. 13. In these calculations, the enhancements occur for
intensities for which an ATI channel closes. This is the case when

EIp + Up = khco. (45)

For an intensity slightly higher than specified by this condition, k + 1 is the


minimum number of photons required for ionization in place of k. Such channel
closings are distinctly visible in the multiphoton-detachment yields of negative
ions (Tang et al., 1991) and have been shown to produce a separate comb of
peaks in the low-energy ATI spectrum (Faisal and Scanzano, 1992). Comparison
of the channel-closing condition (45) with the ATI energy spectrum (22) shows
that at a channel closing electrons may be produced with zero drift momentum p.
In this event, the energy of the k photons is entirely used to overcome the binding
potential raised by the ponderomotive energy, and no energy is left for a drift
66 I'I4. B e c k e r et al. [IV

0.00
iI
o
d
. _

co
" -0.05
O
o

I1) I

N tI

-0.10 t
\

E - plateau
0 a 7.3 eV
• 6.4eV
-0.15 I = I I ~ J

0.4 0.6 0.8 1.0


I/I o

FIG. 12. Comparison of the intensity dependence of ATI electrons with different energies. For
visual convenience, the overall increase in yield with increasing intensity has been subtracted. The
electrons at 6.4eV and 7.3 eV are due to the strongest Freeman resonances, i.e. resonance with
atomic states. Those labeled "plateau" are electrons in the plateau region of the spectra. As a
consequence of the subtraction of the overall increase, the resonance-like behavior corresponds to
those intensities where the respective curves start rising. It is evident that for the plateau electrons
this does not happen at those intensities where the atomic states shift into resonance. Quite to the
contrary, the intensity at which the yield of the plateau electrons starts its rise is reflected in the yield
of the low-energy electrons by a brief halt in their rise. This is indicated by the dashed circles.

motion. An electron having a drift m o m e n t u m near zero has many recurring


opportunities to rescatter. Indeed, the quantum-orbit analysis of the spectra o f
Fig. 13 shows that at the channel closings, and only there, an exceptionally large
number o f orbits are required to reproduce the exact result. All of these orbits
conspire to interfere constructively to produce the observed enhancements. In the
tunneling regime, this can be proved analytically (Popruzhenko et al., 2002).
In Muller's numerical simulations, inspection of the temporal evolution reveals
that at the intensities that produce the enhancements electrons linger about
the ion for m a n y cycles of the field before the final act of rescattering.
A detailed comparison between Muller's numerical simulations and results based
on Eq. (29) has been made by Kopold et al. (2002). This paper also includes
an assessment of the consequences of focal averaging. It is noteworthy that both
approaches predict ATI enhancements also for helium deeply in the tunneling
regime, in spite o f the obvious multiphoton character of the channel-closing
condition (45). Unfortunately, the helium data of Walker et al. (1996) and Sheehy
et al. (1998) do not allow one to draw conclusions about the presence or absence
of enhancements.
The interference interpretation just given requires the existence of a sufficient
number of orbits to contribute to the energy considered. The lower panel o f
IV] ATI: C L A S S I C A L TO Q U A N T U M 67

I 12
! 4~
/ 1"1= 2.626
-13 ~ ~ ,'- "-. ........

-14 "'-'~- '" '" ''• ""


' ,,, , - - 2
-15 ,' ': ,,' - 6
, , - 40
I ',;~ I ~ I ~ I J I
10 20 30 40 50
ffl
~ 9 -12
_ 9
& -13
"o

~, -14
t-
O
i_
tO
-15
o ~", I , ' , i t,rajec] ~
0 10 20 30 40 50

- 12 j oo q = 2.326
-13 e ~

-14 ,....,,

-15"" ,, --- 6 9
',' trajectories 04

10 20 30 40 50
electron energy [eV]

Fic. 13. ATI spectra for Eip = 14.7eV, oJ = 1.55 a.u., and three intensities: at a channel closing
(71 = Up/oJ = 2.526, middle panel), below the channel closing (7/= 2.326, lower panel), and above
(7/= 2.626, upper panel). In each panel, the exact result calculated from Eq. (29) is shown (solid
symbols) and approximations involving the first 2 (dashed line), 6 (dot-dashed line), and 40 (solid
line) quantum orbits in Eq. (37). From Kopold et al. (2002).

Fig. 9 shows that too few orbits contribute for energies above about 8Up. Indeed,
the enhancements observed experimentally are restricted to the lower two-thirds
of the plateau. The interpretation also implies that the enhancements should
disappear for ultrashort pulses, where late returns do not occur. This has been
observed in experiments by Paulus et al. (2002). In numerical simulations o f
H H G based on the three-dimensional TDSE, the same effect has been noticed
by de Bohan et al. (1998).
W h e n the modified K F R matrix element (29) is used to describe data for real
atoms, the ionization potential EIp has to be replaced by an effective (lower)
value that corresponds to the d e f a c t o onset o f the continuum (Paulus et al.,
68 W. B e c k e r et al. [IV

2001 a; Kopold et aL, 2002). It is a fact that, for a Coulomb potential, the actual
onset of the continuum is hard to see and may better be replaced by an effective
value. This is illustrated, for example, by the photoabsorption spectra of Garton
and Tomkins (1967).
Numerical simulations predict very similar enhancements in high-order
harmonic spectra (Becker et al., 1992; Toma et al., 1999; Kuchiev and Ostrovsky,
2001; Kopold et al., 2002) and in nonsequential double ionization (NSDI) of
helium (Muller, 2001c). Experimentally, in argon irradiated by a flat-top pulse
from a Ti:Sa laser, resonant-like enhancement of the 13th harmonic was observed
by Toma et al. (1999). In HHG in one dimension, the dependence of the
enhancements on the shape of the potential and the presence or absence of
excited bound states has been investigated (Figueira de Morisson Faria et al.,
2002a). The results are largely compatible with the quantum-orbit picture.
In a semiclassical framework, the binding potential can be incorporated into
the orbits. This leads to Coulomb refocusing (Ivanov et al., 1996; Yudin and
Ivanov, 2001a): orbits that would miss the ion in the absence of the binding
potential are refocused to the ion in its presence. This emphasizes the importance
of late returns and leads to a substantial increase of rescattering effects without,
however, resonant behavior. If late returns are cut off due to an ultrashort laser
pulse, the rate of NSDI should decrease. Indeed, this has been experimentally
confirmed by comparison of 12-fs and 50-fs pulses (Bhardwaj et al., 2001).

H. QUANTUM ORBITS FOR ELLIPTICAL POLARIZATION

Formulation of a classical model of the simple-man variety to describe


rescattering for an elliptically polarized laser field meets with difficulties. The
problem is that an electron that starts with zero velocity almost never returns
exactly to its starting point if the laser field has elliptical polarization. Formally,
this shows in the saddle-point equations (34)-(36) as follows. For EIp = 0,
Eq. (34) yields k = eA(t0) if real solutions are sought. For linear polarization,
this leaves two equations to be solved for to and tl: Eq. (36) and the x-projection
of Eq. (35). Real solutions are obtained, provided the final momentum p is
classically accessible. In contrast, for elliptical polarization, three equations are
left since now both the x-projection and the y-projection of Eq. (35) have to
be considered. Hence, there is no simple-man model for elliptical polarization,
even when EIp = 0. The same situation occurs for HHG. This does not mean
that there is no HATI or HHG for elliptical polarization: quantum-mechanical
wave-function spreading assures overlapping of the wave packet of the returning
electron with the ion (Dietrich et al., 1994; Gottlieb et al., 1996). The complete
absence of HHG for a circularly polarized laser field is sometimes taken as
confirmation of the rescattering mechanism. This conclusion is not rigorous since
there is still sufficient overlapping. Rather, the absence of HHG is due to angular-
momentum selection rules or, equivalently, destructive interference.
IV] ATI: C L A S S I C A L TO Q U A N T U M 69

~ 5Up 20 a.:u. electric field ellipse

edlrec~ons~. ' +~~'~~~" "= ~ y ~


-18
-~o ~ 7Up -

.m
r -20

. . . . . . . . . . . . . 8.5u j
~"o. -22

" " " " " " -""~" ' " 2


2 -24
o
o exact integration '\ o~ ,,, "~
O
..... 11+212 ,, o ',,, %
~ -26 . . . . . . . 13+412 '\ -',, '~ _
o
....... 15+612 ', ',,,
-
- Isuml 2 2
'\
~
',,,
-28 I , I ~ I ~ I ~ I ~ ~,1 , ]'. . . .
3 4 5 6 7 8 9 10 11
electron energy [ Up]

FIG. 14. ATI spectrum in the direction of the large component of the elliptically polarized driving
laser field (5) for ff = 0.5 (see the field ellipse in the upper right corner of the figure) and electron
energies between 2.5 and 10.5Up. The other parameters are o9 = 1.59eV, EIp = 24.5eV, and
I = 5• 1014 W/cm 2. The open circles give the yields of the individual ATI peaks calculated from
the integral (29) for the zero-range potential. The other curves represent the contributions to the
quantum-path approximation (37) of the shortest trajectories 1 and 2 (dot-dashed line), 3 and 4
(long-dashed line), and 5 and 6 (short-dashed line), as well as the sum of all six (solid line). Note
that some of these curves overlap partly or entirely. The orbits responsible for each part of the
spectrum, viz. 1 and 2, 3 and 4, and 5 and 6, are presented near the margins of the figures. The
position of the ion is marked by a cross; notice that the orbits do not depart from there, but rather
from a point several atomic units away from it. This is the point where the electron tunnels into the
continuum. The electron travels the orbits in the direction of the arrows. Experimental data for a
similar situation are shown in Fig. 15. From Kopold et al. (2000b).

One m i g h t try to f o r m u l a t e a s i m p l e - m a n m o d e l for elliptical p o l a r i z a t i o n by


relaxing the r e q u i r e m e n t that the e l e c t r o n return exactly to the posi t i on o f the
ion or by a d m i t t i n g a n o n z e r o initial velocity, but in d o i n g so a large a m o u n t o f
arbitrariness is unavoidable. Instead, we will just solve the s a d d l e - p o i n t e q u a t i o n s
( 3 4 ) - ( 3 6 ) and accept and interpret the c o m p l e x solutions.
The results o f such a calculation are p r e s e n t e d in Fig. 14. W h a t u s e d to be the
rescattering p l a t e a u for linear p o l a r i z a t i o n has t u r n e d into a staircase for elliptical
polarization. E a c h step can be attributed to one p a r t i c u l a r pair o f orbits, and for
each step the real parts o f such orbits are d i s p l a y e d in the figure. T h e orbits
are closely related to their analogs in the case o f linear polarization, e x h i b i t e d
in Fig. 9. In particular, their cutoffs oscillate with i n c r e a s i n g travel time as
70 W. B e c k e r et al. [IV

FIG. 15. ATI spectrum in xenon for an elliptically polarized laser field with ellipticity ~ = 0.36
and intensity 0.77• 1014W/cm2 for emission at an angle with respect to the polarization axis as
indicated in the upper right. The spectrum has a staircase-like appearance. The respective steps are
shaded differently. For each step, the real parts of the responsible quantum orbits are displayed. The
dots with the crosses mark the position of the atom, and the length scale is given in the upper left
of the figure. From Sali~res et al. (2001).

illustrated in Figs. 8 and 9. The main difference is that for elliptical polarization
the orbits are two-dimensional and encircle the ion. The pair of orbits with
the shortest travel times generates the part of the spectrum preceding the final
(highest-energy) cutoff. However, this part is very weak in relation to the yields
at lower energies. The latter are generated by orbits with longer travel times,
whose contributions for linear polarization are marginal, see Fig. 9.
Intuitively, this staircase structure can be understood as follows. Return o f the
electron to the ion is possible if the electron has a nonzero initial velocity. This
velocity is largely in the direction o f the small component o f the elliptically
polarized field. The larger this velocity is, the smaller is the contribution that
the associated orbit makes to the spectrum. [This can be compared with the
distribution of transverse momenta in a Gaussian wave packet (Dietrich et al.,
1994; Gottlieb et al., 1996).] For the shortest orbit, while the large component
of the field changes sign so that the electron is driven back to the core in this
direction, the small component has the same sign for the entire duration o f the
orbit. Hence, a particularly large initial velocity in this direction is required in
order to compensate the drift induced by the small field component. For the
longer orbits, the small component changes direction, too, during the travel time
and, consequently, a smaller initial transverse velocity suffices to allow the
electron to return to the ion. Support for these qualitative statements can be
found in the orbits depicted in Fig. 14.
The parameters of the two shortest quantum orbits can be read from Fig. 10
IV] ATI: CLASSICAL TO QUANTUM 71

and compared with the case of linear polarization. For elliptical polarization, the
momentum ks has two nonzero components, kxs and kys. Both have substantial
imaginary parts, in particular ky~. This is a consequence of the lack of a
classical simple-man model for elliptical polarization, as discussed above. For the
orbits ( s - 3, 4) (not shown), the imaginary parts are much smaller, in keeping
with the fact that they make a larger contribution to the spectrum.
Figure 15 presents a corresponding measurement of an ATI spectrum and
displays the staircase structure predicted by the theory. The first step (the one
corresponding to energies below 10 eV) is due to direct electrons and does not
concern us here. The other ones correspond to the steps of Fig. 14. The real
parts of representative orbits, calculated from Eqs. (34)-(36), are shown in the
figure. In order to reach a maximum contrast for the steps, the spectrum was
recorded at 30 ~ to the major axis of the polarization ellipse.

I. INTERFERENCE BETWEEN DIRECT AND RESCATTERED ELECTRONS

In the lower part of the plateau, the electron can reach a given energy either
directly or after rescattering so that one expects interference of these two paths.
However, Fig. 9 shows that, for linear polarization and high intensity, the
transition region where both paths make a contribution of comparable magnitude
is very narrow. The situation is more favorable for elliptical polarization:
since the plateau turns into a staircase (Fig.14), the yields of the two paths
remain comparable over a larger energy region. This has permitted experimental
observation of this interference effect in the energy-resolved angular distribution
(EAD) (Paulus et al., 2000).
Figure 16 shows a comparison of the EAD's for linear and for elliptical
polarization at the same intensity. For linear polarization, the standard plateau
in the direction of the laser polarization is very noticeable. The side lobes
corresponding to rainbow scattering, mentioned in Sect. III, are also visible.
For elliptical polarization, the plateau has split into two, one to the left of the
direction of the major axis of the field and another weaker one to its right. The
lower panel of Fig. 16 exhibits (on the right) EAD's of a sequence of ATI peaks
where the interference is best developed and (on the left) compares them with
theoretical calculations from the amplitude (29). The parameters underlying the
calculation do not exactly match the experiment. This is mostly attributable
to the insufficient description of the direct electrons for elliptical polarization.
The theoretical results, however, show the same interference pattern. In order
to make sure that this pattern is really due to interference between direct and
rescattered electrons, the two contributions have been displayed separately for
one of the peaks (s = 17): neither one shows a pronounced dip, only their
coherent superposition does. For more details of the theory we refer to Kopold
(2001).
72 W. B e c k e r et al. [IV

FIG. 16. Upper panels: density plots of measurements of the energy-resolved angular distributions
for Xe at an intensity of 7.7 • 1013 W/cm 2 and a wavelength of 800 nm for (a) linear polarization and
(b) elliptical polarization with ellipticity ~ = 0.36. The direction of the major axis of the polarization
ellipse is at 0 ~ Dark means high electron yield. Yields can only be compared horizontally, not
vertically, since the data were normalized separately for each ATI peak. For linear polarization,
the cutoff is at I OUp = 46eV. Lower panels: (a) theoretical calculation from Eq. (29) of the
angular distribution for the ATI peaks s = 11,... ,21 for elliptical polarization (~ = 0.48). The
other parameters are Eip = 0.436 a.u. Oust below the binding energy of xenon in order to stay away
from a channel closing) and I = 5.7 • 1013 W/cm 2. For the ATI peak s = 17, the contributions of the
direct electrons (dashed line) and the rescattered electrons (dotted line) are displayed separately. The
slight variation in the former is unrelated to the interference pattern of the total yield (solid line),
which results from the coherent sum of the two contributions. (b) Experimental angular distribution
extracted from the upper panel (b) of the ATI peaks s - 1 5 , . . . , 25. From Paulus et al. (2000).
V] ATI: C L A S S I C A L TO Q U A N T U M 73

V. ATI in the Relativistic Regime

A sufficiently intense laser field accelerates an electron from rest to relativistic


velocities Ivl ~ c within one cycle. Such intensities are characterized by the
ponderomotive energy Up becoming comparable with or exceeding the electron's
rest energy m c 2. We will briefly summarize the kinematics o f an otherwise free
electron in the presence of such a field. In other words, we will generalize
the simple-man model of Sect. II.A. 1 to the case o f "relativistic intensity." The
changes are surprisingly few.

A. BASIC RELATWISTIC KINEMATICS

For a four-vector potential A~' = (A0,A), the electron's four-vector velocity is


(Jackson, 1999)
m~t ~ = pU - eA ~, (46)

where flu = y(c, v) with v = v(t) the ordinary velocity d x / d t , and the usual
relativistic factor ~, = [1 - ( v / c ) 2 ] - 1 / 2 (not to be confused with the Keldysh
parameter). The four-velocity satisfies ~2 ~ ~ . ~ ~ ~/t~/~ -- C2 SO that
the four-vector mfi is on the mass shell. Equation (46) is the analog o f the
nonrelativistic Eq. (2).
We will consider a plane-wave field o f arbitrary polarization,

2
A" = Z ai(k. x)e; (47)
i=1

with the four-dimensional wave vector k u = (co~c, k) so that k 2 - 0 and k . ei = O.


The field (47) differs from the field (6) by the fact that the wave fronts are now
given by k - x = const, in place o f t = const., that is, we do no longer make
the dipole approximation. We will assume that the laser field propagates in the
z-direction so that k = Iklez.
The four-vector p" = ( E / c , p ) is the canonical momentum. For a vector
potential whose cycle average vanishes, its spatial components p have the
physical meaning o f the drift m o m e n t u m as in the nonrelativistic case. Since
the electron-field interaction depends only on

u =_ k . x / c o = t - z / c , (48)

the canonical m o m e n t u m Pv -= ( p x , P y , 0) transverse to the propagation direction


as well as p . k = co(p0-pz)/C are constants o f the motion inside the field (47). If
we assume that the laser field (47) is turned on and off as a function o f u = t - z/c,
then Pv a n d p . k are also conserved when the electron enters and leaves the field.
74 W. B e c k e r e t al. [V

As we did for the nonrelativistic simple-man model in Sect. II.A. l, we assume


that the electron is initially, at some space-time instant u0, at rest. Then

Pr = eA(u0) and po-pz = mc for all times. (49)

From the condition that ~2 = C 2, using the conditions (49), we obtain the energy
as a function of u,

e2
E _ 1+ ( A ( u ) - A(uo))2 (50)
~, - mc 2 2m 2c 2

This yields the cycle-averaged kinetic energy

(Ekin) = ( g ) - mc 2 = ~m + Up. (51)

This is exactly the same decomposition into drift energy and ponderomotive
energy as in the nonrelativistic case, Eq. (3). The ponderomotive energy is
still defined by Eq. (4), and the classical bounds of the spectrum discussed
in Sect. II.A.1 are unchanged. However, velocity and canonical momentum are
connected by the relativistic expresion myv = P r - eA, and the cycle average
was performed with respect to u rather than the time t. Since it can be shown
that u is proportional to the electron's proper time, this was, actually, the proper
thing to do (Kibble, 1966).
The fact that p . k is a conserved quantity implies that the electron's velocity
in the propagation direction of the laser field is given by

pz = myvz = mc(y- 1) = Ekin/C. (52)

The presence of this momentum reflects the fact that a laser photon has a
momentum in the direction of its propagation or, alternatively, that the magnetic
field via the Lorentz force causes a drift in the propagation direction or,
alternatively, that the laser field exerts radiation pressure. All three statements
are essentially equivalent. As a consequence, electrons born with zero velocity in
a relativistic laser field are no longer emitted in the direction of its polarization,
but acquire a component in the propagation direction of the laser so that, for
circular polarization, they are emitted in a cone given by the angle

1
tan 0 - IvT(t)l
2m _ . / 2
(53)
Iv~(t)l Ipr/ Vyo~-1
with respect to the propagation direction. The subscript oc characterizes
quantities outside of the laser field. In the derivation, Eqs. (49)-(52) were used.
The angle 0 has been observed by Moore e t al. (1995, 1999) for intensities
V] ATI: CLASSICAL TO QUANTUM 75

of several 1018 W/cm 2 and ~ = 1.053~tm and was used to draw conclusions
regarding the actual (nonzero) value of the initial velocity (McNaught et al.,
1997), which can be introduced as discussed in the nonrelativistic case in
Sect. II.A. 1. There are, however, still some unresolved issues in the interpretation
of these experiments (Taieb et al., 2001).
The cycle average of Eq. (50) can be written in the covariant form

p Z = m Z c 2 - ((eA)2)_ m ,2c 2 > m2c 2, (54)

where p2 and (cA) 2 < 0 are invariant four-dimensional scalar products. This
relation is often used to introduce the so-called "relativistic effective mass" m,.
It occurs very naturally in the context of the Klein-Gordon equation

[(iO~ - eA~) 2 - m2c 2] q / = 0, (55)

which explicitly displays the effective mass. However, one has to keep in
mind that this apparently increased mass is just due to the transverse wiggling
motion of the electron, viz. the ponderomotive energy, and that there is nothing
especially relativistic about it. All the same, envisioning the ponderomotive
energy as a mass increase makes sense since, like the rest mass, it is an energy
reservoir that is not easily tapped.
The classical kinematics just discussed are embedded in quantum-mechanical
calculations, which can be carried out along the lines of the strong-field
approximation (17), taking the relativistic instead of the nonrelativistic Volkov
wave function (Reiss, 1990; Faisal and Rado2ycki, 1993; Crawford and Reiss,
1997). In particular, the stationary-phase approximation is well justified, leading
to a form similar to Eq. (21) (Krainov and Shokri, 1995; Popov et al., 1997; Mur
et al., 1998; Krainov, 1999; Ortner and Rylyuk, 2000).

B. RESCATTERING IN THE RELATIVISTIC REGIME

With increasing laser intensity, the first relativistic effect to become significant-
before the ponderomotive potential becomes comparable with the electronic rest
mass - is the drift momentum (52) in the direction of propagation of the laser
field, which can be traced to the Lorentz force. This has virtually no effect on
the initial process of ionization where the electron's velocity is low, but since it
is always positive it prevents the electron from returning to the ion. Therefore,
with increasing intensity it gradually eliminates the significance of rescattering
processes. This effect can be estimated by calculating the distance by which
the electron misses the ion in the z-direction when it returns to the ion in the
76 W. B e c k e r et al. [VI

x - y plane (approximately at the time/ret ~ T/2). From Eqs. (52) and (51) (where,
for simplicity, we only kept Up), we obtain

Up
(Vz) T/2 ~ )~ (56)
2mc 2

with/l the wavelength of the laser field. Obviously, this distance can exceed the
width of the wave packet of the returning electron to the point where it does not
overlap anymore with the ion, even when Up/mc 2 << 1.
In HHG the consequences have been investigated in a number of recent
theoretical works (see Sect. VI.D) and were found to cause a dramatic drop of the
plateau. The same should be expected for high-order ATI, but to our knowledge,
this has not been explored in detail. However, in the analysis of multiple-
nonsequential-ionization experiments of neon at 2 • 10 ~s W/cm 2 a conspicuous
suppression of the highest charge state has been attributed to the magnetic-field-
induced drift (Dammasch et aL, 2001).

VI. Quantum Orbits in High-order Harmonic Generation

According to the rescattering model, the physics of high-order ATI and high-
order harmonic generation differ only in the third step: elastic scattering versus
recombination. Correspondingly, the description in terms of quantum orbits can
be applied to HHG as well; in fact, quantum orbits were introduced for the first
time in the analysis of HHG by Lewenstein et al. (1994). It is from the practical
point of view that the two processes differ greatly: HHG by one single atom has
never been observed, only HHG by an ensemble of atoms. This introduces phase
matching as an additional consideration, equal in significance to the single-atom
behavior (Sali~res et al., 1999; Brabec and Krausz, 2000).
Below we will consider examples of a quantum-orbit analysis of HHG for
several nonstandard situations. The first example is an elliptically polarized laser
field. A bichromatic elliptically polarized laser field was considered by Milogevi6
et al. (2000), and in Sect. VI.C we concentrate on a special case of such a field: a
two-color bicircular field. Finally, in Sect. VI.D the quantum-orbits formalism is
extended into the relativistic regime. A bichromatic linearly polarized laser field
was investigated by Figueira de Morisson Faria et al. (2000), and a simplified
version of the quantum-orbits formalism was used to deal with problems in the
presence of a laser field and an additional static electric field (Milogevi6 and
Starace, 1998, 1999c) or a laser field and an additional magnetic field (Milo~evi6
and Starace, 1999a,b, 2000).
VI] ATI: CLASSICAL TO QUANTUM 77

A. THE LEWENSTEINMODEL OF HIGH-ORDERHARMONIC GENERATION

The matrix element for emission of a photon with frequency fl and polarization e
in the HHG process in the context of the strong-field approximation (Lewenstein
et al., 1994),

Me(f2) " dtl dto d3k exp [iSn(tl, to, k)] me(tl, to, k), (57)
OO

has the same structure as the corresponding expression (31) for ATI. The
function

me(tl,to, k) : (~0o l e r - e I k - eA(h)) ( k - eA(to)ler. E(to)l ~Po) (58)

is the product of two matrix elements: one that describes the ionization at time to
due to interaction with the laser field, and another one at time tl that corresponds
to recombination into the ground state followed by emission of the high-order
harmonic photon having the polarization e. The difference to ATI is mostly in
the first term of the action:

Sn(tl, to, k) = dr (EIp - ~ ) - ~mm dr [k - eA(r)] 2 + dr EIp, (59)


Oo

which now refers to the emitted photon. The corresponding saddle-point


approximation of Eq. (57) is like the HATI approximation (37), except that the
summation is now over saddle points that are solutions of the system of equations
(34), (35) and (Lewenstein et aL, 1995b, Kopold et al., 2000b)

[k - eA(tl)]2 = 2m(~ - EIp). (60)

The last equation corresponds to the condition of energy conservation at the time
of recombination and replaces the condition (36) of elastic rescattering in HATI.
For a linearly polarized monochromatic field, quantum orbits were employed
from the very beginning for the evaluation of HHG in the Lewenstein model
(Lewenstein et al., 1994, 1995b) and routinely applied in the theoretical analysis
and interpretation (Sali6res et aL, 1999). Conversely, numerical solutions of the
TDSE were analyzed in terms of the short (rl) and the long (r2) quantum orbit,
and the dominance of these two orbits was corroborated (Gaarde et al., 1999;
Kim et al., 2001). The contributions of the long and the short orbit could be
spatially resolved in an experiment by Bellini et al. (1998). Spectral resolution
was achieved by exploiting the dependence of phase matching on the position of
the atomic jet with respect to the laser focus by Lee et al. (2001) and by Salibres
et al. (2001).
78 W. Becker et al. [VI

At the end of Sect. II.B we remarked that the quantum-orbit formalism


is not restricted to periodic fields, but can equally well be applied to finite
pulses. For a periodic field, interference of contributions from different cycles
generates a discrete spectrum. For a finite pulse, it enhances or suppresses
particular frequency intervals. This was dubbed "intra-atomic phase matching"
by Christov et al. (2001) and has been calculated in terms of quantum orbits; in
the context of the TDSE, see Watson et al. (1997). This mechanism underlies
the engineering of a HHG spectrum by tailoring the pulse shape in a feedback-
controlled experiment (Bartels et al., 2000, 2001). Individual HHG peaks could
be enhanced by up to an order of magnitude.
A description of HHG that is practically equivalent to the Lewenstein
model is based on the integral equation (43) and the zero-range potential (42)
(Becker et al., 1990, 1994b). The equivalence implies that the contribution of
"continuum-continuum terms" is insignificant (Becker et al., 1997). The three-
step nature of H H G - direct ATI followed by continuum propagation followed
by laser-assisted recombination- is particularly emphasized in the approach of
Kuchiev and Ostrovsky (1999, 2001), where the integration over the intermediate
momentum k is replaced by a discrete summation over ATI channels. The latter
is carried out by a variant of the saddle-point approach, which is reminiscent of
Regge poles and leads to a complex effective channel number.

B. ELLIPTICALLY POLARIZED FIELDS

High-order harmonic generation by an elliptically polarized field is of great


interest for applications such as the generation of sub-femtosecond pulses
(Corkum et al., 1994). For theoretical calculations in the context of the SFA, see
Becker et al. (1994b, 1997) and Antoine et al. (1996); for a fairly comprehensive
list of references, see Milogevid (2000). Fields having polarization other than
linear generate particularly appealing quantum orbits since they allow them to
unfold in a plane. As an example, Fig. 17 shows a HHG spectrum for the
elliptically polarized laser field (5) (Kopold et al., 2000b; Milogevid, 2000).
The figure confirms that the "exact results" are well approximated by the
contributions of only the six shortest orbits. This figure is the analog of Fig. 14
for HATI. The spectrum exhibits the same staircase structure, and everything
said there applies here as well.

C. H H G BY A TWO-COLOR BICIRCULAR FIELD

The bichromatic m - 2 m laser field


s 1 (gle+e-iCot + r
= gi -2it~ + c.c. , (61)
whose two components are circularly polarized and counter-rotating in the same
plane (e+ = (i 4- i~,)/v/2), is known to generate high harmonics very efficiently;
VI] ATI: CLASSICAL TO QUANTUM 79

20 a.u. electric field ellipse


< >

24 I (4) I
~ ..- / /// _

~ -2 // -
co
.0
.'.~.7.~.'.~.....~.."~
~..~ ~....... // (2) '

~- -28 -1
o -'-. '.1/- - ,/ .

4
Eo -30 "''"''" "" 9- " ...... ""121
0

_9o
~-32 - - " 'sumlZ 13l2'. ,612""... 1112~__t
, , ~ . . . . a . . . . t.,
" , , , ~ ,""., , , ~ , , ,~tl
" '

30 4O 50 60 70
harmonic order
FiG. 17. High-order harmonic spectrum for an elliptically polarized laser field with the same
parameters as in Fig. 14 and harmonic orders between 25 and 77. The open circles are calculated
from the integral (57), and the curves labeled 1 through 6 represent the individual contributions to
the quantum-orbit approximation of the six shortest quantum orbits, numbered as in Fig. 14. The
contributions from quantum orbits 2, 4 and 6 have to be dropped above their intersections with
curves 1, 3 and 6, respectively. The coherent sum of all six orbits is represented by the solid line.
Typical orbits responsible for each part of the spectrum are depicted as in Fig. 14. From Kopold
et aL (2000b).

see Eichmann et al. (1995) for experimental results and Long et al. ( 1 9 9 5 ) for
a theoretical description. We will call this field "bicircular." This high efficiency
was surprising because, for a monochromatic field, the harmonic emission
rate decreases with increasing ellipticity (cf. the preceding subsection) and a
circularly polarized laser field does not produce any harmonics at all. A more
detailed analysis, based on the quantum-orbits formalism, gives an explanation
of this effect (Milo~evi6 et al., 2000, 2001 a,b). The harmonics produced this way
can be of a practical importance because of their high intensity (Milo~evi6 and
Sandner, 2000) and temporal characteristics (attosecond pulse trains; Milo~evi6
and Becker, 2000). The more general case of an roo-s~o (with r and s integers)
bicircular field was considered by Milo~evi6 et al. (2001 a).
For the laser field (61), selection rules only permit emission of circularly
polarized harmonics with frequencies s = (3n + l)6o and helicities +1.
Similar selection rules govern harmonic generation by a ring-shaped molecule
(Ceccherini and Bauer, 2001) or a carbon nanotube (Alon et al., 2000).
80 W. Becker et al. [VI

.m
t--
~ , i . . . . . . . . . r . . . . . . . . . , . . . . . . . . . r . . . . . . . . . i . . . .

-13

ID
I,._
r

o
~ -15
E
~O
r

o
E
~- -17'
4
O

-19 ...........................................................
10 20 30 40 50 60
harmonic order

Flo. 18. Harmonic-emission rate as a function of the harmonic order for the bicircular laser
field (61) with 09 = 1.6eV and intensities I l = 12 = 4x1014 W / c m 2. The ionization potential is
Eip = 15.76 eV (argon). The inset shows the laser electric-field vector in the x - y plane for times
- 89T ~< t ~< 89T, with T = 23r/o9 being the period of the field (61). The arrows indicate the time
evolution of the field. The ionization time to and the recombination time t I of the three harmonics
Q = 19~o, 31o9 and 43~o are marked by asterisks and solid circles, respectively. These times and
harmonics correspond to the dominant saddle-point solution 2 in Fig. 19. In between the ionization
time (asterisks) and the recombination time (solid circles) the x-component of the electric field
changes from its negative maximum to its positive maximum, whereas its y-component remains
small and does not change sign. From Milo~evi6 et al. (2000).

Figure 18 presents an example of the harmonic spectrum for the bicircular


field (61). The results are obtained from Eq. (57) by numerical integration.
Compared with the spectrum of a monochromatic linearly polarized field (see,
for example, the nonrelativistic curve in Fig. 22), the spectrum is smooth.
Furthermore, the cutoff is less pronounced and there are small oscillations after
the cutoff. These features can be explained in terms of the quantum orbits.
Figure 19a shows the first eleven solutions (those having the shortest travel times)
of the system of the saddle-point equations (34), (35) and (60), while Fig. 19b
shows the individual contributions to the harmonic emission rate of the first
eight of these solutions (Milo~evi6 et al., 2000). Obviously, in the plateau region
the contribution of a single orbit, corresponding to solution 2, is dominant by
one order of magnitude, while in the cutoff region more solutions are relevant
(in particular solution 5). This is just the opposite of the standard situation of
the monochromatic linearly polarized field (Lewenstein et al., 1995b) where
essentially two orbits contribute in the plateau and just one in the cutoff region.
Figure 19a shows which solutions are dominant. The probability of harmonic
VI] ATI: CLASSICAL TO QUANTUM 81

0.05 . . . . . . . .

2 4 5 7 8 10 11 (a)
1

6 4 52 13 42 15 413
~,- 0 J(7 7 25 ~
24 ~ 7
21
,,.~'~"'~ 25
E
-- 8 17 50,1 1 4

-0.05
0 05 1 15 2
Re (tl-t0)/T

"~ -11

-13 -- \ '- " " 7

g9 -15 11 ~ ~ ~ ~ . ~ . / "

/
"~___ 4 "" _ _ " , - ~ . \- - " .. .... I
.0
m 9 , \ I

;," "",
t- '...' ,. ~ _ ~ ~ ". " \
',, ,,,s
/ I " ~ ~ ~ ~ ~ ~

O a
,-E -17
9

t-
v
o

o
-19
10 20 30 40 50 60
harmonic order
Fie. 19. Saddle-point analysis of the results of Fig. 18. (a) The imaginary part of the recombination
time tl as a function of the real part of the travel time tl - t o , obtained from the solutions of the
saddle-point equations (34), (35) and (60). Each point on the curves corresponds to a specific value
of the harmonic frequency f2, which is treated as a continuous variable. For the interval of Re(t1 - to)
covered in the figure, eleven solutions were found, which are labeled with the corresponding numbers
at the top and bottom of the graph. Values of the harmonic order that approximately determine the
cutoffs for each particular solution are marked by stars with the corresponding harmonic numbers
next to them. Those values of the harmonic order for which I Im tl] is minimal are identified as
well. (b) The partial contributions to the harmonic-emission rate of each of the first eight solutions
of the saddle-point equations. From Milo~evi6 et al. (2000).
82 W. Becker et al. [VI

/ ,,
,/ "X / \
/ , / \
I \ i X
20 i" "\jl \\

'1 ii '\ x
/ /..---,~.. x
/ XX5

"7'.
-', 0 i

>,,

i " -~"-.-\--~\ ~ " ~


]2.-,,,
/ '\ 3 ""4

-20 /
,/ 1
/ ,/
/ t"

-40 .....................................
-20 0 20 40
x [a.u.]

FIG. 20. Real parts of the quantum orbits for the same parameters as in Fig. 18 and for the
harmonic fl = 43r Five orbits are shown that correspond to the saddle-point solutions 2, 3, 4, 5
and 8 in Fig. 19. The direction of the electron's travel is given by the arrows. In each case, the
electron is "born" a few atomic units away from the position of the ion (at the origin), where its
orbit almost exactly terminates. The dominant contribution to the 43rd harmonic intensity comes
from the shortest orbit number 2, whose shape closely resembles the orbit in the case of a linearly
polarized monochromatic field. From Milo~evi~ et al. (2000).

emission decreases with increasing absolute value of the imaginary part of the
recombination time tl. The possible cutoff of the harmonic spectrum can be
defined as the value of the harmonic order after which I Im tll becomes larger
than (say) 0.01T. The probability of HHG is maximal when I Im tll is minimal.
For each solution in Fig. 19a, these points are marked by asterisks and by the
corresponding harmonic order. As a consequence of wave-function spreading,
the emission rate decreases with increasing travel time ti - t o . This gives an
additional reason why the contribution of solution 2 is dominant in the plateau
region.
Let us now consider the quantum orbits. In Fig. 20 for the fixed harmonic
= 43~o, we present the five orbits that correspond to saddle-point solutions
2, 3, 4, 5 and 8 in Fig. 19. The dominant contribution comes from the shortest
orbit 2 (thick line). It starts at the point (4.06, 0.66) by setting off in the negative
y-direction, but soon turns until it travels at an angle of 68 ~ to the negative y-axis.
Thereafter, it is essentially linear, as would be the case for a linearly polarized
field. This behavior can be understood by inspection of the driving bicircular field
depicted in the inset of Fig. 18, where the start time and the recombination time
of the orbit are marked. During the entire length of the orbit, the field exerts
a force in the positive y-direction. The effect of this force is canceled by the
electron's initial velocity in the negative y-direction. The force in the x-direction
VI] ATI: C L A S S I C A L TO Q U A N T U M 83

0
I

, , i , , i ,

0 0'.2 0.4 0'.6 0.8 1


time [optical cycle]
FIG. 21. Parametric polar plot of the electric-field vector of a group of harmonics during one cycle
of the bicircular field (61) on an arbitrary isotropic scale. The position of the origin is indicated in
the upper and the left margin. The parameters are 11 = 12 = 9.36• 1014 W/cm 2, h~o = 1.6eV, and
E[p = 24.6eV. The plot displays two traces: The circular trace is generated by the ten harmonics
= (3n + 1)co with n = 10. . . . . 19, all having positive helicity. The starlike trace is generated by all
harmonics ~ = (3n + 1)co between the orders 31 and 59, regardless of their helicity. The curve at
the bottom represents the x-component of the field of the latter group over one cycle, the time scale
being given on the horizontal axis. It shows that the field is strongly chirped. The black blob at the
center is due to the fact that the field is near zero throughout most of the cycle, cf. the trace of the
x-component. From Milo~evi6 and Becker (2000).

is m u c h like that in the case o f a linearly p o l a r i z e d driving field. Since H H G by a


linearly p o l a r i z e d field is m o s t efficient, this m a k e s plausible the high efficiency
o f H H G by the bicircular field.
The orbit that c o r r e s p o n d s to solution 3 has a shape similar to that o f
orbit 2, but is m u c h longer. The c o r r e s p o n d i n g travel time is longer, too,
and, consequently, the c o n t r i b u t i o n o f s o l u t i o n 3 to the e m i s s i o n rate o f the
43rd h a r m o n i c is smaller. The other orbits are still l o n g e r and m o r e c o m p l i c a t e d
so that their c o n t r i b u t i o n is negligible.
The electric field o f a g r o u p o f p l a t e a u h a r m o n i c s is displayed in Fig. 21. It
shows interesting behavior, w h i c h again reflects the t h r e e f o l d s y m m e t r y o f the
field (61), see the inset o f Fig. 18. If the g r o u p o f h a r m o n i c s includes h a r m o n i c s
o f either parity, then the field consists o f a s e q u e n c e o f essentially linearly
polarized, strongly c h i r p e d a t t o s e c o n d pulses, each rotated by 120 ~ with r e s p e c t
to the previous one. If, on the other hand, one were able to select h a r m o n i c s o f
definite helicity, i. e. either ~2 = (3n + 1)co or ~ = ( 3 n - 1)co, then one w o u l d
84 W. B e c k e r et al. [VI

obtain a sequence of attosecond pulses with approximately circular polarization.


Both cases are illustrated in Fig. 21.

D. HHG IN THE RELATIVISTIC REGIME

Quantum orbits can also be employed in the relativistic regime starting from
the Klein-Gordon equation (55). Milo~evi6 et al. (2001c, 2002) found that
the relativistic harmonic-emission matrix element has a form similar to that in
Eq. (57), but with the relativistic action (h = c = 1)

Sfa(tl, to, k) = L Cx~du (EIp - m - f2) - L tl du ek(u) + ftOcxDdu (EIp - m), (62)

where
k + ~A(u)
ek(u) = Ek + eA(u) 9 (63)
Ek-~.k

and Ek = ( k 2 + m2) 1/2, u ( t - z)/co. Solving the classical Hamilton-Jacobi


- -

equation for Hamilton's principal function it can be shown that ek(u) is the
classical relativistic electron energy in the laser field. In the relativistic case,
the function m E ( t l , t o , k ) in Eq. (57) consists of two parts: the dominant part
is responsible for the emission of odd harmonics ~ = (2n + 1)~, while
the other one originates from the intensity-dependent drift momentum of the
electron in the field and allows for emission of even harmonics ~ = 2n~.
Similarly to the nonrelativistic case, the integral over the intermediate electron
momentum k can be calculated by the saddle-point method. The stationarity
condition ftto' du Oek(u)/Ok = 0, with Oek/Ok = d r / d t , implies r(t0) = r(tl), so
that the stationary relativistic electron orbit is such that the electron starts from
and returns to the nucleus. As above, the start time and, to a lesser degree, the
recombination time are complex.
In the relativistic case, the stationary momentum k = ks is introduced in
the following way. For fixed to and tl, its component ks• perpendicular to the
photon's direction of propagation f~ is given by

tl
(tl - to) k s • =
L du cA(u). (64)

Introducing .A/I 2 = e 2 fttoI d r / A 2 ( u ) / ( t l - to)- ks2 > O, one h a s

(.A/j2 _ ks2_L)2
k 2 = ks2 + 4 ( m 2 + .A/j2) , (65)
VI] ATI: C L A S S I C A L TO Q U A N T U M 85

-20

~0
~
r
nonrelativistic
= -40
_ _Z [a'u']
-100 -50 0 ~ relativistic
~- -60
t-
O 750 _
ffl
ffl

l
-80
o
.m
t-
O
E
I._

c.- -100
o
T-

O}
0

-120
0 50000 100000 150000
harmonic order

Fie. 22. Harmonic-emission rate as a function of the harmonic order for ultrahigh-order harmonic
generation by an Ar8+ ion (Eip = 422eV) in the presence of an 800-nm Ti:Sa laser having the
intensity 1.5• 1018 W/cm 2. Both the nonrelativistic and the relativistic results are shown. The
corresponding relativistic electron orbit with the shortest travel time that is responsible for the
emission of the harmonic ~ -- 100000o9 is shown in the inset. The arrows indicate which way the
electron travels the orbit. The laser field is linearly polarized in the x-direction and the v • B electron
drift is in the z-direction. From Milo~evi6 et al. (2002).

which yields ek~ as a function of to and tl. The two stationarity equations
connected with the integrals over to and tl are

eks (to) - m - EIp, (66)


~2 - t~k~(tl) + EIp - m. (67)

As in the nonrelativistic case, they express energy conservation at the time of


tunneling to and at the time of recombination t~, respectively. The final expression
for the relativistic harmonic-emission matrix element has the form (37) with (62),
where the summation is now over the appropriate subset o f the relativistic saddle
points ( 6 s , tos, ks) that are the solutions of the system of equations (64)-(67).
In the relativistic case it is very difficult to evaluate the harmonic-emission
rates by numerical integration. For very high laser-field intensities and ultra-high
harmonic orders, this is practically impossible, so that the saddle-point m e t h o d
is the only way to produce reasonable results. Figure 22 presents an example.
The nonrelativistic result is obtained from Eq. (37) where the summation is
over the solutions of the system o f the nonrelativistic saddle-point equations
(34), (35) and (60). It is, o f course, inapplicable for the high intensity o f
86 W. B e c k e r et al. [VII

1.5x1018 W/cm 2 at 800nm and is only shown to demonstrate the dramatic


impact of relativistic kinematics. For the relativistic result, the summation in
Eq. (37) is over the relativistic solutions of Eqs. (64)-(67). The relativistic
harmonic-emission rate assumes a convex shape, and the difference between
the relativistic and nonrelativistic results reaches several hundred orders of
magnitude in the upper part of the nonrelativistic plateau. The origin of this
dramatic suppression is the magnetic-field-induced v x B drift. The significance
of this drift for the rescattering mechanism was emphasized early by Kulyagin
et al. (1996). This is illustrated in the inset of Fig. 22, which shows the real
part of the dominant shortest orbit for the harmonic f2 = 100000o). In order to
counteract this drift so that the electron is able to return to the ion, the electron
has to take off with a very substantial initial velocity in the direction opposite
to the laser propagation. The probability of such a large initial velocity is low,
and this is the reason for the strong suppression. As in the nonrelativistic case,
the electron is "born" at a distance of 7.5 a.u. from the nucleus.
The nonrelativistic harmonic yield shows a pronounced multiplateau structure.
While this is an artifact of the nonrelativistic approximation for the intensity
of Fig. 22, it is a real effect for lower laser-field intensities where relativistic
effects are still small (Walser et al., 2000; Kylstra et al., 2001; Milogevid et al.,
2001 c, 2002). In this case, the three plateaus visible in the nonrelativistic curve of
Fig. 22 are related to the three pairs of orbits, whose contribution to the harmonic
emission rate is dominant in the particular spectral region (see Figs. 2 and 3
of Milogevid et al., 2001c). These are very similar to the pairs of orbits that
we have discussed for the elliptically polarized laser field in Fig. 17. However,
for the very high intensity of Fig. 22, the contribution of the shortest of these
orbits becomes so dominant that the multiplateau and the interference-related
oscillatory structure disappear completely. The reason is that the effect of the
v x B drift increases with increasing travel time; see Eqs. (52) and (56) in
Sect. V.A. This is in contrast to the nonrelativistic case of elliptical polarization,
where longer orbits may be favored because the minor component of the field
oscillates and, therefore, for a longer orbit a smaller initial velocity may be
sufficient to allow the electron to return.

VII. Applications of ATI

Experimental and theoretical advances in understanding A T I - some of which


have been treated in this review - permit its application to the investigation of
other effects. One obvious idea is to exploit the nonlinear properties of ATI.
This is particularly relevant to characterization of high-order harmonics and
measurement of attosecond pulses in the soft-X-ray regime. In this spectral
region (vacuum UV) virtually all bulk non-linear media are opaque. ATI,
in contrast, is usually studied under high- or ultra-high-vacuum conditions.
VII] ATI: CLASSICAL TO QUANTUM 87

Another advantage over conventional nonlinear optics is that the nonlinear effect
of photoelectron emission can be observed from more or less any direction,
whereby different properties of the effect can be exploited.

A. CHARACTERIZATION OF H I G H HARMONICS

The most straightforward approach to characterize high-order harmonics is a


cross-correlation scheme: An (isolated) harmonic of frequency qoo, where q is
an odd integer, produces electrons by single-photon ionization with a kinetic
energy Eq = qhoo - EIp. Simultaneous presence of a fraction of the fundamental
laser beam in the near infrared (NIR) produces sidebands, i.e. electrons with
energies qhoo- Ew + mhoo (m << q). The strength of the sidebands can be
changed by temporally delaying the fundamental with respect to the harmonic
by a time r. Optimal overlapping of the pulses (r = 0) leads to a maximum in
the strength of the sidebands, whereas complete separation entirely eliminates
them. The strength of the sidebands as a function of r can be used to determine
the duration of the harmonic pulse.
For theoretical modeling, the simple ansatz of Becker et al. (1986) can be used,
which assumes that an electron is born in the presence of the laser field with a
positive initial energy Ei, which will be identified with Eq. For Up << hoo, which
is well satisfied for the weak field we will consider, the differential ionization
rate in the field direction is given by (in atomic units)

02F
oc Ipl- m =-~']2 ~f V/2(Eio)2+ moo) O(E - moo - Ei). (68)
OEOff2

Here, ,5'f is the amplitude of the electric field of the fundamental, m is the order
of the sideband, p is the momentum of the photoelectron (]p[ = V/2(Ei + moo)),
and tim is the Bessel function of the first kind. The intensities of the side bands
are not, in general, symmetric. However, for sufficiently weak fields, both fields
can be treated by lowest-order perturbation theory. It follows that a sideband of
,r r m ]
order m is proportional to "~h'~y , where ,5'h is the field strength of the harmonic
radiation. In this case, the cross correlation for a sideband of order m can be
calculated as (X3

Cm(T ) = f g~(t)" E~lml(t - r)dt. (69)


--OO

Figure 23 shows the result of a corresponding calculation, which is compared


with results from a numerical solution of the appropriate one-dimensional
Schr6dinger equation. The agreement is nearly perfect.
Hence, if the NIR pulse is precisely known, the pulse duration of the harmonic
(and even its shape) can be determined by deconvolution of the cross-correlation
88 W. Becker et al. [VII

+ m=+l + m=+2 + m=+3


x rn=-I x m=-2 x m=-3
i i• = i i i i = 1 =

w w w w

- s-10-s 0 s lo -10-5 o s To--- 0-5 0 5 10 - ' - 15


"

delay [number of optical cycles]

Fie. 23. Cross-correlation of near-infrared and soft-X-ray pulses. A harmonic of order q creates
photoelectrons at the kinetic energy qho9- Eip. Sidebands are created by simultaneous irradiation
with the fundamental of frequency o9. Plotted are the heights of the sidebands for various side-band
orders m versus the delay r between the fundamental and the harmonic. The solid line represents
the analytical approximation (69), whereas the points were calculated by numerically solving the
appropriate (one-dimensional) Schr6dinger equation. In each case, the analytical approximation was
normalized to the maximum of the numerical result.

functions. Numerical and experimental investigations of this problem were made


by V6niard et al. (1995) and Schins et al. (1996), respectively.

A. 1. Measurement o f attosecond pulses

Clearly, an experiment as discussed above will not be able to determine


harmonic-pulse durations significantly shorter than that of the fundamental
in the NIR spectral region. In 1996 already, V~niard et al. pointed out that
the cross correlation of harmonic and NIR radiation provides access to the
relative phase of neighboring harmonics. This is an extremely important insight
because the phase dependence of the harmonics as a function of their order
(or frequency) determines whether they are mode-locked and whether the
corresponding pulses - which would constitute attosecond pulses in the soft-
X-ray region if bandwidth-limited- are chirped. In fact, Paul et al. (2001) used
this scheme for the first observation of a train of attosecond pulses.
In order to achieve phase measurement of adjacent harmonics, the conditions
have to be chosen such that only sidebands of order m = +1 are generated
with appreciable amplitude. This calls for intensities of the NIR beam be-
low 1012 W / c m 2. Along with the fact that the NIR field generates only odd-
order harmonics this ensures that only two adjacent harmonics contribute to each
sideband. An electron with energy Eq - q h o o - EIp, with q an even integer,
VII] ATI: CLASSICAL TO QUANTUM 89

1.35 fs,
i

10

2~.50 as
5

0 , ! 9
0 ~1 ,,,~ 3 ', 4 " 5
\ I
\ I
, \ /
,, t i m e [fs] ' \ /
'

FIG. 24. Reconstruction of a train of attosecond pulses synthesized from the five harmonics
q = 11 . . . . . 19. The attosecond pulses are separated by 1.35 fs, which is half the cycle time of the
driving laser. The latter is represented by the dashed cosine function. Reprinted with permission from
Paul et al. (2001), Science 292, 1689, fig. 4. 9 2001 American Association for the Advancement
of Science.

can be generated by absorption of the lower harmonic plus one NIR photon
(Eq -- ( q - 1)ho) + boo) or by absorption of the upper harmonic and emission
of one NIR photon (Eq = (q + 1)hco- ha)). Each of these two channels receives
contributions from two different quantum paths, which are related to the temporal
order of the interaction with the harmonic and the NIR field. (In contrast to
the quantum orbits we considered elsewhere in this chapter, the quantum paths
here are defined in state space rather than position space.) The photoelectron
yield at energy Eq is proportional to the square of the (coherent) sum of the
amplitudes of all four quantum paths. Due to the fact that two paths represent
absorption from the NIR field whereas the other two represent emission into it,
the interference term between these two contributions is essentially proportional
to cos(~q_l --~q+l 4- 2htor). By varying the delay r between the harmonic and
the NIR radiation, the difference ~q-1 --~q+l of the phases of the two harmonics
can be recorded. The result of the corresponding experiment (Paul et al., 2001)
is that the phase of the harmonics depends almost linearly on their frequency.
Hence, the harmonics considered in the experiment (q = 11 to 19) are mode-
locked and make up a train of attosecond pulses of 250 as FWHM duration, see
Fig. 24.

A.2. Isolated attosecond pulses


With respect to applications, isolated attosecond pulses appear more useful than
a train of pulses separated by half the period of the fundamental. Isolated
attosecond pulses could be generated by sufficiently short fundamental pulses,
i.e. pulses of about 5 fs, which consist of less than two optical cycles (few-cycle
90 W. B e c k e r et al. [VII

regime). Then, however, the spectral width of the harmonics will be so broad
that it is no longer possible to identify individual sidebands as necessary for the
method of Paul et al. (2001).
Nevertheless, Drescher et al. (2001) and Hentschel et al. (2001) succeeded
in performing measurements of the harmonic-pulse length with a resolution of
1.8 fs and 150 as, respectively. The experimental setup, in principle, resembles
that of Paul et al. with the difference that higher intensities of the NIR ra-
diation are used for the photoionization cross correlation. In addition, only
photoelectrons ejected perpendicularly to the laser polarization are detected.
The motivation for choosing these conditions can be deduced from a classical
analysis of trajectories of electrons that were injected into the electric field of
the few-cycle NIR pulse by absorption of a harmonic photon. If the duration of
the X-ray pulse is shorter than the optical period T in the NIR, then the final
kinetic energy of the photoelectrons depends on the phase tot0 when the injection
took place, i. e. it exhibits a modulation with a period of T/2. By delaying
the fundamental with respect to the harmonic, the modulation can be recorded.
This was done in the experiment of Drescher et al. (2001). Hentschel et al.
(2001) realized that the width of the photoelectrons' kinetic energy distribution
also exhibits such a modulation, and is measureable with much higher precision
than the center of mass of the distribution. For the two approaches, it is not the
envelope of the fundamental that enters the correlation function, but rather the
optical period. The restriction to photoelectrons emitted perpendicularly to the
laser polarization suppresses the influence of effects related to the emission and
absorption of photons from the laser field, i.e. the sidebands which were crucial
for the experiment of Paul et al. (2001).

B. THE "ABSOLUTE PHASE" OF FEW-CYCLE LASER PULSES

The need for highest intensities and extremely broad bandwidths in several areas
of the natural sciences is driving the development to shorter and shorter laser
pulses. At a FWHM duration shorter than a few optical cycles the time variation
of the pulse's electric field depends on the phase q~ of the carrier frequency with
respect to the center of the envelope, the so-called "absolute phase." The electric
field should be written as

E(t) = C0(t)ex cos(rot + ~), (70)

where the function g0(t) is maximal at t = 0. Clearly, for a long pulse the phase q~
can be practically eliminated by resetting the clock. For a short pulse, however,
the shape of the field (70) strongly depends on this phase, which, therefore, will
influence various effects of the laser-atom interaction. This is one reason for the
significance of this new parameter of laser pulses. The precise knowledge and
control of the absolute phase will pave the way to new regimes in coherent X-ray
VII] ATI: CLASSICAL TO QUANTUM 91

FIG. 25. Evidence of absolute-phase effects from few-cycle laser pulses. In this contingency map,
every laser shot is recorded according to the number of photoelectrons measured in the left and
the right arm of the "stereo" ATI spectrometer. The number of laser shots with electron numbers
according to the coordinates of the pixel is coded in grey shades. For visual convenience the darkest
shades were chosen for medium numbers of laser shots. (The most frequent result of the laser pulses
was about 5 electrons in each of both arms.) The signature of the absolute phase is an anticorrelation
in the number of electrons recorded with the left and the right detector. In the contingency map they
form a structure perpendicular to the diagonal. Shown here is a measurement with krypton atoms
for circular laser polarization, a pulse duration of 6 fs, and an intensity of 5 x 1013 W/cm 2. From
Paulus et al. (2001b).

generation and attosecond generation; for an overview see Krausz (2001). In


addition, such extremely well-defined laser pulses are likely to have applications
for the coherent control of chemical reactions and other processes. Another
reason is that phase control of femtosecond laser pulses has already had a huge
impact on frequency metrology. This is because phase-stabilized femtosecond
lasers can be viewed as ultra-broadband frequency combs that can be used to
measure optical frequencies with atomic-clock precision; see, e.g., Jones et al.
(2000).
With current laser technology, only femtosecond laser oscillators can be phase-
stabilized (Reichert et al., 1999; Apolonski et al., 2000), which is sufficient for
frequency metrology. Strong-field effects require amplified laser pulses. Nisoli
et al. (1997) demonstrated that it is possible to generate powerful (>500~tJ)
laser pulses in the few-cycle regime. However, these are not stabilized and,
accordingly, the absolute phase changes in a random fashion from pulse to pulse.
92 W. B e c k e r et al. [IX

In a recent experiment, Paulus et al. (2001b) were able to detect effects due
to the absolute phase by performing a shot-to-shot analysis of the number of
photoelectrons emitted in opposite directions. To this end, a field-free drift tube is
placed symmetrically around the target gas. Each end of the tube is equipped with
an electron detector. Because of its characteristic appearance, this was dubbed a
stereo-ATI spectrometer.
A characteristic feature of few-cycle pulses such as (70) is that, depending
on the absolute phase, the peak electric-field strength (and thus also the vector
potential) is different in the positive and negative x-directions. Recall from
Eq. (2) that the electron's drift momentum depends on the vector potential at
its time of birth. Therefore, depending on the value of the absolute phase,
such a laser pulse creates more electrons in one direction than in the other.
A theoretical analysis of the photoelectrons' angular distribution was given by
Dietrich et al. (2000) and Hansen et al. (2001) for the nonperturbative intensity
regime. Interestingly, the effect is predicted to be much less pronounced in the
perturbative regime (Cormier and Lambropoulos, 1998). Equivalent to a left-
right asymmetry of the photoelectron angular distribution is that the number
of electrons emitted to the left vs. those emitted to the right is anticorrelated:
A laser shot for which many electrons are seen at the right detector is likely to
produce only a few that go left, and vice versa. This can be proved by correlation
analysis. Each laser shot is sorted into a contingency map according to the
number of electrons recorded at both detectors. Anticorrelations can then be seen
in structures perpendicular to the diagonal, see Fig. 25.

VIII. Acknowledgments

We learned a lot in discussions with S.L. Chin, M. DSrr, C. Faria, S.P.


Goreslavskii, C.J. Joachain, M. Kleber, V.P. Krainov, M. Lewenstein, A. Lohr,
H.G. Muller, S.V. Popruzhenko, and W. Sandner. This work was supported in
part by Deutsche Forschungsgemeinschaft and Volkswagen Stiftung.

IX. References
Agostini, P., Fabre, E, Mainfray, G., Petite, G., and Rahman, N.K. (1979). Phys. Rev. Lett. 42, 1127.
Agostini, E, Antonetti, A., Breger, E, Crance, M., Migus, A., Muller, H.G., and Petite, G. (1989).
J. Phys. B 22, 1971.
Alon, O.E., Averbukh, V., and Moiseyev, N. (2000). Phys. Rev. Lett. 85, 5218.
Ammosov, M.V., Delone, N.B., and Krainov, V.P. (1986). Zh. Eksp. Teor. Fiz. 91, 2008 [Soy. Phys.-
JETP 64, 1191].
Antoine, P., l'Huillier, A., Lewenstein, M., Sali6res, P., and CarrY, B. (1996). Phys. Rev. A 53, 1725.
Antoine, Ph., Gaarde, M., Sali~res, P., CarrY, B., l'Huillier, A., and Lewenstein, M. (1997). In
"Multiphoton Processes 1996" (P. Lambropoulos and H. Walther, Eds.), Institute of Physics
Conference Series No. 154. Institute of Physics Publishing, Bristol, p. 142.
IX] ATI" C L A S S I C A L TO Q U A N T U M 93

Apolonski, A., Poppe, A., Tempea, G., Spielmann, Ch., Udem, Th., Holzwarth, R., H~insch, T.W.,
and Krausz, E (2000). Phys. Rev. Lett. 85, 740.
Bao, D., Chen, S.G., and Liu, J. (1996). Appl. Phys. B 62, 313.
Bartels, R., Backus, S., Zeek, E., Misoguti, L., Vdovin, G., Christov, I.P., Murnane, M.M., and
Kapteyn, H.C. (2000). Nature (London) 406, 164.
Bartels, R., Backus, S., Christov, I., Kapteyn, H., and Murnane, M. (2001). Chem. Phys. 267, 277.
Bashkansky, M., Bucksbaum, P.H., and Schumacher, D.W. (1988). Phys. Rev. Lett. 60, 2458.
Becker, W., Schlicher, R.R., and Scully, M.O. (1986). s Phys. B 19, L785.
Becket, W., Schlicher, R.R., Scully, M.O., and W6dkiewicz, K. (1987). s Opt. Soc. Am. B 4, 743.
Becker, W., Long, S., and McIver, J.K. (1990). Phys. Rev. A 42, 4416.
Becker, W., Long, S., and McIver, J.K. (1992). Phys. Rev. A 46, R5334.
Becker, W., Lohr, A., and Kleber, M. (1994a). s Phys. B 27, L325. Corrigendum: 28, 1931.
Becker, W., Long, S., and McIver, J.K. (1994b). Phys. Rev. A 50, 1540.
Becker, W., Lohr, A., and Kleber, M. (1995). Quantum Semiclass. Opt. 7, 423.
Becket, W., Lohr, A., Kleber, M., and Lewenstein, M. (1997). Phys. Rev. A 56, 645.
Becker, W., Kleber, M., Lohr, A., Paulus, G.G., Walther, H., and Zacher, E (1998). Laser Phys.
8, 56.
Beigman, I.L., and Chichkov, B.N. (1987). Pis "ma Zh. Eksp. Teor. Fiz. 46, 314 [JETP Lett. 46, 395].
Bellini, M., LyngA, C., Tozzi, A., Gaarde, M.B., H/insch, T.W., l'Huillier, A., and Wahlstr6m, C.-G.
(1998). Phys. Rev. Lett. 81, 297.
Berson, I.J. (1975). J. Phys. B 8, 3078.
Bhardwaj, V.R., Aseyev, S.A., Mehendale, M., Yudin, G.L., Villeneuve, D.M., Rayner, D.M.,
Ivanov, M.Yu., and Corkum, P.B. (2001). Phys. Rev. Lett. 86, 3522.
Blondel, C., Delsart, C., Dulieu, E, and Valli, C. (1999). Eur. Phys. J. D 5, 207.
Borca, B., Frolov, M.V., Manakov, N.L., and Starace, A.E (2001). Phys. Rev. Lett. 87, 133001.
Bordas, C., Paulig, E, Helm, H., and Huestis, D.L. (1996). Rev. Sci. Instrum. 67, 2257.
Brabec, T., and Krausz, F. (2000). Rev. Mod. Phys. 72, 545.
Bryant, H.C., Mohagheghi, A., Stewart, J.E., Donahue, J.B., Quick, C.R., Reeder, R.A., Yuan, V.,
Hummer, C.R., Smith, W.W., Cohen, C., Reinhardt, W.P., and Overman, L. (1987). Phys. Rev.
Lett. 58, 2412.
Bucksbaum, P.H., Bashkansky, M., Freeman, R.R., Mcllrath, T.J., and DiMauro, L.F. (1986). Phys.
Rev. Lett. 56, 2590.
Bucksbaum, P.H., Bashkansky, M., and Mcllrath, T.J. (1987). Phys. Rev. Lett. 58, 349.
Bunkin, F.V., and Fedorov, M.V. (1966). Zh. Eksp. Teor. Fiz. 49, 1215 [Soy. Phys.-JETP 22, 844].
Ceccherini, E, and Bauer, D. (2001). Phys. Rev. A 64, 033423.
Chelkowski, S., and Bandrauk, A.D. (2000). Laser Phys. 10, 216.
Chin, S.L., Yergeau, E, and Lavigne, P. (1985). J. Phys. B 18, L213.
Christov, I.P., Bartels, R., Kapteyn, H.C., and Murnane, M.M. (2001). Phys. Rev. Lett. 86, 5458.
Corkum, P.B. (1993). Phys. Rev. Lett. 71, 1994.
Corkum, P.B., Burnett, N.H., and Ivanov, M.Y. (1994). Opt. Lett. 19, 1870.
Cormier, E., and Lambropoulos, P. (1997). J. Phys. B 30, 77.
Cormier, E., and Lambropoulos, P. (1998). Eur. Phys. J. D 2, 15.
Cormier, E., Garzella, D., Breger, P., Agostini, P., Ch6riaux, G., and Leblanc, C. (2001). J. Phys. B
34, L9.
Crawford, D.P., and Reiss, H.R. (1997). Opt. Express 2, 289.
Dammasch, M., D6rr, M., Eichmann, U., Lenz, E., and Sandner, W. (2001). Phys. Rev. A
64, 061402(R).
de Bohan, A., Antoine, P., Milo~evi6, D.B., and Piraux, B. (1998). Phys. Rec. Lett. 81, 1837.
Delande, D., and Buchleitner, A. (1994). Adv. At. Mol. Opt. Phys. 34, 85.
Delone, N.B., and Krainov, V.P. (1994). "Multiphoton Processes in Atoms." Springer, Berlin.
Delone, N.B., and Krainov, V.P. (1998). Usp. Fiz. Nauk 168, 531 [Phys. Usp. 41, 469].
94 W. B e c k e r et al. [IX

Demkov, Yu., and Ostrovskii, V.N. (1989). "Zero-Range Potentials and their Applications in Atomic
Physics." Plenum, New York.
Dietrich, E, Burnett, N.H., Ivanov, M., and Corkum, EB. (1994). Phys. Rev. A 50, R3585.
Dietrich, E, Krausz, E, and Corkum, EB. (2000). Opt. Lett. 25, 16.
DiMauro, L.E, and Agostini, E (1995). Adv. At. Mol. Opt. Phys. 35, 79.
D6rner, R., Mergel, V., Jagutzki, O., Spielberger, L., Ullrich, J., Moshammer, R., and Schmidt-
BScking, H. (2000). Phys. Rep. 330, 95.
D6rr, M., Potvliege, R.M., and Shakeshaft, R. (1990). Phys. Rev. A 41,558.
Drescher, M., Hentschel, M., Kienberger, R., Tempea, G., Spielmann, Ch., Reider, G.A., Corkum, P.B.,
and Krausz, F. (2001). Science 291, 1923.
Du, M.L., and Delos, J.B. (1988). Phys. Rev. A 38, 1896, 1913.
Duchateau, G., Cormier, E., Bachau, H., and Gayet, R. (2001). Phys. Rev. A 63, 053411.
Dykhne, A.M. (1960). Zh. Eksp. Teor. Fiz. 38, 570 [Soy. Phys.-JETP 11, 411].
Eberly, J.H., Javanainen, J., and Rz~2ewski, K. (1991). Phys. Rep. 204, 331.
Eberly, J.H., Grobe, R., Law, C.K., and Su, Q. (1992). Adv. At. Mol. Opt. Phys. Suppl. 1, 301.
Ehlotzky, E (2001). Phys. Rep. 345, 175.
Eichmann, H., Egbert, A., Nolte, S., Momma, C., Wellegehausen, B., Becker, W., Long, S., and
Mclver, J.K. (1995). Phys. Rev. A 51, R3414.
Faisal, EH.M. (1973). J. Phys. B 6, L89.
Faisal, F.H.M., and Rado2ycki, T. (1993). Phys. Rev. A 47, 4464.
Faisal, EH.M., and Scanzano, P. (1992). Phys. Rev. Lett. 68, 2909.
Faisal, EH.M., Filipowicz, P., and Rz~• K. (1990). Phys. Rev. A 41, 6176.
Ferray, M., l'Huillier, A., Li, X.E, Lompr6, L.A., Mainfray, G., and Manus, C. (1988). J. Phys. B
21, L31.
Figueira de Morisson Faria, C., Milo~evi6, D.B., and Paulus, G.G. (2000). Phys. Rev. A 61, 063415.
Figueira de Morisson Faria, C., Kopold, R., Becker, W., and Rost, J.M. (2002a). Phys. Rev. A
65, 023404.
Figueira de Morisson Faria, C., Schomerus, H., and Becker, W. (2002b). physics~0206028. Phys.
Rev. A, to be published.
Filipowicz, P., Faisal, F.H.M., and Rz~.ewski, K. (1991). Phys. Rev. A 44, 2210.
Fittinghoff, D.N., Bolton, P.R., Chang, B., and Kulander, K.C. (1992). Phys. Rev. Lett. 69, 2642.
Freeman, R.R., Bucksbaum, P.H., Milchberg, H., Darack, S., Schumacher, D., and Geusic, M.E.
(1987). Phys. Rev. Lett. 59, 1092.
Gaarde, M.B., Salin, E, Constant, E., Balcou, Ph., Schafer, K.J., Kulander, K.C., and l'Huillier, A.
(1999). Phys. Rev. A 59, 1367.
Gaarde, M.B., Schafer, K.J., Kulander, K.C., Sheehy, B., Kim, D., and DiMauro, L.F. (2000). Phys.
Rev. Lett. 84, 2822.
Gallagher, T.E, and Scholz, T.J. (1989). Phys. Rec. A 40, 2762.
Garton, W.R.S., and Tomkins, F.S. (1967). Astrophys. J. 158, 839.
Goreslavskii, S.P., and Popruzhenko, S.V. (1996). Zh. Eksp. Teor. Fiz. 110, 1200 [JETP 83, 661].
Goreslavskii, S.P., and Popruzhenko, S.V. (1999a). Phys. Lett. A 249, 477.
Goreslavskii, S.P., and Popruzhenko, S.V. (1999b). J. Phys. B 32, L531.
Goreslavskii, S.P., and Popruzhenko, S.V. (2000). Zh. Eksp. Teor Fiz. 117, 895 [JETP 90, 778].
G6rlinger, J., Plagne, L., and Kull, H.-J. (2000). Appl. Phys. B 71, 331.
Gottlieb, B., Kleber, M., and Krause, J. (1991). Z. Phys. A 339, 201.
Gottlieb, B., Lohr, A., Becker, W., and Kleber, M. (1996). Phys. Rev. A 54, R1022.
Gribakin, G.E, and Kuchiev, M.Yu. (1997). Phys. Rev. A 55, 3760.
Gutzwiller, M. (1990). "Chaos in Classical and Quantum Mechanics." Springer, Berlin.
Hansch, P., Walker, M.A., and Van Woerkom, L.D. (1997). Phys. Rev. A 55, R2535.
Hansen, J.P., Lu, J., Madsen, L.B., and Nilsen, H.M. (2001). Phys. Rev. A 64, 033418.
Hauge, E.H., and Stovneng, J.A. (1989). Rev. Mod. Phys. 59, 917.
IX] ATI: C L A S S I C A L TO Q U A N T U M 95

Hentschel, M., Kienberger, R., Spielmann, Ch., Reider, G.A., Milosevic, N., Brabec, T., Corkum, P.,
Heinzmann, U., Drescher, M., and Krausz, E (2001). Nature (London) 414, 509.
Hertlein, M.E, Bucksbaum, P.H., and Muller, H.G. (1997). J. Phys. B 30, L 197.
Hu, S.X., and Keitel, C.H. (2001). Phys. Rev. A 63, 053402.
Ivanov, M.Yu., Brabec, Th., and Burnett, N. (1996). Phys. Rev. A 54, 742.
Jackson, J.D. (1999). "Classical Electrodynamics," 3rd edition. Wiley, New York.
Jarofi, A., Kamifiski, J.Z., and Ehlotzky, E (1999). Opt. Commun. 163, 115.
Joachain, C.J., DSrr, M., and Kylstra, N. (2000). Adv. At. Mol. Opt. Phys. 42, 225.
Jones, D.J., Diddams, S.A., Ranka, J.K., Stentz, A., Windeler, R.S., Hall, J.L., and Cundiff, S.T.
(2000). Science 288, 635.
Kamifiski, J.Z., Jarofi, A., and Ehlotzky, E (1996). Phys. Rev. A 53, 1756.
Keldysh, L.V. (1964). Zh. Eksp. Teor. Fiz. 47, 1945 [Soy. Phys.-JETP 20, 1307].
Kibble, T.W.B. (1966). Phys. Rev. 150, 1060.
Kim, J.-H., Lee, D.G., Shin, H.J., and Nam, C.H. (2001). Phys. Rev. A 63, 063403.
Kopold, R. (2001). Ph.D. Dissertation. Munich Technical University. In German.
Kopold, R., and Becker, W. (1999). J. Phys. B 32, L419.
Kopold, R., Becker, W., and Kleber, M. (1998). Phys. Rev. A 58, 4022.
Kopold, R., Becker, W., and Kleber, M. (2000a). Opt. Commun. 179, 39.
Kopold, R., Milo~evi6, D.B., and Becker, W. (2000b). Phys. Rev. Lett. 84, 3831.
Kopold, R., Becker, W., Kleber, M., and Paulus, G.G. (2002). J. Phys. B 35, 217.
Krainov, V.P. (1999). J. Phys. B 32, 1607.
Krainov, V.P., and Shokri, B. (1995). Laser Phys. 5, 793.
Kramer, T., Bracher, C., and Kleber, M. (2001). Europhys. Lett. 56, 471.
Krausz, E (2001 ). Phys. World 14, 41.
Kroll, N.M., and Watson, K.M. (1973). Phys. Rev. A 8, 804.
Krsti6, E, and Mittleman, M.H. (1991). Phys. Rev. A 44, 5938.
Krsti6, ES., Milo~evi6, D.B., and Janev, R.K. (1991). Phys. Rev. A 44, 3089.
Kruit, P., Kimman, J., Muller, H.G., and van der Wiel, M.J. (1983). Phys. Rev. A 28, 248.
Kuchiev, M.Yu. (1987). Pis 'ma Zh. Eksp. Teor. Fiz. 45, 319 [JETP Lett. 45, 404].
Kuchiev, M.Yu., and Ostrovsky, V.N. (1999). J. Phys. B 32, L189.
Kuchiev, M.Yu., and Ostrovsky, V.N. (2001). J. Phys. B 34, 405.
Kulander, K.C., and Lewenstein, M. (1996). In "Atomic, Molecular, & Optical Physics Handbook"
(G.W. Drake, Ed.). American Institute of Physics Press, Woodbury, NY, p. 828.
Kulander, K.C., Schafer, K.J., and Krause, J.L. (1993). In "Super-Intense Laser-Atom Physics"
(B. Piraux, A. l'Huillier and K. Rz~ewski, Eds.), Vol. 316 of NATO Advanced Studies Institute,
Series B: Physics. Plenum, New York, p. 95.
Kull, H.-J., G6rlinger, J., and Plagne, L. (2000). Laser Phys. 10, 151.
Kulyagin, R.V., Shubin, N.Yu., and Taranukhin, V.D. (1996). Laser Phys. 6, 79.
Kylstra, N.J., Worthington, R.A., Patel, A., Knight, EL., Vfizquez de Aldana, J.R., and Roso, L.
(2000). Phys. Rev. Lett. 85, 1835.
Kylstra, N.J., Potvliege, R.M., and Joachain, C.J. (2001). J. Phys. B 34, L55.
Lambropoulos, P., Maragakis, E, and Cormier, E. (1998). Laser Phys. 8, 625.
Lee, D.G., Shin, H.J., Cha, Y.H., Hong, K.H., Kim, J.-H., and Nam, C.H. (2001). Phys. Rev. A
63, 021801 (R).
Lein, M., Gross, E.K.U., and Engel, V. (2001). Phys. Rev. A 64, 023406.
Leubner, C. (1981). Phys. Rev. A 23, 2877.
Lewenstein, M., Balcou, Ph., Ivanov, M.Yu., l'Huillier, A., and Corkum, P.B. (1994). Phys. Rev. A
49, 2117.
Lewenstein, M., Kulander, K.C., Schafer, K.J., and Bucksbaum, P.H. (1995a). Phys. Rev. A 51, 1495.
Lewenstein, M., Sali6res, P., and l'Huillier, A. (1995b). Phys. Rev. A 52, 4747.
l'Huillier, A., Lompr6, L.A., Mainfray, G., and Manus, C. (1983). Phys. Rev. A 27, 2503.
96 W. B e c k e r et al. [IX

Lindner, E, Dreischuh, A., Grasbon, E, Paulus, G.G., and Walther, H. (2001). IEEE J. Quantum
Electron., to be published.
Lohr, A., Kleber, M., Kopold, R., and Becker, W. (1997). Phys. Rev. A 55, R4003.
Long, S., Becker, W., and McIver, J.K. (1995). Phys. Rev. A 52, 2262.
Mainfray, G., and Manus, C. (1991). Rep. Prog. Phys. 54, 1333.
Manakov, N.L., and Fainshtein, A.G. (1980). Zh. Eksp. Teor. Fiz. 79, 751 [Sot). Phys.-JETP 52, 382].
Manakov, N.L., and Rapoport, L.E (1975). Zh. Eksp. Teor. Fiz. 69, 842 [Soy. Phys.-JETP 42, 430].
Manakov, N.L., Frolov, M.V., Starace, A.E, and Fabrikant, I.I. (2000). J. Phys. B 33, R141.
McNaught, S.J., Knauer, J.P., and Meyerhofer, D.D. (1997). Phys. Rev. Lett. 78, 626.
McPherson, A., Gibson, G., Jara, H., Johann, U., Luk, T.S., McIntyre, I.A., Boyer, K., and
Rhodes, C.K. (1987). J. Opt. Soc. Am. B 4, 595.
Milo~evi6, D.B. (2000). J. Phys. B 33, 2479.
Milo~evi6, D.B., and Becker, W. (2000). Phys. Rev. A 62, 011403(R).
Milo~evi6, D.B., and Ehlotzky, E (1998a). Phys. Rev. A 57, 5002.
Milo~evi6, D.B., and Ehlotzky, E (1998b). Phys. Rev. A 58, 3124.
Milo~evi6, D.B., and Ehlotzky, E (1998c). J. Phys. B 31, 4149.
Milo~evi6, D.B., and Sandner, W. (2000). Opt. Lett. 25, 1532.
Milo~evi6, D.B., and Starace, A.E (1998). Phys. Rev. Lett. 81, 5097.
Milo~evi6, D.B., and Starace, A.E (1999a). Phys. Rev. Lett. 82, 2653.
Milo~evi6, D.B., and Starace, A.E (1999b). Phys. Rev. A 60, 3160.
Milo~evi6, D.B., and Starace, A.E (1999c). Phys. Rev. A 60, 3943.
Milo~evi6, D.B., and Starace, A.E (2000). Laser Phys. 10, 278.
Milo~evi6, D.B., Becker, W., and Kopold, R. (2000). Phys. Rev. A 61, 063403.
Milo~evi6, D.B., Becker, W., and Kopold, R. (2001a). In "Atoms, Molecules and Quantum Dots
in Laser Fields: Fundamental Processes" (N. Bloembergen, N. Rahman and A. Rizzo, Eds.),
Conference Proceedings Vol. 71. Italian Physical Society/Editrice Compositori, Bologna, p. 239.
Milo~evi6, D.B., Becker, W., Kopold, R., and Sandner, W. (2001b). Laser Phys. 11, 165.
Milo~evi6, D.B., Hu, S., and Becker, W. (2001c). Phys. Rev. A 63, 011403(R).
Milo~evi6, D.B., Hu, S.X., and Becker, W. (2002). Laser Phys. 12, 389.
Mohideen, U., Sher, M.H., Tom, H.W.K., Aumiller, G.D., Wood II, O.R., Freeman, R.R., Bokor, J.,
and Bucksbaum, EH. (1993). Phys. Rev. Lett. 71,509.
Moore, C.I., Knauer, J.E, and Meyerhofer, D.D. (1995). Phys. Rev. Lett. 74, 2439.
Moore, C.L., Ting, A., McNaught, S.J., Qiu, J., Burris, H.R., and Sprangle, E (1999). Phys. Rev.
Lett. 82, 1688.
Moshammer, R., Feuerstein, B., Schmitt, W., Dorn, A., Schr6ter, C.D., Ullrich, J., Rottke, H.,
Trump, C., Wittmann, M., Korn, G., Hoffmann, K., and Sandner, W. (2000). Phys. Rev. Lett.
84, 447.
Muller, H.G. (1999a). Phys. Rev. A 60, 1341.
Muller, H.G. (1999b). Phys. Rev. Lett. 83, 3158.
Muller, H.G. (1999c). Laser Phys. 9, 138.
Muller, H.G. (2001a). Opt. Express 8, 44.
Muller, H.G. (2001b). Opt. Express 8, 86.
Muller, H.G. (2001c). Opt. Express 8, 417.
Muller, H.G., and Kooiman, EC. (1998). Phys. Rev. Lett. 81, 1207.
Muller, H.G., Tip, A., and van der Wiel, M.J. (1983). J. Phys. B 16, L679.
Mur, V.D., Karnakov, B.M., and Popov, V.S. (1998). Zh. Eksp. Teor. Fiz. 114, 798 [J. Exp. Theor.
Phys. 87, 433].
Mur, V.D., Popruzhenko, S.V., and Popov, V.S. (2001). Zh. Eksp. Teor. Fiz. 119, 893 [J Exp. Theor.
Phys. 92, 777].
Nandor, M.J., Walker, M.A., and Van Woerkom, L.D. (1998). J Phys. B 31, 4617.
Nandor, M.J., Walker, M.A., Van Woerkom, L.D., and Muller, H.G. (1999). Phys. Rev. A 60, R1771.
IX] ATI: C L A S S I C A L TO Q U A N T U M 97

Nisoli, M., De Silvestri, S., Svelto, O., Szip6cs, R., Ferencz, K., Spielmann, Ch., Sartania, S., and
Krausz, E (1997). Opt. Lett. 22, 522.
Nurhuda, M., and Faisal, F.H.M. (1999). Phys. Rev. A 60, 3125.
Ortner, J., and Rylyuk, V.M. (2000). Phys. Rev. A 61, 033403.
Parker, J.S., Moore, L.R., Meharg, K.J., Dundas, D., and Taylor, K.T. (2001). J. Phys. B 34, L69.
Patel, A., Protopapas, M., Lappas, D.G., and Knight, P.L. (1998). Phys. Rec. A 58, R2652.
Paul, P.M., Toma, E.S., Breger, P., Mullot, G., Aug6, E, Balcou, Ph., Muller, H.G., and Agostini, P.
(2001). Science 292, 1689.
Paulus, G.G. (1996). "Multiphotonenionisation mit intensiven, ultrakurzen Laserpulsen." Utz,
Mfinchen.
Paulus, G.G., Becker, W., Nicklich, W., and Walther, H. (1994a). J. Phys. B 27, L703.
Paulus, G.G., Nicklich, W., and Walther, H. (1994b). Europhys. Lett. 27, 267.
Paulus, G.G., Nicklich, W., Xu, H., Lambropoulos, P., and Walther, H. (1994c). Phys. Rev. Lett.
72, 2851.
Paulus, G.G., Becker, W., and Walther, H. (1995). Phys. Rev. A 52, 4043.
Paulus, G.G., Zacher, F., Walther, H., Lohr, A., Becker, W., and Kleber, M. (1998). Phys. Rev. Lett.
80, 484.
Paulus, G.G., Grasbon, E, Dreischuh, A., Walther, H., Kopold, R., and Becker, W (2000). Phys.
Rev. Lett. 84, 3791.
Paulus, G.G., Grasbon, E, Walther, H., Kopold, R., and Becker, W. (2001a). Phys. Rev. A
64, 021401 (R).
Paulus, G.G., Grasbon, E, Walther, H., Villoresi, P., Nisoli, M., Stagira, S., Priori, E., and
De Silvestri, S. (2001b). Nature (London) 414, 182.
Paulus, G.G., Grasbon, E, Walther, H., Nisoli, M., Stagira, S., Sansine, G., and De Silvestri, S.
(2002). To be published.
Perelomov, A.M., and Popov, V.S. (1967). Zh. Eksp. Teor. Fiz. 52, 514 [Soy. Phys.-JETP 25, 336].
Perelomov, A.M., Popov, V.S., and Terent'ev, M.V. (1966a). Zh. Eksp. Teor. Fiz. 50, 1393 [Soy.
Phys.-JETP 23, 924].
Perelomov, A.M., Popov, V.S., and Terent'ev, M.V. (1966b). Zh. Eksp. Teor. Fiz. 51, 309 [Soy.
Phys.-JETP 24, 207].
Popov, V.S., Mur, V.D., and Karnakov, B.M. (1997). Pis 'ma Zh. Eksp. Teor. Fiz. 66, 213 [JETP Lett.
66, 229].
Popruzhenko, S.V., Goreslavskii, S.P., Korneev, P.A., and Becker, W. (2002). Phys. Rev. Lett.
89, 023001.
Protopapas, M., Keitel, C.H., and Knight, P.L. (1997a). Rep. Progr. Phys. 60, 389.
Protopapas, M., Lappas, D.G., and Knight, P.L. (1997b). Phys. Rev. Lett. 79, 4550.
Raczyfiski, A., and Zaremba, J. (1997). Phys. Lett. A 232, 428.
Reichert, J., Holzwarth, R., Udem, Th., and H~insch, T.W. (1999). Opt. Commun. 172, 59.
Reichle, R., Helm, H., and Kiyan, I.Yu. (2001). Phys. Rev. Lett. 87, 243001.
Reiss, H.R. (1980). Phys. Rev. A 22, 1786.
Reiss, H.R. (1990). J. Opt. Soc. Am. B 7, 574.
Reiss, H.R. (1992). Prog. Quantum Electron. 16, 1.
Reiss, H.R. (1996). Phys. Rev. A 54, R1765.
Robustelli, D., Saladin, D., and Scharf, G. (1997). Helv. Phys. Acta 70, 96.
Rottke, H., Wolff, B., Brickwedde, M., Feldmann, D., and Welge, K.H. (1990). Phys. Rev. Lett.
64, 404.
Salibres, P., l'Huillier, A., Antoine, Ph., and Lewenstein, M. (1999). Adv. At. Mol. Opt. Phys. 41, 83.
Salibres, P., Carr6, B., le D6roff, L., Grasbon, F., Paulus, G.G., Walther, H., Kopold, R., Becker, W.,
Milo~evi6, D.B., Sanpera, A., and Lewenstein, M. (2001). Science 292, 902.
Schins, J.M., Breger, P., Agostini, P., Constantinescu, R.C., Muller, H.G., Bouhal, A., Grillon, G.,
Antonetti, A., and Mysyrowicz, A. (1996). J. Opt. Soc. Am. B 13, 197.
98 W. B e c k e r et al. [IX

Schomerus, H., and Sieber, M. (1997). J. Phys. A 30, 4537.


Schulman, L. (1977). "Techniques and Applications of Path Integration." Benjamin, New York.
Scrinzi, A., Geissler, M., and Brabec, Th. (1999). Phys. Rev. Lett. 83, 706.
Sheehy, B., Lafon, R., Widmer, M., Walker, B., DiMauro, L.E, Agostini, P.A., and Kulander, K.C.
(1998). Phys. Rev. A 58, 3942.
Smirnov, M.B., and Krainov, V.P. (1998). J. Phys. B 31, L519.
Smyth, E.S., Parker, J.S., and Taylor, K.T. (1998). Comput. Phys. Commun. 114, 1.
Spence, D.E., Kean, EN., and Sibbett, W. (1991). Opt. Lett. 16, 42.
Taieb, R., V6niard, V., and Maquet, A. (2001). Phys. Rev. Lett. 87, 053002.
Tang, C.Y., Bryant, H.C., Harris, P.G., Mohagheghi, A.H., Reeder, R.A., Sharifian, H., Tootoonchi, H.,
Quick, C.R., Donahue, J.B., Cohen, S., and Smith, W.W. (1991). Phys. Rev. Lett. 66, 3124.
Toma, E.S., Antoine, Ph., de Bohan, A., and Muller, H.G. (1999). J. Phys. B 32, 5843.
van de Sand, G., and Rost, J.M. (2000). Phys. Rev. A 62, 053403.
van Linden van den Heuvell, H.B., and Muller, H.G. (1988). In "Multiphoton Processes" (S.J. Smith
and P.L. Knight, Eds.), Vol. 8 of Cambridge Studies in Modern Optics. Cambridge University
Press, Cambridge, p. 25.
V~izquez de Aldana, J.R., and Roso, L. (1999). Opt. Express 5, 144.
V~zquez de Aldana, J.R., Kylstra, N.J., Roso, L., Knight, EL., Patel, A., and Worthington, R.A.
(2001). Phys. Rev. A 64, 013411.
V~niard, V., Ta'ieb, R., and Maquet, A. (1995). Phys. Rev. Lett. 74, 4161.
V6niard, V., Taieb, R., and Maquet, A. (1996). Phys. Rev. A 54, 721.
Volkov, D.M. (1935). Z. Phys. 94, 250.
Walker, B., Sheehy, B., Kulander, K.C., and DiMauro, L.E (1996). Phys. Rev. Lett. 77, 5031.
Walser, M.W., Keitel, C.H., Scrinzi, A., and Brabec, T. (2000). Phys. Rev. Lett. 85, 5082.
Walsh, T.D.G., Ilkov, EA., and Chin, S.L. (1994). J. Phys. B 27, 3767.
Watson, J.B., Sanpera, A., Burnett, K., and Knight, P.L. (1997). Phys. Rev. A 55, 1224.
Weber, Th., Giessen, H., Weckenbrock, M., Urbasch, G., Staudte, A., Spielberger, L., Jagutzki, O.,
Mergel, V., Vollmer, M., and D6rner, R. (2000a). Nature (London) 405, 658.
Weber, Th., Weckenbrock, M., Staudte, A., Spielberger, L., Jagutzki, O., Mergel, V., Afaneh, E,
Urbasch, G., Vollmer, M., Giessen, H., and D6rner, R. (2000b). Phys. Rev. Lett. 84, 443.
Weingartshofer, A., Holmes, J.K., Caudle, G., Clarke, E.M., and Krfiger, H. (1977). Phys. Rev. Lett.
39, 269.
Weingartshofer, A., Holmes, J.K., Sabbagh, J., and Chin, S.L. (1983). J. Phys. B 16, 1805.
Wildenauer, J. (1987). J. Appl. Phys. 62, 41.
Yang, B., Schafer, K.J., Walker, B., Kulander, K.C., Agostini, P., and DiMauro, L.E (1993). Phys.
Rev. Lett. 71, 3770.
Yergeau, E, Chin, S.L., and Lavigne, P. (1987). J. Phys. B 20, 723.
Yudin, G.L., and Ivanov, M.Yu. (2001a). Phys. Rev. A 63, 033404.
Yudin, G.L., and Ivanov, M.Yu. (200 l b). Phys. Rev. A 64, 013409.
ADVANCES IN ATOMIC, M O L E C U L A R , A N D O P T I C A L PHYSICS, VOL. 48

DARK OPTICAL TRAPS FOR


COLD ATOMS
NIR FRIEDMAN, ARIEL K A P L A N and NIR D A V I D S O N
Weizmann Institute of Science, Department of Physics of Complex Systems, Rehovot, Israel

I. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
II. Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
A. Trapping Cold Atoms using Repulsive Dipole Forces . . . . . . . . . . . . . . . . . 101
B. Loading Atoms into Dark Optical Traps . . . . . . . . . . . . . . . . . . . . . . . . . . 103
C. Hollow Laser Beams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
III. Multiple-Laser-Beams Dark Optical Traps . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
A. Adding Beams Incoherently: Light-sheets and Hollow-beam Traps . . . . . . . . 107
B. Evanescent-wave Traps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
C. Using Interference: Dark Optical Lattices . . . . . . . . . . . . . . . . . . . . . . . . . 111
IV. Single-Beam Dark Optical Traps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
A. Generating Dark Volumes using Refractive Optical Elements: Axicon Traps . . 113
B. Creating Single-beam Dark Traps with Diffractive Optical Elements . . . . . . . 115
C. Scanning-beam Dark Optical Traps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
D. Comparing Different Traps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
V. Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
A. Manipulations in Phase Space: Cooling and Compression . . . . . . . . . . . . . . 127
B. Precision Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
C. Dynamics of the Trapped Atoms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
VI. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
VII. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

I. I n t r o d u c t i o n

Trapping of neutral atoms has become possible in the last 15 years, thanks to the
advances in laser cooling techniques (Adams and Riis, 1997). Several kinds of
traps were developed and investigated (Balykin et al., 2000), the most useful ones
being magnetic, magneto-optical and dipole force traps. Optical dipole traps use
the interaction between the electric field of the light and an electric dipole, which
is induced in the atom by this field. This force is weaker than the magneto-optical
and magnetic forces, and typical dipole trap depths are below 1 mK. These traps
offer the possibility to trap atoms in all internal states, as well as a possibility
to lower the dissipative component of the interaction between the atoms and the
trapping light by increasing the detuning of the laser from the atomic resonance
(which also unavoidably reduces the depth of the trapping potential). These traps
and their properties have been described in a recent review by Grimm et al.
(2000).

99 Copyright 9 2002 Elsevier Science (USA)


All rights reserved
ISBN 0-12-003848-X/ISSN 1049-250X/02 $35.00
1O0 N. Friedman et al. [I

Fla. 1. Fluorescence images of atoms in optical dipole traps. (a) Atoms trapped inside a
red-detuned trap; imaging is performed when the trapping beam is on. Atoms close to the focus of
the trapping beam are not observed, since they experience a large ac Stark shift, which shifts them
out of resonance with the probe beam. (b) Same as (a), but imaging is performed a short time after
the trapping beam is shut off. All atoms are detected. (c,d) Atoms trapped inside a blue-detuned trap,
with the same detuning and laser power as those of the red-detuned trap. All atoms are observed,
irrespective of whether the trapping beam is on (c) or off (d), indicating reduced perturbation in this
dark trap.

To further reduce the interaction between the atoms and the trapping laser,
dark optical traps were developed, in which the atoms are trapped by a repulsive
dipole interaction with the laser. The repulsive force is achieved by detuning the
trapping laser above the atomic resonance, such that there is a phase difference
between the field and the induced dipole. In these dark optical dipole traps,
cold atoms are trapped inside a dark region which is surrounded by a repulsive
dipole potential wall. To visualize the lower perturbations that are observed by
atoms inside a dark optical trap, consider Fig. 1, in which fluorescence images
of trapped atomic clouds are presented, for both attractive and repulsive optical
dipole traps. In the attractive (red-detuned) trap, atoms around the b e a m focus
are not observed, since they are shifted out of resonance with the fluorescence
beams, due to the large ac Stark shift induced by the trapping laser. W h e n the
measurement is repeated just after the trapping b e a m is shut off, these atoms
are easily observed. For the repulsive (blue-detuned) dark trap, which had the
same laser power, the same detuning and the same dimensions, there is no
II] DARK OPTICAL TRAPS FOR COLD ATOMS 101

difference between the measurements with and without the trapping beam. This
result indicates that perturbations due to the trapping beam are largely suppressed
for dark traps. Throughout this review we will present quantitative evidences for
this suppression.
Experimentally, dark traps (also called "blue-detuned" traps) are harder to
realize than attractive (or "red-detuned") dipole traps, where already a single
focused beam constitutes a trap. Several configurations have been demonstrated,
the first of which used multiple laser beams to form a confinement by light in
three dimensions, or used repulsive light structures to support the cold atoms
against gravity, which provided the trapping potential in the vertical direction.
Later, simpler trap designs were realized, which use only a single laser beam,
and traps with optimized properties were designed for various applications.
In this review, we describe dark optical traps, the techniques in which they
are experimentally realized, and their main applications in atomic physics. We
start with a short background section discussing the optical dipole force and
the process of loading cold atoms into the trap. We also describe hollow laser
beams, methods by which they can be formed, and their application as atom
guides, which confine atoms in two dimensions. In Sect. III we discuss dark
traps created using multiple laser beams, including gravito-optical traps, traps
that use evanescent light fields as the confining walls, and far-detuned dark
optical lattices. In Sect. IV, single-beam dark optical traps are considered, with
special emphasis on the optical techniques used in their construction. In both
sections, the focus is on trap designs that have been experimentally realized,
and the measured properties of the trapped atomic ensembles are described. In
Sect. V, the main applications of dark traps are discussed, including advanced
laser-cooling methods for dense atomic samples, the use of dark traps as a
favorable environment for precision spectroscopy, and the study of the dynamics
of trapped atoms.

II. Background
A. TRAPPING COLD ATOMS USING REPULSIVE DIPOLE FORCES

Light can exert a force upon atoms by momentum exchange due to photon
absorption and emission. The force that a laser field applies on an atom is usually
separated into two terms which correspond to a scattering force and a dipole
force. The scattering force results from absorption followed by spontaneous
emission of a photon in a random direction. The average momentum exchanged
is therefore one photon momentum, hkL, in the direction of the absorbed photon
(kf = 2:r//~, where/~ is the wavelength of the laser). In the case of the dipole
force, the photon is emitted in a stimulated way into a mode of the laser field. The
momentum transfer in this case is the vector difference between the momenta
of the absorbed and emitted photons.
102 N. Friedman et al. [II

For a two-level atom, the forces can be calculated by using the solutions of
the optical Bloch equations, while translational degrees of freedom are taken into
account (Cohen-Tannoudji et al., 1992). The resulting expression for the average
scattering force is then

Fscattering = hkL ~ (~2 + ( ),2/4 ) + (ff22/2) = hkL ),s, (1)


where ), is the natural linewidth of the atomic transition, 6 = 00c-000 is
the detuning of the laser from the atomic resonance, f2 is the Rabi frequency
that characterizes the atom-field interaction, and ),s is the spontaneous photon
scattering rate. The average dipole force can be written as
-h6( ~7(Q2(r)/2) )
Fdipole(r) = ~ 62 + (),2/4) + (ff22(r)/2) . (2)
This force depends on the sign of the detuning. If the laser is detuned below the
atomic resonance (red-detuned), the force will attract atoms into regions with
higher intensity. If the laser is detuned above resonance (blue-detuned) the force
will be repulsive, and atoms will be repelled by the light into lower-intensity
regions. The dipole force is conservative, and can be written as a gradient of the
potential
7 ( f22(r'/2 ) h6 ( l(r)/Is ) . (3,
Udipole(r) = - - In 1 + 62 + (),2/4) = m In 1 + 1 + (462/), 2)

In the last expression, the Rabi frequency is given in terms of measurable


quantities: ~2 = (),2/2)(i/ls), where I is the intensity of the laser, and Is is the
saturation intensity of the transition, given by Is = 2:r2h),c/3)t 3. For a large
detuning of the laser, 6 >> ),, f2, which is the usual case in optical dipole traps,
the potential has a simpler form:
h), 2 I(r) _ 3:rrc2 ~I(r).
Udipole(r)- 86 Is 2003 (4)

The last term is equal to Eq. (12) of Grimm et al. (2000), which was derived for
the classical oscillator model of the atom, using the rotating wave approximation.
The same result can also be derived in the dressed-state model, where the
combined Hamiltonian for the atom and the laser field is solved. (See, for
example, Cohen-Tannoudji et al., 1992). Equation 4 indicates that the dipole
potential is proportional to the laser intensity and inversely proportional to its
detuning. Comparing the expressions for the dipole force and scattering rate,
under the above approximations, yields the relationship
Udipole _ ~, (5)
h),s ),
which means that a trap with a reduced scattering rate can be made by increasing
the detuning while maintaining the ratio I/6.
II] DARK OPTICAL TRAPS FOR COLD ATOMS 103

In the case of multi-level atoms, Eq. (4) should be modified to include


the electric dipole interaction between the ground-state and all the excited
states, with their respective detunings and transition strengths. In practice, only
energy levels which are close to resonance with the laser frequency have to be
considered. In the case of far blue-detuned dipole traps, levels above the first
electronic excited state might have a considerable influence on the potential, if
the laser frequency is close to resonance with the transition. The laser detuning is
limited at the blue side by the ionization energy of the atom. These consideration
do not apply to red-detuned traps, where the contribution of transitions other
than the lowest one are much smaller. As an example, we consider a blue-
detuned trap for Rb atoms. The ionization energy of the ground state (5S1/2)
corresponds to a photon wavelength of ~300 nm. Note that ionization from the
excited state (5Pj) will occur at ~480 nm. If a trap is realized at this wavelength,
excited-state ionization will lead to trap loss when laser cooling is performed on
the trapped atoms. The lowest transitions from the ground state are the D lines
at 795 nm and 780nm. The next line (5S1/2 ~ 6P1/2) is at 421.5 nm, and its
transition strength is about 100 times lower than that of the D lines. Hence, this
line becomes relevant only for a laser which is detuned about 2 nm with respect
to it. For very large detunings, which are comparable to the optical frequency, a
multiplicative correction factor of the order of unity is needed in the potential
calculation, as a correction to the rotating wave approximation.

B. LOADING ATOMS INTO DARK OPTICAL TRAPS

The usual loading scheme of atoms into optical dipole traps starts with a
magneto-optical trap (MOT) (Raab et al., 1987), which traps atoms from a
vapor or an atomic beam and cools them to a typical temperature of 100 ~tK.
Since most dipole traps are relatively shallow and small as compared with the
MOT, it is advantageous to further cool the atoms and increase their density
in order to enhance the loading efficiency. The loading of red-detuned dipole
traps was thoroughly investigated both for trapping laser detunings of few nm
(or -3 • 105 V) (Kuppens et al., 2000), for larger detunings (~ 1 • 107 V) (Han et al.,
2001), and also for CO2 laser traps (O'Hara et al., 2001), where the detuning
is comparable to the atomic resonance frequency. In the CO2 trap, spatial and
phase-space densities much higher than that of a MOT are achieved, a fact that
led recently to the first demonstration of a Bose-Einstein condensate (BEC)
created in an all-optical way (Barrett et al., 2001).
There are some differences in the loading process between red- and blue-
detuned dipole traps. For red-detuned traps, the atoms are loaded into a region
of high trapping light intensity, which may interfere with the loading process.
The trap may reduce the optical cooling efficiency, since it results in a spatially
inhomogeneous Stark shift of the cooling line. The level shifts may influence also
104 N. Friedman et al. [II

the photon reabsorption since both the spontaneous emission and the absorption
spectra are altered. Combined with the trap's potential, which is usually steeper
than that of the MOT, it can result in a higher atomic density in the trap.
For blue-detuned traps, the loading seems to be simpler since the atoms
are loaded into a dark region, hence their interaction with the trapping light
is much smaller than their interaction with the MOT. This is supported by
several experimental observations. First, in many experiments the density and
temperature of the atoms loaded into the dark trap are very close to those in the
MOT. Second, the number of trapped atoms almost does not change if the trap is
present during the whole loading stage, or is turned on just at the end of the MOT
operation. These findings suggest that the loading is purely geometrical - those
atoms that are inside the dark "box" will stay there, those outside the box will not
be trapped, while the trap does not interfere with the operation of the MOT. The
different loading mechanism of bright and dark dipole traps is emphasized by an
experiment where atoms are first loaded from a MOT into a red-detuned dipole
trap, and are then transferred into an overlapping dark trap. In this manner, an
increase of about x2.5 in the number of atoms in the dark trap was observed,
as compared to direct loading of the dark trap from the MOT (Friedman et al.,
200 l a). This indicates that the red-detuned trap can enhance the atomic density,
while the dark trap leaves it almost unchanged. In this way, a dark trap may be
used as a probe for investigating loading into other traps, since it can sample
the atomic density and temperature at a given time. The advantage is that the
measurement can be performed at a later time, when the MOT atoms that were
not trapped have expanded and have fallen out of the detection region.

C. HOLLOWLASER BEAMS

As an introduction to the discussion of three-dimensional dark optical traps, it is


useful to first treat hollow laser beams, which can serve as two-dimensional traps,
or guides, for cold atoms. Here, we will describe how such hollow beams can
be produced, and the main results of atom guiding in such beams. This subject
has been discussed thoroughly by Balykin (1999).
A hollow beam has a light distribution with a minimum (ideally equal to
zero) along its axis. With a laser beam detuned above the atomic resonance,
such beams act as linear guides for cold atoms, where the atoms propagate
inside the dark "tube" created by the dipole potential of the beam. It is possible
to distinguish between two types of hollow beams. The first type contains
structurally stable beams (modes), which have a constant intensity cross section
that scales in size as they propagate. The second type of hollow beams is
not structurally stable but remains hollow along a relatively large propagation
distance, which can be made long enough for many applications.
II] DARK OPTICAL TRAPS FOR COLD ATOMS 105

Laguerre-Gaussian modes: An example for structurally stable beams are


Laguerre-Gaussian modes, LG/, which form a complete basis set of solutions
of the paraxial wave equation 1. For p = 0, l ;~ 0, the intensity distribution of
this beam has the form of an annulus, and is given by

jr[l[!w(z) 2 ~w(z)2 exp W(Z) 2 , (6)

where P is the total laser power, the waist size is w ( z ) = w0v/1 + (2/ZR) 2, and
zR = :rw2/)~ is the Rayleigh range. The phase of the beam changes linearly with
the azimuthal angle, 0, and there is a phase singularity on the beam axis. These
modes were extensively studied in the last decade, mainly due to their special
property of having orbital angular momentum lh per photon. For a review of this
subject see (Allen et al., 1999).
LG modes were generated experimentally from high-order Hermite-Gaussian
modes (which are easy to produce directly from laser resonators), by a
mode converter composed of two cylindrical lenses (Beijersbergen et al., 1993).
Such a cylindrical-lens mode converter was used to produce a LG g mode that
was part of a 3D dark optical trap (Kuga et al., 1997), as will be discussed in
Sect. III.A.
LG modes were generated also by using a computer generated hologram,
with a "fork" in the grating pattern. When illuminated with a plane-wave (or
a TEM00 laser beam), a "charge-one" phase singularity will be created in the
beam, centered around the fork defect. The resulting 1st diffraction order is a
good approximation of the required LG~ field distribution, although it is not a
pure mode (Heckenberg et al., 1992).
Such computer-generated holograms were used to create hollow beams which
served as guides for cold atoms (Kuppens et al., 1998; Schiffer et al., 1998).
In these experiments, metastable neon atoms were guided inside a focused LG~
hollow beam, a first demonstration of guiding cold atoms in free space 2. Atoms
were guided along a distance of 30cm and focused to a spot size o f - 6 . 5 ~tm.
Polarization-gradient cooling (PGC) in the transverse direction was applied in
order to further increase the phase-space density (see Sect. V.A.1). In later
experiments, a BEC was adiabatically transferred into a hollow LG 1 beam and
its propagation inside this optical guide was studied (Bongs et al., 2001).
A TEM00 Gaussian beam will be projected with a high efficiency onto a
LGt0 beam when it passes through a spiral phase element having a transmission

1 Another example for structurally stable hollow beams is constituted by high-order Bessel beams,
which are non-diffractive and were considered as atom guides (Arlt et al., 200 l b).
2 Cold atoms have been guided previously with light propagating inside hollow fibers (Balykin,
1999).
106 N. F r i e d m a n et al. [III

function of e ilO (Beijersbergen et al., 1994). Alternatively, intra-cavity spiral


phase elements can force a pure helical mode output directly from the laser (Oron
et al., 2000).
Output from a hollow fiber: A conceptually different method to generate a
hollow beam is to use the output beam from a hollow optical fiber (Yin et al.,
1997). Here, the laser is coupled to a high-order mode which propagates in the
fiber's cylindrical core. When the mode exits from the fiber's end it is collimated
with a strong lens to form the hollow beam, which has an intensity distribution
that is similar to the LG~ mode. Efficient guiding of falling cold atoms in such
a beam was demonstrated along a vertical distance of 11 cm (Xu et al., 1999,
2001).
Axicons: An axicon is a refractive optical element with a conical surface.
A TEM00 laser beam incident on its center will be refracted into a conical
beam. If a second axicon with the same base angle is then placed into the beam
it will collimate it, resulting in a light "tube." The intensity cross section of
this beam is similar to that of a high-order LG mode, but it has a constant
phase and it is not structurally stable. However, such tubes can remain hollow
over relatively large propagation distances, large enough for guiding cold atoms.
This was recently demonstrated (Song et al., 1999), by transporting a cloud of
-~108 atoms through an 18 cm long optical guide, with a diameter of 1 mm. In
another work, a continuous low-velocity atomic beam was guided inside a hollow
laser beam which was produced using axicons (Yan et al., 2000).
Axicons have some advantages in producing hollow beams for cold-atom
guiding and trapping, as demonstrated by Manek et al. (1998). First, the
characteristic width of the guide walls, w, can be very narrow, and the radius
of the dark center, r, can be large. In the above experiment, for example,
R = r/w ~ 11 was demonstrated, equivalent to a LGt0 mode with a very large l
that is much more difficult to realize efficiently 3. Second, the optical setup is
very simple: a Gaussian laser beam is used as the input, and simple commercially
available optical elements are utilized. When illuminated with a LG mode with
l > 0, the axicon-lens system converts it into a hollow beam with a much thinner
ring (hence much larger R), which still has the orbital angular momentum of the
original beam (Arlt et al., 2001a).

III. Multiple-Laser-Beams Dark Optical Traps

During the last decade, several types of dark optical traps were proposed and
demonstrated. The first dark traps were formed by incoherently adding several
laser beams that served as the trap's walls. These first traps were shallow and

3 For LG modes, R _~ v/1. Hence R - 11 will require l as large as 120.


III] DARK OPTICAL TRAPS FOR COLD ATOMS 107

had a relatively small volume, hence captured only a modest number of atoms.
However, these first experiments demonstrated the main advantages of dark traps,
namely low photon scattering rates and long coherence times of the trapped
atoms. Later improvements in trap design led to traps with larger volumes, in
which large number of atoms could be captured and manipulated. In this chapter,
traps based on multiple laser beams are described in a comparative way. The main
dark-trap designs are considered, and the performance of experimentally realized
traps is discussed. In order to compare traps made for different atomic species
on a common basis, trap properties are given in a normalized way: the detuning
is normalized by y (the linewidth of the relevant excited state) and the trap depth
is normalized by the recoil energy of the atom, Erec = hZk~/2m, where m is the
mass of the atom.

A. ADDING BEAMS INCOHERENTLY: LIGHT-SHEETS AND HOLLOW-BEAM TRAPS

In the first dark optical dipole trap for cold atoms, realized in 1995 in Stanford
(Davidson et al., 1995), sodium atoms where trapped using light from an Ar-ion
laser. This trap consisted of two elliptical light sheets (generated by focusing
a Gaussian beam with a cylindrical lens) intersecting at 90 ~ and forming a
"V"-shaped cross section. Confinement was provided by gravity in the vertical
direction and by the beams' divergence in the longitudinal direction. The two
beams had powers of 4 and 6 W, were linearly polarized, and had different wave-
lengths (488 and 514.5 nm) so that they did not interfere in the overlap region and
formed a smooth potential. The large detuning of the trap beams, ~ 107 y, resulted
in a relatively low potential of ~10E~ec, and hence a low number of trapped
atoms (~3000), but also an extremely low photon scattering rate, calculated as
~ 10-3 s-1 , and a long lifetime of 5 s, limited by the background vacuum. A spec-
troscopic measurement of the hyperfine splitting of Na, which was performed on
the trapped atoms (see Sect. V.B), yielded a coherence time of 7 s. This coher-
ence time was 300 times longer than that achieved in a red-detuned trap having
the same potential height and a larger detuning, emphasizing the advantage of
a dark trap for precision measurements. The coherence time was limited by
inhomogeneous broadening, since different atoms acquire different Stark shifts,
depending on the velocity distribution and the dynamics of the trapped atoms.
In a later work (Lee et al., 1996), this trap was improved by intersecting two
such "V" traps at a right angle, resulting in an inverted-pyramid trap (see Fig. 2).
The polarization of the beams in the second pair was rotated by 90 ~ with respect
to the first in order to prevent interference in the trapping region, which would
lead to trap loss. A much larger number of atoms (4.5 • 105) were confined in this
trap, and its shape produced coupling between the motion in all three dimensions.
This allowed cooling in three dimensions by applying one-dimensional Raman
cooling (see further discussion in Sect. V.A.2). The issue of atom dynamics
108 N. F r i e d m a n et al. [III

FIG. 2. Schematic illustration of the inverted pyramid dark optical trap, which is composed of
four blue-detuned light sheets. (From Lee et al., 1996, Phys. Rev. Lett. 76, 2658, Fig. 1).

FIG. 3. Schematic illustration of the dark hollow beam trap, which is based on a hollow
Laguerre-Gaussian beam and two plugging beams. (From Kuga et al., 1997, Phys. Rev. Lett.
78, 4713, Fig. 2).

and its possible implication for cooling and spectroscopy will be discussed later,
Sect. V.C.
A different trap configuration was demonstrated by Kuga et al. (1997): it
consisted of a hollow laser beam (LG 3) and two additional "plug" beams that
confined the atoms in the propagation direction of the hollow beam (see Fig. 3).
The hollow beam had a power of 600 mW and a radius of 600 ~m. A detuning
of ~1047 resulted in a potential of ~100Erec, higher than the typical energy of
PGC-cooled Rb atoms. The deep potential, combined with a very large volume
of 2 • 10-3 cm 3, enabled the loading of a much larger number of atoms (1 • 108),
with a very good loading efficiency of about one-third from the MOT. However,
the relatively small detuning resulted in a high photon scattering rate of~100 s-1 ,
which limited the lifetime of the trap to 150 ms due to heating of the atoms above
the potential barrier. To reduce this heating effect, pulsed PGC was applied,
resulting in a longer lifetime of 1.5 s (Torii et al., 1998).
There have been several proposals for traps using a vertical hollow beam
combined with a horizontal plug beam that supports the atoms against gravity,
so forming a gravito-optical trap (Morsch and Meacher, 1998; Yin et al., 1998).
The hollow beam can be produced in either of the ways discussed in the previous
section, and can be focused to create a conical shape, forming a funnel that
III] DARK OPTICAL TRAPS FOR COLD ATOMS 109

potentially increases the loading efficiency from a MOT into the dipole trap.
These proposed schemes suggest the use of inelastic reflection of atoms from
the trapping light (reflection Sisyphus cooling, see Sect. V.A.3) to reduce the
relatively high kinetic energy of atoms that fall from the MOT into the trap,
and also to balance the heating due to spontaneous photon scattering. A similar
configuration was used by Webster et al. (2000): 103 Cs atoms were trapped
above the focus of a vertical LG 1 beam from an Ar-ion laser. No plug beam was
used here, but the loss of atoms through the very small hole at the bottom of the
trap is negligible due to the relatively high temperature of the atoms. The trap
was loaded from a magnetic trap, in which evaporative cooling was performed
to lower the kinetic energy of the atoms below the dipole trap potential. Since
no reflection cooling occurs at this large detuning, gravity was balanced by a
magnetic field gradient, such that atoms were falling very slowly into the dark
trap.

B. EVANESCENT-WAVE TRAPS

In the early 1990s, normal-incidence reflection of cold atoms from an evanescent


wave was demonstrated by Kasevich et al. (1990). The evanescent wave is
produced by total internal reflection of a linearly polarized blue-detuned laser
beam at the surface of a dielectric (glass) prism, which forms a steep potential
wall above the surface, of the form U ( z ) = Uo exp(-2z/A). Here,

A = (7)
2~ V/n 2 sin 2 0 - 1

is the characteristic interaction length scale, where n is the refractive index of


the prism, and 0 is the angle of incidence. U0 is the dipole potential on the
prism surface, which can be calculated as discussed in Sect. II.A. Atoms with a
kinetic energy lower than the potential height are reflected from the evanescent
light sheet without hitting the surface. The photon scattering per bounce is given
by np = 7 m o • where v• is the vertical component of the atom's velocity
before it enters the interaction region. A is typically much smaller than the
size of a focused Gaussian beam, resulting in a much steeper potential, which
is advantageous for supporting atoms against gravity while minimizing their
interaction with the supporting sheet of light. The height of the potential barrier
is reduced due to the attractive van der Waals force between the atoms and the
surface of the prism, which becomes relevant at distances of ~~./2:r.
In another experiment, a gravitational atom cavity was realized, in which
atoms bounced several times on an evanescent wave which was formed on a
curved glass surface (Aminoff et al., 1993). This "atomic trampoline" can be
regarded as a dark optical trap, with a potential depth of 5000Erec in the vertical
direction (created by the evanescent wave), and ~30Erec in the radial direction
110 N. F r i e d m a n et al. [III

repumping l l
beam ,~::z~,,..~]

atoms in MOT V ~ ; [
1 1 evan.
atoms in GOST " ~ T wave

P D

v F=4
9.2 GHz
,~ F=3
dielectric vacuum
EW laser W ~
beam hollow beam

(a) (b)
FIG. 4. (a) Schematic illustration of the experimental setup of the gravito-optical surface trap.
The trap is formed by an evanescent light wave that supports atoms against gravity, and a hollow
beam that provides confinement in the horizontal direction. (From Ovchinnikov et al., 1997, Phys.
Reu. Lett. 79, 2225, Fig. 1). (b) The reflection cooling cycle, with the relevant energy levels for the
Cs D 2 line. An atom moves towards the mirror in the lower hyperfine level (F -- 3). Close to the
classical turning point, it may undergo a spontaneous Raman transition to the upper level ( F - 4), by
scattering a photon from the evanescent wave. It is then reflected from the mirror along the lower
potential observed by the upper level. The cycle is closed by spontaneous scattering of a photon
from a repumping beam, which takes the atom back to the lower level. (From Engler et al., 1998,
Appl. Phys. B 67, 709, Fig. 2).

(induced by the surface curvature). The trap's lifetime was ~lOOms, limited
mainly by scattering o f stray light from the surface, which either heats the atoms
or optically pumps them into a state for which the potential is weaker, hence the
effective area o f the mirror is smaller.
In order to form a trap with a longer lifetime and a better confinement, a
vertical hollow beam was added to confine the atoms in the horizontal direction,
as shown in Fig. 4a (Ovchinnikov et al., 1997). In this trap, the evanescent wave
was produced using a 6 0 - m W laser b e a m which was detuned by only ~ 2 0 0 7 from
the lower hyperfine level of the atomic ground state, creating a potential barrier
of 104Erec for that level (taking into account the van der Waals interaction
with the surface). For the upper hyperfine level, the detuning is larger by the
hyperfine splitting ( - 1 7 5 0 y in Cs), hence the potential is lower. The different
potential for these two levels is used in the reflection Sisyphus cooling to be
discussed in Sect. V.A.3. The 1 2 0 - m W hollow b e a m had a radius o f 360[am
and a characteristic ring width o f w0 - 18 [am. It was detuned by 0.3 n m from
resonance (6 = 2 x 104) ,) and produced a potential barrier o f ~500Erec. The high
potential barrier and large volume m a d e it possible to trap a large n u m b e r o f
Cs atoms: 2 x 105 in the first experiment, which was later increased up to 2 x 107
by improving the loading scheme o f the M O T ( H a m m e s et al., 2000). From a
III] DARK OPTICAL TRAPS FOR COLD ATOMS 111

measurement of the heating rate of the trapped atoms, the upper bound for the
total photon scattering rate in this trap was calculated to be 50 s -1.
The large number of atoms at relatively low temperature and high density
make this trap a good candidate for evaporative cooling towards quantum
degeneracy in an all-optical way in a dark trap. Evaporative cooling of Cs atoms
was indeed demonstrated by lowering the potential height, as will be discussed in
Sect. V.A.4. Due to the very different confinement in the vertical and horizontal
directions, this trap has some promising applications in the investigation of
quantum degeneracy of a 2D gas (Petrov et al., 2000). For this purpose, the
atoms can be confined more tightly in the vertical direction by adding a red-
detuned evanescent wave that will create a narrower potential well (Engler et al.,
1998), or by adding a vertical standing wave above the surface. In this scheme,
an atom falls on the evanescent wave and at the turning point is transferred by
a spontaneous Raman transition to a different internal state, which is uncoupled
from the evanescent wave but is coupled to the standing-wave potential that has
a minimum at this point. Such a scheme was demonstrated experimentally with
metastable Ar atoms, which were preferentially loaded into a single potential well
of a red- or blue-detuned standing wave with an increase of 100 in spatial
density (Gauck et al., 1998). This scheme can be extended to alkali atoms by
a proper choice of laser polarizations, as discussed by Spreeuw et al. (2000).
Finally, the evanescent-wave surface trap can also be used to explore atom-
surface interaction, which is of theoretical as well as practical importance for
various atom-optics devices such as waveguides and traps close to surfaces.

C. USING INTERFERENCE: DARK OPTICAL LATTICES

Cold atoms can be trapped in the dark nodes of interference patterns of blue-
detuned light. Only three-dimensional (3D) blue-detuned optical lattices will
provide 3D trapping, while even a 1D red-detuned optical lattice is capable
of confining atoms in three dimensions. In a first demonstration of a 3D dark
optical lattice, lithium atoms were trapped in a non-dissipative way (Anderson
et al., 1996). The lattice was formed by the interference of four intersecting laser
beams, detuned either to the red (10 nm) or to the blue (1 nm, or - 1 x 105 y) of
the atomic resonance. Typical lifetime of the trapped atoms was on the order
of 50ms, limited by heating due to photon scattering with a scattering rate of
~100s -1. This scattering rate is relatively high as compared to far-detuned dark
optical traps since the lattice potential height was comparable to the kinetic
energy of the atoms, so no effective "darkness factor" was obtained 4.

4 Note that a higher potential is needed for trapping the lighter Li atoms, since the typical momenta
of laser-cooled alkalis are comparable, resulting in a higher kinetic energy for the lighter atoms.
112 N. Friedman et al. [Ill

A much lower scattering rate was demonstrated for metastable Ar atoms


trapped in the lowest energy band of a 3D dark optical lattice (Muller-Seydlitz
et al., 1997). Atoms in higher bands suffer higher loss rates due to heating caused
by photon scattering since their broader wavefunctions have a larger overlap with
the trapping light. Quantitatively, the scattering rate for atoms in the (l, m, n) band
can be estimated within a harmonic approximation of the potential around a
node, as !

ysO, 9 (t + m + n + 3 ) .

With the experimental parameters, gs was as low as 6 S-1 for the (0, 0, 0) ground
band. Another band-dependent loss mechanism is tunneling of a trapped atom
to a neighboring lattice site, and eventually out of the lattice by diffusion. The
detuning of the lattice beams was 6 = 2.5x 105,/, and the trap depth was
54Erec, not much higher than the initial kinetic energy of the atoms. Due to
the low potential, only the lowest bands with l, m, n ~< 3 are bound and initially
populated, with a total number of 104 atoms. The RMS momentum of the atoms
remaining in the trap decreased as a function of trapping time, indicating the
band-dependent loss described above. The population in each band was resolved
by ramping down the potential slowly such that higher lying bands were released
first, and atoms in the lower state were released in a later, resolvable time.
The decay time for the lowest band was found to be 0.31 s, as compared to
0.13 s for the next-higher band, leading to a preparation of an atomic sample in
the ground band after about 0.45 s of storage in the lattice. Since the preparation
was done by selection and not by cooling to the ground state, only a very small
number of about 50 atoms were trapped in the ground state.
Muller-Seydlitz et al. (1997) constructed the lattice by three orthogonal
standing waves with mutually orthogonal linear polarizations. Since the atoms
are trapped in a J = 0 state, the dipole potential is independent of the local
polarization of the light and is proportional to the sum of intensities. For atoms
in a J ;~ 0 state, the relative phases of the three orthogonal standing waves have to
be stabilized. Alternatively, standing waves with a frequency difference between
them can be used such that on average the polarization is linear (DePue et al.,
1999; Chin et al., 2001), or an inherently stable configuration with only four
beams can be used, as did Anderson et al. (1996).
A related proposal describes a way of producing a 1D blue-detuned optical
lattice using two counter-propagating Gaussian beams with different waists such
that confinement is achieved also in the radial direction (Zemanek and Foot,
1998). Here, as opposed to the 3D lattice, the intensity in the nodes is zero
only at z = 0, where the waists of the two waves are located, and the intensities
are equal. Out of this plane the two waves have different divergence angles due
to the different waist sizes, resulting in a gradual increase in the intensity at
the well bottom and a decrease in the darkness factor. Since the confinement in
IV] DARK OPTICAL TRAPS FOR COLD ATOMS 113

the radial direction results from non-perfect destructive interference, the radial
potential depth is much smaller than that in the longitudinal direction.
Dark optical lattices are promising candidates for performing precision
measurements, such as of the electron's permanent electric dipole moment, as
discussed recently by Chin et al. (2001), and to be described in Sect. V.B. Finally,
far blue-detuned optical lattices have attracted much attention as a possible
system for the realization of quantum information processing (Brennen et aL,
1999). Here, the main advantages are the low interaction of an atom with its
environment leading to long coherence time, and the possibility to control the
interactions between individual atoms to induce entanglement.

IV. Single-Beam Dark Optical Traps

In this section we describe dark optical traps created with a single laser beam.
These traps are simpler to align than traps using several beams, hence it is
easier to optimize the trap properties. The simplicity of these traps also permits
easy manipulation and dynamical control of the trapping potential, its size and
its shape. As opposed to the 2D case, where any desired light distribution can
be generated using diffractive or refractive optical elements, there is no simple
procedure to design an arbitrary 3D light distribution. Actually, there are not
enough degrees of freedom in the design of an optical element to achieve a
full 3D arbitrary light distribution 5. Nevertheless, with the combined use of
refractive and holographic optical elements, it is possible to extend the methods
described in Sect. II.C in order to produce light distributions which are suitable
for trapping atoms in the dark using a single laser beam. Such light distributions,
which comprise of a dark volume completely surrounded by light, were realized
using either combinations of axicons and spherical lenses, diffractive optical
elements, or rapidly scanning laser beams. The following subsections describe
the various trap designs, and analyze the properties of the resulting traps. In the
last subsection, traps of different classes are compared quantitatively.

A. GENERATING DARK VOLUMES USING REFRACTIVE OPTICAL ELEMENTS:


AXICON TRAPS

A combination of axicons and spherical lenses is capable of transforming a


TEM00 Gaussian beam into a light distribution that is suitable for 3D dark optical
trapping. The first example was actually a gravity-assisted trap, where a cone of

5 This issue has been discussed, for example, by Piestun and Shamir (1994), Spektor et al. (1996)
and Shabtayet al. (2000), who designed hologramsthat produce some specific 3D light distributions,
and discussed the limits of these design procedures.
114 N. F r i e d m a n et al. [IV

Fic. 5. Schematic illustration of the conical beam optical trap. The conical hollow beam, directed
upwards, is generated with two equal axicons (an "axicon telescope") and a spherical lens. (From
Ovchinnikov et al., 1998, Europhys. Lett. 43, 510, Fig. l a).

light facing upward was realized (Ovchinnikov et al., 1998). The optical setup,
illustrated in Fig. 5, consisted of two axicons with identical angles (an "axicon
telescope") followed by a spherical lens. As opposed to the funnel trap described
previously (Sect. III.A), the apex of the cone in this case is not hollow and
no plug beam is needed. In the experiment, a 250-mW laser beam was used
to form the conical trap with an opening angle of 150mrad. The detuning of
the laser was 3 GHz (6 = 560),) with respect to the lower hyperfine level of
Cs (F = 3), producing a potential barrier o f ~ 3 x 10SErec in the focal plane (apex
of the cone), decreasing linearly with height. The trap was loaded from a MOT
located 5 mm above the apex. A Cs atom that falls 5 mm acquires a kinetic energy
o f - 8x 103Erec and hence is easily confined by the trap potential. Six molasses
beams were left on during the experiment, to realize a combined polarization gra-
dient cooling and reflection Sisyphus cooling. The measured loading efficiency
from the MOT into the trap was very high, ~ 80%, resulting in -8 x 105 atoms in
the trap, at a temperature of-10Trec. The trapped atomic cloud had an elongated
shape with a length of about 1 mm in the vertical direction, and a radial size of
100 ~tm. Two drawbacks of this trap are the high photon scattering rate from the
conical beam, ~3 x 103 s-1 , and the poorly defined potential shape near the apex
[probably caused by interference (Webster et al., 2000)]. This can be inferred
from the elongated shape of the resulting atomic cloud, which does not compress
towards the apex as would be expected from the equipartition of energy.
Recently, a related scheme producing a 3D single-beam dark trap has been
demonstrated by two groups (Cacciapuoti et al., 2001; Kulin et al., 2001). This
scheme uses a single axicon placed between two spherical lenses (see Fig. 6).
IV] D A R K OPTICAL TRAPS F O R C O L D ATOMS 115

FIG. 6. Schematic illustration of the experimental setup for a single-beam dark optical trap, based
on an axicon and two spherical lenses. The axicon and the first lens provide a virtual image of the
trap (with radius r) indicated by the dashed lines, which is imaged by the second lens to form the
dark trap (with radius R). (From Cacciapuoti et al., 2001, Eur. Phys. J. D 14, 373, Fig. 1).

After passing through the first lens, a divergent Gaussian beam hits the axicon.
The second lens focuses the resulting divergent conical beam into a ring o f
light with a dark center. Since the divergence angle o f the beam is larger than
the opening angle o f the cone of light, the trap is closed also along the beam
propagation direction. Traps in a large range o f dimensions (40 ~tm < r < 740 ~tm)
were experimentally realized, demonstrating the flexibility of this design 6. In
the second experiment (Kulin et al., 2001), this optical scheme was used to
produce a large trap, with r = 740 ~tm, L = 150 mm, and V = 8 x 10 -2 cm 3. The
optical darkness factor (defined as the ratio between the light intensity inside the
trap and the intensity in the focal plane ring) was about 1 : 1000, while along
the axis there was a residual peak, probably caused by light going through the
apex of the axicon. Since the opening angle of the conical beam is not zero, the
two cusps on the optical axis are not of the same height, the one closer to the
lens being about 10 times higher. The lowest points in the potential barrier are
located off-axis, close to the far end o f the trap. This trap was used to capture
Xe atoms in a metastable state. The high trap volume resulted in a good loading
efficiency o f 50% from the MOT which contained a few million atoms. The
lifetime o f the trap was only about 2 0 m s , limited by the gravitational energy
of the heavy Xe atoms in the large trap, which is larger than the height of the
potential barrier at the far end o f the trap.

B. CREATING SINGLE-BEAM DARK TRAPS WITH DIFFRACTIVE OPTICAL ELEMENTS

A different approach to produce dark volumes surrounded by light is to create


a destructive interference in some region in space between two coherent light
fields which have different propagation characteristics. Outside this dark region
the intensity will rise in all directions due to the different propagation constants.

6 Actually, the trap dimensions can be changed by moving only the axicon, which does not change
the location of the focal plane.
1 16 N. Friedman et al. [IV

FIG. 7. (a) Schematic illustration of the experimental setup for the single-beam dark optical trap
based on a Jr-phase plate element. A destructive interference between the inner part of the beam
(shifted by Jr radians) and the outer ring is formed around the focus of the third lens. (b) Contour
map of the calculated light intensity distribution around the focus, for the parameters described in
the text. The dark minimum is labeled m, and three bright maxima are labeled M. The trap height
is given by the two saddle-points at r _~ 0.017 mm, Z _~ 2504-1 mm. (From Ozeri et al., 1999, Phys.
Rev. A 59, R1750, Fig. 1).

The :r-phase plate trap: The simplest trap o f this kind is realized by placing
a circular phase-plate element into a Gaussian laser beam (Chaloupka et al.,
1997; Ozeri et al., 1999), as shown in Fig. 7a. The phase plate imposes a phase
difference o f exactly Jr radians between the central and outer parts o f the beam,
which have equal intensities. W h e n the b e a m is focused by a lens, destructive
interference between the two parts ensures a dark region around the focus, which
is surrounded by light from all directions, as required. Let b denote the radius
of the inner (phase-shifted) circle, a the outer (clipped) radius o f the Gaussian
beam, and w0 its waist. Equal intensities in the two regions are achieved when

b = w 0 v / - l n {~1 [1 + exp(-a2/w2)] }. (9)

The resulting intensity distribution calculated numerically using the Fresnel


diffraction integral, for the parameters used by Ozeri et al. (1999), are presented
in Fig. 7b. The radius o f the trap is similar to the waist of the focused Gaussian
beam in the absence o f the phase element, w0, focus = ) f / Z w o , w h e r e f is the focal
length o f the focusing lens. The length o f the trap is L ~ 2ZR = 2Yt'w2,focus//],,
which gives a volume o f 7

V 2:rK4wZ, focusZe = 2;rZK4w4,focus/) ~ = ; ( wo)4 ~3. (10)

7 Note that this is the maximal volume of the trap. The trapped atoms usually fill a smaller volume,
depending on the ratio between their kinetic energy and the height of the trapping potential.
IV] DARK OPTICAL TRAPS FOR COLD ATOMS 117

Here, K is a constant on the order of 1-3, which depends on the clipping radius a.
The volume of many single-beam dark traps can be expressed in the form

V : C . (f/wo~ ~3, (11)

where the constant C depends on the exact realization of the trap. Note that the
volume of a focused Gaussian red-detuned trap is given by V -- 2zR. Jrw2, focus,
which can be written in the form of Eq. (11), with C = 2/jr 2. With the
experimental parameters w0 = 6mm, a = 5 m m (resulting in K ~ 2.3),
f = 250 mm, and/l = 799 nm, the volume of the dark trap is V ~ 9• 10-6 cm 3.
The potential height of the trap is determined by the minimal light intensity
on the trap surface, which is ~ 10% of the peak intensity of the unaltered
Gaussian beam.
In the experiment described by Ozeri et al. (1999), 105 85Rb atoms were
confined in a trap that was realized using this optical setup. The Jr phase plate
was produced by evaporating a thin dielectric layer of an exact thickness on
a glass plate (Davidson et al., 1999). An optical darkness ratio of 750 was
measured, which depends on the amount of light scattered into the dark region,
the degree to which condition 9 is fulfilled, the deviations of the incoming beam
from a Gaussian, and the deviations of the phase shift from Jr. These effects
were studied in detail by Chaloupka and Meyerhofer (2000). In the experiment
of Ozeri et al. (1999), a 1-W laser beam was used, which could be detuned
in the range 4 • 104)' < 6 < 1x 106) ' above resonance. In this range the
lifetime of the trap was 300 ms, limited by collisions with background atoms. For
smaller detunings, the lifetime decreased linearly with 6, which is consistent with
heating-induced lifetime, since the heating rate is proportional to 6 -2 whereas
the trap depth is proportional to 6 -! .
A similar optical configuration was used to optically trap high-energy electrons
with an intense single laser beam, using the ponderomotive force (Chaloupka
and Meyerhofer, 1999). A laser intensity distribution with a local minimum
is essential for trapping electrons with light, since the electrons are always
repelled by the laser field towards the intensity minimum. In this experiment,
the Jr phase shift was realized with a segmented wave plate which can endure a
very high laser intensity.
Spontaneous Raman scattering rate: The amount of interaction between
the atoms and the trapping light was accurately determined by measuring the
spontaneous Raman scattering rate that results from the trapping light (Cline
et al., 1994). This is a very useful experimental technique which permits the
measurement of even very low scattering rates, hence we will present it in some
detail. For this measurement, the trapped atoms were first prepared in the lower
hyperfine level of the ground state, F - 2. The number of atoms in F = 3 after a
variable time t, N3(t), was measured by detecting the fluorescence after a short
pulse of a resonant laser beam. For normalization, the total number of atoms in
118 N. Friedman et al. [IV

0.51 . . . . . . . . ,~ . 9

!~~;i ' . . . . . . oo, 000


Ume (msec)
FIG. 8. A measurement of spontaneous Raman scattering rate for atoms in the z-phase plate dark
optical trap. At t = 0 all atoms are pumped to the lower hyperfine level. The fraction of atoms in
the upper hyperfine level (F = 3, for 85Rb) is measured as a function of time in the trap. The fit
to Eq. (12) gives a Raman scattering time of 164 ms, from which the total scattering rate can be
calculated. (From Ozeri et al., 1999, Phys. Rev. A 59, R1750, Fig. 4).

the two sublevels of the ground state, N3(t)+ Nz(t), can be measured by turning
on also the repumping beam (which is resonant with F = 2) during the detection
pulse. Actually, N3(t) and N2(t) are measured in the same experimental run. Since
the detection beam excites a closed transition, it accelerates the atoms in F = 3,
and these are rapidly shifted out of resonance. Then the repumping laser is turned
on and the number of atoms in F = 2 (which were not accelerated) is measured.
This normalized detection scheme is insensitive to shot-to-shot fluctuations in
atom number as well as fluctuations in frequency and intensity of the detection
laser (Khaykovich et al., 2000). Typical experimental data for the F = 3 fraction
as a function of time in the trap, for a detuning of 0.5 nm (4• 104),), are shown
in Fig. 8. The data are well fitted by the function

N3(t)+ N2(t) = c 1 - e x p --r-~R ' (12)

where c is the equilibrium fraction of atoms in F = 3 at long times, and


rSR is the measured spontaneous Raman scattering time, which is 164 ms in
this case. The total scattering rate ?'~. is given by 1/(qrSR), where q is the
branching ratio for a spontaneous Raman transition which ends in the upper
hyperfine level. The branching ratio depends on the detuning of the trapping
laser and its polarization.
For laser trap detunings larger than the fine-structure splitting of the excited
state (e.g. 15 nm in Rb), destructive interference exists between the transition
amplitudes for Raman scattering, summed over the intermediate (excited) states
(Cline et al., 1994). In this case, most spontaneous scattering events leave the
internal state of the atom unchanged. Hence, the probability for a spontaneous
Raman transition is strongly suppressed, and the spin-relaxation times are largely
increased.
The "bottle beam" trap: Another way to create a destructive interference
between two different light fields emerging from a single laser beam is
IV] DARK OPTICAL TRAPS FOR COLD ATOMS 119

FIG. 9. Contour map of the calculated light intensity distribution for the CR-BPE dark trap.
O indicates the trap center, A the transverse maximum, B the axial maximum, and C the lowest
barrier height. (From Ozeri et al., 2000, J. Opt. Soc. Am. B 17, 1113, Fig. 2).

described by Arlt and Padgett (2000). A computer-generated hologram produces


a superposition of the LG ~ and LG ~ modes from an incoming Gaussian beam,
such that the two modes have equal on-axis intensities in the focal plane
and an opposite phase. The relative phase between the modes changes with
propagation due to the different Gouy phase around the focal plane, hence the
zero-intensity region is surrounded by regions of higher intensity in all directions.
The dimensions and shape of this trap as well as the height of the potential barrier
are very similar to those of the Jr phase-plate trap.
"Diffractive axicon" traps: The main disadvantages of the traps described
above are their low volumes and high aspect ratios, which leads to low loading
efficiency of atoms from a MOT. It is possible to optimize the trap parameters
by combining diffractive optical elements and axicons to form a more symmetric
trap with a large volume while still enjoying the advantages of a single-laser-
beam setup. A main optical element for this optimization is the concentric-rings
binary phase element (CR-BPE), which is an extension of the Jr phase plate
(Ozeri et al., 2000). The CR-BPE is composed of concentric phase rings with a
Jr phase difference between sequential rings, thus creating a radial grating with
a uniform spacing. This radial grating functions similar to a refractive axicon,
however, it produces two cones of light corresponding to the :kl diffraction
orders, as opposed to only one cone in the axicon. To form a dark trap, the
CR-BPE is illuminated with a Gaussian beam of waist w0, which is then
focused by a lens. At the focal plane, the interference between the two cones
of light results in a ring with a dark center, and due to the different propagation
characteristics, a dark volume surrounded by light is formed. A numerical
solution of the Fresnel diffraction integral for this case is shown in Fig. 9. This
light-intensity map reveals that out of the focal plane the interference pattern
of the two diffraction orders is not smooth but rather contains radial fringes,
forming channels with reduced dipole potential height. In particular, the lowest
120 N. F r i e d m a n et al. [IV

point in the potential barrier (point C in Fig. 9) is - 1 0 times lower than the
transverse potential height (point A) and 60 times lower than the maximal
on-axis intensity (point B). Hence, this trap is relatively shallow, and a very
large detuning cannot be used. The radius of this trap can be estimated as
r = (Jr/2)Mwo, focus, and its length as L - JrMzR, where M = wold, and d is the
width of a phase ring in the radial grating. Since M and w0, focus can be controlled
nearly independently, it is possible to design a trap with a smaller aspect ratio
and a larger volume than the previous ones. Specifically, the volume of this trap
is given by

V ~ ~1jrr2L = ]-~
1 ~ 4 M 3 w 2 0, focusZR = 1-~jrm3
()4
f ~3. (13)
W0
This trap was realized with the following parameters" w0 = 400 [xm, d = 50 ~tm
a n d f = 16mm, giving a volume o f - l . 6 • -4 cm 3. The CR-BPE was formed
as a binary surface-relief phase element, as described by Ozeri et al. (2000).
With a laser power of 120 mW and a detuning of 6 = 1 • 1057 above resonance,
-3 • 106 Rb atoms were loaded into the trap, with a loading efficiency of 5% from
the MOT. The total photon scattering rate was determined from a measurement
of the spontaneous Raman scattering rate to be 10 s-~ . For larger detunings, part
of the atoms were trapped only in two dimensions, and escaped from the trap
through the lowest point in the potential barrier.
An improved trap configuration was recently demonstrated by adding an
axicon telescope before the CR-BPE of the above setup (Kaplan et al., 2002a).
This configuration maximizes the trap depth for a given laser power and
trap dimensions, and greatly reduces the light-induced perturbations to the
trapped atoms. These properties are achieved by surrounding a large dark volume
with a light envelope with (a) an almost minimal surface area for a given
volume, (b) the minimal wall thickness that is allowed by diffraction, and
(c) an almost constant wall height over the entire envelope. The stiffness of the
trap walls, combined with the large detuning allowed by the efficient intensity
distribution, yield a very low calculated spontaneous photon scattering rate for
the trapped atoms. The optical configuration for the creation of this trap is
illustrated in Fig. 10a. The thin ring of light created by the axicon telescope is
diffracted into two cones by the CR-BPE. When these two diffraction orders are
imaged, a light distribution is generated that consists of two equal hollow cones
attached at their bases and completely surrounding a dark region. The height of
the confining potential of this double-conical trap can be approximated by

U(z) L 1
= 9 (14)
U ( z - O) 4 [ ~L _ z] V/1 + (Z/ZR) 2 '

where L is the total trap length. The first term accounts for the linear decrease in
the trap radius away from the focal plane (in both directions) and the second term
IV] DARK OPTICAL TRAPS FOR COLD ATOMS 121

FIG. 10. The double-conical single-beam dark optical trap. (a) Schematic illustration of the optical
setup. A CR-BPE is placed after an "axicon telescope." The +1 and -1 diffraction orders are imaged
to generate the trap (in black). (b) Contour map of the calculated light intensity distribution for the
trap. Note the very thin walls, which provide a very good darkness factor. (c) Measured intensity
cross sections at different planes along the trap's axis [parameters of the optical setup are different
than in (b)]. The two diffraction orders are observed, and the inner one provides the trap walls.
At z = 0 the two orders overlap exactly.

accounts for the diffraction o f the f o c u s e d b e a m . U s i n g a s p h e r i c a l - l e n s t e l e s c o p e


before the a x i c o n s (see Fig. 10a), the b e a m waist on the p h a s e e l e m e n t is c h o s e n
to b a l a n c e these two effects, and to m a i n t a i n a n e a r l y c o n s t a n t intensity a l o n g z:

1)3/2 p (15)
I = ~ rw0, focus'

w h i c h is o b t a i n e d for L ~ 7ZR 8 (r is the trap radius in the focal plane). T h e

8 Two exceptions are z = 0, where the two diffraction orders overlap, yielding a double potential
height, and z "~ L / 2 , where the singularity in Eq. (14) yields an extremely high potential.
122 N. F r i e d m a n et al. [IV

FIG. 11. (a) Measured cross sections of the time-averaged light intensity in three positions along
the scanning-beam optica! trap, together with a schematic drawing of the trap formation. (b) Lifetime
of atoms trapped in a scanning-beam trap, as a function of the scanning frequency of the trapping
beam, for two trap radii. (From Friedman et al., 2000a, Phys. Reo. A 61, 031403(R), Figs. 1,2).

dimensions of the trap can be changed while maintaining this optimal ratio
between L and zR. The trap volume is given by

w2
1 ~ r 2 L ~ 7yg2r2 o,focus
Z ~ 5 5 /l " (16)

This optical configuration was demonstrated experimentally, and a trap with


w0,focus = 22.6gm, r = 1.47mm and L -- 13mm was generated. The
aspect ratio of this trap was 1"4.4, which is very small compared to the other
configurations discussed above. Figure 10b presents a calculated cross section of
the trap's potential in the x - z plane, and Fig. 10c shows the measured intensity
cross sections along the trap axis. The measured intensity is nearly constant over
the entire envelope, evidence for the optimal use of the laser power.

C. SCANNING-BEAM DARK OPTICAL TRAPS

Trapping particles with a time-dependent potential was successfully applied to


trap ions in oscillating electric fields (Major and Dehmelt, 1968), cold atoms
in magnetic traps (Petrich et al., 1995), and larger particles in optical tweezers
(Sasaki et al., 1992). It is possible to extend this approach to construct a
dynamical optical trap for cold atoms with a rapidly scanning laser beam. If the
scan frequency is high enough, the optical dipole potential can be approximated
as a time-averaged quasi-static potential. For a blue-detuned laser beam, and
a radius of rotation, r, larger than the waist of the focused beam, w0,focus, a
dark volume suitable for 3D trapping is obtained, as shown in Fig. 1 la. Let us
calculate the properties of the trap for a circular scan of the beam in the focal
IV] DARK OPTICAL TRAPS FOR COLD ATOMS 123

plane. The time-averaged intensity in the focal plane ring (which corresponds to
the radial potential barrier) is given by

2)1/2 i0 2P (17)
/r = ~ 4---R ~ 5~rwo, focus'

where P is the laser power, I0 is the peak intensity of the static focused beam,
and R = r/wo, rocus is the resolution of the scan. The two lowest points in the
potential barrier of the trap are located on the optical axis, and the intensity at
these points (which corresponds to the trap depth) is

Io P
Iz -- 2eR2 Jrer 2 . (18)

Hence, the trap depth depends only on the rotation radius and not on the
resolution. The length of the trap is L = 2zRv/2R 2 - 1, which gives a trapping
volume of 4
V ~ 89 = 2zR2v/ZR2 - 1 -- ~3 (19)
3Jr

The scanning is realized either with two perpendicular acousto-optic scan-


ners (AOS) (Friedman et al., 2000a) or with two mechanical scanners (Rudy
et al., 2001), by electronically modulating the deflection angle. The possibility
of electronically controlling the shape of the trap is an important advantage of the
scanning-beam trap. It is possible to create traps of different shapes by changing
the modulating signal, as was demonstrated by creating "optical billiards" for
cold atoms (Milner et al., 2001; Friedman et al., 200 lb), as will be discussed in
Sect. V.C. Moreover, it is possible to dynamically change the shape and size of
the trap during an experiment, as demonstrated by compression of a cold atomic
cloud to very high densities (see Sect. V.A).
A scanning-beam optical trap was demonstrated by Friedman et al. (2000a)
using a linearly polarized laser with a power of 200mW, scanned at a
rate of 100 kHz. Stable trapping of atoms was obtained for a detuning in the range
1 x 104)' < 6 < l x 1 0 6 ) ' above resonance, and a trap radius of r = 24-105 ~tm
(R = 1.5-6.5). For r = 100 ~tm and 6 = 1.2 x 1057' the potential height was
"~60Erec as compared to the atomic kinetic energy which was -25Erec. More than
106 8SRb atoms were loaded into this trap from a MOT, and 3.5x 106 atoms
were loaded into a deeper trap with 6 = 4 x 104. The lifetime of the trap
was 600ms, limited by collisions with background gas. Several measurements
were performed in order to prove that the potential can be regarded as a time-
averaged potential. The radial and axial oscillation frequencies of atoms in the
trap were measured and found to be in very good agreement with the predicted
frequencies for a time-averaged potential. In addition, the trap stability was
studied by measuring the trap lifetime as a function of the scanning frequency
124 N. F r i e d m a n et al. [IV

of the laser beam (see Fig. 11 b). A large trap (r = 67 ~tm) was found to be stable
for scan frequencies larger than -20 kHz, while for a smaller trap (r = 32 ~tm)
stable trapping was achieved only for higher frequencies (above -60 kHz). As the
scan frequency is decreased, stability is reduced in a gradual way. One reason for
this gradual behavior is the velocity distribution of the trapped atoms, since slow
atoms are easier to trap even with a slowly scanning beam. However, numerical
simulations for a monoenergetic ensemble reveal that the stability of the trap
depends not only on the velocity of the atoms, but also on the exact position in
phase space, and on the shape of the trap. The stability of the scanning-beam trap
is an interesting subject, which will probably be further studied. A spin-relaxation
time of rSR = 380 ms was measured for the trapped atoms by investigating the
spontaneous Raman scattering. Using the branching ratio for the experimental
parameters, this corresponds to a total photon scattering rate of 7 s-1 .
In another realization of a scanning-beam optical trap (Rudy et al., 2001), a
500-mW laser beam was scanned with mechanical scanners at lower frequencies
(2-11kHz). This trap had larger dimensions (w0 = 200~tm, r = 1.5mm),
which made it more stable at low scanning frequencies, but demanded lower
detuning ( o n l y - 1 x 103),) to achieve a sufficient height of the trapping
potential. To increase the height of the potential barrier, the scanning beam
was re-imaged after passing through the vacuum chamber and crossed the
original in an orthogonal direction, such that a nearly spherical trapping volume
was achieved, with a potential height of 100Erec. Up to 8x 105 Na atoms
were loaded into the trap with a lifetime o f - 5 0 m s , limited by heating
due to spontaneous photon scattering from the trapping beam, calculated
to be 500 s -1.
In the realization of scanning-beam traps, two contradicting requirements
exist: a fast scan and a high resolution. Mechanical scanners are usually limited
to scan frequencies below 10 kHz. Acousto-optic scanners are capable of much
faster scans, up to a few 100 kHz, but with nonlinear scans the resolution is
decreased at high frequencies due to the chirp of the acoustic grating over the
laser beam. As demonstrated recently by Friedman et al. (2000b), this limitation
can be corrected through the use of two counter-propagating acoustic waves,
such that the chirp is canceled.

D. COMPARING DIFFERENT TRAPS

The appropriate design considerations, and the advantages or disadvantages


of the various trap schemes, depend on the specific experimental details (the
application, the atomic species, etc.). Hence, no absolute preference of one
configuration over the others, nor a general recipe for an optimization procedure
can be presented. Nevertheless, the purpose of this section is to shed some light
IV] D A R K OPTICAL TRAPS F O R C O L D ATOMS 125

on the common design considerations and trade-offs usually met when choosing
a certain scheme and optimizing its performance.
As a specific example, we assume that the requirement is to trap most o f the
atoms from a magneto-optical trap (MOT) into the dipole trap, and minimize
the spontaneous photon scattering rate o f the trapped atoms. We assume a laser
with a fixed power P -- 1 W and a sample o f alkali atoms, laser-cooled in a
MOT to a temperature o f ks T ,.~ 25Erec, and forming a nearly spherical cloud
with radius -0.5 mm. We will compare the performance o f three dark traps o f the
different classes described previously: a trap based on a Laguerre-Gaussian (LG)
beam (Kuga et al., 1997), a trap based on a scanning beam (Friedman et aL,
2000a), and one based on a diffractive axicon element (the double-conical trap,
Kaplan et al., 2002a).
Adopting a criterion o f > 90% geometrical loading efficiency from the MOT,
we choose a radius r - 0.5 mm for all traps. The beam waist is chosen as
w0 = 50 ~tm (and therefore R = 10) for the scanning-beam trap, and w0 = 10 ~tm
(and therefore R = 50) for the double-conical trap. The length o f the latter is an
independent parameter, chosen as L = 3 m m to optimize the power distribution
as explained in Sect. IV.B. For the LG trap, a LG 3 mode is assumed, with
w0 - 0.5 mm. For comparison, we look also at a red-detuned trap, formed by
two focused Gaussian beams, intersecting at a right angle 9 (Adams et al., 1995).
We neglect the enhanced loading efficiency of red-detuned traps (Kuppens et al.,
2000) (see Sect. II.B), and assume for the crossed red-detuned trap w0 = 0.6 mm,
which corresponds to > 90% overlap with the MOT.
Following Grimm et al. (2000), we introduce the parameter to, defined as
the ratio of the ensemble-averaged potential and kinetic energies of the trapped
atoms, tr = <Epot>/ (Ek >, and refer to it as the "darkness factor" o f the dark trap.
Assuming a trapped atomic gas in thermal equilibrium, and neglecting gravity,
the ensemble-averaged potential energy is given by

fdr(U(r)-Uo) e x p [ - U ( r ) - U0]
<Epot)= kbT ,
fdrexp[-U(r)-kSTUo] (20)
with U(r) the three-dimensional dipole potential function, U0 the potential at
the bottom o f the trap, and the integration taken over the entire trap volume.
3
The ensemble-averaged kinetic energy is (Ek> = ~ksT. To generalize the
darkness factor also for red-detuned traps, we use a modified definition o f

9 We choose a crossed trap, and not a simpler focused Gaussian beam trap, since with a single
focused beam a trap radius of 0.6 mm will result in an extremely large axial size of > 1 m.
126 N. F r i e d m a n et al. [IV

Table I
Required detuning, calculated atomic darkness factor and mean spontaneous photon scattering rate
for Rb atoms confined in the traps analyzed in the text

Trap Detuning [nm] to' ()'s) [s-1 ]

Red-detuned trap -0.7 4.9 166.9


"Laguerre-Gaussian" trap 0.23 0.6 87.6
Scanning-beam trap 0.19 0.2 21.2
Double-conical "optimized" trap 4.69 0.02 0.09

to' = ] U o / ( E k ) + tr For r e d - d e t u n e d traps U0 < 0, while U0 = 0 for m o s t


b l u e - d e t u n e d traps 10.
The e n s e m b l e - a v e r a g e d s p o n t a n e o u s p h o t o n scattering rate is then given in
terms o f the darkness factor by

( Ys ) = - ~Y -~
3 kb T . K.t. (21)

The detuning in the c o m p a r i s o n is c h o s e n such that the trap depth is 3 times larger
than the m e a n kinetic energy o f the atoms. Since a fixed laser p o w e r is
assumed, less efficient traps w o u l d require s m a l l e r detunings to provide the s a m e
trap depth. In Table I, the required d e t u n i n g is p r e s e n t e d for Rb a t o m s in each
o f the different traps, with the p a r a m e t e r s discussed above. Also s h o w n are the
numerically calculated to', and the m e a n s p o n t a n e o u s p h o t o n scattering rate, (Ys),
which d e t e r m i n e s the heating and d e c o h e r e n c e rates o f the trapped atoms.
As expected, all b l u e - d e t u n e d traps have a better d a r k n e s s factor than the red-
detuned trap. Their scattering rates are smaller as well, even for traps requiring
a smaller detuning. The d o u b l e - c o n i c a l trap has a significantly better d a r k n e s s
factor (to' = 0.02) than all other schemes. The a d v a n t a g e o f the d o u b l e - c o n i c a l
trap is even larger w h e n the scattering rate is considered, since the i m p r o v e d
darkness factor is c o m b i n e d with the efficient distribution o f optical p o w e r that
enables an increased detuning ll. We p e r f o r m e d a similar calculation also for a

10 For example, for an harmonic trap given by U = Uo + ax 2 + by 2 + cz 2, the ensemble-averaged


potential and kinetic energies are equal, and therefore the darkness factor is always tr ~ = 1 for a
blue-detuned trap, independent of laser power, detuning or atomic temperature, but tr t cx Uo/kB T for
a red-detuned trap with U0 >> kBT. This is the case for the crossed red-detuned trap considered
here, for which tr = 1.1, while tr r = 4.9.
11 A scattering rate comparable to that achieved with the double-conical trap could be achieved with
the LG trap, with a similar trap resolution R-50. But since for a LG mode R-x/7 an extremely high-
order mode would be required, which is experimentally impractical. On the other hand, as a result
of the large contribution of light in the out-of-focus regions, the darkness ratio of the scanning-beam
trap depends only very moderately on R.
V] DARK OPTICAL TRAPS FOR COLD ATOMS 127

3D blue-detuned optical lattice, composed of three orthogonal standing waves.


Here, a power of gi W was taken for each of the standing waves, and other
parameters were w0 = 737 ~tm, 6 = 0.25 nm, such that the lattice potential at
the MOT radius (0.5 mm) was 3 times the kinetic energy of the atoms. We then
assumed further cooling of the atoms to the ground state of the lattice, and
calculated the scattering rate, averaged over the lattice (inside the MOT volume).
The result is (Ys) -- 29.1 s-1, similar to that in the rotating-beam trap and
considerably higher than in the double-conical trap. In the next section, other
considerations for applying dark traps for precision measurements are discussed,
such as collision-induced level shifts, which are dramatically reduced inside a
dark lattice (Chin et al., 2001).
So far, we have neglected the effects of gravity. Assuming gravity is in the
radial direction, the relatively large dimensions of our traps result in a large
maximal gravitational potential of ~500Erec (larger than the assumed trap depth),
which can seriously affect the loading efficiency, and increase the temperature of
the atoms. We neglect this excess energy, assuming that it is efficiently dissipated
by some cooling mechanism (e.g. PGC) during the loading process. In this case,
the only effect of gravity is to modify the distribution of the atoms inside the
trap. Including gravity in our calculations results in an increase of 10-60% for
the scattering rate for Rb atoms in the above dark traps. Similar calculations
were performed also for Na atoms. Since Na atoms have a much smaller mass
than Rb atoms, and a similar resonance frequency, their recoil energy is much
higher, hence deeper traps are needed. For a fixed power of the trap laser
smaller detunings are required, resulting in higher scattering rates. An additional
consequence of the smaller mass is that the scattering rate of trapped atoms is
less affected by gravity. In our calculations, the inclusion of gravity yields an
increase of no more than 5% in the scattering rate for Na atoms (and even less
for the lighter Li atoms).

V. Applications
Dark optical traps offer a relatively interaction-free environment for the confined
cold atoms. This property has opened up a way for several applications, including
precision spectroscopic measurements and the preparation and investigation of
atomic samples at high spatial and phase-space densities. Recently, the dynamics
of atoms inside a dark optical trap has been studied as a versatile experimental
realization of the well-known billiard system. All these applications are discussed
in this section, with a focus on their experimental demonstrations.

A. MANIPULATIONS IN PHASE SPACE: COOLING AND COMPRESSION

In this subsection we describe the main cooling mechanisms that have been
realized inside dark optical traps. These include polarization-gradient cooling,
128 N. Friedman et al. [V

Raman cooling and evaporative cooling, and also the reflection cooling mecha-
nism, which is unique for blue-detuned traps. Cooling inside a trap is favorable
since density increases as the atoms get colder, hence the gain in phase-space
density is larger. Moreover, coupling between the different directions which
results from trap anisotropy or collisions between the atoms makes it possible to
achieve 3D cooling while performing laser cooling along only one dimension.
As opposed to magnetic traps, in an optical dipole trap it is possible to trap and
cool atoms independently of their magnetic sublevel, so that a BEC composed
of different m-states can be achieved (Barrett et al., 2001), it is possible to
investigate a BEC in an arbitrary magnetic field (Inouye et al., 1998) and to
study spinor condensates composed of atoms at different m-states (Stenger et al.,
1998). Moreover, in some atomic species such as Cs, the lowest ground state
cannot be trapped in a magnetic trap, and those states that can be trapped suffer
from a very high inelastic loss rate which limits the achievable phase-space
density below the BEC transition (S6ding et al., 1998). As a result, experimental
effort is directed towards cooling Cs atoms in optical traps.
Experimental and theoretical effort has been devoted to find the limitations of
the various cooling schemes, and many heating and loss mechanisms have been
identified and investigated. [Relevant examples for the present discussion include
Bali et al. (1994), Castin et al. (1998), Winoto et al. (1999), Wolf et al. (2000),
Kerman et al. (2000), Kuppens et al. (2000)]. A detailed description of these
mechanisms is beyond the scope of this review. Here, we will briefly describe
the limiting mechanisms which are relevant for cooling inside dark optical traps.
At the end of this subsection we will discuss compression of the trapped atomic
cloud, as it can lead to better starting conditions for evaporative cooling, and is
also interesting for measurements of cold collisions.

A.1. Polarization-gradient cooling

Polarization-gradient cooling is among the most widely used sub-Doppler laser


cooling techniques for non-trapped atoms. It was applied also to cool atoms
inside bright and dark optical dipole traps. Inside a bright trap, the position-
dependent light shift induced by the trapping light might be comparable to the
shift induced by the cooling laser, hence cooling efficiency is reduced. In dark
traps, this effect is suppressed due to the lower interaction with the trapping
light.
A first investigation of the effect of a dark trap on the PGC process was
performed with metastable neon atoms guided inside a focused, blue-detuned
hollow beam (Kuppens et al., 1998). A cold atomic beam was adiabatically
focused by the hollow guiding laser beam, and hence heated from -5Erec to
more than 1000Erec. When 2D PGC was applied just before the focus of the
beam, sub-Doppler cooling of the atoms was observed. A residual interaction of
the atoms with the guiding beam alters the optical pumping of the cooling light.
V] DARK OPTICAL TRAPS FOR COLD ATOMS 129

As a result, a higher intensity is required in the cooling beams to overcome the


light shift of the guiding beam, and cooling efficiency is reduced with increasing
depth of the guiding potential.
PGC was also applied to 3D dark traps. In some experiments it was used
to suppress heating due to photon scattering of the trapping light which was
relatively close to resonance (Torii et al., 1998; Ovchinnikov et al., 1998). Torii
et al. (1998) applied pulsed PGC to reduce loss of atoms due to inelastic light-
assisted collisions. Here, as well, the achievable temperature was slightly higher
than in free-space PGC with the same parameters and atomic densities. In the
work of Ovchinnikov et al. (1998) the PGC beams were continuously on, but the
cooling had an effective low duty cycle because no repumping laser was used,
hence atoms spent most of the time in a dark state and the trapping laser supplied
a slow repumping. The main limit to optical cooling inside traps is density-
dependent heating and loss, which are discussed at the end of this subsection.

A.2. R a m a n cooling

Raman cooling was first used to cool untrapped alkali atoms below the photon-
recoil limit in one dimension (Kasevich and Chu, 1992) and then in two
and three dimensions (Davidson et al., 1994). This cooling method is based
on accumulating cold atoms in a velocity dark state. In alkalis, the scheme
is realized by transferring atoms between the two hyperfine levels of the
ground state, using pulses of counter-propagating Raman laser beams. The
parameters of these pulses are chosen such that atoms in different velocity
classes (except a velocity class around zero) are transferred to the upper level.
A repumping beam transfers the atoms back to the lower level, via spontaneous
emission. This process is continued until most of the atoms accumulate in the
lower level, in the dark velocity class near v = 0.
This scheme was successfully applied to sodium atoms trapped in the inverted-
pyramid dark trap described in Sect. III.A (Lee et al., 1996). Since the trap
mixes the motion in the three spatial dimensions, cooling in 3D was achieved
by applying the Raman beams in only one dimension. In the trap, 4.5 • 105 atoms
were cooled to a temperature of 0.4Trec at a final density of 4• 1011 cm -3. The
phase-space density was increased by a factor of 320 as compared to the MOT, to
a final value of n/laB ,~ 6• l 0 -3 [with AdB the thermal de-Broglie wavelength],
which is the highest yet achieved inside a dark optical trap. Since the atoms
are trapped, their velocity might change during their interaction with the Raman
pulse, resulting in motional sidebands which reduce the velocity selectivity of the
Raman pulses, and limit the achievable temperature as compared with untrapped
atoms. In a later experiment, a modified scheme was used to simultaneously
Raman-cool the atoms and optically pump them into one magnetic sublevel of
the lower hyperfine state (Lee and Chu, 1998). The modified cooling scheme
resulted in a slightly higher temperature and also some loss of trapped atoms,
130 N. Friedman et al. [V

such that the final phase-space density was comparable to that of the previous
experiment.
For very tight traps, the motional sidebands are resolved, enabling the
realization of Raman sideband cooling schemes. Raman sideband cooling was
applied so far only to atoms trapped inside red-detuned optical lattices (Hamann
et al., 1998; Perrin et al., 1998; Vuleti6 et al., 1998; Kerman et al., 2000), but it
can be applied also for dark optical lattices to cool atoms to the lowest vibrational
level. Cooling inside optical lattices [using either Raman sideband cooling or
PGC (Winoto et aL, 1999)] can also be used to produce a cold and dense source
of atoms, to improve loading into optical dipole traps (Han et al., 2001).

A.3. Reflection Sisyphus cooling


Reflection Sisyphus cooling is a unique cooling scheme for atoms in blue-
detuned dipole traps. It exploits the fact that a different potential height is
observed by different internal energy levels of the atom. By scattering a photon
during a reflection, an atom may lose some of its kinetic energy to the light field.
This cooling mechanism was first proposed (Ovchinnikov et al., 1995; S6ding
et al., 1995) and realized (Desbiolles et al., 1996) for traps consisting of
evanescent waves. The cooling cycle begins with a bouncing atom that enters
the potential in an internal level which feels a high potential value. Close to the
classical turning point the atom has a large chance to perform a spontaneous
Raman transition into another internal state for which the potential is lower.
As a result, the atom loses part of its potential energy to the light field, and is
reflected with a lower velocity. The cooling cycle is closed by optically pumping
the atom back to the initial state when it is out of the potential, without changing
its potential energy.
In the case of alkali atoms, the two hyperfine levels of the ground electronic
state are used as illustrated in Fig. 4b, for a Cs atom. The average energy
loss per reflection, for motion perpendicular to the surface, is given by
A E • 1 7 7 ~ - ~2 ~6hf
, ~ qnp, where 6hf is the frequency difference between the
two hyperfine levels and q is the branching ratio as defined in Sect. IV.B.
As an example, consider reflection cooling performed inside the gravito-optical
surface trap (see Sect. III.B). With the experimental parameters of Ovchinnikov
et al. (1997), a Cs atom initially loses on average 6% of its kinetic energy
per bounce. The mean time between reflections is rr = 2v• (with g the
gravitational acceleration), thus the cooling rate in the case of an evanescent
wave is independent of v• and given by

AE• q 6hf mgA


fi - rrE• - 3 6 h(b + Ohf ) Y" (22)

For Cs, one has q = 0.25, dihf = 2Jr • 9.2 GHz, and )' = 2Jr • 5.3 MHz. With
the experimental values of 6 = 1 GHz and A = 300nm, the cooling rate is
V] DARK OPTICAL TRAPS FOR COLD ATOMS 131

/3 ~ 5 • 10-77 ~ 2st x 2.5 Hz. This cooling rate is limited by the relatively long
time between inelastic collisions, and is much lower than the cooling rates of
Doppler cooling and polarization-gradient cooling. The equilibrium temperature
can be estimated by equating the average cooling and heating per bounce. With
the above experimental parameters, a temperature of T ~ 10Trec ,~ 2~tK
is expected, which is similar to that achieved in PGC. This temperature was
experimentally obtained for a sample with a small number of atoms. Higher
temperatures were obtained in dense samples, due to multiple photon scattering,
caused mainly by stray light from the hollow beam. Although the Sisyphus
cooling acts only along the vertical direction, the horizontal direction is also
cooled to the same temperature. This coupling is probably due to evanescent-
wave diffusive reflection from the non-perfect surface of the prism.
Reflection cooling was also demonstrated with traveling waves, in the Axicon
conical trap (Ovchinnikov et al., 1998). Here, cooling was observed only for
larger detunings of the conical beam, 6 = 30GHz. The larger detuning is
required in order that the condition np < 1 be fulfilled [np is the number
of photons scattered per bounce, see Sect. III.B], but it lowers the cooling
rate (22). In traveling-wave dark optical traps with a larger spring constant, such
as the scanning-beam trap, the lower cooling rate might be partially compensated
by the high rate of reflection from the walls, which is of the order of the
oscillation frequency in the trap potential.

A.4. Evaporative cooling


Despite the progress of optical cooling schemes, the only way by which BEC
has been achieved until now is evaporative cooling. For a detailed review
on evaporative cooling, see Ketterle and VanDruten (1996). Briefly, cooling
is initiated by inducing a loss of the more energetic atoms from a dense
sample, which is followed by rethermalization via elastic collisions, to a lower
equilibrium temperature. By gradually decreasing the cutting temperature, the
phase-space density of the trapped sample increases while the number of trapped
atoms and their temperature decrease. The common way to produce a BEC uses
evaporative cooling of atoms inside a magnetic trap. Evaporative cooling was
demonstrated for atoms inside a red-detuned optical trap (Adams et al., 1995),
and very recently 87Rb atoms were cooled below the condensation limit in a
quasi-electrostatic crossed trap (two crossed CO2 laser beams), creating for the
first time a BEC in an all-optical way (Barrett et al., 2001).
Evaporative cooling was studied also inside dark optical traps, with Cs atoms
inside the gravito-optical surface trap (see Sect. III.B) (Hammes et al. 2000,
2001). The starting conditions provided by efficient loading and reflection
cooling w e r e 10 7 atoms, at a temperature of 52Trec (10~tK), and a density of
6 • 1011 c m -3. The sample was unpolarized in the seven sublevels of the F = 3
lower hyperfine level, hence the corresponding phase-space density was -10 -5.
132 N. Friedman et al. [V

Evaporation was forced by lowering the dipole potential of the evanescent


wave by gradually increasing its detuning. This simultaneously reduced photon
scattering and loss due to light-assisted collisions. In the experiment, the laser
detuning was changed from ~l.5x103y up to ~5x104y within 4.5 seconds.
At the end of the evaporation ramp, the temperature of the atoms dropped to
1.5Trec (300 nK), the number of atoms in the trap was ~3 z 104, and their density
was ~6x 10 l~ cm -3. The phase-space density was increased by a factor of 30
to 3 x 10 -4. Since the potential in the vertical direction is linear with z due to
gravity, the density increases with decreasing temperature as T -1, only slightly
less than the T -3/2 dependence in a harmonic potential. The phase-space density
is thus proportional to N T -5/2, as was confirmed experimentally. The regime
of runaway evaporation was not reached in the experiment. The evaporation was
limited by heating caused by residual on-resonant light from the evanescent-wave
laser, which was greatly reduced by passing the laser beam through a heated cell
of Cs vapor. Another limiting mechanism was the interaction between the atoms
and the imperfect surface of the prism at low potentials.

A.5. Compression

When the volume of a trap is reduced, the atomic density and temperature
are increased. The increase in density leads to better starting conditions for
evaporative cooling since the cooling rate is limited by the elastic collision
rate or, in the hydrodynamic limit, also by the trap oscillation frequency (Han
et al., 2001; Ketterle and VanDruten, 1996), which are both increased with
compression. Compression of magnetic traps is a common procedure in many
BEC production schemes. In optical traps, compression might have a larger
effect since even higher densities and oscillation frequencies can be achieved.
For example, in the first realization of an all-optical BEC formation (Barrett
et al., 2001), the evaporative cooling time was only 2 s, as compared to > 10 s
usually needed in magnetic traps.
Compression was demonstrated in the scanning-beam dark optical trap
described in Sect. IV.C. Since the shape of the trap is controlled electronically,
it can be changed easily at a desired rate. As an example, when the trap
radius was decreased from 100 ~tm to 27 ~m in 150ms, a 350 times adiabatic
increase in spatial density was observed. These results were improved by adding
a PGC pulse during the compression, which resulted in a cloud of 106 atoms
at a density of 2x1013 cm -3, an axial temperature of 75Trec and a radial
temperature of 110Tr~. This represented a x 130 increase in spatial density and
a x 16 increase in phase-space density over the initial conditions, to a value of
1.2x 10-4.
An interesting effect in this context is that even an adiabatic change in the
potential shape can lead to a change in the maximal phase-space density if the
potential functional dependence is modified, and the elastic collision rate is high
V] DARK OPTICAL TRAPS FOR COLD ATOMS 133

enough to allow for thermalization. For a potential which is approximated as


U ( x , y , z ) = axX nx + ayy ny + azz nz, an adiabatic change in the exponents will lead
to a change in phase-space density which can be expressed as exp( ]"final- ~'initial),
where ), = nx 1 + ny 1 + nz 1. This effect was first observed for magnetically
trapped hydrogen by Pinkse et al. (1997). In the scanning-beam trap experiment,
it was demonstrated by decreasing the final radius to 24 ~m, for which the trap
is nearly harmonic in all directions (as compared to nx = ny = 28, nz -- 6 in
the initial 100 ILtmtrap). In this experiment, 3.5 • 106 atoms were compressed to
a final density of 5 • 1013 cm -3, and the phase-space density was increased by
a factor of ~4, in agreement with the above calculation. Loss of atoms during
the compression was negligible, ensuring that no evaporation process caused the
observed gain in phase-space density.
Compression can be accomplished also by mechanical movement of a lens,
as was recently demonstrated for a red-detuned crossed dipole trap (Han et al.,
2001), and can be applied to dark optical traps based on axicons or diffractive
optical elements, as well.

A.6. H e a t i n g a n d loss

In this section we briefly discuss the main heating and loss mechanisms for dark
optical traps. Heating can result either from interaction with the trapping light
itself, or due to photon re-absorption when laser cooling is applied to dense
atomic samples. Loss may result from interaction of atoms with the environment
and with the trapping light, or from inelastic collisions between atoms. Finally,
light-assisted collisions, where two colliding atoms interact with the trap light,
also contribute to trap loss.
Photon scattering: A major source for heating in optical traps is photon
scattering from the trapping light. To minimize the scattering rate, traps with
large detunings are favorable, and dark traps have an advantage over bright traps
(see Sect. IV.D). From a practical viewpoint, it is important to reduce the amount
of stray light scattered into the dark region. Very small amounts of residual
on-resonance light in the trap laser beam might also lead to heating, and should
be filtered in cases where the latter should be kept minimal.
Other heating sources: Intensity or pointing instabilities of the trapping laser
(Savard et al., 1997), and quantum diffractive collisions with background gas in
the vacuum chamber (Bali et al., 1999) cause heating in both bright and dark
optical traps. However, in some experiments the measured heating rate exceeds
the estimated rate based on these processes (Han et al., 2001), indicating that
some other heating mechanisms may exist.
Density-dependent heating: In optically dense samples, reabsorption of
spontaneously scattered photons during a laser-cooling process causes heating,
and limits the attainable equilibrium temperature. A quantitative estimation of
this effect was obtained experimentally for PGC performed on Cs atoms in
134 N. Friedman et al. [V

free space (Boiron et al., 1996). In this experiment, a density-dependent increase


in temperature, at a rate of 0.6~tK/10 l~ cm -3, was measured. Density-dependent
heating limits also sub-recoil laser-cooling techniques, such as Raman cooling
(Perrin et al., 1999), since the dark state is only dark with respect to the cooling
laser photons, but not with respect to spontaneously emitted photons, which have
a different frequency spectrum.
The effect of photon re-absorption can be reduced in elongated traps with
a large surface to volume ratio, since the photons have a higher probability to
escape from the trap before being rescattered. This was demonstrated for a red-
detuned dipole trap (Boiron et al., 1998), where atoms at a density of 1012 cm -3
where cooled to a temperature of 2 ~tK, about 30 times colder than a free-space
sample with the same density. Cooling becomes limited again when the optical
density of the cloud in the transverse (smallest) direction is larger than unity, as
was demonstrated by applying PGC in a compressed dark trap (Friedman et al.,
2000a). A cooling pulse was able to cool the atoms to their free-space molasses
temperature up to a density which resulted in a radial optical density of the
order of one, while for higher densities optical cooling failed. Density-dependent
heating was observed also for reflection cooling in the evanescent-wave trap
(Hammes et al., 2000), for densities that corresponded to an optical density larger
than unity even in the small direction.
Recently, suppression of density-dependent heating was observed and investi-
gated in tightly confining red-detuned optical lattices, using either PGC (Winoto
et al., 1999) or Raman sideband cooling (Han et al., 2000; Vuleti6 et al., 1998;
Kerman et al., 2000). The proposed mechanisms for this suppression should also
be valid for a tightly confining dark trap.
Loss: The loss of atoms from a trap can be approximated as

dN_dt aN(t)-fifvn2(r't)d3r-cJvn3(r't)d3r' (23)

where n is the atomic density, and integration is over the trap volume (Grimm
et al., 2000). The first term corresponds to loss processes which do not depend on
the atomic density, mainly collisions of the trapped atoms with hot background
atoms in the vacuum chamber. In ultra-high vacuum chambers, the background-
limited lifetime can be of the order of tens of seconds. In shallow traps, heating
of the atoms is also translated into trap loss.
In tightly confining traps, the high density results in two-body loss mech-
anisms, which are described by the second term in Eq. (23). The third term
corresponds to three-body loss which plays a role only at very high densities.
The density-dependent loss can be observed and quantified by measuring the
decay of the number of trapped atoms as a function of time. As an example, the
decay curves for atoms from the compressed scanning-beam dark trap (Friedman
et al., 2000a) are presented in Fig. 12a, for atoms in either the lower or the upper
V] DARK OPTICAL TRAPS FOR COLD ATOMS 135

06 Ur
! F,=2
F,=3
I
10~ I s+p

~ 104 I
I (!)
I
U=(R) I
' S+S
I~ oro o'.5 I'.o ~15 i.o
Time [sl R

(a) (b)
FIG. 12. (a) Measured decay of the number of trapped atoms from a scanning-beamtrap, for atoms
in either of the two hyperfine levels of the ground state. A high density was obtained by compression
of the trap, and resulted in two-body collision loss. (From Friedman et al., 2000a, Phys. Rev. A 61,
031403(R), Fig. 4). (b) Diagram illustrating the light-assisted collision loss mechanism, for two
atoms colliding in the presence of a blue-detuned optical field. A pair of atoms in the ground
state (1) approach each other. At the Condon point (Rc) the laser is in resonance with a repulsive
molecular excited state. The pair might be excited by the laser (2), and reach the turning point (Rtp).
Then, the atoms are repelled (3) and, if not brought back to the ground state by the laser, they share
a gain in kinetic energy which is asymptotically equal to h6. (From Suominen, 1996, J. Phys. B 29,
5981, Fig. 4b).

ground-state hyperfine level. The data are well fitted by the solution to Eq. (23)
neglecting the third term (e = 0). The two-body loss coefficients found from the
fit are flF=3 = 2.0• 10-ll cm3/s and flF=2 = 1.2• 10 -ll cm3/s. The larger two-
body collision loss from the upper hyperfine level is due to hyperfine exchange
collisions, since the energy difference between the two ground-state hyperfine
levels in this case (85Rb) is "-4 • 105Erec, which is much higher than the trap depth
of ~ 103Erec. For this reason, it is important to keep atoms in the lower hyperfine
level in high-density traps.
The two-body loss from the lower hyperfine level is attributed to light-assisted
collisions. When two atoms collide in the presence of a light field, absorption
of a photon will transfer them into an excited molecular state. In the case of
red-detuned light, excitation is possible into an attractive molecular state, which
gives rise to loss processes (Weiner et al., 1999; Suominen, 1996) like radiative
escape and photoassociation. Blue-detuned light can excite the colliding atoms
into a repulsive molecular state (Bali et al., 1994) (see Fig. 12b). The atoms are
then accelerated along the repulsive potential curve and obtain a kinetic energy
which is asymptotically equal to the detuning of the exciting laser from the
atomic resonance. This energy is usually much larger than the potential barrier
of the trap, hence both atoms will be lost. Light-assisted binary collisions in the
presence of blue-detuned light were further investigated in the gravito-optical
surface trap (Hammes et al., 2000). Another blue-detuned beam was applied
136 N. Friedman et al. [V

to the trap, and the induced two-body loss coefficients,/3, were measured for
different detunings and intensities of this "catalysis" laser. The loss was found
to be proportional to//62, in the range 5-80 GHz.

B. PRECISIONMEASUREMENTS

The strong suppression of Doppler and time-of-flight broadening due to the ultra-
low temperatures, and the possibility to obtain very long interaction times, are
obvious advantages of using cold atoms for precision optical and rf spectroscopic
measurements. To obtain long atomic coherence times, spontaneous scattering
of photons and energy-level perturbations caused by the trapping laser should
be reduced. This is achieved by increasing the laser detuning from resonance
and trapping the atoms in dark traps. These advantages were demonstrated
already in the first experimental realization of a dark optical trap for cold
atoms (Davidson et aL, 1995), where the ground-state hyperfine splitting
of sodium, dihf, was measured with a very long coherence time of 4s,
yielding a linewidth of 0.125Hz. The magnetic-field insensitive transition
between the IF = 1,mF = 0) and IF = 2, mF = 0) states was excited with a
~l.77-GHz linearly polarized rf wave. A magnetic field parallel to the rf
polarization direction separated the required transition from the magnetic-field
sensitive transitions. During the experiment, the trap was loaded with atoms
which were optically pumped to F = 1. Then, the rf sequence was applied
and the number of atoms making the transition was measured by a state-
selective fluorescence detection of atoms in the F = 2 state. The rf transition
was excited using Ramsey's method of separated oscillatory fields (Ramsey,
1956) by applying two Jr/2 pulses separated by the measurement time T.
The resulting central Ramsey fringes are shown in Fig. 13a, together with a
sinusoidal fit. The fit yields a fringe contrast of 43%, which was found to decay
exponentially with T, with a decay constant of 4.4 s. The uncertainty in the line
central frequency was + 1.3 mHz, for 200 data points collected during 900 s. This
is higher than the shot-noise-limited frequency sensitivity of the Ramsey method,
which is given by A v = (4:r2NtT) -1/2, where N is the number of atoms and
t is the integration time.
The accuracy resolution of the spectroscopic measurement are limited by the
interaction of atoms with the trapping laser field. First, this interaction causes
an average shift in the line center, since it shifts the energy levels of the atom
(in proportion to I/6). This ac Stark shift is different for the two levels used in
the experiment, due to the (very small) difference in the detuning. (For linearly
polarized light, the dipole matrix elements are identical for all sub-levels of the
ground state.) As a result, the ac Stark shift of the hyperfine transition is lower
than the optical Stark shift by a factor of 6/6hf, which is ~ 4.5 • 10 4 in this case.
In the above experiment, a linear dependence of the Stark shift on the trapping
V] DARK OPTICAL TRAPS FOR COLD ATOMS 137

Hyperfine light shift (Hz)


0 0.11 0.22 0.33 0.44
350 I I I
.-. 300 1"0-i; .
r
= 250

~.~ 200
8 0.5
" 150
8
100
O

u_ 50
0.0 "- ~
.... "'"'" ~ 1 7 6 "'" , o . . ~, ~ ,,.

0 I . I,, I
-0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0 5000 10000 15000 20000
f- 1,771,626,130 Hz Light shift (Hz)

(a) (b)
FIG. 13. (a) Central Ramsey fringes of the IF = 1,mf = 0 ) ~ IF = 2, mf = 0 ) rf transition,
measured in the "V"-shaped trap with a 4s measurement time. (b) Calculated ac Stark shift
distributions for atoms stored in blue-detuned dipole traps. The dotted line corresponds to the
"V"-shaped trap, and the solid line is for a more ergodic inverted pyramid trap (composed of three
laser beams). (From Davidson et al., 1995, Phys. Reo. Lett. 74, 1311, Fig. 4).

laser intensity was observed, resulting in a 270-mHz shift of the line center in
the displayed data. Second, it is important to note that this shift is not equal for
atoms which have different trajectories in the trap, yielding an inhomogeneous
distribution of Stark shifts that limits the coherence time of the trap. Hence,
coherence time is related to the dynamics of the trapped atoms. A more chaotic
trap will increase temporal averaging between the atoms and lead to a narrower
distribution and longer coherence times. This averaging effect was calculated
using numerical Monte Carlo simulations; the results are presented in Fig. 13b. In
our discussion, we neglected the contribution of spontaneous photon scattering to
the decoherence rate. This is justified because )'s is smaller than the ac Stark shift
by the factor t~hf/Y ~ 170-2000 for most alkalis. This means that by the time it
takes for a spontaneous scattering event to occur, the inhomogeneous phase shift
of the rf transition due to the ac Stark shift is already hundreds of 2Jr radians.
The Stark shift in the dark optical trap is a great limitation for its performance
as an atomic clock 12. On the other hand, it seems that a dark optical trap is a
very good candidate for precision experiments in atomic physics, such as parity-
nonconservation and permanent electric dipole moment (EDM) measurements
(Bijlsma et al., 1994). Such tabletop experiments are very appealing for tests

12 This limitation can be partially solved by using an additional, very weak, optical field, which
is spatially mode-matched to the trap laser and whose detuning is in the middle of the hyperfine
splitting. The relative Stark shift introduced by this laser compensates that introduced by the trap
(Kaplan et al., 2002b).
138 N. Friedman et al. [V

of the standard model and extensions of it. In some of these theories, for
example, a non-zero EDM value is predicted which is within experimental reach.
Measurement of EDM with optically trapped cold atoms was proposed and
discussed in two recent papers (Romalis and Fortson, 1999; Chin et al., 2001),
and dark optical traps were found to be a promising tool for performing such
measurements with much higher sensitivity than is currently available. The use
of cold atoms can overcome the two limiting factors of current experiments,
namely systematic errors due to the atomic velocity in beam experiments, and
leakage currents in cell experiments.
The EDM measurement is based on measuring a possible energy shift between
two Zeeman sublevels when a static electric field is applied. Therefore, any
perturbation of the Zeeman sublevels should be kept minimal. In dark optical
dipole traps, the limits on the accuracy of the measurement are due to interactions
between the trapping light and the atoms, which cause frequency shifts between
the Zeeman sublevels. The three leading terms of these interactions are (Romalis
and Fortson, 1999): a vector shift caused by a residual circular polarization of
the trapping laser; tensor shifts which result from the interaction of the atoms
with the trapping light in the presence of the static electric field; and a third-
order effect which represents interaction of an induced electric dipole with
the laser field through magnetic-dipole or electric-quadrupole interaction. The
enhanced cross section for cold atomic collisions may result in frequency shifts
(Gibble and Chu, 1993; Bijlsma et al., 1994) and should also be avoided. These
limiting factors have been analyzed by Romalis and Fortson (1999) for the case
of cesium atoms confined in either a red- or a blue-detuned dipole trap. The
vector light shift can be lowered by reducing the residual circular polarization
of the trapping beam, and aligning the beam propagation direction perpendicular
to the static magnetic field (which defines the quantization axis). When the
trapping laser is detuned above resonance, destructive interference between the
amplitudes of the vector light shift from two resonance lines lowers the total
shift 13. The tensor shift is eliminated at a "magic angle" (54.7 ~ which satisfies
3 cos 2 r 1 = 0) between the direction of the electric field of the light and the
quantization axis, or by measuring a IF, mF) ~ I F , - m F ) transition (as suggested
by Chin et al., 2001), for which the tensor shift, which depends on m~, vanishes.
In another work, Chin et al. (2001) have made a detailed error analysis
for a specific experimental realization of an EDM measurement, for Cs atoms
in a dark optical lattice. The calculation was made for a lattice which is
realized with a 532-nm laser (detuning of 4 x 107}I above the Cs 6S1/2 ---+ 6P3/2
transition). The optimal trap depth is "~130Erec, and the photon scattering rate
for atoms in the lattice ground state is -~7x 10-3 s-1. The proposed lattice

13 For Cs, it actually vanishes for two wavelengths: 464 and 474nm, but no sufficiently strong
laser lines are available at these wavelengths.
V] DARK OPTICAL TRAPS FOR COLD ATOMS 139

FIG. 14. (a) Proposed experimental configuration for a measurement of the electron's electric
dipole moment using Cs atoms trapped in a dark optical lattice. The lattice is composed of three
linearly polarized standing waves, with different frequencies. The polarizations of the beams are
perpendicular to the quantization axis defined by the external electric and magnetic fields along
the z-axis. (b) Ground-state tunneling rate (solid line) and scattering rate for blue- (dotted) and
red-detuned (dashed) 3D lattices. Lattice detuning is assumed constant, and the lattice depth is
changed by increasing the laser intensity. The arrow marks the operation point of the proposed
scheme, where tunneling and scattering rates are equal for a blue-detuned lattice. (From Chin et al.,
2001, Phys. Rev. A 63, 033401, Figs. 1, 2).

configuration has 3 linearly polarized standing waves at different directions (see


Fig. 14a), which have a frequency difference of few MHz between them. In
this way, the polarization is effectively linear in every point in the lattice. The
optical lattice is a very flexible trap, and its parameters (beam polarization and
propagation directions) are chosen in a way which minimizes systematic errors
and decoherence processes. The energy difference between the IF = 3, m F = 2)
and IF = 3, m F = --2) states can be measured in the Ramsey method, using pulses
of circularly polarized light to induce two-photon Raman transitions inside the
F = 3 level of the ground state. The main advantage of using a lattice is the
great reduction in collision rate, since atoms are isolated in different lattice sites,
and the collision rate is then limited by tunneling between neighboring sites.
In Fig. 14b, tunneling and scattering rates for the ground states of a red-
and a blue-detuned lattice are plotted. When the lattice is made deeper while
keeping its detuning constant, the tunneling rate is decreased but the scattering
is increased (see Eq. 8). The working point is chosen so as to equalize the
two rates. As can be seen from the figure, the tunneling rate at the working
point is 20 times lower in the blue-detuned lattice than in the red-detuned one.
For a singly occupied lattice with the chosen parameters, and atoms cooled
to the ground state (e.g. by Raman sideband cooling, see Sect. V.A.2), the
collision rate is lowered by a factor of 106 as compared to free atoms with the
same bulk density and kinetic energy. Another advantage of the lattice is that
140 N. Friedman et al. [V

a different atomic species (Rb, for example) can be trapped in the same lattice,
and serve as a "co-magnetometer," to investigate and correct potential systematic
effects induced by the external fields. It is estimated that using this system with
108 trapped atoms, a 1 s coherence time, and a measurement time of 8 hours,
the sensitivity of the EDM measurement will be 100 times better than current
experiments. The main limitations are stringent requirements on the intensity and
polarization stability of the trapping laser, and the 3rd-order polarizability effect.
Another application for which dark optical traps are considered (Kulin et al.,
2001) is parity-nonconservation measurements (PNC). For these experiments,
the radioactive alkali atom francium appears as a good candidate, since it is
predicted to have a large PNC effect (18 times larger than Cs). In the last few
years, Fr was efficiently trapped in a MOT (Simsarian et al., 1996; Lu et al.,
1997), and its energy structure was investigated spectroscopically (Sprouse et al.,
1998). Precision spectroscopy experiments on Fr atoms in a dark optical trap
seem feasible, and might improve experimental tests of the standard model.
Dark traps are useful also for spectroscopic measurements of extremely
weak optical transitions. While preserving long atomic coherence times, those
traps can provide large spring constants and tight confinement of trapped
atoms, which ensure good spatial overlap even with a tightly focused excitation
laser beam. Therefore, the atoms can be exposed to a much higher intensity
of the excitation laser, yet being relatively unperturbed by the trapping light.
This yields an increased sensitivity for very weak transitions, and especially for
multi-photon transitions. This property was demonstrated by measuring a two-
photon transition in cold Rb atoms trapped in a scanning-beam optical trap, with
a very weak probe laser (Khaykovich et al., 2000). A spectroscopy scheme which
exploits the long spin-relaxation time of the dark trap was used.
In this scheme (see Fig. 15a), atoms with two ground-state hyperfine levels
(Ig,), Ig2)) are stored in the trap in a level Ig,) that is coupled to the upper
(excited) state, [e), by an extremely weak transition which is excited with a
laser. An atom that undergoes the weak transition may be shelved, through
a spontaneous Raman transition, in ]g2), which is uncoupled to the excited
level. After waiting long enough, a significant fraction of the atoms will be
shelved in Ig2)- The detection benefits from multiply excited fluorescence of
a strong cycling transition from the shelved level ]g2). Thanks to the use of
a stable ground state ([g2)) as a "spin shelf," the quantum amplification is
limited only by spin-relaxation processes which are strongly suppressed in a
dark trap. This scheme was realized on the 5S1/2 ----+ 5D5/2 two-photon transition
in cold and trapped 85Rb atoms (see Fig. 15a for the relevant energy levels). The
trapped atoms were optically pumped to the lower (F = 2) hyperfine level. The
spectroscopy was made with an extremely weak (25 ~tW), frequency-stabilized
laser beam, which excited the two-photon transition. The scanning-beam trap was
loaded and then compressed to a lower radius (Friedman et aL, 2000a), to best
V] DARK OPTICAL TRAPS FOR COLD ATOMS 141

• • 5D,a le)

6P,~__~_~ / F'--4 F,=3 b)


6p,~ - ~ a - ~ / 777.9tim 0.45 ,.',,
420.:t
/// o
.g ,.. ': F'--2
5P~ ~ 0.40 ,~ q' I'=
~ 5P,a q
: k } ~ ' ~ V'=l
0.35 9 4

780.24mn
F=3 ig~...,~ 0.30
F--2 I ~ 5Sla 5 10 15 20 25
Frequency [MHz]

(a) (b)
FIG. 15. (a) Energy levels of 85Rb and the relevant transitions for a two-photon spectroscopy
experiment inside a scanning-beam dark trap. Spectroscopy of the Igl) ~ ]e) transition
( 5 S 1 / 2 , F = 2 --~ 5 D 5 / 2 , F t in the case of 85Rb) is performed. Atoms which undergo the transition
are shelved in the level ]g2) (5S1/2, F = 3), from which they are detected using a cycling transition
(to 5 P 3 / 2 , F = 4). (b) Measured spectrum of the 5 S I / 2 , F = 2 ~ 5 D 5 / 2 , F t = 4,3,2, 1 two-photon
transition. Each point corresponds to an experimental cycle, in which atoms are exposed during
500ms to a 25~tW excitation laser. (From Khaykovich et al., 2000, Europhys. Lett. 50, 454,
Figs. 1, 4b).

overlap with the excitation laser. After 500 ms, the fraction of atoms transferred
to F = 3 was measured. In Fig. 15b, the measured F = 3 fraction is presented as
a function of the frequency of the two-photon laser. The hyperfine splitting of the
excited state was resolved, and the measured frequency differences and relative
line strengths are in a good agreement with theory and previous measurements
(performed with much higher laser intensities). A transition rate as small as
0.09 s -I was measured, with a "quantum rate amplification," due to spin shelving,
o f ~ 1 0 7 . This optical spectroscopy technique can be applied for other weak
(forbidden) transitions such as optical clock transitions (Ruschewitz et al., 1998;
Kurosu et aL, 1998) and parity-violating transitions where a much lower mixing
with an allowed transition could be used.

C. DYNAMICS OF THE TRAPPED ATOMS

In a dark optical trap, atoms move freely in the dark region, and reflect elastically
from the trapping potential. The similarity of this system to the well-studied
billiard problem has recently led to the realization of "atom-optics billiards"
(Milner et al., 2001; Friedman et aL, 2001b) in which cold atoms move inside
dark traps of various shapes. The motion of the atoms is governed by the shape of
the trap, and can exhibit different types of dynamics, from regular to chaotic. The
billiards were formed as scanning-beam traps, using two perpendicular acousto-
142 N. F r i e d m a n et al. [V

FIG. 16. Classical numerical simulations of two-dimensional trajectories of Rb atoms in


atom-optics billiards of various shapes. A 16 ~tm blue-detuned laser beam scans along the boundary
(shown in the figure) at a 100kHz rate. (a) Circular billiard: an example of a nearly periodic
trajectory, which requires an increasinglylong time to sample all regions on the boundary. (b) "Tilted"
Bunimovich stadium: the motion is chaotic, every trajectory would reach a certain region on the
boundary with a comparable time scale. (c) Elliptical billiard: an example of an "internal" trajectory,
which is confined by a hyperbolic caustic and thus excluded from a certain part of the boundary.
The other type of trajectories ("external," not shown) reaches every region on the boundary. (From
Friedman et al., 2001b, Phys. Rev. Lett. 86, 1518, Fig. 1).

optic scanners, as described in Sect. IV.C. In some cases, a stationary blue-


detuned standing wave was applied along the optical axis, to confine the atomic
motion to two-dimensional planes. In this case the atoms are tightly confined near
the node planes of the standing wave, forming "pancake"-shaped traps separated
b y - 4 0 0 n m (half the wavelength of the standing-wave laser), all of them with
a nearly identical billiard potential in the radial direction. The dynamics of
the trapped atoms was probed by opening a small hole in the boundary, and
measuring the decay of atoms from the billiard through this hole. The hole is
opened by switching off one of the AOSs, synchronously with the scan. The
decay of atoms through the hole depends on their dynamics, as illustrated by a
comparison of the atomic trajectories for a circular billiard (Fig. 16a) and a tilted
Bunimovich stadium billiard (two semicircles with different radii, connected
with straight lines, Fig. 16b). For the circular billiard, in which the motion is
integrable (regular), nearly periodic trajectories exist (see Fig. 16a) that require
an increasingly long time to sample all regions on the boundary 14. This yields
many time scales for the decay through a small hole on the boundary, and
results in an algebraic decay at the long-time limit (Bauer and Bertsch, 1990).
For the tilted-stadium billiard, phase space is chaotic, hence each trajectory
samples every point on the boundary with an equal probability. This results in
a pure exponential decay (Bauer and Bertsch, 1990; Alt et al., 1996), with a

14 Exactly periodic trajectories that are completely stable have only a zero measure, and hence can
be neglected.
V] DARK OPTICAL TRAPS FOR COLD ATOMS 143

FIG. 17. Survival probability of atoms in the gravitational wedge billiard, 300 ms after opening the
hole, as a function of the wedge half-angle. Lower curve: experimental data. Upper curve: results of
a classical numerical simulation (divided by 2 and shifted up by 0.1). The inset shows the measured
intensity cross section at the focal plane of the billiard (with a hole) (From Milner et al., 2001,
Phys. Rev. Lett. 86, 1514, Figs. 3, 4).

(I/e) decay timescale of rc = ( ~ A ) / ( v L ) , with A the billiard area, v the atomic


velocity and L the length of the hole (Bauer and Bertsch, 1990)15.
In a recent experiment (Milner et al., 2001), cold Cs atoms were trapped
inside a "gravitational wedge" billiard (shown as inset to Fig. 17). The
dynamical behavior of this billiard can be tuned from stability to chaos with a
single parameter, the vertex half-angle of the wedge, 0. Numerical simulations
demonstrate that for 0>45 ~ the system is fully chaotic. For 0 < 45 ~ the system
has a mixed phase space, i.e. it has chaotic regions coexisting with stable periodic
and quasiperiodic trajectories. For 0 = 90~ (n = 3, 4 , . . . ) , the chaotic part of
phase space is minimal, and most of the phase space is regular. In between these
angles, the fraction of chaos is larger. The billiard was realized experimentally
by loading Cs atoms from a MOT into a scanning-beam optical trap with the
required wedge shape. (This trap is actually gravity-assisted, similar to the
"V"-shape trap described in Sect. III.A). A hole was opened at the apex of the
wedge, and the number of atoms left in the trap after 3 0 0 m s was measured.
In Fig. 17, the measured survival probability in the billiard is presented as a
function of 0. As expected, maxima are observed in the survival probability
for 0 = 22.5 ~ and 30 ~ (90~ and 90~ whereas for intermediate angles
and above 45 ~ the motion is less stable. These observations are in very good
agreement with theory, and with the results of classical numerical simulations
of the system, which are also shown. The factor of 2 discrepancy between the
simulation and the experiment is attributed either to collisions between atoms,

15 A "tilted" stadium (and not the more common Bunimovich stadium) is used in the experiment
in order to reduce the number of nearly stable trajectories (Vivaldi et al., 1983).
144 N. Friedman et al. [V

FIG. 18. Survival probability of atoms in billiards of various shapes, as a function of time after
opening a hole in the boundary. (a) Elliptical billiard. The solid symbols denote the unperturbed
case, in which the surviving fraction for the ellipse with the hole on the long side (solid squares)
decays much faster than for the hole on the short side (solid circles). The open symbols show the
case in which 10~ts velocity randomizing molasses pulses are applied every 3 ms. (b) Decay of
atoms from circular and stadium billiards: The decay from the stadium billiard (solid circles) shows
a nearly pure exponential decay. For the circle (solid diamonds) the decay curve flattens, indicating
the existence of nearly stable trajectories. The solid lines represent numerical simulations, including
all the experimental parameters, and no fitting parameters. The dashed line represents exp(-t/rc),
where r c is the escape time calculated for the experimental parameters. The insets show CCD-camera
images of the billiards' cross sections at the beam focus. The size of the images is 300x 300 ~tm.
(From Friedman et al., 2001b, Phys. Rev. Lett. 86, 1518, Figs. 2, 4).

which are not included in the simulation and may decrease the stability, or to
imperfections in the trapping beam.
In another set of experiments, billiards of various shapes were investigated
(Friedman et al., 2001b). First, "macroscopic" separation in phase space was
measured, for the elliptical billiard. Here, phase space is divided into two separate
regions (Koiller et al., 1996): "external" trajectories that are confined outside
elliptical caustics (smaller than the billiard itself but with the same focal points),
and "internal" trajectories confined by hyperbolic caustics, again with the same
focal points, as shown in Fig. 16c. Hence, if a hole exists at the short side of
the ellipse (upper inset of Fig. 18a), atoms in those trajectories remain confined
and never reach the hole. Alternatively, all trajectories, excluding a zero-measure
amount, reach the vicinity of a hole on the long side of the ellipse (lower
inset of Fig. 18a) and hence the number of confined atoms decays indefinitely.
Figure 18a shows the measured survival probability for cold Rb atoms in the
elliptical optical billiard with a hole on either the long or the short side. At
long times, the survival probability for the hole on the short side is much higher
than for the hole on the long side, as expected from the discussion above. Next,
a controlled amount of randomization was introduced to the atomic motion, by
exposing the confined atoms to a series of short PGC pulses (using the six MOT
beams). During each pulse, an atom scatters "--30 photons, and hence its direction
of motion is completely randomized, whereas the total velocity distribution
remains statistically unchanged. The measured decay curves for this case, with a
V] DARK OPTICAL TRAPS FOR COLD ATOMS 145

PGC pulse every 3 ms, are also shown in Fig. 18a, for the two hole positions of
the ellipse. As seen, for the hole on the long side, the randomizing pulses cause
little change. However, for the hole on the short side, a complete destruction
of the stability occurs, and the decay curves for the two hole positions coincide
approximately. The stability decrease was found to be gradual, and the pulse rate
required to significantly reduce the stability is approximately one pulse per rc,
the average 1/e decay time.
Next, decay curves were measured for the circular and the tilted-stadium
billiards. Since the atoms were loaded into the billiard from a cloud in
thermal equilibrium, their velocity was distributed around zero, with a measured
RMS velocity distribution of 1 It)recoil. To better approximate a mono-energetic
ensemble of atoms, and to reduce the relative contribution of gravitational
energy (~< 50Erecoil), the loading scheme was modified. After loading from the
MOT into the trap, the atoms were illuminated with a short pulse of a strong,
on-resonance pushing beam perpendicular to the billiard beam. Following this
pushing beam, the center of the velocity distribution was shifted to 20t)recoil,
while the RMS width was barely changed. The hole was opened 50ms after
the push, to allow for a randomization of the direction of the transverse velocity
through collisions with the billiard's boundaries. The measured decay curves for
the circular and the tilted-stadium billiards, with equal area and hole size, are
shown in Fig. 18b. The decay from the circular billiard is slower, indicating the
existence of nearly stable trajectories, whereas that of the stadium is a nearly
pure exponential. Also shown in the figure are the results of full numerical
simulations that contain no fitting parameters. The simulations include the
measured three-dimensional atomic and laser-beam distributions, atomic velocity
spread, laser-beam scanning, and gravity. It is seen that there is fairly good
agreement between the simulations and the data.
As opposed to ideal billiards which have an infinite potential wall, optical
billiards inherently have a soft-wall potential, which may affect the dynamics
(Rom-Kedar and Turaev, 1999; Gerland, 1999; Sachrajda et al., 1998). For
example, a soft wall may introduce stable regions into an otherwise chaotic
phase space, and create "islands of stability" immersed in a chaotic sea. This
structure greatly affects the transport properties of the system, since trajectories
from the chaotic part of phase space are trapped for long times near the boundary
between regular and chaotic motion (Zaslavski, 1999). This was demonstrated
in a recent experiment by Kaplan et aL (2001), who compared the decay
from billiards with hard and soft walls. The softness of the billiards' wall
was experimentally changed by varying w0 of the scanning beam. In Fig. 19,
experimental results for the decay from a tilted stadium billiard with a harder
(w0 = 14.5~m) and a softer (w0 = 24~tm) wall are presented. When the
hole is located entirely inside the big semi-circle (Fig. 19a), the soft wall
causes an increased stability, and a slowing down in the decay curve. When the
146 N. F r i e d m a n et al. [V

FIG. 19. The effect of wall softness: measured survival probability of atoms in a tilted-stadium
shaped billiard, with two different values of wall softness: w0 = 14.5 ~tm (o), and w0 = 24.5 ~tm (+).
(a) The hole is located inside the big semicircle. The smoothening of the potential wall causes a
growth in stability, and a slowing down in the decay curve. (b) The hole includes the singular point,
no effect for the change in w0 is seen. Also shown are results of numerical simulations, with the
experimental parameters and no fitting parameters. The dashed line represents exp(-t/rc), the decay
curve for an ideal (hard-wall) billiard. The insets show measured intensity cross sections for the
soft-wall billiards, in the beam's focal plane. The size of the images is 300• ~tm. (From Kaplan
et al., 2001, Phys. Rev. Lett. 87, 274101, Fig. 1).

hole includes the singular point, where the straight line and semi-circle meet,
(Fig. 19b), no effect for the change in w0 is seen. These results can be explained
by the formation of a stable island around the trajectory that connects both
singular points, and a "sticky" region around it. This explanation is supported
by numerical simulations, which predict the formation of a stable island around
the singular trajectory when the wall becomes soft. Around this stable island
there is a "sticky" area in which chaotic trajectories spend a long time before
escaping back to the chaotic part of phase space. Similar decay measurements
and simulations for a circular atom-optics billiard showed no dependence on w0
in the range 14.5-24 ~m, and no dependence on the hole position. In another
work, it was shown that adding a force field across the billiard can also stabilize
specific orbits in otherwise chaotic billiards (Andersen et al., 2002a).
Two theoretical works (Liu and Milburn, 1999, 2000) are related to the optical
billiard system. In these works, the classical and quantum dynamics of a gas of
cold atoms trapped inside a circular hollow laser beam or a hollow fiber was
investigated, when the intensity of the trapping light is periodically modulated. In
this system, chaotic dynamics exists for certain values of the modulation index,
and causes atoms to accumulate in rings corresponding to fixed points of the
system.
The ability to form billiards of arbitrary shape which can also be varied
dynamically, together with the precise control of parameters offered by laser-
cooling techniques, provide a powerful tool for the study of dynamical quantum
effects. These effects are expected to become important at lower temperatures
and smaller billiards. Very interesting in this context is the investigation of
the properties of a BEC trapped in an integrable or a chaotic billiard. Other
VI] DARK OPTICAL TRAPS FOR COLD ATOMS 147

problems in nonlinear dynamics can also be addressed using atom-optics billiard,


including the influence of many-body interactions, external fields and noise on
the dynamics of the atoms. The precise control of the atomic motion in these
optical traps can also find useful application in stochastic cooling of atoms
(Raizen et al., 1998) and precision spectroscopy. In stochastic cooling, the
existence of chaotic motion is an important requirement for mixing the velocities
of the atoms after a cooling step. In precision spectroscopy, chaotic motion may
reduce inhomogeneous line broadening, as described in the previous section. The
performance of an echo scheme, in which a Jr-pulse is added in the middle of
the measurement time (between the two Jr/2 pulses in the Ramsey method),
should also depend on the dynamics (Andersen et al., 2002b), and may be
used to investigate connections between dynamics and decoherence (Jalabert and
Pastawski, 2001).

VI. Conclusions

In this chapter, we have reviewed the main configurations that are used to form
dark optical dipole traps, and their principal applications. The formation of a
dark, blue-detuned trap is less obvious than that of a red-detuned one, and, as
discussed in Sects. III and IV, some of the effort in the last few years was
directed towards the generation of improved schemes, which are also easier
to implement experimentally. This effort has led to the development of traps
with larger volumes, better loading efficiencies, more efficient use of the laser
power and a larger darkness factor. We believe that dark optical traps have
matured and will now enter into more applications, in which their advantages
will be important. These will include precision spectroscopic measurements,
where the reduced interaction with the trapping field is crucial, and investigation
of atomic dynamics inside atom-optics billiards, both as a model system
for quantum and mesoscopic dynamics and as a tool to further improve the
accuracy of spectroscopic measurements. Other applications may benefit from
the ability to confine atoms with reduced interactions, including quantum
information processing in dark optical lattices, and possibly also quantum-optics
experiments which require long relaxation times of the atomic spins, such as
slow light (Hau et al., 1999; Kash et al., 1999), stopped light (Liu et al., 2001;
Phillips et al., 2001), and entangled atomic samples (Julsgaard et al., 2001).
Finally, the special light distributions employed to trap cold atoms can also be
used to trap electrons (Chaloupka and Meyerhofer, 1999), Rydberg atoms (Dutta
et al., 2000) and molecules (Seideman, 1999) using the ponderomotive force, and
to manipulate larger objects when used as dark optical tweezers (Sasaki et al.,
1992). In all these cases, the advantages of dark traps are twofold: The ability to
trap dark-field seeking objects (e.g. electrons, or metallic beads) and the reduced
light intensity to which the trapped object is exposed.
148 N. Friedman et al. [VII

VII. References

Adams, C.S., and Riis, E. (1997). Prog. Quantum Electron. 21, 1.


Adams, C.S., Lee, H.J., Davidson, N., Kasevich, M., and Chu, S. (1995). Phys. Rev. Lett. 74, 3577.
Allen, L., Padgett, M.J., and Babiker, M. (1999). Prog. Opt. 39, 291.
Alt, H., Gr~if, H.D., Harney, H.L., Hofferbert, R., Rehfeld, H., Richter, A., and Schard, P. (1996).
Phys. Rev. E 53, 2217.
Aminoff, C.G., Steane, A.M., Bouyer, P., Desbiolles, P., Dalibard, J., and Cohen-Tannoudji, C.
(1993). Phys. Rev. Lett. 71, 3083.
Andersen, M.E, Kaplan, A., Friedman, N., and Davidson, N. (2002a). J. Phys. B: At. Mol. Opt.
Phys. 35, 2183.
Andersen, M.F., Kaplan, A., and Davidson, N. (2002b). quant-ph/0204082.
Anderson, B.P., Gustavson, T.L., and Kasevich, M.A. (1996). Phys. Rev. A 53, R3727.
Arlt, J., and Padgett, M.J. (2000). Opt. Lett. 25, 191.
Arlt, J., Kuhn, R., and Dholakia, K. (2001 a). J. Mod. Opt. 48, 783.
Arlt, J., Dholakia, K., Soneson, J., and Wright, E.M. (200 l b). Phys. Rev. A 63, 063602.
Bali, S., Hoffmann, D., and Walker, T. (1994). Europhys. Lett. 27, 273.
Bali, S., O'Hara, K.M., Gehm, M.E., Granade, S.R., and Thomas, J.E. (1999). Phys. Rev. A 60, R29.
Balykin, V.I. (1999). Adv. Atom Mol. Opt. Phys. 41, 181.
Balykin, V.I., Minogin, V.G., and Letokhov, V.S. (2000). Rep. Prog. Phys. 63, 1429.
Barrett, M.D., Sauer, J.A., and Chapman, M.S. (2001). Phys. Rev. Lett. 87, 010404.
Bauer, W., and Bertsch, G.F. (1990). Phys. Rev. Lett. 65, 2213.
Beijersbergen, M.W., Allen, L., van der Veen, H.E.L.O., and Woerdman, J.P. (1993). Opt. Commun.
96, 123.
Beijersbergen, M.W., Coerwinkel, R.P.C., Kristensen, M., and Woerdman, J.P. (1994). Opt. Commun.
112, 321.
Bijlsma, M., Verhaar, B.J., and Heinzen, D.J. (1994). Phys. Rev. A 49, R4285.
Boiron, D., Michaud, A., Lemonde, P., Castin, Y., Salomon, C., Weyers, S., Szymaniec, K., Cognet, L.,
and Clairon, A. (1996). Phys. Rev. A 53, R3734.
Boiron, D., Michaud, A., Fournier, J.M., Simard, L., Sprenger, M., Grynberg, G., and Salomon, C.
(1998). Phys. Rev. A 57, R4106.
Bongs, K., Burger, S., Dettmer, S., Hellweg, D., Arlt, J., Ertmer, W., and Sengstock, K. (2001). Phys.
Rev. A 63, 031602.
Brennen, G.K., Caves, C.M., Jessen, P.S., and Deutsch, I.H. (1999). Phys. Rec. Lett. 82, 1060.
Cacciapuoti, L., de Angelis, M., Pierattini, G., and Tino, G.M. (2001). Eur. Phys. J D 14, 373.
Castin, Y., Cirac, J.I., and Lewenstein, M. (1998). Phys. Rev. Lett. 80, 5305.
Chaloupka, J.L., and Meyerhofer, D.D. (1999). Phys. Rev. Lett. 82, 4538.
Chaloupka, J.L.,'and Meyerhofer, D.D. (2000). J. Opt. Soc. Am. B 17, 713.
Chaloupka, J.L., Fisher, Y., Kessler, T.J., and Meyerhofer, D.D. (1997). Opt. Lett. 22, 1021.
Chin, C., Leiber, V., Vuletic, V., Kerman, A.J., and Chu, S. (2001). Phys. Rev. A 63, 033401.
Cline, R.A., Miller, J.D., Matthews, M.R., and Heinzen, D.J. (1994). Opt. Lett. 19, 207.
Cohen-Tannoudji, C., Dupont-Roc, J., and Grynberg, G. (1992). "Atom-Photon Interactions." Wiley,
New York.
Davidson, N., Lee, H.J., Kasevich, M., and Chu, S. (1994). Phys. Rev. Lett. 72, 3158.
Davidson, N., Lee, H.J., Adams, C.S., Kasevich, M., and Chu, S. (1995). Phys. Rev. Lett. 74, 1311.
Davidson, N., Ozeri, R., and Baron, R. (1999). Rev. Sci. Instrum. 70, 1264.
DePue, M.T., McCormick, C., Winoto, S.L., Oliver, S., and Weiss, D.S. (1999). Phys. Rev. Lett.
82, 2262.
Desbiolles, P., Arndt, M., Szriftgiser, P., and Dalibard, J. (1996). Phys. Rev. A 54, 4292.
Dutta, S.K., Guest, J.R., Feldbaum, D., Walz-Flannigan, A., and Raithel, G. (2000). Phys. Rev. Lett.
85, 5551.
VII] DARK OPTICAL TRAPS FOR COLD ATOMS 149

Engler, H., Manek, I., Moslener, U., Nill, M., Ovchinnikov, Yu.B., Schl6der, U., Schfinemann, U.,
Zielonkowski, M., Weidemfiller, M., and Grimm, R. (1998). Appl. Phys. B 67, 709.
Friedman, N., Khaykovich, L., Ozeri, R., and Davidson, N. (2000a). Phys. Rev. A 61, 031403.
Friedman, N., Kaplan, A., and Davidson, N. (2000b). Opt. Lett. 25, 1762.
Friedman, N., Kaplan, A., and Davidson, N. (2001a). unpublished.
Friedman, N., Kaplan, A., Carasso, D., and Davidson, N. (2001b). Phys. Rev. Lett. 86, 1518.
Gauck, H., Hartl, M., Schneble, D., Schnitzler, H., Pfau, T., and Mlynek, J. (1998). Phys. Rev. Lett.
81, 5298.
Gerland, U. (1999). Phys. Rev. Lett. 83, 1139.
Gibble, K., and Chu, S. (1993). Phys. Rev. Lett. 70, 1771.
Grimm, R., Weidemuller, M., and Ovchinnikov, Y.B. (2000). Adv. Atom Mol. Opt. Phys. 42, 95.
Hamann, S.E., Haycock, D.L., Klose, G., Pax, P.H., Deutsch, I.H., and Jessen, P.S. (1998). Phys.
Rev. Lett. 80, 4149.
Hammes, M., Rychtarik, D., Druzhinina, V., Moselner, U., Manek-Honninger, I., and Grimm, R.
(2000). J. Mod. Opt. 47, 2755.
Hammes, M., Rychtarik, D., and Grimm, R. (2001). C.R. Acad. Sci. IV 2, 625.
Han, D.J., Wolf, S., Oliver, S., McCormick, C., DePue, M.T., and Weiss, D.S. (2000). Phys. Rev.
Lett. 85, 724.
Han, D.J., DePue, M.T., and Weiss, D.S. (2001). Phys. Rev. A 63, 023405.
Hau, L.V., Harris, S.E., Dutton, Z., and Behroozi, C.H. (1999). Nature, 397, 594.
Heckenberg, N.R., McDuff, R., Smith, C.P., Rubinsztein-Dunlop, H., and Wegener, M.J. (1992). Opt.
Quantum Electron. 24, $951.
Inouye, S., Andrews, M.R., Stenger, J., Miesner, H.-J., Stamper-Kurn, D.M., and Ketterle, W. (1998).
Nature 392, 151.
Jalabert, R.A., and Pastawski, H.M. (2001). Phys. Rev. Lett. 86, 2490.
Julsgaard, B., Kozhekin, A., and Polzik, E.S. (2001). Nature 413, 400.
Kaplan, A., Friedman, N., Andersen, M., and Davidson, N. (2001). Phys. Rev. Lett. 87, 274101.
Kaplan, A., Friedman, N., and Davidson, N. (2002a). J. Opt. Soc. Am. B 19, 1233.
Kaplan, A., Andersen, M.E, and Davidson, N. (2002b). physics~0204082. Phys. Rev. A, in press.
Kasevich, M., and Chu, S. (1992). Phys. Rev. Lett. 69, 1741.
Kasevich, M.A., Weiss, D.S., and Chu, S. (1990). Opt. Lett. 15, 607.
Kash, M.M., Sautenkov, V.A., Zibrov, A.S., Hollberg, L., Welch, G.R., Lukin, M.D., Rostovtsev, Y.,
Fry, E.S., and Scully, M.O. (1999). Phys. Rev. Lett. 82, 5229.
Kerman, A.J., Vuleti6, V., Chin, C., and Chu, S. (2000). Phys. Rev. Lett. 84, 439.
Ketterle, W., and VanDruten, N.J. (1996). Adv. Atom. Mol. Opt. Phys. 37, 181.
Khaykovich, L., Friedman, N., Baluschev, S., Fathi, D., and Davidson, N. (2000). Europhys. Lett.
50, 454.
Koiller, J., Markarian, R., Oliffson Kamphorst, S., and Pinto de Carvalho, S. (1996). J. Stat. Phys.
83, 127.
Kuga, T., Torii, Y., Shiokawa, N., and Hirano, T. (1997). Phys. Rev. Lett. 78, 4713.
Kulin, S., Aubin, S., Christe, S., Peker, B., Rolston, S.L., and Orozco, L.A. (2001). J. Opt. B:
Quantum Semiclass. Opt. 3, 353.
Kuppens, S., Rauner, M., Schiffer, M., Sengstock, K., Ertmer, W., van Dorsselaer, EE., and
Nienhuis, G. (1998). Phys. Rev. A 58, 3068.
Kuppens, S.J.M., Corwin, K.L., Miller, K.W., Chupp, T.E., and Wieman, C.E. (2000). Phys. Rev. A
62, 013406.
Kurosu, T., Zinner, G., Trebst, T., and Riehle, E (1998). Phys. Rev. A 58, R4275.
Lee, H.J., and Chu, S. (1998). Phys. Rev. A 57, 2905.
Lee, H.J., Adams, C.S., Kasevich, M., and Chu, S. (1996). Phys. Rev. Lett. 76, 2658.
Liu, C., Dutton, Z., Behroozi, C.H., and Hau, L.V. (2001). Nature 409, 490.
Liu, X.M., and Milburn, G.J. (1999). Phys. Rev. E 59, 2842.
150 N. F r i e d m a n et al. [VII

Liu, X.M., and Milburn, G.J. (2000). Phys. Rev. A 61, 053401.
Lu, Z.-T., Corwin, K.L., Vogel, K.R., Wieman, C.E., Dinneen, T.P., Maddi, J., and Gould, H. (1997).
Phys. Rev. Lett. 79, 994.
Major, EG., and Dehmelt, H. (1968). Phys. Rev. 170, 91.
Manek, I., Ovchinnikov, Y.B., and Grimm, R. (1998). Opt. Commun. 147, 67.
Milner, V., Hanssen, J.L., Campbell, W.C., and Raizen, M.G. (2001). Phys. Rev. Lett. 86, 1514.
Morsch, O., and Meacher, D.R. (1998). Opt. Commun. 148, 49.
Muller-Seydlitz, T., Hartl, M., Brezger, B., Hansel, H., Keller, C., Schnetz, A., Spreeuw, R.J.C.,
Pfau, T., and Mlynek, J. (1997). Phys. Rev. Lett. 78, 1038.
O'Hara, K.M., Granade, S.R., Gehm, M.E., and Thomas, J.E. (2001). Phys. Rev. A 63, 043403.
Oron, R., Davidson, N., Friesem, A.A., and Hasman, E. (2000). Opt. Commun. 182, 205.
Ovchinnikov, Y.B., Soding, J., and Grimm, R. (1995). JETP Lett. 61, 21.
Ovchinnikov, Y.B., Manek, I., and Grimm, R. (1997). Phys. Rev. Lett. 79, 2225.
Ovchinnikov, Y.B., Manek, I., Sidorov, A.I., Wasik, G., and Grimm, R. (1998). Europhys. Lett.
43, 510.
Ozeri, R., Khaykovich, L., and Davidson, N. (1999). Phys. Rev. A 59, R1750.
Ozeri, R., Khaykovich, L., Friedman, N., and Davidson, N. (2000). J. Opt. Soc. Am. B 17, 1113.
Perrin, H., Kuhn, A., Bouchoule, I., and Salomon, C. (1998). Europhys. Lett. 42, 395.
Perrin, H., Kuhn, A., Bouchoule, I., Pfau, T., and Salomon, C. (1999). Europhys. Lett. 46, 141.
Petrich, W., Anderson, M.H., Ensher, J.R., and Cornell, E.A. (1995). Phys. Rev. Lett. 74, 3352.
Petrov, D.S., Holzmann, M., and Shlyapnikov, G.V. (2000). Phys. Rev. Lett. 84, 2551.
Phillips, D.E, Fleischhauer, A., Mair, A., and Walsworth, R.L. (2001). Phys. Rev. Lett. 86, 783.
Piestun, R., and Shamir, J. (1994). Opt. Lett. 19, 771.
Pinkse, P.W.H., Mosk, A., Weidemfiller, M., Reynolds, M.W., Hijmans, T.W., and Walraven, J.T.M.
( ! 997). Phys. Rev. Lett. 78, 990.
Raab, E.L., Prentiss, M., Cable, A., Chu, S., and Pritchard, D.E. (1987). Phys. Rev. Lett. 59, 2631.
Raizen, M.G., Koga, J., Sundaram, B., Kishimoto, Y., Takuma, H., and Tajima, T. (1998). Phys. Rev.
A 58, 4757.
Ramsey, N.E (1956). "Molecular beams." Oxford University Press, London.
Rom-Kedar, V., and Turaev, D. (1999). Physica D 130, 187.
Romalis, M.V., and Fortson, E.N. (1999). Phys. Rev. A 59, 4547.
Rudy, P., Ejnisman, R., Rahman, A., Lee, S., and Bigelow, N.P. (2001). Opt. Express 8, 159.
Ruschewitz, E, Peng, J.L., Hinderthfir, H., Schaffrath, N., Sengstock, K., and Ertmer, W. (1998).
Phys. Rev. Lett. 80, 3173.
Sachrajda, A.S., Ketzmerick, R., Gould, C., Feng, Y., Kelly, P.J., Delage, A., and Wasilewski, Z.
(1998). Phys. Rev. Lett. 80, 1948.
Sasaki, K., Koshioka, M., Misawa, H., Kitamura, H., and Masuhara, H. (1992). Appl. Phys. Lett.
60, 807.
Savard, T.A., O'Hara, K.M., and Thomas, J.E. (1997). Phys. Rev. A 56, R1095.
Schiffer, M., Rauner, M., Kuppens, S., Zinner, M., Sengstock, K., and Ertmer, W. (1998). Appl.
Phys. B 67, 705.
Seideman, T. (1999). J. Chem. Phys. 111, 4397.
Shabtay, G., Zalevsky, Z., Levy, U., and Mendlovic, D. (2000). Opt. Lett. 25, 363.
Simsarian, J.E., Ghosh, A., Gwinner, G., Orozco, L.A., Sprouse, G.D., and Voytas, P.A. (1996).
Phys. Rev. Lett. 76, 3522.
S6ding, J., Grimm, R., and Ovchinnikov, Y.B. (1995). Opt. Commun. 119, 652.
S6ding, J., Gu6ry-Odelin, D., Desbiolles, P., Ferrari, G., and Dalibard, J. (1998). Phys. Rev. Lett.
80, 1869.
Song, Y., Milam, D., and Hill, W.T. (1999). Opt. Lett. 24, 1805.
Spektor, B., Piestun, R., and Shamir, J. (1996). Opt. Lett. 21,456.
VII] DARK OPTICAL TRAPS FOR COLD ATOMS 151

Spreeuw, R.J.C., Voigt, D., Wolschrijn, B.T., and van den Heuvel, H.B.V. (2000). Phys. Rev. A
61, 053604.
Sprouse, G.D., Orozco, L.A., Simsarian, J.E., and Zhao, W.Z. (1998). Nucl. Phys. A 639, 316c.
Stenger, J., Inouye, S., Stamper-Kurn, D.M., Miesner, H.-J., Chikkatur, A.P., and Ketterle, W. (1998).
Nature 396, 345.
Suominen, K.A. (1996). J. Phys. B 29, 5981.
Torii, Y., Shiokawa, N., Hirano, T., Kuga, T., Shimizu, Y., and Sasada, H. (1998). Eur. Phys. J. D
1,239.
Vivaldi, E, Casati, G., and Guarneri, I. (1983). Phys. Rev. Lett. 51,727.
Vuleti6, V., Chin, C., Kerman, A.J., and Chu, S. (1998). Phys. Rev. Lett. 81, 5768.
Webster, S.A., Hechenblaikner, G., Hopkins, S.A., Arlt, J., and Foot, C.J. (2000). J. Phys. B 33, 4149.
Weiner, J., Bagnato, V.S., Zilio, S., and Julienne, P.S. (1999). Rev. Mod. Phys. 71, 1.
Winoto, S.L., DePue, M.T., Bramall, N.E., and Weiss, D.S. (1999). Phys. Rev. A 59, R19.
Wolf, S., Oliver, S.J., and Weiss, D.S. (2000). Phys. Rev. Lett. 85, 4249.
Xu, X., Kim, K., Jhe, W., and Kwon, N. (2001). Phys. Rev. A 63, 063401.
Xu, X.Y., Minogin, V.G., Lee, K., Wang, Y.Z., and Jhe, W.H. (1999). Phys. Rev. A 60, 4796.
Yan, M., Yin, J.E, and Zhu, Y.E (2000). J. Opt. Soc. Am. B 17, 1817.
Yin, J., Zhu, Y., and Wang, Y. (1998). Phys. Lett. A 248, 309.
Yin, J.P., Noh, H.R., Lee, K.I., Kim, K.H., Wang, Y.Z., and Jhe, W. (1997). Opt. Commun. 138, 287.
Zaslavski, G.M. (1999). Phys. Today 52(8), 39.
Zemanek, P., and Foot, C.J. (1998). Opt. Commun. 146, 119.
This Page Intentionally Left Blank
A D V A N C E S IN ATOMIC, M O L E C U L A R , A N D O P T I C A L PHYSICS, VOL. 48

MANIPULATION OF COLD A TOMS IN


HOLLOW LASER BEAMS
HEUNG-RYOUL NOH, XINYE XU* and WONHO JHE**
School of Physics and Center for Near-field Atom-photon Technology, Seoul National University,
Seoul 151-742, South Korea

I. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
II. Theoretical Models for Cold Atoms in Hollow Laser B e a m s . . . . . . . . . . . . . . . . 154
A. Strict Kinetic Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
B. Dressed-atom Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
III. Generation Methods for Hollow Laser Beams . . . . . . . . . . . . . . . . . . . . . . . . . . 160
A. Mode-Conversion Method with Cylindrical Lens . . . . . . . . . . . . . . . . . . . . . 160
B. Computer-Generated Hologram Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
C. Spiral Phase-plate Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
D. Geometric Optics Method with Axicons . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
E. Micro-imaging Method for Hollow Fiber Modes . . . . . . . . . . . . . . . . . . . . . . 166
E Near-field Diffraction Method for Hollow Fiber Modes . . . . . . . . . . . . . . . . . 166
IV. Cold Atom Manipulation in Hollow Laser Beams . . . . . . . . . . . . . . . . . . . . . . . 170
A. Atomic Guidance in Hollow Laser Beams . . . . . . . . . . . . . . . . . . . . . . . . . . 170
B. Atomic Fountain with Hollow Laser Beams . . . . . . . . . . . . . . . . . . . . . . . . . 176
C. Atomic Traps with Hollow Laser Beams . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
V. Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
VI. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

I. Introduction

Atom optics has become an active and attractive research field, and numerous
novel atom-optical components that use optical or magnetic methods have been
developed [1-3]. Although magnetic atom optics is a promising approach to
realize coherent atom optics or miniaturized atom-optical elements, atom optics
utilizing optical schemes also provides unique and versatile tools for such stud-
ies. In particular, optical atom optics becomes more powerful when combined
with microscopic atom-optical elements on the surface or even with magnetic
atom-optical techniques.
Cold atoms have been manipulated by optical dipole potentials produced by a
hollow-core optical fiber (HOF)[4-7] and a hollow laser beam (HLB)[8-12].

* Present address: JILA, University of Colorado, Boulder, CO.


** Corresponding author: whjhe@snu.ac.kr

153 Copyright 9 2002 Elsevier Science (USA)


All rights reserved
ISBN 0-12-003848-X/ISSN 1049-250X/02 $35.00
154 H.-R. Noh et al. [II

Guidance of atoms by HLB, in particular, has advantages over that by the


evanescent waves in HOF, since the van der Waals attraction due to the fiber
walls can be ignored and collisions with background gas are much less probable.
Moreover, an HLB configuration can be controlled more conveniently than one
with HOE Therefore, the development of efficient optical manipulation of cold
atoms is of much interest and importance in applications related to the transfer
of trapped cold atoms to a region of dense atoms or to a lower-dimensional
space, which can be used for such experiments as high-resolution spectroscopy,
atom lithography, atom microscopy, atomic fountains, and cold atoms in confined
space. In this review, we present quantitative experimental and theoretical studies
associated with optical manipulation of trapped cold atoms by HLB.

II. Theoretical M o d e l s for Cold A t o m s in H o l l o w L a s e r B e a m s

In this section, we present two theoretical models describing the motion of


cold atoms in laser light, in particular in HLB [13,14]. The optical dipole
force, the radiation-pressure force, and the diffusion coefficients are derived. We
assume a three-level A-atom that consists of two hyperfine-structure ground-
states (]g,) = I1) and Ig2) = 12)) and one excited state (]e) = ]3)). The atom is
assumed to interact with a single-mode traveling laser beam of frequency coL.
The detunings of the laser wave with respect to the atomic resonant transitions
are 6/ = col-coj (j = 1,2), where col (092) is the atomic transition frequency
between the ground state I1) (]2)) and the upper state ]3). The hyperfine-structure
splitting between two ground-state levels is denoted by 6hfs = co~ --092. The laser
field is chosen as a spatially inhomogeneous monochromatic laser wave given

by E = E0 cos (k- r - col t), ( 1)

where E0 = E0(r) is the coordinate-dependent amplitude of the laser field and


k = coL/c is the magnitude of the wave vector.

A. STRICT KINETIC THEORY

In this interaction scheme, the atomic Hamiltonian can be expressed by

]/2 V 2
H = Ho - ~ - D . E, (2)

where Ho describes the atomic energy levels E 1 , E 2 , E 3 , and the last term
describes the dipole interaction between the atom and the laser field. The
complete description of the time evolution of an atom can be then given by
the equations for the atomic density matrix in the Wigner representation. The
II] MANIPULATION OF COLD ATOMS IN HLBs 155

dipole interaction is considered as usual in the rotating-wave approximation [ 15].


To simplify the derivation of the equations of motion, we first write down the
initial microscopic density-matrix equations neglecting the spatial variation of
the laser beam amplitude E0:

d /933 = i/r
( ei(k'r-6't) P(-)13 -- e-i(k'r-'5~t) P(-)'~31)
+iK'2 (e i(k'r-6zt) p~;) --e -i(k'r-62t) p(-))
32
- 2( Yl + Y2)P33

-dt- P22 = itr e i(k r eat) P32 ..(+) + 2 72


..(+) _ ei(k.r-(ht) /d23 ~(n) p(n)
33 d2n,

dld [ l l --iK'l
(e_i(k.r_(~,t)
p~l )-
ei(k.r_61t )
pi3 )) + 2y,
/ *(n)
p(n)
33
dan, (3)
d = iKl ei(k.r_6~ t ) ( p l l ) - p(+))
-dt/331
- 33 + itr e i(k'r-6zt) P~l)--( Y1 + 72)P31,
d
dt P32 = iir ei(k .r-,~t)( P~2) _ p(+))
33 + iK'l
ei(k.r-O,t) p(-)
12 - - ( Yl + Y2)P32,

d Pi2 = i1r e -i(k" r-r t)p(+)


dt 32 _ iic2ei(k.r-r

Here the density matrix elements are defined with respect to the time-
dependent, stationary atomic eigenfunctions such that

pr = (a IP ( r , p + 89 t)I b), pI~) = (a p (r, p + n h k , t) I b>,

where (a,b) - (1,2,3) and n is a unit vector that defines the direction of the
spontaneous photon emission. The halves of the Rabi frequencies, ~ , and the
partial spontaneous decay rates, )9, are defined by (j = 1,2)

_ f2j dj3" Eo 2~. = W J p - 4 ~23 m 3


2- 2h ' 3 hc 3" (4)

In Eqs. (3), the function ~(n) describes the angular anisotropy of the spon-
taneously emitted photons. In our simplified model that neglects the atomic-
level degeneracy, the function ~(n) can be chosen to be isotropic such that
9 ( n ) - 1/4sz and f ~(n)dZn - 1. Note that the microscopic equations for the
considered model scheme do not include the integral term for the ground-state
coherence Pl2.
Assuming the interaction time exceeds the spontaneous decay time (tint >>
l"sp = 1/Wsp), one can expand the density-matrix elements in powers of photon
momentum hk. Moreover, when Tint >> Tsp , the Wigner density-matrix elements
become the functionals of the Wigner distribution function, w = w(r, p , t ) =
}-~3a .
= l Paa As a result, one can derive the Fokker-Planck kinetic equation for
156 H.-R. Noh et al. [II

the Wigner distribution function w(r, p,t), to the second order in the photon
momentum hk,

_ 02
0W (914' 0 (Fw)+ E ~(Diiw). (5)
Ot + v Or Op Opi

For a three-level atom, the Rabi frequencies are f2j = r v/[/(2[s)s 1/2 (j = 1,2),
where F = 27' = 2(7'1 + 7'2), I is the intensity of the laser beam, Is = ,rrhcF/(3A 3)
is the saturation intensity, fj. = @/(3areohc3F)l(eldj .eLlgj)l 2 is the relative
transition strength, ~. is the resonant transition frequency from le) to ]gj), and
eL is the polarization unit vector of the electric field EL. Note that the saturation
parameters can be written as G1 = j i G and G2 = J)G, where the reduced
saturation parameter are given by G = I/Is. Consequently, for large positive
detunings and slowly moving atoms (]k. v I << 61,62), the dipole force, the
spontaneous force, and their respective diffusion tensors can be obtained as the
following simple expressions [ 13]:

h VG
Fd = - ~ (ql ~1 + q262) --~-a ' (6)

Fr = hi{l-"Co' (7)
1 G
Dii = -i-~h2k2Fc0 , (8)

gd = hF 2 (ql ~1 -t- q262) e l ' (9)


8

where Co = ql/fl + qz/f2 + 3G/2 + 4Cl/l -'2, Cl = qlOZ/fl + q262/f2, the optical
potential is obtained by Ua = - f Fa. dr, and qj = ))/(Yl + 7'2) defines the relative
spontaneous-emission rates for two photon-emission channels with j = 1,2.
The above general relations can now be applied directly to our problem of cold
atoms guided in a hollow laser beam. For an HLB, the dipole gradient force in
Eq. (6) pushes atoms to the central dark hollow region at blue laser detunings. We
assume that HLB is linearly polarized and described by the intensity distribution

I=I(p,z) 2 W2 exp ( ) -2p2


-4Ppz315w --~- , w = wo v / l + ~ ( z / z R ) 2, zR=jrw2/~,
(10)
where P is the laser power, w = w(z) and w0 are the beam waists at distance z
and at z = O, respectively, and zR is the Rayleigh length. The parameters w0 and
can be obtained by comparing the measured intensity distribution with Eq. (10).
Note that at a given coordinate z, the radial intensity distribution has a maximum
II] M A N I P U L A T I O N OF C O L D ATOMS IN HLBs 157

Table I
The mean relative transition strengths, J), j = 1,2, o f several alkali atoms.

Line ft. 23Na(87Rb, 39K) 85Rb 133Cs

2 2 2
D2-1ine 3q j j j
or a + or a - j~
2j 2j 2
j
1 1 1
D 1-line fl j J J
1 1
Jr or a + or o- 3~ j J 31

Table II
Mean branch ratio qi, i = 1,2 from the excited state 13) decay to the ground state li) for the excitation
o f the ground state lJ) by the H L B , j = 1,2.

Line Initial State IFj) qi 23Na(87Rb, 39K) 8SRb 133Cs

13 20 3
D2-1ine ~ = FI qI T8 ~
5 7 l
q2 l--g ~ 2[
1 5 7
Fj=F2 ql $ ~ 3-6
5 22 29
q2 ~ 2--7 3-6
Dl-line Fj = F l ql ~4 ~13 I
5 14 1
q2 ~ ~-7
1 10 7
/~ = F2 ql ~ ~ 1--8
2 17 11
q2 j 2--7 1-8

value of Im = 2 P / ~ e w 2 at the distance Pm = w / v / ~ . For Eq. (10), the reduced


saturation parameter becomes
I p2 (2p2) 4P
G = ~ = Go ~-5 exp - - - ~ , Go - ~ w 2 i S . (11)

For numerical calculation of atomic dynamics, the values o f f for several alkali
atoms are listed in table I. We find that in the case of the traveling-wave laser
field, the values o f f are independent of the polarization o f the laser field. In the
case of D2-1ine transitions, for example, J] =j~ = 2 for all alkali atoms as listed
in table I. In addition, table II presents the relative spontaneous emission rates,
qj (j = 1,2), from le) to [ g j ) [ 1 4 , 1 6 ] .

B. DRESSED-ATOM MODEL

In this subsection we analyze the atomic interaction in the dressed-atom


representation by using the secular approximation, 6j >> Qj >> F, ( j = 1,2),
158 H.-R. Noh et al. [II

where F is the spontaneously decay rate of the excited state [e) [1 4]. Furthermore,
under the electric-dipole and rotating-wave approximations, the dressed-atom
Hamiltonian at point r within the manifold En can be written as [ 17,1 8]
2
H,(r) = Z {-h(coL - 6j) bjbf - [dj- EL(r)bfaL + dj . t~(r)bja~]} + hcoLa+taL,
j=l
(12)
where the first term describes the atomic energies, the second term is the atomic
dipole interaction Hamiltonian, and the last term represents the laser field. Here
bj = Igj) (el is the lowering operator, bf = ]e) (gj] is the raising operator, dj is the
electric-dipole-moment operator, eL(r) is the laser field, and a~ (aL) is the cre-
ation (annihilation) operator of a photon. In this model, the laser field is also cho-
sen as a spatially inhomogeneous monochromatic laser wave as given in Eq. (1).
In a manifold E~, each eigenstate is given by a linear superposition of three
basis states,
]i(n)) = aTl ]gl,n + 1) +a72 ]g2,n + 1) + a73 ]e,n), (i = 1,2,3), (13)
where ~-~=, lai~ 12 = 1 according to the time-independent Schr6dinger equation,
H~ [i(n))- E,i li(n)). Therefore, the matrix form of H~ becomes

H,, = h 0 n~OL + 62 Q2/2 , (1 4)


Ql/2 ff22/2 n~OL

with s = - 2 v / n + ldj. tc(r)/h. The coefficients in Eq. (13) are then


a~. = b~. ~ 1
v/l + ~--~2=l (bg.) 2' ai3-- V/1 + ~-~2= 1(bg.) 2 , (15)

where
ff2j/2
bij = n~oc + 6j - E n i / h ' (16)
with i = l, 2, 3 and j = 1,2. Consequently, the eigen-energies of Hn,
E~i = nhO)L + Ei, (i = 1,2, 3), are given by

E ~ ~ nh~oL + h6~ + ~ (i = 1,2), (17)

En3 ~ nhOJL- A~." (18)


"T I_Ij
j=l
For the intensity distribution of the laser beam described by Eq. (1 0), the energy
levels of the dressed atoms are shown in Fig. 1 for the case of 6j > 0, which
shows the dressed-level energies, as a function of the position p of the atom,
belonging to the manifolds E, and En-1 separated by the laser energy hCOL. At
II] MANIPULATION OF COLD ATOMS IN HLBs 159

2(n))

En / ~ [gl'n+l)/ ~ l(n))

n~ ' le, n) ' 3(n))

V-
03 L

En_l _/2 , le, n - 1 ) [3(n-l))


(n-l) cot

P Pm o Pm P
FIG. 1. Dressed states for a three-level A-atom interacting with a HLB for positive detunings
(aj, 62 > 0).

the center of HLB, ff2j tends to vanish and thus the atomic levels approach the
uncoupled-state levels.
Taking into account the coupling of the dressed atom with the vacuum-field
reservoir, responsible for spontaneous emission between adjacent manifolds,
one can write a master equation for the density matrix o of the dressed
atom, which describes both the internal free evolution of the dressed atom
and the relaxation due to the atom-vacuum coupling. If we denote three
reduced populations by ;ri(r), corresponding to three dressed states ]i(n)) defined
by ~i(r) = ~-~n (i(n) l~ the evolution of 3-gi is described by
~ = ~-]~i~j(-FijJrj. + Fji:ri) with (i, j) = (1,2,3). Here the rate of transfer Fik
from ]k(n)) to ]i(n - 1)) is given by Fik = ~ - ] 2 Fjaoak3"
2 2 The resulting steady-
state solutions are then calculated as

MS~t = q , f 262
qlf262 + qzf162, (19)

qzflb2 (20)
:r~t = qlf262 + qefl 62,
;r~t = JiJ~GZF4 qlflb2 + qzfzb2 (21)
6462 62 q l j~ 62 + q zj] 62.
160 H.-R. Noh et al. [III

Consequently, the radiation force, the momentum diffusion tensor, and the dipole
force can be obtained as [ 15,17]

Fr=hk --~ i'


.=

Dii : gh
1 2 k 2 rp~', (23)
2
Fa = - Z ygstvui, (24)
i=1

where (dN/dt)i = ~-~2=1 rki"7"gst, D st3 -- ~ - ~ : 1 zcSta2i3' and Ui = fhr2G/(86j)


Note that by explicit substitution of Eqs. (15), (16) and (19)-(21) in Eqs.
(22)-(24), one can recover Eqs. (6)-(8) obtained by the strict kinetic theory. As a
result, both independently derived results can be equally employed in numerical
simulation of atomic dynamics in HLB [ 13,14].

III. Generation Methods for Hollow Laser Beams

A hollow laser beam (HLB) is a laser beam whose intensity along the central
axis vanishes, having a doughnut-shaped intensity distribution. HLBs include
a ring-shaped TEM~)l mode, high-order Laguerre-Gaussian (LG) beams, high-
order Bessel beams, and vortex solitons. Here we review several methods that
have been developed to generate an HLB, such as the vortex grating method [ 19],
transverse-mode selection [20], direct production from a laser [21 ], the optical
holographic method [22], computer-generated holography [23], mode conversion
from Hermite-Gaussian to LG by use of two cylindrical lenses [24,25], spiral
phase-plate methods [26,27], geometrical methods with axicons [11,28] or a
double-cone prism [29], and use of a hollow-core optical fiber (HOF)[30].

A. MODE-CONVERSION METHOD WITH CYLINDRICAL LENS

Both the Hermite-Gaussian (HG) and the Laguerre-Gaussian (LG) modes form
complete sets of solutions to the paraxial wave equation [31]. The rectangularly
symmetric HG modes are described by the product of two independent Hermite
polynomials, describing the field distribution in the x and y directions. They
are characterized by integer subscripts m and n representing the order of the two
polynomials, that is, the number of nodes in the electromagnetic field. In contrast,
the circularly symmetric LG modes are similarly denoted by LGlp, where l is the
number of 2Jr cycles in phase around the the circumference and (p + 1) is the
number of nodes across the radial field distribution.
III] MANIPULATION OF COLD ATOMS IN HLBs 161

Since the LG mode possesses an azimuthal phase dependence of exp(i/q~), it


has a helical wavefront and null intensity along the propagation axis. Therefore,
the LG/ beam has an orbital angular momentum of lh per photon [32] as
well as a spin angular momentum +h (-h) for a o+-polarized (o--polarized)
light beam. The existence of the orbital angular momentum has led to a
number of exciting studies such as transfer of the orbital angular momentum
to macroscopic objects[33-35], second-harmonic generation[36], interaction
with atomic systems [37,38], high-resolution spectroscopy in a magneto-optical
trap [39], and blue-detuned optical dipole traps (see Sect. IV.C, and the chapter
by Friedman et al. in this volume).
The propagation of laser beams with an HG or an LG mode can be described
in the usual language of Gaussian beams. In the vicinity of the beam waist,
a Gaussian beam experiences a phase shift compared to that of a plane wave
of the same frequency. This phase shift lp is called the Gouy phase shift [31]
and is given by ~p(z) = (n + m + 1 ) a r c t a n ( z / z R ) for the HGm,n mode, whereas
it is ~p(z) = (2p + l + 1)arctan(z/zR) for the LGtp mode, where z is the distance
along the axis from the beam waist in each case and zR is the Rayleigh range.
If the Gaussian beam is focused by a cylindrical lens, the situation becomes
more complex since the Rayleigh ranges in the x - z and y - z planes are not
equal, Zp.x ~ ZRy. Such a beam is called an elliptical Gaussian beam, and the
corresponding Gouy phase shift for the HGm,n mode is given by

l(Z)
~p(z) = (m + ~) arctan
ZRx
+(n+ 89
(z)
~
ZRy
. (25)

Note that it is the Gouy phase shift occurring in the presence of a cylindrical
lens that forms the basis of the mode converter.
The generation of an LG beam was first demonstrated by Beijersbergen
et al. [24,25] by transforming an HG mode of arbitrarily high order to an
LG mode. They used a mode converter that consisted of two cylindrical lenses.
Unlike other methods discussed in the remaining part of this section, this method
can produce pure LG modes. Figure 2 shows how the HG~,0 mode rotated at 45 ~
with respect to the x- or y-axis is equivalent to the sum of the HG~,0 and HG0,~
modes, and how these two modes are related to the LG~ mode. Specifically, the
LG~ mode can be formed by a superposition of HG1,0 and HG0,1 modes with a
phase difference of Jr/2.

B. COMPUTER-GENERATEDHOLOGRAM METHOD

One can generate an HLB by using a computer-generated hologram (CGH),


which is created by the interference between an electromagnetic field of interest
and a reference laser beam. An optical holographic method was carried out by
Lee et al. [22] to generate a nondiverging hollow beam, which is similar to a J1
162 H.-R. Noh et al. [III

FIG. 2. Generation of LG1 mode HLB by mode-conversionmethod (from Padgett et al., 1996,
Am. J. Phys. 64(1), 77, Fig. 2, reprinted with permission).

Bessel beam. By using the CGH method, Heckenberg et al. generated a TEM~)1
doughnut mode [40], a TEM~0 doughnut mode [41], and later an LG~ mode [42].
In particular, by using the LG beam, they demonstrated trapping of reflective and
absorptive microscopic particles, which cannot be trapped by using a Gaussian
spot due to the strong repulsive forces.
Paterson and Smith [23] produced high-order Bessel beams by using an
axicon-type CGH, where an azimuthal phase factor, exp(inq)), is added to the
phase of the hologram. The Bessel beam [43] is one of the propagation-invariant
waves and has an amplitude proportional to Jn(por)exp(in(~), where J,, is the
nth-order Bessel function of the first kind, r is the radial coordinate, q~ is the
azimuthal coordinate, and P0 is the radial spatial frequency. The zero-order
Bessel beam has a sharp intensity peak at its center, while higher-order Bessel
waves have zero-intensity minima at their centers. Paterson and Smith calculated
the amplitudes of the waves produced by an axicon-type hologram by using the
Kirchhoff integral, and experimentally demonstrated the production of Bessel
beams of orders 1 and 10.
Clifford et al. [44] generated LG laser modes with an azimuthal mode index l
ranging from 1 to 6 (p = 0) by using an external cavity diode laser. The
transmittance function is given by T(r, 4))= exp[i6H(r, r in polar coordinates,
where 6 is the amplitude of the phase modulation, and the holographic pattern
is given by
H(r, r = ~ mod lq~ - --~r cos 0, 2~ , (26)

with mod (a, b) = a - b int(a/b). As the azimuthal index l increases, the inner
III] MANIPULATION OF COLD ATOMS IN HLBs 163

dark region of the light becomes larger and the outer ring becomes narrower.
The conversion efficiency was as high as 40% and the efficiency was claimed
to increase by using a phase hologram and blazing it to maximize the power
in a chosen diffracted order. In general, when a hologram is irradiated by a
fundamental Gaussian mode, the output becomes a superposition of an infinite
number of LG modes with the same l and different p. The fraction of p = 0 mode
was 78.5% in the first diffracted order. In an analogous method, they also
generated multi-ringed (p > 0) LG modes with azimuthal index l = 1 [45].

C. SPIRAL PHASE-PLATE METHOD

Beijersbergen et al. [26] demonstrated that a spiral phase plate can convert a
Gaussian laser beam into an LG mode with a phase singularity on its axis.
A spiral phase plate is a transparent plate whose thickness increases in proportion
to the azimuthal angle q~ around a point in the middle of the plate. If u(p, q~,z)
is the complex amplitude of the incident beam, the amplitude u I directly after
the plate is given by u' = u exp(-iA/r where Al is the height of the step in
wavelengths given by Al = A n h / X , h is the step height at r = 0, An is the
difference of refractive index between the plate and its surrounding, and X is the
vacuum wavelength.
Beijersbergen et al. chose an acrylic (PMMA, n = 1.49) as a phase plate,
where h = 0.72mm or Al = 577 at 633 nm wavelength. To make Al = 1, the
plate was immersed in a liquid with nearly the same index of refraction. The
effective step size was tuned by controlling the temperature, and they obtained
Al = 1 with An = 8.7x 10-4. They used an LG ~ or LG~ beam as an incident
laser that passed through a phase plate, and the output beam was imaged by a
lens in the focal plane. For each incident beam, the output beam was obtained
with various values of Al ranging from -1 to 2.5 with a step of 0.5. For the
LG ~ mode, a nearly LG~ beam was obtained ( A / = 1). For the LG 1 mode, which
itself is already a helical mode, a mode similar to the input beam was obtained
when Al = 2.
Turnbull et al. [27] generated free-space LG modes at millimeter-wave
frequencies (-100 GHz) by using a spiral phase plate. Due to the large frequency
difference of-104 with respect to the optical field, the orbital angular momentum
is also -104 times smaller. The phase plate was made of polyethylene, which has
a refractive index of 1.52 at millimeter-wave frequencies. They could generate
LG 1 and LG 2 modes with phase plates of step height 6.7 mm and 13.4 mm,
respectively.

D. GEOMETRIC OPTICS METHOD WITH AXICONS

Since the time Herman and Wiggins [46] used an axicon [47] to produce a
164 H.-R. Noh et al. [III

FIG. 3. Generation of HLB by axicons (from Manek et al., 1998, Opt. Commun. 14"/, 67, Fig. l,
reprinted with permission).

propagation-invariant zero-order Bessel beam, hollow laser beams have been


produced by axicons [11,28] or a double-cone prism [29].
Manek et al. [28] generated an HLB for atom trapping by using an axicon in
combination with a spherical lens (Fig. 3). They used this method in a recent
demonstration of a gravito-optical surface trap for Cs atoms that was based
on evanescent-wave cooling (See Sect. IV.C). The axicon has one flat and one
conical surface (base angle 0 ~ 10mrad), and is used in combination with a
spherical lens (achromatic doublet, f = 100 mm). For small base angles 0 << 1,
the ring diameter is given by d = 2 0 ( n - 1)D ,,~ 800 ~tm, where n = 1.51 is the
refractive index of the glass substrate and D = 85 mm is the distance between
the axicon and the focal plane. The diameter of the ring in the focal plane can
be changed easily by varying D, i.e. by simply moving the axicon along the
optical axis. While the diameter of the incoming laser beam does not affect the
resulting ring diameter, it does determine the 1/e-width of the ring because of
diffraction [48].
Manek et al. produced a one-to-one image of the ring-shaped beam profile
with a second achromatic doublet ( f ' = 150 mm). In order to clean the dark inner
region, the dark spot (an opaque disk of 725 ~tm diameter) was also introduced.
The beam profile of the hollow beam in the image plane was recorded by a
CCD camera. They observed the conical propagation of the hollow beam with
a divergence of ~50 mrad, which was comparable to the propagation of a usual
Gaussian laser beam with the corresponding waist in the focal plane. With
III] MANIPULATION OF COLD ATOMS IN HLBs 165

FIG. 4. Generation of HLB by a double-cone prism.

the dark spot inserted, the intensity at the center was ~0.1% of the maximum
intensity of the ring.
Song et al. [11] used a series of axicons (each with a 3 ~ base angle) and
a simple lens. A hollow beam generated thereby has different focal points for
the inner and outer walls. Therefore, two additional axicons were required to
control the core diameter and wall thickness. An important byproduct of this
arrangement was that most of the diffraction originating from the tip of the first
axicon was located outside the core. Since the LG beams are not generated by
axicons, this HLB generally does not propagate indefinitely, nor is the dark core
preserved when the beam is focused by a simple lens. Nevertheless, a HLB that
is usable for tens of centimeters can be produced routinely, which is useful for
guiding atoms.
A double-cone prism was also used for the generation of HLB by Ito
et al. [29]. Figure 4 explains the conversion mechanism by a double-cone prism.
A Gaussian beam is divided into two parts by the first refraction at the apex of the
prism, and then a nondivergent doughnut-shaped hollow beam appears from the
other side after the second refraction. When a prism with a length of 4.3 mm
and a full apex angle of 90 ~ is used, the inner and outer diameters of the
doughnut-shaped light beam are 0.6 and 1.4 mm at e -1 intensity, respectively.
Although the characteristic feature is not better than can be produced with
multiple axicons, this method is very simple and convenient and is useful, for
example, for generating evanescent waves at the conical hollow prism.
An axicon was also employed for the transformation of an LG beam to
a high-order Bessel beam by Arlt and Dholakia [49]. If an LG mode with
azimuthal mode index l is used to illuminate an axicon placed at its beam
waist, an approximation to a Bessel beam of order I is generated (Fig. 5). First,
they obtained LG modes by the computer-generated hologram method [42]. The
LG beam had a waist of w0 = 2.5 mm and the axicon was positioned at its
beam waist. They generated Bessel beams with orders l = 1 to 4. The radius of
the inner ring of the generated first-order Bessel beam was only rm = 21.2 ~tm,
and it propagated about Zma• = 29 cm without any spread. This should be
compared with an LG beam with I = 1 and the same ring size at its waist, which
would have a Rayleigh range of only about 4 mm. The conversion efficiency was
almost 100%, limited only by the CGHs used to produce the LG beams.
166 H.-R. Noh et al. [III

FIG. 5. Generation of a high-order Bessel beam within the shaded region by illuminating an
axicon with an LG mode (from Arlt et al., 2000, Opt. Commun. 177, 297, Fig. 1, reprinted with
permission).

E. MICRO-IMAGING METHOD FOR HOLLOW FIBER MODES

A hollow-core optical fiber (HOF) has many interesting applications in sensors


[50], harmonic generation [51 ], and in particular, atom optics [4,6,52]. Yin et al.
[30] obtained an HLB by using a micro-collimation technique for the output
beam of a micron-sized hollow optical fiber. The principle of this method is
very simple: for a fiber waveguide consisting of a hollow cylindrical core, some
low-order modes can be guided in the hollow core, such as the LP01, LPll, LP21
and LP31 modes [53]. Therefore, when one uses a microscope objective with a
short focal length to image the output intensity distribution at the facet of a
hollow fiber, a simple HLB can be obtained.
The inner and outer diameters of the hollow core of the fiber were 7 ~tm
and 14.6 ~tm, respectively, and the outer diameter of the cladding of the fiber
was 123.4~tm. The relative refractive index difference, An = (n 2 - n Z ) / Z n 2,
was 0.0018 and n2 - 1.45, where nl and n2 are the refractive index of the core
and the cladding, respectively. The numerical aperture is about 0.124. Figure 6
shows the relationship between the dark spot size (DSS) and the propagation
distance Z of the dark HLB. It can be observed that (i) the DSS of the dark
hollow beam collimated by a M-20• objective is about 50 ~tm at Z = 100 mm
and about 100 ~tm at Z = 500 mm, and (ii) the relative divergent angle in the near
field of HOF is about 6.5 • 10-5, whereas the divergent angle in the far field is
4.0 x 10-4. If one uses an HOF having a slightly larger hollow core, an HLB with
a smaller DSS and better propagation invariance may be obtained.

F. NEAR-fiELD DIFFRACTION METHOD FOR HOLLOW FIBER MODES

Instead of using an imaging lens, an HLB can be obtained by superposing two


orthogonal LPll guided modes, which have a node line. Let us first describe the
characteristics of near fields diffracted by an HOF and the resulting generation of
an HLB [54,55]. In the weakly guiding approximation, the excited mode is well
III] MANIPULATION OF COLD ATOMS IN HLBs 167
eoo
7o0
(a)
E= L S O O M - 2 0 X Objective
~4oo

J
0')
(/) soo
t'~ 2o0
100
0 ~,,. t .... t .... i, ~., 1 .... I , ,

~
0 2oo 4oo 000 Boo 1000

Z(mm)

700 (b)
~eoo
E=1 r~o~ M - 4 0 X Objective

CO
IZI z~0
lO0

0 100 2o0 moo 400 500

Z(mm)

FIG. 6. The relationship between the DSS and the propagation distance Z of the HLB measured
with (a) M-20• and (b) M-40• objective lens.

described by the dominant transverse component, the so-called linear-polarized


LPlm mode, where l (m) is the azimuthal (radial) mode number [56]. For angular
frequency r wave number k, and propagation constant fi, a guided mode in a
HOF with hollow diameter 2a and core thickness d = b - a is given in cylindrical
coordinates by its transverse component [52,53],

I Al/(ur) sin[(10 + 0)] (r < a),


Et(r, O) = (BJ/(ur) + CN/(ur))sin[(10 + 0)] (a ~< r ~< b), (27)
DKl(wr) sin[(/0 + 0)] (r > b),

where u = V/fi 2 - k 2, u = v/k2n, 2 - fi2, and w = v/fi 2 - kZn22. Here Jl and Nl


(It and Kl) are the (modified) Bessel functions of the first and the second kind
of order l, respectively, nl (n2) is the refractive index of the core (cladding),
and q) is a phase constant. All the coefficients and fl can be determined by the
boundary conditions at a and b, and the solutions are then given as the LPtm
modes according to the number of radial nodes.
Since one knows the electric fields (Eq. 27) on the facet of the HOF (z = 0),
one can calculate the diffraction pattern at (x,y,z) using the Huygens-Fresnel
integral [57]:

z 1 exp(ikp)
E(x,y,z) = - ~ ~ ~ E~ P - ik p2 clxo dyo, (28)
168 H.-R. Noh et al. [III

(~tm) J~/ (~trn)~//


O, ,-11, ,

10 0 r (~tm) 10 0 r (~tm)

(a) LPol (b) LPll


FIG. 7. Development of the radial intensity distributions due to the diffraction of (a) LP01 mode
and (b) LP11 mode near the facet of HOF (z = 0 at the facet).

with (x0,Y0) the coordinate of a source point and p = v/z 2 + (x--X0) 2 + ( y - y 0 ) 2.


For a given LPlm mode represented by El~ at z = 0, one can obtain

Elm(r, O,z) -- 2~r(-i) t sin(10 + q~0)


(29)
x U~~ - X2~z]Jt(2~r)~d~,

where (r, 0) are the cylindrical coordinates of (x, y). Here U)~ that is, the Fourier
transform of Et~ can be obtained analytically as described in Ref. [54].
Figure 7a explains how the LP01 mode diffracts in free space near the HOF: the
two peaks at z - 0, representing a cross section of ring-shaped mode, diminish
away while an additional central peak grows. In Fig. 7b one can find that the
peaks of LPll also diminish whereas another pair of peaks grow. In this case,
however, there still does exist a dark column along the central axis. A doughnut-
shaped HLB can then be produced as follows. For a given propagation constant,
there are four degenerate LPll modes, whose respective polarizations and angular
variations are described by ~ sin(0 + q~),)3 cos(0 + q~),~ cos(0 + r and)3 sin(0 + q~)
[58]. One can easily find that superposition of the first two or the last two modes
can produce an azimuthally symmetric mode. The output beam then forms a
doughnut-shaped HLB since it is just a superposition of the output fields of two
orthogonal LPll modes. The resulting combined beam in front of the HOF may
look like a single linear-polarized beam with its plane of polarization rotated
by 45 ~ with respect to the horizontal plane. One should note, however, that each
beam can be adjusted separately, which is important in exciting modes that differ
from one another.
The first two images in Fig. 8 represent the independent mode patterns of
two perpendicular modes at z = 0 before they are merged, and the last image
III] MANIPULATION OF COLD ATOMS IN HLBs 169

Fie. 8. Superposition of two orthogonal LPll modes to obtain HLB. Transverse intensity profiles
at z = 0 are shown: (a) and (b) present the images before superposition, (c) shows the superposed
image. Polarization and angular variation of the corresponding electric field can be described by, for
example, (a) ~ sin(0 + r and (b) ~ cos(0 + r

Fl~. 9. Characteristic dimensions of the diffraction-limited dark hollow spot, measured in terms
of the dark-spot radius Rmax and the halfwidth w, which is in good agreement with the numerical
simulation.

shows their combined pattern, which is similar to that of the LP01 mode. At a
distance z = 250 ~m, the peak-to-peak distance is about 17 ~tm and the dark spot
size is about 8.2 ~tm. Its azimuthal isotropy was checked by measuring the beam
profiles along eight different radial axes and they showed good uniformity within
the maximum uncertainty of about 7%.
In a later experiment, the same group [59] also obtained the diffraction-limited
dark spot near the HOF facet. Figure 9 shows the development of the radial
intensity and the size of the dark hollow region, which is equivalent to the
calculations presented in Fig. 7b. The minimum size of the dark spot, Rmax, is
about 2 txm, the halfwidth of the dark spot, w, is under 1 ~tm, and the inner peaks
are diverging with a 40-mrad diffraction angle. The experimental results were
in good agreement with the numerical calculations obtained by the Rayleigh-
Sommerfeld theory and the weak guiding approximation. The small dark spot
may be applicable as an atomic lens which focuses atoms to a small spot or as
an optical dipole-force microtrap by combining several beams.
170 H.-R. Noh et al. [IV

IV. Cold Atom Manipulation in Hollow Laser Beams

A. ATOMIC GUIDANCE IN HOLLOW LASER BEAMS

A.1. HLB guidance of Rb


Xu et al. [12] performed optical guiding of trapped cold atoms by a hollow
laser beam (HLB) produced by micro-collimation and micro-imaging technique
as discussed in Sect. III.E. The atomic guiding direction was downward along
the gravity (+z direction), whereas the HLB propagated along the - z direction
(counterpropagating scheme) or along the +z direction (copropagating scheme).
A Ti:sapphire laser was used as the guiding laser source with a maximum output
power of 1.8 W. It was coupled to the core of the HOF with a coupling efficiency
of about 30%. The typical HLB power used for guiding atoms was 250 mW. They
used a micron-sized HOF with a hollow diameter of 4 Ftm, core thickness 2 Ixm,
and length 25 cm. In both guiding schemes they obtained identical values for
the radius of the maximum-intensity ring pm(z) that varies linearly with the
distance z [pm(Z) -- pro(0)- aZ, where pm(0) = 1.4 mm is the value at the trap
center (z - 0) and a = 1.27(4)x 10-3].
They used a standard vapor-cell magneto-optical trap (MOT) of 85Rb atoms.
The number of trapped atoms was typically 2x 107, and the trap diameter
was about l mm so that the loading efficiency of the atoms trapped into the
HLB was 98%. By time-of-flight measurement, the temperature of atoms in
the MOT was found to be about 140FtK, which was further cooled down to
16 txK by polarization-gradient cooling. After sub-Doppler cooling, the cooling
and repumping lasers were blocked by mechanical shutters, and the HLB was
simultaneously introduced to the atoms to guide their gravitational falling. The
number and the temperature of guided atoms were detected by observing the
probe-induced fluorescence with a photomultiplier. The probe laser beam was
placed horizontally at 105 mm below the trap center.
Figure 10 shows time-of-flight signals of guided cold atoms in both guiding

.~ 2.5] (a) 16GHz--j ~m~ 2.5 (b)


2.0 t 0OHz 2.0

0.5 ree ~ 0.5


0 ,~
"E00 0;05 0;10 0;15 0"8 0 0.~12 0.~14 0.~160.~180.20 0.22. 0J24
0 "~ . ~ - - " > " - . -"~. -~.- 9. " ; - " ~ ---=- . . -.~-,e-----. --

-time of Flight (s) Time of Flight (s)


FIG. 10. Typical TOF signals of atoms guided by a single HLB. (a) In the copropagating scheme,
the laser detuning 62 is 1, 2, 6, 10 and 16 GHz, respectively. (b) In the counterpropagating scheme,
62 is 6, 10 and 16 GHz, respectively.
IV] MANIPULATION OF COLD ATOMS IN HLBs 171

0- 9 Copropagating
~..~ Counterpropagating
"~ 40-
30,

20,
.-~ 10.
(-9 O
-4-~_ o 2 ~, 6 8 io i2 i4 i6
Detuning~i2(GHz)
FIc. 11. Guiding efficiency as a function of the detuning 62 in the (a) copropagating and (b)
counterpropagating schemes. The solid curves represent numerical simulation results.

schemes at various laser detunings with respect to the 5S1/2,F - 2 ~ 5P3/2


transition line. For comparison, the detected signal without the HLB for the freely
falling atoms is also shown. In particular, it is observed that the number of atoms
guided by the copropagating HLB is about 20-fold enhanced with respect to
that without the HLB at 2 GHz detuning. In this case, the guided atoms also
become accelerated along the +z direction due to the increased radiation pressure
at small detunings (Fig. 10a). In the counterpropagating case, on the other hand,
the guided atoms are decelerated as the detuning is decreased (Fig. 10b).
Figure 11 presents experimental and numerical guiding efficiencies versus
detuning in the copropagating (a) as well as in the counterpropagating (b)
scheme. Note that in numerical simulation the HLB was assumed, to a good
approximation, as the lowest Laguerre-Gaussian (LG~) mode given by Eq. (10)
in Sect. II.A. It can be observed that at small detuning, atoms are most
efficiently guided in the copropagating scheme (for example, the maximum
guiding efficiency is about 50% at a detuning of 2GHz). Note that when
atoms are released from the molasses without HLB, only 2.5% of the initial
trapped atoms are detected. On the other hand, the counterpropagating guiding
is generally less efficient as seen in Fig. 11. However, for large detunings, both
schemes provide similar guiding efficiencies and the maximum efficiency of 23%
is obtained at 10 GHz detuning in the counterpropagating scheme.
Note that the efficiency of the copropagating guiding can be further enhanced
with a higher laser power at a given atomic temperature (e.g., 60% with 500 mW
at 16 9K) or with a lower atomic temperature at a given power (e.g., 75% with
1 ~tK at 250mW). Moreover, if the quality of the HLB is improved by more
careful fiber treatment and optical alignment to excite the LG~ mode, the guiding
efficiency is also expected to be much enhanced. Xu et al. find that the maximum
guiding efficiency can be 80% (50%) in the copropagating (counterpropagating)
scheme. In particular, the guiding efficiency at large detunings in both schemes
can be increased twice. They also find that if the coherence of guided atoms
such as Bose condensate is to be preserved, both schemes may be exploited
172 H.-R. Noh et al. [IV

with higher power, larger detuning, and colder atoms. For example, simulation
shows that when the HLB has 500 mW power and 220 GHz detuning, more than
30% of the atoms at 2 ~tK temperature can be guided over a distance of 30 cm
without any spontaneous emission and with an average photon scattering rate
much less than 1 s-1.

A.2. HLB guidance o f Ne* and Rb condensate

Schiffer et al. [ 10,60] have demonstrated guiding and focusing of atoms in the
dark region of a holographically generated HLB. They performed experiments
with metastable neon atoms. A laser-decelerated and compressed atomic beam,
having a high brightness of 5• 1012/sr/s [61], is injected into the dark region
of a blue-detuned doughnut mode. The longitudinal velocity is 28 rn/s with a
rms width of 4 m/s. The atoms are injected into the doughnut mode through
a 30-mm-radius hole in a dielectric mirror at 45 ~ Behind the mirror, the flux
is 1.4 • 10 6 atoms per second with a transverse Gaussian velocity distribution,
having the spread a. = 7.8 cm/s=2.5Vrec, where Ure c = hk/m is the photon
recoil velocity. They used a computer-designed blazed phase hologram that was
produced by a direct laser-writing technique [62].
Figure 12 shows the normalized atomic intensity in the focal plane as a
function of the doughnut-mode well depth for a waist w0 of 50 ~tm and a power
of 300mW. For a shallow potential, only a minor part of the atomic beam
was captured by the guiding potential when entering the doughnut mode. By
increasing potential height, however, a growing number of atoms was trapped
and guided. The experimental curve in Fig. 12 provides slightly lower guiding

1/ [1/OHz]
0.00 0.02 0.04 0.06 0.08
,,
| i !

I-, I
2000

1500
0 50 I00
o
N
I000

Q 500
z

:-AO ...... 51 ......... i ..... 110 i,.

Urnax[ 103Eree]

FIG. 12. The dependence of the atomic intensity enhancement on the maximum light shift for
atom guiding in a blue-detuned doughnut-mode HLB (open circles) and in a red-detuned Gaussian
mode (solid circles). The dashed and solid curves present the results of the numerical simulation.
The plot for a wider range is shown in the inset (from Schiffer et al., 1998, Appl. Phys. B 67, 705,
Fig. 2, reprinted with permission).
IV] MANIPULATION OF COLD ATOMS IN HLBs 173

Fie. 13. Atoms loaded into a HLB waveguide from a Rb BEC with different evolution times
inside the waveguide (from Bongs et al., 2001, Phys. Rev. A 63(3), 031602, Fig. 4, reprinted with
permission).

efficiency compared to the simulation result. For Umo~ - 8500Eree, the atomic
intensity is enhanced by a factor of 1600. In this case, (60+ 10)% of all atoms
injected into the doughnut mode are captured and guided. The inset of Fig. 12
shows that for high Umo~ the enhancement factor is expected to reach values
well above 3000. For Uma~ = 2000Erec, the flux is 8 • 105 atoms/s, the density is
2 • 108 cm -3, and the intensity is 6 • 1011 (cm -2 s -1). In particular, they achieved
a factor of 10 improvement in atomic flux by using the doughnut mode, and a
factor of 80 with a TEM~)5 mode.
Recently, the transfer of Bose-Einstein condensate (BEC) into a quasi-lD
waveguide created by a blue-detuned HLB was also demonstrated by the same
group [63]. The combined optical dipole and magnetic (DM) trap consists of a
waveguide added to the 3D potential of a Ioffe-type magnetic trap. It allows for a
natural connection between the magnetic trap and a pure 1D waveguide geometry
created by an LG (TEMPI) laser beam. With a power of P = 1 W at 532 nm
and a beam waist of r0 ~ 10 ~tm, a dipole potential at the focal plane with a
maximum value of ~120~tK and a transverse oscillation frequency of ~6 kHz
(corresponding to 570nK) for 87Rb atoms can be realized. They investigated
the transfer process of BEC into a blue-detuned dipole waveguide and studied
the subsequent evolution of the ensemble in a quasi-lD waveguide. Figure 13
shows guiding for evolution times up to 500 ms. On time scales above 40 ms,
the conversion of mean-field energy into kinetic energy is nearly complete and
the ensemble is expected to expand with constant velocity, keeping its parabolic
density distribution. In these experiments they demonstrated that a fully coherent
transfer is possible, and they observed a mean-field-dominated expansion of the
ensemble for adiabatic loading conditions.

A.3. HLB guidance o f Cs


Song et al. [11] demonstrated guiding of ~108 Cs atoms through an 18-cm-
long, 1-mm-diameter core HLB. The generation of HLB used in their guiding is
174 H.-R. Noh et al. [IV

2.0 I ~ ' ' ' I ' '

1.5

x
v
1.0

0.5

o.o . . . . . I ~ c~|
0 50 100 150
UKE (=,)

Fl6. 14. Number of atoms in the HLB tunnel as a function of time, measured from longitudinal
images (squares) with the repumper, those without the repumper (diamonds), and those in the
presence of the repumper but with the kicking beam placed 10 cm below the MOT (triangles) (from
Song et al., 1999, Opt. Lett. 24(24), 1805, Fig. 4, reprinted with permission).

described in Sect. III.D. The source of cold atoms was a vapor-loaded Cs MOT
containing - 1 0 9 atoms. The guide beam and the major axis of the MOT were
aligned and oriented in the vertical direction. The atoms are transferred to the
tunnel by turning the MOT beams off. They then monitored the evolution of the
cloud of atoms (size, location, and density) by the absorption shadow the cloud
of atoms casts on a CCD camera.
They showed that the center of mass of the cloud in the transverse
direction has an acceleration o f - 1 5 m / s 2, appreciably larger than the gravita-
tion g ( ~ 9.8 m/s2). They also found that when the guide beam is directed against
gravity and detuned by less than 1 GHz, it is possible to levitate the center of
mass of the cloud. In the one-dimensional case, in which the atoms are restricted
to radial motion, they find that the average acceleration /asc)ensemble ~ 1/V/A
to first order, in qualitative agreement with the observations (A is the laser
detuning). They also measured the number of atoms in the tunnel for times well
beyond 35 ms by taking images in the longitudinal direction (Fig. 14). With the
r e p u m p e r , - 1 0 s atoms make it to the bottom of the chamber, whereas without
the repumper, they detect no atoms at the bottom. As the triangles in Fig. 14
show, nearly all the atoms are kicked out of the beam before hitting the window.
Although only 3% of the atoms go through the tunnel as shown in Fig. 14
(A ~ 1.5 GHz), the efficiency was about 10% when they reduced the detuning
to increase the atomic speed.
IV] M A N I P U L A T I O N OF C O L D ATOMS IN HLBs 175

FiG. 15. (a) CCD image of the low-velocity rubidium beam split and guided by the HLB.
(c) Spatial profile of the rubidium beam taken at the place indicated by the arrow. (b) and (d)
show the corresponding image and the spatial profile without the HLB (from Yan et al., 2000,
J. Opt. Soc. Am. B 17(11), 1817, Fig. 4, reprinted with permission).

A. 4. H L B s p l i t t i n g o f Rb

Yan et al. [64] guided a continuous low-velocity atomic beam and achieved
incoherent splitting of the atomic beam with the HLB. A low-velocity intense
source of atoms (LVIS) was generated by the method demonstrated by Lu
et al. [65]. An HLB was generated by an axicon setup that converts a Gaussian
beam from a Ti:sapphire laser into the HLB as explained in Sect. III.D [28].
A convergent HLB with a full convergence angle o f 5 ~ was generated from a
collimated HLB with a dark center diameter o f ~1 cm.
With the intercepting angle a = 0, they investigated the effect o f the HLB for
the guiding. The result was that, owing to the HLB guiding and collimating or
focusing, the rubidium flux was increased by ~20% and the spatial width was
reduced from 0.9 mm to 0.7 mm. Figure 15a shows the CCD image of the LVIS
split and guided by an HLB with a convergence angle o f ~ 5 ~ and intercepting the
LVIS at a ~ 8 ~ The top of the image was about 4 m m below the center o f the
MOT. Spatial separation of the LVIS is observed: one atomic beam is propagating
downward and a second atomic beam is produced by the atoms guided by the
HLB along the propagating direction of the HLB. This effectively realizes an
incoherent atomic beam splitter. For comparison, Fig. 15b shows the CCD image
of the LVIS without the HLB. Figures 15c and 15d plot the cross-sectional spatial
profiles of the rubidium atomic beams taken at a distance o f 9 m m below the
MOT. The maximum atomic flux guided by the HLB is 50% o f the flux intensity
of the free-traveling rubidium beam without the HLB.
176 H.-R. Noh et al. [IV

B. ATOMIC FOUNTAIN WITH HOLLOW LASER BEAMS

The development of an atomic fountain based on laser-cooled atoms [66, 67] has
created prospects for an improved accuracy and stability of frequency standards.
In such a clock, one approach to solve the line shift due to cold collisions is to
use laser light for guiding the upward-launched atoms [68]. This is because the
guiding can enhance the number of atoms which come back into the microcavity
without increasing the atomic densities.
Optical guiding of an atomic fountain by using a cylindrical HLB was recently
demonstrated by Kim et al. [69]. They generated an HLB by using the micro-
imaging method described in Sect. III.E. It was collimated by the objective lens
and propagated downwards toward the center of the Rb MOT. The power of
the guiding laser was 250 mW and the beam waist was 3 mm. The HLB was
nearly collimated in order to remove the dipole force of the guiding direction,
which can cause broadening of the spatial distribution of guided atoms. With an
intensity of 3 mW/cm 2 in each beam, the typical diameter of an atomic cloud
in the MOT was about 1 mm, and the number of trapped atoms was typically
2 • 107. Cold atoms were then launched upwards in a rather simple way by rapidly
varying the vertical magnetic field, resulting in the atomic Zeeman shift. After
1-ms acceleration, the detuning of the laser beams was changed from -2.5F to
-70F, lowering the atomic temperature to 33.7 ~tK in the frame moving upwards.
A typical launching velocity of ascending atoms was 1.4 m/s and the atoms were
launched up to 10 cm.
The number of guided atoms was detected by observing with a photomultiplier
tube the fluorescence induced by a horizontally placed probe laser at 10.5 cm
below the center of the MOT. They observed that 0.5% of the launched atoms
were detected without the HLB. On the other hand, a tenfold enhancement
of the HLB-guided atomic fountain was clearly obtained without appreciable
heating. In Fig. 16, curve (a) is the time-of-flight (TOF) signal of atoms
launched without the HLB, while curve (b) is the TOF signal with the HLB
at a detuning of 19 GHz. From this TOF signal, one can deduce the guiding
efficiency of atoms and the temperature. Without the HLB, the temperature was
about 33.7 (-+-2.1) ~tK; with the HLB it was about 34.4 (• ~K.
To characterize the enhancement due to the guiding HLB, they introduced the
enhancement factor, defined as the ratio of the number of atoms guided with
the HLB to that without HLB. In Fig. 17, the curve with solid squares shows
the relationship between the enhancement factor and the detuning measured
with respect to the 5S1/2,F = 2 ~ 5/93/2 transition line. The inset shows
the enhancement of the guiding efficiency for larger detuning. The curve with
open circles shows how the number of scatterings, or the heating, is changed
with the detuning. They observe that for small detuning the enhancement factor
is more than 35, but there is serious heating. As the detuning increases, the
enhancement factor as well as the heating decrease. Note that the number
IV] MANIPULATION OF COLD ATOMS IN HLBs 177

10
P = 250 mW
.~,08
6z = 19 G H z
.I=I
Vlaunch---- 1.4 m / s
5
HLB
"r - " 0 6

1:1 04
Q

HLB ~{j~~
9 o2

O0 _ .~' . _ ~ _.. ~ ,, _, . . . 7 . . . ,,. _ l.~.-. ,, ,.._ ,_ T , - ~ , ....... ~._~L I ' " "I

0 05 010 015 0 20 0 25 0 30 0 35 0 40 0 45

-time of Right (s)

FIG. 16. TOF signals in the HLB-guided atomic fountain experiment: (a) without the HLB;
(b) with the HLB.

-- Enhancing Factor 40~


40 - - o - - Number of Scatterings 30J 2000EZ
I__ 3
O
o 30 /'~ ~0t i~---....__ 1500~"
U_ #~o " 0 50 loo O
1000(/)
o#/~o
/ ~o '.%"-.-..
t- 20
.m

I::: ~' \ ,', ..... ,.,,,,,... ,.,,,


t- 10 500
t- / %'000 -
uJ OOOOOOOOOOOOOOOOOOOOOOOOOOOOOO "
0 0
g 1'0 1'5 i0
6 2 [GHz]
FIG. 17. Dependence of the enhancement factor (squares) and the number of scatterings (circles)
on the detuning 6. The inset shows the enhancement factor for larger detuning.

of scatterings decreases more rapidly (,~(~-2) than the enhancement factor.


At a detuning of 19GHz, the enhancement factor is over 10 and an atom
experiences spontaneous emissions about 40 times during the launching and
falling processes. According to the calculation, however, they found that the
heating due to spontaneous emissions was not so serious.
In order to reduce the loss of atomic coherence, an HLB with a large
detuning may be used. For example, if a 15-W Ar-ion laser is used for a
tenfold enhancement of guiding efficiency, the average rate of spontaneously
178 H.-R. Nob et al. [IV

scattered photons is calculated to be l0 -3 Hz. While the number of atoms


being guided in the fountain is increased, the HLB introduces inhomogeneous
energy shifts of the ground-state hyperfine levels. In a trap based on a sheet
of blue-detuned light supporting against gravity, a Stark shift of 270mHz is
obtained for 4 s trapping time, which is larger than the line shift due to cold-
atom collisions [70]. One possibility for reducing the light shift in the HLB is to
use a Bessel beam of much higher order for the HLB, or to use the evanescent
waves of a hollow optical fiber. Since the evolution of atoms in an HLB depends
on the shape of the HLB, it is suggested that the ensemble-averaged heating and
the light shift will be changed with the shape of the HLB [71]. In principle, if
the potential of the HLB is square, then atoms in the HLB may not feel any
scattering or light shift.

C. ATOMIC TRAPS WITH HOLLOW LASER BEAMS

The first optical dipole trap consisted of a strongly focused Gaussian laser
beam, detuned to the red side of the atomic resonance line [72]. Since then,
several schemes have been realized to reduce the scattering rate that limits the
trap lifetime [73], such as a far-off-resonance trap (FORT)[74] and far-detuned
traps operating with Nd:YAG [75] or CO2 [76] lasers. Since the large scattering
rate or the ac-Stark energy shift is a serious obstacle to precision spectroscopy,
blue-detuned optical dipole traps have been designed to provide lower scattering
rate and less energy shift by using the fact that atoms mostly remain in the region
of low laser intensity.
Before the advent of HLB, Davidson et al. [70] at Stanford demonstrated
the first blue-detuned dipole trap by using two sheets of Ar + laser beams. The
beams intersect at 90 ~ forming a V-shaped cross section, and provide a strong
confinement perpendicular to the beam propagation axis. Confinement along the
laser propagation axis is provided by the divergence of the focused light sheets.
In this trap, a 1/e coherence decay time of 4.4 s was obtained. A linewidth of
0.125 Hz and a Ramsey fringe contrast of 43% were also obtained. The coherence
time was 300 times longer than that achieved in a red-detuned Nd:YAG laser
dipole trap with a comparable trap depth. In a later experiment, they used two
pairs of light-sheets, forming an inverted pyramid [77]. After Raman cooling,
4.5• 105 atoms were loaded at a temperature of 1.0~tK to a final density of
4 • 10 ll cm -3 with 1/e lifetime 7.0 s. In this subsection, we discuss several types
of atom traps that use blue-detuned hollow laser beams.

C. 1. Single-HLB trap of Rb
Kuga et al. [8] have demonstrated a novel optical dipole trap using an LG laser
beam. Precooled Rb atoms were trapped in the dark core of the doughnut beam
(2D trap). Because there was no restoring force along the axial direction, they
IV] MANIPULATION OF COLD ATOMS IN HLBs 179
, - 9 . ,,,

10 a

o
L..
i1)
.io
E
r 107

, ! | i ~ , t

100 200 300


trapping tln'~ (ms)

FIG. 18. Decay of atoms from the 2D (circles) and 3D (squares) trap (from Kuga et al., 1997,
Phys. Rev. Lett. 78(25), 4713, Fig. 4, reprinted with permission).

added two "plugging" laser beams to make a three-dimensional optical trap


(3D trap). To obtain the plugging beams, they recycled the doughnut-shaped
HLB that was divided into two beams and redirected to the trap with a separation
of 2 mm. The HLB was 1.5 mm in diameter and the plugging beam had a
diameter of 0.7 mm.
An LG-mode HLB was generated from a Hermite-Gaussian (HG) beam by
the cylindrical-lens mode conversion method (Sect. III.A), which was produced
by a Ti:sapphire (TS) laser pumped by an all-line Ar-ion laser. A tungsten wire
of 20ram diameter was inserted into the TS laser cavity and its position was
adjusted to generate the HG03 laser beam. It was converted to the LG30 mode
by an astigmatic mode converter composed of a pair of cylindrical lenses (focal
l e n g t h f = 25 ram) separated a distance d = x/2j.
Figure 18 shows the decay of the number of trapped atoms as a function of
the trapping time. The lifetime of the 3D trap was determined to be 150 ms. The
decay of the 2D trap was faster than that of the 3D trap and did not fit a simple
exponential function. By extrapolating the decay curve to zero trapping time,
they could estimate that one-third of the atoms in the MOT was initially loaded in
the dipole trap. The temperature of the trapped atoms was approximately 18 p~K,
almost independent of the trapping time. These results also suggested that the
lifetime of the 2D trap was determined mainly by the radiation pressure. From
the decay curve of the 2D trap, they estimated that the photon scattering rate
was -~100 S-1 .
When they used an LG~ beam for trapping, the lifetime was only a few tens of
milliseconds, shorter than that with the LG 3 beam. The reason is that a lower-/
doughnut beam has a smaller dark spot. Therefore, they expected that the lifetime
of the novel trap would be considerably extended by using a doughnut beam that
is not only intense at a large detuning but also high in I. Such a beam can be
efficiently produced from a Gaussian laser beam by a blazed phase hologram
[Sect. III.B].
In a later experiment, they applied polarization-gradient cooling (PGC) to
extend the lifetime of the trap [78]. They estimated the number of atoms initially
180 H.-R. Noh et al. [IV

loaded in the dipole trap to be 2x 108 (loading efficiency from the MOT to the
dipole trap was 30%). The time constant of r = 1.5 s was consistent with losses
due to collisions with background gas. This indicates that the trap loss due to
heating could be efficiently suppressed by pulsed PGC. The temperature of the
trapped atoms was 13 gK, close to the temperature of 10 ~tK achieved by PGC
just before the dipole trap was turned on.

C.2. Single-HLB trap of Cs


Ovchinnikov et al. [9] have presented a gravito-optical surface trap (GOST)
in which they used evanescent-wave cooling to store Cs atoms just above a
flat dielectric surface. Horizontal confinement was provided by the conservative
optical dipole potential due to a cylindrical HLB, far blue-detuned from the
atomic resonance. The evanescent-wave (EW) cooling mechanism in the GOST
is based on the splitting of the 2S1/2 ground state of Cs into two hyperfine
sublevels with F = 3, 4. As an inelastic reflection takes place when the atom
enters the EW in the F = 3 lower ground state, it is pumped into the less
repulsive F = 4 upper ground state by scattering an EW photon during the
reflection process. The EW that forms the bottom of the GOST was produced
on the flat horizontal surface of a fused-silica prism by total internal reflection
of the 60-mW beam delivered by a laser diode. The upward directed HLB used
for horizontal confinement in the GOST was produced by imaging the 300-mW
output of a TS laser with a spherical lens (achromatic doublet, focal length
100mm) in combination with an axicon (Sect. III.D]. The GOST was loaded
from a standard MOT placed right at the center of the HLB about 800 gm above
the prism surface.
They measured the lifetime of atoms in the GOST by recapturing them in the
MOT after a variable time. Figure 19 shows the results they obtained for two
different values of the rest-gas pressure in the apparatus (p = 4.2x 10-l~ mbar

e , , , ,

e -1 o ~

Zo e -2

~- e-S
e -4
-5 I I
I
e o 5 lo 15 20 25
time (s)
FIG. 19. Number of atoms in the GOST versus storage time at a rest gas pressure of
4.2x 10 - l ~ mbar (solid circles) and 7.6x 10 - l ~ mbar (open circles). The solid lines are exponential
fits (from Ovchinnikov et al., 1997, Phys. Rev. Lett. 79(12), 2225, Fig. 3, reprinted with permission).
IV] MANIPULATION OF COLD ATOMS IN HLBs 181

and p = 7.6 • 10-l~ mbar). In about 1 s, the decay is found to be exponential with
1/e lifetimes of (6.0-+-0.1)s and (3.2• respectively. These results suggest
that, besides small transfer losses of ~30% observed in the first second, losses
from the GOST are essentially due to rest-gas collisions. From the TOF method,
they obtained the vertical and horizontal temperatures To = (3.0+0.1) ~tK and
Th = (3.1-+-0.3) ~tK after 4 s of storage and cooling in the GOST. Both have equal
values within experimental uncertainty. In order to study the cooling dynamics,
they also performed a series of measurements on the temperatures To and Th as
a function of the storage time in the GOST. They obtained the EW cooling rate
to be 1/fi = (380-+-20)ms, which is in good agreement with the predicted value
of 1/fl = (400• ms [16].
Ovchinnikov et al. also demonstrated a gravito-optical dipole trap, which
used an intense blue-detuned conical HLB together with gravity to confine an
ensemble of cesium atoms in a dark spatial region, producing a conical atom trap
(CAT) [79]. The conical trapping beam with an opening angle of 0 ~ 150 mrad
is generated by using two identical axicons: in a telescope-like arrangement, a
Gaussian laser beam is transformed into a collimated HLB. This tubular beam
is then focused with a spherical lens to generate the conical HLB with its
apex located in the focal plane. In the focal plane, the beats profile is roughly
Gaussian with a diameter of about 100 ~tm. Within a few millimeters, the beam
profile evolves into a ring shape in resemblance to a higher-order LG mode [8].
Along the symmetry axis, the intensity distribution is approximately Gaussian
and decreases to 1/e of its peak value (maximum optical potential of ~50 mK at
3 GHz detuning) within an estimated distance of ~700 ~tm.
They have demonstrated the storage of Cs atoms in the CAT with a lifetime
of several seconds, limited only by collisions with the residual gas in the
vacuum chamber. After a variable storage time in the CAT, atoms were retrapped
into the MOT and a fluorescence image was taken with a CCD camera. The
corresponding experimental results are shown in Fig. 20 for two different values
of the residual gas pressure in the vacuum chamber (p - 2.5 x 10-1~ mbar and
6.9• 10-1~ mbar). An exponential decay is observed with lifetimes of 7.8s and
2.8 s, respectively. This indicates that collision with the residual gas is also
the predominant loss mechanism of atoms in the CAT. The measurement also
demonstrates a very high transfer efficiency from the MOT into the CAT of
about 80%. This corresponds to an absolute number of ~8 • 105 atoms initially
stored in the CAT. The atoms in the CAT are cigar shaped, with the diameter of
about 100 ~tm and the length of about 1 mm. The peak density is on the order
of 1011 cm -3 with about 10 6 atoms in the CAT.
Webster et al. [80] demonstrated a dipole trap by using a blue-detuned
LG-mode HLB based on the geometry proposed by Morsch and Meacher [81 ].
An LG-mode doughnut-shaped beam was produced by a mode converter
(Sect. III.A) from an HG beam emitted by an argon-ion laser. The doughnut-
182 H.-R. Noh et al. [IV

! | ! |

-!
e

e -2

z
e -B

-4 i 9
e
o 5 10 15 20 25
t i m e (s)

FIG. 20. Measurement of the storage time in the CAT as a function of storage time. The measured
lifetime is 7.8 s at a background pressure of 2.5 • 10-1~ mbar (solid circles). At 6.9x 10-l~ mbar (open
circles), a lifetime of 2.8 s is observed (from Ovchinnikov et al., 1998, Europhys. Lett. 43(5), 510,
Fig. 2, reprinted with permission).

shaped beam was expanded by a telescope to a beam waist of 3.5 m m and then
focused by a lens of focal length 25 mm, located one focal length from the centre
of the trap. To make the temperature o f the atomic cloud less than the height o f
the potential barrier, and to make the cloud small enough to fit inside the cone,
the cesium atoms were evaporatively cooled in a magnetic trap [82] to give a
temperature of 1 ~tK and a cloud diameter in the horizontal plane of 150 ~tm.
The size of the trapped cloud was roughly 100 ~tm radially and 200 ~tm axially,
elongated in the direction of the beam axis. To increase the trap depth, they
noted that a better method for achieving a higher dipole potential would be to
use a holographic technique for producing the HLB, since holography enables
one to reproduce an arbitrary waveform from a Gaussian laser beam with an
efficiency greater than 90% [60]. In particular, they demonstrated application to
a rubidium BEC. A condensate was formed in a magnetic trap positioned above
the apex of a focused doughnut-mode HLB and then released so that, falling
downwards, it would be funnelled towards the focused region, which thus realizes
a condensate propagating through a small orifice. This may permit investigation
of the superfluid flow in Bose condensates produced by evaporative cooling in
the magnetic trap.

C.3. C r o s s e d - H L B trap o f Rb

Xu et al. [83] constructed a blue-detuned optical dipole trap by intersecting two


horizontal, cylindrical HLBs at a right angle in the center o f a Rb MOT. The
polarizations of the beams were chosen to be orthogonal in the crossed region
in order to suppress standing-wave effects. The detuning the HLBs was 20 GHz
from the 5 8 1 / 2 , F = 3 ~ 5P3/2,F t = 4 transition line of 85Rb, and the trap depths
were about 10 ~tK in the x-direction and 90 ~tK in the z-direction, respectively.
IV] MANIPULATION OF COLD ATOMS IN HLBs 183

o~ 100.
",~176
or 8O
] 9 Experimental data
--
9 6o
o "., .... S i m u l a t i o n result
a::
4O
t:D "-..
c
~.
9 20 "-J[.,
D. ...
tO "" . . . . . . . . ;b'" "e. . . . . . .
'- 0
2'o 4'o 8'o ~oo
9 , 9

0 60
Trapping time (ms)
FIG. 21. The trapping efficiency of a crossed-HLB trap as a function of the trapping time. The
two horizontal HLBs have powers of 200 and 400 roW, the detuning of each HLB is 20 GHz, and
the initial temperature of atoms is 16 gK.

They estimated that 60% of the atoms in the MOT were initially loaded in the
HLB trap.
The number of atoms trapped, and their temperature, could be deduced by
TOF measurements. They found that the temperature of the trapped atoms was
about 7 gK and this value was close to the minimum height of potential barrier of
10 gK, where about 105 atoms stayed inside the trap for 100 ms and the lifetime
of trapped atoms was about 20ms as shown in Fig. 21. Figure 21 also shows
the trapping efficiency as a function of the trapping time. For comparison, the
simulation results based on Eqs. 6 - 9 in Sect. II.A are also shown as the dotted
curve in Fig. 21. Since the detuning was much larger than the splitting between
the hyperfine-structure levels of the excited state, the three-level interaction mode
was quite good for the simulation.

C.4. Single-beam hollow optical traps


There has been interesting progress in the generation of light beams having a
null intensity region surrounded by light walls in all three directions with simple
geometries. In this subsection, we discuss theoretical and experimental results on
the application of such HLBs to the blue-detuned dipole traps reported thus far.
Ozeri et al. [84] trapped atoms in a blue-detuned dipole trap that uses a
single, three-dimensional HLB detuned by 0 . 1 - 3 0 n m above the 5S~/2 ~ 5P3/2
transition in 85Rb. To produce a light distribution that is completely surrounded
by light, a collimated beam is passed through a circular phase plate that imposes
a phase difference of exactly Jr radians between the central and the outer part of
the beam, and then focused by a lens. Destructive interference between these two
parts yields a dark region around the focus, which is surrounded by light in all
directions. To form the circular Jr-phase plates, they evaporated a thin dielectric
layer (e.g., MgF2, with a refractive index of n = 1.38) through a 4-mm diameter
radial mask on a glass substrate by using a commercial optical coating machine.
To generate a phase difference of exactly Jr radians, the thickness of the dielectric
184 H.-R. N o h et al. [IV
9 ' ........ i ' i 9 i " I 9 I 9 I 9

400-

350.
9 I

300
0 II

r
250
E
200 84 101 . . . . . . .

E
;:; 150

100 "6
50. 103 .......... . .
0 200 4O0 6OO 8OO 1000

time (rnsec)
u 9 i 9 i 9 9 i ' " ' 9 ! 9
0 5 10 ~'~ ~o ~'~ ~o 35

detuning (nm)
FIG. 22. Lifetime versus detuning. The inset shows a typical lifetime measurement for
6 = 5 nm (dots). The exponential fit (solid line) gives the lifetime of 340ms (from Ozeri e t a l . ,
1999, P h y s . Reu. A 59(3), R1750, Fig. 3, reprinted with permission).

coating should be d = ~,/[2(n- 1)] = 676.3 nm for ~, = 514.5 nm. The darkness
factor, defined as the ratio between the light intensity at the center of the trap
and the light intensity at the first surrounding ring, was measured to be 1/750.
A TEM00 Gaussian laser beam was magnified into a collimated Gaussian beam
with w0 = 6 mm, passed through the :r-phase plate, and was focused with a
f - 250-mm lens into the vacuum chamber.
Figure 22 shows the measured lifetime for various trapping beam wavelengths.
For detunings 6 > 0.5 nm the 1/e trap lifetime is nearly independent of 65 at
around 300ms and is inversely proportional to the background pressure in the
vacuum chamber. This indicates that the trap lifetime is governed by collisions
with background atoms. For 6 > 0.5 nm, the trap lifetime is approximately
proportional to 6. This is consistent with a heating-induced lifetime, since the
heating rate is proportional to 0 -2, whereas the trap depth is approximately
proportional to 6 -1. The peak density was measured to be 6.8• l0 II atoms/cm 3.
In a later experiment [85], they made a similar dipole trap with a much larger
volume and a more symmetric shape than before. They achieved this by
simultaneously exploiting two diffraction orders of a properly designed binary
phase element (BPE)[86], which was composed of concentric phase rings with
a phase difference of ;r between subsequent rings, thus creating a radial grating
with uniform spacing. Figure 23 depicts the 1st a n d - l s t diffraction orders from
the radial grating, each having a diffraction efficiency of 40.5%, which together
form an outgoing cone of light. When focused, the cone appears in the focal
plane as a narrow ring. The ring closes upon itself in both sides of the focal plane
owing to the beam divergence to form a dark region completely surrounded by
light.
Zem~nek and Foot [87] proposed a blue-detuned dipole trap by using two
IV] MANIPULATION OF COLD ATOMS IN HLBs 185

FIG.23. Generation of a dark, hollowoptical trap by a binary phase element (BPE). The 1st (grey)
and-lst (dashed) diffracted beams are focused and generate a dark region around the focal plane
(from Ozeri et al., 2000, J. Opt. Soc. Am. B 17(7), 1113, Fig. 1, reprinted with permission).

counterpropagating Gaussian waves with different beam waists. At the nodes,


the field intensity of the standing wave is not completely cancelled at all radial
positions across the beam. This creates an intensity dip in both the axial and
radial directions that can be used as an atomic trap with blue-detuned laser light.
In particular, the intensity dips in the standing wave correspond to an array of
small traps. They found that the bigger the difference in beam waists is, the
bigger is the trap width in proportion to the width of the smallest beam. In
particular, the radial width changes more rapidly with axial position. They also
found that the bigger the difference in beam waists, the wider and deeper the
trap becomes.
Arlt and Padgett [88] demonstrated that a CGH can be used to form an optical
beam with a well-localized intensity null at its focus. The beam is a superposition
of two LG modes that are phased in such a way that they interfere destructively
to give a beam focus that is surrounded in all directions by regions of higher in-
tensity. When two LG modes are appropriately adjusted to give the same on-axis
intensity, the relative phases of these two modes can be controlled so that they
interfere destructively at their common focus to give zero on-axis intensity. To
demonstrate the generation of such an optical bottle beam, they chose to produce
the required superposition of LG modes by employing a CGH. They observed
good agreement between the calculated and observed beam cross sections and
profiles. They concluded that there were no holes due to unwanted interference
effects and the optical bottle was complete. The residual on-axis intensity was
measured to be less than 1% of the surrounding radial maximum.
Cacciapuoti et al. [89] proposed a single-beam blue-detuned optical trap
by using an axicon in combination with two spherical lenses. A collimated
laser beam passes through a converging lens and then enters the axicon. If the
distance d between the first lens and the axicon is greater than the lens focal
length f , the optical rays produce a ring-shaped intensity distribution in the back
focal plane of the first lens. Since this radiation field is imaged by the second lens
in the image plane, it is possible to observe a dark region surrounded by light.
The generated optical potential is cylindrically symmetric and bounded in all
directions. They experimentally obtained an optical bottle beam for a refractive
index of the axicon n = 1.51, base angle a ___ 100 mrad, f = 12.5 cm, d = 14 cm,
186 H.-R. Nob et al. [IV

focal length of the second lens f l = 10.0 cm, and distance between the axicon
and the second lens d I = 93 cm. As a result, they observed a ring radius of
43 gm in the image plane, which is in qualitative agreement with the theoretical
prediction of 90 gm.

C.5. Rotating-beam hollow optical trap


Friedman et al. [90] demonstrated a single-beam blue-detuned hollow optical
trap, based on a rapidly rotating laser beam that traps atoms in a dark volume
completely surrounded by light. The rotating beam optical trap (ROBOT) is com-
posed of a linearly polarized, tightly focused Gaussian laser beam ( 1 / e 2 radius
of w0 = 16 gm), which is made to rotate rapidly (up to 400 kHz) by using two
perpendicular acousto-optic scanners. If the rotation frequency is high enough,
the optical dipole potential can be approximated as a time-averaged quasistatic
potential. With a laser power of 200mW, a detuning of 6 = 1.5 nm above the
D2 line of 85Rb, and the radius of rotation r = 100~tm, the minimal potential
height is estimated to be 22 gK, which scales approximately a s r - 2 .
The ROBOT was loaded from a compressed 85Rb MOT [91] having ~3 x 108
atoms, with temperature 9 gK and peak density 1.5 x 101~ cm -3. The measured
1/e lifetimes are shown in Fig. 24 for two rotation radii. For r = 67 gm,
the trap becomes stable for for > 20kHz, where the lifetime is -600ms. For
r = 32 gm, on the other hand, stable trapping is achieved only forfrot > 60 kHz.
To increase the density of the atomic sample adiabatically, the trap was slowly
compressed from rinitial = 100~tm to rfinal = 27~tm in 150ms. Furthermore,
the density was increased by optically cooling the atomic sample. The lowest
final temperatures were T: = 28 ~tK and Tx = 41 ~tK. From the measured
temperatures and oscillation frequencies, the peak atomic density is calculated to
be 2 • cm -3. As an application, they also measured the weak transition rate
in the dark ROBOT [92]. They measured the 5S1/2 -~ 5D5/2 transition rate of

750"
@

9 o o
e' ~ 5O0 o
o

9 o

l
.z 250
o o 9 r=67p m
o o r=32~m
" O --
O

|~ , i 9 i 9 | "~ i , , , i 9 i 9

0 20 40 60 80 100 120 140 160


Rotation frequency [kHz]

FIG. 24. Atomic lifetime in ROBOT as a function of the frequency of the trapping beam (from
Friedman et al., 2000, Phys. Rec. A 61(3), 031403, Fig. 2, reprinted with permission).
IV] MANIPULATION OF COLD ATOMS IN HLBs 187

400

300 "'.,.,
CO
E "~" .,..,"~" '~"..,
200 ."1 ss "I
O3 ,," 1 so
,r ." , 1.45 ~ ~.
E~4o
100
~ ."

1.35

1 .all

1.zs
1'2~ 5-1"0-1'5 2'0 2"5 3"o 35 4"0 45
' Idetunin~ (OHz I

lb
"IV B,

0 20 30 40 50
detuning (GHz)
Ff6. 25. Number of atoms captured by RODiO as a function of trap radius (from Rudy et al.,
2001, Opt. Exp. 8(2), 159, Fig. 3, reprinted with permission).

85Rb atoms using an extremely weak (25 gW) probe laser beam. The measured
transition rate was as small as 0.09 s-~. Note that the striking feature was the
long spin-relaxation times combined with tight confinement of the atoms in a
dark, hollow optical dipole trap.
Rudy et al. [93,94] demonstrated a Rotating Off-resonant Dipole Optical
(RODiO) trap, a similar all-optical dynamical dark trap. A circular scan of radius
1-5 mm was created by using two perpendicular linear scans phase-shifted by
90 ~ In this way, they created an average trap potential with a spatially averaged
height of ~< 240 gK. After the RODiO trap was loaded, it was left on for a
variable time duration, At. Figure 25 shows the atomic confinement at different
scan radii taken after 30 ms at A - 30 GHz. For small radii, the averaged potential
well was deep, but the trap beam cut into the initial trapped atoms, expelling
many atoms. As shown in the inset, the decay time is faster in this regime because
the interaction time with the light is significant. When the radius is increased,
since the interaction time as well as the time-averaged scattering force decrease,
the number of trapped atoms rises sharply.
They also examined the dependence of confinement time rRoDio on the scan
rate by implementing a two-dimensional RODiO at 2, 5 and 11 kHz scan rates.
For the 11-kHz case, they found rRODiO ~ 40ms at the optimum conditions
of A = 30GHz. At 5kHz, rRODiO ~ 13 ms and retention was observed
beyond 25 ms. At 2 kHz, rRODiO ~ 6 ms and finite retention was found well
beyond 10 ms. In particular, they discussed the merits of RODiO over other blue-
188 H.-R. Noh et al. [VI

detuned dipole traps, such as the flexibility with respect to the number of laser
beams, the realization of an arbitrarily sized dark region, and the completely dark
interior.

V. A c k n o w l e d g m e n t

The authors acknowledge the support of the Creative Research Initiatives of the
Korean Ministry of Science and Technology.

VI. R e f e r e n c e s

1. Adams, C.S., Siegel, M., and Mlynek, J. (1994). Phys. Rep. 240, 143-210.
2. Balykin, V.I. (1999). Adv. At. Mol. Opt. Phys. 41, 181-260.
3. Folman et al., this volume.
4. Renn, M.J., Montgomery, D., Vdovin, O., Anderson, D.Z., Wieman, C.E., and Cornell, E.A.
(1995). Phys. Rev. Lett. 75, 3253-3256.
5. Renn, M.J., Donley, E.A., Cornell, E.A., Wieman, C.E., and Anderson, D.Z. (1996). Phys. Rev.
A 53, R648-R651.
6. lto, H., Nakata, T., Sakaki, K., Ohtsu, M., Lee, K.I., and Jhe, W. (1996). Phys. Rev. Lett.
76, 4500-4503.
7. Ito, H., Sakaki, K., Ohtsu, M., and Jhe, W. (1997). Appl. Phys. Lett. 70, 2496-2498.
8. Kuga, T., Torii, Y., Shiokawa, N., and Hirano, T. (1997). Phys. Rev. Lett. 78, 4713-4716.
9. Ovchinnikov, Yu.B., Manek, I., and Grimm, R. (1997). Phys. Rev. Lett. 79, 2225-2228.
10. Kuppens, S., Rauner, M., Schiffer, M., Sengstock, K., and Ertmer, W. (1998). Phys. Rev. A
58, 3068-3079.
11. Song, u Milam, D., and Hill IIl, W.T. (1999). Opt. Lett. 24, 1805-1807.
12. Xu, X., Kim, K., Jhe, W., and Kwon, N. (2001). Phys. Rev. A 63, 63401.
13. Xu, X., Minogin, V.G., Lee, K., Wang, Y.Z., and Jhe, W. (1999). Phys. Rev. A 60, 4796-4804.
14. Xu, X., Wang, Y.Z., and Jhe, W. (2000). J. Opt. Soc. Am. B 17, 1039-1050.
15. Minogin, V.G., and Letokhov, V.S. (1987). "Laser Light Pressure on Atoms." Gordon and Breach,
New York.
16. S6ding, J., Grimm, R., and Ovchinnikov, Yu.B. (1995). Opt. Comm. 119, 652-662.
17. Cohen-Tannoudji, C., Dupont-Roc, J., and Grynberg, G. (1992). "Atom-Photon Interactions,
Basic Processes and Applications." Wiley, New York.
18. Dalibard, J., and Cohen-Tannoudji, C. (1985). J. Opt. Soc. Am. B 2, 1707-1720.
19. Mamaev, A.V., Saffman, M., and Zozulya, A.A. (1996). Phys. Rev. Lett. 77, 4544-4547.
20. Wang, X., and Littman, M.G. (1993). Opt. Lett. 18, 767-769.
21. Harris, M., Hill, C.A., and Vaughan, J.M. (1994). Opt. Commun. 106, 161-166.
22. Lee, H.S., Atewart, B.W., Choi, K., and Fenichel, H. (1994). Phys. Rev. A 49, 4922-4927.
23. Paterson, C., and Smith, R. (1996). Opt. Commun. 124, 121-130.
24. Beijersbergen, M.W, Allen, L., van der Veen, H.E.L.O., and Woerdman, J.P. (1993). Opt.
Commun. 96, 123-132.
25. Padgett, M., Arlt, J., Simpson, N., and Allen, L. (1996). Am. J. Phys. 64, 77-82.
26. Beijersbergen, M.W., Coerwinkel, R.P.C., Kristensen, M., and Woerdman, J.P. (1994). Opt.
Commun. 112, 321-327.
27. Turnbull, G.A., Robertson, D.A., Smith, G.M., Allen, L., and Padgett, M.J. (1996). Opt. Commun.
127, 183-188.
VI] M A N I P U L A T I O N O F C O L D A T O M S IN H L B s 189

28. Manek, I., Ovchinnikov, Yu.B., and Grimm, R. (1998). Opt. Commun. 147, 67-70.
29. Ito, H., Sakaki, K., Jhe, W., and Ohtsu, M. (1997). Phys. Rev. A 56, 712-718.
30. Yin, J.E, Noh, H.R., Lee, K.I., Kim, K.H., and Jhe, W. (1997). Opt. Commun. 138, 287-292.
31. Siegman, A.E. (1986). "Lasers." University Science, Mill Valley, CA, pp. 685-695.
32. Allen, L., Beijersbergen, M.W., Spreeuw, R.J.C., and Woerdman, J.P. (1992). Phys. Rev. A
45, 8185-8189.
33. He, H., Fries, M.E.J., Heckenberg, N.R., and Rubinsztein-Dunlop, H. (1995). Phys. Rev. Lett.
75, 826-829.
34. Simpson, N.B., Dholakia, K., Allen, L., and Padgett, M.J. (1997). Opt. Lett. 22, 52-54.
35. Paterson, L., MacDonald, M.P., Arlt, J., Sibbett, W., Bryant, EE., and Dholakia, K. (2001).
Science 292, 912-914.
36. Dholakia, K., Simpson, N.B., Padgett, M.J., and Allen, L. (1996). Phys. Rev. A
54, R3742-R3745.
37. Power, W.L., Allen, L., Babiker, M., and Lembessis, V.E. (1995). Phys. Rev. A 52, 479-488.
38. Lai, W.K., Babiker, M., and Allen, L. (1997). Opt. Commun. 133, 487-494.
39. Snadden, M.J., Bell, A.S., Clake, R.B.M., Riis, E., and McIntyre, D.H. (1997). J. Opt. Soc. Am.
B 14, 544-552.
40. Heckenberg, N.R., McDuff, R., Smith, C.P., Rubinsztein-Dunlop, H., and Wegener, M.J. (1992).
Opt. Quantum. Electron. 24, $951-$962.
41. Heckenberg, N.R., McDuff, R., Smith, C.P., and White, A.G. (1992). Opt. Lett. 17, 221-223.
42. He, H., Heckenberg, N.R., and Rubinsztein-Dunlop, H. (1995). J. Mod. Opt. 42, 217-223.
43. Durnin, J., Miceli, J.J., and Eberly, J.H. (1987). Phys. Rev. Lett. 58, 1499-1501.
44. Clifford, M.A., Arlt, J., Courtial, J., and Dholakia, K. (1998). Opt. Commun. 156, 300-306.
45. Arlt, J., Dholakia, K., Allen, L., and Padgett, M.J. (1998). J. Mod. Opt. 45, 1231-1237.
46. Herman, R.M., and Wiggins, T.A. (1991). J. Opt. Soc. Am. A 8, 932-942.
47. McLeod, J.H. (1954). J. Opt. Soc. Am. 44, 592-597.
48. B61anger, P.A., and Rioux, M. (1978). Appl. Opt. 17, 1080-1086.
49. Arlt, J., and Dholakia, K. (2000). Opt. Commun. 177, 297-301.
50. Sudo, S., Yokoyama, I., Yaska, H., Sakai, Y., and Ikegami, T. (1990). IEEE Photon. Tech. Lett.
2, 128-131.
51. Christov, I., Kapteyn, H., and Murnane, M. (1998). Opt. Express 3, 360-365.
52. Marksteiner, S., Savage, C.M., Zoller, E, and Rolston, S. (1994). Phys. Rev. A 50, 2680-2690.
53. Ito, H., Sakaki, K., Nakata, T., Jhe, W., and Ohtsu, M. (1995) Opt. Commun. 115, 57-64.
54. Won, C., Yoo, S.H., Oh, K., Paek, U.C., and Jhe, W. (1999). Opt. Commun. 161, 25-30.
55. Yoo, S.H., Won, C., Kim, J.A., Kim, K., Shim, U., Oh, K., Paek, U.C., and Jhe, W. (1999).
J. Opt. B 1,364-370.
56. Gloge, D. (1971). Appl. Opt. 10, 2252-2258.
57. Goodman, J.W. (1968). "Introduction to Fourier Optics." McGraw-Hill, New York.
58. Okoshi, T. (1982). "Optical fibers." Academic Press, New York.
59. Shin, Y.I., Kim, K., Kim, J.A., Noh, H.R., Jhe, W., Oh, K., and Paek, U.C. (2001). Opt. Lett.
26, 119-121.
60. Schiffer, M., Rauner, M., Kuppens, S., Zinner, M., Sengstock, K., and Ertmer, W. (1998). Appl.
Phys. B 67, 705-708.
61. Schiffer, M., Christ, M., Wokurka, G., and Ertmer, W. (1997). Opt. Commun. 134, 423-430.
62. Gale, M.T., Rossi, M., Schutz, H., Ehbets, P., Herzig, H.P., and Prongu, D. (1993). Appl. Opt.
32, 2526-2533.
63. Bongs, K., Burger, S., Dettmer, S., Hellweg, D., Arlt, J., Ertmer, W., and Sengstock, K. (2001).
Phys. Rev. A 63, 031602.
64. Yan, M., Yin, J., and Zhu, Y. (2000). J. Opt. Soc. Am. B 17, 1817-1820.
65. Lu, Z.T., Corwin, K.L., Renn, M.J., Anderson, M.H., Cornell, E.A., and Wieman, C.E. (1996).
Phys. Rev. Lett. 77, 3331-3334.
190 H.-R. Noh et al. [VI

66. Kasevich, M., Riis, E., DeVoe, R.G., and Chu, S. (1989). Phys. Rev. Lett. 63, 612-615.
67. Clairon, A., Salomon, C., Guellati, S., and Phillips, W.D. (1991). Europhys. Lett. 16, 165-170.
68. Szymaniec, K., Davies, H.J., and Adams, C.S. (1999). Europhys. Lett. 45, 450-455.
69. Kim, K., Noh, H.R., Yeon, Y.H., Xu, X., Jhe, W., and Kwon, N. (2001). J. Korean Phys. Soe.
39, 877-880.
70. Davidson, N., Lee, H.J., Adams, C.S., and Chu, S. (1995). Phys. Rev. Lett. 74, 1311-1314.
71. Song, Y. (1999). Ph.D. dissertation, University of Maryland.
72. Chu, S., Bjorkholm, J., Ashkin, A., and Cable, A. (1986). Phys. Rev. Lett. 57, 314-317.
73. Grimm, R., Weidemfiller, M., and Ovchinnikov, Yu.B. (2000). Adv. At. Mol. Opt. Phys.
42, 95-170.
74. Miller, J.D., Cline, R.A., and Heinzen, D.J. (1993). Phys. Rev. A 47, R4567-R4570.
75. Adams, C.S., Lee, H.J., Davidson, N., Kasevich, M., and Chu, S. (1995). Phys. Rev. Lett.
74, 3577-3580.
76. Takekoshi, T., and Knize, R.J. (1996). Opt. Lett. 21, 77-79.
77. Lee, H.J., Adams, C.S., Kasevich, M., and Chu, S. (1996). Phys. Rev. Lett. 76, 2658-2661.
78. Torii, Y., Shiokawa, N., Hirano, T., Kuga, T., Shimizu, Y., and Sasada, H. (1998). Eur. Phys. J.
D 1,239-242.
79. Ovchinnikov, Yu.B., Manek, I., Sidorov, A.I., Wasik, G., and Grimm, R. (1998). Europhys. Lett.
43, 510-515.
80. Webster, S.A., Hechenblaikner, G., Hopkins, S.A., Arlt, J., and Foot, C.J. (2000). J. Phys. B
33, 4149-4155.
81. Morsch, O., and Meacher, D.R. (1998). Opt. Commun. 148, 49-53.
82. Hopkins, S.A., Webster, S., Arlt, J., Bance, P., Cornish, S., Marag6, O., and Foot, C.J. (2000).
Phys. Rev. A 61, 032707.
83. Xu, X., Kim, K., and Jhe, W. (2000). J. Korean Phys. Soc. 37, 661-664.
84. Ozeri, R., Khaykovich, L., and Davidson, N. (1999). Phys. Rev. A 59, R1750-R1753.
85. Ozeri, R., Khaykovich, L., Friedman, N., and Davidson, N. (2000). J. Opt. Soc. Am. B
17, 1113-1116.
86. Davidson, N., Ozeri, R., and Baron, R. (1999). Rev. Sci. lnstrum. 70, 1264-1267.
87. Zemfinek, P., and Foot, C.J. (1998). Opt. Commun. 146, 119-123.
88. Arlt, J., and Padgett, M.J. (2000). Opt. Lett. 25, 191-193.
89. Cacciapuoti, L., de Angelis, M., Pierattini, G., and Tino, G.M. (2001). Euro. Phys. J. D
14, 373-376.
90. Friedman, N., Khaykovich, L., Ozeri, R., and Davidson, N. (2000). Phys. Rev. A 61, 031403.
See also the chapter by Friedman et al. in this volume.
91. Khaykovich, L., and Davidson, N. (1999). J. Opt. Soc. Am. B 16, 702-709.
92. Khaykovich, L., Friedman, N., Baluschev, S., Fathi, D., and Davidson, N. (2000). Europhys.
Lett. 50, 454-459.
93. Rudy, P., Ejnisman, R., Rahman, A., Lee, S., and Bigelow, N.P. (1997). OSA Tech. Dig. 12, 67.
94. Rudy, P., Ejnisman, R., Rahman, A., Lee, S., and Bigelow, N.P. (2001). Opt. Express 8, 159-165.
A D V A N C E S IN A T O M I C , M O L E C U L A R , A N D O P T I C A L P H Y S I C S , VOL. 48

CONTINUOUS STERN-GERLACH
EFFECT ON ATOMIC IONS
G(J/NTHER WERTH l, H A R T M U T Hf4"FFNER l and WOLFGANG Q U I N T 2
l Johannes Gutenberg University, Department of Physics, 55099 Mainz, Germany,"
2Gesellschaft fiir Schwerionenforschung, 64291 Darmstadt, Germany

I. I n t r o d u c t i o n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
II. A Single Ion in a P e n n i n g Trap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
III. C o n t i n u o u s S t e r n - G e r l a c h Effect ................................. 206
IV. D o u b l e - T r a p Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
V. C o r r e c t i o n s and S y s t e m a t i c Line Shifts ............................. 212
VI. C o n c l u s i o n s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
VII. O u t l o o k . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
VIII. A c k n o w l e d g e m e n t s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
IX. R e f e r e n c e s ................................................ 216

I. I n t r o d u c t i o n

In 1922 Stern and Gerlach succeeded in spatially separating a beam of silver


atoms into two beams, utilizing the force exerted in an inhomogeneous magnetic
field on the magnetic moment of the unpaired electron in silver. This so-called
Stern-Gerlach effect was the first observation of the directional quantization of
the quantum-mechanical angular momentum, and represents a cornerstone of
quantum physics (Stern and Gerlach, 1922).
Apart from its historical role the effect has been used in numerous atomic
beam experiments to determine the magnetic moments of electrons bound in
atomic systems: An atomic beam enters a first inhomogeneous magnetic field
where one spin direction is separated from the other. The polarized atoms then
enter a region with a homogeneous magnetic field where they are subjected to
a radio-frequency (rf) field which changes the spin direction. Finally, a second
inhomogeneous magnetic field region analyzes the spin direction. Variation of
the frequency w of the rf field and recording the spin-flip probability as a
function of w yields a resonance curve. Together with a measurement of the
field strength in the homogeneous magnetic field, the magnetic moment can be
derived. Usually the size of the magnetic moment ~ is given in units of the Bohr
magneton ktB = eh/(2m) and expressed by the g-factor:

= gskts, (1)

191 Copyright 9 2002 Elsevier Science (USA)


All rights reserved
ISBN 0-12-003848-X/ISSN 1049-250X/02 $35.00
192 G. Werth et al. [I

where s is the spin quantum number. Accurate g-factor measurements represent


a critical test of atomic physics calculations for complex systems (Veseth, 1980,
1983; Lindroth and Ynnerman, 1993).
There has been an extensive discussion that the Stern-Gerlach effect could
be employed for neutral atoms only. For charged particles the magnetic force
is overshadowed by the Lorentz force acting on a moving charged particle in
a magnetic field. Proposals to use the acceleration of electrons moving along
the field lines of an inhomogeneous B-field to separate the spin directions
(Brillouin, 1928) have been discarded by Bohr (1928) and Pauli (1958) on the
basis of Heisenberg's uncertainty principle. Nevertheless attempts have been
made, however unsuccessful, to separate the two spin states in an electron beam
by using the different signs of the force on the spin exerted by a longitudinal
inhomogeneous magnetic field which results in a deceleleration or acceleration
of the electrons (Bloch, 1953). Recently, new proposals have come up to perform
Stern-Gerlach experiments on electron beams which w o u l d - in contrast to
the analysis by Bohr and Pauli - result in a high degree of spin separation
under carefully chosen initial conditions (Batelaan et al., 1997; Garraway and
Stenholm, 1999).
Dehmelt has pointed out that confining a charged particle by electromagnetic
fields provides a way to circumvent Bohr's and Pauli's reasoning (Dehmelt,
1988): the force of an inhomogeneous magnetic field on the spin of a particle
which oscillates in a parabolic potential well leads to a small but measurable
difference of its oscillation frequency for different orientations of the spin. Thus
a precise measurement of the oscillation frequency gives information on the
spin direction. Dehmelt et al. used this effect for monitoring induced changes
of the spin direction of an electron by observing the corresponding changes
in the electron's oscillation frequency. Since the sensitive electronics monitors
the trapped particle's spin direction continuously, Dehmelt coined the term
"continuous Stern-Gerlach effect" for this technique. In a series of experiments
he and his coworkers have applied this method to measure the magnetic moment
of the electron and the positron, which culminated in the most precise values of
a property of any elementary particle (Van Dyck et al., 1987). The experimental
value
gexp = 2.002319304 3766(87) (2)
agrees to 10 significant digits with the result of calculations based on the theory
of quantum electrodynamics (QED) for free particles (Hughes and Kinoshita,
1999),
gth = 2.002 319 304432 0(687) (3)
and provides one of the most stringent tests of QED. The theoretical result for
the free-electron g-factor can be expressed in a perturbation series as

gfree = 2 Ao +A1 +A2 ~ +A3 +A4 ~ +-" , (4)


I] CONTINUOUS STERN-GERLACH EFFECT ON ATOMIC IONS 193

--~,1017
10TM

,,,,"."'.".'".'".""'.". . . . . . . . . . . . . . . . . . . .
W"1012101310141015

1011 /
10lo
109
10 20 30 40 50 60 70 80 90
nuclearchargeZ
Fie. 1. Calculated expectation values for the electric field strength for hydrogen-like ions of
different nuclear charge Z. (Courtesy Thomas Beier.)

where a ,-~ ~-~


i is the fine-structure constant. The coefficients An in Eq. (4) have
been calculated by evaluating the Feynman diagrams of different orders using a
plane-wave basis set; they are of the order of unity.
In contrast to a free electron, an electron bound to an atomic nucleus
experiences an extremely strong electric field. The expectation value of the
field strength ranges from 109V/cm in the helium ion (Z = 2) to 1015 V/cm
in hydrogen-like uranium (Z = 92) (Fig. 1), and gives rise to a variety of new
effects. The largest change of the bound electron's g-factor was analytically
derived by Breit (1928) from the Dirac equation:

gBreit = (5)

The conditions of extreme electric fields also necessitate changes to be made in


the methods of calculations for the QED contributions to the electron's magnetic
moment. In a perturbative treatment a series expansion in (Za) is made in
addition to that in a. The expansion parameter Z a, however, i s - at least for
large Z - no longer small compared to 1. In addition, the expansion coefficients
can be large. For small values of the nuclear charge Z the perturbation expansion
may give reliable results, and calculations were performed which include terms
up to order (Za) 2 (Grotch, 1970a; Close and Osborn, 1971; Karshenboim et al.,
2001). For more accurate theoretical predictions, non-perturbative methods have
been developed where the solutions of the Dirac equation for the hydrogen-like
ion rather than those for the free case are used as a basis set (Beier et al.,
2000). The most recent summary of the status of QED for bound systems has
194 G. Werth et al. [I

20 40 60 80 100
10 0 100

10 -1 10 -1

I. 10-2 jag free 10 .2


O _

o 10 .3 10 .3
9;'~ AgBsQED 1. order in ot 9 . ~" 9
' 104 : /. 9
9 )r
10 -4
~o )r
o
.~ 10 -s )r 10 -~
hue.
+ + + +
0 10 .6 + 10 .6

10 .7 +
+~ AgBs QF.D2. order in a (estimate~ 10 .7

E
o
tJ
lO-S
10 .9

10-1o
+
+
Y 10 ~
10 .9

10-1o
+

1 0 -11 10-~1
I , I , I , I , I

0 20 40 60 80 100

nuclear charge Z
FIG. 2. Contributions to the g-factor of a bound electron in hydrogen-like ions for different nuclear
charges Z. (Courtesy Thomas Beier.)

Table I
Theoretical contributions to the g-factor in 12C5+

Contribution Size Reference

Dirac Theory 1.998 721 3544 Breit (1928)


QED, free (all orders) +0.002 319 3044 Hughes and Kinoshita (1999)
QED, bound, order (a/K) +0.000 000 844 6( 1) Yerokhin et al. (2002)
QED, bound, order ( a / ~ ) 2, (Za) 2 term -0.000 000 0012(3) Yerokhin et al. (2002)
Recoil (in Z a expansion) +0.000000087 6(1) Yelkovsky (2001 )
Finite size correction +0.0000000004 Beier et al. (2000)

been published by Beier (2000). A graphical representation of the bound-state


contributions to the electron g-factor is shown in Fig. 2.
In this contribution we describe an experiment which for the first time applies
the "continuous Stern-Gerlach effect" to an atomic ion (Hermanspahn et aL,
2000). We measured the magnetic moment of the electron bound to a nucleus
with zero nuclear magnetic moment, hydrogen-like carbon (12C5+). For the
bound-state contributions of order a(Za)" the existing calculations (Grotch,
1970a) deviate already for Z = 6 by as much as 10% from the non-perturbative
calculations to all orders in (Za). The numerical results of the calculations
for C 5+ are summarized in Table I. The leading term comes from the solution
II] CONTINUOUS STERN-GERLACH EFFECT ON ATOMIC IONS 195

of the Dirac equation and deviates from the value g = 2 for the free electron
(Breit, 1928). The next-largest part is the well-known QED contribution for the
free electron (Hughes and Kinoshita, 1999). The bound-state contributions of
order a (calculated to all orders in Z a) are given with error bars which represent
the numerical uncertainty of the calculations. The quoted uncertainty of the
aZ(za) 2 term is an estimate of the contribution from non-calculated higher-
order terms. Finally, nuclear recoil contributions have been calculated to lowest
order in Z a by Grotch (1970b), Faustov (1970) and Close and Osborn (1971).
Recently, Shabaev (2001) presented formulas for a non-perturbative calculation
(in Za), and Yelkovsky (2001) presented further numerical results. The nuclear-
shape correction was considered numerically by Beier et al. (2000), and recently
(for low Z) also analytically by Glazov and Shabaev (2001). The sum of the
different contributions leads to a theoretical value for the g-factor in hydrogen-
like carbon of
gtheor(12C5+) = 2.001 041 590 1 (3). (6)

II. A Single Ion in a Penning Trap

The experiment is carried out on a single C 5+ ion confined in two Penning


traps. In a Penning trap a charged particle is stored in a combination of a
homogeneous magnetic field B0 and an electrostatic quadrupole potential. The
magnetic field confines the particle in the plane perpendicular to the magnetic
field lines, and the electrostatic potential confines it in the direction parallel to
the magnetic field lines. In our experiment we use two nearly identical traps
placed 2.7 cm apart in the magnetic field direction. They consist of a stack of 13
cylindrical electrodes of 2r0 = 7 mm inner diameter. The difference between the
traps is that in one trap the center electrode is made of ferromagnetic nickel while
all others are machined from OFHC copper. Figure 3 shows a sketch of this setup.
The nickel ring distorts the homogeneity of the superimposed magnetic field in
the corresponding trap while the field remains homogeneous in the other trap
(see Fig. 3). As will become evident below, the inhomogeneity of the field is the
key element to analyze the direction of the electron spin through the continuous
Stern-Gerlach effect. Therefore we call the corresponding potential minimum
"analysis trap" while we call the one in the homogeneous magnetic field
"precision trap."
Each trap uses five of these electrodes to create a potential well, which serves
for axial confinement. We apply a negative voltage U0 to the center electrode
while we hold the two endcap electrodes at a distance z0 from the center at
196 G. Werth et al. [II

FIc. 3. Sketch of the electrode structure and potential distribution of the double trap.

ground potential. The potential 9 inside this configuration can be described in


cylindrical coordinates r , z , 0 by an expansion in Legendre polynomials Pi:

Uo ~ (~); (7)
(~)( r , I.~q) : -~- Z Ci -~ Pi(COS L~),
i=0

where d 2 - (z 2 + r 2 / 2 ) / 2 is a characteristic dimension of the trap (Gabrielse


et al., 1989). Two correction electrodes are placed between the center ring
and the endcaps. The coefficient (74 which is the dominant contribution to the
trap anharmonicity can be made small by proper tuning of the voltages applied
to the correction electrodes. Essentially then the potential depends on the square
of the coordinates, and is a harmonic quadrupole potential

Uo z 2 - r2/2
,I, - (8)
2 d2

We optimize the trap by changing the voltages on the correction electrodes until
the ion oscillation frequency is independent of the ion's oscillation amplitude
as characteristic for a harmonic oscillator. With this method we can reduce the
II] CONTINUOUS STERN-GERLACH EFFECT ON ATOMIC IONS 197

axial
motion

magnetron
m

T-B

cyclotron
motion
FIG. 4. Ion oscillation in a Penning trap.

dominant high-order term C4, the octupole contribution, to less than 10-5. The
ion's frequency in the harmonic approximation is then given by

qUo
O)z = Md 2 . (9)

Radial confinement is achieved by the homogeneous magnetic field directed


along the trap axis. This results in three independent oscillations (axial,
cyclotron, and magnetron oscillation) for the ion motion, as depicted in Fig. 4.
The fast radial oscillation frequency of the ion in the Penning trap is a perturbed
cyclotron frequency ~o~. It differs from the cyclotron frequency

o)~ = q B (10)
M
of a free particle with charge q and mass M because of the presence of the
electric trapping field, and is given by

, toe v/~O2 co2 (11)


coc = - ~ + 4 2

It can be expressed also as


~o~ = ayc - OOm, (12)

where (-/)mis the magnetron frequency, a slow drift of the cyclotron orbit around
the trap center, given by

Om-2 V7--5-" (13)


198 G. Werth et al. [II

For calibration of the magnetic field we use the cyclotron frequency of the
trapped ion. It can be derived either from Eq. (12) or more reliably from the
relation

since this equation is independent of trap misalignments to first order (Brown and
Gabrielse, 1986). In this case the measurement of the magnetron frequency (_gm
is required in addition to a measurement of to~ and COz.
The traps are enclosed in a vacuum chamber placed at the bottom of a helium
cryostat at a temperature of 4 K and located at the center of a superconducting
NMR solenoid. The helium cryostat provides efficient cryopumping. As an upper
limit we estimate the vacuum in the container to be below 10 -16 mbar. The
estimation was derived from the measurements on a cloud of highly charged
ions, whose storage time would be limited by charge exchange in collisions with
neutral background particles. We observed no ion loss in a cloud of 30 hydrogen-
like carbon ions stored for 4 weeks. Together with the known cross section for
charge exchange with helium as the most likely background gas at 4 K we obtain
an upper limit of 10 -16 mbar for the background gas pressure. The magnetic
field of the superconducting magnet is chosen to be 3.8 T. At this field strength
the precession frequency of the electron spin is 104 GHz. Microwave sources of
sufficient power and spectral purity are commercially available at this frequency.
We load the trap by bombarding a carbon-covered surface with electrons. This
process releases ions and neutrals of the element under investigation as well as
of other elements present on the surface. Higher charge states are obtained by
consecutive ionisation by the electron beam. We detect the ions by picking up
the current induced by the ion motions in the trap electrodes. For this purpose
superconducting resonant circuits and amplifiers are attached to the electrodes.
Upon sweeping the voltage of the trap, and thus the ions' axial frequencies, the
ions get in resonance with the circuit and their signal is detected. Figure 5 shows
such a spectrum, where we identified different elements and charge states. We
eliminate unwanted ion species by exciting their axial oscillation amplitude with
an rf field until the ions are driven out of the trap.
Ions of the same species have different perturbed cyclotron frequencies in
the slightly inhomogeneous magnetic field of the precision trap, because they
have different orbits. Therefore, for small ion numbers, single ions can be
distinguished by their different cyclotron frequencies. Figure 6 shows a Fourier
transform of the induced current from 6 stored ions. We excite the ions' cyclotron
motion individually and thus eliminate them from the trap until a single ion is
left. Typical cyclotron energies for signals as shown in Fig. 6 are of the order of
several eV.
In order to reduce the ion's kinetic energy we use the method of "resistive
cooling" which was first applied by Dehmelt and collaborators (Wineland and
Dehmelt, 1975; Dehmelt 1986). The ion's oscillation is brought into resonance
II] CONTINUOUS S T E R N - G E R L A C H EFFECT ON ATOMIC IONS 199

mass-to-charge ratio m/q


3,4 3,2 3,0 2,8 2,6 2,4 2,2 2,0
| ! , | l | , ,

0,4 a) with impurity ions


1606+

= 0,3

"~ 0,2

0,0 l : I , I , I , I , I , I , I ,

b) without impurity ions


=" 0,3

"~ 0,2
r

,w

~ o,1

0,0_ 14
trapping potential [Volt]
FIG. 5. Mass spectrum of trapped ions after electron bombardment of a carbon surface showing
different charge states of carbon ions as well as impurity ions (a) before and (b) after removal of
unwanted species.

with the circuits attached to the electrodes. The induced current through the
impedance of the circuit leads to heating of the resonance circuit, and the ion's
kinetic energy is dissipated to the surrounding liquid helium bath (Fig. 7). This
leads to an exponentially decreasing energy with a time constant r given by
q2
r -1 - R, (15)
Md 2
where R is the resonance impedance of the circuit.
For the axial motion we use superconducting high-quality circuits (Q = 1000 at
1 MHz in the precision trap and Q = 2500 at 365 kHz in the analysis trap).
With the resonance impedances of R = 23 M ~ (analysis trap) and 10 Mr2
(precision trap) the cooling time constants are 8 0 m s and 235 ms, respectively.
For cooling the cyclotron motion we employ a normal-conducting circuit at
24 MHz (Q = 400) with a resonance impedance of 80 kf~. Here we reach cooling
time constants of a few minutes. Figure 8 shows the exponential decrease of the
induced currents from the ion oscillations as the result of axial cooling.
200 G. Werth et al. [II

.~ 30

"f~ 25
6 C 5+ions I

~ o
, i
J , I , I ~ I

76000 76500 77000 77500 78000

cyclotron frequency [Hz] - 24 MHz


FIG. 6. Fourier transform of the voltage induced in one of the trap electrodes from the cyclotron
motion of 6 stored 12C5+ ions. The inhomogeneous magnetic field of the trap causes ions at different
positions to have slightly different cyclotron frequencies.

| Superc0nductang

l Cryogenic
Ion trap GaAsamphfier
I
FIG. 7. Principle of resistive cooling.

We do not cool the magnetron motion in a similar way because it is


metastable: in the radial plane the ion experiences an electrostatic force towards
the negatively biased center electrode. Ion loss is prevented by the presence of
the magnetic field. Thus the potential energy is an inverted parabola. Therefore
reduction of the ion's magnetron energy results in an increase in the magnetron
radius. It is, however, essential to reduce the magnetron radius because of
the magnetic field inhomogeneities. This is achieved in a well-defined way
by coupling the magnetron motion to the axial motion by a radio-frequency
field at the sum frequency of both oscillations (Brown and Gabrielse, 1986;
Cornell et al., 1990). In the quantum-mechanical picture for the ion motion, the
absorption of a photon from this field increases the quantum number of the axial
oscillation by 1 while that of the magnetron oscillation is decreased by 1. An
analysis of the absorption probabilities in the framework of a harmonic oscillator
II] CONTINUOUS STERN-GERLACH EFFECT ON ATOMIC IONS 201

t--
~

ms

t',-

T =4 K ~~,..,__....~
_

cooling time [s]


FIG. 8. Exponential energy loss of the axial motion of trapped C 5+ ions by resistive cooling.

yields that the quantum numbers tend to equalize. This leads to the expectation
value (Em) for the magnetron energy,

I<Em>I= boom ((Am) + 1) __ ho)m ((kz) + .~) 1

_ 1 (-Om (16)
o)m h ooz ( <kz ) -k- ~ ) = --~z ( Ez ) .
r

The axial oscillation is continuously kept in equilibrium with the cooling circuit
and we thus reduce the magnetron orbit to about 10 ~tm.
The mean kinetic energy of a single ion is often expressed in terms of
temperature. This is justified by the statistical equilibrium of the ion and the
resonant circuit. The statistical motion of the electrons in the resonance circuit
causes Johnson noise in the trap-electrode voltages, which in turn leads to
varying energies of the ion as a function of time (Fig. 9). Extracting a histogram
of the cyclotron energies results in a Boltzmann distribution (Fig. 10) with a
temperature of 4.9 K close to the temperature of the environment. Calculating
the temporal autocorrelation function of the energy gives, as expected, an
exponential (Fig. 11) with a time constant well in agreement with the measured
cooling time constant.
In order to calibrate the magnetic field at the ion's position with high precision
from Eq. (14) the three oscillation frequencies have to be measured. Because
of their different orders of magnitude (~Oc~/2;r = 24 MHz, Ogz/2:r = 1 MHz,
O)m/2~ 18 kHz) the required precision is different, o~ is determined from the
=

Fourier transform of the current induced in a split electrode. Figure 12 shows


that the relative linewidth of the resonance, well described by a Lorentzian, is of
the order of 10-9 and the center frequency can be determined with an accuracy
of 10-1~ In order to obtain sufficient signal strength the energy of the cyclotron
oscillation has to be raised to about 1 eV. Due to the inhomogeneity of the
202 G. Werth et al. [II

FIG. 9. Noise power of the induced voltage in a trap electrode from the cyclotron oscillation of
a single trapped ion while its frequency is continuously kept in resonance with an attached tank
circuit.

FIG. 10. Histogram of the probabilities for cyclotron energies. The curve can be well fitted to a
Boltzmann distribution, giving a temperature of 4.9(1) K.
II] CONTINUOUS STERN-GERLACH EFFECT ON ATOMIC IONS 203

%,
%,
%, C(t) = C o exp(-th:)
%,
%,
%,
".%
-c = 5.40 + 0.07 min
%,
%
"~,
-%
%
,~,

%
4
%.
".,,
%. e.

1~0
I ' I ' [ ' ' I '

0 2 4 15 8
Time [min]

FIG. 11. The time-correlation function of the noise in fig. 10 shows an exponential decrease. The
time constant o f 5 . 4 0 ( 7 ) m i n corresponds to the time constant for resistive cooling of the cyclotron
motion.

~
.,... 30

~ 25

2O
.,,..

~ lO
5

0
-- ' i 7 - - - 1 - , - - - ~ . . . . . . -w--- - T - - - "T~ ...... 1 i 1

-0.6 -0.4 -0.2 0.0 0.2 0.4 0.6

reduced cyclotron frequency [Hz] - 24 075 552.802 6 Hz

FIG. 12. High-resolution Fourier transform of the induced noise at the perturbed cyclotron
frequency. A Lorentzian fit gives a fractional width o f 1.4>< 10 -9.

magnetic field this changes the mean field strength along the cyclotron orbit.
This has to be considered in the final evaluation of the measurements.
The axial frequency OJz is determined while the ion is in thermal equilibrium
with the resonance circuit. At a given temperature the thermal noise voltage in
the impedance Z(e)) of the axial circuit, given by

Unoise = v/4kTRe[Z(oo)] 6v, (17)

excites the ion motion within the frequency range 6v of the ion's axial resonance.
This motion in turn induces a voltage in the endcap electrodes, however at a
phase difference of 180 ~ as can been shown by modeling the system as a driven
harmonic oscillator. Consequently the sum of the thermal noise voltage and
204 G. Werth et al. [II

FIG. 13. Axial resonance of a single trapped C5+ ion. The noise voltage across a tank circuit
shows a minimum at the ion's oscillation frequency.

FIG. 14. High-resolution Fourier transform of the axial noise near the center of the resonance
frequency of the axial detection circuit.

the induced voltage leads to a reduced total power around the axial frequency
of the ion. This appears as a m i n i m u m in the Fourier transform of the axial
noise as shown in Fig. 13. A spectrum with a resolution of 1 0 m H z (Fig. 14)
shows that the center frequency can be determined to about 24 mHz. A different
approach to explain the appearance of a m i n i m u m in the axial noise spectrum
was taken by Wineland and Dehmelt (1975) considering the equivalent electric
circuit of an oscillating ion in the trap. The magnetron frequency tOm is
measured by sideband coupling to the axial motion. If the ion is excited at
the difference between the axial and magnetron frequencies, the ion's axial
II] CONTINUOUS STERN-GERLACH EFFECT ON ATOMIC IONS 205

energy increases, leading to an increased current at the ion's axial frequency.


We detect this as a peak in the detection circuit signal. The uncertainty in this
frequency determination is below 100 mHz.
Imperfections in the trap geometry may change the motional frequencies.
These changes have been calculated by different authors (Brown and Gabrielse,
1982; Kretzschmar, 1990; Gerz et al., 1990; Bollen et al., 1990). Considering
only an octupole contribution to the trap potential, characterized by a coeffi-
cient Ca in the potential expansion (5), these shifts amount to

tOz ~ Ez-3 Ec +6Em, (18)

A6Oct-
tO,c C4~ 1 (~cc)2[-3Ez + -~3 (~cc)2Ec - 6Era] (19)

~Om qUo E~ +6Em . (20)

Ez, Ee and Em are the energies in the axial, cyclotron and magnetron degrees
of freedom. The tuning of the trap potential results in a coefficient C4
as small as 10-5 . For the energies of the motions in thermal equilibrium the
corresponding frequency shifts are below 10-l~ and need not be considered here.
The coefficient C6 of the dodekapole contribution to the trapping potential has
been calculated to an accuracy of 10-3 for our trap geometry. The frequency
shifts arising from this perturbation scale with (E/qUo) 2 and are negligible here.
The residual inhomogeneity of the magnetic field in the precision trap arising
from the nickel ring electrode in the analysis trap 2.7 cm away changes the value
of the oscillation frequencies of an ion with finite kinetic energy as compared
to the ion at rest. A series expansion of the B-field in axial direction,

Bz = Bo + Blz + B2z2 + " " (21)

gives frequency shifts

Am~_ 1 B2 1
mtO2z Bo hoOc
- ~ Ec+Ez+2E~j (22)

ACOz 1 B2 1
[Ee -Em], (23)
~Oz mm2z Bo hoe
Ao)m 1 B2 1
[2Ec - Ez - 2Era]. (24)
(Dm mm 2 Bo hmc

The size of the inhomogeneity term B2 is measured by application of a bias


voltage between the endcap electrodes. This shifts the ion's position in the axial
206 G. Werth et al. [III

direction by a calculable amount, and the cyclotron frequency is measured at each


position. We obtain B2 = 8.2(9)~tT/mm 2. The shift in the perturbed cyclotron
frequency which is of most interest here is dominated by the axial energy Ez,
and amounts to A~o~/o9~ = 7x 10 -9 for an axial temperature of 100 K.

III. Continuous Stern-Gerlach Effect

The g-factor of the bound electron as defined by Eq. (1) can be determined by
a measurement of the energy difference between the two spin directions in a
magnetic field B:
A E = hVL = g ~ s B , (25)

where VL is the Larmor precession frequency. We induce spin flips by applying


magnetic dipole radiation which is blown into the trap structure by a microwave
horn.
For the detection of an induced spin flip we follow a route developed in the
determination of the g-factor of the free electron (Dehmelt, 1986; Van Dyck
et al., 1986): the quadrupole potential of the Penning trap depends on the square
of the coordinates (6) leading to a linear force acting upon the charge of the
stored ion. Considering the force upon the magnetic moment of the bound
electron by the inhomogeneous field in the analysis trap we get

F - -27(g 9B). (26)

The nickel ring in the analysis trap creates a bottle-like magnetic field distortion
which can be described in first approximation by

B=B0+2B2 2 b-z . (27)

The odd terms vanish in the expansion because of mirror symmetry of the field.
The corresponding force on the magnetic moment in axial direction is

F= = - 2 1 z z B 2 z , (28)

which is linear in the axial coordinate. It adds to the electric force from the
quadrupole trapping field acting on the particle's charge. Since both forces are
linear in the axial coordinate the ion motion is still described by a harmonic
oscillator (Fig. 15). The axial frequency, however, depends on the direction of
the magnetic moment/t with respect to the magnetic field:

ltzB2
~ 6COz = % 0 + 9 (29)
% = %0 + ~ M (OzO
The value of B2 in our set-up was calculated using the known geometry and
magnetic susceptibility of the nickel ring electrode. We also determined it
III] C O N T I N U O U S S T E R N - G E R L A C H E F F E C T ON ATOMIC IONS 207

z-a~s

FIo. 15. Axial parabola potential for an ion in a quadrupole trap including the magnetic potential
for the spin-magnetic moment in a bottle-like magnetic field. The strength of the potential depends
on the spin direction. Upper curve: spin down; lower curve: spin up.

spin up
,~ ~ J , ---- ~---0.7Hz , r

, ,

i
i
_.

-10 ,
-5 ~ ,
0
I '
5
r , 1~0

axial frequency- 364 423.07 [Hz]

FIG. 16. Axial frequencies of a single C5+ ion for different spin directions. The averaging time
for each resonance line was 1 min.

experimentally by applying a bias voltage between the endcap electrodes o f the


analysis trap and measuring the cyclotron frequency of the ion at different axial
positions. The calculated and experimental values for B2 in our experiment agree
within their uncertainties of 10% and yield B2 - 1T/cm 2. For hydrogen-like
carbon the frequency difference &Oz/2:r between the two spin states amounts to
0.7 Hz at a total frequency of tOzO/2:r- 365 kHz.
As evident from Fig. 16, the axial frequency can be determined to better
than 100mHz. Fig. 17 demonstrates that after 1 min. averaging the expected
frequency difference between the two spin states becomes obvious. Driving
the spin-flip transition, we can distinguish the two possible axial frequencies,
0.7 Hz apart as calculated from the trap parameters. Varying the frequency o f
the microwave field and counting the number of induced spin flips per unit time
yields a resonance curve as shown in Fig. 18. The shape of this resonance is
asymmetric due to the inhomogeneity of the magnetic field. The general shape
of the Larmor resonance in an inhomogeneous magnetic field has been derived
208 G. Werth et al. [III

)" 1,0
quantum jumps

0,5 ~,.1 I 111 I l


0,0

i u~
down

oo

~ -1,o~

' go ' Ibo


time [min]

FI6. 17. Center of the axial frequency for a single C 5+ ion when irradiated continuously with
microwaves at the Larmor precession frequency showing two distinct values which correspond to
the two spin directions of the bound electron.

3O

E
E 20

2~

~L_
lO

E
m m m m
m m m m

I I I I I
-5 0 5 10 15
Larmor precession frequency- 103 958 [MHz]
FIG. 18. Number of observed spin flips per unit time vs. the frequency of the inducing field. The
solid line is a fit according to Eq. (29).

by Brown (1985) as a complex function of the trap parameters, the ion's energy
and the field inhomogeneity. However, assuming that the ion's amplitude z(t) is
IV] CONTINUOUS STERN-GERLACH EFFECT ON ATOMIC IONS 209

constant during the time 1/AwL, the line profile is given by a 6-function averaged
over the Boltzmann distribution of the energy:

Z ( ~~ ) =
io (
dE 6 ~oL - O)Lo
( '"))".-""
1 + m---~z
(30)

-- O(O)L
{ A--O09LO)
) L exp - (-OL--O')LO}Ao)L"

Here O(~oL - coL0) is the step function, which is 0 for o)L < coL0 and 1 for col > cot0,
and e is a linewidth parameter so that the Larmor frequency depends as
ooL = ~Oo(1 + ez 2) on the axial coordinate. A least-squares fit of this function
to the data points of Fig. 18 yields the Larmor frequency with a relative
uncertainty of 10-6 (Hermanspahn et al., 2000). This is sufficient to measure
the binding correction to the g-factor in C 5+. The bound-state QED corrections
for C 5+, however, are 4 • 10-7 and were not observed in this measurement.

IV. Double-Trap Technique


The limitation in accuracy of the experiment described above stems from the
inhomogeneous magnetic field as required for the analysis of the spin direction
via the continuous Stern-Gerlach effect. In fact the inhomogeneity of the field
was chosen to be as small as possible, but still large enough to be able to
distinguish the two spin directions.
We obtained an improvement of three orders of magnitude in the accuracy of
the measured magnetic moment by spatially separating the processes of inducing
spin flips and analyzing the spin direction (H/iffner et al., 2000). This is achieved
by transferring the ion after a determination of the spin direction from the
analysis trap to the precision trap. The voltages at the trap electrodes are changed
in such a way that the potential minimum in which the ion is kept is moved
towards the precision trap. The transport takes place in a time of the order of 1 s,
which is slow compared to any oscillation period of the ion and is therefore
adiabatic. Once in the precision trap, the ion's motional amplitudes are prepared
by coupling the ion to the resonant circuits. We then apply the microwave field
to induce spin flips. After the interaction time, typically 80 s, and an additional
cooling time, the ion is moved back to the analysis trap. Here the spin direction
is analyzed again. In principle one measurement of the axial frequency would be
sufficient to determine whether it has changed by 0.7 Hz as compared to the value
before transport into the precision trap. If, however, the ion is not brought back
with the same radial motional amplitudes to the analysis trap, the axial frequency
may have changed by as much as 1 Hz. This is because of the magnetic moment
connected with the cyclotron and magnetron motion. To circumvent this problem
210 G. Werth et al. [IV

spin flips in 1,0


0,4
analysis trap

0,2

~" 0,0
~ 0,4
-0,2
(D :' ~.~
", ,' =I 0,2
: :' ~r ~
o -0,4 ',, :,

, % 0,0 -
)$
..~
-0,6
-0,2
-0,8
i i i I i I I i i 1 i i 1 1 i
0 1 2 3 4 5 6 7 0 l 2 3 4 5 6 7
m i c r o w a v e excitations m i c r o w a v e excitations

FIG. 19. D e t e r m i n a t i o n o f the spin direction in the analysis trap after transport from the precision
trap. A change in axial frequency o f about 0.7 Hz indicates that the spin was up (left) or d o w n (right)
when the ion left the precision trap.

we induce an additional spin flip in the analysis trap to determine without doubt
the spin direction after return to the analysis trap. Figure 19 shows several cycles
for a spin analysis. The total time for a complete cycle is about 30 min.
While the ion is in the precision trap its cyclotron frequency COc = ( q / M ) B
is measured simultaneously with the interaction with the microwaves. This
ensures that the magnetic field is calibrated at the same time as the possible
spin flip is induced. The field of a superconducting solenoid fluctuates at the
level of 10-8-10 -9 on the time scale of several minutes. Figure 20 shows a
measurement of the cyclotron frequency of the ion in the precision trap over
a time span of several hours. Every 2 min the center frequency of the cyclotron
resonance was determined. The change in cyclotron frequency has approximately
a Gaussian distribution with a full-width-at-half-maximum of 1.2x 10 -8. This
may impose a serious limit on the precision of measurements as in the case of
high-precision mass spectrometry using Penning traps (Van Dyck et al., 1993;
Natarajan et al., 1993). However, the simultaneous measurement of cyclotron and
Larmor frequencies eliminates most of this broadening. Using Eqs. (10) and (25)
we obtain the g-factor as the ratio of the two measured frequencies
(DL m
g=2~--. (31)
mc M
The mass ratio of the electron to the ion can be taken from the literature. In our
case of 12C5+, Van Dyck and coworkers (Farnham et al., 1995) measured it with
high accuracy using a Penning trap mass spectrometer.
We measure the induced spin flip rate for a given frequency ratio of the
microwave field and the simultaneously measured cyclotron frequency. When we
IV] CONTINUOUS STERN-GERLACH EFFECT ON ATOMIC IONS 211

80-

7O

6O
e--

-I~ 50
.~.
._~ 4o
v/v = 1.2. 10-8I
30
Q.

20

10

0 A.~ I I I I I ' I ' I . . . . i. I

-0,6-0,5-0,4-0,3-0,2-0,1 0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7


change of reduced cyclotron frequency [Hz]

FIG. 20. Distribution of magnetic field values measured by the cyclotron frequency of a trapped
ion in a period of several hours. Data were taken every 2 min. The distribution is fitted by a Gaussian
with a full width of 1.2 x 10-8.

35

,-, 3O

~>' 25
.m

-0

.0 20
o
r
i,..

15
.m

.~ 10

~' ~ ~ ' I ' 1 ' I ' w , ~ I

-60 -40 -20 0 20 40 60 x l 0 6

Wmw/We - 4 3 7 6 . 2 1 0 499

Fie. 21. Measured spin-flip probability vs. ratio of Larmor and cyclotron frequencies. The data
are least-squares fitted to a Gaussian.

plot the spin flip probability, i.e. the n u m b e r o f successful at t empt s to c h a n g e the
spin direction d i v i d e d by the total n u m b e r o f attempts, we obtain a r e s o n a n c e line
as s h o w n in Fig. 21. The m a x i m u m attainable p r o b a b i l i t y is 50% w h e n the
a m p l i t u d e o f the m i c r o w a v e field is high e n o u g h . To avoid those saturation effects
we take care to k e e p the a m p l i t u d e o f the m i c r o w a v e field at a level that the
m a x i m u m p r o b a b i l i t y for a spin flip at r e s o n a n c e f r e q u e n c y is b e l o w 30%. In
addition we can take saturation into a c c o u n t using a s i m p l e r a t e - e q u a t i o n m o d e l .
212 G. Werth et al. [V

In contrast to the single-trap experiment the lineshape is now much more


symmetric. For a constant homogeneous magnetic field in the precision trap the
lineshape would be a Lorentzian with a very narrow linewidth determined by
the coupling constant ~, to the cooling circuit. However, the observed lineshape
can be well described by a Gaussian. The fractional full width is 1.1 • 10 -8. This
reflects the variation of the magnetic field during the time the ion spends in the
precision trap which is of the same order of magnitude (see Fig. 20). The line
center can be determined from a least squares fit to 1 x 10-l~

V. Corrections and Systematic Line Shifts


The main systematic shifts of the Larmor and cyclotron resonances arise from
the fact that the field in the precision trap is not perfectly homogeneous.
As mentioned above, the ferromagnetic nickel ring placed 2.7 cm away in
the analysis trap causes a residual inhomogeneity in the precision trap. The
expansion coefficient from Eq. (21) gives B2 = 8 ~tT/mm 2, three orders of
magnitude smaller than in the analysis trap. Therefore we still have to consider
an asymmetry in the line profile. Performing such an analysis gives a maximum
deviation as compared to the symmetric Gaussian fit of 2• 10 -1~ In addition,
the inhomogeneity of the magnetic field causes a shift of the line with the ion's
energies. In order to obtain a sufficiently strong signal of the induced current
from the cyclotron motion in the precision trap, the ion's energy has to be raised
to about 1 eV. This finite cyclotron energy has a large magnetic moment and thus
shifts the axial frequency as compared to vanishing cyclotron energy even in the
precision trap by about 1 Hz. To account for this shift we grouped our data of
the spin flip probabilities according to the different axial frequency shifts in the
precision trap corresponding to different cyclotron energies, and extrapolated
the ratios o)L/coc to zero cyclotron energy (Fig. 22). We find a slope of
A(~oL/~oc)/Ec = -1.09(5)• 10-9 eV -l . Other systematic shifts are less important:

0',
-10-
0
c-q -20-
t~

~" -30-

-40-

2 4 6 8 10 1 14 16

Cyclotron energy E+ [eV]

FIG. 22. Extrapolation of measured frequency ratios to vanishing cyclotron energy.


VI] CONTINUOUS STERN-GERLACH EFFECT ON ATOMIC IONS 213

Table II
Systematic uncertainties (in relative units) in the g-factor determination of 12C5+

Contribution Relative size

Asymmetry of resonance 2 x 10-10


Electric field imperfections 1 x 10- l 0
Ground loops in apparatus 4 x 10 -11
Interact. with image charges 3 x 10 -11
Calibration of cyclotron energy 2 x 1O-11
Sum 2.3 x 10- l 0

From the residual imperfection of the electric trapping field (C4 - l0 -5) we
calculate a shift of the cyclotron frequency of 1 x 10 -10. O f the same order of
magnitude are frequency shifts caused by changes of the trapping potential due
to ground loops when the computer controls are activated. The interaction of
the ion with its image charges changes the frequencies by 3 x 10-1~ but can be
calculated with an accuracy of 10%. Relativistic shifts are of the order of 10-l~ at
typical ion energies, but do not contribute to the uncertainty at the extrapolation
to zero energy. A list of uncertainties of these corrections is given in Table II.
The quadrature sum of all systematic uncertainties amounts to 3x 10-~~ The
final experimental value for the frequency ratio WL/OOCin 12C5+ is

COL
- 4376.2104989(19)(13). (32)
O9C

The first number in parentheses is the statistical uncertainty from the extrap-
olation to vanishing cyclotron energy, the second is the quadrature sum of the
systematical uncertainties. Taking the value for the electron mass in atomic units
(M(12C) = 12) from the most recent CODATA compilation (Mohr and Taylor,
1999) we arrive at a g-factor for the bound electron in 12C5+ of

gexp( 12 C 5+) = 2.001 041 596 3 (10)(44). (33)

Here the first number in parentheses is the total uncertainty of our experiment,
and the second reflects the uncertainty in the electron mass.

VI. Conclusions

A comparison of the experimentally obtained result of Eq. (33) to the theoretical


calculations presented in Table I shows that the bound-state QED effects of
214 G. Werth et al. [VII

order a / ~ in hydrogen-like carbon are verified at the level of 5 x 10-3. Bound


QED contributions of order (a/:r) 2 are too small to be observed. The nuclear
recoil part has been verified to about 5%.
It is believed that uncalculated terms of higher-order QED contributions do not
change the theoretical value beyond the presently quoted uncertainties. Taking
this for granted we can use experimental and theoretical numbers to determine
a more accurate value for the atomic mass of the electron, since this represents
by far the largest part in the total error budget (Beier et al., 2001). Using Eqs.
(6) and (33) we obtain from Eq. (31) the electron's atomic mass as

m = 0.000 548 579909 3(3). (34)

This is in agreement with the CODATA electron mass (Mohr and Taylor, 1999)
based on a direct determination by the comparison of its cyclotron frequency to
that of a carbon ion in a Penning trap (Farnham et al., 1995):

m : 0.000 548 579 911 0(12). (35)

VII. Outlook

The continuous Stern-Gerlach effect, using the frequency dependence of the


axial oscillation on the spin direction of an ion confined in a Penning trap
when an inhomogeneous field is superimposed, is a powerful tool to measure
magnetic moments of charged particles with great precision. This accurate
knowledge of magnetic moments is very important for tests of QED calculations.
The g - 2 experiment on free electrons by Dehmelt and coworkers (Van Dyck
et al., 1987) was a first example, followed now by the first application to an
atomic ion. The method described above is applicable to any ion having a
magnetic moment on the order of a Bohr magneton, provided it can be loaded
into the trap. For a given axial frequency and magnetic inhomogeneity B2, the
frequency splitting depends as 1/x/-qM on the mass M of the ion and its charge
state q (Fig. 23). This will impose technical limitations when working with
heavier hydrogen-like ions. Currently the stability of the electric trapping field
limits the maximal resolution of the axial frequency measurements: a jitter of
the trapping voltage by 1 ~tV, typical for state-of-the-art high-precision voltage
sources, induces frequency changes of 100 mHz for trap parameters as in our
case. However, materials with higher magnetic susceptibilities than nickel, such
as Co-Sm alloys, produce a larger magnetic inhomogeneity and therefore a
larger frequency splitting, allowing to proceed to heavier ions. In addition, the
induced magnetic inhomogeneity scales with the cube of the inverse radius of
the ring electrode. Thus a reduction in size of the analysis trap increases the
VII] CONTINUOUS S T E R N - G E R L A C H EFFECT ON ATOMIC IONS 215

2,8

2,4

2,o
rn--,I
t'q

1,6
N
<3
1,2

0,8

0,4
4He + 12C5+ 1607+ 9 9 9 9

t
9 9 n uu 9

o,o , t, ,t ,
; 8 ,'2 ' ,'6 2'0 ' 2'4 ' 2'8
(q * m)'/2

FIG. 23. Difference in axial frequency for two spin directions in a bottle-like magnetic field for
various hydrogen-like ions. The parameters B0 = 3.8T, B2 = I T/cm2, and tOz/2ar = 365kHz
are those of our experiment. The frequency difference scales linear with the magnetic field
inhomogeneity B2.

frequency difference for the two spin states significantly. This would also have
the advantage that it reduces the amount of ferromagnetic material placed in the
analysis trap, and so helps to improve the homogeneity in the precision trap. To
further improve the homogeneity in the precision trap the distance between the
two traps can be increased. Finally, shim coils may be used to make the field
in the precision trap more homogeneous. We believe that we can maintain the
presently achieved precision with other ions as well, and hope to even increase
it when we apply some of the measures for improvement. This would result in a
more significant test of higher-order bound-state QED contributions since they
increase quadratically with the nuclear charge (see Fig. 2).
The method can also be applied to more complicated systems: a measurement
of the electronic g-factor in lithium-like ions would test not only the QED cor-
rections in these systems but also correlation effects with the remaining electrons
which change the g-factor significantly. When applied to hydrogen-like ions with
non-zero nuclear spin the transition frequencies between spin states depend on
the nuclear magnetic moment. Measuring the different transition frequencies
yields the magnetic moment of the nucleus. This would be of special interest,
because all nuclear magnetic moments so far have been determined using neutral
atoms or singly ionized ions. The effective magnetic field seen by the nucleus in
these systems differs from the applied magnetic field by shielding effects of the
electron cloud. In a measurement on hydrogen-like ions this shielding is strongly
216 G. Werth et al. [IX

reduced, and comparison with data obtained on neutral systems would, for the
first time, test atomic-physics calculations on electron shielding.

VIII. Acknowledgements

The measurements described above are performed in close collaboration to


GSI/Darmstadt. We gratefully acknowledge financial support from its Atomic
Physics group (Prof. H.-J. Kluge). Several doctoral and diploma students were
and are actively involved in the experiments: Stefan Stahl, Nikolaus Her-
manspahn, Jose Verdfi, Tristan Valenzuela, Slobodan Djekic, Michael Diederich,
Markus Immel, and Manfred T6nges. We appreciated stimulating discussions
with our colleagues: Thomas Beier, Andrzej Czarnecki, Ingvar Lindgren, Savely
Karshenboim, Vasant Natarajan, Hans Persson, Sten Salomonson, Vladimir
Shabaev, Gerhard Soft, Alexander Yelkovsky, and Vladimir Yerokhin.
The experiments are part of the TMR network ERB FMRX CT 97-0144
"EUROTRAPS" of the European Community.

IX. R e f e r e n c e s

Batelaan, H., Gay, T.J., and Schwendiman, J.J. (1997). Phys. Rec. Lett. 79, 4517.
Beier, Th. (2000). The g j factor of a bound electron and the hyperfine structure splitting in
hydrogenlike ions. Phys. Rep. 339, 79-213.
Beier, Th., et al. (2000). The g-factor of an electron bound in a hydrogenlike ion. Phys. Rev. A
62, 032510, pp. 1-31.
Beier, Th., et al. (2001). New determination of the electron's mass. Phys. Rev. Lett. 88, 011603-1-4.
Bloch, E (1953). Experiments on the g-factor of the electron. Physica 19, 821-831.
Bohr, N. (1928). The magnetic electron. Collected Works of Niels Bohr (J. Kalckar, Ed.), Vol. 6.
North-Holland, Amsterdam, p. 333.
Bollen, G., et al. (1990). The accuracy of heavy-ion mass measurements using time of flight-ion
cyclotron resonance in a Penning trap. J Appl. Phys. 68, 4355-4374.
Breit, G. (1928). The magnetic moment of the electron, Nature 122, 649.
Brillouin, L. (1928). Proc. Natl. Acad. Sci. U.S.A. 14, 755.
Brown, L.S. (1985). Geonium lineshape. Ann. Phys. 159, 62-98.
Brown, L.S., and Gabrielse, G. (1982). Precision spectroscopy of a charged particle in an imperfect
Penning trap. Phys. Rec. A 25, 2423-2425.
Brown, L.S., and Gabrielse, G. (1986). Geonium Theory: Physics of a single electron or ion in a
Penning trap. Rev. Mod. Phys. 58, 233.
Close, EE, and Osborn, H. (1971). Relativistic extension of the electromagnetic current for composite
systems. Phys. Lett. B 34, 400-404.
Cornell, E.A., et al. (1990). Mode coupling in a Penning trap: :r pulses and a classical avoided
crossing. Phys. Rev. A 41, 312-315.
Dehmelt, H. (1986). Continuous Stern-Gerlach effect: Principle and idealized apparatus. Proc. Natl.
Acad. Sci. US.A. 53, 2291.
Dehmelt, H. (1988). New continuous Stern-Gerlach effect and the hint of "the" elementary particle.
Z. Phys. D 10, 127-133.
IX] CONTINUOUS STERN-GERLACH EFFECT ON ATOMIC IONS 217

Farnham, D.L., Van Dyck, R.S., and Schwinberg, P.B. (1995). Determination of the electron's atomic
mass and the proton/electron mass ratio via Penning trap mass spectrometry. Phys. Rev. Lett. 75,
3598-3601.
Faustov, O. (1970). The magnetic moment of the hydrogen atom, Phys. Lett. B 33, 422-424.
Gabrielse, G., Haarsma, L., and Rolston, S.L. (1989). Open endcap Penning traps for high-precision
experiments. Int. J. Mass Spectrosc. Ion Proc. 88, 319-332.
Garraway, B.M., and Stenholm, S. (1999). Observing the spin of a free electron. Phys. Rev. A 60,
63-79.
Gerz, Ch., Wilsdorf, D., and Werth, G. (1990). A high precision Penning trap mass spectrometer.
Nucl. Instrum. Methods B 47, 453-461.
Glazov, D.A., and Shabaev, V.M. (2001). Finite nuclear size correction to the bound-state g factor
in a hydrogenlike atom, Phys. Lett. A 297, 408-411.
Grotch, H. (1970a). Electron g factor in hydrogenic atoms. Phys. Rev. Lett. 24, 39-45.
Grotch, H. (1970b). Nuclear mass correction to the electronic g factor. Phys. Rev. A 2, 1605-1607.
H/iffner, H., et al. (2000). High-accuracy measurement of the magnetic moment anomaly of the
electron bound in hydrogenlike carbon. Phys. Rev. Lett. 85, 5308-5311.
Hermanspahn, N., et al. (2000). Observation of the continuous Stern-Gerlach effect on an electron
bound in an atomic ion. Phys. Rev. Lett. 84, 427-430.
Hughes, V.W., and Kinoshita, T. (1999). Anomalous g values of the electron and muon. Rev. Mod.
Phys. 71, 133-139.
Karshenboim, S., Ivanov, V.G., and Shabaev, V.M. (2001). Can. J. Phys. 79, 81-86.
Kretzschmar, M. (1990). A theory of anharmonic perturbations in a Penning trap. Z. Naturf 45a,
965-978.
Lindroth, E., and Ynnerman, A. (1993). Ab initio calculations of g j factors for Li, Be +, and Ba +.
Phys. Rev. A 47, 961-970.
Mohr, P.J., and Taylor, B.N. (1999). CODATA recommended values of the fundamental physical
constants: 1998, J. Phys. Chem. Ref Data 28, 1713-1852.
Natarajan, V., et al. (1993). Precision Penning trap comparison of nondoublets: atomic masses of H,
D, and the neutron. Phys. Rev. Lett. 71, 1998-2001.
Pauli, W. (1958). Prinzipien der Quantentheorie. In "Handbuch der Physik" (S. Fliigge, Ed.), Vol. 5.
Springer, Berlin, p. 167.
Shabaev, V.M. (2001). QED theory of the nuclear recoil effect on the atomic g factor. Phys. Rev. A
64, 052104-1-14.
Stern, O., and Gerlach, W. (1922). Der experimentelle Nachweis der Richtungsquantelung im
Magnetfeld, Z Phys. 9, 349-352.
Van Dyck, R.S., Schwinberg, P.B., and Dehmelt, H.G. (1986). Electron magnetic moment from
Geonium spectra: Early experiments and background concepts. Phys. Rev. D 34, 722-736.
Van Dyck, R.S., Schwinberg, P.B., and Dehmelt, H.G. (1987). New high-precision comparison of
electron and positron g factors. Phys. Rev. Lett. 59, 26-29.
Van Dyck, R.S., Farnham, D.L., and Schwinberg, P.B. (1993). Tritium-helium-3 mass difference
using the Penning trap mass spectroscopy. Phys. Rev. Lett. 70, 2888-2891.
Veseth, L. (1980). Spin-extended Hartree-Fock calculations of atomic g j factors. Phys. Rev. A 11,
421-426.
Veseth, L. (1983). Many-body calculations of atomic properties: I. g j factors. J. Phys. B 16,
2891-2912.
Wineland, D.J., and Dehmelt, H.G. (1975). Principles of the stored ion calorimeter. J. Appl. Phys.
46, 919-930.
Yelkovsky, A. (2001). Recoil correction to the magnetic moment of a bound electron. E-print archive,
hep-ph/O108091 (http://xxx.lanl.gov).
Yerokhin, V.A., Indelicato, P., and Shabaev, V.M. (2002). Self-energy correction to the bound-electron
g factor in H-like ions. E-print archive, physics~0205245 (http://xxx.lanl.gov).
This Page Intentionally Left Blank
A D V A N C E S IN ATOMIC, M O L E C U L A R , A N D O P T I C A L PHYSICS, VOL. 48

THE CHIRALITY OF BIOMOLECULES


R O B E R T N. C O M P T O N 1,2 a n d R I C H A R D M. P A G N I 1
IDepartment of Chemistry and 2Department of Physics, University of Tennessee,
Knoxville, Tennessee 37996

I. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
II. Fundamental Nature of Chirality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
III. True and False Chirality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
IV. Galaxies, Plants, and Pharmaceuticals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
V. Plausible Origins of Homochirality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
VI. A s y m m e t r y in Beta Radiolysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
A. Amplification and Degradation M e c h a n i s m s . . . . . . . . . . . . . . . . . . . . . . . 247
VII. Possible Effects of the Parity-Violating Energy Difference (PVED) in
Extended Molecular Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
VIII. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
IX. Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
X. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258

I. Introduction

It is often said that the feature that distinguishes the physical sciences from the
biological sciences is the precision with which the terms and concepts in each
field are defined and measured. Terms in the physical sciences are precisely
defined while those in the biological sciences are often much less so. Even the
definition of life, which is the sine qua non of biology, is difficult to pin down in
all cases. Is a virus or a prion living, for example? Nonetheless, there are features
that all living things have in common. Most of the molecules that serve structural
and functional roles in all living systems exist in only one of two seemingly
identical forms. This property is called homochirality, and the molecules that
possess it are said to be optically active. This chapter will investigate the present
state of understanding of how homochirality in living systems may have come
to be on the earth. We refer to the fact that overwhelmingly only one of these
identical forms is found in life. Before doing this, however, it will be necessary to
find out what homochirality means and what are the properties of the molecules
that potentially posses it.

II. Fundamental Nature of Chirality

Chirality is a term familiar to physicists, mathematicians and chemists. What fol-


lows below is a condensed description of chirality from a chemist's perspective.

219 Copyright 9 2002 Elsevier Science (USA)


All rights reserved
ISBN 0-12-003848-X/ISSN 1049-250X/02 $35.00
220 R.N. Compton and R.M. Pagni [II

FIG. 1. Left and right hands.

A more detailed discussion may be found in the excellent book on the subject
by Eliel et al. (1994).
The English word chirality is derived from the Greek word Zetp (kheir)
meaning hand, and refers to an interesting geometrical property of left and right
hands (and feet) and other three-dimensional objects. Neither the left nor the
right hand has any internal symmetry but both are related to each other via a
mirror plane. If you take your left hand and reflect it into a mirror, what you
see in the mirror is an image of your right hand (Fig. 1). Likewise, reflection
of your right hand into the mirror also yields an image of your left hand. All
objects (except vampires) have a mirror image. In most cases, such as with
spheres (balls) and planes (pieces of papers), the mirror image is superposable
or congruent with the original object. This means that the mirror image object
is identical in every respect (constituents and geometry) to the original. What
makes the relationship of the left and right hands different is the fact that they
cannot be superposed one on top of the other even though the connectivities of
the fingers to the palms are the same in both cases. No matter how hard one
tries to attain the superposition, there will always be mismatches. The left and
right h a n d s - and any other objects with non-superposable mirror i m a g e s - are
said to be chiral.
There are myriad objects in the physical world that are chiral. Coils, screws
and springs exist in chiral forms (right- and left-handed helices), for example.
Perhaps the most well-known of these is the alpha (right-handed) helix of DNA.
II] THE CHIRALITY OF BIOMOLECULES 221

H H
H\ .... H
CH4 HmC--H
H / C "~H
I
H

formula bond connectivities; tetrahedron representa- shorthand tetrahe-


no three-dimensional tion; three-dimensional dral representation
structure implied structure implied

FIG. 2. Various representations of the molecule CH 4.

On a larger scale, vines coil up (twine) trees and shrubs in one of the two
mirror image forms whose direction of coiling (helicity) is species dependent.
When only one of the two mirror image forms exists, the system is said to be
homochiral. Homochirality at all size scales is a hallmark of biology. Many
seashells also coil in only one direction, with the direction of coiling being
species dependent. Spiral staircases are chiral as well. The massless photon with
a unit spin and moving in a straight line at the speed of light carries a spin
angular momentum (+h). Thus, a single photon can be thought of as "right-" or
"left-" handed. Linearly polarized light is an equal mixture (racemic) mixture of
these photons. Electrons also have one-half integer spin, but at rest the electron
and its mirror image are superposable. When electrons are in motion, their
mirror images are not superposable, thus yielding spin-polarized particles if the
z component of the spin angular momentum can be along or against the direction
of propagation of the electron. Thus, for elementary particles such as photons
and electrons, it is the combination of spin and translation that is responsible for
their chirality (physicists speak of helicity) and circular polarization.
Many molecules, including biological ones, are chiral. To see how this is possi-
ble, consider first the structure and properties of the simplest organic molecule,
methane (CH4) (Fig. 2). This molecule consists of a central carbon atom (C)
connected or bonded to four hydrogen atoms (H). Each bond connecting a carbon
atom to a hydrogen atom is represented by a straight line. When a carbon atom
is bonded to four other atoms, it exists in a tetrahedral environment in which the
carbon resides in the center of an imaginary Platonic tetrahedron and the four hy-
drogen atoms reside at apices of the tetrahedron. This yields a three-dimensional
structure with four identical carbon-hydrogen bond lengths and six identical
hydrogen to carbon to hydrogen bond angles, in agreement with experiment.
Rather than draw a cumbersome tetrahedron every time a structure contains a
tetrahedral carbon atom, it is convenient to draw it in a different, simplified
manner: In this picture of methane two straight lines from C to H represent bonds
in the plane of the paper, while a solid triangle represents a carbon-hydrogen
bond projecting from the paper, where the carbon resides, to the hydrogen in front
of the paper and a dashed triangle represents a C - H bond projecting behind the
paper. Methane is thus a highly symmetric molecule with Td symmetry. It should
222 R.N. Compton and R.M. Pagni [II

H~ ....,\\ H Ho, t,,. /H

i _ i / c "~H H~'~C ~ H

FIG. 3. Methane and its mirror image.

H H H H
I I I I
HmC ~H ~ H~G ~F H--CmF ~ Br--CmF
I / I I
H H CI C1

CHIRAL

FIG. 4. Methane and several of its derivatives.

be noted, however, that the picture of methane with Ta symmetry represents the
time-averaged or equilibrium structure. As the bonds in molecules are constantly
vibrating, an instantaneous picture of methane would be of lower symmetry.
Now consider methane as was done above for left and right hands, i.e., reflect
the image of methane into a mirror (Fig. 3). It is not difficult to see, either in
one's mind or with models, that the mirror image object is superposed exactly
on the original methane molecule. Methane thus lacks chirality and is said to be
achiral.
What happens when one of the hydrogen atoms in methane is replaced with a
different atom (or a group of atoms)? Consider fluoromethane, CH3F, where the
replacement is a fluorine atom (Fig. 4). This molecule also contains a tetrahedron
carbon atom at its center but with C3~ symmetry has lower symmetry than
methane. Nonetheless, CH3F and its mirror image are superposable on one
another and the molecule is achiral. Likewise, when a second atom is replaced
with a still different atom (or a group of atoms), an achiral molecule results. For
example, chlorofluoromethane (CHzC1F), where the second replacement is with
a chlorine atom (C1), is also achiral.
When a third replacement is made with a still different substituent, a chiral
molecule is formed. Bromochlorofluoromethane, CHBrC1F, where the third
substituent is a bromine atom (Br), is a fascinating example of such a molecule
(Fig. 5). Here the mirror image molecule cannot be superposed on the original.
It is always possible to match up the carbon and two of its attached atoms (and
their bonds) in the two structures, but the other two attached atoms (and their
bonds) will always be out of alignment. Thus CHBrC1F is chiral.
II] THE CHIRALITY OF BIOMOLECULES 223

H\ F/,..... / H

J C -,.~ICI CIw~'" C ~ B r
Br~

FIG. 5. Bromochlorofluoromethane and its mirror image.

It is worth noting that the achiral CH4, CH3F and CH2FC1 all have one or more
planes of symmetry while the chiral CHBrC1F does not. It can be shown that
all achiral molecules have at least one improper axis of symmetry Sn, defined
as rotation by 2sr/n radians followed by reflection in a plane perpendicular to
the rotation axis. A plane of symmetry is thus equivalent to S I and a center of
symmetry is equivalent to $2.
The two forms of CHBrC1F are stereoisomers of one another, i.e., they have
the same connectivities of atoms and yet the very same atoms are oriented
differently in space. The carbon atoms in the stereoisomers of CHBrC1F are
called stereogenic centers. Stereoisomers such as the two forms of CHBrC1F
that are also mirror images of one another are called enantiomers. Interestingly,
enantiomers have identical chemical properties except when they react with other
molecules which are also enantiomeric; reaction or interaction with chiral forces
may also yield a difference in behavior. Enantiomers also have identical physical
properties except in the way they interact with plane-polarized or circularly
polarized light or other chiral objects. From this observation it would appear to
be difficult to separate, i.e., resolve, mixtures of enantiomers into their individual
forms. Fortunately, as described below, there are methods to accomplish this
task.
If one has samples of each pure enantiomer, one enantiomer would rotate
the plane of monochromatic plane-polarized light a given number of degrees
in the clockwise direction while the other enantiomer would rotate the light
the same number of degrees but in the counterclockwise direction. The device
in which these measurements are made is called a polarimeter. Molecules that
rotate plane-polarized light are said to be optically active, and the wavelength
dependence of the optical activity is called optical rotary dispersion (ORD).
A collection of achiral molecules does not rotate light and is thus said to be
optically inactive. An individual achiral molecule can also rotate the plane of
polarization, but the rotation averaged over many non-oriented molecules goes
to zero. A 1:1 mixture of two enantiomers, which is called a racemic mixture,
also is optically inactive because the rotation of one enantiomer is canceled by
224 R.N. Compton and R.M. Pagni [II

the opposite rotation of the other enantiomer. The enatiomeric excess is defined
as the difference in amounts of the two enantiomers in a sample divided by their
sum times 100. A pure enantiomer will have an enatiomeric excess of 100% and
is thus homochiral; a racemic mixture will have an enantiomeric excess of 0%.
Enantiomeric excess is often abbreviated as ee.
CHBrC1F is prepared in racemic form by the reaction of achiral CHBrzC1 with
achiral HgF2 (Hine et al., 1956):

2CHBr2C1 + HgF2 --~ 2CHBrC1F + HgBr2.

Generally speaking when achiral or racemic compounds react with one another,
the resulting chiral products will be racemic. Only when a r e a g e n t - or
c a t a l y s t - is optically active will a chiral product be optically active as well.
Chiral forces such as circularly polarized light may also induce optical activity
(enantiomeric excess) in the product as well. Chiral reagents, chiral catalysts
or chiral forces are required to bring about enantiomeric excess in reaction
products. Thus, at first glance it is difficult to see how optically active biological
compounds such as amino acids and carbohydrates came to be on the earth
when they resulted in the first place from the chemical combination of achiral
molecules.
CHBrC1F can be prepared in optically active or resolved form in two ways:
(1) by the reaction of a precursor that is already optically active (Doyle and
Vogl, 1989) and (2) by the reaction of a chiral but optically inactive precursor
in a chiral, optically active environment (Wilen et al., 1985). It always takes an
optically active molecule or chiral force to produce a product that is optically
active.
There is nothing unique about CHBrC1E Any molecule containing a tetrahe-
dral carbon or other central atom such as silicon (Si) bonded to four different
substituents, regardless of their nature, will exist in enatiomeric forms. Thus the
property of chirality, which is manifested in chemical compounds, is in reality
geometric in nature. Many other geometric orientations of molecules in space
also yield enantiomeric forms.
CHBrC1F (and like molecules) has permanent chirality because its stereogenic,
tetrahedral carbon atom is attached to four different substituents. The innate
chirality of CHBrC1F does not depend on the lengths of its four bonds or
the fact that the lengths are constantly changing due to molecular vibrations.
CH4, on the other hand, behaves differently. It is achiral because its averaged
equilibrium structure of Td symmetry possesses six planes of symmetry and
six $4 axes of symmetry as well. Because it is vibrating, any instantaneous
structure of CH4 lacks Td symmetry and may be chiral if all four bond lengths are
different. In another moment this chiral structure will disappear and be replaced
by another, perhaps with a mirror-image relationship to the first. This constantly
shifting chirality will disappear when averaged through an ensemble of methane
II] THE CHIRALITY OF BIOMOLECULES 225

molecules, thus yielding a net optical rotation of zero degrees. Even crystals and
liquids made up of achiral molecules may display instantaneous chirality which
averages to zero.
Monochromatic, linearly polarized light (LPL) may be viewed as a su-
perposition of left-handed and right-handed circularly polarized light (CPL)
(Michl and Thulstrup, 1986). Alternately, one could describe LPL as a racemic
mixture of photons with spin angular momentum oriented along and against the
direction of propagation. When LPL passes through a sample of enantiomers,
one circular component of the light passes through the medium faster than the
other component because the real parts of the refractive indices, nL and ne,
of the left-handed and right-handed CPL through the molecule are different
(Atkins, 1978). This is the origin of optical activity. The angle of rotation
is equal to ( n L - nR)Jrl/~,, where l is the length of the sample and Jl is the
wavelength of the light. If the rotation of the plane of polarization is clockwise
as seen by an observer looking at the source of light, this is designated in
two ways: (1) with a + sign or (2) with a lower case "d" which stands for
dextrorotatory. Counterclockwise rotation is designated with a - sign or with
a lower case 'T' which stands for levorotatory. One enantiomer of CHBrC1F
will thus be (+)-CHBrC1F and the other (-)-CHBrC1F or d-CHBrC1F and
1-CHBrC1E Unfortunately, there is no simple method to relate the direction of
the optical rotation that the ensemble of enantiomers generates with the absolute
configuration of the enantiomer. Absolute configuration is the property of the
enantiomer that distinguishes the orientation of its substituents in space from
that of its mirror image partner.
Amino acids are the building blocks of peptides, proteins and enzymes which
serve an extraordinary number of structural and catalytic roles in cells. Enzymes,
for example, are the biological agents that catalyze virtually all cellular reactions,
most often with amazing speed and selectivity. Structurally, an amino acid that
is used biologically consists of an amino group NH2, consisting of a nitrogen
atom N bonded to two hydrogen atoms, and a carboxylic acid group COOH,
consisting of a carbon atom b o n d e d - actually double-bonded- to one oxygen
atom O and a hydroxyl group OH, and to a central CHR group, where R varies
in structure. As there are 20 amino acids commonly used in cellular chemistry,
there are 20 different R groups. All of the amino acids except glycine with R = H
contain a stereogenic center at the CHR carbon atom. Two of the amino acids,
isoleucine and threonine, contain a second stereogenic center in the R group
itself.
H
I H2N~CHR~COOH
H2N~ l ~COOH
R
AMINO ACID
226 R.N. Compton and R.M. Pagni [II

..... //COOH
H l, HOOC ....,,\H
CH3W~"C ~ H2N/C "~CH3
NH2
one enantiomer of alanine other enantiomer of alanine

I rotate rotate
structure structure

COOH COOH
_

H2ND-- C ~ H H~ ' - C ~ N H 2
_

CH3
_

CH3
I project I project
into paper into paper

COOH COOH
H2N~
I
C ~H
I
H~ C ~NH2
I
CH3 I
CH3
L-alanine D-alanine

FIG. 6. The enantiomers of alanine.

To see the consequences of the amino acids having a stereogenic center at the
CHR group, consider alanine with R = CH3 (methyl group, where a carbon atom
is bonded to three hydrogen atoms) (Fig. 6). With one stereogenic center alanine
exists in two enantiomeric forms, as shown in Fig. 6. It is common to draw the
enantiomers of the amino acids in the so-called Fischer projection, named after
Emil Fischer, the great German chemist. Here the COOH and CH3 groups of
one of the enantiomers are aligned vertically with the COOH group at the top.
This forces the hydrogen and NH2 group attached to the stereogenic center to
occupy horizontal positions. If one imagines the stereogenic carbon to be in the
plane of the paper, the two vertical groups lie behind the paper and the two
horizontal groups lie in front. At this point the vertical and horizontal groups
and their bonds are projected onto the plane of the paper, resulting in the Fischer
projection of one of the enantiomers of alanine. The enantiomer which places the
NH2 group to the left of the vertical axis in the Fischer projection is assigned the
L absolute configuration. The other enantiomer, with the NH2 group to the right
of the vertical axis, is assigned the D absolute configuration. All 19 amino acids
with a stereogenic center at the CHR carbon thus have L and D enantiomers. It
II] THE CHIRALITY OF BIOMOLECULES 227

is important to note that the upper case D and L refer to absolute configuration
while the lower case "d" and "1" refer to the clockwise/counterclockwise rotation
of plane polarized light through the enantiomers; knowing one quantity does
not tell you anything about the other. A strange fact of living material is the
almost exclusive use of L-amino acids in cellular chemistry. The origin of this
homochirality is difficult to understand if one assumes that the amino acids were
synthesized on the primitive earth from achiral and racemic compounds without
any apparent intervening chiral force. Chiral products generated from achiral
or racemic compounds are themselves racemic. Homochirality is an accepted
tenant in any theory involving the origins of life. Science continues to ponder
the seemingly mystical question of the origins of specific homochirality: Did it
occur by chance or some fundamental bias?
Carbohydrates or sugars are a second class of important biological molecules.
Most people are familiar at least with the names of the carbohydrates glucose,
fructose, lactose, and sucrose (table sugar). Although they generally serve as
energy sources in cells, they have other functions as well. Glucose, for example,
is the building block for both cellulose, the fibrous material of plants, and
starch. Other carbohydrates can be found in cell walls. Ribose and deoxyribose,
two other carbohydrates, are building blocks of the nucleic acids which are
responsible for the genetic code and ultimately for all of the chemistry that takes
place in cells.
The name carbohydrate is a misnomer as it implies a hydrated form of carbon,
i.e., a compound in which carbon is bonded to water. This arises because sugars
have formulas of the type Cn(HzO)m, where H20 is the formula for water, and
n and m are integers. Instead, most carbohydrates are polyhydroxyaldehydes;
here hydroxy implies an OH group and aldehyde a CHO group, which can
only occur at the terminus of a molecule. Thus carbohydrates are aldehydes
containing two or more, thus poly, OH groups, one at each of the non-aldehydic
carbons: HOCHz(CHOH)nCHO. Because most carbohydrates contain more than
one stereogenic center, it is important to see what stereochemical consequences
there are for such molecules.
Consider the two stereoisomers of the carbohydrate 2,3,4-trihydroxybutanal
(HOCHzCHOHCHOHCHO) in Fischer projection: D-erythrose and L-erythrose,
each with stereogenic centers at carbons 2 and 3 (Fig. 7). These mirror-image
isomers are enantiomeric because they are non-superimposable. The enantiomer
at the left has the D absolute configuration because its OH group at carbon 3
(the one attached to the stereogenic center farthest removed from the aldehyde
group) is to the right of the vertical axis of carbon atoms, while its partner to
the right has the L absolute configuration because its carbon-3 OH group is to
the left of the vertical axis. As D- erythrose and L-erythrose are enantiomeric,
their properties and relationships are identical to those described above for
enantiomers with one stereogenic center.
228 R.N. Compton and R.M. Pagni [II

CHO OHC
I I
H-- C2 --OH HO-- C2--H
I I
H--C3--OH HO-- ~3--H
I
CH2OH HOH2C
D-erythrose ~L-erythrose

FIG. 7. The enantiomers of the carbohydrate erythrose.

CHO OHC
[ I
HO-- C2--H H-- C2--OH
I I
H--~3--OH HO-- ~3--H
CH2OH HOH2C
D-threose L-threose

FIG. 8. The enantiomers of the carbohydrate threose.

There are two other stereoisomers that have the same bonding pattern as
D-erythrose and L-erythrose (Fig. 8). One can imagine forming them from the
erythrose enantiomers by interchanging the H and OH groups at carbon 2. The
resulting structures, which are called D-threose and L-threose, are enantiomeric
with one another because, even in Fischer projection, it is easy to see that the
molecules are mirror images of one another and yet are not superimposable.
As with D-erythrose, D-threose has the D absolute configuration because its
OH group at carbon 3 is to the right of the vertical axis of carbon atoms, while
its mirror image partner has the L absolute configuration because its carbon-3
OH group is to the left of the axis.
What is new for molecules that have two stereogenic centers is the relationship
of the erythrose enantiomers to the threose enantiomers. They are clearly
stereoisomers of one another because the bond connectivities in the erythrose
isomers are the same as those in the threose isomers but the atoms are oriented
differently in space. Furthermore, neither erythrose isomer is superposable
on or has a mirror-image relationship with the threose enantiomers and vice
versa. Stereoisomers that do not have a mirror-image relationship are called
II] THE CHIRALITY OF BIOMOLECULES 229

D-threose L-threose D-erythrose L-erythrose

COOH HOOC COOH HOOC


I I I I
HOmCmH HmC~OH H--C ~OH HOmC--H
I I ....... f . . . . . . . .
....... ~ .......

H~CmOH HO~C~H HmCmOH HO--CmH


I I I I
COOH HOOC COOH HOOC

1-tartaric acid d-tartaric acid meso-tartaric acid


(dashed line represents the
plane of symmetry $1)

FIG. 9. Conversion of threose and erythrose into tartaric acid.

diastereomers. Thus, the erythrose enantiomers are diastereomeric with the


threose enantiomers. Unlike enantiomers whose physical and chemical properties
are identical except under very special circumstances, diastereomers have
different chemical and physical properties. One can take advantage of this fact
to easily separate diastereomers from each other, a process that is not feasible for
enantiomers. Methods that are available to separate the threose diastereomeric
pair from the erythrose diastereomeric pair, for example, will not separate the
threose enantiomers from each other or the erythrose enantiomers from each
other.
In the examples just described, molecules with two stereogenic centers
afforded 4 stereoisomers consisting of 2 pairs of enantiomers. In general,
a structure with n stereogenic centers will yield a maximum number of
stereoisomers equal to 2 n, consisting of 2n/2 or 2 "-~ pairs of enantiomers. Three
centers thus will yield a maximum of 4 pairs of enantiomers, 4 centers yield
8 pairs etc. The maximum number will not be obtained if the molecule of interest
possesses an internal improper axis of symmetry such as S] or $2. To visualize
this, consider a known classical reaction of carbohydrates that converts both
termini of D- and L-erythrose and D- and L-threose into carboxylic acid groups
(COOH) (Fig. 9). The threose isomers yield a pair of enantiomeric molecules
called d- and 1-tartartic acid, d because one enantiomer rotates plane-polarized
light to the right and 1 because the other enantiomer rotates the light to the
left. Neither of these enantiomers possesses an improper axis of symmetry in
any of its various conformations. It is thus not possible to superpose d-tartaric
acid onto 1-tartaric acid. The situation is different for D- and L-erythrose. When
the ends of both molecules are converted into COOH, a single molecule called
meso-tartaric acid is produced, i.e., both enantiomers yield the same product. It
is easy to see from the two Fischer projections of meso-tartaric acid in Fig. 9
that each possesses a place of symmetry (S1), which renders the two structures
superposable and thus equivalent. Tartaric acid with two stereogenic centers
230 R.N. Compton and R.M. Pagni [III

represents then a set of three stereoisomers consisting of a pair of enantiomers,


d,l-tartaric acid, and meso-tartaric acid which is diastereomeric with the first two.
meso-Tartaric acid has the unusual property of having two stereogenic centers
and yet being achiral.
As with the amino acids (except glycine which lacks a stereogenic center),
carbohydrates such as glucose, fructose, ribose and many others, regardless of
how many stereogenic centers each possesses, occur as pairs of enantiomers with
D and L absolute configurations. Emil Fischer, who elucidated the structure of
glucose and many other sugars, never knew whether D-glucose or L-glucose
was naturally occurring. Determination of absolute configurations came later
with the development of X-ray diffraction methods and, more recently, circular
dichroism (CD). Circular dichroism is the differential absorption of LCPL and
RCPL of an enantiomer undergoing an electronic or vibrational transition. Today
it is known that all naturally occurring carbohydrates have the D absolute
configuration. What mechanism did nature use to achieve homochirality in the
carbohydrates?
Before moving on to a discussion of true and false chirality, there is one
additional issue of nomenclature that must be discussed because it will appear
on occasion in the remainder of the chapter. D and L, as already seen, refer
to the absolute configuration of an entire molecule, be it with one stereogenic
center such as in alanine or with many centers such as in glucose which
has 4. It is actually preferable to assign the absolute configuration of every
stereogenic center. The unambiguous methodology for doing this, which will
not be presented here, was developed by Cahn, Ingold and Prelog (1966). Each
stereogenic center is assigned either the (upper case) R (rectus _= right) absolute
configuration or the S (sinister--left) absolute configuration.

III. True and False Chirality

Thus far we have defined a static three-dimensional object to be chiral if its


coordinates are non-superposable on that of its mirror image. In the language
of group theory a stationary object is chiral if it has no mirror plane, center
of inversion, or improper axis of rotation. However, as we discussed earlier
for the case of an electron, an object in motion requires further consideration
since the chirality is now defined by both space- and time-dependent quantities.
Barron has provided a broader definition of molecular chirality through the
concept of "true and false" chirality (Barron, 1981, 1986a,b). "True chirality
is shown by systems existing in two distinct enantiomeric states that are
interconverted by space inversion, but not by time reversal combined with any
proper spatial rotation." This can be illustrated through considerations of the
operations involved. The parity operator P represents the spatial inversion, SI,
of the positions of all particles in an object through a fixed origin (mirror
III] THE CHIRALITY OF BIOMOLECULES 231

reflection involves spatial inversion) plus a rotation through Jr, Rn, about an
axis perpendicular to the mirror plane. This is illustrated by the S-(-) and R-(+)
enantiomers of bromochlorofluromethane"

F Mirror Reflection F
.,,,~ Br Brm,.... /
C "'~H SI + R~ . . . . C
H"" \
Cl

Classical time reversal, T, represents the reversal of the motions of all the
particles in the system. Barron presented a vivid picture of the distinction
between "true" and "false" chirality by considering two rotating cones, one in a
state of linear translation and the other rotating but translationally at rest (Barron,
1986a,b). A stationary rotating cone exists as two mirror image states which are
interconverted by SI. Since P plus R:r also interconverts these two cones, the
object exhibits "false chirality." However, a spinning cone translating along the
axis of the cone, although interconverted by SI, is not interconverted by T and R
and thus is said to possess "true chirality." This is further illustrated by replacing
the spinning cones with rotating ammonia molecules:

FALSE CHIRALITY

,,, H H

TRUE CHIRALITY

,...... "N" S..4-...[.-~I H~. H ~ ,......"I~ T


HH~'(~ H .N. H~'~~ H ~ H

Thus the instantaneous snapshot (i.e., time short compared to the tunneling time)
of a translating and rotating ammonia molecule represents a truly chiral system.
However, the fact that ammonia is rapidly converting (tunneling) from one form
to the other, requires that the two states would have definite chirality but not
definite parity. The two true eigenstates of definite parity are described by the
symmetric or antisymmetric combination of the two mirror-image forms of the
ammonia molecule. If the molecule is in one state, after a time r it will exist in
the other state. Thus, if observed over a time long compared to r, the probability
of being in either state is equal (racemization). This also illustrates the so-called
Hund's paradox (Hund, 1927), where a quantum-mechanical chiral molecule is
232 R.N. Compton and R.M. Pagni [III

therefore thermodynamically unstable (i.e., able to exist in either energetically


degenerate form). Hund evoked the long tunneling time between enantiomers to
obtain stability.
Right- and left-circularly polarized light (RCPL and LCPL) consists of
massless particles of unit spin (I = 1) angular momentum along (+h, RCPL)
or against (-h, LCPL) the direction of propagation. Circulary polarized photons
exhibit true chirality since P interconverts RCPL and LCPL, but T does not.
Likewise, translating particles with half-integer spin such as electrons and
neutrons are also fundamentally chiral. The possible effects of spin-polarized
electrons originating from beta decay will be discussed at length later.
Barron's criteria can be applied to other combined sources. As a final example
consider the combination of a static magnetic field, B, and static electric field, E:

E SI E ~ E T E
.,,q_..............~
~ D , ~ ~ D . - -" Y

B B B

where E represents a time-even polar vector and B a time-odd axial vector.


Therefore, E changes sign under spatial inversion, SI, and B does not, whereas
the reverse is true for time reversal operation, T. As illustrated above, false
chirality is seen for this system because R~ plus T has the same effect as does SI.
Avertisov, Goldanskii and Kuz'min (1991) have used the prescriptions of
Barron for true and false chirality to provide a convenient summary of most of
the physical situations in nature which are capable of producing an enantiomeric
excess in chemical reactions. They divide the "advantage factor" for mirror
symmetry breaking into a "local advantage" for those which might have existed
in a particular region on the Earth and "global advantage" as those due directly
or indirectly to the electroweak interaction. They define an advantage factor, g, as
the relative difference in the rate constants, k L and k o, for the mirror-conjugated
reactions, i.e. k L _ kD

g = kL + k D 9

Neither a static magnetic field (SMF), static electric field (SEF) or their
combinations are chiral. Gravitational fields (GF) are also achiral. However,
certain combinations of these fields together with rotations are chiral, as seen
in Table I. All of the advantage factors shown are dependent upon a molecular
factor, Z, determined by the structure of the molecule under consideration. The
chiral influence of longitudinally polarized [3 particles also depends upon the
electron helicity h, and the relative difference in the spin-polarized electron
scattering cross section ((yL _ ( j D ) / / ( o L _+_ (yD).
A report of asymmetric induction/resolution occurring in static magnetic
fields (Zadel et al., 1994) has been retracted (Breitmaier, 1994) and was shown
IV] THE CHIRALITY OF BIOMOLECULES 233

Table I
Chirality of field combination with rotation

Advantage factors

Local
Circularly polarized light 10-4 to I 0 -1

Rotation + SMF + SEF <10 -4

Rotation + SMF + GF <10 -4

Linearly polarized light + SMF <10 -4

Global
Weak neutral currents IO-20ZrZ5kBT~ 10-13 to 10-17
Beta particles Xqh3A(oL-oD)/(OL +oD)~ 10-9 to 10-11

to be theoretically impossible (Barron, 1994), in agreement with experiment.


Recently, Rikken and coworkers have shown, however, that enantioselective
photochemistry can occur using plane-polarized light in a parallel magnetic
field which exhibits true chirality (Rikken and Raupach, 2000; Raupach et al.,
2000). This research was recently "highlighted" by van Wfillen (2001). The
enantiomeric excesses were quite small even in the presence of large magnetic
fields. It is generally believed that the Earth's magnetic field is too small to have
produced an enantiomeric advantage factor large enough to have influenced the
development of homochirality of biological molecules.

IV. Galaxies, Plants, and Pharmaceuticals


Chirality exists in our universe at many levels of size. On the grand scale, spiral
galaxies are chiral. When observed from the earth onto the galactic plane, they
have an "S" or "Z" shape. Of course, when observed side-on, these shapes are not
apparent. A non-receding, stationary, S-shaped galaxy would be achiral because
its mirror image, which is Z-shaped, is superposable on the original, even if the
galaxy were rotating about the galactic center. Likewise, a stationary Z-shaped
galaxy would be achiral. Because of the Big Bang origin of the universe, all
galaxies are receding from the earth. Thus, the receding and rotating S- and
Z-shaped galaxies are chiral due to the combined linear and rotational motion.
Interestingly, a study of more than 500 spiral galaxies reveals an equal number
of S- and Z-shaped galaxies (Kondepudi and Durand, 2001; Iye and Sugai,
1991; Sugai and Iye, 1995), although there is a significant asymmetry in certain
subclasses of the galaxies (Sugai and Iye, 1995). There is also no apparent
segregation of S- and Z-shaped galaxies based on their recessional velocities.
234 R.N. Compton and R.M. Pagni [IV

FIG. 10. Twisted eucalyptus tree.

Many plants also exhibit handedness. Charles Darwin, the father of the theory
of evolution, made detailed observations of the "twining" of certain climbing
plants, that is, their propensity to grow in a spiral or corkscrew manner (Darwin,
1906). In nature, twining plants are homochiral; honeysuckle twines to the left
and morning glory twines exclusively to the right. Most other plants grow in a
right-handed spiral and most seashells also exhibit a right-handed helicity. The
human umbilical cord, on the other hand, twists in either direction. The left tusks
of the Narwhal, a species of toothed whale, twist in a counterclockwise direction.
Figure 10 shows the right-hand twist exhibited by a large Blue Gum tree (genus
Eucalyptus) located in the Botanical Gardens of Christchurch, New Zealand.
IV] THE CHIRALITY OF BIOMOLECULES 235

CH 3 CH3 CH3 CH3


o

i
_
i
~cg3 3 ~"CH3 H3

(R)-carvone (S)-carvone (R)-limonene (S)-limonene

OH
,I
O H OCH2CHCH2NHCH(CH3)2
II ,I
H2NCCH2m C mCOOH
I
NH2

asparagine propranolol

O O

OH
I,
HOCH2CCH2CI
I
H

1-chloro-2,3-propanediol (R)-thalidomide (S)-thalidomide

FIG. 11.

Many other examples of homochirality may be found in the book by Gardner


(1990).
In addition to amino acids and sugars, many other life-supporting molecules
are homochiral as are many pharmaceuticals. The chemistry of the human body
is vitally dependent upon the subtle stereochemical differences of enantiomers.
Different sensations, for example, are often exhibited by naturally occurring
enantiomers. To about 90% of people (R)-carvone (see Fig. 11; the asterisk
locates the stereogenic center) tastes and smells like spearmint (from spearmint
oil) while the S enantiomer smells and tastes like caraway (from caraway
seeds). The R enantiomer of limonene, a terpene, tastes like orange while
the other enantiomer tastes like lemon. The S enantiomer of asparagine, a
chemical (alkaloid) found in asparagus, tastes bitter while its R mirror twin
is sweet to the taste. These differences are not always observed, however. The
236 R.N. Compton and R.M. Pagni [V

naturally occurring D-glucose and the unnatural L-glucose are equally sweet. If
an inexpensive synthesis of it were available, L-glucose would be an ideal non-
nutritive sweetener.
Enantiomerically pure drugs are becoming increasingly important even though
they are generally more expensive to prepare than their racemic counterparts
(Deutsch, 1991; Stinson, 1995, 1998, 2000, 2001). The world-wide market
for single-enantiomers drugs already exceeds $120 billion per year. Often
one enantiomer is therapeutic while the other is benign or even harmful.
Separate enantiomers of barbituric acid derivatives can function as a narcotic
or anticonvulsant. (-)-Propranolol acts as a beta blocker for heart disease while
its enantiomer acts as a contraceptive. The S form of 1-chloro-2,3-propanediol is
a pharmaceutical while the R form is poisonous. One enantiomer of the steroid
estrone is a sex hormone while the other is inactive.
Thalidomide, which possesses one stereogenic center (at the starred position),
is undoubtedly the most notorious compound where one enantiomer is reported
to be therapeutic while the other is harmful. The R enantiomer of this
compound has been used to alleviate the effects of morning sickness of pregnant
women while the S enantiomer gives rise to teratoidism, i.e., birth defects, in
developing fetuses. It has been estimated that 12 000 children have been born
with thalidomide-induced birth defects in the 1950s, primarily in Europe and
South America. This molecule was not approved for use in the United States
except for research purposes. It had been presumed that if enantiomerically pure
(R)-thalidomide were given to pregnant women the problem of birth defects
would be avoided. However, it is now known that thalidomide racemizes rapidly
in blood plasma so that providing an enantiomerically pure form of thalidomide
would not alleviate the problem of teratoidism. Thalidomide may still prove
useful in certain circumstances. The compound suppresses the growth of blood
vessels (angiogenesis) which may prove useful in the treatment of certain cancers
and AIDS-related conditions. Research is now underway on the development of
resolved derivatives of thalidomide which are therapeutic but do not racemize
easily.
Approximately 530 synthetic chiral drugs are now marketed worldwide.
Because of the technical difficulties and corresponding expense of making
enantiomerically pure drugs, only about 60 of these compounds are marketed
in their enantiomerically pure form. Increasing pressure from the scientific
community along with tighter restrictions imposed by the U.S. Food and Drug
Administration means that more enantiomerically pure pharmaceuticals will
become available in the near future.

V. Plausible Origins of Homochirality


Louis Pasteur, the 19th-century chemist and microbiologist, was the first person
to realize that many of the molecules of life are homochiral. We know today
V] THE CHIRALITY OF BIOMOLECULES 237

that carbohydrates, which are found in starch, cellulose, and the nucleic acids,
exist in nature exclusively with the D absolute configuration. Most amino acids,
which are the building blocks of proteins and enzymes, have the L absolute
configuration. This is true regardless of the type or complexity of the organism,
be they found in a bacterium, betel (palm or pepper), beetle, beagle, or
the Beatles. Thus specific homochirality of biomolecules is a fundamental
characteristic of life. Any plausible theory of the origins of life on earth must take
this fact into account. Two general types of theories have been invoked to explain
the origin of homochirality on the earth: (1) those involving a chance or random
selection of the homochiral molecules that were incorporated into the first living
organisms; and (2) those that require an internal or external chiral force to bring
about the asymmetric synthesis of the homochiral molecules. As will be seen in
due course, many theories in the second grouping also have a random quality
about them because the chiral influence itself may appear randomly. Leaving
aside the accidental "chance scenario," theories have concentrated on how chiral
forces may influence the prebiotic asymmetric synthesis of biological molecules
in nature.
Ideas that specific homochirality has resulted from some prebiotic asymmetric
synthesis have become more credible in recent years following the analysis
of organic compounds brought to the earth in meteors. Cronin and Pizzarello
(1997) for example, found a 7 to 9% excess of L amino acids in the Murchison
meteorite found in Australia. It is believed that this enantiomeric excess did not
arise by terrestrial contamination because some of the amino acids found in the
meteorite do not exist in the biosphere. The Murchison meteorite, a carbonaceous
chondrite, was created about 4.5 billion years ago in the asteroid belt, strongly
suggesting an asymmetric prebiotic chemical synthesis of the carbon-containing
molecules in the universe. The "normal" laboratory synthesis of amino acids, of
course, yields a racemic mixture. For example, an electric discharge (a mimic
of lightning) in a mixture of methane, nitrogen, ammonia, and water produced
racemic mixtures of the amino acids, alanine (1.7% yield), asparagine (0.024%),
and glutamine (0.051%); achiral glycine was found in 2.1% yield (Miller, 1959).
Assuming an extraterrestrial asymmetric synthesis of the amino acids found in
the Murchison meteorite, some chiral influence has been at work on the meteorite
when it was first synthesized or during the intervening 4.5 billion years. It is
plausible that the source of homochiral molecules necessary for the formation
of life on the earth has an extraterrestrial origin. Even if this be true, the
question remains as to how the asymmetric synthesis occurred. Several possible
explanations follow.
Circularly polarized light is a plausible chiral influence. It is well known that
the photochemistry of racemic mixtures of molecules initiated with circularly
polarized light can lead to an excess of one enantiomer over the other, either
in product or recovered reactant (Feringa and van Delden, 1999; Bonner,
238 R.N. Compton and R.M. Pagni [V

1996; Rau, 1983; Balavoine et al., 1974). This photochemistry can occur
by photoequilibration of a racemic mixture of molecules without any loss of
reactant or by the selective destruction of one enantiomer over the other. Both
processes take advantage of the fact that enantiomers have different absorptivities
(cross sections) when exposed to LCPL and RCPL. Both methods, however,
have serious drawbacks. In the first instance enantiomeric excesses tend to
be low because differences in absorptivities usually are very small while the
absorptivities are generally large. In the second case, very high enantiomeric
excesses can be obtained but only at the expense of destroying most of the
reactant. There are several notable photoreactions in the literature initiated by
CPL. Shimizu and Kawanishi have studied the two-photon photochemistry of
D,L, i.e., racemic tartaric acid with 351-nm CPL from a XeF excimer laser
(Shimizu and Kawanishi, 1996a,b; Shimizu, 1997). Irradiation with RCPL led
to a selective destruction of the L enantiomer while the concentration of the
D enantiomers changed little; this led to a maximum enantiomeric excess of
recovered tartaric acid of 7.5%. Irradiation with LCPL was unfortunately not
carried out.

COOH
I COOH
~HOH 351-nm RCPL from
XeF excimer laser
r
HO H
+ CO2, CO, H 2
CHOH H OH
I
COOH COOH

racemic D,L- D-tartaric acid


tartaric acid

Inoue and coworkers have also investigated a photoreaction using a different


intense light source, CPL with 190 nm synchrotron radiation from a storage ring
(Inoue et al., 1996). They studied the photointerconversion of racemic trans-
cyclooctene, a chiral molecule with no stereogenic centers, and the achiral cis-
cyclooctene:

CPL

R enantiomer S enantiomer

trans-cyclooctene cis-cyclooctene

Although enantiomeric enrichment was found, the effect was small. Photolysis of
the racemic amino acid, D,L-leucine, with RCPL and LCPL in aqueous solution
(in water) at 212.8 nm also afforded small, but meaningful enantiomeric excesses
V] THE CHIRALITY OF BIOMOLECULES 239

in recovered reactant, a result with implications for the origins of homochirality


on the earth (Flores et al., 1977).
Thus circularly polarized light in nature could provide a mechanism for the
prebiotic synthesis of optically active amino acids, carbohydrates, and other
relevant molecules. A number of sources of circularly polarized light on earth
and in the universe have been identified. Sunlight passing through the atmosphere
generates slightly elliptically polarized light (not quite circular), LCPL before
sunrise and RCPL after sunset (Angel et al., 1992). The polarization can reach
- and +0.2%, respectively. At sunrise and sunset the polarization is zero. This
is clearly a small effect that averages to zero over the course of a day. However,
the opposite sides of a mountain ranging from north to south would experience
more or less of a given elliptical polarization. Supernovae are also known
to emit CPL at right angles to the normal linear Bremsstrahlung generated
from electrons rotating in a circle. CPL is emitted perpendicular to the plane
of Bremsstrahlung: RCPL in one direction and LCPL in the other direction.
Thus, in the sunrise/sunset effect of our sun and in supernovae explosions, both
right- and left-CPL are produced so that any advantage in the photosynthesis of
biomolecules would have to occur through a chance encounter with one form of
the light (Bonner, 1997).
There has also been observation of strong circular polarization of infrared
radiation in regions such as Orion OMC-1, where star formation is occurring
(Bailey et al., 1998). The circular polarization was attributed to Mie scattering of
linearly polarized light from either spherical grains of dust or nonspherical grains
that had been aligned by a magnetic field. The percentage of circular polarization
varied from -5% to +17% depending upon the region observed. Although only
polarized radiation at 2.2 ~m was detected, calculations show that the circular
polarization would persist at shorter wavelengths. Bailey and coworkers noted
that this type of light may be responsible for the synthesis of homochiral
biomolecules because enantioselective synthesis of amino acids using CPL of
much shorter wavelength (<200 nm) has been reported (Norden, 1997). Recent
calculations also suggest that photochemistry initiated with elliptically polarized
infrared radiation may also be possible (Fujima et al., 1999). Unfortunately,
infrared photochemistry is usually a multi-photon process because a single
infrared photon does not possess enough energy to break chemical bonds.
We are investigating the use of intense, short-pulsed, high-power circularly
polarized infrared light from a free-electron laser to initiate enantioselective
photochemistry.
All of the chiral influences discussed thus far are local. When averaged over
space and time, these asymmetries average to approximately zero. One source of
CPL that is chiral over space and time comes from the beta decay of radioactive
atoms. This effect will be described shortly. Many review articles describe other
physical processes which could have been responsible for the mirror symmetry-
240 R.N. Compton and R.M. Pagni [V

breaking in the biological world (Avertisov et al., 1991; Ponnamperuma and


MacDermott, 1994; Mason, 1988, 2000; Quack, 1989; Frank et al., 2000; Cline,
1996; Bonner, 1988; Avalos et al., 1998, 2000a; Buschmann et al., 2000).
Of particular note are the possible effects of parity-nonconserving electroweak
interaction because the phenomenon is innately homochiral.
All matter interacts through four known forces (in order of strength):
nuclear, electromagnetic, weak, and gravitational. Of these forces only the weak
interaction is fundamentally chiral because it can distinguish between left and
right. The weak interaction is responsible for radioactive beta decay, the slow but
spontaneous emission of electrons from metastable nuclei. Electrons possess an
intrinsic spin of l h (h = h/27c) and spin-angular momentum (z component) of
1v~h . An electron in motion can have its spin-angular momentum vector aligned
along or against its direction of propagation, thus yielding a circularly polarized
particle. Thus an electron in motion is chiral and its spin describes a right- or
left-handed helix as it moves through space. In the 1950s it was predicted (Lee
and Yang, 1956) and observed (Wu et al., 1957) that the weak nuclear force does
not conserve parity. The consequence of parity nonconservation in beta decay is
that leptons emitted from the nucleus are primarily spin polarized in one helicity.
Beta rays are primarily left-handed. Positron decay, on the other hand, gives rise
to right-handed anti-electrons. Thus, it is often said that we live in a left-handed
universe since it is made of matter rather than antimatter (Barrow and Silk, 1983).
Shortly after the discovery of parity violation, the predicted effects of the chiral
electroweak interaction in atomic physics were tested. Since the nucleus of an
atom contains a chiral force, all atoms are thereby chiral and should exhibit
optical activity and circular dichroism (Khirplovich, 1991). This force will also
mix states of opposite parity, giving rise to the observation of "parity-forbidden
transitions" in atoms (Bouchait and Bouchait, 1997). In combination with theory,
these experimental measurements have provided a sensitive test of the standard
model of nuclear physics at low energy (Masterson and Wieman, 1995). Similar
predictions can be made concerning the effects of the electroweak interaction
in molecular physics, i.e., optically active diatomic molecules and the mixing
of parity-forbidden states; however, such effects have yet to be reported. The
observation of the parity-violating effects in atoms poses a broad question in the
context of the subject of this article: can the weak interaction play a role in chem-
istry in general and the origins of homochirality in biomolecules in particular?
There are at least two ways that the weak interaction could influence molecules
and their chemical reactivity: (1) by altering the energies and energy levels of
molecules (parity-violating energy difference, PVED); and (2) by the interaction
of left-handed electrons with matter following beta decay. We examine these
effects separately.
The parity-violating (P-odd) weak interaction is the only known chiral force in
non-decaying atoms and molecules. Weak interaction effects in molecules of the
V] THE CHIRALITY OF BIOMOLECULES 241

type observed for atoms have not been reported. Instead, research in P-odd effect
in molecules has concentrated on the prediction and detection of PVED between
R and S enantiomers. Rein (1974) and Letokhov (1975) first suggested that the
electroweak interaction would lift the degeneracy between enantiomers predicted
by normal quantum electrodynamics. A number of calculations of the PVED for
small molecules have been published (Hegstrom et al., 1980; Zel'dovich et al.,
1977; Mason, 1984; Mason and Tranter, 1984; MacDermott et al., 1992; Bakusov
et al., 1998; Lazzeretti and Zanasi, 1997; Zanasi et al., 1999; Laerdahl and
Schwerdtfeger, 1999; Laerdahl et al., 2002; Berger et al., 2001). The Hamiltonian
operator describing the PVED for a molecule containing i electrons and n nuclei
is given by

Hpv= H~v+ H2v


=(GF
-~) ZOw'5pn(ri)+-~) Z di " [nPn(ri).
I11(I~ + 1)

The Fermi electroweak coupling constant, GF (2.22255• 10-14 a.u.-1.16637•


10-l~ eV), sets the scale for the small magnitude of this effect. 111 is the
nuclear spin vector operator, K11 and tr are nuclear parameters for nucleus n,
y5 _- i y I y Z y 3 y 4 represents a pseudoscalar chirality operator formed from a
product of Dirac g-matrices, p11(ri) is the normalized nucleon density, and Qw
is the weak charge given by

Qw = -N11 + Z11(1 - sin20w),

where N11 and Z11 are the numbers of neutrons and protons, respectively, and
0w is the Weinberg mixing angle. An excellent review of the state of affairs
of the theoretical calculations of the parity-violating energy difference for small
molecules has been published by Buschmann, Thede and Heller (2000). In the
1980s, ab initio calculations gave PVED values on the order of 10-14 J/mol
and, interestingly, concluded that the L-amino acids and D-sugars were lower
in energy. Until recently, this stabilization energy was believed to be too small
to be of significance to prebiotic asymmetric synthesis. Although autocatalytic
processes such as those evoked by Kondepudi and Nelson (1983, 1984a,b, 1985)
could play a role in providing enantiomeric excesses over long times, the small
energy differences expected did not encourage further speculation. However,
more recent calculations, particularly those by Quack and co-workers (see Berger
and Quack, 2000 and others contained therein) and Schwerdtfeger et al. (see
Laerdahl, Wesendrup and Schwerdtfeger, 2000) are about an order of magnitude
larger and give new interest to possible PVED effects in prebiotic asymmetric
chemistry. However, both of these groups have shown that the magnitude and
indeed the sign of the PVED for alanine depend upon its exact molecular
conformation (shape that may change by rotating atoms or group of atoms
242 R.N. Compton and R.M. Pagni [V

around single bonds) and therefore may not lead to a preferential stabilization
of one enantiomer over the other. There is an obvious critical need for an
experimental measurement of the PVED. Unfortunately, a measurement on the
order of 10- 1 4 J/mol is beyond any present experimental technique. Fortunately,
the PVED increases with the magnitude of the atomic number. Specifically, the
PVED scales with Z 6 (Laerdahl and Schwerdtfeger, 1999). Thus, molecules with
atoms having large atomic numbers will show the largest PVED and may provide
an experimental test for PVED.
There have been a number of experimental attempts to measure PVED. Most
notably, high-resolution infrared absorption spectroscopy has been used on two
occasions. Arimondo, Glorieux and Oka (1977) observed inverted Lamb-dip
ro-vibrational transitions for the two enantiomers of camphor and found them to
agree to within 10-8. Recently, Daussy et al. (1999) demonstrated that the C - F
stretching mode of the two CHFBrC1 enantiomers agreed to within 13 Hz. This
places an upper limit on AEpv/E of <4x 10 -13. Although these measurements
are five orders of magnitude more sensitive than the results for camphor, they
are still not sensitive enough. Theoretical studies on CHFBrC1 give an energy
difference of only 2 mHz, (-8 x 10-18 eV) for the C - F frequency (Laerdahl et al.,
2002).
Keszthelyi (1994) used M6ssbauer spectroscopy to establish an upper limit
of 4x 10 -9 for the energy difference of 1- and d-tris(1,2-ethanediamine)iridium
complexes. A Z 6 scaling was used to estimate the energy difference of
1.1 x 10-12 eV between the two enantiomers. Thus the measured upper limit is
three orders of magnitude higher than the expected Z-scaling result. This report
contains very few details on the experimental procedures used, and no M6ssbauer
spectra were shown.
Although all of the L amino acids are predicted to lie lower in energy than
the corresponding D amino acids, the differences are so small that it is hard to
envision how this could lead to a dominance of the L amino acids in biology. A
large amplification mechanism would be required. However, the PVED scales
with Z 6, as noted earlier. Thus, for larger molecules containing atoms with
larger atomic numbers, the energy differences between enantiomers should
be sufficiently large to be observable with currently available spectroscopic
techniques such as nuclear magnetic resonance, M6ssbauer and perhaps infrared
spectroscopies. Lahamer et al. (2000), for instance, reported an energy difference
between enantiomers of an iron complex using M6ssbauer spectroscopy. Due to
the fact that these experiments were performed on crystalline material, the energy
difference may be due to crystalline effects, although the difference in energy
(-10 -l~ eV) is close to that predicted by scaling the PVED with atomic number.
In 1991 Abdus Salam (1991), who along with Glashow and Weinberg
united the electromagnetic and weak force theory into one electroweak theory,
speculated that the Z ~ interactions might provide an explanation of the
VI] THE CHIRALITY OF BIOMOLECULES 243

domination of L amino acids in living systems in a manner which has


not previously been considered. His explanation was mainly heuristic and
involved quantum-mechanical cooperative and condensation effects similar to
those in Cooper pairing and Bose-Einstein condensation. He postulated that
these effects could give rise to second-order phase transitions below a critical
temperature Tc. Realizing that phase transitions as a general rule occur at
a low temperature he speculated as to the location for the production of
L amino acids and suggested possible laboratory tests of these ideas. In
a recent paper, Wang et al. (2000) presented three experimental tests of
Salam's ideas using single crystals of L- and D-alanine (and valine). In one
measurement they report a X phase transition at 270+ 1 K for both sets of crystals
using differential scanning calorimetry. The L-enantiomer was reported to be
lower in energy. They also reported temperature-dependent mass susceptibilities
which are different for L- and D-alanine. These differences are attributed
to a PVED-dependent difference in the intramolecular geometry of chiral
density. Finally, Raman spectroscopy measurements show that the second-order
Ca-H deformation modes at 2606 cm -l and 2724 cm -I for D-alanine vanish at
270 K but reappear at 100 K whereas L-alanine shows no such effect. All of these
experiments are presented as experimental evidence of the possible relevance of
the Z ~ interaction as presented by Salam at a phase transitions of Tc ~ 270 K.
One notes the proximity of this 270 K temperature to that of the freezing point of
water, the solvent used to recrystallize the samples. Verifications of these results
are crucial and are ongoing in our laboratories.

VI. Asymmetry in Beta Radiolysis

Because the energy difference between a pair of enantiomers is extremely small


regardless of the conformations of the molecules, a significant amplification or
deracemization mechanism would have had to operate in the prebiotic world for
an extended period of time for the homochirality of biological molecules to have
evolved. This is indeed a difficult task. A perhaps more viable way in which
PVED could have brought about homochirality is through beta decay. Many
experiments have been performed to test this hypothesis, often with ambiguous
or contradictory results.
Vester et al. (1959) were the first to suggest a connection between the
asymmetry of beta ([3) decay and the evolution of homochirality in the biological
world. Their mechanism involves the energetic spin-polarized [3-rays and spin-
polarized secondary electrons producing a net left-circularly polarized )'-rays
(Bremmsstrahlung) which could be a source of asymmetric photochemical
reactions. In an extensive set of experiments using radioactive phosphorus (32p),
strontium (9~ and europium (152Eu), all beta-particle emitters, to initiate
asymmetric reactions of organic compounds, the same researchers found no
244 R.N. Compton and R.M. Pagni [VI

H
,I
HzN~C~COOH
I
R

R
O
(+)-camphor tyrosine H O ~ H 2 ~

J
CH 2

] tryptophan
CH3CH2~COOH
H

]~~/~ leucine (CH3)2CHCH


2
2-phenylbutyric acid
aspartic a c i d HOOCCH2

FIG. 12.

positive results (Ulbricht and Vester, 1962). Garay (1968), on the other
hand, found D-tyrosine, an amino acid (Fig. 12), to degrade more rapidly
in water containing radioactive strontium chloride (9~ than the naturally
occurring L-tyrosine enantiomer. Unfortunately, the method of analysis that
Garay u s e d - UV spectroscopy- is not definitive. Darge et al. (1976), likewise,
found positive results. In these experiments a frozen aqueous solution of
racemic D,L-tryptophan, another amino acid, and radioactive phosphate (32po43)
underwent self-radiolysis for 12 months; an astounding 19% enantiomeric excess
of the D enantiomer was reported. As with the Garay experiments, the methods
of analysis- UV spectroscopy and polarimetry- are not definitive. When the
Garay and Darge experiments were repeated (with considerable modification)
on D,L-tyrosine, D,L-tryptophan, and other amino acids by Bonner (Bonner,
1974; Bonner and Flores, 1975, Bonner et al., 1979) using gas chromatography,
where the enantiomers are separated and quantitated, as the analytical tool, no
asymmetric degradation of any amino acid was found even though extensive
overall degradation of the amino acids had occurred. Calvin and colleagues
(Bernstein et al., 1972) and then Bonner (Bonner et al., 1978) also examined
the internal beta radiolysis of several 14C-labeled amino acids (14C is a beta-
particle emitter), Calvin by polarimetry and Bonner by gas chromatography;
no positive results were obtained. Bonner did have success, however, with
polarized electrons not from beta decay but using longitudinally polarized
electrons (120 KeV) from a linear accelerator (Bonner et al., 1975, 1976/77).
VI] THE CHIRALITY OF BIOMOLECULES 245

The target was a solid sample of racemic leucine, a still different amino acid.
Antiparallel electrons afforded a slight excess of L-leucine (~1% enantiomeric
excess) after about 50% degradation of the sample, while parallel electrons
afforded a slight excess of the D enantiomer (~0.9% enantiomeric excess) after
about 75% degradation. More recently, Akaboshi et al. (1999) looked for the
asymmetric decomposition of the racemic amino acid, D,L-aspartic acid, in
tritiated water (3H20 , 3H--tritium, a beta emitter). When the recovered amino
acid was examined by HPLC, another technique that separates and quantitates
the enantiomers, no enantiomeric excess was detected.
Except for Bonner's results using polarized electrons generated in a linear
accelerator, none of the other experiments involving amino acids, all of
which involved beta radiolysis, led to asymmetric degradation. Several related
experiments did give positive results. Beta radiolysis of polycrystalline amino
acids yields free radicals (species with one unpaired electron) which are detected
by electron-spin resonance spectroscopy, a very sensitive technique. Irradiation
of D- and L-alanine with internal or external beta radiation affords an excess
of radicals from the D enantiomer (Akaboshi et al., 1978, 1981, 1982; Conte
et al., 1986). Beta particles passing through the liquid enantiomers, R- and
S-2-phenylbutyric acid scatter differently as evidenced by emitted Cerenkov
radiation (Garay and Ahlgren-Beckendorf, 1990). The scattering of polarized,
low-energy electrons from enantiomers such as D- and L-camphor also occurs
asymmetrically (Campbell and Farago, 1985; Blum and Thompson, 1998).
To date there is no unambiguous evidence that the beta radiolysis of amino
acids leads to the faster degradation of one enantiomer as compared to the other,
i.e. deracemization. There is evidence, on the other hand, for the racemization of
enantiomers of amino acids by beta radiation. The self beta radiolysis of optically
pure 14C-labeled amino acids over the course of decades led to the compounds'
partial racemization. (Bonner, 1996). Radiolysis of L-alanine, D-2-aminobutyric
acid, L-norvaline, L-norleucine, and D- and L-leucine, all amino acids, with
y-rays from a cobalt source (6~ results in partial racemization of the amino
acids in all cases (Bonner and Lemmon, 1978a,b). Thus any mechanism
involving radiolysis that builds up enantiomeric excess in recovered amino acid
would also provide a mechanism for its loss.
There is a way in which beta radiolysis has led to significant amplification of
enantiomeric excess. This involves the behavior of sodium chlorate (NaC103) in
water, for which relevant background material will now be presented.
NaC103 is an achiral molecule consisting of a spherical sodium cation (Na +)
and a chlorate anion (C103) of C3v symmetry. Solutions of this compound in
water are likewise achiral and thus optically inactive. The colorless, transparent
crystals of this molecule, on the other hand, are chiral because they belong to the
chiral P213 space group. Crystals of NaC103 thus exist in enantiomeric forms
which are easily distinguished by polarimetry (Pagni and Compton, 2002).
246 R.N. Compton and R.M. Pagni [VI

When an aqueous solution of NaC103 is evaporated in a Petri dish, well-


formed crystals form; a distribution of (+) and (-) crystals, and thus an
enantiomeric excess, can be determined by counting crystals. Kondepudi has
shown that in the absence of a perturbation, a random distribution of (+) and (-)
crystals is produced, as expected (Kondepudi et al., 1993). This is a consequence
of the fact that the seed crystals are generated in one enantiomeric form or the
other randomly. Stirring, surprisingly, has a dramatic effect on the distribution
of crystals (Kondepudi et al., 1990). Under these circumstances a Petri dish
will generally contain exclusively all (+) or all (-) crystals. The direction of
stirring has no bearing on which enantiomers are produced in a dish. Although
the mechanism by which this bifurcation occurs has been debated (Kondepudi
and Sabanayagam, 1994; Martin et al., 1996), it is clear that, when stirring is
involved, the initially formed crystals provide the seed(s) for the subsequent
production of all crystals that arise in the dish. Even though bifurcation has
occurred with stirring, a 50:50 distribution of (+) and (-) crystals is still obtained
when one counts all the crystals from a large number of dishes.
[3 radiolysis of an evaporating aqueous solution of NaCIO3 has a profound
influence on the distribution of the resulting (+) and (-) crystals (Mahurin et al.,
2001). In two sets of experiments (54 Petri dishes) carried out in front of an
intense 9~176 source, enantiomeric excesses (based on numbers of crystals)
of +32.1% and +42.6% favoring the (+) crystals were obtained. The enantiomeric
excess distribution per dish versus frequency was now bifurcated asymmetrically,
with dishes having enantiomeric excesses of +100% being four times more
prevalent than those with enantiomeric excesses o f - 1 0 0 % . There was still a
smattering of enantiomeric excesses between these two extreme values. Because
the [3 particles are approximately 80% left-handed, one can rationalize these
results in the following manner: The more common left-handed electrons induce
the preferential formation of (+) crystals which then seed the solution to produce
more of the same (+) crystals. By the same token, the less common right-handed
electrons initiate the formation of (-) seed crystals. The details of how this
remarkable result occurs are not currently known. Random crystallization still
competes with the particle-induced crystallization. This explains the smattering
of enantiomeric excesses occurring between +100% a n d - 1 0 0 % .
Similar, but less extensive experiments were carried out in front of the positron
source (22Na). The energetic, positively charged electrons are predominantly
right-handed. In this instance an overall enantiomeric excess o f - 5 5 . 2 % favoring
the (-) crystals was obtained. Petri dishes with 100% of the (-) crystals were
about four times more prevalent than dishes with 100% of the (+) crystals. The
results with positrons were in essence the mirror image of those obtained with
beta particles. The chirality of one lepton thus strongly influenced the chirality
of the crystals.
Before proceeding to the next section, it is worth describing a recent result
VI] THE CHIRALITY OF BIOMOLECULES 247

that presents an interesting juxtaposition to the results of Kondepudi described


above. Stirring of the aqueous NaC103 solutions had a profound effect on the
distribution of + and - crystals, but in a random fashion. Rib6 and coworkers
(2001) have now found a clear correlation. The researchers found that certain
achiral porphyrins form chiral and optically active aggregates in solution. When
the aggregations are carried out without stirring, a 50:50 distribution of + and -
aggregates are generated. When the aggregate-forming solutions are stirred in
one direction, 88% of the experiments yield + aggregates whereas, when stirred
in the other direction, 85% of the experiments y i e l d - aggregates. There is thus
a distinct correlation between the direction of stirring and the chirality of the
aggregates that result. Stirring of the solutions creates vortices in the fluids which
possess true chirality according to the definition of Barron. The mechanism by
which the chirality of a vortex is transferred to that of the aggregate is not clear.
It is also unclear how this very interesting observation can be applied to the
origin of optically active molecules in the prebiotic world because the optically
active vortices in the primitive aqueous environment would have been generated
randomly.

A. AMPLIFICATION AND DEGRADATION MECHANISMS

If the physics and chemistry described above generally yield enantiomers with
very small enantiomeric excesses, a large amplification mechanism is required to
attain the homochirality of the biological world. This is particularly true if PVED
was the initial source of asymmetry because its effect is exceedingly small.
Where crystallization and aggregation phenomena are concerned, however, the
initial enantiomeric excesses may be substantial and only a modest amplification
may be required to attain homochirality.
The crystallization effect is best exemplified with sodium chlorate. Seeding
of an evaporating aqueous sodium chlorate solution with (+) and (-) crystals
yields (+) or (-) crystals, exclusively. Stirring has the same effect but the
direction of the induced homochirality is random. Circularly polarized electrons
and positrons generated from radioactive nuclei, as noted above, have a similar
effect on the distribution of (+) and (-) sodium chlorate crystals.
Although it is difficult to see how sodium chlorate crystals could have played
a role in the origin of biological homochirality because they are synthetic,
naturally occurring quartz, which exists in enantiomeric forms, could have played
such a role. Unfortunately, there is essentially no excess of one enantiomeric
form of quartz over the other on the earth although local excesses may exist
(Frondel, 1978; McBride, 2001). Interestingly, asymmetric adsorption of the
alanine enantiomers onto (+) and (-) quartz has been demonstrated by Bonner
(Bonner et al., 1974). Until the recent work of Soai, which will be described
below, quartz had not served as a medium for an asymmetric chemical reaction.
248 R.N. Compton and R.M. Pagni [VI

Calcite, a form of calcium carbonate (CaCO3), also occurs naturally. Unlike


quartz, calcite is achiral. Its adjacent crystal faces turn out to be optically active
and of opposite handedness, however. Recent experiments have shown that one
face of calcite selectively adsorbs L enantiomers of a few amino acids from
aqueous solution while the other face selectively adsorbs the D enantiomers
(Hazen et al., 2001). On a related note, naturally occurring clay has been
suggested as the environment in which life evolved (Cairns-Smith, 1982). Even
if this were so, clay could not have been the progenitor of homochirality because
it is achiral.
The photochemistry of homochiral crystals, which are made up of achiral
molecules, usually affords photoproducts with significant enantiomeric excesses.
The following example illustrates this fact. The so-called dibenzobarrelene
exists in two crystalline space groups: Pbca, which is achiral, and P212121,
which is chiral. Photolysis in the achiral crystalline form generates a dibenzo-
semibullvalene containing four stereogenic centers in racemic form, identical to
what is obtained in solution (Evans et al., 1986):

CO2CH(CH3)2
(CH3)2CHO2C~ (CH3)2CHO2C CO2CH(CH3)2
hv

dibenzobarrelene dibenzosemibullvalene

Photolysis of homochiral crystals of the dibenzobarrelene in the P212121


dimorph, on the other hand, yields the photoproduct with 100% enantiomeric
excess. It is difficult, however, to see how results of this type could be responsible
for the origin of biological homochirality. As with quartz, one expects to get
a 50:50 mixture of crystals on crystallization, with photolysis of (+) crystals
yielding one enantiomer of the photoproduct and photolysis of (-) crystals
yielding the other enantiomer.
Frank (1953) was the first person to propose a kinetic model for the
spontaneous synthesis of optically active products from achiral precursors. In this
model both enantiomers of the products are produced initially. Each enantiomer
in turn acts as a catalyst for its own production and an inhibitor for the production
of the other enantiomer. Under some circumstances the system becomes unstable
and the subsequent course of the reaction is subject to small perturbations. The
net result is to generate an excess of one enantiomer, with the one in excess
being produced randomly. Kondepudi and Nelson (1983, 1984a,b, 1985) have
proposed a related model in which two reagents A and B react reversibly to
VI] THE CHIRALITY OF BIOMOLECULES 249

give enantiomeric products R and S, each of which catalyzes its own formation
reversibly and with each other irreversibly to produce a product P:

A + B ~,-~--R, A+B~----S,
A + B + R ~,-~-2R, A + B + R ~,-~--2S, R + S--,P.

This system is open because A and B are continually added while P is continually
removed. As with Frank's model, this one becomes unstable under certain
conditions, thus yielding one enantiomeric product in excess. If this system is
biased, however, with an initial excess of R or S, even ever so slightly due to
PVED, one pathway becomes dominant. If R is in excess initially, the pathway
leading to production of R will be favored, and vice versa if S is in excess
initially. Calculations show that in periods as small as tens of thousands of years,
the initial very small enantiomeric excess due to PVED can be amplified to
enantiomeric excesses close to 100% favoring the L-amino acid. It is unfortunate
that there is no possibility of testing this mechanism.
Soai and coworkers have developed in recent years a remarkable autocatalytic
reaction which also results in significant amplification of chirality. (Shibata et al.,
1996, 1998; Soai et al., 1999, 2000; Soai and Shibata, 1999; Sato et al., 2000a,
2000b.) In this chemistry the chiral product also functions as a catalyst for its
own production. Although this chemistry could not have occurred in the prebiotic
world, prebiotic chemistry based on the model of this chemistry could have
played a role in the origin of homochirality. Let us begin by examining the
reaction under consideration.
Treatment of achiral pyrimidine-5-carboxaldehyde which contains a carbon-
oxygen double bond (carbonyl group) with achiral diisopropylzinc [isopropyl =
-CH(CH3)2] affords a zinc alkoxide product. Treatment of the zinc alkoxide
in turn with water yields an alcohol product. The initial reaction involves the
transfer of an isopropyl group from the zinc reagent to the carbonyl carbon of
the pyrimidine reagent. Both the initially formed zinc alkoxide and the isolated
alcohol contain a single stereogenic center. Because there are no chiral forces or
optically active reagent involved in the chemistry, both products are racemic.

0 OZnCH(CH3)2
~ ~ "- ~ Jc"CH(CH3)2
H + [(CH3)2CH]2Z
n ...- -~- ~
N "N
pyrimidine-5- diisopropylzinc zinc alkoxide
carboxaldehyde /

OH
] ~ H CH(CH3)2
~
N
alcohol
250 R.N. Compton and R.M. Pagni [VI

The zinc alkoxide still has a reactive isopropyl group which also adds to the
carbonyl group of the pyrimide-5-carboxaldehyde to yield a zinc dialkoxide with
a second stereogenic center. Recall that a product with two stereogenic centers
yields a maximum of 4 stereoisomers. In this instance a pair of enantiomers
(RR and SS) in racemic form, and a meso compound (RS=SR) are generated
in unequal amounts. On treatment with water this mixture of zinc dialkoxide
stereoisomers yields the same alcohol as shown above in racemic form.
CH(CH3)2 CH(CH3) 2
,I ,I
Ar--- C ~ O ~ Z n ~ O ~ C ~Ar Ar = pyrimidine ring
I I
H H

zinc dialkoxide
RS = SR (meso)
RR and SS (enantiomeric pair)
The system becomes very interesting when a small amount of optically
active alcohol is added to diisopropylzinc before addition of the carboxaldehyde.
The added alcohol reacts with the zinc reagent to yield the optically active
(S)-zinc alkoxide. If the added alcohol has the S absolute configuration, so
does the zinc alkoxide. Before addition of the carboxaldehyde, the system
contains mostly achiral diisopropylzinc and a little (S)-zinc alkoxide. When
the carboxaldehyde is then added, it reacts preferentially with the optically
active zinc alkoxide. This can happen in two ways, creating either an S
or an R stereogenic center in the zinc dialkoxide. For steric reasons, the
second addition yields the S absolute configuration at the new center. By a
disproportionation reaction the (S)-zinc dialkoxide reacts with diisopropylzinc
to generate more (S)-zinc alkoxide which is then used to add to still more
pyrimidine-5-carboxaldehyde. Overall then, the addition of a little (S)-alcohol
to the systems yields lots of (S)-alcohol.
(CH3)2CHx (CH3)2CHx
Ar,, .)C -- OH + [(CH3)2CH]2Zn ~ Ar,,')C--OZnCH(CH3)2 + Propane
H H
(S)-alcohol (S)-zinc alkoxide

(CH3)2CHx,
...._ (CH3)2CHN /QCH3)2CH
Ar,"~C --OZnCH(CH3)2
"- ~C_O_ZnOC~,,H~....x
O
[ (S,S)-zinc dialkoxide I
Ar H original ~ new
S center J S center
J
~ [(CH3)2CH]2Zn

2 (S)-zinc alkoxide
VI] THE CHIRALITY OF BIOMOLECULES 251

Table II
Amplification of chirality in the Soai reactions

Initiator % Enantiomeric excess % Enantiomeric excess of


alcohol after one reaction

Leucine 2 (L) 21 (R)


Leucine 2 (D) 20 (S)
Valine 1 (L) 51 (R)
Valine 1 (D) 47 (S)
2-Butanol 0.1 (S) 73 (S)
2-Butanol 0.1 (R) 76 (R)
Quartz 100 (+) 89 (S)
Quartz 100 (-) 85 (R)
NaC103 100 (+) 98 (S)
NaCIO3 100 (-) 98 (R)

Even more amazing results are obtained when the added alcohol is only
slightly enriched in one enantiomer (2% enantiomeric excess). After one
reaction, the alcohol has enantiomeric excess = 10%. If this alcohol in turn
is used to initiate a second reaction, the resultant alcohol has enantiomeric
excess = 57%. After 4 such cycles, the original 2% enantiomeric excess has
become 88%, a 44-fold amplification of chirality.
Similar but even more spectacular results are also obtained on the reaction
of pyrimidine carboxaldehydes using the amino acids leucine (2% enantiomeric
excess favoring D or L) and valine (1% D or L), alcohols (0.1% enantiomeric
excess R or S), (+) and (-) quartz and (+) and (-)-NaC103 as initiators (Table II).
Amplifications of close to a thousand have been obtained after one reaction cycle
in some cases.
The results discussed above provide clear theoretical and experimental
evidence that autocatalysts can lead to enantioamplification. Autocata|ysis is,
however, not the only mechanism by which enantiomer enrichment may have
occurred in the prebiotic world. Yamagata has proposed that amplification may
have occurred by an accumulation of small enrichments during each step of
the synthesis of a polymer such as a protein or nucleic acid (Yamagata, 1966;
Yamagata et al., 1980). The starting material in this scheme is a racemic mixture
of a monomer to be incorporated into the polymer; the monomer will be called
R-M and L-M. Now allow R-M to react with itself to make a dimer (R-M)2.
A similar reaction converts L-M into (L-M)2, the enantiomer of (R-M)2. Because
of PVED, the two reactions occur at minutely different rates. Repeat the reactions
again to form (R-M)3 and (L-M)3. These also occur at slightly different rates.
252 R.N. Compton and R.M. Pagni [VII

Thus, if homopolymers (R-M)n and (L-M)n are ultimately formed, the PVED
kinetic effect will have been amplified n - 1 times. Although this amplification
would occur several thousand times in forming a protein, it would occur tens of
millions of times to form DNA. In forming the protein the PVED kinetic effect
occurs within the D- and L-amino acids, 20 of which are commonly used in life,
whereas for DNA, the differences occur in cyclic forms of the carbohydrates D-
and L-deoxyribose.
The simple scheme involving the build-up of a polymer from a single pair of
enantiomers is too simple because R-M and L-M can also react with one another
although at rates different from the R-M + R-M and L-M + L-M reactions. This
produces a pair of dimer enantiomers, (R-M)(L-M) and (L-M)(R-M). The PVED
kinetic effect will also operate when R-M and L-M react with one another. If
this "random" polymerization continues, one will obtain in theory 2 n pairs of
enantiomers for a polymer with n monomeric units. Each pair of enantiomers
will be subject to the cumulative PVED effect; each pair of enantiomers will be
formed in unequal amounts.
If the starting pool of reactants consists of 20 different racemic molecules, as
would be the case for the amino acids, the chemistry becomes incredibly complex
because each of the 40 reactants can react to form 1600 dimeric products
consisting of 800 pairs of enantiomers. The dimers would then yield 64000
trimeric products.

VII. Possible Effects of the Parity-Violating


Energy Difference (PVED) in Extended Molecular Systems

As discussed above, the PVED for individual molecules made up of elements


with small atomic numbers Z is expected to be exceedingly small. However, since
the initial work of Yamagata (1966) a number of researchers have considered the
possible effects of "amplification" of molecular PVED in extended molecular
systems such as crystals and polymers. MacDermott (1995), for example,
has attempted to relate the PVED expected for chiral quartz crystals to an
enantiomeric excess of quartz crystals on the Earth. This comparison suffers
for a lack of sufficient data on quartz crystals and possible local enantiomeric
excesses. Others have considered the kinetic effects of PVED on reaction
stereochemistry (Szabo-Nagy and Keszthelyi, 1999a; Avalos et al., 2000a,b).
From equilibrium thermodynamics, the reaction rate for either R-R or S-S
molecules is proportional to e -PVED/kT. Thus, the ratio of reaction rates for R-R
or S-S molecules is roughly proportional to PVED/kT, or epsilon for brevity.
Accordingly, a chemical reaction of achiral starting materials would yield an
enantiomeric excess, (R-S)/(R+S), of chiral products on the order of e. The
influence of this small energy difference for polymers or crystals is expected
to increase with the number of monomers or molecules in the bulk material.
VII] THE CHIRALITY OF BIOMOLECULES 253

This is expected to yield an enantiomeric excess favoring the lower-energy


enantiomer roughly in proportion to n times c. Avalos et al. (2000b) have
recently provided a complete review of this subject. All of these studies
involve a statistical analysis of chiral crystal growth. Optical rotation or circular
dichroism is used to characterize the enantiomeric excess before and after
crystallization. Crystallization from low-Z materials such as (+)- and (-)-sodium
ammonium tartrate show no appreciable enantiomeric selection (Szabo-Nagy
and Keszthelyi, 1999b). That is, crystallization of a racemic mixture of sodium
ammonium tartrate gave a Gaussian distribution of (+) and (-) crystals centered
about an enantiomeric excess of zero. However, in the cases of crystallization
of tris(1,2-ethanediamine)cobalt and tris(1,2-ethanediamine)iridium(III), both
containing atoms with large Z, the authors obtained a Gaussian distribution of
crystals which was broadened and shifted from zero relative to that of triply
distilled water or the starting racemic material. Using a statistical analysis, these
authors calculated an e of 8.3 • 10-j4 for the cobalt complex and 4.5 • 10-11 for
the iridium complex. The estimated value for the sodium ammonium complex
was ~<10-17. These data were presented as evidence for the expected increase
with atomic number Z discussed above. The interested reader should refer to
the review article by Avalos et al. (2000b) to find an in-depth analysis of some
of the controversy surrounding these interpretations.
The Z-dependence of asymmetric crystallization of chiral crystals has received
other attention. Crystallization of potassium dodecatungstosilicate in aqueous
solution is found to give a small excess of (+) crystals (Mason and Tranter,
1984). The actinide compound sodium uranyl acetate also crystallizes into a
chiral crystal (Suh et al., 1997). Interestingly, in 1954, before the discovery
of parity violation, Havinga (1954) found that the crystallization of allyl
ethyl methylanilinium iodide gives mainly dextrorotatory crystals. Our own
studies involving beta radiolysis of sodium chlorate crystals discussed earlier
also revealed a rather strange crystallization behavior of sodium bromate
which crystallizes in the same chiral space group as sodium chlorate. Careful
examination of numerous crystallizations yielding thousands of sodium bromate
crystals in a wide variety of locations (Wake Forrest, NC, USA; Christchurch,
New Zealand; Statesville, GA, USA; Knoxville and Oak Ridge, TN, USA)
under extreme conditions of cleanliness (e.g., autoclaving of solutions, filtering,
glove box conditions in rare gas or pure nitrogen atmosphere) revealed a
consistent ee of >95% favoring (+) crystals. Clearly a bias exists toward
the growth of dextrorotatory crystals. This was not seen in the case of
crystallization of the sodium chlorate solutions. In view of the results described
herein, one could logically evoke the PVED as an explanation of this result.
Currently, we are extending these measurements to sodium iodate, which
contains iodine with Z = 53. Unfortunately, this is a difficult task because the
chiral crystals are extremely small.
254 R.N. Compton and R.M. Pagni [VII

A large number of interesting, but complicated, studies, too numerous to


discuss in detail here, have been carried out on the polymerization of amino
acids and their derivatives and other molecules in the light of the origins of life
and homochirality. The interested reader is referred to the reviews of Feringa
(Feringa and van Delden, 1999) and Bonner (1988) for details. A few pertinent
points are in order about this work. (1) Peptide synthesis, which involves the
polymerization of amino acids and their derivatives or the linking of smaller
peptide units, may be self-replicating (Lee et al., 1996; Bonner, 1972) provided
the growing chain adopts an a-helical structure. In a recent paper, Saghatelian
et al. (2001) have reported such a chiroselective amplification in which a
32-residue peptide replicator demonstrated efficient amplification of homochiral
products from a racemic mixture of peptide products through a chiroselective
autocatalytic cycle. This work showed that even within a simple helical
molecular architecture, a peptide replicator can contain information capable
of self-replication, homochirality and resistance toward accumulation of errors
(stereochemical mutations). Although this article does not address the question
of specific homochirality, the chiral forces in nature as discussed above (CPL,
beta rays, electroweak force, etc.) may act as the subtle influence which provides
the initial bias toward a specific handedness. (2) Polymerization of racemic
amino-acid derivatives does not lead to any measurable enantioamplification in
the polymer or unreacted monomer. However, when one of the enantiomers of the
amino acid is in excess, enantioamplification may occur if the growing polymer
adopts an a-helical structure (Matsura et al., 1965). These results also suggest
an autocatalytic mechanism is operating in which the c~ helix preferentially
assists the incorporation of the correct amino-acid enantiomers. (3) Degradation
of the polymer via reaction with water (hydrolysis) may compete with the
polymerization itself. When the polymer being hydrolyzed is enantiomerically
impure, amplification of ee is observed in the reaction products (Blair et al.,
1981).
The non-linear effect is another way in which an initially small ee can be
amplified (Girard and Kagan, 1998). This is generally observed in the case where
an enantiomerically impure catalyst yields products with a larger ee. Although
there are many ways in which this phenomenon can occur, let us look at one
example.
Consider the case in which a reaction is catalyzed by a catalyst consisting
of a metal M and a ligand L existing in enantiomeric forms LR and Ls. The
reaction of interest creates a product with a single stereogenic center. The
reaction of LR and Ls with M yields two complexes, MLR and MLs. One of
the complexes produces the R enantiomer of the product, PR, and the other
produces Ps. In this case the enantiomeric excess of the product (eeproduct) is
linearly proportional to the enantiomeric excess of the ligand (eeligand) and thus
VII] THE CHIRALITY OF BIOMOLECULES 255

does not result in amplification. Thus, eeproduct= (eeo)(eeligand), where eeo is the
maximum enantiomeric excess that is attainable for the reaction of interest.

kR ke
M + LR ~ MLR---+ PR, M + Ls ~ MLs ' Ps.

On the other hand, if the metal contains two ligands, three catalysts, MLRLR,
MLsLs and MLsLR, are possible, each of which catalyzes the asymmetric
reaction to yield, respectively, PR, Ps and racemic product, each at its own
distinctive rate. The expression relating eeproduct to eeligand is now more complex
in that it contains/3, which is the ratio of the amount of meso complex, MLRLs,
to the sum of the homochiral complexes, MLRLR and MLsLs, and g, which is
the ratio of the rate constant for the reaction of the meso complex, kRs, to that
of the homochiral complex MLRLR, kRR (kRR = kss):

M + LR + Ls ~ MLRLR + MLsLs + MLRLs,


f I I
+ kaa + kss + kRs
+
PR Ps Racemic product
l+/q
eeproduct = (eeo)(eeligand)
1 +g/~'

where MLRLs kRs


/~ = MLRLR + MLsLs' g- kRR

When /3 = 0, i.e., the meso complex is not present, or g = l, i.e., the meso
and homochiral complexes have the same rate of reactivity, the new equation for
eeproduct reverts to the original one where a linear response is anticipated. When
g < 1, a positive non-linear effect (amplification) is expected which reaches
its maximum effect at g = 0, where the meso complex MLRLs is unreactive.
Values of g > 1 result in a negative non-linear or deamplification effect. In order
to calculate the exact positive, zero or negative non-linear effect it is necessary
to know K = (MLRLs)2/(MLRLR)(MLsLs), the equilibrium distribution of the
complexes. If the complexes are formed randomly, K = 4, for example. It should
be noted here that the chemistry of Soai described in detail above is almost
certainly an example of a positive non-linear amplification of chirality involving
a diastereomeric zinc complex of unknown structure. Diastereomeric catalysts
are required for this behavior to occur; enantiomeric complexes are excluded.
As seen above, there is significant theoretical speculation and experimental
evidence for enantioamplification. What is not always appreciated is that
racemization chemistry will always compete. Degradation chemistry will also
remove the substrate from the racemic pool being amplified. As seen earlier in
256 R.N. Compton and R.M. Pagni [VII

the article, beta radiation racemizes amino acids. Amino acids also racemize in
water with a half-life of about 6000 years, the exact value depending on the
structure of the amino acid (Bada, 1972).

COOH COOH
I water I
H2N - C --H -... H-- C --NH 2 1;1/2- 6000 years
I /
R R
L amino acid D amino acid

It is possible to date bone and teeth in archaeological samples using this fact.
Dating is difficult, however, because the half-life of racemization is environment
dependent. The optically active amino acids in the Murchison meteorite, for
example, may have remained that way for 109 years.
Carbohydrates, as noted near the beginning of the article, are polyhydroxy-
aldehydes and ketones. As is the case with all aldehydes and ketones with
similar structure, they readily form enols and enolates reversibly in the presence
of acid and base, respectively. In so doing the stereogenic center next to the
carbonyl group (C=O) is lost initially and reformed in two ways: with the original
absolute configuration or with the other absolute configuration. By such facile
pathways, D-glyceraldehyde is converted into L-glyceraldehyde, thus racemizing
the molecule. The enol and enolate also react to give an isomeric ketone,
2,3-dihydroxy-2-propanone, with no stereogenic centers. The racemization of
amino acids in waters also occurs through intermediate enols and enolates:

H\ /OH
C
II

CHO
aci
~ HO/ C \ CH2OH ~ . ~ d CHO
I enol I
H-- C --OH HO-- C --H
I I
CH2OH ~ CH2OH
D-glyceraldehyde ~ H\ /O- L-glyceraldehyde
C
II
/C\
HO CH2OH
enolate

HOCH2\
enol, enolate /c=o
HOCH2

If the carbohydrate contains more than one stereogenic center, racemization


does not occur since the absolute configuration of only the center next to the
carbonyl group is altered. Instead a diastereomer is formed if the enol/enolate
IX] THE CHIRALITY OF BIOMOLECULES 257

regenerates the center, and an isomeric compound if the center is not regenerated.
D-glucose, for example, yields the diastereomer D-mannose and the isomer
D-fructose when treated with base in water.
CHO CHO
I I HOCH2~ //O
H - - ~ --OH HO-- ~ - - H
HO-- C --H base HO-- C --H HO-- C --H
I ~ I + I
H--C --OH H--C --OH H--C --OH
I I J
H - - C --OH H - - C --OH H - - C --OH
I I I
CH2OH CH2OH CH2OH
D-glucose D-mannose D-fructose

Thus most carbohydrates will not racemize but will quite readily be converted
into diastereomeric and isomeric compounds. How such chemistry would affect
the amplification of chirality is unclear but important.
Interestingly, the carbohydrates, ribose and deoxyribose, that exist in nucleic
acids are not susceptible to the above chemistry because they exist in cyclic
forms that mask the carbonyl group. Under such circumstances enols and
enolates cannot form.
In conclusion it is fair to say that any mechanism that amplifies a small
enantiomeric excess into a larger one must do so at rates faster than mechanisms
that racemize or degrade the compound of interest.

VIII. Conclusions

Considerable progress has been made in the last few decades in understanding the
origin and evolution of the molecules of life (Brack, 1998; Mason, 1991). This
cannot be said about the origins of specific homochirality. As this article attests,
the field is active and many interesting results have been obtained, but there are
still a very large number of plausible explanations for specific homochirality.
Hopefully future research will reject some explanations and refine others to the
point where a likely cause is found. This may never occur, however, because the
time scales of some theories are vast and thereby difficult to study experimentally.
The origin may have arisen from a random event and is thus not explainable by
any theory or testable by experiment, because of our inability to go back and
see it transpire in the first place. Thus, the origins of specific homochirality may
remain one of the many mysteries of life.

IX. Acknowledgment

The authors thank the National Science Foundation for support of this work, and
graduate student Rodney Sullivan for his assistance.
258 R.N. C o m p t o n a n d R . M . P a g n i [X

X. References

Akaboshi, M., Noda, M., Kawai, K., Maki, H., and Kawamoto, K. (1978). Orig. Life 9, 181.
Akaboshi, M., Noda, M., Kawai, K., Maki, H., Ito, Y., and Kawamoto, K. (1981). Orig. Life 11, 23.
Akaboshi, M., Noda, M., Kawai, K., Maki, H., and Kawamoto, K. (1982). Orig. Life 12, 395.
Akaboshi, M., Kawai, K., Tanaka, Y., and Fujii, N. (1999). In "Advances in Biochirality" (G. Palyi,
C. Zucchi and L. Caglioti, Eds.). Elsevier, Amsterdam, p. 389.
Angel, J.R., Illing, R., and Martin, P.G. (1992). Nature 238, 389.
Arimondo, E., Glorieux, P., and Oka, T. (1977). Opt. Commun. 23, 369.
Atkins, P.W. (1978). "Physical Chemistry." W.H. Freeman, San Francisco.
Avalos, M., Babiano, R., Cintas, P., Jim6nez, J.L., Palacios, J.C., and Barron, L.D. (1998). Chem.
Rev. 98, 2391.
Avalos, M., Babiano, R., Cintas, P., Jim6nez, J.L., and Palacios, J.C. (2000a). Tetrahedron Asym.
11, 2845.
Avalos, M., Babiano, R., Cintas, P., Jim6nez, J.L., and Palacios, J.C. (2000b). Chem. Comm., p. 887.
Avertisov, V.A., Goldanskii, V.I., and Kuz'min, V.V. (1991). Phys. Today (July), p. 33.
Bada, J.L. (1972). J. Am. Chem. Soc. 94, 1371.
Bailey, J., Chrysostomon, A., Hough, J.H., Gledhill, T.M., McCall, A., Clark, S., Menard, F., and
Tamara, M. (1998). Science 281,672.
Bakusov, A., Ha, T.K., and Quack, M. (1998). J. Chem. Phys. 109, 7263.
Balavoine, G., Moradpour, A., and Kagan, H.B. (1974). J. Am. Chem. Soc. 96, 5152.
Barron, L.D. (1981). Mol. Phys. 43, 1395.
Barron, L.D. (1986a). Chem. Phys. Lett. 123, 423.
Barron, L.D. (1986b). J. Am. Chem. Soc. 108, 5339.
Barron, L.D. (1994). Science 266, 1491.
Barrow, J.D., and Silk, J. (1983). "The Left Hand of Creation: The Origin and Evolution of the
Expanding University." Basic Books, New York.
Berger, R., and Quack, M. (2000). Chem. Phys. Chem. 1, 57.
Berger, R., Quack, M., and Stohner, J. (2001). Angew. Chem. Int. Ed. 40, 1667.
Bernstein, W.J., Lemmon, R.M., and Calvin, M. (1972). In "Molecular Evolution, Prebiological and
Biological" (D.L. Rolfing and A.I. Oparin, Eds.). Plenum, New York, p. 151.
Blair, N.E., Dirbas, F.M., and Bonner, W.A. (1981). Tetrahedron 37, 27.
Blum, K., and Thompson, D.G. (1998). In "Advances in Atomic, Molecular, and Optical Physics,"
Vol. 38 (B. Bederson and H. Walther, Eds.). Academic Press, New York, p. 39.
Bonner, W.A. (1972). In "Exobiology" (C. Ponnamperuma, Ed.). North-Holland, Amsterdam, p. 170.
Bonner, W.A. (1974). J. Mol. Evol. 4, 23.
Bonner, W.A. (1988). In "Topics in Stereochemistry," Vol. 18 (E.L. Eliel and S.H. Wilen, Eds.).
Wiley-Interscience, New York, p. 1.
Bonner, W.A. (1996). In "Physical Origin of Homochirality in Life" (D.B. Cline, Ed.). American
Institute of Physics, Woodary, NY, p. 17.
Bonner, W.A. (1997). Orig. Life 21, 59.
Bonner, W.A., and Flores, J.J. (1975). Orig. Life 6, 187.
Bonner, W.A., and Lemmon, R.M. (1978a). J. Mol. Evol. 11, 95.
Bonner, W.A., and Lemmon, R.M. (1978b). Bioorg. Chem. 7, 175.
Bonner, W.A., Kavasmaneek, ER., Martin, ES., and Flores, J.J. (1974). Science 186, 143.
Bonner, W.A., Van Dort, M.A., and Yearian, M.R. (1975). Nature 258, 419.
Bonner, W.A., Van Dort, M.A., Yearian, M.R., Zeman, H.D., and Li, G.C. (1976/77). Isr. J. Chem.
15, 89.
Bonner, W.A., Lemmon, R.M., and Noyes, H.P. (1978). I Org. Chem. 43, 522.
Bonner, W.A., Blair, N.E., and Flores, J.J. (1979). Nature 28, 150.
Bouchait, M.A., and Bouchait, C. (1997). Rep. Prog. Phys. 66, 1351.
X] THE CHIRALITY OF BIOMOLECULES 259

Brack, A., ed. (1998). "The Molecular Origins of Life." Cambridge University Press, Cambridge.
Breitmaier, E. (1994). Angew. Chem. Int. Ed. Engl. 33, 1461.
Buschmann, H., Thede, R., and Heller, D. (2000). Angew. Chem. Int. Ed. 39, 4033.
Cahn, R.S., Ingold, C., and Prelog, V. (1966). Angew. Chem. Int. Ed. Engl. 5, 385.
Cairns-Smith, A.G. (1982). "Genetic Takeover." Cambridge University Press, Cambridge.
Campbell, D.M., and Farago, P.S. (1985). Nature 318, 52.
Cline, D.B., ed. (1996). "Physical Origin of Homochirality in Life." American Institute of Physics,
Woodary, NY.
Conte, E., Fanfani, G., Pieralice, M., Amerotti, R., and D'Addabbo, A. (1986). Orig. Life 17, 51.
Cronin, J., and Pizzarello, S. (1997). Science 275, 951.
Darge, W., Lacko, A., and Thiemann, W. (1976). Nature 261,522.
Darwin, C. (1906). "The Movements and Habits of Climbing Plants." John Murray, London.
Daussy, C., Marrel, T., Amy-Klain, A., Nguyen, C.T., Borde, C.J., and Chardonnet, C. (1999). Phys.
Rev. Lett. 83, 1554.
Deutsch, D.H. (1991). Chemtech 157.
Doyle, T.R., and Vogl, T. (1989). J. Am. Chem. Soc. 111, 8510.
Eliel, E.L., Wilen, S.H., and Mander, L.N. (1994). "Stereochemistry of Carbon Compounds." Wiley-
Interscience, New York.
Evans, S.V., Garcia-Garibay, M., Omkaram, N., Scheffer, J.R., Trotter, J., and Wireka, E (1986).
J. Am. Chem. Soc. 108, 5648.
Feringa, B.L., and van Delden, R.A. (1999). Angew. Chem. Int. Ed. 38, 3418.
Flores, J.J., Bonner, W.A., and Massey, G.A. (1977). J. Am. Chem. Soc. 99, 3622.
Frank, EC. (1953). Biochem. Biophys. Acta 11,459.
Frank, P., Bonner, W.A., and Zare, R.N. (2000). On the one hand but not the other: the challenge
of the origin and survival of homochirality in prebiotic chemistry, In "Chemistry for the 21st
Century" (E. Keinan and I. Schechter, Eds.). Wiley-VCH, Weinheim, ch. 11.
Frondel, C. (1978). Am. Minerol. 63, 17.
Fujima, L., Gonzales, L., Hoki, K., Manz, J., and Ohtsuki, Y. (1999). Chem. Phys. Lett. 306, 1.
Garay, A.S. (1968). Nature 219, 338.
Garay, A.S., and Ahlgren-Beckendorf, J.A. (1990). Nature 346, 451.
Gardner, M. (1990). "The New Ambidextrous Universe." W.H. Freeman, New York.
Girard, C., and Kagan, H.B. (1998). Angew. Chem. Int. Ed. 37, 2923.
Havinga, E. (1954). Biochem. Biophys. Acta. 13, 171.
Hazen, R.M., Filley, T.R., and Goodfriend, G.A. (2001). Proc. Natl. Acad. Sci. U.S.A. 98, 5487.
Hegstrom, R.A., Rein, D.W., and Sandars, P.G.H. (1980). J. Chem. Phys. 73, 2329.
Hine, J., Dowel, A.M., and Singley Jr, S.E. (1956). J. Am. Chem. Soc. 78, 479.
Hund, E (1927). Z. Phys. 43, 805.
Inoue, Y., Tsuneishi, H., Hakushi, T., Yagi, K., Awazu, K., and Onuki, H. (1996). Chem. Comm.
2627.
Iye, M., and Sugai, H. (1991). Astrophys. J. 374, 112.
Keszthelyi, L. (1994). J. Biol. Chem. 20, 241.
Khirplovich, I. (1991). "Parity Nonconservation in Atomic Phenomena." Gordon and Breach,
Amsterdam.
Kondepudi, D.K., and Durand, D.J. (2001). Chirality 13, 351.
Kondepudi, D.K., and Nelson, G.W. (1983). Phys. Rev. Lett. 50, 1023.
Kondepudi, D.K., and Nelson, G.W. (1984a). Phys. Lett. A 106, 203.
Kondepudi, D.K., and Nelson, G.W. (1984b). Physica A 125, 465.
Kondepudi, D.K., and Nelson, G.W. (1985). Nature 314, 438.
Kondepudi, D.K., and Sabanayagam, C. (1994). Chem. Phys. Lett. 217, 364.
Kondepudi, D.K., Kaufman, R., and Singh, N. (1990). Science 250, 975.
260 R.N. Compton and R.M. Pagni [X

Kondepudi, D.K., Bullock, K.L., Digits, J.A., Hall, J.K., and Miller, J.M. (1993). J. Am. Chem. Soc.
115, 10211.
Laerdahl, J.K., and Schwerdtfeger, P. (1999). Phys. Rev. A 60, 4439.
Laerdahl, J.K., Wesendrup, R., and Schwerdtfeger, P. (2000). Chem. Phys. Chem. 1, 60.
Laerdahl, J.K., Schwerdtfeger, P., and Quincy, H.M. (2002). Phys. Rev. Lett., submitted.
Lahamer, A.S., Mahurin, S.M., Compton, R.N., House, D., Laerdahl, J.K., Lein, M., and
Schwerdtfeger, E (2000). Phys. Rev. Lett. 85, 4473.
Lazzeretti, P., and Zanasi, R. (1997). Chem. Phys. Lett. 279, 249.
Lee, H.A., Granja, J.R., Martinez, J.A., Severin, K., and Ghadiri, M.R. (1996). Nature 382, 525.
Lee, T.D., and Yang, C.N. (1956). Phys. Rev. 104, 254.
Letokhov, V.S. (1975). Phys. Lett. A 53, 275.
MacDermott, A. (1995). Orig. Life Evol. Biosphere 25, 191-199.
MacDermott, A.J., Tranter, G.E., and Trainor, S.J. (1992). Chem. Phys. Lett. 194, 152.
Mahurin, S., McGinnis, M., Bogard, J.S., Hulett, L.D., Pagni, R.M., and Compton, R.N. (2001).
Chirality 13, 636.
Martin, B., Tharrington, A., and Wu, X.-L. (1996). Phys. Rev. Lett. 77, 2826.
Mason, S. (1988). Chem. Soc. Rev. 17, 347.
Mason, S.E (1984). Nature 311, 19.
Mason, S.E (1991). "Chemical Evolution." Clarendon Press, Oxford.
Mason, S.E (2000). In "Circular Dichroism: Principles and Applications," 2nd edition (N. Berora,
K. Nakanishi and R. Woody, Eds.). Wiley, New York, p. 37.
Mason, S.E, and Tranter, G.E. (1984). Mol. Phys. 53, 1091.
Masterson, B.P., and Wieman, C.E. (1995). In "Atomic Parity Nonconservation in Precision Tests of
the Standard Electroweak Model" (?. Langacker, Ed.). World Scientific, Singapore, p. 545.
Matsura, K., Inoue, S., and Tsurata, T. (1965). Macromol. Chem. 85, 284.
McBride, M. (2001). Personal communication.
Michl, J., and Thulstrup, E.W. (1986). "Spectroscopy with Polarized Light." VCH, New York.
Miller, S.L. (1959). Ann. N.Y. Acad. Sci. 69, 260.
Norden, B. (1997). Nature 266, 567.
Pagni, R.M., and Compton, R.N. (2002). Cryst. Growth Design 2, 249.
Ponnamperuma, C., and MacDermott, A. (1994). Chem. Britain (June), p. 487.
Quack, M. (1989). Angew. Chem. Int. Ed. 28, 571.
Rau, H. (1983). Chem. Rev. 83, 535.
Raupach, E., Rikken, G.L.J.A., Train, C., and Mal6zeux, B. (2000). Chem. Phys. 261,373.
Rein, D.W. (1974). J. Mol. Evol. 4, 15.
Rib6, J.M., Crusats, J., Sagu6, E, Claret, J., and Rubires, R. (2001). Science 292, 2063.
Rikken, G.L.J.A., and Raupach, E. (2000). Nature 405, 932.
Saghatelian, A., Yokobayashi, Y., Soltani, K., and Ghadiri, M.R. (2001). Nature 409, 797.
Salam, A. (1991). J. Mol. Evol. 33, 105.
Sato, I., Kadowaki, K., and Soai, K. (2000a). Angew. Chem. Int. Ed. 39, 1510.
Sato, I., Omiya, D., Saito, T., and Soai, K. (2000b). J. Am. Chem. Soc. 122, 11239.
Shibata, T., Morika, H., Hayase, T., Choji, K., and Soai, K. (1996). J. Am. Chem. Soc. 118, 471.
Shibata, T., Yamamoto, J., Matsumoto, N., Yonekubo, S., Osanai, S., and Soai, K. (1998). J. Am.
Chem. Soc. 120, 12157.
Shimizu, Y. (1997). J. Chem. Soc., Perkin Trans. 1, p. 1275.
Shimizu, Y., and Kawanishi, S. (1996a). Chem. Comm., p. 819.
Shimizu, Y., and Kawanishi, S. (1996b). Chem. Comm. p. 1333.
Soai, K., and Shibata, T. (1999). In "Advances in Biochirality" (G. Palyi, C. Zucchi and L. Caglioti,
Eds.). Elsevier, Amsterdam, p. 125.
Soai, K., Osanai, S., Kadowaki, K., Yonekubo, S., Shibata, T., and Suto, I. (1999). J. Am. Chem.
Soc. 121, 11235.
X] THE CHIRALITY OF BIOMOLECULES 261

Soai, K., Shibata, T., and Sato, I. (2000). Acc. Chem. Res. 33, 382.
Stinson, S.C. (1995). Chem. Eng. News (Oct. 9 issue), p. 44.
Stinson, S.C. (1998). Chem. Eng. News (Sept. 21 issue), p. 83.
Stinson, S.C. (2000). Chem. Eng. News (Oct. 23 issue), p. 55.
Stinson, S.C. (2001). Chem. Eng. News (May 14 issue), p. 45.
Sugai, H., and lye, M. (1995). Mont. Notice R. Astron. Soc. 276, 327.
Suh, I.-H., Park, K.H., Jensen, W.P., and Lewis, D.E. (1997). J. Chem. Ed. 74, 800.
Szabo-Nagy, A., and Keszthelyi, L. (1999a). In "Advances in Biochirality" (G. Palyi, C. Zucchi and
L. Caglioti, Eds.). Elsevier, Amsterdam, p. 367.
Szabo-Nagy, A., and Keszthelyi, L. (1999b). Proc. Natl. Acad. Sci. U.S.A. 96, 4252.
Ulbricht, T.L.V., and Vester, E (1962). Tetrahedron 18, 629.
van Wfillen, L. (2001). Chem. Phys. Chem. 2, 107.
Vester, F., Ulbricht, T.L.V., and Krauch, H. (1959). Naturwissenschafien 46, 68.
Wang, W., Yi, E, Ni, Y., Zhao, Z., Jin, X., and Tang, Y. (2000). J. Biol. Phys. 26, 51.
Wilen, S.H., Bunding, K.A., Kaschenes, C.M., and Wieder, M.J. (1985). J. Am. Chem. Soc. 107, 6997.
Wu, C., Ambler, E., Hayword, R., Hoppes, D., and Hudson, R. (1957). Phys. Rev. 105, 1413.
Yamagata, Y. (1966). J. Theor. Biol. 11,495.
Yamagata, Y., Sakihama, H., and Nakano, K. (1980). Orig. Life 70, 349.
Zadel, G., Eisenbraun, C., Wolff, G.-J., and Breitmaier, E. (1994). Angew. Chem. Int. Ed. Engl.
33, 454.
Zanasi, R., Lazzeretti, D., Ligabue, Andrea, Ligabue, Alessandro, and Soncici, A. (1999). Phys. Rev.
E. 59, 3382.
Zel'dovich, B.Y., Saakyan, D.B., and Sobel'man, I.I. (1977). JETP Lett. 25, 94.
This Page Intentionally Left Blank
ADVANCES IN ATOMIC, M O L E C U L A R , A N D O P T I C A L P H Y S I C S , VOL. 48

MICROSCOPIC ATOM OPTICS."


FROM WIRES TO A N A TOM CHIP
RON FOLMAN, PETER KROGER and JORG SCHMIEDMAYER
Physikalisches Institut, Universit?it Heidelberg, 69120 Heidelberg, Germany

JOHANNES DENSCHLA G
Institut fiir Experimentalphysik, Universita't Innsbruck, 6020 Innsbruck, Austria

CARSTEN HENKEL
Institut fiir Physik, Universitdt Potsdam, 14469 Potsdam, Germany

I. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
II. Designing Microscopic Atom Optics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
A. Magnetic Interaction ....................................... 265
B. Electric Interaction ........................................ 283
C. Traps and Guides formed by C o m b i n i n g the Interactions . . . . . . . . . . . . . . . 286
D. Miniaturization and Technological Considerations . . . . . . . . . . . . . . . . . . . . 289
III. Experiments with Free-Standing Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . 292
A. Magnetic Interaction ....................................... 292
B. Charged Wire Experiments ................................... 300
IV. Surface-Mounted Structures: The Atom Chip ......................... 303
A. Fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
B. Loading the Chip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
C. Atom Chip Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
V. Loss, Heating and Decoherence .................................. 324
A. Loss Mechanisms ......................................... 324
B. Heating ................................................ 330
C. Decoherence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335
VI. Vision and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
A. Integrating the Atom Chip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
B. Mesoscopic Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
C. Q u a n t u m Information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
VII. Conclusion ................................................ 351
viii. Acknowledgement ........................................... 351
IX. References ................................................ 352

I. I n t r o d u c t i o n
Scientific and technological progress in the last decades has proven that
miniaturization and integration are important steps towards the robust application
of fundamental physics, be it electronics and semiconductor physics in integrated
circuits, or optics in micro-optical devices and sensors. The experimental effort
described in this work aims at achieving the same for matter wave optics.

263 Copyright 9 2002 Elsevier Science (USA)


All rights reserved
ISBN 0-12-003848-X/ISSN 1049-250X/02 $35.00
264 R. Folman et al. [I

Matter wave optics beautifully illustrates quantum behavior. Realizations


using neutral atoms are attractive because of the well established techniques of
coherently manipulating internal and external degrees of freedom, and their weak
coupling to the environment. Miniaturizing electric and magnetic potentials is
essential to building versatile traps and guides for atoms at a scale < 1 ~tm which
will enable controlled quantum manipulation and entanglement. Integration with
other quantum optics, micro-optics and photonics techniques will allow for
robust creation, manipulation and measurement of atomic quantum states in these
microtraps. In our vision we see a monolithic integrated matter wave device
which will allow us to establish a new experimental toolbox and enable new
insights into fundamental quantum physics, for example in issues such as deco-
herence, entanglement and nonlinearity, low-dimensional mesoscopic systems,
and degenerate quantum gases (Bosons and Fermions) beyond mean-field theory.
A successful implementation may lead to widespread applications from highly
sensitive sensors (time and acceleration) to quantum information technology.
The goal of this review is to sum up the 10 year long exciting journey into
the miniaturization and integration of matter wave optics resulting in devices
mounted on surfaces, so called atom chips. It brought together the best of two
worlds: the vast knowledge of quantum optics and matter wave optics and the
mature techniques of microfabrication.
The first experiments started in the early 1990s with the guiding of atoms
along free-standing wires and investigating the trapping potentials in simple ge-
ometries. This later led to the microfabrication of atom-optical elements down to
1 ~m size on atom chips. Very recently the simple creation of Bose-Einstein con-
densates in miniaturized surface traps was demonstrated, and the first attempts
to integrate light optics on the atom chip are in progress. Even though there are
many open questions, we firmly believe that we are only at the beginning of a new
era of robust quantum manipulation of atomic systems with many applications.
The review is organized as follows. We begin in Sect. II by describing
microscopic atom-optical elements using current-carrying and charged structures
that act as sources for electric and magnetic fields which interact with the atom.
In the following sections we describe first the experiments with free-standing
structures - the so called atom wires (Sect. III), investigating the basic principles
of microscopic atom optics, and then the miniaturization on the atom chip
(Sect. IV). In Sect. V we discuss one of the central open questions: what happens
with cold atoms close to a warm surface, how fast will they heat up, and how
fast will they lose their coherence? The role of technical noise, the fundamental
noise limits and the influence of atom-atom and atom-surface interactions are
discussed. We conclude with an outlook of what we believe the future directions
to be, and what can be hoped for (Sect. VI).
The scientific progress regarding manipulation of atoms close to surfaces
has been enormous within the last decade. Besides the atom wire and atom
II] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 265

chip described here, it covers a whole spectrum: from reflection experiments on


atom mirrors to studying Van der Waals interactions and quantum reflection;
from using micromagnets to trap atoms to employing evanescent light field
traps. Many of these have been reviewed recently and will not be included
here. We will almost exclusively concentrate on manipulation of atoms with
static microscopic electric and magnetic fields created by charged and/or current-
carrying (microscopic) structures. For related experiments and proposals, which
are not discussed in this review, we refer the reader to the excellent reviews
referenced throughout the text, e.g. Dowling and Gea-Banacloche (1996), Grimm
et al. (2000), Hinds and Hughes (1999).

II. Designing Microscopic Atom Optics


Neutral atoms can be manipulated by means of their interaction with magnetic,
electric, and optical fields. In this review the emphasis is put on the magnetic
and the electric interaction. The designing of traps and guides using charged and
current-carrying structures and the combination of different types of interaction
to form devices for guided matter wave optics are discussed. It is shown how
miniaturization of the structures leads to great versatility where a variety of
potentials can be tailored at will. We start with some general statements and
then focus on the concepts that are important for surface-mounted structures
and address issues of miniaturization and its technological implications.

A. MAGNETIC INTERACTION

A particle with total spin F and magnetic moment/u = gFttBF experiences the
potential Vmag = - / ~ " B = --gFktBmFB, (1)

where /re is the Bohr magneton, gF the Land6 factor of the atomic hyperfine
state, and mF the magnetic quantum number. In general, the vector coupling/u. B
results in a complicated motion of the atom. However, if the Larmor precession
(mL = ItBB/h) of the magnetic moment is much faster than the apparent change of
direction of the magnetic field in the rest frame of the moving atom, an adiabatic
approximation can be applied. The magnetic moment then follows the direction
of the field adiabatically, mF is a constant of motion, and the atom is moving in
a potential proportional to the modulus of the magnetic field B = ]B].
Depending on the orientation of/u relative to the direction of a static magnetic
field, one distinguishes two cases:
(1) If the magnetic moment is pointing in the same direction as the magnetic
field (Vmag < 0), an atom is drawn towards increasing fields, therefore it is in
a strong-field seeking state. This state is the lowest energy state of the system.
266 R. Folman et al. [II

Minima o f the potential energy are found at maxima of the field. Maxima of the
magnetic field in free space are, however, forbidden by the Earnshaw theorem 1.
This means that for trapping atoms in the strong-field-seeking state, a source
of the magnetic field, such as a current-carrying material object or an electron
beam, has to be located inside the trapping region.
(2) If the magnetic moment of an atom is pointing in the direction opposite
to the magnetic field (Vmag > 0), the atom is repelled from regions with high
magnetic fields; it is then in the metastable weak-field seeking state. In this case,
minima o f the modulus of the field correspond to potential minima. Because a
minimum of the modulus of the magnetic field in free space is not forbidden
by the Earnshaw theorem, traps of this type are most c o m m o n for neutral atom
trapping. Losses from the traps are a potential problem (see Sect. V), especially
when non-adiabatic transitions to the energetically lower high-field-seeking states
become likely in regions of low or even vanishing fields.

A. 1. Kepler guide
A possible realization of a trap for an atom in the strong-field-seeking state is
a current-carrying wire with the atom orbiting around it (Vladimirskii, 1961;
Schmiedmayer, 1992, 1995a,b; Schmiedmayer and Scrinzi, 1996a,b; Denschlag,
1998; Denschlag et al., 1999b). The interaction potential is given by 2

1
Vmag = - / I / - B = - ~ Iw-%'la,r (2)

where Iw is the current through the wire, % is the azimuthal unit vector in
cylindrical coordinates, and /z0 = 4 ; r m m G / A is the vacuum permeability.
This potential has the 1/r form of a Coulomb potential, but the coupling
/ u - B is vectorial. Using the adiabatic approximation, Vmag corresponds to a
2-dimensional scalar ( l / r ) potential, in which atoms move in Kepler orbits 3.
In the quantum regime, the system looks like a 2-dimensional hydrogen atom
in a (nearly circular) Rydberg state. The wire resembles the "nucleus" and

l The Earnshaw theorem can be generalized to any combination of electric, magnetic and
gravitational fields (Wing, 1984; Ketterle and Pritchard, 1992).
2 This and all other expressions for magnetic and electric fields in this section are given in the limit
of an infinitely thin wire, unless stated otherwise.
3 From corrections to the adiabatic approximation to the next order, we obtain an effective
Hamiltonian for the orbital motion of the atom where the Coulomb-like binding potential is corrected
by a small repulsive 1/r2 interaction (Shapere and Wilczek, 1989; Aharonov and Stern, 1992; Stern,
1992; Littlejohn and Weigert, 1993; Schmiedmayer and Scrinzi, 1996a,b). As a result, the Kepler-
like orbits show an additional precession around the wire. A very similar potential can be realized
for small polar molecules with a permanent dipole moment interacting with the electric field of a
charged wire (Sekatskii and Schmiedmayer, 1996).
II] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 267

FIG. 1. Guiding neutral atoms using a current carrying wire. (a) Guiding the atoms in their strong
.field seeking state as they circle around the wire. (b) Atoms in the weakfield seeking state can be
held in a 2-dimensional magnetic quadrupole field which is created by adding a constant bias field
to the wire field. Typical trajectories of atoms are shown on the right-hand side of the figure.

the atom takes the place of the "electron ". Considerable theoretical work has
been published on the quantum mechanical treatment of this system showing
a hydrogen-like energy spectrum (Pron'kov and Stroganov, 1977; Blfimel and
Dietrich, 1989, 1991; Voronin, 1991 ; Hau et al., 1995; Burke et al., 1996; Berg-
Sorensen et al., 1996) with a characteristic quantum defect (Schmiedmayer and
Scrinzi, 1996a,b).
The magnetic field, the potential, and typical classical trajectories are
presented in Fig. la.

A.2. Side guide


Originally, Frisch and Segr6 (1933) presented the idea that a straight current-
carrying wire (Iw) and a homogeneous bias field (Bb) pointing in a direction
orthogonal to the wire form a quadrupole field with a well-defined 2-dimensional
field minimum (Fig. lb). The bias field cancels the circular magnetic field of the
wire along a line parallel to the wire at a distance

r0 = ~-~)
{ ~to s
Bb
(3)
Around this line the modulus of the magnetic field increases in all directions and
forms a tube with a magnetic field minimum at its center. Atoms in the weak-
field seeking state can be trapped in this 2-dimensional quadrupole field and can
268 R. Folman et al. [II

| | |

| | | | | | |

Fie. 2. Upper left: potential for a side guide generated by one wire and an external bias field
perpendicular to the wire direction. The external bias field can be replaced by two extra wires (lower
left). Upper right: field configuration for a two-wire guide with an external bias field perpendicular
to the plane containing the wires. This external bias field may also be replaced by surface mounted
wires (lower right).

be guided along the side of the wire, i.e. in a side guide. At the center of the
trap the magnetic field gradient is

(4)
dr ro Iw ro

As long as the bias field is orthogonal to the wire, the two fields cancel exactly,
and trapped atoms can be lost due to Majorana transitions between trapped and
untrapped spin states (see Sect. V.A). This problem can be circumvented by
adding a small B-field component Bip along the wire direction which lifts the
energetic degeneracy between the trapped and untrapped states. This potential is
conventionally called a Ioffe-Pritchard trap (Gott et al., 1962; Pritchard, 1983;
Bagnato et al., 1987). At the same time, the potential form of the guide near the
minimum changes from linear to harmonic. The guide is then characterized by
the curvature in the transverse directions

dT 2 r~ = -~- gipI 2 T2 g i p . (5)

In the harmonic oscillator approximation, the trap frequency is given by

2Jr 2:r V M -ff~r2 J (3( --r0 m-Oip ' (6)

where M is the mass of the atom.


When mounting the wire onto a surface, the bias field has to have a component
parallel to the surface in order to achieve a side guide above the surface. The bias
field can be formed by two additional wires on each side of the guiding wire. The
direction of the current flow in these wires has to be opposite to the current in
the guiding wire (Fig. 2). This is especially interesting because the wires can be
mounted on the same surface (chip), and a self-sufficient guide can be obtained.
II] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 269

A.3. Two-wire guides


A.3.1. Counter-propagating currents. A different way to create a guide is by
using two parallel wires with counter-propagating equal currents Iw with a bias
field which has a component Bb orthogonal to the plane containing the two wires
(Fig. 2) (Thywissen et al., 1999a).
The important advantage of this configuration is that the two wires and
therefore the atom guide can be bent in an arbitrary way in the plane
perpendicular to the bias field, whereas the single-wire guide direction is
restricted to angles close to the line perpendicular to the bias field. If there is
an additional bias field Bip applied along the wires, a Ioffe-Pritchard guide is
obtained. Again, two added wires can replace the external bias field (see Sect. IV
for an example of an experimental implementation).
The field generated by the wires compensates the bias field Bb at a distance

ro= -Y- 1, (7)

where d is the distance between the two wires. When Bb > 2~Iw/Jrd, the field
from the wires is not capable of compensating the bias field. Two side guides
are then obtained, one along each wire in the plane of the wires.
In the case Bb < 2ltolw/:rd, the gradient in the confining directions is given
by dB r0
d r ro = --~ Iw d (8)

If there is a field component Bip along the wire, the position of the guide is
unchanged. However, the shape of the potential near its minimum is parabolic:
the curvature in the radial direction is given by

d2B
dr 2 r0 = BipI2 de. (91

In the special case of r0 = d/2, the gradient and, for the case of a non-vanishing
Bip, the curvature of the potential at the minimum position, are exactly equal to
the corresponding magnitudes for the single-wire guide.

A.3.2. Co-propagating currents. The magnetic fields formed by two parallel


wires carrying equal co-propagating currents vanishes along the central line
between the wires and increases and changes direction like a 2-dimensional
quadrupole. The wires form a guide as shown in Fig. 3 allowing atoms to
be guided around curves (Mfiller et al., 1999). It is even possible to hold
atoms in a storage ring formed by two closed wire loops (Sauer et al., 2001)
270 R. Folman et al. [II

FIG. 3. Atoms are guided in a two-wire guide that is self-sufficient without external bias fields.
Insets (a), (b) and (c) show the magnetic field contour lines for no bias, horizontal bias, and vertical
bias fields, respectively. Courtesy E. Cornell.

FIG. 4. Potential for a two-wire guide formed by copropagating currents. The plots show from
left to right the equipotential lines for increasing bias fields. As the field is raised, two (quadrupole)
minima approach each other in the vertical direction and merge at the characteristic bias field denoted
by B = 1 into a harmonic (hexapole) minimum. At higher bias fields this minimum splits into a
double (quadrupole) well again; this time the splitting occurs in the horizontal direction.

(Sect. III.A.7). When aiming at miniaturized, surface-mounted structures, the


fact that the potential minimum is located between the wires rather than above
them, has to be considered.
When a bias field parallel to the plane of the wires is added, the potential
minimum moves away from the wire plane and a second quadrupole minimum
is formed at a distance far above the wire plane where the two wires appear as a
single wire carrying twice the current (see side guide in Sect. II.A.2). Depending
on the distance d between the wires with respect to the characteristic distance

/
dsplit =
\
5-~)
Ito Iw
Bb
(lO)
one observes three different cases (Fig. 4): (i) If d/2 < dsplit, two minima are
created one above the other on the axis between the wires. In the limit of d
going to zero, the barrier potential between the two minima goes to infinity and
the minimum closer to the wire plane falls onto it; (ii) if d/2 = dsplit, the two
minima fuse into one, forming a harmonic guide; (iii) if d/2 > dsplit, t w o minima
are created, one above each wire. Splitting and recombination can be achieved
by simply increasing and lowering the bias field (Denschlag, 1998; Zokay and
Garraway, 2000; Hinds et al., 2001).
II] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 271

Table I
Typical potential parameters for wire guides, based on tested atom chip componentsa

Atom Wire Bias fields Potential Ground state


current
Bb Bip Depth Distance Gradient Frequency Size Lifetime
[mA]
[G] [G] [mK] [~tm] [kG/cm] [kHz] [nm] [ms]

Side guide b
Li 1000 80 2 5.4 25 32 100 120 > 1000
Li 500 200 10 13 5 400 570 50 > 1000
Li 200 400 30 27 1 4000 3300 21 7
Rb 1000 80 1 5.4 25 32 41 53 >1000
Rb 500 200 4 13 5 400 250 21 > 1000
Rb 200 400 20 27 1 4000 1100 10 > 1000
Rb 1000 2000 50 130 1 20000 3600 6 > 1000

Two-wire guideC (counter-propagating currents)


Li 1000 80 2 5.4 25 32 100 120 > 1000
Li 500 200 10 13 5 400 570 50 > 1000
Li 100 130 l0 8.7 1.5 870 1200 34 5
Rb 1000 80 1 5.4 25 32 41 53 >1000
Rb 500 200 4 13 5 400 250 21 > 1000
Rb 100 130 5 8.7 1.5 870 490 15 185

a The parameters are given for the two different atoms lithium and rubidium, both assumed to be
in the (internal) ground state with the strongest confinement (F = 2, m F = 2). For both types
of guide, small bias-field components Bip pointing along the guide were added in order to get a
harmonic bottom of the potential and to enhance the trap life time that is limited by Majorana
spin flip transitions (see Eq. 18 in Sect. V). It was confirmed in a separate calculation that the
trap ground state is always small enough to fully lie in the harmonic region of the Ioffe-Pritchard
potential. See also Fig. 2.
b Side guide created by a thin current-carrying wire mounted on a surface with an added bias field
parallel to the surface but orthogonal to the wire.
c Two-wire guide created by two thin current-carrying wires mounted on a surface with an added
bias field orthogonal to the plane of the wires. In these examples the two wires are 10 ~tm apart.

Finally we mention a proposal by Richmond et al. (1998) where a tube


c o n s i s t i n g o f two i d e n t i c a l , i n t e r w o u n d s o l e n o i d s c a r r y i n g e q u a l b u t o p p o s i t e
c u r r e n t s c a n be u s e d as a w e a k - f i e l d - s e e k e r guide. T h e m a g n e t i c field is a l m o s t
z e r o t h r o u g h o u t the c e n t e r o f the t u b e , b u t it i n c r e a s e s e x p o n e n t i a l l y as o n e
a p p r o a c h e s the w a l l s f o r m e d b y the c u r r e n t - c a r r y i n g w i r e s . H e n c e , c o l d l o w -
f i e l d - s e e k i n g a t o m s p a s s i n g t h r o u g h the t u b e s h o u l d b e r e f l e c t e d b y the h i g h
m a g n e t i c fields n e a r the w a l l s , w h i c h f o r m a m a g n e t i c m i r r o r .
Examples of typical guiding parameters for the a l k a l i a t o m s l i t h i u m a n d
r u b i d i u m t r a p p e d in s i n g l e a n d t w o - w i r e g u i d e s are g i v e n in Table I. T r a p
f r e q u e n c i e s o f the o r d e r o f 1 M H z or a b o v e c a n be a c h i e v e d w i t h m o d e r a t e
272 R. Folman et al. [II

currents and bias fields. The guided atoms are then located a few ~tm from the
wire (above the surface).

A.4. Simple traps


An easy way to build traps is to start from the guides discussed above, and
close the trapping potential with 'endcaps'. This can be accomplished by taking
advantage of the fact that the magnetic field is a vector field, and the interaction
potential is scalar (Eq. 1). By varying the angle between the wire and the bias
field, one can change the minimum of the potential and close the trap. Simple
geometries are either a straight guide and an inhomogeneous bias field, or a
homogeneous bias field in combination with a bent wire.

A.4.1. Straight guide and an inhomogeneous bias field. Traps formed by


superposing an inhomogeneous bias field and the field of a straight wire are based
on quadrupole fields because the complete change of direction in addition to the
inhomogenity is needed to close the trap. An interesting fact is that a current-
carrying wire on the symmetry axis of a quadrupole field can be used to 'plug'
the zero of the field. In this configuration a ring shaped trap is formed (Fig. 5a)
that has been demonstrated experimentally (Denschlag, 1998; Denschlag et al.,
1999a). In the Ttibingen (formerly Munich) group of C. Zimmermann a modified
version of this type of trap with the wire displaced from the quadrupole axis

FIG. 5. Creating wire traps: The upper row shows the geometry of various trapping wires,
the currents and the bias fields. The lower column shows the corresponding radial and axial
trapping potential. (a) A straight wire on the axis of a quadrupole bias field creates a ring-shaped
3-dimensional non-zero trap minimum. (b) A "U"-shaped wire creates a field configuration similar
to a 3-dimensional quadrupole field with a zero in the trapping center. (c) For a "Z"-shaped wire a
Ioffe-Pritchard type trap is obtained.
II] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 273

(Fortagh et al., 1998, 2000) was used to create a Bose-Einstein condensate on


an atom chip (Ott et al., 2001).

A.4.2. Bent wire traps: the U- and Z-trap. 3-dimensional magnetic traps can be
created by bending the current-carrying wire of the side guide (Cassettari et al.,
1999; Reichel et al., 1999; Haase et al., 2001). The magnetic field from the bent
leads creates endcaps for the wire guide, confining the atoms along the central
part of the wire. The size of the trap along this axis is then given by the distance
between the endcaps. Here we describe two different geometries:
(1) Bending the wire into a "U"-shape (Fig. 5b) creates a magnetic field
that in combination with a homogeneous bias field forms a 3-dimensional
quadrupole trap 4. The geometry of the bent leads results in a field
configuration where a rotation of the bias field displaces the trap minimum
but the field always vanishes completely at this position.
(2) A magnetic field zero can be avoided by bending the wire ends to form a
"Z" (Fig. 5c). Here, one can find directions of the external bias field where
there are no zeros in the trapping potential, for example when the bias field is
parallel to the leads. This configuration creates a Ioffe-Pritchard type trap.
The potentials for the U- and the Z-trap scale similarly as for the side guide,
but the finite length of the central bar and the directions of the leads have to
be accounted for. Simple scaling laws only hold as long as the distance of the
trap from the central wire is small compared to the length of the central bar
(Cassettari et al., 1999; Reichel et al., 1999; Haase et al., 2001). Bending both
Z leads once more results in 3 parallel wires. This supplies the bias field for a
self-sufficient Z-trap.

A.4.3. Crossed wires. Another way to achieve confinement in the direction


parallel to the wire in a side guide is to run a current ll < lw through a second
wire that crosses the original wire at a right angle (Reichel et al., 2001). Ii
creates a magnetic field B~ with a longitudinal component which is maximal at
the position of the side guide that is closest to the additional wire. Adding a
longitudinal component to the bias field, i.e. rotating Bb, results in an attractive
potential confining the atoms in all three dimensions. As a side effect position
and shape of the potential minimum are altered by the vertical component of B~.
Figure 6 illustrates this type of trap configuration. Experiments of the Munich
group have proven this concept to be feasible (see Sect. IV.C.1 and Fig. 34) and
it was suggested to use the two-wire cross as a basic module for implementing
complex trapping and guiding geometries.

4 The minimum of the U-trap is displaced from the central point of the bar, in a direction opposite
to the bent wire leads. A more symmetric quadrupole can be created by using 3 wires in an
H configuration. There the side guide is closed by the two parallel wires crossing the central wire
orthogonally. The trap is then in between the two wires, along the side guide wire.
274 R. Folman et al. [II

Flo. 6. Two geometries of crossed-wire traps: different cuts through the potential are displayed
without and with a longitudinal bias field component in the left and right column, respectively. The
1-dimensional plots show the potential along the direction of the side guide; in the contour plots
the wire configuration is illustrated by light gray bars. Courtesy J. Reichel.

A.5. Weinstein-Libbrecht traps

Even more elaborate designs for traps than those described previously can be
envisioned. For example, Weinstein and Libbrecht (1995) describe planar current
geometries for constructing microscopic magnetic traps (multipole traps, Ioffe-
Pritchard traps and dynamical traps). We focus here on the Ioffe-Pritchard trap
proposals. Figure 7 shows four possible geometries: (a) three concentric half
loops; (b) two half loops with an external bias field; (c) one half loop, one full
loop and a bias field; (d) two full loops with a bias field and external Ioffe
II] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 275

a) 111 b) I1

111 1!
cl 111 d]
| |

| |

111
FIG. 7. Four planar (and pseudoplanar) Ioffe trap configurations, as described in the text. Courtesy
J. Weinstein/K. Libbrecht.

bars. The first of these (a) is essentially a planar analog of the nonplanar Ioffe-
Pritchard trap with two loops and four bars. Configuration (b) replaces one of the
loops with a bias field. Configuration (c) is similar to (b) but provides a steeper
trapping potential on-axis and weaker trapping in the perpendicular directions;
this makes an overall deeper trap with greater energy-level splitting for given
current and size. (d) is a hybrid configuration, which uses external (macroscopic)
|offe bars to produce the 2-dimensional quadrupole field, while deriving the on-
axis trapping fields from two loops and a bias field. Typical energy splittings
in the range of 1 MHz are achievable using experimentally realistic parameters
(Drndi6 et al., 1998).

A.6. Arrays of traps


The various tools for guiding and trapping discussed above can be combined
to form arrays of magnetic microtraps on atom chips. Particularly suitable for
this purpose is the technique of the crossed wires which requires, however, a
multilayer fabrication of the wires on the surface. Arrays of traps and their
applications, especially in quantum information processing, are discussed in
Sect. VI.

A. 7. Moving potentials
Introducing time-dependent potentials facilitates arbitrary movement of atoms
from one location to another. There are different proposals for possible
276 R. Folman et al. [II

Yl
IH2
FIG. 8. Magnetic 'conveyor belt': The wires are configured in a way that allows to transport atoms
from one trap to another along a side guide. Together with a homogeneous time-independent bias
field, the currents IQ, IH1, and /H2 are used for the confining fields of the source and collecting
traps, I0 is the current through the side guide wire. The currents IM1 and IM2 alternate sinusoidally
with a phase difference of Jr/2 and provide the moving potential. Courtesy J. Reichel.

implementations of such 'motors' or 'conveyor belts', one of which has already


been demonstrated experimentally by Hansel et al. (2001b): Using solely
magnetic fields it is based on an approximation of the crossed-wire configuration.
Atoms trapped in a side guide potential are confined in the longitudinal direction
by two auxiliary meandering wires (Fig. 8). By running an alternating current
through both auxiliary wires with a relative phase difference of Jr/2, the potential
minimum moves along the guide from one side to the other in a controllable way.
In Sect. IV we present experimental results of the above scheme.

A. 8. Beam splitters
By combining two of the guides described above, it is possible to design
potentials where at some point two different paths are available for the atom.
This can be realized using different configurations (examples are shown in Fig. 9)
some of which have already been demonstrated experimentally (see Sect. IV).

A.8.1. Y-beam splitters. A side guide potential can be split by forking an


incoming wire into two outgoing wires in a Y-shape (Fig. 9a). Similar potentials
have been used in photon and electron interferometers 5 (Buks et al., 1998).
A Y-shaped beam splitter has one input guide for the atoms, that is the central
wire of the Y, and two output guides corresponding to the right and left wires.
Depending on how the current Iw in the input wire is sent through the Y, atoms
can be directed to the output arms of the Y with any desired ratio. This simple
configuration has been investigated by Cassettari et al. (2000) (see Sects. III.A.3
and IV.C.3 for experimental realizations). Its disadvantages are: (1) In a single-
wire Y-beam splitter the two outgoing guides are tighter and closer to the surface
than the incoming guide. The changed trap frequency and the angle between

5 The Y-configuration has been studied in quantum electronics by Palm and Thyl6n (1992) and
Wesstr6m (1999).
II] M I C R O S C O P I C ATOM OPTICS: F R O M WIRES TO ATOM CHIP 277

FIG. 9. Different wire geometries for a beam splitting potential: The plots show the wire
arrangement on the surface of an atom chip, and the directions of current flow and the additional bias
field. Each picture also shows a typical equipotential surface to illustrate the shape of the resulting
potential. (a) A simple Y-beam splitter consisting of a single wire that is split into two: The output
side guides are tighter and closer to the surface than the input guide. Note that a second minimum
closer to the chip surface occurs in the region between the wire splitting and the actual split point
of the potential; (b) a two-wire guide split into two single-wire guides does not exhibit this 'loss
channel'. (c) Here, the output guides have the same characteristics as the input guide, minimizing
the backscattered amplitude. The vertical orientation of the bias field ensures exact symmetry of
the two output guides. (d) In an X-shaped wire pattern the splitting occurs because of tunneling
between two side guides in the region of close approach of the two wires.

incoming and outgoing wires lead to a change o f field strength at the guide
minimum and can cause backscattering from the splitting point. (2) In the Ioffe-
Pritchard configuration (i.e. with an added longitudinal bias field), the splitting
is not fully symmetric due to different angles of the outgoing guides relative to
the bias field. (3) A fourth guide leads from the splitting point to the wire plane,
i.e. to the surface of the chip.
The backscattering and the inaccessible fourth guide o f the Y-beam splitter
may be overcome, at least partially, by using different beam splitter designs,
like those shown in Fig. 9b,c. The configuration in Fig. 9b has two wires which
run parallel up to a given point and then separate. If the bias field is chosen
so that the height of the incoming guide is equal to the half distance d/2 o f
the wires (d/2 = dsplit as defined in Eq. 10 in Sect. II.A.3), the height o f the
potential m i n i m u m above the chip surface is maintained throughout the device
(in the limit of a small opening angle) and no fourth port appears in the splitting
region. The remaining problem o f the possible reflections from the potential in
the splitting region can be overcome by the design presented in Fig. 9c. Here, a
guide is realized with two parallel wires with currents in opposite directions and
a bias field perpendicular to the plane of the wires. This type of design creates a
truly symmetric beam splitter where input and output guides have fully identical
characteristics.
278 R. Folman et al. [II

A.8.2. X-beam splitters. A different possible beam splitter geometry relies


on the tunneling effect: Two separate wires are arranged to form an X, where
both wires are bent at the position of the crossing in such a way that they do not
touch (see Fig. 9d). An added horizontal bias field forms two side guides that are
separated by a barrier that can be adjusted to be low enough to raise the tunneling
probability considerably at the point of closest approach. If the half distance
between the wires becomes as small as dsplit (Eq. 10), the barrier vanishes
completely, resulting in a configuration that is equivalent to a combination of two
Y-beam splitters (Miiller et al., 2000). The choice of the parameters in the wire
geometry, the wire current and the bias field governs the tunneling probability
and thereby the splitting ratio in this type of beam splitter. The relative phase shift
between the two split partial waves in a tunneling beam splitter allows to combine
two beam splitters to form a Mach-Zehnder interferometer. Another advantage
of the X-beam splitter is that the potential shape in the inputs and outputs stays
virtually the same all over the splitting region as opposed to the Y-beam splitter.
For a detailed analysis of the tunneling X-beam splitter see Andersson et al.
(1999).

A.8.3. Quantum behavior o f X- and Y-beam splitters. For an ideal symmetric


Y-beam splitter, coherent splitting for all transverse modes should be achieved
due to the definite parity of the system (Cassettari et al., 2000). This was
confirmed with numerical 2-dimensional wave packet propagation for the lowest
35 modes. The 50/50 splitting independent of the transverse mode is an important
advantage over four-way beam splitter designs relying on tunneling such as the
X-beam splitter described above. For the X-beam splitter, the splitting ratios for
incoming wave packets are very different for different transverse modes, since
the tunneling probability depends strongly on the energy of the particle. Even
for a single mode, the splitting amplitudes, determined by the barrier width and
height, are extremely sensitive to experimental noise.

A.9. Interferometers

Following the above ideas of position-dependent multiple potentials and time-


dependent potentials which are able to split minima in two and recombine them,
several proposals for chip-based atom interferometers have been put forward
(Hinds et al., 2001; H/insel et al., 2001c; Andersson et al., 2002).

A.9.1. Interferometers in the spatial domain. To build an interferometer for


guided atoms (Andersson et al., 2002) two Y-beam splitters can be joined back
to back (Fig. 10a). The first acts as splitter and the second as recombiner. The
eigenenergies of the lowest transverse modes along such an interferometer in
II] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 279

FIG. 10. Basic properties of the guided matter wave interferometer: (a) Two Y-beam splitters are
joined together to form the interferometer. (b) Transverse eigenfunctions of the guiding potentials
in various places along the first beam splitter. When the two outgoing guides are separated far
enough, i.e. no tunnelling between left and right occurs, the symmetric and antisymmetric states
become degenerate. (c) Energy eigenvalues for the lowest transverse modes as they evolve along the
interferometer. One clearly sees that pairs of transverse eigenstates form disjunct interferometers.
(d) The wavefunction of a cold atom cloud starts out in the vibrational ground state of a guide or
trap. The wavefunction splits when the guide divides, leaving a part of the wavefunction in each
arm of the interferometer. If the phases of the two parts evolve identically on each side, then the
original ground state is recovered when the two parts of the wavefunction are recombined. But if a
phase difference of Jr accumulates between the two parts (for example due to different gravitational
fields acting on them), then recombination generates the first excited state of the guide with a node
in the center. Courtesy E. Hinds. (e) 2-dimensional plots of a wave packet propagating through a
guided matter wave interferometer for 10) and 11) incoming transverse modes, calculated by solving
the time-dependent Schr6dinger equation in two spatial dimensions (x, z, t) for realistic guiding
potentials, where z is the longitudinal propagation axis. The probability density of the wave function
just before entering, right after exiting the interferometer, and after a rephasing time t are shown
for a phase shift of ~ . One clearly sees the separation of the two outgoing packets due to the
energy conservation in the guide, e.g. for n = 0 the first excited outgoing state is slower than the
ground state.
280 R. Folman et al. [II

2-dimensional geometry 6 are depicted in Fig. 10c. From the transverse mode
structure one can see that there are many disjunct interferometers in Fock space.
Each of them has two transverse input modes (]2n) and [2n + 1), n being the
energy quantum number of the harmonic oscillator) and two output modes. In
between the two Y-beam splitters, the waves propagate in a superposition o f
]n)l and ]n)r in the left and right arm, respectively. With adiabaticity fulfilled,
the disjunct interferometers are identical.
Considering any one of these interferometers, an incoming transverse state
evolves after the interferometer into a superposition o f the same and the
neighboring transverse outgoing state (Fig. 10c), depending on the phase
difference acquired between [n)l and In)r during the spatial separation o f the
wave function 7. While the propagation remains unchanged if the emerging
transverse state is the same as the incoming state, a transverse excitation
or de-excitation translates into an altered longitudinal propagation velocity
(Ao _~ +oo/k where hk is the m o m e n t u m o f a wave packet moving through
the interferometer and to/2:r is the transverse trapping frequency), since
transverse oscillation energy is transferred to longitudinal kinetic energy, and
vice versa.
As presented in Fig. 10e, integrating over the transverse coordinate results in a
longitudinal interference pattern observable as an atomic density modulation. As
all interferometers are identical, an incoherent sum over the interference patterns
of all interferometers does not smear out the visibility of the fringes.

A.9.2. Interferometers in the time domain. Two different proposals are based
on time-dependent potentials (Hinds et al., 2001 ; Hfinsel et al., 2001c). These
proposals differ from the interferometer in the spatial domain in several ways:
(1) The adiabaticity of the process may be controlled to a better extent due to
easier variation of the splitting and recombination time. (2) The interferometers
are based on a population of only the ground state. (3) The interference signal
amounts to different transverse state populations in the recombined single
minimum trap, whereas the above proposal anticipates a spatial interference
pattern which may be easier to detect.
The first proposal (Hinds et al., 2001) is based on a two parallel wire
configuration with co-propagating currents (see Sect. II.A.3). Changing the bias
field in this configuration as a function of time produces cases (i), (ii), and (iii)
discussed in Sect. II.A.3 depending on the strength of the bias field as compared

6 In 2-dimensional confinement the out of plane transverse dimension is either subject to a much
stronger confinement or can be separated out. For experimental realization see Gauck et al. (1998),
Spreeuw et al. (2000), Hinds et al. (2001), Pfau (2001).
7 The relative phase shift Aq~between the two spatial arms of the interferometer can be introduced
by a path length difference or by adjusting the potentials to be slightly different in the two arms. In
general, Ar is a function of the longitudinal momentum k.
II] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 281

to the critical bias field Be = -Y


~o (-3-)"
1~ Starting with Bb < Bc and an atom cloud in
the ground state of the upper minimum, a coherent splitting of the corresponding
wave function is achieved when Bb is raised to be larger than Bc. As shown in
Fig. 10d, the symmetry of the wave function now depends on the relative phase
shift introduced between its two spatially separated parts. Thus, when the bias
field is lowered again to Bb - Be, a superposition of the symmetric and the
antisymmetric state forms in the recombined guide.
If the spatial resolution of the detection system is not sufficient to distinguish
between the two output states, the following scheme is proposed: The node plane
of the excited state is rotated by 90 ~ by turning an additional axial bias field while
the guides are combined. If after such an operation the bias field is lowered,
atoms in the ground state go to the upper guide whereas the population of the
excited state is found in the lower guide.
The second proposal (H~insel et al., 2001c) utilizes the crossed-wire concept
introduced in Sect. II.A.4. Here, in contrast to the interferometer described above,
the splitting of the atomic wavefunction occurs in one dimension whereas the
confinement in the other two dimensions is the constant strong confinement of a
side guide. Longitudinally, the atoms are trapped by two currents running through
wires crossing the side guide wire. The resulting Ioffe-Pritchard potential well
is split into a double well and then recombined by a third crossing wire carrying
a time-dependent current flowing in the opposite direction.
Starting with a wavefunction in the ground state of the combined potential,
a relative phase shift introduced between the two parts of the potential after
splitting leads to a wavefunction in a (phase-shift dependent) superposition
of the ground and first excited states upon recombination. A state-selective
detection then displays a phase-shift dependent interference pattern. A detailed
analysis of realistic experimental parameters has shown that in this scheme
non-adiabatic excitations to higher levels can be sufficiently suppressed. The
position and size of the wavefunction are unchanged during the whole process.
Therefore, the interferometer is particularly well suited to test local potential
variations.

A. 10. Permanent magnets

Although beyond the scope of this chapter, we mention configurations with


permanent magnets (Sidorov et al., 1996; Meschede et al., 1997; Saba et al.,
1999; Hinds and Hughes, 1999; Davis, 1999). Though less versatile in the sense
of not enabling the ramping up and down of fields, permanent magnets might
reward us with advantages such as less noise, strong fields, and large-scale
periodic structures. As described in Sect. V, technical noise in the currents which
induce the magnetic fields may have severe consequences in the form of heating
and decoherence. In the framework of extremely low decoherence, such as that
282 R. F o l m a n et al. [II

B, G
(a) (b)
/ ~ 2so
200
150
100
50
Z, m
0.5 1 1.5 2 2.5 3 3.5
I B, G

25o~ (c)

lmm lOOi \ \ / ....


50 " ~ ' ~ '

X
0.5 1 1.5 2

FIc. 11. (a) Two pairs of differently sized magnetic sheets (bottom) are magnetized using
current-carrying wires wound around them. The choice of the direction of current flow in these
wires establishes the direction of magnetization: the arrows show a possible configuration for which
the equipotential lines are plotted (top). (b) The field produced by the sheet pairs measured in the
symmetry plane. (c) Scaling of the field due to the combined inner and outer pair of sheets in the
plane of symmetry. Courtesy M. Prentiss.

demanded by quantum computation proposals, permanent magnets might be a


better choice.
An interesting tool is a magnetic atom mirror formed by alternating magnetic
dipoles (Opat et al., 1992), creating an exponentially growing field strength as
the mirror is approached. This situation can be achieved by running alternating
currents in an array of many parallel wires or by writing alternating magnetic
domains into a magnetic medium such as a hard disk or a video tape. This has
been demonstrated by Saba et al. (1999) and may achieve a periodicity of the
order of 100 nm. Current-carrying structures have the disadvantage of large heat
dissipation, especially when the structure size is in the submicron region.
Another possibility is based on a combination of current-carrying wires and
magnetic materials; this was experimentally demonstrated at Harvard in the
group of M. Prentiss: Two pairs of ferromagnetic foils that were magnetized
by current-carrying wires wound around them were used for magnetic and
magnetooptic trapping (Vengalattore et al., 2001). The setup and the potential
achieved is illustrated in Fig. 11. The advantages of such a hybrid scheme
over a purely current-carrying structure are larger capturing volumes of the
traps, less heat dissipation, and enhanced trap depths and gradients because
the magnetic field of the wires is greatly amplified by the magnetic material.
II] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 283

The magnets can still be switched by means of time-dependent currents through


the wires.

B. ELECTRIC INTERACTION

The interaction between a neutral atom and an electric field is determined by


the electric polarizability a of the atom. In general, a is a tensor. For the simple
atoms we consider, i.e. atoms with only one unpaired electron in an s-state, the
electric polarizability is a scalar and the interaction can be written as

Vpol(r ) = - ~ l aE 2 (r). (11 )

B.1. Interaction between a neutral atom and a charged wire


We now consider the interaction of a neutral polarizable atom with a charged wire
(line charge q) inside a cylindrical ground plate (Hau et al., 1992; Schmiedmayer,
1995a; Denschlag and Schmiedmayer, 1997; Denschlag et al., 1998).
The interaction potential (in cylindrical coordinates) given by

1 ) 2 2aq2
Vp~ = - 4n'e0 r2 (12)

is attractive. It has exactly the same radial form ( 1/r 2) as the centrifugal potential
barrier (VL = LZ/2Mr 2) created by an angular momentum Lz. VL is repulsive. The
total Hamiltonian for the radial motion is

H = 2M +2Mr 2 4Jre0 r2 (13)


_ p2 Lz2 _ LcZrit
+ (14)
2M 2Mr 2 '
- - -

where Lcrit = ~ a ]q]/2zrc0 is the critical angular momentum characteristic for


the strength of the electric interaction. There are no stable orbits for the atom
around the wire. Depending on whether Lz is greater or smaller than Lcrit , the
atom either falls into the center and hits the wire (]Lz ] < Lcrit) or escapes from the
wire towards infinity ([Lzl > Lcrit). In the quantum regime, only partial waves
with hl < Lcrit (l is the quantum number of the angular momentum Lz) fall
towards the singularity and thus the absorption cross section of the wire should
be quantized (Fig. 12).
To build stable traps and guides one has to compensate the strongly attractive
singular potential of the charged wire. This can be done either by adding a
284 R. Folman et al. [II

12 - "" 60 2101

10 ........... 58 - 208 '

8 56- 206

6 "
; . ~
54 It" 204

....... .. ~ kRw = 0 202


t::~ 4 ii.:
9
i{ 52 t~,i,~,-~ -- kR w = 0.1
9 " " "" .......... kR w = 1 ,'r
50 " M":" ..... kR w = 2 200
~'tP ..... kR w = 5

0 48 ' 198
0 1i 2, 3i 4i 5 25 21 6 2 1 7 2I8 2t9 30 ,
100 ,
101 ,
102 ,
103 ,
104 105
Line charge q in units of mcrit

FIG. 12. Theoretical absorption cross section for a charged wire. The calculations are made for
several different relative thicknesses (kRw) of the wire; the charge is given in units of the angular
momentum mcrit = Lcrit/h.

repulsive potential, for example from an atom mirror or an evanescent wave


(see Sect. II.C.1), or by oscillating electric fields (see Sect. II.B.2).

B.2. Stabilizing the motion with an oscillating electric charge:


the Kapitza wire

The motion in the attractive electric potential can be stabilized by oscillating


the charges. The mechanism is similar to the RF Paul trap (Paul, 1990) where
an oscillatory part of the electric fields creates a 3-dimensional confinement
for ions. An elementary theoretical discussion of the motion in a sinusoidally
varying potential shows that Newton's equations of motion can then be integrated
approximately, yielding a solution that consists of a fast oscillatory component
superimposed on a slow motion that is governed by an effective potential (Landau
and Lifshitz, 1976).
An example of a 2-dimensional atom trap based on a charged wire with
oscillating charge was proposed by Hau et al. (1992). By sinusoidally varying the
charge on a wire, it is possible to add an effective repulsive 1/r 6 potential which
stabilizes the motion of an atom around the wire. Sizeable electrical currents
appear when the charge of a real wire (with capacitance) is rapidly varied.
Magnetic fields are produced which interact with the magnetic moment of an
atom. This leads to additional potentials which have not been taken into account
in the original calculations.
Another AC-electrical trap with several charged wires was proposed by
Shimizu and Morinaga (1992). Their setup is reminiscent of a quadrupole mass
filter and consists of 4 to 6 charged electrodes that are grouped around the
trapping center.
II] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 285

B.3. Guiding atoms with a charged optical fiber


Stable orbits for the motion of an atom around a line charge are obtained if the
atom is prevented from hitting the wire by a strong repulsive potential near the
surface of the wire. Such a strong repulsion can be obtained by the exponential
light shift potential of an evanescent wave that is blue-detuned from an atomic
resonance. This can be realized by replacing the wire with a charged optical
fiber with the cladding removed and the blue-detuned light propagating in the
fiber (Batelaan et al., 1994). The fiber itself has to be conducting or coated with
a thin (<</l) conducting layer to allow uniform charging. For the simple case of
a TE01 mode propagating in the fiber, the light shift potential is independent of
the polar angle and the combined guiding potential is given by

Vguid(r) = AKZ(Br)- ( 1 ) 22~


4:re0 r2 , (15)

where A and B are constants that depend on specifics of the optical fiber as
well as on light power, wavelength and atomic properties (Batelaan et al., 1994).
K0 is the modified Bessel function of the second kind. Figure 13 shows a typical
example of such a potential. Cold atoms are bound in radial direction by the
effective potential but free along the z-direction, the direction of the charged
optical fiber.

2 ~ T 1

g o

~ -2 -

\ / ............. Van der Waals


-4. ~ ~ ...... evanescent wave -
~ electric interaction

0.5 1.0 1.5 2.0


radius [l~m]

FIG. 13. Typical radial potential for a neutral lithium atom trapped around a charged (5 V) optical
quartz fiber (diameter 0.5 ~tm) with 1-mW light and a detuning of A/F = 3 • 105. The attractive
potential (1/r 2) is created by the interaction of the induced dipole moment with the electric field of
the charged fiber. The repulsion is due to the evanescent wave from blue-detuned light propagating
in the fiber. Close to the wire surface the Van der Waals interaction becomes important.
286 R. Folman et al. [II

C. TRAPS AND GUIDES FORMED BY COMBINING THE INTERACTIONS

C. 1. Charged wire on a mirror


As we have seen above, a static charged wire alone cannot form the basis for
stable trapping. Cylindrical solutions such as the charged light fiber have the
disadvantage that they cannot be m o u n t e d on a surface. An alternative solution
would be to mount a charged wire onto the surface o f an atom mirror. The
combination o f the attractive 1/r 2 potential with the repulsive potential o f the
atom mirror s Vm(z) gives:

Vguid(r)=Vm(Z)--( 1 ) 22~
4;re0 r2 , (16)

where z is the height above the mirror and r the distance from the wire. This
creates a potential tube for the atoms as shown in Fig. 14 which can be viewed
as a wave guide for neutral atoms.
Typical parameters for guides f o r m e d by a magnetic mirror and a charged wire
are given by Schmiedmayer (1998). They can be very similar to the magnetic
guides discussed in Sect. II.A. Using typical mirror parameters (Roach et al.,
1995; Sidorov et al., 1996), one can easily achieve deep and narrow guides with
transverse level spacings in the kHz range for both light (Li) and heavy (Rb)
atoms.
In a similar fashion microscopic traps can be created by mounting a charged
tip (point) at or close beneath the atom mirror surface. A point charge on the
surface o f an atom mirror creates an attractive 1/r 4 interaction potential:

1 ) 2aq 2
Vpol(r) = - 8yt60 r4 (17)

where q is the tip charge. Together with the atomic mirror it forms a microscopic
cell for the atoms. It can be viewed as the atom-optical analog to a quantum dot
(Schmiedmayer, 1998; Sekatskii et al., 2001).
This approach o f combining a charged structure with an atom mirror is
compatible with well-developed nanofabrication techniques. This opens up

8 There are two main types of atom mirrors: The first type utilizes evanescent waves (e.g. above
the surface of a reflecting prism) of blue-detuned light which repels the atoms. Here the potential
takes the form Vm(z) = VOexp(-tCmZ) where trm is of the order of the light wave number and z is
the distance from the mirror (Cook and Hill, 1982). The second type is based on a surface with
alternating magnetic fields. Here, ~ = 2;r/lcm is the periodicity of the alternating magnetic field.
The approaching atom experiences an exponentially increasing field, and consequently the weak-
field seekers are repelled (Opat et al., 1992; Roach et al., 1995; Sidorov et al., 1996; Hinds and
Hughes, 1999).
II] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 287

10 // - - ,

.Eo. 5 _ .
/'-'\ 1 -
-

-10 _ "~ _

-50 0 50 [! ' . . . . ]
potential [ne_V]50~ 1

=X -5 __...~
/'i......
5 10 15 20 25 30
distance from mirror [gm]

Flo. 14. Typical potential for a neutral atom guide. The attractive potential (1/r 2) is created by the
interaction of the induced dipole moment in the electric field of the charged wire mounted directly
on the surface of an atomic mirror. The action of the atomic mirror (evanescent wave or magnetic
mirror) prevents the atom from reaching the surface and creates a potential tube close to the surface
illustrated by the contour graph. The two adjacent plots give the potential in a direction orthogonal
to the charged wire and orthogonal to the mirror surface. Distances are given from the location of
the charged wire and the surface of the atom mirror.

a wide variety of possibilities ranging from curved and split guides to


interferometers or even complex networks.

C.2. Combined electric-magnetic state-dependent traps


The magnetic guides and traps (Sect. II.A) can be modified by combining
them with the electric interaction, thereby creating tailored potentials depending
on internal (e.g. spin) states. For example, supplementary electrodes located
between independent magnetic traps can be used to lower the magnetic barrier
between them by the attractive electric potential the electrodes create. Since
the magnetic barrier height depends on the magnetic substate of the atom,
whereas the electric potential does not, this allows state-selective operation. This
is especially interesting since it can lead to implementing quantum information
processing with neutral atoms in microscopic trapping potentials where the
logical states are identified with atomic internal levels (see Sect. VI).
A simple example, showing such a controllable state dependence, is a
magnetic wire guide approached by a set of electrodes (Fig. 15a). Applying a
high voltage to the electrodes introduces an electrostatic potential which provides
confinement along the direction parallel to the magnetic side guide, and also
shifts the trapping minimum towards the surface, possibly breaking the magnetic
288 R. F o l m a n et al. [II

FIG. 15. State-dependent potential: (a) top view of an actual chip design; the wire in the center
is used as a side guide wire, the additional electrodes create a spatially oscillating electric field
providing confinement also along the wire. The contour plot shows a typical potential configuration
for 7Li a t o m s in the IF = 2, m F = 2) magnetic substate using experimentally accessible parameters.
Dark areas correspond to attractive potentials, the trap minima are located 50 gm above the surface.
(b,c) The side views show that only one state (IF = 2, m F = 2)) is trapped in the combined
potential (b), while the other (IF = 1, m F = - 1 ) ) is not, because the weaker magnetic barrier to the
surface is compensated by the attractive electric potential (c). The parameters used in a simulation
with the electromagnetic field solver MAFIA were lw = 500 mA, B b = 20 G for the side guide and
a voltage of 600 V on the electrodes.

potential barrier in the direction p e r p e n d i c u l a r to the surface itself. T h e charge


can be adjusted in such a way that d e p e n d i n g on the strength o f the m a g n e t i c
barrier created by the wire current, the a t o m s either i m p a c t onto the surface
or are trapped above it. Since the strength o f the m a g n e t i c barrier d e p e n d s on
the m a g n e t i c substate o f the a t o m or, m o r e precisely, d e p e n d s linearly on the
q u a n t u m n u m b e r mF, this can be e x p l o i t e d to f o r m a state-selective m a g n e t i c
trap (Fig. 15b,c).

C.3. The e l e c t r i c m o t o r

In general, electric fields are always p r e s e n t in m a g n e t i c wire traps since an


electric potential difference is n e e d e d to drive a c u r r e n t t h r o u g h a wire w i t h finite
resistance. For large wires, the voltages in q u e s t i o n are low and if the distances
o f the a t o m s f r o m the wire are large e n o u g h , the attractive electric interaction
II] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 289

FIG. 16. Two-wire guide configuration with currents of 1 A running in opposite directions with a
vertical bias field of 150 G. The combined magnetic and electric potential is shown in contour plots
perpendicular to the wire and along the wire at the minimum height (inset). The parabolic potential
shape offers the possibility to drive the atoms (87Rb) along the wire. In the example, the voltages
applied to the wires are chosen to be 0 V with respect to ground in the wire center.

can be neglected. However, for micron-sized wires, one finds that if the current-
carrying wire becomes long, at some point the voltage is strong enough to create
a significant driving force for the atoms or even to destroy the traps.
On the other hand, one can actually exploit this effect and turn it into an
'electric motor' by using the electric potential gradient inside the magnetic
minimum to accelerate and decelerate the atoms at will. Figure 16 illustrates
the mechanism used for the motor for the example of a two-wire guide with a
vertical bias field. The wires carry counter-propagating currents, and the electric
interaction is zero in the middle of the guide (see inset) where both wires have
the same voltage. By adding a homogeneous electric potential relative to ground,
the zero electric field point may be moved at will to achieve any acceleration or
deceleration rate. A constant acceleration is obtained when the zero electric field
point is maintained at a constant distance from the position of the atoms.

D. MINIATURIZATION AND TECHNOLOGICAL CONSIDERATIONS

To achieve very robust and highly controlled atom manipulation one would like
to localize atoms in steep traps or guides which can be fabricated with high
precision. The large technological advances in precise nanofabrication, with the
290 R. Folman et al. [II

achievable size limit on chips smaller than 100 nm, makes the adaptation of these
processes for mounting the wires onto surfaces very attractive.

D. 1. Miniaturization

The main motivations behind miniaturization and surface fabrication are:


9 Large trap level spacings help to suppress heating rates. To achieve the
necessary large trapping gradients and curvatures with reasonable power
consumption, miniaturization is unavoidable (see Sect. II.D.4).
9 The tailoring resolution of the potentials used for atom manipulation is
given by the resolution of the fabrication of the structures used. It is,
for example, important for the realization of atom-atom entanglement by
controlled collisions as suggested by Calarco et al. (2000) (see Sect. VI)
to reduce the distances between individual trapping sites to the micron
regime or below. This would be virtually impossible with (large) free-standing
structures.
9 Nanofabrication is a mature field which allows one to place wires on a surface
with great accuracy (< 100 nm). Surface-mounted structures are very robust
and the substrate serves as a heat sink allowing larger current densities (see
Sect. II.D.4). In addition, nanofabrication allows parallelism in production of
manipulating elements (scalability).
9 Nanofabrication also allows us to contemplate the integration of other
techniques on the chip (see Sect. VI for details).

D.2. Finite size effects

The formulae presented in Sects. II.A to II.C are exact only for infinitely small
wire cross sections. In the case of a physical wire with a finite cross section,
they are a good approximation only as long as the height above the wire is
greater than the width of the wire. For experiments requiring a trap height smaller
than the width of the wire, finite size effects have to be taken into account. In
Fig. 17, we present examples of calculations showing how the trap gradient is
limited by finite size wires. One clearly sees that at trap heights of the order of
the width of the wire the resulting gradient starts to deviate from the expected
value. The effect is small for wires with a square cross section, while it becomes
considerably more important when rectangular wires with high ratios of width
to thickness are used.

D.3. Van der Waals interaction

The Van der Waals interaction becomes important at distances of the order
of a few 100nm from the surface. The interaction can be strong enough to
II] M I C R O S C O P I C ATOM OPTICS: F R O M W I R E S TO ATOM CHIP 291

0 30 , lar

025

0 20

-- 015

82
0 05

, i , i

1 2 3 4

Distance from wire surface [in units of d]

FIG. l 7. Deviations from the field of an infinitely thin wire become important as the surface of
a physical wire is approached. The plot shows the trap gradient for a side guide (see Sect. II.A.2)
when differently shaped wires are used. The solid line corresponds to a circular cross section as a
reference since the field outside the wire equals that of an infinitely thin wire at the wire center.
A wire with a square cross section (dotted line) shows very small deviations, while broad and thin
wires (dashed lines) deviate more and more as the thickness/width ratio decreases. Here, all wires
are chosen to have the same cross section d 2. Therefore, the widths of the rectangular wires are
2d and x/]--dd = 3.2d for the ratios 1:4 and 1:10, respectively.

significantly alter the trapping potentials (Grimm et al., 2000). Traps much
closer than 100 nm from the surface will be very hard to achieve since the Van
der Waals potential attracts the atoms to the surface and increases with 1/d 3
(in the non-retarded regime where the distance d is smaller than the optical
wavelength).

D. 4. Current densities

A limiting factor in creating steep traps and guides is the maximally tolerable
current density of a current-carrying structure. Considering a side guide potential
created by a wire with finite width d and a constant thickness, the highest
possible gradient is achieved at a distance from the wire comparable with d. The
bias field needed for such a trap is given by the ratio of the m a x i m u m current
that can be pushed through the wire and d; therefore the bias field is proportional
to the m a x i m u m current density j. This leads to the conclusion that the highest
possible gradient is given by j / d which favors smaller wires. If a square wire
cross section d 2 is assumed, the m a x i m u m gradient is proportional to j . Even in
this case, smaller d will allow for larger gradients b e c a u s e j has been observed to
increase with smaller wire cross sections. The drive for smaller width is stopped
292 R. Folman et al. [III

at a distance of about 100 nm where surface decoherence effects (see Sect. V)


and Van der Waals forces may be too strong to endure.

D.5. Multi-layer chips


Last, one should also note that as more complex operations are demanded from
the atom chip (see Sect. VI), it will have to move on from a 2-dimensional
structure into a 3-dimensional structure in which not only current- and charge-
carrying wires are embedded, but also light elements and wave guides. These
highly complex devices will force upon the fabrication a whole range of material
and geometrical constraints.

III. Experiments with Free-Standing Structures

The basic principles of microscopic atom optics have been demonstrated using
free-standing structures: current-carrying and charged wires. The interaction
potentials are in general shallow, typically only a few mK deep. Hence
experiments use cold atoms from a MOT or a well collimated atom beam
(even the moderate collimation of 1 mm over 1 m results in a typical transverse
temperature of <1 mK).
Free-standing wire structures can be installed close to a standard six beam
MOT without significantly disturbing its operation (as long as the wire is thin
enough), and offer large optical access which has advantages when probing the
dynamics of the atoms and their spatial distribution within the wire potentials.
They have the disadvantages that they are not very sturdy, they deform easily
due to external forces, and they cannot be cooled efficiently to dissipate energy
from ohmic heating. This limits the achievable confinement and the potential
complexity of wire networks. Nevertheless there are some special potentials
which can only be realized with free-standing wires.

A. MAGNETIC INTERACTION

As discussed in Sect. II.A there are two possibilities for magnetically trapping a
particle with a magnetic moment: traps for strong field seekers and traps for
weak field seekers. In the following we describe experiments with magnetic
microtraps which are based on small, free-standing wires or other magnetic
structures. Typical wire sizes range from 10 ~tm to a few mm and the wires
carry electrical currents of up to 20 A. All experiments but the first example
start with a conventional MOT of alkali atoms (lithium or rubidium) which is
initially situated a few mm away from the magnetic field producing structures.
This distance prevents the atoms in the MOT from coming into contact with the
III] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 293

structure surface where they would be adsorbed. It also provides the necessary
optical access for the MOT laser beams.
To load the magnetic wire traps and guides, the MOT laser light is simply
switched off and the magnetic trap fields are turned on. The loading rate into the
miniature magnetic traps has been enhanced in some experiments (a) by optically
pumping the unpolarized MOT atoms to the right trapping state (Key et al.,
2000); (b) by first loading the MOT atoms in a size-matched magnetic trap which
is then further adiabatically compressed (Vuletic et al., 1996, 1998; Key et al.,
2000; Fortagh et al., 1998; Haase et al., 2001); (c) by moving the MOT closer
to the trapping region shortly before the light is turned off, which can be done
with an additional magnetic bias field (Denschlag et al., 1999b). In this way the
efficiency of transferring the atoms into the miniature magnetic traps reached
between 1 and 40 %. In general the spatial distribution of the trapped atoms was
imaged with a CCD camera by shining a resonant laser beam onto the atoms and
detecting its absorption or the atomic fluorescence.

A. 1. Magnetic strong-field-seeking traps." the Kepler guide


A magnetic strong-field-seeker trap for cold neutral atoms was demonstrated
in two experiments: in 1991 by guiding an effusive beam of thermal sodium
atoms (mean v e l o c i t y - 6 0 0 m / s ) along a 1-m long current-carrying wire
(Schmiedmayer, 1992, 1995a,b) and in 1998 with cold lithium atoms loaded from
a MOT (Denschlag et al., 1999b).
The setup of the beam experiment is given in Fig. 18. The atom beam is
emitted from a 1-mm diameter nozzle in a 100~ oven and is collimated to
1 mrad. Introducing a small bend in the wire (-1 mrad), one can guide some of
the atoms along the wire around the beam stop. The atomic flux was measured
with a hot wire detector. The guiding wire was 150 ~tm thick and carried 2 A of
electrical current.
In the second experiment, lithium atoms were cooled in a MOT (1.6mm
diameter FWHM) to about 200 ~tK (which corresponds to a velocity of about
0.5 m/s). By shifting the MOT onto a 50 ~tm thick wire and releasing the atoms
from the MOT, about 10% of the unpolarized atomic gas could be trapped
magnetically in orbits of about 1 mm diameter around the wire that carried about
1 A of current. Monte Carlo calculations indicate that by optically pumping the
atoms and optimizing the trap size and current through the wire, it should be
possible to guide over 40% of the atoms from a thermal cloud with the Kepler
guide. The loading efficiency is limited to this amount, because atoms in highly
eccentric orbits hit the wire and are lost.
The bound atoms are guided along the wire corresponding to their initial
velocity component in this direction. Consequently, a cylindrical atomic cloud
forms that expands along the wire. After 40 ms of guiding, the atoms typically
had propagated over a 2cm distance along the wire (see left-hand panel of
294 R. Folman et al. [III

FIG. 18. (a) Experimental setup: The schematics at the bottom show in detail the relative geometric
arrangement between the apertures, the movable beam shutter used to bend the wire, and how the
wire is mounted. (b) Guiding of Na atoms along the 1-m long, 150-~tm diameter tungsten wire
(at detector position 0 indicated by the vertical line). Experimental count rates, n(I)- n(O) (left),
and Monte Carlo simulations (right), are shown for 0.0, 0.50 and 1.00mrad bends in the wire.
The different symbols represent currents of 0.5 A (circles), 1.0 A (diamonds), 1.5 A (crosses) and
2.0 A (triangles) through the wire. The thick line shows the fraction of atoms of the direct beam
that gets to the detector when no current is on (right-hand vertical axis). Its form corresponds to
the shadow of the bender that is cast onto the detector.

Fig. 19). For long guiding times the bound atoms leave the field of view, and the
fluorescence signal of the atoms decreases. The top left view images of Fig. 19
show a round atom cloud that is centered on the wire suggesting that atoms circle
around it.
By studying the ballistic expansion of the bound atoms after switching off
the guiding potentials, the m o m e n t u m distribution of the guided atoms can be
extracted. The center panel of Fig. 19 shows a picture sequence demonstrating
how the atomic cloud expands as a function of time. Starting from a well-
localized cylindrical cloud of guided atoms at t = 0 the spatial atomic distribution
transforms into a doughnut-like shape. This shows that there are no zero-velocity
atoms in the Kepler guide. In order to be trapped in stable orbits around the wire
the atoms need sufficient angular m o m e n t u m and therefore velocity. Atoms with
too little angular m o m e n t u m hit the wire and are lost.
Guiding in the Kepler guide is very sensitive to the presence of uncompensated
bias fields. Such additional magnetic bias fields, even if homogeneous, destroy
the rotational symmetry of the Kepler potential and angular m o m e n t u m is
not conserved anymore. Over the course of time, the Kepler orbits become
increasingly eccentric and thus finally hit the current-carrying wire leading to
loss, which was confirmed by Monte Carlo calculations. The right-hand panel
of Fig. 19 shows the results of an experiment investigating the dependence of the
III] M I C R O S C O P I C ATOM OPTICS: FROM WIRES TO ATOM CHIP 295

FIG. 19. Left: guiding of atoms along a current-carrying wire in their strong field seeking state
(Kepler guide). Pictures of the atomic clouds are shown, taken in axial and transverse directions with
respect to the wire. For times shorter than 15 ms the expanding cloud of untrapped atoms is also
visible. The location of the wire is indicated by a line (dot). The pictures show a 2-cm long section
of the wire that is illuminated by the laser beams. Center: Atomic distribution after free expansion
of 0 to 9 ms for atoms that have been guided in Kepler orbits around the wire. The expanded cloud
is doughnut-shaped due to the orbital motion of the atoms around the wire. Right: Experimentally
measured stability of the Kepler guide as a function of the magnitude of bias fields. The signal is
proportional to the number of atoms trapped in the guide after an interaction time of 20 ms.

magnetic trap stability on the magnetic bias field. The remaining atom number in
the Kepler guide was measured after 20 ms interaction time. It clearly decreases
with increasing bias field strength: the larger the bias field, the faster the atoms
get lost (Denschlag, 1998). In case of a weak disturbance the orbits can be
stabilized by an additional 1/r 2 potential which leads to a precession of the
orbits.

A.2. Magnetic weak-field-seeking traps and guides


The development of miniature weak-field-seeker traps, as discussed in Sects.
II.A.2 and II.A.3, lays the foundations of miniaturized atom optics on chips. Here
and in the following sections we restrict our discussion explicitly to experiments
with free-standing structures. Surface-mounted guides and traps are discussed in
Sect. IV.
In the following experiments the circular symmetric magnetic field of a
straight current-carrying wire is combined with a magnetic bias field as described
in Sect. II.A.2. The two fields cancel each other along a line that is parallel to the
wire creating a magnetic field minimum (side guide). In the simplest case, the
bias field can be created by an additional wire (Fig. 20a) (Fortagh et al., 1998) or
by an homogeneous external field (Fig. 20b) (Denschlag, 1998; Denschlag et al.,
1999b). Four wires also create a 2-dimensional quadrupole field (Fig. 20c) (Key
et al., 2000).
The experiments of the group of C. Z i m m e r m a n n (Fortagh et al., 1998) used
additional endcap ('pinch') coils (see Fig. 20a) to confine the atoms also in
the direction along the wire. They succeeded in adiabatically transferring and
296 R. F o l m a n et al. [III

FIG. 20. Three realizations of magnetic quadrupole traps with straight wires. (a) Trap realized
by Fortagh et al. (1998) with a thin wire (50~tm) glued onto a thick wire (1 mm). The current
through both wires flows in opposite directions. (b) A homogeneous bias field is combined with
a single straight wire (Denschlag, 1998; Denschlag et al., 1999b). (c) Four wires with alternating
current direction produce a quadrupole field minimum in the center. In the experiment the four wires
were embedded in a silica fiber (Key et al., 2000). (d) Images of atoms in guide (b).

compressing the magnetic t r a p - reaching a relatively high transfer efficiency


of 14% from the MOT into a microtrap without losing phase space density. In
experiments in Innsbruck (Denschlag, 1998; Denschlag et al., 1999b) and Sussex
(Key et al., 2000) (Figs. 20b and c, respectively) cold atoms released from a MOT
were guided along the wires at a distance of one to two centimeters (Fig. 20d). In
addition, the vertical Sussex experiment used one bottom pinch coil to confine
the falling atoms from exiting the guide. The atoms bounced back and were
imaged at the top exit.
By choosing appropriate bias field strengths and wire currents, a wide range
of traps with different gradients have been realized, and the scaling properties
(see Sect. II.A.4) were studied. With a fixed trap depth (given by the magnitude
of the bias field Bb) the trap size and its distance from the wire can be controlled
by the current in the wire. The trap gets smaller and steeper (gradient ~ B 2 / I )
for decreasing the current in the wire, which was confirmed experimentally
(Denschlag, 1998; Denschlag et al., 1999b). For example, a trap with a gradient
of 1000 G/cm can be achieved with a moderate current of 0.5 A and an offset
field of 10 G. The trap is then be located 100 ~tm away from the wire center.
A different weak-field-seeker trap has been experimentally realized by placing
a current-carrying wire right through the minimum of a magnetic quadrupole
field (Denschlag, 1998; Denschlag et al., 1999a). If the wire is aligned along the
direction of the symmetry axis of the quadrupole field, a ring-shaped potential
is obtained with a non-vanishing minimum field strength (see Sect. II.A.4 and
Fig. 5a).
III] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 297

A.3. Beam splitters

Although free-standing wire experiments are certainly limited in their architec-


tural complexity because of mechanical stability, some variations of the straight
wire geometry have been explored. By combining two free-standing wires one
can form a "Y" or fork, which can be used as an atomic beam splitter (see
Fig. 21) (Cassettari et al., 1998; Denschlag et al., 1999a). Choosing an arm of
the fork through which an electrical current is conducted, the atomic flow can
be switched from one arm to the other. If current is sent through both arms, the
atom beam is split in two.

FIG. 21. Atomic beam switch for guided atoms using a "Y"-shaped current-carrying wire. By
controlling the current through the arms, one can send cold lithium atoms along either arm or split
the beam in two. The images here show the switch operated in the Kepler guide mode and the
"weak-field-seeker" mode.

A.4. Free-standing bent wire traps

Experiments with free-standing wires that are bent in shape of a "U" or "Z" have
been reported by Denschlag et al. (1999a), Haase (2000) and Haase et al. (2001).
Bending the wire has the effect of putting potential endcaps on the wire guide,
which turns it into a 3-dimensional weak-field-seeker trap (see Sect. II.A.4.2).
A simple Z-wire trap achieves trapping parameters similar to the ones currently
used in BEC production, here, however, with moderate currents and very low
power consumption (see Sect. IV.C.4). In their experiment, Haase et al. used
a 1-mm thick copper wire, with the central bar being about 6 mm long. The
wire can carry 25 A without any sign of heating. Figure 22c shows the scaling
298 R. Folman et al. [III

FIG. 22. (a) Schematic description of the experiment. Camera 1 is looking along the central bar
of the magnetic trap and camera 2 along the leads. In addition to the two laser beams shown in the
figure, there is the third MOT beam parallel to the central bar. (b) The Z-wire held by two Macor
blocks is mounted on a flange. (c) The cloud of trapped atoms monitored by camera 1. By changing
the bias field Bb from 5 to 52 G, the trap size and position change. Also, the trap frequency increases
from 30 to 1600Hz. The experiment confirms the predicted scaling laws concerning trap distance,
frequency and bias field.

properties of the Z-trap. The atomic cloud can be compressed by raising the bias
field or by lowering the wire current.

A.5. The tip trap

Vuletic et al. (1996, 1998) have demonstrated a miniature magnetic quadrupole


trap (the tip trap) by mounting small coils on a combination of permanent
magnets and ferromagnetic pole pieces (see Fig. 23). In this way they exploited
the fact that for a given magnetic field Bo the maximum possible field gradient
scales like Bo/R where R is the geometric size of the smallest relevant element.
The central element of the tip trap is a 0 . 6 5 m m steel pin of which one
tip is sharpened to a radius of curvature of 10gm. Thus with R = 1 0 g m
and Bo -- 1000G the magnetic field gradient exceeded 105 G/cm. Working
with lithium atoms, this gradient implies a ground-state size of the atomic
wavefunction smaller than the wavelength of the optical transition at 671 nm.
The microtrap is loaded by adiabatic transport and compression: The atoms of
the lithium MOT are transferred to a volume-matched, but still relatively shallow
magnetic potential after turning off the MOT light. By adiabatically changing the
currents through the miniature coils the magnetic trap compresses its size by a
III] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 299

first electromagnet second electromagnet 4.1 0 5 ........


(tip coil) (counter coil) ~ b) a) . 9

J .................
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .J. . . / / / x ~ 8 ~ .... J 3"105
2.105
..~ ~p. 21 mm mm

] lCm 1.101t
I ~1" I=- permanentmagnet
J with steelpin -2 ' (~ " :~ ' ~, " 6 ' 8 " 1()' 1'2' 1 ' 4 ' 1 6
distance of light sheet from pin tip [ram]

FIG. 23. Left: Setup of the tip trap ofVuletic e t a l . . A sharp steel pin is magnetized by a permanent
magnet and exposed to a variable magnetic field that is generated by two electromagnets. Right:
Observed shape of the atomic cloud (a) in the shallow field after loading from a magneto-optical
trap and (b) after compression in the steep potential of the tip trap at a current in the tip coil of
1.2 A. Courtesy V. Vuletic.

factor of 6.5 within 100 ms. A total of 3% of the MOT atoms could be transferred
to the microtrap at moderate currents of 3 A through the tip trap coils.

A.6. Scattering experiments with a current-carrying wire


In 1995 the Melbourne group (Rowlands et al., 1995, 1996a,b) performed an
experiment where a beam of laser cooled cesium atoms, after being released
from a MOT, is scattered off a current-carrying wire. As the atoms pass through
the static inhomogeneous magnetic field of the wire they are deflected by a
force 27(/tB) dependent on the magnetic substate of the atom (see Fig. 24).

30 i , ,-, , 9 , , , 9 , , , 9 , 9 , 9 ,, , , ' |

20 ~
..l'..
trap 4
v ent into page)
~= o

~5 - l o

~ 20

30 ,,m mF - ". .

-40
-8 -6 .4 -2 0 2 4 s 8 10 12 14
horizontal distance (mm)

FIG. 24. Computer simulation of trajectories of cesium atoms deflected by the magnetic field from
a wire carrying 20 A. The solid lines indicate the trajectories for atoms in the nine possible magnetic
substates, assuming zero initial velocity. The broken lines are for atoms in the m F = + F substates
with initial transverse velocities of-t-1 cm/s. Courtesy P. Hannaford.
300 R. Folman et al. [III

With currents of up to 45 A through the wire, the positions of the atoms in the
individual magnetic substates are resolved and deflection angles as large as 25 ~
are observed. State preparation of the atoms using optical pumping increases the
number of atoms deflected through essentially the same angle.

A. 7. A storage ring f o r neutral atoms

Very recently Sauer et al. (2001) have demonstrated a storage ring for neutral
atoms using a two-wire guide (Sect. II.A.3). A pair of wires (separation-840 gm)
which forms a ring of 2cm diameter, produces a 2-dimensional quadrupole
magnetic field (see Fig. 25). The wires carry currents of 8 A in the same direction
which produces a field minimum between the two wires with a field gradient
of 1800 G/cm and a trap depth of 2.5 mK for the F = 1, mF = -1 ground
state of 87Rb (weak-field seeker). The ring is loaded from a MOT via a similar
second two-wire waveguide. The MOT is turned off and the second waveguide is
ramped up in 5 ms. Approximately 106 laser-cooled rubidium atoms (longitudinal
temperature 3 ~tK) fall 4 cm under gravity along the guide after which they enter
the storage ring with a velocity of about 1 m/s. To transfer the atoms to the ring,
the current in the guide is ramped off while simultaneously increasing the current
in the ring. Using fluorescence imaging the position and the number of the atom
cloud can be probed. Up to seven revolutions of the atoms in the ring have been
observed.

FIG. 25. (a) Schematic of the storage ring. (b) Cross section of the overlap region. The trap
minimum is shifted from between the guide wires to the ring wires by adjusting the current.
(c) Contour plot of a two-wire potential. The contours are drawn every 0.5 mK for the wire distance
d = 0.84mm and I = 8 A. (d) Successive revolutions in the storage ring. The points represent
experimental data, the curve is a theoretical model. Courtesy M. Chapman.

B. CHARGED WIRE EXPERIMENTS

Two types of experiments have used the 1/r 2 potential (Eq. 12) of a charged
III] M I C R O S C O P I C ATOM OPTICS: FROM WIRES TO ATOM CHIP 301

wire. One investigated the effect of a charged wire in atom interferometry. The
other investigated atomic motion in the singularity of the 1/r 2 potential. Here,
laser cooled atoms fall into the attractive singularity and are lost as they hit the
charged wire.

B.1. A charged wire a n d interferometry

Shimizu et al. (1992) used a straight charged wire to shift (deflect) the
interference patterns of a matter wave interferometer in a Young's double slit
configuration. In a recent experiment of the same group (Fujita et al., 2000)
(Fig. 26, at right), this work is expanded by combining a binary matter wave

FIG. 26. Experimental set-ups and data for interferometry and holography experiments with
charged wires. Left: charged-wire interferometerfor metastable helium. Different voltages applied to
the electrodes: the data sets are plotted with a vertical offset. The dotted horizontal lines indicate the
zero level for the respective measurements. Courtesy J. Mlynek. Right: Selective atom holography:
switching between atomic images "r and "~". For the upper figure the wire array is uncharged,
whereas for the lower figure it is electrically charged. The squares in the lower part of each figure
are nondiffracted atom patterns. Courtesy E Shimizu.

hologram with an array of straight charged wires. By changing the electric


potential applied to the electrodes on the hologram the holographic image
patterns can be shifted or erased, and it is even possible to switch between
two arbitrary holographic image patterns 9. These experiments were performed
using laser-cooled metastable neon in the l s3 state. After releasing them from
the MOT, the atoms fell under gravity onto a double slit or a binary hologram.
A few centimeters further down the atoms formed an interference pattern which
was detected by a multi-channel plate (MCP).
The binary hologram pattern held an array of 513 regularly spaced parallel wires
of platinum on its surface. Each electrode was either grounded or connected to

9 Similarly it was suggested by Ekstrom et al. (1992) that charged patches on a grating can be used
to modify the diffraction properties.
302 R. Folman et al. [III

FIG. 27. (a) Two classical trajectories: An atom falls into the 1/r2 singularity of an electrically
charged wire if the atomic angular momentum Lz < Lcrit. If Lz > Lcrit it scatters and escapes from
the singularity. (b) When moved onto the wire the atom trap decays exponentially, as can be seen
by monitoring the atomic fluorescence signal. Charging the wire (100 V ~ 6.4 pC/cm) creates an
attractive 1/r2 potential and enhances the decay rate. Inset: experimental steps. Loading of the trap,
shifting it onto the wire and observing its decay. (c) Dependence of the trap decay rates on the wire
charge for different wire thicknesses. The decay rate for uncharged wires is proportional to their
actual diameters. For increasing charges the absorption rate becomes a linear function of the charge,
a characteristic of the 1/r2 singularity. The slope is independent of the wire diameter.

a terminal. The width and the spacing of each wire was 0.5 ~tm and the holes for
the binary hologram in between the wires were 0.5 ~tm • 0.5 ~tm in size. The
electric field E generated between two wires shifted the energy of the neon atom
b y - o t E 2 / 2 . When two adjacent electrodes had the same potential, the atoms in
the gap were unaffected. If they had different potentials, the atoms accumulated
an additional phase while passing through the hole.
In an experiment in Konstanz, Nowak et al. (1998) sent a collimated thermal
beam of metastable helium atoms onto a charged wire (tungsten, 4 ~tm diameter)
where it was diffracted (see Fig. 26, left). 1.3 m further downstream they
observed an interferometric fringe pattern which depended on the wire charge
and on the de Broglie wavelength. The data agreed well with the theoretical
predictions for scattering polarizable particles off a 1/r 2 potential.

B.2. A charged wire in gas o f cold atoms: studying a singular p o t e n t i a l

The motion in a 1/r 2 singularity can be studied by placing a cloud of cold atoms
in the potential of a charged wire. In this experiment the n u m b e r of cold lithium
atoms of a M O T is monitored while the atoms move in the 1/r 2 potential of the
wire (Denschlag et al., 1998). At extremely low light levels the M O T acts as a
box holding a gas o f atoms. Atoms falling into the attractive 1/r 2 singularity are
lost as they hit the wire. This loss m e c h a n i s m leads to an exponential decay of
the trapped atom n u m b e r (see Fig. 27b).
The corresponding loss rate is characteristic for the 1/r 2 singularity and its
IV] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 303

strength. Atoms with angular momentum Lz < Lcrit (see Eq. (14) in Sect. II.B)
fall into the singularity. The loss rate is a linear function of q because Lcrit is
proportional to the line charge q and the atoms are uniformly distributed over
angular momentum states (see Fig. 27c). This is actually true only for high
charges, since for lower q, the finite thickness of the charged wire becomes
apparent. The MOT decay rate for an uncharged wire is proportional to its actual
diameter. The radii of the wires in the experiments ranged between 0.7 ~tm and
5 ~tm. A detailed analysis of the absorption data reveals that Van der Waals forces
also contribute to the atomic absorption rate (Denschlag, 1998). This effect was
found to be important for thin wires with diameters of less than 1 ~tm. Hence
this system should allow for detailed future studies of Van der Waals interaction
and retardation in nontrivial boundary conditions.
The 1/r 2 potential would be especially interesting to study in the quantum
regime where the de Broglie wavelength of the atoms is much larger than the di-
ameter of the charged wire; the quantization of angular momentum then begins to
play a role (Denschlag and Schmiedmayer, 1997). This can be used for example
in order to build an angular momentum filter for atoms (Schmiedmayer, 1995a).

IV. Surface-Mounted Structures: The Atom Chip

Free-standing structures, as those described in the previous section, are extremely


delicate, and one arrives quickly at their structural limit, when miniaturizing
traps and guides. Wires mounted on a surface are more robust, can be made
much smaller, and heat is dissipated more easily which allows significantly
more current density to be sent through the wires. This together with strong
bias fields allows for tighter confinement of atoms in the traps. Consequently,
ground-state sizes <10nm become feasible. Existing accurate nanofabrication
technology provides rich and well established production procedures, not only
for conducting structures, but also for micromagnets. Optical elements such as
micro-optics, photonic crystals and microcavities can also be included to arrive
at a highly integrated device. The small ground-state size of such microtraps
implies that we know the exact location of the atom relative to other structures
on the surface to the precision of the fabrication process (typically <100 nm),
allowing extremely close sites to be addressed individually for manipulation and
measurement.
We have named nanofabricated surfaces for cold atom manipulation 'atom
chips' in reminder of the similarity of these atom-optical circuits to electronic
integrated circuits (Folman et al., 2000). In designing atom chips one attempts to
bring together the best of two worlds: the well-developed techniques of quantum
manipulation of atoms, and the mature world of nanofabrication in electronics
and optics, to build complex experiments utilizing the above techniques.
304 R. Folman et al. [IV

In the following we describe the atom chip and its present experimental status.
Future goals will be addressed in Sect. VI.

A. FABRICATION

There are many different techniques of atom manipulation which can be


integrated into an atom chip. Present atom chip experiments follow a simple
scheme based on wires that carry currents or charges. These allow to miniaturize
the free-standing devices discussed in Sect. III. We will focus here on these
simple integrated structures, leaving issues of further integration to the outlook
in Sect. VI.
To build an atom chip one has to solve the following problems: first of all, the
microstructures have to withstand high current densities and high electric fields.
This requires structures with low electric resistance. The material of choice for
the wires is gold, though other materials such as copper are also used. For
the substrate one wants good heat conductivity with high electric insulation
withstanding large electric fields (created at sharp (r ~ l gm) corners even by
small voltages), and ease of fabrication. Typical materials are silicon, gallium
arsenide, aluminum nitride, aluminum oxide and sapphire (A1203), though glass
has also been used.
Another requirement lies in the fact that cold atoms have to be collected
and then transferred towards the small traps on the chip. If one wants to avoid
transferring the atoms from a distant MOT, the chip has to be either transparent
or reflecting, to allow lasers to address the atoms from all directions near the
surface. Nevertheless, experiments exist in which the atoms have been brought
from a distance to a chip (Ott et al., 2001; Gustavson et al., 2002).
Presently, atom chips are built mainly using two technologies: thin film hybrid
technology, or plain nanofabrication which is the first step of the two-stage hybrid
technology.

A.1. Thin film hybrid technology


In this approach one starts from an insulating substrate (e.g. sapphire) and
patterns, using lithographic techniques, a layout of the desired structure onto
a thin (<100nm) metallic layer. In the second stage, the wires are grown by
electroplating: Metal ions from a solution are deposited onto the exposed metallic
layer, which is now charged. With this process one obtains wires with quite
large cross sections (typical structure widths are 3 to 100~tm) that support
high currents. However, miniaturization will be limited to a few micron wire
width. Furthermore, surface roughness is quite large, which makes such surfaces
less suitable for the reflection MOT and atom detection. These drawbacks and
the expected shadows from large etchings between wires, have been dealt with
successfully by covering the chip with an insulating layer and then with a metallic
IV] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 305

FIG. 28. Electroplating. Left: cross section of the Munich-group chip. The metallic layer on top of
the wires gives the chip enhanced surface quality in order to form a mirror MOT. Right: the layout
of the chip. The magnetic 'conveyor belt' explained in Sect. II is visible. The wires are connected
to the chip pads from the outside by means of wire bonding. Recently, this chip was used to achieve
Bose-Einstein condensation. Courtesy J. Reichel.

reflection layer (e.g. the Munich chip as shown in Fig. 28) (Reichel et al., 2001).
This, however, carries the price of not enabling atoms to be closer than some
20 gm from the wires themselves. A technical advantage of electroplating is that
it wastes less gold or copper because one avoids evaporation of large amounts
of metal, which mostly cover the evaporation chamber and not the chip.
Atom chips fabricated using using this technique have been used sucessfully
by the groups at Harvard (M. Prentiss), Munich (J. Reichel and T.W. HS.nsch),
JILA, Boulder (D. Anderson and E. Cornell) and Ttibingen (C. Zimmermann).

A.2. Nanofabrication
Atom chip structures can also be fabricated into an evaporated conductive layer
with state-of-the-art processes used for electronic chips. To the best of our
knowledge, this approach is only used by the Heidelberg (formerly Innsbruck)
group. In these atom chips a 1-5 gm gold layer is evaporated onto a 0.6-mm
thick semiconductor substrate (GaAs or Si). As GaAs or Si tend to leak currents,
especially in the presence of light, a thin isolating layer of SiO2 is put between
the substrate and the gold layer. The chip wires are defined by 2-10 gm wide
grooves from which the conductive gold has been removed. This leaves the chip
as a gold mirror that can be used to reflect MOT laser beams (the 10-gm grooves
impede the MOT operation only in a very slight way). The mirror surface quality
is very high, achieving an extremely low amount of scattered light. The chips
were produced at the microfabrication centers of the Technische UniversitS, t Wien
and of the Weizmann Institute of Science, Rehovot, see Fig. 29.
Atom chips fabricated with this method have the advantage that the structure
size is only limited by nanofabrication (<100nm). The drawback is that the
conductive layer cannot be too thick. This is due mainly to restrictions on the
available thickness of the photoresist used in the process. The thin wires support
only smaller currents, and therefore only smaller traps closer to the surface can
306 R. Folman et al. [IV

FIG. 29. Nanofabricated atom chips (Heidelberg). Left: a mounted chip, ready to be put into the
vacuum chamber. The mechanical clamp contacts to the pads are visible. The mounting also includes
cooling in order to remove heat produced by the currents. Center (from top to bottom): details of
fabrication and assembly: (i) a chip in the middle of the fabrication process, after some gold has
been evaporated and before the photoresist has been removed. The visible wires have a cross section
of 1• 1 ~tm2; (ii) an electron-microscope view of the surface: a 'T' junction of a 10-~tm wide wire
is visible as well as the 10 ~tm etchings which define it; (iii) typical design of the U- and Z-shaped
wires placed underneath the chip to help in the initial loading process; the wires can support >50 A
of current in DC operation without degrading a p < 10-11 mbar vacuum. Right: a typical design of
an atom chip. On both sides contact pads are visible. The center of the chip is used for loading
the atoms, which are then released into the physics areas: on top, a magnetic guide with arrays
of electric leads, on the bottom, a spiral formed by two parallel wires enables atom guiding in all
directions on the chip.

be built. This d i s a d v a n t a g e can be c o r r e c t e d by a d d i n g larger wires b e l o w the


chip surface, as p r e s e n t e d in Fig. 29.
At this stage it is hard to j u d g e w h a t is the best fabrication process. T h e r e
are still m a n y o p e n questions. For e x a m p l e , is there a sizable difference in
the resistivity b e t w e e n e v a p o r a t e d gold and e l e c t r o p l a t e d g o l d ? For a direct
c o m p a r i s o n , one w o u l d have to unify all other p a r a m e t e r s such as substrates a n d
i n t e r m e d i a t e layers. A n o t h e r q u e s t i o n c o n c e r n s the final fabrication r e s o l u t i o n
one w i s h e s to realize. A s s u m i n g one aspires to achieve the s m a l l e s t possible
trap height above the chip surface for sake o f low p o w e r c o n s u m p t i o n and h i g h
potential tailoring resolution, the limit will be at a h e i g h t b e l o w w h i c h surface
i n d u c e d d e c o h e r e n c e b e c o m e s too strong (see Sect. V). This height, t o g e t h e r
with the finite size effects d e s c r i b e d in Sect. II, will d e t e r m i n e the fabrication
resolution needed. S m o o t h n e s s o f r e s o l u t i o n will also be r e q u i r e d as fluctuating
wire widths will c h a n g e the current density and therefore the trap f r e q u e n c y in a
way that m a y h i n d e r the transport o f B E C due to potential hills. Finally, as m u l t i -
layer chips u s i n g m o r e elaborate 3 - d i m e n s i o n a l designs are i n t r o d u c e d e.g. for
IV] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 307

wire crossings and more complicated structures including photonic elements,


it may be that conductor layers thicker than a few microns will have to be
abandoned. In order to fully exploit the potential of the atom chip in the future,
the technology used will have to be such that all elements could be made with
a suitable process into a monolithic device.
Finally, we note that although usual current densities used in the experiments
range between 106 and 107 A/cm 2 (higher with smaller cross sections and
depending on pulse time, work cycle and heat conductivity of substrate),
densities of up to 108 A/cm 2 have been reported for cooled substrates (Drndi6
et al., 1998). Gold wires have been found to be the best, achieving superior
performance even when compared to superconductors.

B. LOADING THE CHIP

In general there are two different approaches to loading cold atoms into the chip
traps:
(i) Collect and cool the atoms at a different location and transport the cold
ensemble to the surface traps. This may be achieved using direct injection from
a cold atomic beam coming from a low-velocity intense source (LVIS) (Mfiller
et al., 1999, 2000, 2001) or a released MOT whereby the atoms are pulled by
gravity (Dekker et al., 2000). Transferring the atoms with magnetic traps has
been achieved by (Ott et al., 2001), and a Bose-Einstein condensate (BEC) has
been loaded using optical tweezers (Gustavson et al., 2002).
(ii) Cool and trap atoms close to the surface in a surface MOT, and transfer
the atoms from there to the microtraps on the chip (Reichel et al., 1999; Folman
et al., 2000). For this method the atom chip has to be either transparent or
reflecting.
In the following, we describe experiments performed at Heidelberg (resp.
Innsbruck), Sussex and Munich using the second approach. Further on, several
experiments using the first approach will also be discussed (see Figs. 37 and 38).

B. 1. Mirror M O T

The first problem to solve is how to obtain a MOT configuration close to a


surface. This problem has an easy solution if we recall that a circularly polarized
light beam changes helicity upon reflection from a mirror. To the best of our
knowledge, the idea of a reflection MOT was first put into practice with a
pyramid of mirrors and one beam (Lee et al., 1996), as presented in Fig. 30a.
Almost in parallel, a single planar surface with four beams impinging at 45 ~
degrees onto the surface was used, thus realizing an eight beam MOT (Pfau
308 R. Folman et al. [IV

FI6. 30. (A) a 'pyramid MOT' is obtained when one single laser beam is retro-reflected by
a four-sided pyramid in the center of a magnetic quadrupole. The reflections ensure the correct
helicities of the laser beams when the quadrupole field (field lines) has the same symmetry as the
pyramid. (B) The mirror MOT is generated from the pyramid by leaving out 3 of the 4 reflecting
walls. Two MOT beams (I and II) impinge from opposite directions on the reflecting surface of
the atom chip. The correct MOT configuration is ensured together with the magnetic quadrupole
field rotated 45 ~ to the atom chip surface as illustrated by the field lines. The magnetic field can
be obtained either by a set of external quadrupole coils, or by a U-shaped wire on the chip. Top:
the Sussex mirror MOT chip setup with the external quadrupole coils on the mounting, inside the
vacuum; two parallel wires embedded in a fiber are positioned on the surface of the mirror, forming
a two-wire guide and a time-dependent interferometer (see Sect. II). Two small 'pinch' coils visible
at the edges of the mirror provide longitudinal confinement. Courtesy E. Hinds.

et al., 1997; S c h n e b l e et al., 1999; G a u c k et al., 1998; S c h n e b l e et al., 2 0 0 1 ) 1 0 .


The surface MOT most common today derives f r o m the p y r a m i d MOT. The
M O T b e a m c o n f i g u r a t i o n is g e n e r a t e d f r o m f o u r b e a m s b y r e f l e c t i n g o n l y t w o

10 In another version of this experiment, an evanescent field just above the extremely thin metal
surface, formed by light beams impinging on the back of the surface, was used as an atom mirror.
This allowed to produce a MOT with reasonable surface induced losses even at the extreme proximity
of 100nm from the surface.
IV] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 309

beams off the chip surface (see Fig. 30b) (Reichel et al., 1999; Folman et al.,
2000). The magnetic quadrupole field for the MOT can be obtained either by a
set of external quadrupole coils, or by superimposing a homogeneous bias field
with the field generated from a U-shaped wire on or below the chip ('U-MOT').
External quadrupole coils generate the correct magnetic field configuration if one
of the reflected light beam axes coincides with the coil axis. If the U-MOT is
used, the reflected light beams must lie in the symmetry plane of the U. Trapping
in the U-MOT has the advantage that the MOT is well aligned with respect to
the chip and its microtraps. If the mirror MOT is sufficiently far from the surface
(a few times the MOT radius), its loading rate and final atom number are very
similar to a regular free space MOT under the same conditions (laser power,
vacuum, supply of cold atoms, etc.). In agreement with earlier observations using
wires, the shadows (diffraction patterns) from the 10~tm etchings in the gold
surface of the nanofabricated atom chip do not disturb the MOT significantly
(Denschlag, 1998; Denschlag et al., 1999b).
Such atom chip mirror MOTs have been loaded from an atomic beam in Inns-
bruck/Heidelberg (Folman et al., 2000), from the background vapor in Munich
(Reichel et al., 1999), and in a double MOT system in Innsbruck/Heidelberg
(>108 atoms at lifetime <100 s), using either external coils or the U-wire for
the quadrupole field. In addition, at Sussex and Harvard surface MOTs were
realized using permanent and semi-permanent (magnetizable cores) magnetic
structures.
As an example we describe the Innsbruck/Heidelberg lithium setup. Figure 31
(overleaf) shows a top view of the mirror MOT just above the chip with
some of its electric connections. For the transfer into the U-MOT, the large
external quadrupole coils are switched off while the current in the U-shaped
wire underneath the chip is switched on (up to 25 A), together with an external
bias field (8 G). This forms a nearly identical, but spatially smaller quadrupole
field as compared to the fields of the large coils. By changing the bias field,
the U-MOT can be compressed and shifted close to the chip surface (typically
1-2 mm). The laser power and detuning are changed to further cool the atoms,
giving a sample with a temperature of about 200 ~tK.

B.2. Transferring atoms to the chip surface

After the U-MOT phase, atoms are cooled using optical molasses, optically
pumped and transferred into a matched magnetic trap, typically produced by
a thick Z-shaped wire plus bias field. From there atoms are transferred closer
and closer to the chip and loaded sequentially into smaller and smaller traps. In
general, it is favorable to lower the trap towards the surface by increasing the
magnetic bias field. This way the trap depth increases and less atoms are lost
due to adiabatic heating during compression. Unfortunately this is not feasible
310 R. Folman et al. [IV

FIG. 31. Loading of cold atoms close to the surface of an atom chip. Top left: Picture of the
mirror MOT, taken from above; the cloud is visible at the center while the electrical contacts can be
seen at the edges. Top right: schematic of the MOT beams and quadrupole coils. Center row: Atoms
trapped in the U-MOT created by a current in the large U-shaped wire underneath the chip and a
homogeneous bias field. Bottom row: Atoms in a magnetic trap generated by the U-wire field; from
left to right, the columns show the top, front and side (direction of bias field) views respectively, the
far right column shows the schematics of the wire configuration; current-carrying wires are marked
in black. The front and side views show two images: the upper is the actual atom cloud and the lower
is the reflection on the gold surface of the chip. The distance between both images is an indication
of the distance of the atoms from the chip surface. The pictures of the magnetically trapped atomic
cloud are obtained by fluorescence imaging using a short laser pulse (typically < 1 ms).

all the way: Finite size effects limit small traps to thin wires, at the price o f not
being able to p u s h high currents.
The basic transfer principle from a large wire to a small wire is to first
switch on the current for the s m a l l e r trap, and then to r a m p d o w n the c u r r e n t
in the b i g g e r trap m a i n t a i n e d by a thicker wire (Fig. 32). F u r t h e r c o m p r e s s i o n
is a c h i e v e d by using s m a l l e r and s m a l l e r currents. Care has to be taken that
the transfer is adiabatic, especially with r e s p e c t to the m o t i o n o f the p o t e n t i a l
m i n i m u m . B y an appropriate c h a n g e o f the bias field, the c o m p r e s s i o n o f the
a t o m s in the shrinking trap can be p e r f o r m e d v e r y smoothly. T r a n s f e r r i n g into
IV] M I C R O S C O P I C A T O M O P T I C S : F R O M W I R E S TO A T O M C H I P 311

Fl6. 32. Principle of compressing and loading wire guides. The position of the surface-mounted
wires and equipotential lines of the trapping potential are shown. Top row: the transfer from two
large 200 ~tm wires to one small 10 ~tm wire. In (a)-(c) the current in the small wire is constant
at 300 mA and the bias field is constant at 10 G. The current in the two large wires is decreased
from 2 A in each wire to zero. This transfers the atoms to the small wire. (d) By increasing the bias
field the trap can be compressed further. Bottom row: the transfer from one large 200 ~tm wire to
one small 10 ~tm wire. In (a)-(d) the current in the large wire drops from 2 A to zero. The thick
line shows how the trap center moves during transfer. A much weaker confinement during transfer
is obtained in this configuration.

m o r e c o m p l i c a t e d potential configurations one has to avoid the o p e n i n g o f escape


routes for the trapped atoms.
For an adiabatic transfer o f relatively hot atoms, the m a i n loss is due to
heating: w h e n c o m p r e s s i n g by lowering the current, high-lying levels m a y
eventually spill over the potential barrier. Significant loss occurs if the trap depth
is m u c h smaller than 10 times the t e m p e r a t u r e o f the atomic ensemble. Other
loss m e c h a n i s m s are described in Sect. V. For t h e r m a l clouds, typical achieved
transfer efficiencies from the M O T to the m a g n e t i c chip trap can be as high as
60%.
As a detailed e x a m p l e we describe the loading o f the first o f the I n n s b r u c k
experiments ( F o l m a n et al., 2000). After a c c u m u l a t i n g atoms in a m i r r o r M O T
and transferring t h e m to the U-MOT, the laser b e a m s are switched off and
the quadrupole field generated by the U - s h a p e d wire below the chip surface
312 R. Folman et al. [IV

FIG. 33. Compressing a cloud of cold atoms on an atom chip: Top row: view from the top;
center row: front view; bottom row: side view. The first three columns show atoms trapped on the
chip with the two U-shaped wires. The compression of the trap is accomplished by increasing the
bias field. The last row displays images from a Z-trap created by 300 mA current through the 10 gm
wire in the center of the chip. The pictures of the magnetically trapped atomic cloud are obtained
by fluorescence imaging using a short (<1 ms) molasses laser pulse.

serves as a magnetic trap for weak-field seekers (Fig. 3 l: U-trap). The magnetic
trap is lowered further towards the surface of the chip by increasing the bias
field. Atoms are now close enough to be trapped by the chip fields. Next, a
current of 2 A is sent through two 2 0 0 g m U-shaped wires on the chip, and
the current in the U-shaped wire located underneath the chip is ramped down
to zero. This procedure brings the atoms closer to the chip, compresses the
trap considerably, and transfers the atoms to a magnetic trap formed by the
currents on the chip surface. This trap is further compressed and lowered towards
the surface (typically <100~tm) by increasing the bias field (Fig. 33). From
there the atoms are transferred to a microtrap created by a 10 g m Z-shaped
wire.
In the lowest height and most compressed trap achieved to date, a 1 x 1 ~tm2
Z-shaped wire is used with a current of 100 mA (Heidelberg). With a bias field
of 30 G the atoms are trapped at a height of about 7 ~tm above the surface and
at an angular oscillation frequency co ~ 2Jr x 200 kHz (magnetic field gradients
of 50 kG/cm) for several tens of ms (see Eq. 6 for typical trap frequencies). At
such a small trap height several problems come into play: First, with the present
IV] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 313

2 mm distance between the bends of the Z-shaped wire, the Ioffe-Pritchard


configuration is nearly lost and one is left with a single-wire quadrupole field
where atoms can suffer Majorana flips. Two easy remedies would involve smaller
Z lengths or a slight tilt of the bias field direction. Second, the trap is so tight
that the number of atoms that survive the transfer and compression is small.
This limitation should not be applicable in the case of a BEC as there are no
high-lying states where atoms run the risk of spilling over the finite trap barrier.
A third problem has to do with the observation of the atoms: even with negligible
stray light from surface scattering or blurring by atomic motion, it is found that
direct observation of extremely tight traps close to the surface (<20 ~tm) is very
hard. The signal suppression is probably due to large Zeeman shifts in the cloud,
which together with optical pumping processes dramatically reduce the scattered
light. In such a case, one can observe the atoms after trapping by 'pulling' them
up, away from the surface into a less compressed trap. This may be done simply
by increasing the wire current or decreasing the bias field.

B.3. Observing atoms on the chip

A simple way to observe the trapped atoms is by fluorescence imaging. For this,
one illuminates the cloud with near-resonant molasses laser beams for a short
time (typically much less than l ms). The scattered light is imaged by CCD
cameras as shown in Fig. 31 and Fig. 33. One should use short enough exposure
times to avoid blurring of the image due to atomic motion. One also has to
select the camera positions wisely to avoid stray light caused by scattering off
the grooves in the atom chip surface. Furthermore, it is important that the metal
surface itself shows minimal light scattering. Here, the excellent surface quality
of evaporation on semiconductor surfaces is advantageous.
A different possibility is to use absorption imaging. If the probe beam is
directed parallel to the chip surface, the surface quality is not as critical, and
one does not have to take care of diffraction peaks from the grooves. Such
absorption imaging is used by the Munich, Tfibingen, Heidelberg, and MIT
groups. Profiting from an excellent surface mirror quality, the Heidelberg group
has also implemented absorption imaging with laser beams reflected from the
chip surface, which allows direct distance measurements. More sophisticated
methods such as phase contrast imaging will be important for more complicated
atom-optical devices on atom chips, where non-destructive observation very
close to the chip surface becomes essential. For an overview of these methods,
we refer the reader to the many BEC review papers (see, for example, Ketterle,
1999).
Finally, future light optical elements incorporated on the chip, such as
microspheres or cavities, will allow for much better detection sensitivity, possibly
at the single-atom level (see Sect. VI.A.3). Such work has been started in several
laboratories.
314 R. Folman et al. [IV

C. ATOM CHIP EXPERIMENTS

Since the first attempts two years ago, the atom chip has now become a 'tool
box in development' in numerous labs around the world. To the best of our
knowledge these include (in alphabetical order) the groups at Boulder/JILA
(D. Anderson and E. Cornell), CalTech (H. Mabuchi), Harvard (M. Prentiss),
Heidelberg (J. Schmiedmayer), MIT (W. Ketterle), Munich (J. Reichel and
T.W. H~insch), Orsay (C. Westbrook and A. Aspect), Sussex (E. Hinds), and
Tfibingen (C. Zimmermann). Unfortunately, we will not be able to present in
detail all the extensive work done, nor will we be able to touch upon other
surface-related projects, such as the atom mirror.

C.1. Traps

The simplest traps (i.e 3-dimensional confinement) are usually based on a straight
wire guide with some form of longitudinal confinement, which is produced either
by external coils or by wires on the chip (Sect. II.A.4). Additional wires for on-
board bias fields may also be added.
As an example, we start with the simple microtraps realized in Inns-
bruck/Heidelberg with lithium (Folman et al., 2000) and Munich with rubidium
atoms (Reichel et al., 1999; Reichel et al., 2001). Here, the traps are based on
wires of 1 to 30 ~tm width with which surface-trap distances below 10 ~tm were
achieved. The wires used are either U-, Z- or H-shaped.
In these experiments, the compression of traps and guides was also inves-
tigated (Folman et al., 2000; Reichel et al., 2001). This is done by ramping
up the bias magnetic field. In this process one typically achieves gradients of
>25 kG/cm. With lithium atoms, trap parameters with a transverse ground-state
size below 100nm and angular frequencies of 2:r x 200kHz were achieved
(Folman et al., 2000), thus reaching the parameter regime required by quantum
computation proposals (Calarco et al., 2000; Briegel et al., 2000).
In addition, an on-board bias field for the thin wire trap was also created
by sending currents through two U-shaped wires in the opposite direction
with respect to the thin wire current, substituting the external bias field
(see Sect. II.A.2). Hence, trapping of atoms on a self-contained chip was
demonstrated (Folman et al., 2000).
An example of a different configuration was realized in Munich with rubidium
atoms. Three-dimensional trapping was achieved by crossing two straight wires
and choosing an appropriate bias field direction, as discussed in Sect. II.A.4
(Reichel et al., 2001) (Fig. 34). Here the additional wire actually provides the
endcaps that were previously provided in the Z- and U-shaped traps by the same
wire. This type of trap will be useful for the realization of arrays of nearby
traps. In Tfibingen and Sussex longitudinal confinement has been achieved by
additional coils.
IV] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 315

FIG. 34. Ioffe-Pritchard trap created by two intersecting wires. The left-hand column corresponds
to I l > 12 and ]B0,y] > [B0,x[; in the right-hand column both relations are reversed. Top row:
conductor pattern; the thickness of the arrows corresponds to the magnitude of the current; dashed
arrows indicate the bias field direction. Middle row: calculated contours of the magnetic field
modulus ]B(x,y)] indicating how the long trap axis turns; the left potential continuously transforms
into the right one when the parameters are changed smoothly. Bottom row: absorption images
corresponding to the two situations. Courtesy J. Reichel.

Finally, the splitting o f a s i n g l e trap into t w o has b e e n d e m o n s t r a t e d in


H e i d e l b e r g , M u n i c h a n d Sussex. S u c h a t i m e - d e p e n d e n t p o t e n t i a l is p r e s e n t e d
in Fig. 35; as e x p l a i n e d in Sect. I|, it m a y f o r m the basis o f an i n t e r f e r o m e t e r .
It is also the first step in c r e a t i n g m u l t i - w e l l traps or a r r a y s o f traps.

FIG. 35. Top view of a thermal 200 ~tK cloud of lithium atoms in a double-well potential 40 ~tm
above the chip surface. The minima are separated by 350 ~tm. The imaging flash light pulse is 100 ~ts
long. The splitting may be done as slow as needed in order to achieve adiabaticity.
316 R. Folman et al. [IV

More sophisticated designs have been suggested by Weinstein and Libbrecht


(1995) (see Sect. II.A.5) and fabricated (e.g. Harvard: Drndid et al., 1998).

C.2. Guiding and transport

To achieve mesoscopic atom optics on a chip, it is essential to have reliable


means of transporting atoms. One such device is an atomic guide using a single
wire with a bias field. Such an experiment is shown in Fig. 36a. The Z-trap is
transformed into an L-shaped guide by re-routing the current from one of the
Z leads. The atoms expand along the guide due to their thermal velocity (Folman
et al., 2000). Similarly, it was demonstrated that one can directly load the guide
from a larger magnetic trap on the chip and skip the small surface trap.
It is also possible to achieve a continuously loaded magnetic guide using a
leaky microtrap (see Fig. 36b). This is achieved by lowering the barrier between
the trap and the guide, the barrier being simply the trap end cap, whose height
may be controlled by changing the current in the microtrap (Brugger et al.,
2000).
However, there are limitations to such a simple guide. Using a homogeneous
external bias field, the guide has to be straight (linear), since the bias field must
be perpendicular to the wire as discussed in Sect. II.A.2. This considerably limits
the potential use of the whole chip surface. A possible solution is to create the
bias field using on-chip wires (3-wire configuration shown in Fig. 2) or the
two-wire guide configuration discussed in Sect. II.A.3, in which the currents
are counter-propagating and the bias field is perpendicular to the chip surface.
A first experiment was conducted by M. Prentiss' group at Harvard (Dekker
et al., 2000). Here, cesium atoms were dropped from a MOT onto a vertically
positioned chip, on which a two-wire guide managed to deflect the atoms from
their free fall (see Fig. 37). Furthermore, a four-wire guide was realized whereby
the two extra wires served as the source for the bias field (see also Fig. 2).

Fie. 36. (a) Cold atoms in a microtrap (left) and released all at once into a linear guide (right).
(b) Continuous loading of an atom guide by leaking atoms from a reservoir created by a U-trap into
the guide by ramping down the current in the U. Propagation is due to thermal velocity.
IV] MICROSCOPIC A T O M O P T I C S : F R O M W I R E S TO A T O M C H I P 317

FIG. 37. Vertical bias field: this Harvard experiment realized two-wire vertical guides, enabling
the guiding of atoms in a variety of directions. Left: setup. Right: absorption of probe beam versus
position along x at 3.5 mm below the output of the guide. The left and right peaks are attributed to
unaffected atoms and atoms deflected by the outside of the guide potential, respectively. The open
triangles are the data collected while the guide is turned off. Courtesy M. Prentiss.

G u i d i n g a l o n g a c u r v e d two-wire guide has been achieved in H e i d e l b e r g , and


e x p e r i m e n t s to guide a t o m s along a spiral are in progress (for the chip design
see Fig. 29).
Several e x p e r i m e n t s have achieved g u i d i n g w i t h o u t any bias field by trapping
the weak-field seekers in the m i n i m u m existing exactly in b e t w e e n two parallel
wires with c o - p r o p a g a t i n g currents (see Sect. II.A.3). In Fig. 38, we p r e s e n t such

FIG. 38. The JILA setup in which a 'low-velocity intense source' (LVIS) was used to directly load
the two-wire guide. The data show the need for strong potentials with which the magnetic guide
can overcome the kinetic energy in order to deflect the atoms, thereby bypassing the beam block.
Courtesy E. Cornell.
318 R. F o l m a n et al. [IV

Fic. 39. Moving atoms using a magnetic conveyor belt. (a) Potential for various phases of the
movement. The numbers indicate the position of the atoms as shown in the absorption images in
column (b). (c) Linear collider experiment: left, time evolution of the centers of mass of the two
clouds; right, absorption images of the colliding atoms. Courtesy J. Reichel.

a setup (Mfiller et al., 1999). Another similar use of this principle (in this case,
not surface-mounted), in which a storage ring has been realized is presented
in Sect. III.A.7. Although advantageous for the lack of bias fields, this concept
may be hard to implement on miniaturized atom chips as the atoms would be
extremely close to the surface for 1-2-~tm thick wires.
Guiding with semi-permanent magnets has also been achieved (Rooijakkers
et al., 2001; Vengalattore et al., 2001). These materials enhance the magnetic
fields coming from current-carrying wires (see Sect. II.A.10 for a description).
Completely permanent magnets are also being contemplated to avoid current
noise. However, to the best of our knowledge, only atom mirrors have thus far
been realized this way (Hinds and Hughes, 1999).
A further limitation of the guides described above is that they rely on thermal
velocity. Much more control can be achieved by transporting atoms using moving
potentials, as described in Sect. II.A.7. Such a transport device was implemented
in an experiment in Munich. Using the movable 3-dimensional potentials of their
'motor', atoms can be extracted from a reservoir and moved or stopped at will
(Hfinsel et al., 2001b) (Fig. 39). This considerably improves the possibilities of
the chip, as demonstrated by the 'linear collider' shown in Fig. 39c, in which
the motor was used to split a cloud in two and then to collide the two halves
(Reichel et al., 2001).
IV] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 319

C.3. Beam splitters


As discussed in Sect. II.A.8, one may combine the wire guides as described in
the previous section to build more complicated atom-optical elements. One such
element is a beam splitter. A simple configuration is a Y-shaped wire (Fig. 40a)
which creates a beam splitter with one input guide for the atoms, the central wire
of the Y, and two outputs guides, the right and left arms. The atoms are split by

FIG. 40. Beam splitter on a chip. (a) Chip outline; (b) fluorescence images of guided atoms. Two
large U-shaped 200 ~tm wires are used to load atoms into the input guide of a 10-~tm Y-beam splitter.
In the first two pictures in (b), a current of 0.8 A is driven only through one side of the Y, therefore
guiding atoms either to the left or to the right; in the next two pictures, taken at two different bias
fields (12 G and 8 G, respectively), the current is divided in equal parts and the guided atoms split
into both sides. (c) Switching atoms between left and right is achieved by changing the current ratio
in the two outputs and keeping the total current constant as before. The points are measured values
while the lines are obtained from Monte Carlo simulations with a 3-G field along the input guide.
The kinks in the lines are due to Monte Carlo statistics.
320 R. F o l m a n et al. [IV

means of symmetric scattering off the potential hill, which they encounter at the
splitting point. Such a beam splitter on an atom chip was realized by Cassettari
et al. (2000) in Innsbruck. Atoms were released from a chip microtrap and guided
into the beam splitter. Depending on how the current in the input wire is sent
through the Y, atoms can be switched from one output arm of the Y to the other,
or directed to the two outputs with any desired ratio (Fig. 40). Similar beam
splitters have been widely used for the splitting of guided electron waves in solid-
state quantum electronics devices. For example, two Y splitters were put back
to back to form an Aharonov-Bohm type interferometer (Buks et al., 1998).
A four-port beam splitter has been realized at JILA by the group of E. Cornell
and D. Anderson by making a near X-shape out of two wires which avoid a full
crossing (Mtiller et al., 2000). In this experiment, two input guides formed by
two current-carrying wires, merge at the point of closest approach of the wires
so that the two minima merge into one, and then again split into two independent
minima.

C. 4. B E C on a chip

A degenerate quantum gas in a microtrap is an ideal reservoir from which atoms


can be extracted for the experiments on the chip. For example a BEC will take a
similar role as source of bosonic matter waves as the Fermi sea has in quantum
electronics. A clear advantage of a BEC is the higher efficiency of the transfer
to the smallest compressed surface traps, which involves high compression,
leading to large losses for thermal atoms if the trap depth is not appropriate.
The condensate occupies the trap ground state and should follow any adiabatic
compression of the trap. Second, a BEC in a microtrap also provides the initial
atomic state needed to initiate delicate quantum processes such as interference
or even a well-defined entanglement between atoms in two nearby traps.
In the last year three groups in Tfibingen, Munich and Heidelberg succeeded
in making and holding a Bose-Einstein condensate in a surface trap (Ott et al.,
2001; HS.nsel et al., 2001a; Reichel, 2002), and the MIT group managed to
transfer a BEC to a surface trap, and load it into it (Leanhardt et al., 2002).
These experiments showed that making the BEC in a surface trap can be much
simpler. For example in very tight microtraps the BEC is formed in much shorter
time as the tightness of the traps allows for fast thermalization and consequently
fast evaporative cooling which relaxes the vacuum requirements, permitting the
use of a very simple one-MOT setup to collect the atoms (H~insel et al., 2001 a).
In the Tfibingen experiment (Ott et al., 2001), a relatively large condensate
of 4 • 105 87Rb atoms has been formed at a height of some 200 ~tm above the
surface. The experiment made use of a pulsed dispenser as an atom source,
allowing ultra high vacuum (2x 10-11 mbar) while the dispenser was off. This
enabled the use of a simple single-MOT setup. In the experiment, atoms were
transferred magnetically from a distant six-beam MOT to the chip using two
IV] M I C R O S C O P I C ATOM OPTICS: F R O M WIRES TO ATOM CHIP 321

FIG. 41. Left: the Ttibingen setup. The first pair of coils (right) produced the MOT and then the
atoms were conveyed to the trap formed by the second pair of coils. The chip mounting is visible
within the second pair of coils. Right: absorption images of the compression and final cooling stage;
(a) compression into the microtrap; (b) RF cooling in the microtrap; (c) release of the condensate
after 5, 10, and 15 ms time of flight. Courtesy C. Zimmermann.

adjacent pairs of coils (Fig. 41). In the chip trap, condensation was reached
after 10 to 30 s of forced RF evaporative cooling. Aside from being the first
surface BEC, the chip used in Tfibingen with its 25-mm long wires provides a
highly anisotropic BEC (aspect ratio 105), approaching a quasi one-dimensional
regime. In recent work, the BEC was taken to a height of only 20~tm without
observing substantial heating (Fortagh et al., 2002). The smallest structure
holding the BEC was a 3 x 2 . 5 / t m 2 cross section copper wire with a current
of 0.4 A. The BEC had a lifetime of 100 ms in the compressed trap (limited by
3-body collisions) and a 1 s lifetime once it was expanded into a larger trap.
The second experiment producing a BEC in a microtrap was performed in
Munich (H~insel et al., 2001a) (Fig. 42). Here an attractively simple setup
with a continuous dispenser discharge was used. Consequently, the vacuum
background pressure was high (10 -9 mbar) and evaporative cooling had to be
achieved quickly. RF cooling times were as short as 700 ms thanks to the strong
compression in the microtrap which results in a high rate of elastic collisions.
The final BEC included some 6000 atoms at a height o f 70 ~tm. The trapping
wire was 1 . 9 5 m m long and had a cross section of 50• the current
density approached 106 A / c m 2. Strong heating of the cloud was observed in this
experiment but the source remains elusive (possibly, current noise). A beautiful
feature of this experiment is the use of the magnetic 'conveyor belt' described
before (Sects. II.A.7 and IV.C.2, Fig. 39), in order to transport the BEC during
a time of lOOms over a distance o f 1.6 m m without destroying it (see Fig. 42).
Furthermore, the ability of the 'motor' to split clouds was used to show that a
322 R. Folman et al. [IV

FI6. 42. Munich atom chip BEC experiment. Left: schematics of the simple vapor cell apparatus.
Center: time-of-flight images showing the formation of a BEC. Right: (a) the BEC is transported
in a movable 3-dimensional potential minimum; (b) at the end it is released and is observed falling
and expanding. Courtesy J. Reichel.

BEC survives such a splitting. Two such halves were then released into free fall
exhibiting interference fringes as they overlapped.
In the third experiment, performed in Heidelberg, typically 3 • 105 87Rb atoms
were condensed in a Z-wire Ioffe-Pritchard trap, created by a structure under-
neath the chip, and subsequently transferred to a Z-trap on the chip (Fig. 43).
First, more than 3• 108 atoms are loaded into a mirror MOT ( < 1 0 -li torr)
created by external quadruople coils using a double MOT configuration with a
continuous push beam. The atoms are then transferred into a U-MOT, where they
are compressed and after molasses-cooling loaded into a Z-wire trap. The BEC
is formed by forced RF evaporation in typically 20 seconds. Creating the BEC
using a wire structure underneath the chip allows to place other surfaces close
to the BEC while maintaining the high precision of a microtrap for manipulating
the cold atoms. This will open up the possibility to study surfaces with the cold
atoms and to transfer the BEC to surface traps based on dipole forces in light
fields created by micro-optic elements and evanescent fields.
The MIT group transported a BEC of the order of 10 6 Na atoms into an
auxiliary chamber and loaded it into a magnetic trap formed by a Z-shaped
wire (Gustavson et al., 2002) (Fig. 44). This was accomplished by trapping
the condensate in the focus of an infrared laser and translating the location
of the laser focus with controlled acceleration. This transport technique avoids
the optical and mechanical access constraints of conventional condensate
experiments. The BEC was consequently loaded into a microstructure (Leanhardt
et al., 2002).
Finally, we would like to note that currently other groups are also working
IV] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 323

FIG. 43. Heidelberg atom chip BEC experiment. (a) Schematics of the double MOT setup. Atoms
from a lower vapor cell MOT are transferred to a UHV mirror MOT using a continuous push beam.
(b) Photograph of the upper (UHV) chamber. (c) The mounted chip and the U- and Z-shaped wire
structure underneath the chip (inset). (d) Thermal cloud, BEC with thermal background, and pure
BEC released and expanded for 15 ms.

FIG. 44. Transfer of a BEC to a microtrap. Left: schematics of the setup with the science chamber
housing the Z-trap on the far left and the BEC production chamber on the right. Right: condensates
in the science chamber (a) optical trap and (b) Z-trap. The condensate was (c) released from an
optical trap and imaged after 10 ms time of flight and (d) released from a wire trap and imaged after
23 ms time of flight. (e) Schematic of the Z-trap. Courtesy W. Ketterle.

towards BEC in s u r f a c e t r a p s , a n d w e e x p e c t to s e e m a n y d i f f e r e n t s u c c e s s f u l
experiments shortly.
324 R. Folman et al. [V

V. Loss, Heating and Decoherence

For atom chips to work, three main destructive elements have to be put under
control:
(i) Trap loss: It is crucial that we are able to keep the atoms inside the trap as
long as needed.
(ii) Heating: Transfer of energy to our quantum system may result in excitations
of motional degrees of freedom (e.g. trap vibrational levels), and conse-
quently in multimode propagation which would render the evolution of the
system ill-defined.
(iii) Decoherence or dephasing as it is sometimes referred to also originates from
coupling to the environment. While heating requires the transfer of energy,
decoherence is more delicate in nature (Stern et al., 1990). Nevertheless, the
effect is just as harmful because superpositions with a definite phase relation
between different quantum states are destroyed. This has to be avoided, e.g.
for interferometers or quantum information processing on the atom chip.
In discussing these three points, we focus on the particularities of atoms in
strongly confined traps close to the surface of an atom chip. The small separation
between the cold atom cloud and the 'hot' macroscopic environment raises the
intriguing question of how strong the energy exchange will be, and which limit
of atom confinement and height above the surface can ultimately be reached.
We review theoretical results showing that fluctuations in the magnetic trapping
potential give a fairly large contribution to both atom loss and heating. In
addition, thermally excited near fields are also responsible for loss and may
impose limits on coherent atom manipulation in very small (~tm-sized) traps
on the atom chip. Estimates for the relevant rates are given, and we outline
strategies to reduce them as much as possible. Experimental data are not yet
reliable enough to allow for a detailed test of the theory, but there are indications
that field fluctuations indeed influence the lifetime of chip traps (Hansel et al.,
2001a; Fortagh et al., 2002).

A. Loss MECHANISMS

A.1. Spilling over a finite potential barrier

Compression of a thermal atom cloud can lead to losses when the cloud
temperature rises above the trap depth. The smallest losses occur if the
compression is adiabatic. Atoms then stay in their respective energy levels as
the level energy increases. They can nevertheless be lost during trap compression
because of the finite trap depth. It should be noted that this loss occurs for the
highest energies in the trap and can also be used to evaporatively cool the cloud
(see Luiten et al., 1996 for a review).
V] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 325

A.2. Majorana flips


If the atomic magnetic moment is not able to follow the change in the direction
of the magnetic field, the spin flips, and a weak-field-seeking atom can be turned
into a strong-field seeker which is not trapped (Majorana, 1932; Gov et al., 2000).
This occurs when the adiabatic limit (Larmor frequency col much larger than trap
frequency co) does not hold. Majorana flips thus happen at or near zeros of the
magnetic field. For this reason, additional bias fields are employed to 'plug the
hole' in the center of a quadrupole field.
For a magnetic field configuration with a zero, loss can be reduced if the
atoms circle around it. The loss rate is then inversely proportional to the angular
momentum because the latter determines the overlap with the minimum region
(Bergeman et al., 1989; Hinds and Eberlein, 2000).
In Ioffe-Pritchard traps with nonzero field minimum Bip, there is a finite
residual loss rate that has been calculated by Sukumar and Brink (1997). For
a model atom with spin 1/2 in the vibrational ground state, one gets
~CO
y = -~ exp(-It[IBip/hco)

__ 6• s_l co/2 ~ ((itll/itB)(Bip/1G)) (18)


100 kHz exp - 14 co/2;r 100 kHz '
where /~11 is the component of the magnetic moment parallel to the trapping
field. Note the exponential suppression for a sufficiently large plugging field
Bip, typical of nonadiabatic (Landau-Zener) transitions. Choosing a Larmor
frequency coc = ItljBip/h > 10co, one gets a lifetime larger than '~104 trap
oscillation periods. A ratio coL/co > 20 already pushes this limit to ~ 108.

A.3. No&e-induced flips


Fluctuations in the magn~/tic trap fields can also induce spin flips into untrapped
states, and lead to losses. These fluctuations are produced by thermally excited
currents in the metallic substrate or simply by technical noise in the wire
currents. Fluctuations of electric fields and of the Van der Waals atom-surface
interaction have been shown to be less relevant for typical atom traps (Henkel
and Wilkens, 1999; Henkel et al., 1999).
The trapped spin is perturbed via the magnetic dipole interaction and flips at
a rate given by Fermi's Golden Rule:
1
Y - 2h 2 Z (il/tklf)(fl/t'li)S~'(COL), (19)
k,l = x,y,z

where S~t(COL) is the noise spectrum of the magnetic fields, taken at the Larmor
frequency COL.We use the following convention for the noise spectrum:

SO(co) = 2 dr e i~~ (Bi(t + r)Bj(t)) , (20)


oo
326 R. Folman et al. [V

Table II
Trace of the geometric tensor Y/j that determines the loss due to the thermally fluctuating magnetic
near field, according to the rate (22) a

Geometry Half-space Layer Wire

Tr Yij ~/h Jrd/h 2 jr2 a2/(2h 3)

a The metallic layer has a thickness d, assumed much smaller than the distance h to the trap center.
The wire has a radius a << h, and h << 6 is assumed where 6 is the skin depth of the metal. Taken
from Henkel and P6tting (2001). A more accurate calculation of Tr Y/j corrects the results of table 2
by a factor of 1/2 for the half-space and the layer (Henkel and Scheel, 2002).

where (...) is a time average (experiment) or an ensemble average (theory). The


rms noise is thus given by an integral over positive frequencies

Z
~dco O
(Bi(t)Bj(t))= ~-~SB(o)). (21)

For example, the rms magnetic field in a given bandwidth Af for a white noise
spectrum $8 is given by Brms = v/AfSB. The spectrum SB thus has units G2/Hz.

A.3.1. Thermally excited currents. An explicit calculation of the magnetic


noise due to substrate currents ('near field noise') yields the following estimate
for the loss rate (Henkel et al., 1999; Henkel and P6tting, 2001):

g "-~ 75 S-1 ( t 2 / ~ B ) 2 ( T s / 3 0 0 K) (TRY/)- x 1 gm), (22)


(e/ecu)

where 1/e is the substrate conductivity (for copper, ~)Cu = 1.7x 10-6f~cm) and
Ts is the substrate temperature. Note that the Larmor frequency COLactually does
not enter the loss rate. The 'geometric tensor' YO has dimension (1/length) and
is inversely proportional to the height h of the trap center above the surface
(Table II). The loss rate (22) is quite large for a trap microns above a bulk metal
surface. One can reduce the loss by two orders of magnitude when bulk metal in
the vicinity of the trap is replaced by microstructures. For a thin metallic layer
of thickness d, the loss rate (22) is proportional to d/h 2, and for a thin wire
(radius a), a faster decrease o( a2/h 3 takes over (Table II).
The estimates of Table II apply only in an intermediate distance regime,
d,a << h << 6((OL): on the one hand, when the trap distance h is smaller than
the size of the metallic structures, one recovers a 1/h behaviour characteristic for
V] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 327

FIG. 45. Loss rates in a magnetic trap above a copper surface. Results for two different Larmor
frequencies tOL/2Jr = 1 MHz (curve a) and 100 MHz (curve b) are shown. The arrows mark the
corresponding skin depths 6(~OL). Equation (22) applies to the region h << 6(~OL). See Henkel et al.
(1999) for details. Parameters: spin S = 89 magnetic bias field aligned parallel to the surface. The
loss rate due to the black-body field is about 10-13 s-I at 100MHz (not shown). Reproduced from
Henkel et al. (1999), Appl. Phys. B 69 (1999) 379, Fig. 3, with permission. 9 Springer-Verlag.

a metallic half-space; on the other hand, steeper power laws take over at large
distances, when h gets comparable to or larger than the skin depth

0 = 160 gm v/0/~)Cu (23)


(~( coL ) = [JOcoL V/coL /2Yr 1 MHz"

Recall that the skin depth characterizes the penetration of high-frequency


radiation into a metal. This crossover can be seen in Fig. 45 where the flip
rate (22) is plotted vs. the trap height h for a metallic half-space. For details,
we refer to Henkel et al. (1999) and Henkel and P6tting (2001). Note that an
increase of the Larmor frequency only helps to reduce the substrate-induced flips
in the regime where h >> 6(coc). Equation (23) shows that this requires, for
h _~ 1 gm, quite large Larmor frequencies coL/2~ >> 10 GHz, meaning large
magnetic (bias) fields.

A.3.2. Technical noise. Additional loss processes may be related to fluctuations


in the currents used in the experiment, for example in the chip wires and in
the coils producing the bias and compensation fields. Let us focus on the wire
current, and denote by St(co) its noise spectrum. Neglecting the finite wire size,
the magnetic field Bw = ~ I w / 2 ~ h is given by Eq. (3), and we find the following
upper limit for the noise-induced flip rate:
~2 )2 (~/~B)2
]1 "~ - ~ ( ~lO SI((DL) ~ 1.3 S-1 SI((DL) (24)
-~/ (h/1 ~tm) 2 SSN '

where the reference value Ssy = 3.2 x 10 - 1 9 A2/Hz corresponds to shot noise at a
wire current of 1 A (Ssy = 2elw). This estimate is pessimistic and assumes equal
328 R. Folman et al. [V

noise in both field components parallel and perpendicular to the static trapping
field. Nevertheless, it highlights the need to use 'quiet' current drivers for atom
chip traps. In future chip traps with strong confinement, it may turn out necessary
to reduce current noise below the shot-noise level. This can be achieved with
superconducting wires or permanent magnets, as discussed in Sect. II.A. 10 and
reviewed by Hinds and Hughes (1999). See also Varpula and Poutanen (1984).

A. 4. Collisional losses
A.4.1. Background collisions. Here, collisions between background gas atoms
and trapped atoms endow the latter with sufficient energy to escape the trap. In
order to estimate the loss rate per atom ~', let us assume that the background gas
is dominated by hydrogen molecules and at room temperature. We then get:

]r = nbg Ubg O --- 3.8 • 10 -3 s-l P bg O" (25)


10 -l~ mbar 1 nm 2'

w h e r e Pbg is the background pressure. Typical collision cross sections o are in


the l nm 2 = 100 ~2 range (Bali et al., 1999). As a general rule, one gets a trap
lifetime of a few seconds in a vacuum of 10 -9 mbar. It is clear that vacuum re-
quirements will become more stringent as longer interaction times are required.

A.4.2. Collisions o f trapped atoms. For traps in UHV conditions, and especially
for highly compressed traps and high-density samples, the dominant collisional
loss mechanisms involve collisions between trapped atoms. The scattering of
two atoms leads to a loss rate per atom scaling with the density, while 3-body
collision rates scale with the density squared.
Spin exchange. This process corresponds to inelastic two-body collisions
where the hyperfine spin projections mF are conserved, but not the spins F
themselves. In the alkali atoms 7Li, 23Na, and 87Rb, for example, a collision
between two weak-field-seeking states IF = 1, mF = -1) can lead to the
emergence of two strong-field seekers ]2,-1 ) that are not trapped. This transition
requires an excess energy of the order of the hyperfine splitting to occur, which
is typically not available in cold atom collisions. Exothermic collisions between
the weak-field seekers 11,-1), 12, +1), and 12, +2) are not suppressed, however.
The corresponding rate constant is proportional to n ( a s - aT) 2 where as (at)
are the scattering lengths in the singlet (triplet) diatomic potential (C6t6 et al.,
1994). For 87Rb, these scattering lengths accidentally differ very little, leading
to a very small spin flip rate (Moerdijk and Verhaar, 1996; Burke et al., 1997).
As a consequence, 87Rb is practically immune to spin exchange and can form
stable condensates, even of two hyperfine species (Myatt et al., 1997; Julienne
et al., 1997; Kokkelmans et al., 1997). Spin-polarized samples consisting only
of 12, +2) cannot undergo spin exchange because of mF conservation, the other
available states having smaller F. For more details, we refer to the review by
Weiner et al. (1999) and references therein.
V] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 329

Spin relaxation. This process also results from inelastic two-body collisions,
but does not conserve mF. Spin relaxation is caused by a flip of the nuclear spin
and occurs at a lower rate because of the smaller nuclear magnetic moment. In
87Rb for example, the trapped weak-field seeker I1,-1) may be changed into the
untrapped strong-field seeker I1, + 1). More details can be found in the theoretical
work by Burke et al. (1997), Julienne et al. (1997), Timmermans and C6t6
(1998), the experimental work of Gerton et al. (1999) and S6ding et al. (1998),
and the review paper by Weiner et al. (1999).
Three-body recombination. In this process, two atoms combine to form a
molecule. Although the molecules may have a definite magnetic moment and
still be trapped, the reaction releases the molecular binding energy that is shared
as excess kinetic energy between the molecule and the third atom. The binding
energy being typically quite large (larger than 100 ~teV), both partners escape the
trap (Fedichev et al., 1996; Moerdijk and Verhaar, 1996; Moerdijk et al., 1996;
Esry et al., 1999). For references to experimental work, see Burt et al. (1997),
S6ding et al. (1999) and Weiner et al. (1999). We expect three-body processes to
be the dominant collisional decay channel in strongly compressed traps because
the collision rate per atom increases with the square of the atomic density.

A.5. Tunneling
Traps very close to the surface might also show loss due to tunneling of atoms out
of the local minimum of the trap towards the surface. The rate can be estimated
by
?'~ e)
1
exp\- (,/
2m[U(z)- E]/dz, (26)
rrier width ~

where U ( z ) - E is the height of the barrier above the energy of the trapped
particle. Tunneling will therefore only be important for states close to the
top of the potential barrier. Low-lying states in traps where the magnetic
field magnitude rises for long distances will have very little tunneling. Even
for atom waveguide potentials as close as 1 ~tm from the surface, tunneling
lifetimes of more than 1000s have been estimated (Pfau and Mlynek, 1996;
Schmiedmayer, 1998).

A.6. Stray light scattering


Residual light can flip the atomic spin via optical pumping. For resonant light,
this happens at a rate of the order of ['(Istray/Isat) where F is the linewidth of
the first strong electric dipole transition (typically, F/2Jr _~ 5 MHz) and/sat the
saturation intensity (typically a few mW/cm2). It is highly desirable to perform
atom chip experiments 'in the dark': a shielding from any stray light at the level
10-6/sat is required for manipulations on a scale of seconds. For more detailed
estimates, we refer to the review by Grimm et al. (2000) on optical traps.
330 R. F o l m a n e t al. [V

Table III
Loss mechanisms for the atom chip (overview)

Mechanism Scaling a Magnitude a Remedy

Spilling over deep trap potential


Background collisionsb Pbg 0.01 s-1 ultra-high vacuum
Majorana flips c m e -C~176 ~ 1 s-1 avoid B = 0
Near-field noise d Ts/~h a 10 s-1 little metal
Current noise e Si(tOL)/h 2 _~ 1 s-1 low-noise drivers
2-body spin exchange f n 10-4 s- l spin polarize
2-body spin relaxation g n 10-2-10 -4 s- 1
3-body collisions h n2 10-9-10 -7 s-1
Tunneling 10-3 s- 1 deep potential, wide
barriers
Stray light /stray keep in the dark

a The columns 'Scaling' and 'Magnitude' refer to loss rates per atom at typical atom chip traps:
density n = 101~ -3, height h = 10~tm, trap frequency to/2Jr = 100kHz, Larmor frequency
tOL/2Jr = 5 MHz.
b Eq. (25).
c Flip rate (18) from trap ground state.
d Eq. (22). The exponent ct = 1,2, 3 for metal half-space, layer, and wire (see table 2). The estimate
10s -l is for a half-space.
e Eq. (24). SI(m)/SsN = 100.
f Experimental result for 87Rb (Myatt et al., 1997).
g Experimental result for Cs and 7Li, respectively (S6ding et al., 1998; Gerton et al., 1999).
h Experimental results for 87Rb and 7Li, respectively (Burt et al., 1997; S6ding et al., 1999; Gerton
et al., 1999).

To s u m m a r i z e , an o v e r v i e w o f the p r e v i o u s loss m e c h a n i s m s is g i v e n in
Table III. We e x p e c t that on the route t o w a r d s ~tm-sized traps with high c o m p r e s -
sion, inelastic collisions and m a g n e t i c field n o i s e will d o m i n a t e the trap loss.

B. HEATING

In the t r e a t m e n t o f loss m e c h a n i s m s in the p r e v i o u s section, h e a t i n g w a s


m e n t i o n e d in relation to adiabatic c o m p r e s s i o n w h e r e s o m e a t o m s gain e n e r g i e s
larger than the trap depth. H e r e w e discuss a different f o r m o f h e a t i n g , in
w h i c h the a t o m e x c h a n g e s e n e r g y w i t h the e n v i r o n m e n t . S u c h h e a t i n g d o e s not
n e c e s s a r i l y cause the a t o m to be lost, but it is still v e r y h a r m f u l as e x c i t a t i o n s o f
vibrational d e g r e e s o f f r e e d o m lead to an ill-defined q u a n t u m state o f the s y s t e m .
In the case o f the a t o m s y s t e m and the chip e n v i r o n m e n t , the e n v i r o n m e n t is
always hot c o m p a r e d to the system. E n e r g y e x c h a n g e thus i n c r e a s e s b o t h the
system's m e a n e n e r g y and its e n e r g y spread. In the f o l l o w i n g , w e first d e s c r i b e
V] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 331

the influence of position and frequency noise using the harmonic oscillator
model, then turn to substrate and technical noise, and finally touch upon the
issue of heating due to light fields.

B.1. Harmonic oscillator model


Let us consider the trap potential to be a one-dimensional harmonic potential
with angular frequency 6o and with a ground-state size of a0 = ( h / ( 2 M o ) ) ) 1/2,
where M is the mass of the vibrating atom. Assume for simplicity that the atom
is initially prepared in the oscillator ground state 10). Heating can occur as a
result of a fluctuating trap either in frequency or position (see, for example,
Gehm et al., 1998; Turchette et al., 2000). These processes may be described by
transition rates to higher excited states of the oscillator. For example, fluctuations
in the trap position (amplitude noise) are equivalent to a force acting on the atom.
They drive the transition 0 ~ 1 between the ground and first excited vibrational
states, with an excitation rate given by (Gehm et al., 1998; Henkel et al., 1999)

['0---* 1 -- -•2 SF(CO)


SF(O0) -- 4ho)M
(27)

that is determined by the noise spectrum of the force at the oscillator frequency
SF(~O). The rate of energy transfer to the atom ('heating rate') is simply F0~lho)
or SF(O))/4M. Note that this estimate remains valid for an arbitrary initial state.
We may make contact with the work of Gehm et al. (1998) by noting that
fluctuations Ax of the trap center are equivalent to a force

F = Me)2Ax. (28)

In terms of the fluctuation spectrum of the trap center Sx(~O), the excitation
rate (27) is thus given by

Me) 3 09 2
F0___, 1 = Sx((D ) = -x-Sx/ao((_D), (29)
4h ~5

which is equivalent to the heating rate (12) of Gehm et al. (1998), given our
definition (20) of the noise spectrum.
Fluctuations of the trap frequency are described by the Hamiltonian MxZo)Aco
and heat the atom by exciting the 0 ~ 2 transition. The corresponding transition
rate is (Gehm et al., 1998)

ro_~2 = 1S~o(2o)) (30)

and involves the frequency noise spectrum at twice the trap frequency. Using
the rates given by Gehm et al. (1998), one can show that the heating rate due
332 R. Folman et al. [V

to frequency fluctuations is equal to F0~2(4(E) + hog), where the mean energy


(E) = 89 in the ground state.
In the following, we differentiate between thermal fluctuations and technical
ones. To get the total heating rate, one simply adds the force fluctuation spectra
SF(~) of all the relevant sources (e.g. electromagnetic noise from radio stations).

B.2. Thermal fluctuations


Magnetic fields generated by thermally excited currents in the metallic substrate
correspond to a force given by the gradient of the Zeeman interaction - p . B.
An explicit calculation of the magnetic gradient noise gives the following force
spectrum (Henkel and Wilkens, 1999; Henkel et al., 1999)

SF(O)) -- 3-~ h3 , (31)


where p is the magnetic moment and /tll its component parallel to the static
trapping field. The expression (31) applies to a planar metallic substrate (half-
space) and an oscillation perpendicular to its surface. Again, the noise spectrum
is actually frequency independent as long as h << 6(~o) where 6(09) is the
skin depth (23). The average magnetic moment is taken in the trapped spin state
[see Henkel et al. (1999) for details]. We thus obtain the following estimate for
the excitation rate (27):

F0~l '~ 0.7s -1 (~/ptB)Z(Ts/300 K) (32)


(M/amu)(oo/2er 100 kHz)(e/eCu)(h/1 ~tm) 3"
Fig. 46 shows a plot of heating rates for varying substrates and trap heights
according to Eq. (32).

FIG. 46. Heating rate for a trapped spin above copper and glass substrates. Parameters: trap
frequency to/2Jr = 100 kHz, M = 40 amu, magnetic moment/t =/t B = 1 Bohr magneton, spin S - 1.
The heating rate due to the magnetic black-body field (not shown) is about 10-39 s-1 . For the glass
substrate, a dielectric constant with Re e = 5 and a resistivity e = 1011 ~2cm are taken. Reproduced
from Henkel et al. (1999), Appl. Phys. B 69 (1999) 379, Fig. 5, with permission. 9 Springer-Verlag.
V] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 333

For lithium atoms, a typical trap frequency of 100kHz and a height of


h = 101am, we estimate a heating rate of 10 -4 s-1. For time scales typical of
atom chip experiments (1-100ms), thermal fluctuations thus lead to tolerable
heating only for traps with h > 100 nm.

B.3. Technical noise


Heating due to technical noise may arise from fluctuations in the currents used in
the experiment. Noise in the chip wire currents and in the bias and compensation
fields, for example, randomly shifts the location of the trap center. Let us focus on
fluctuations in the chip wire current Iw. Neglecting finite size effects, the current
and the bias field Bb produce the magnetic trap at a height of h = IAOIw/2~Bb
(Eq. 3). The conversion from the current noise spectrum SI(O)) to the force
spectrum required for the heating rate (27) is simply

SF((,O)=(IA~176
2SrBb
2$I ((o), (33)
and we end up with an excitation rate

F0--,1 = 1.4 s-l (M/amu)(oo/2ar 100 kHz) 3 SI(oo)/SsN (34)


(Bb/1 G) 2"
The reference SSN for the current noise is again the shot-noise level at lw = 1 A.
Note that this rate increases with the trap frequency: while a strong confinement
suppresses heating from thermal fields (Eq. 32), the inverse is true for trap
position fluctuations. This is because in a potential with a large spring constant,
position fluctuations translate into large forces (Eq. 28). Typical trap parameters
(o)/2:z = 100kHz, Bb = 50G) lead for 7Li atoms to an excitation rate of
-~4• 10-3 s-1 • S1(~o)/SsN. This estimate shows that even for very quiet currents
technical noise is probably the dominant source of heating on the atom chip.
The fluctuations of the trap center (location proportional to Iw/Bb) can be
reduced by correlating the currents of the bias field coils and the chip wire so that
they have the same fluctuations, up to shot noise. Heating due to fluctuations in
the trap frequency may then be relevant, as co is proportional to BZ/Iw (Eq. 6).
Let us again calculate an example. For a fixed ratio Iw/Bb (due to correlated
currents), we find for the relative frequency fluctuations
A(o AI
- (35)
co Iw
and hence an excitation rate (30)

F0-+2 ,-,o 10. 7 S-1 (~o/2.rr 100 kHz) 2 SI(2O)). (36)


(Iw / 1 A) 2 SSN
Typical atom chip parameters (a)/2oz = 100 kHz, Iw = 1 A) lead to F0+2 "~
10-7 s-1 • Si(2co)/SsN, which is negligible when compared to the rate obtained
in Eq. (34).
334 R. Folman et al. [V

B.4. Light heating

Another source of heating are the external light fields with which the atoms are
manipulated and detected. Here the Lamb-Dicke parameter r/ is a convenient
tool, where 27ra0
r/- ~ (37)

is the ratio between the ground-state size of the trap a0 and the wavelength of
the impinging wave. This becomes clear if we remember that the probability not
to be excited P0~0 is simply the well-known Debye-Waller factor

exp(-Ak2a 2) ~ exp(-r/2), (38)

where Ak ~ k is the momentum loss of the impinging photon. Hence, if the


atoms are confined below the photon wavelength (the so-called Lamb-Dicke limit
t/ < 1), they will not be heated by light scattering. Loss via optically induced
spin flips is still relevant, however, as discussed in Sect. V.A.6 and reviewed by
Grimm et al. (2000).
In Table IV we give an overview of the heating mechanisms discussed above.
For microscopic traps, we expect noise from current fluctuations and (to a lesser
extent) from the thermal substrate to be the dominant origins of heating. Note the
scaling with the trap frequency: trap fluctuations due to technical noise become
more important for guides with strong confinement.
In this subsection, we have restricted ourselves to heating due to single-
atom effects. Collisions with background gas atoms also lead to heating and
rate estimates have been given by Bali et al. (1999). Finally, in an on-
chip Bose condensate, fluctuating forces may be expected to drive collective

Table IV
Heating mechanisms for the atom chip (overview)

Mechanism Scaling a Magnitude b Remedy

Near-field noise b Ts/co~)h 3 10-4 s -l


Current noise c co3S I / B 2 ~ coSl/h 2 1 s- l correlate currents
Trap frequency noise d co2S1/I2w ~ S l / h 4 10-5 s-l
Light scattering 1/co,~2 reduce stray light

a The columns 'Scaling' and 'Magnitude' refer to transition rates from the ground state of a typical
atom chip trap: lithium atoms, height h = 10~tm, trap frequency co/2sr = 100kHz. Harmonic
confinement is assumed throughout.
b Eq. (32), for a metal half-space.
c Eq. (34). Note the scalings c o - B Z / I w and h ~ I w / B b for trap frequency and height.
d Eq. (36).
V] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 335

and quasiparticle excitations, leading to a depletion of the condensate ground


state (Henkel and Gardiner, 2002). This area deserves further study in the
near future.

C. DECOHERENCE

We now turn to the destruction of quantum superpositions or interferences


due to the coupling of the atom cloud to the noisy chip environment. This is
an important issue when coherent manipulations like interferometry or qubit
processing are to succeed on the atom chip. With chip traps being ever closer to
the chip substrate, thermal and technical magnetic noise is expected to contribute
seriously to decoherence, as it does to loss and heating processes.
The theoretical framework for describing decoherence makes use of the
density matrix for the trapped atoms. Its diagonal elements give the occupation
probabilities, or populations, in some preferred basis, usually the stationary trap
states. Their evolution has been discussed in the previous subsections in terms
of simple rate equations and constants. Decoherence deals with the decay of
off-diagonal elements, or coherences, of the density matrix. Their magnitude
can be related to the fringe contrast one obtains in an interference experiment.
Magnetic fluctuations typically affect both populations and coherences: field
components perpendicular to the trapping fields redistribute the populations,
and parallel components suppress the coherences. The latter case illustrates
that decoherence can occur even without the exchange of energy, because
it suffices that some fluctuations randomize the relative phase in quantum
superposition states (Stern et al., 1990). Such fluctuations are sometimes called
'phase noise'.
In this subsection we consider first the decoherence of internal atomic states
and then describe the impact of fluctuations on the center-of-mass. In the same
way as for the heating mechanisms, we leave aside the influence of collisions on
decoherence, nor do we consider decoherence in Bose-Einstein condensates.

C. 1. Internal states

The spin states of the trapped atom are promising candidates for the implemen-
tation of qubits. Their coherence is reduced by transitions between spin states,
induced by collisions or noise. The corresponding rates are the same as for the
loss processes discussed in Sect. V.A.
In addition, pure phase noise occurs in the form of fluctuations in the
longitudinal magnetic fields (along the direction of the trapping field). These shift
the Larmor frequency in a random fashion and hence the relative phase between
spin states. The corresponding off-diagonal density matrix element (or fringe
contrast) is proportional to (exp(iA~)) where Aq~ is the phase shift accumulated
336 R. Folman et al. [V

due to noise during the interaction time t. A 'decoherence rate' Ydec can be
defined by (A~ 2) S0)(o)--+ 0)
'/dec -- (39)
2t 4 '

where S~(o)) is the spectrum of the frequency fluctuations. Two spin states
[mF), Imp-), for example, 'see' a frequency shift (p(t)- g l l B ( m F - m~F)ABil(t)/h,
that involves the differential magnetic moment and the component ABII(t ) of
the magnetic field noise parallel to the trap field. The spectrum S~(co) is then
proportional to the spectrum of the magnetic field fluctuations.
Equation (39) is derived in a rotating frame where the phase shift has zero
mean and making the assumption that the spectral density Sr is essentially
constant in the frequency range o) ~< 1/t. The noise then has a correlation time
much shorter than the interaction time t. We consider, as usual in theory, that
A~ is a random variable with Gaussian statistics, and get a fringe contrast

(e iaq~) = e -ydect (40)

that decays exponentially at the rate (39).


Let us give an estimate for the decoherence rate due to magnetic noise. If
AB(r, t) are the magnetic fluctuations at the trap center, the shift of the Larmor
frequency is given by

(il~llli)
AwL(t) - ~ABil(r, t). (41)

Here, the average magnetic moment is taken in the spin state ]i) trapped in the
static trap field, thus picking the component ABI[ parallel to the trap field. The
noise spectrum of this field component, for thermal near field noise, is of the
same order of magnitude as for the perpendicular component (Henkel et al.,
1999) and depends only weakly on frequency. We thus get a decoherence rate
comparable to the loss rate (22), typically a few l s-1 . The same argument can be
put forward for fluctuations in the wire current and the bias field. Assuming a flat
current noise spectrum at low frequencies, we recover the estimate (24) for spin
flip loss (a few l s-1). Therefore, keeping the atoms in the trap, and maintaining
the coherence of the spin states requires the same effort.
We finally note that near field magnetic noise also perturbs the coherence
between different hyperfine states that have been suggested as qubit carriers.
Although these states may have the same magnetic moment (up to a tiny
correction due to the nuclear spin), excluding pure phase noise, their coherence
is destroyed by transitions between hyperfine states. The corresponding loss rate
(relevant, e.g., for optical traps) has been computed by Henkel et al. (1999) and
is usually smaller than the spin flip rate.
V] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 337

C.2. Motional decoherence


The decoherence of the center-of-mass motion of a quantum particle has been
put forward as an explanation for the classical appearance of macroscopic objects
since the work of Zeh (1970) and Zurek (1991) (see also the book by Giulini
et al., 1996). It has been shown that the density matrix of a free particle subject
to a random force field in the high-temperature limit evolves into a diagonal
matrix in the position basis (Zurek, 1991)

p(z, zt, t) ~ p(z z,,O)exp I (z-z')ZDt 1


' h2 9
(42)

Here, the distance z - z t denotes how 'off-diagonal' the element is, and D is the
momentum diffusion coefficient. The coherence length thus decreases like

h
~:~ - v / ~ . (43)

At the same time, the momentum spread Ap "-" (2Dt) 1/2 increases, so that the
relation Ap~c -~ h is maintained at all times. At long times, ~c will be limited by
the thermal de Broglie wavelength at the equilibrium temperature. However, this
regime will not be reached on atom chips for typical experimental parameters.
For a particle trapped in a potential, the density matrix tends to a diagonal
matrix in the potential eigenstate basis if the timescale for decoherence is
large compared to the oscillation time 2zr/co. This regime typically applies
to the oscillatory motion in atom chip waveguides. The regime in which the
two timescales are comparable has been discussed by Zurek et al. (1993) and
Paz et al. (1993); it leads to the 'environment-induced selection' of minimum
uncertainty states (coherent states for a harmonic oscillator).
In the following we discuss different decoherence mechanisms for a typical
separated path atom interferometer on the atom chip.

C.3. Longitudinal decoherence


We focus first on the quasi-free motion along the waveguide axis (the z-axis),
using the free particle model mentioned above. Decoherence arises again from
magnetic field fluctuations due to thermal or technical noise. The corresponding
random potential is given by (41):

V(r, t) = -(il/tll Ii)ABII(r, t), (44)


where we retain explicitly the position dependence. Henkel and P6tting (2001)
have shown that for white noise, the density matrix in the position representation
behaves as
p(z,z', t) = p(z,z', O) e x p (-Ydec(Z - z t ) t), (45)
338 R. Folman et al. [V

where the decoherence rate Ydec(s) depends on the spatial separation s = z - z'
between the two parts of the atomic wave function being observed:

1 -C(s)
Ydec(S) = 2],l 2 Sv(h; co ~ 0). (46)

Here, C(s) is the normalized spatial correlation function of the potential (equal
to unity for s = 0), and the noise spectrum Sv(h; co ---+ 0) characterizes the
strength of the magnetic noise at the waveguide center.
For an atom chip waveguide perturbed by magnetic near field noise, the
decoherence rate is of the order of

0)
y = 2h 2 (47)

and hence comparable to the spin flip rate (19, 22). Decoherence should thus
typically occur on a timescale of seconds. The correlation function C(s) is
well approximated by a Lorentzian, as shown by Henkel et al. (2000), and the
decoherence rate (46) can be written as

ys 2
Ydec(S)- $2 + 12, (48)

where lc is the correlation length of the magnetic noise. This length can be
taken equal to the height h of the waveguide above the substrate (Henkel et al.,
2000). This is because each volume element in the metallic substrate generates a
magnetic noise field whose distance-dependence is that of a quasi-static field (a
1/r 2 power law). Points at the same height h above the surface therefore see the
same field if their distance s is comparable to h. At distances s >> h, the magnetic
noise originates from currents in uncorrelated substrate volume elements, and
therefore C(s) ~ O. The corresponding saturation of the decoherence rate (48),
Ydec(s >> lc) ---. y, has also been noted, for example, by Cheng and Raymer
(1999).
Decoherence due to magnetic noise from technical sources will also happen
at a rate comparable to the corresponding spin flip rate, as estimated in Eq. (24).
The noise correlation length may be comparable to the trap height because the
relevant distances are below the photon wavelength at typical electromagnetic
noise frequencies, so that the fields produced by wire current fluctuations are
quasi-static, and the same argument applies. The noise correlation length of
sources like the external magnetic coils will, of course, be much larger because
these are far away from the waveguide. These rough estimates for the spatial
noise properties of currents merit further investigation, in particular at the shot-
noise level (Henkel et al., 2002).
V] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 339

I F(s,t) I
1

0.8
=

0.6

0.4

0.2

s/l c
1 2 3 4 5 6

FIc. 47. Illustration of spatial decoherence in an atomic wave guide. The spatially averaged
coherence function F(s, t) = f d z p(z + s,z, t) is plotted vs. the separation s for a few times t. Space
is scaled to the field correlation length lc and time to the 'scattering time' 1/y _= 1/Ydec(~ ).
A Lorentzian correlation function for the perturbation is assumed. Reproduced from Henkel and
P6tting (2001), Appl. Phys. B 72 (2001) 73, Fig. 3, with permission. 9 Springer-Verlag.

Spatial decoherence as a function of time is illustrated in Fig. 47 where the


density matrix p(z + s,z, t) averaged over z is plotted. Note that this quantity
will be directly proportional to the visibility of interference fringes when two
wavepackets with a path difference s interfere. One sees that for large splittings
s >> lc, the coherence decays rapidly on the timescale 1/y given in Eq. (47). This
is because the parts of the split wavepacket are subject to essentially uncorrelated
noise. In a typical waveguide at height h = 10 Ftm, fringe contrast is thus lost
after 0.1-1 s (the spin lifetime) for path differences s >> 10 ~tm. Increasing the
height to h = 100 ~tm decreases y by at least one order of magnitude as shown by
Eq. (22). In addition, the correlation length grows to 100 ~tm, and larger splittings
remain coherent. Alternatively, one can choose smaller splittings s << lc which
decohere more slowly because the interferometer arms see essentially the same
noise potential. Note, however, that the spin lifetime will always be the upper
limit to the coherence time of the cloud.
The previous theory allows to recover the decoherence model of Eq. (42) at
long times t > 1/y. In this limit, only separations s < lc have not yet decohered,
and we can make the expansion

s2
~dec(S) ~ ~ [2 (49)

for the decoherence rate (48). From the density matrix (45), we can then read
off the momentum diffusion constant D hZy/l 2. =

C.4. Transverse decoherence


We finally discuss the decoherence of a spatially split wavepacket in an atom
chip interferometer, as described in Sect. II.A.9.
340 R. Folman et al. [V

C.4.1. Amplitude noise. The excitation of transverse motional states in each


arm suppresses the coherence of the superposition at the same rate as the heating
processes discussed in Sect. V.B (about 1 s-l). Note that due to the transverse
confinement, the relevant noise frequencies are shifted to higher values compared
to the longitudinal decoherence discussed before.

C.4.2. Phase noise. The coherence between the spatially separated interferom-
eter arms is suppressed in the same way as the longitudinal coherence discussed
in Sect. V.C.3. To show this, we use an argument based on phase noise, and
focus again on magnetic field fluctuations, either of thermal or technical origin.
Magnetic fluctuations affect both the bottom of the trap well and the transverse
trap frequency, but are only relevant when they differ in the spatially separated
arms. The well bottoms get differentially shifted from an inhomogeneous bias
field, e.g., while the trap frequency shifts due to changes in the field curvature.
We generalize formula (39) to a phase shift Aq) that is the accumulation of
energy-level differences AE(t) along the paths in the two arms. The decoherence
(or dephasing) rate is thus given by

SAE(O0 ~ O)
Ydec = 4h 2 , (50)

where SAE(CO ~ 0) is the spectral density of the energy difference, extrapolated


to zero frequency.
To make contact with the density matrix formulation of Eq. (45), we write
AE(t) = E R ( t ) - EL(t) where ER, L(t) are the energy shifts in the right and left
interferometer arms that are 'seen' by an atom travelling through the waveguide.
We find

<AE(t) AE(t')) = (ER(t) ER(t')) + (EL(t) Ec(t')) - <ER(t) Ec(t')) - <EL(t) ER(t')) ,
(51)
where the last two terms contain the correlation between the noise in both arms.
They may therefore be expressed through the normalized correlation function
CRC =-- C(s) with s the separation between the left and right arms. The reasonable
assumption that both arms 'see' the same white noise spectrum, say SE(W),
yields

<AE(t) A E ( t ' ) ) = [I - C(s)] SE(o) ~ 0) 6(t - t'), (52)

1 - C(s)
Ydec = ] I d e c ( S ) = 2h 2 SE(o) ~ 0), (53)

where we recover the decoherence rate (46) obtained for the quasi-free
longitudinal motion. We also recover the trivial result that the contrast stays
constant if both interferometer arms are subject to the same noise amplitude
(perfect correlation C ( s ) - 1).
V] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 341

The previous argument shows that transverse and longitudinal coherence are
affected in a similar way by magnetic noise. Again, near field noise is a serious
threat due to its short correlation length. Since the decoherence rate is so
small that Ydec(CX3)t << 1 for interaction times not longer than a few hundred
ms, the phase noise remains small even for widely separated arms subject to
decorrelated noise (separation larger than the guide height). This is a worst-
case estimate: a more careful approach would take into account the form of
the interferometer, where the arm separation is not constant. Current noise
should neither be underestimated. It is certainly possible to reduce dephasing
by feeding the same current through both left and right wire guides, as shown
by Eq. (53). But this does not seem to help at the shot-noise level because each
electron randomly follows one or the other wire. The wire current fluctuations
are thus uncorrelated, leading to a transverse decoherence rate comparable to
the longitudinal decoherence rate. Both rates are thus of the order of the flip
rate (24), typically a few s-1 .
Let us estimate as another example the dephasing due to technical noise in
a magnetic field gradient. This may be introduced by an imperfect Helmholtz
configuration or coil misalignment. For small gradients b, we have

AE(t) =(ktll) s. b(t) (54)

where s is the spatial separation between the interferometer arms. To be precise,


b(t) = VBx(t) gives the gradient of the bias field component along the direction
of the (static) trapping field. Ignoring a possible anisotropy in the gradient noise,
we find the estimate (/tl I) 2 S 2
)tdec(S) ~'~ 4h 2 Sb(o9 ~ 0), (55)

where Sb(og) is related to the power spectrum of the current difference in


the Helmholtz coils. We may take as the worst case completely uncorrelated
Helmholtz currents, and a magnetic gradient b ~_ B b / R where R is the size of
the Helmholtz coils. The dephasing rate is then of the order of

Ydec(S) ~ 10-6 S-1 Q/b)2 $2 (Bb/G)2 Sl(fO --+ O) (56)


tA2 R 2 (Ib/m) 2 SSN

where Ib and Bb are the Helmholtz current and the bias field. The experimentally
reasonable parameters Ib = 1 A, s = 100 ~tm, R = 10 cm, Bb = 10 G yield the
small value I/dec(S) "~ 10-10 S -1 X SI(O) --+ 0 ) / / S s N . We note that the residual
gradient of imperfect Helmholtz coils is usually less than 0.1 G/cm which is an
order of magnitude below the estimate Bb/R - 1 G/cm taken here.
Finally, let us estimate the phase noise due to fluctuations in the spring constant
of the guide potential. Even in the adiabatic limit where the transitions between
transverse quantum states are suppressed (no heating), these fluctuations shift
342 R. Folman et al. [VI

Table V
Decoherence mechanisms for atom chip interferometers (overview)

Mechanism Scaling Magnitude a Remedy

Substrate fields b
s << h Tss2/eh a+2 << 10s -1 little metal,

s >> h Ts/eh a 10 s -1 small splitting

Current noise c w 3 S I / B 2 ~ ogSi/h 1 s -1 correlate currents

Bias fluctuationsd s 2B b2S I / R 2 I~2 ~ s2SI/R 4 10 -8 s -1

Trap frequency noise e ~ 2 3 i / I 2 ~ Si/h 4 10 -5 s -1

a 'Magnitude' refers to the decoherence rate Ydec(S) for a typical guided interferometer: lithium
atoms, height h = 10 gm, separation s - 10 ~tm, transverse guide frequency o9/2:r = 100 kHz. Along
the waveguide axis, the atomic motion is free.
b Exponent a = 1,2, 3 for metal half-space, layer, and wire (Eq. 22 and table 2).
c Eq. (34).
d Eq. (56). The bias field scales as B b ~ Ib/R where R is the size of the bias coils.
e Eq. (36).

the energy of the guided state. In the harmonic approximation, we have for the
ground state of the guide AE = l hA~o where Ao9 is the relative shift of the
vibration frequency. This gives a dephasing rate

Ydec : ~S,,,(Oo --~ 0). (57)

We have neglected noise correlations between the interferometer arms that


would reduce decoherence because of correlated phase shifts in both arms. The
rate (57) is of the same order as the heating rate (30, 36) due to frequency noise
(~_ 10-5 s-~). It thus appears that fluctuations of the trap frequency have a larger
impact than bias field gradients, but still they lead to negligible dephasing.
In Table V we give an overview of the different decoherence mechanisms
discussed in this subsection. For interferometers with large path differences
(compared to the waveguide height), we expect current shot noise and thermal
near field fluctuations to be the dominant sources of decoherence. They give
quite 'rough' potentials (small correlation length) and perturb both the quasi-
free motion along the waveguide axis and the relative phase between spatially
separated wavepackets in an interferometer. An increase in the trap frequency
does not help, rather the amount of metallic material in the vicinity of the guide
should be kept to a minimum.

VI. Vision and Outlook


Much has been achieved in the field of micro-optics with matter waves in the
last 10 years. We have seen a steady development from free-standing wires to
micron-size traps and guides, from trapping thermal atoms to the creation of
VI] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 343

BEC on an atom chip. Where to go from here? What can we expect from future
integrated matter wave devices? There are still many open questions before we
can assess the full promise of integrated microscopic atom optics.
In the following we try to pinpoint the relevant future developments and
directions. Some of them, like the study of the influence of the warm thermal
surface and the fundamental noise limits on lifetime, heating and coherence of
atoms, are already under way. Hopefully in a few years we will know how far
micromanipulation of atoms on chips can be pushed.

A. INTEGRATING THE ATOM CHIP


A. 1. Chip fabrication technology
We will see continued development of atom chip fabrication techniques. De-
pending on how close to the surface one is able to place atoms before significant
decoherence occurs, the commonly used technology will be either state-of-the-art
nanofabrication with scale limits below 100 nm or thicker and larger wires built
by a combination of less demanding techniques. Another limitation would be
smoothness of fabrication: as fluctuations in wire widths would cause changing
current densities and consequently changing trap frequencies, potential 'hills'
may appear which may be large enough to hinder the transport of a BEC or
control its phase evolution.
In the near future many advances in fabrication techniques are expected.
One of the first steps will be to build multilayer structures that will enable for
example crossing wires in order to realize more elaborate potentials and give
more freedom for atom manipulation.
Thin film magnetic materials should allow to build permanent magnetic
microscopic devices, which can be switched on and off for loading and
manipulation of atoms. Such structures would have the advantage that the
magnetic fields are much more stable, and consequently one can expect much
longer coherence times, when compared with current generated fields.

A.2. Integration with other techniques


With cold atoms trapped close to a surface, integration with many other
techniques of atom manipulation onto the atom chip is possible.
One of the first tasks will be to integrate present day atom chips with
existing micro-optics (see for example Birkl et al., 2001) and solid-state optics
(photonics), for atom manipulation and detection. We envision for example
microfabricated wave guides and/or microfabricated lenses on the atom chip for
bringing to and collecting light from atoms in the atom-optical circuits.
Light can also be used for trapping (Grimm et al., 2000). Having cold atoms
close to a surface will allow efficient transfer and precise loading of atoms into
light surface traps, which would be otherwise difficult because of their small
volume and inaccessible location. For example, an atom chip with integrated
344 R. F o l m a n et al. [VI

micro-optics will allow to load atoms into evanescent-wave guides and traps,
as proposed by Barnett et al. (2000). Such traps and guides would be a way
to circumvent the decoherence caused by Johnson noise in a warm conducting
surface (Sect. V).
With the standing wave created by reflecting light off the chip surface one
will be able to generate 2-dimensional traps with strong confinement in one
direction, resembling quantum wells, as demonstrated by Gauck et al. (1998).
Adding additional laser beams or additional electrodes on the surface restricts the
atomic motion further, yielding 2-dimensional devices as in quantum electronics
(Imry, 1987). Similarly one can build and load optical lattices close to the surface
where each site can be individually addressed by placing electrodes on the chip
next to each site.
In principle, many other quantum optical components can be integrated on the
atom chip. For example, high-Q cavities combined with microtraps will allow
atoms to be held inside the cavity to much better than the wavelength of light
providing a strong coupling between light and atoms. For recent experimental
work concerning the manipulation and detection of atoms in cavities, we refer
the reader to Berman (1994), Pinkse et al. (2000), Hood et al. (2000), Osnaghi
et al. (2001) and Guth6hrlein et al. (2001).
Regarding cavities one can think of examining a wide variety of technologies
ranging from standard high-Q cavities consisting of macroscopic mirrors to
optical fiber cavities (with Bragg reflectors or with mirrors on the ends); from
photonic band gap structures to microcavities like microspheres and microdiscs
fabricated from a suitable transparent material. One proposed implementation is
presented in Fig. 48 (Mabuchi et al., 2001).

FIG. 48. A proposed implementationof an integrated nanofabricated high-Q cavity from CalTech.
The cavity is made of a 2D photonic crystal utilizing holes with diameters of order 100nm. A
Weinstein-Libbrecht-type Ioffe magnetic trap will hold the atom in the cavity. CourtesyH. Mabuchi.
VI] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 345

A.3. Atom detection

For future applications, it would be advantageous to have a state-selective single-


atom detector integrated on the atom chip. Such detectors could be based on
different methods. The most direct method would be to detect the fluorescent
light of the atom using surface-mounted micro-optics. More accurate non-
destructive methods could be based on measuring an optical phase shift induced
by an atom in a high-Q cavity.

A.3.1. Single-atom detection using near field radiation. To detect light scattered
from single atoms near a chip surface, the main challenge will be to minimize the
stray light scattered from the surface. One possible solution may be to collect a
large fraction of the light scattered by the atom using near field apertures and/or
confocal microscope techniques. An atom could also be used to couple light
between two wave guides, as used in some micro-optic detectors for molecules
and directional couplers in telecommunication.

A.3.2. Detecting single atoms by selective ionization. This may be achieved


using a multistep process up to a Rydberg state. The electron and the location
from where it came can then be detected with a simple electron microscope.
Using a dipole blockade mechanism as discussed by Lukin et al. (2001) one
should be able to implement an amplification mechanism, which will allow 100%
detection efficiency (Schmiedmayer et al., 2002).

A.3.3. Transmission of resonant light through a small cavity. Such a scheme


may be used to detect single atoms even for moderate Q values of the cavity.
The cavity could be created by two fibers with high reflectivity coatings at the
exit facets, or even by a DBR fiber cavity with a small gap for the cold atoms.
Fiber ends molded in a lens shape could considerably reduce the light losses due
to the gap. Having atoms localized in steep traps should allow a small gap that
would reduce the losses even further.

A.3.4. Transmission of light through a high-Q cavity. Here, the transmission


is modified by the presence of single atoms, and the light may be quite far
from atomic resonance and the atoms are still detected with high probability.
The basic mechanism of this detector is that atoms inside the cavity change the
dispersion for the light. The high Q value makes it possible to detect very small
modifications of the dispersion. In addition the cavities can be incorporated into
integrated optics interferometers to measure the phase shift introduced by the
presence of the atoms. Off resonant detection would allow for nondestructive
atom detection (see for example Domokos et al., 2000).
346 R. Folman et al. [VI

B. MESOSCOPIC PHYSICS

The potentials created on an atom chip are very similar in scale and confinement
to the potentials confining electrons in mesoscopic quantum electronics (Imry,
1987). There electrons move inside semiconductor structures, in our case
atoms move above surfaces in atom-optical circuits. In both cases they can be
manipulated using potentials in which at least one dimension is comparable to
the de Broglie wavelength of the guided, trapped particle. To find similarities
and differences between mesoscopic quantum electronics and mesoscopic atom
optics will probably become a very rich and fascinating research field.
Electrons in semiconductors interact strongly with the surrounding lattice.
It is therefore hard to maintain their phase coherence over long times and
distances. An atomic system on the contrary is well isolated. Furthermore, atoms
(especially in a BEC) can be prepared in such a way that the temperature is
extremely low with respect to the energy level spacing. The consequence is that
phase coherence is maintained over much longer times and distances. This might
enable us to explore new domains in mesoscopic physics, which are hard to reach
with electrons.

B.1. Matter wave optics in versatile potentials


A degenerate quantum gas in the atom chip will allow us to study matter wave
optics in confined systems with non-trivial geometries, such as splitters, loops,
interferometers, etc. One can think of building rings, quantum dots connected
by tunnel junctions or quantum point contacts (Thywissen et al., 1999b), or
even nearly arbitrary combinations thereof in matter wave quantum networks.
For many atomic situations the electronic counterparts can easily be identified.
Atom chips will allow to probe a wide parameter range of transverse ground
state widths, confinement and very large aspect ratios of 105 and more. Atomic
flow can be monitored by observing the expansion from an on-board reservoir
along the conduit. Further perturbations and corrugations can be added at any
stage to the potential by applying additional electric, magnetic or light fields to
modify the quantum wire or quantum well. In this manner we can also explore
how disorder in the guides may change the atomic behavior.
In the following, we give details regarding three exemplary matter wave
potentials on the atom chip.

B.2. Interferometers
In the near future it will be essential to develop and implement interferometers,
and to study through them the decoherence of internal states and external
motional states. Atom chip interferometers have been discussed in detail in
Sect. II.A.9. They can be built either in the spatial (Andersson et al., 2002)
or in the temporal domain (Hinds et al., 2001; H~insel et al., 2001 c). Integrated
VI] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 347

on an atom chip, they are very sensitive devices that may be used to measure
inertial forces or even to perform computation (Andersson and Barnett, 2000).
Coherence properties in more complicated networks can be studied by observing
interference and speckle patterns.
Interferometers can also serve as probes for the understanding of surface-
atom interactions, allowing for a quantitative test of the limits imposed on the
atom chip by the warm surface for both internal state and external (motional)
state coherence. Since many of the important parameters scale with the spin flip
life time in a trap (see Sect. V), a first important step would be to measure the
(BEC) lifetime in a microtrap as a function of distance to the surface. Aside from
heating and spin flips, the surface also induces 'phase noise'. Interferometers will
be able to measure this subtle effect as a function of surface material type and
temperature as well as atom-surface distance and spatial spread of the atomic
superposition, through a reduction in the fringe visibility. Finally, by coupling
microtraps (atomic quantum dots) to one of the interferometer arms, similar to
the mesoscopic electron experiments (Buks et al., 1998), subtle interaction terms
may be investigated, e.g., 1/r second-order dipole interactions as discussed by
O'Dell et al. (2000).
Internal state superpositions of atoms close to surfaces can be studied using
internal state interferometers. Using Raman transitions or microwave transitions
we can create superpositions, observe their lifetime and put theoretical estimates
to the test.

B.3. Low-dimensional systems

Much is known about the behavior of fermions in low-dimensional strongly con-


fining systems (one- and two-dimensional systems) from mesoscopic quantum
electronic experiments. By designing low-dimensional experiments using atoms
(weakly interacting bosons or fermions) we expect to obtain further insight also
about electronic phenomena.
The role of interactions inside an atomic matter wave can range from minimal
in a very dilute system to dominating in a very dense system. Low-dimensional
systems are especially interesting in this context, since it is expected that the
interactions between the atoms will change for different potentials. The study of
the dependence of the interactions (scattering length) on the dimensionality and
the degree of confinement of the system (Olshanii, 1998; G6rlitz et al., 2001;
Petrov et al., 2000) will benefit from the variety of potentials available on the
atom chip.

B. 4. Non-linear phenomena

Another example of an interesting regime for the study of atom-atom interaction


or non-linearity are multi-well potentials. Again, as mentioned in the context
348 R. Folman et al. [VI

of interferometers, the splitting of a cloud of atoms into these multi-sites can


be either temporal or spatial. Here calculations beyond mean field theory are
relevant, and new insight may be acquired. For example, one expects a crossover
from coherent splitting to number splitting in different potential configurations,
depending on the height of the potential barrier, the density, and the scattering
length (Menotti et al., 2001; Vardi and Anglin, 2001; Orzel et al., 2001). This
phase transition has already been observed experimentally by Greiner et al.
(2002).

B.5. Boundary between macroscopic and microscopic description

Let us end this subsection concerning mesoscopics by noting that the ability
to change the number of atoms in a system, or alternatively to address specific
atoms in an interacting ensemble, will allow us to probe the boundary between
the macroscopic and microscopic description. Starting from a large system,
we will try to gain more and more control over the system parameters,
imprinting quantum behavior onto the system. On the other hand we can
try and build larger and larger systems from single quantum objects (called
qubits in modern lingo), and keep individual control over the parameters.
Success in such an undertaking would bring us much closer to implementing
quantum information transfer and quantum information processing as discussed
below.

C. QUANTUMINFORMATION

The implementation of quantum information processing requires (DiVincenzo,


2000): (i) storage of the quantum information in a set of two-level systems
(qubits), (ii) the processing of this information using quantum gates, and
(iii) reading out the results. For a review of quantum computation we refer the
reader to Bouwmeester et al. (2000).
We believe that quantum optical schemes where the qubit is encoded in
neutral atoms can be implemented using atom optics on integrated atom
chips (Schmiedmayer et al., 2002). These promise to combine the outstanding
features of quantum optical proposals, in particular quantum control and
long decoherence times, with the technological capabilities of engineering
microstructures implying scalability, a feature usually associated with solid-state
proposals. Let us review some of the requirements:
- The qubit. Using neutral atoms, the qubit can be encoded in two internal, long-
lived states (e.g. two different hyperfine electronic ground states). Single-qubit
operations are induced as transitions between the hyperfine states of the atoms.
These are introduced by external fields, using RF pulses like in NMR or in
Ramsey-Bord6 interferometers, Raman transitions or adiabatic passage.
VI] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 349

One method to realize a qubit is to write the qubits into single atoms, which
requires selective cooling and filling of atoms into the qubit sites. However,
recently it was proposed that single qubits can also be written into an ensemble
of atoms using 'dipole blockade' (Lukin et al., 2001). This may be simpler
as it avoids the need for single-atom loading of traps. As will be pointed out
below, the dipole blockade mechanism can also be used to manipulate the
qubit.
- E n t a n g l i n g qubits. The fundamental two-qubit quantum gate requires state-
selective interaction between two qubits, which is more delicate to implement.
A two-qubit quantum gate is a state-dependent operation such as a control
NOT gate:

Io) 1o) Io) 1o),


Io)1 ) FO)ll),
11>1o> I1>11>,
I1)[1) ~ I1)1o). (58)
A good way to implement such a quantum gate is by state-selective
interactions, which can be switched on and off at will. This interaction can be
between the qubits themselves, or mediated by a 'bus'. Neutral atoms naturally
interact with each other. To achieve different phase shifts for different qubit
states, either the interaction between the qubits has to be state selective, or it
has to be turned on conditioned on the qubit state. There are different ways
to implement quantum gates in atom optics: depending on the interaction, we
distinguish between (a) the generic interactions between the atoms, like the
Van der Waals interaction (Jaksch et al., 1999; Calarco et al., 2000; Briegel
et al., 2000) and (b) interactions which can be switched on and off, like
induced electric dipole-dipole interactions (Brennen et al., 1999; Brennen and
Deutsch, 2000), including highly excited Rydberg states (Jaksch et al., 2000;
Lukin et al., 2001).
- Dipole-blockade quantum gates between m e s o s c o p i c atom ensembles. Lukin
et al. (2001) devised a technique for the coherent manipulation of quan-
tum information stored in collective excitations of many-atom mesoscopic
ensembles by optically exciting the ensemble into states with a strong atom-
atom interaction. Under certain conditions the level shifts associated with
these interactions can be used to block the transitions into states with more
than a single excitation. The resulting dipole-blockade phenomenon closely
resembles similar mesoscopic effects in nanoscale solid-state devices. It
can take place in an ensemble with a size that can exceed many optical
wavelengths and can be used to perform quantum gate operations between
distant ensembles, each acting as a single qubit.
- Cavity QED. The 2-qubit processing operation may be realized through a
direct interaction (entanglement) between two atoms or through an inter-
350 R. F o l m a n et al. [VI

FIG. 49. A possible implementation of a neutral atom qubit processor on an atom chip which
includes a reservoir of cold atoms in a well-defined state (for example a BEC or a degenerate Fermi
gas). From there the atoms can be transported using guides or moving potentials to the processing
sites. Either single atoms or small ensembles of atoms are then loaded into the qubit traps. Each
qubit can be addressed either by bringing light to each individual site separately, or by illuminating
the whole processor and addressing the single qubits by shifting them in and out of resonance
using local electric or magnetic fields created by the nanostructures on the atom chip. We think that
electric fields are preferable, since magnetic fields might produce qubit-state dependent phase shifts,
which have to be corrected. A different method would also be to address the single sites using field
gradients like in NMR.

mediate 'bus'. A light mode of a high-Q cavity can serve as such a 'bus'
acting on an array of atoms trapped inside the cavity (Pellizzari et al., 1995).
Atoms in high-Q cavities which in turn are connected with fibers, can also
act as a converting device between 'flying' qubits (photons) which transverse
distances, and storage qubits (atoms). The same principle can be used for
entangling atoms in different cavities for a 'distributed' computation process
(van Enk et al., 1998, 1999). In all of the above, the atom chip promises to
enhance the feasibility of accurate a t o m - c a v i t y systems.
- Input~Output. Even without high-Q cavities, an integrated atom chip, with
atoms trapped in well-controlled microtraps and with individual site light
elements, can probably provide input/output processes by making use o f
techniques such as light scattering from trapped atomic ensembles (Duan
et al., 2001), slow light (Hau et al., 1999; Vitali et al., 2000), stopped light
(Phillips et al., 2001; Liu et al., 2001; Fleischhauer and Lukin, 2002) or
macroscopic spin states (Duan et al., 2000; Julsgaard et al., 2001).
Let us summarize the road map for quantum computation with the atom chip:
one would need to implement
(a) versatile traps to accurately control atoms up to the stage of entanglement;
VIII] MICROSCOPIC ATOM OPTICS: FROM WIRES TO ATOM CHIP 351

(b) controlled loading of single qubits (atoms or excitations) into these traps in
well-defined internal and external states;
(c) manipulation and detection of individual qubits;
(d) control over decoherence; and
(e) scalability to be able to achieve controlled quantum manipulation of a large
number of qubits.
A schematic view of a possible realization is shown in Fig. 49.

VII. Conclusion

Neutral-atom manipulation using integrated microdevices is a new and extremely


promising experimental approach. It combines the best of two worlds: the ability
to use cold a t o m s - a well-controllable quantum system, and the immense
technological capabilities of nanofabrication, micro-optics and microelectronics
to manipulate and detect the atoms.
In the future, a final integrated atom chip will have a reliable source
of cold atoms with an efficient loading mechanism, single-mode guides for
coherent transportation of atoms, nanoscale traps, movable potentials allowing
controlled collisions for the creation of entanglement between atoms, high-
resolution light fields for the manipulation of individual atoms, and internal
state-sensitive detection of atoms. All of these, including the bias fields and
possibly even the light sources and the read-out electronics, could be on-board a
self-contained chip. Such a robust and easy to use device would make possible
advances in many different fields of quantum physics: from applications such
as clocks, sensors and implementations of quantum information processing and
communication, to new experimental insight into fundamental questions relating
to decoherence, disorder, non-linearity, entanglement, and atom scattering in low-
dimensional physics.

VIII. Acknowledgement

Foremost we would like to thank all the members of the Innsbruck, now
Heidelberg, atom chip group for their enthusiasm and the enormous effort they
put into the experiments. We would like to thank our long-time theoretical
collaborators Peter Zoller, Tommaso Calarco and Robin C6t6. The atom chips
for the Innsbruck-Heidelberg experiments were fabricated by Thomas Maier at
the Institut ft~r Festk6rperelektronik, Technische Universit~t Wien, Austria, and
by Israel Bar-Joseph at the Sub-micron center, Weizmann Institute of Science,
Israel. We would also like to extend a warm thanks to the entire atom chip
community for responding so positively to our requests for information and
figures. Our work was supported by many sources, most notably the Austrian
352 R. Folman et al. [IX

Science Foundation (FWF), projects S065-05 and SFB F15-07, the Deutsche
Forschungsgemeinschaft Schwerpunktprogramme: 'Quanten Informationsverar-
beitung' and 'Wechselwirkungen in ultrakalten Atom- und Molekfilgasen', and
the European Union, contract numbers IST-1999-11055 (ACQUIRE), HPRI-CT-
1999-00069 (LSF), TMRX-CT96-0002, and HPMF-CT- 1999-00235.

IX. R e f e r e n c e s

Aharonov, Y., and Stern, A. (1992). Phys. Rev. Lett. 69, 3593.
Andersson, E., and Barnett, S.M. (2000). Phys. Rev. A 62, 052311.
Andersson, E., Fontenelle, M.T., and Stenholm, S. (1999). Phys. Rev. A 59, 3841.
Andersson, E., Calarco, T., Folman, R., Andersson, M., Hessmo, B., and Schmiedmayer, J. (2002).
Phys. Rev. Lett. 88, 100401.
Bagnato, V.S., Lafyatis, G.P., Martin, A.G., Raab, E.L., Ahmad-Bitar, R.N., and Pritchard, D.E.
(1987). Phys. Rev. Lett. 58, 2194.
Bali, S., O'Hara, K.M., Gehm, M.E., Granade, S.R., and Thomas, J.E. (1999). Phys. Rev. A 60, R29.
Barnett, A.H., Smith, S.P., Olshanii, M., Johnson, K.S., Adams, A.W., and Prentiss, M. (2000). Phys.
Rev. A 61, 023608.
Batelaan, H., Abfalterer, R., Wehinger, S., and Schmiedmayer, J. (1994). In "Technical Digest,
EQEC V, Amsterdam," p. 13.
Berg-Sorensen, K., Burns, M., Golovchenko, J., and Hau, L. (1996). Phys. Rev. A 53, 1653.
Bergeman, T.H., McNicholl, P., Kycia, J., Metcalf, H., and Balazs, N. (1989). J Opt. Soc. B 6, 2249.
Berman, P., ed. (1994). "Cavity Quantum Electrodynamics," Supplement 2 of Advances in Atomic,
Molecular, and Optical Physics. Academic Press, New York.
Birkl, G., Buchkremer, F.B.J., Dumke, R., and Ertmer, W. (2001). Opt. Commun. 191, 67.
Blfimel, R., and Dietrich, K. (1989). Phys. Lett. A 139, 236.
Blfimel, R., and Dietrich, K. (1991). Phys. Rev. A 43, 22.
Bouwmeester, D., Ekert, A., and Zeilinger, A., eds. (2000). "The Physics of Quantum Information."
Springer, Berlin.
Brennen, G.K., and Deutsch, I.H. (2000). Phys. Rev. A 61, 062309.
Brennen, G.K., Caves, C.M., Jessen, P.S., and Deutsch, I.H. (1999). Phys. Rev. Lett. 82, 1060.
Briegel, H.-J., Calarco, T., Jaksch, D., Cirac, J.I., and Zoller, P. (2000). J. Mod. Opt. 47, 415.
Brugger, K., Calarco, T., Cassettari, D., Folman, R., Haase, A., Hessmo, B., Krfiger, P., Maier, T.,
and Schmiedmayer, J. (2000). J. Mod. Opt. 47, 2789.
Buks, E., Schuster, R., Heiblum, M., Mahalu, D., and Umansky, V. (1998). Nature 391,871.
Burke, J.P., Bohn, J.L., Esry, B.D., and Greene, C.H. (1997). Phys. Rev. A 55, R2511.
Burke Jr, J.P., Greene, C.H., and Esry, B.D. (1996). Phys. Rev. A 54, 3225.
Burt, E.A., Ghrist, R.W., Myatt, C.J., Holland, M.J., Cornell, E.A., and Wieman, C.E. (1997). Phys.
Rev. Lett. "/9, 337.
Calarco, T., Hinds, E.A., Jaksch, D., Schmiedmayer, J., Cirac, J.I., and Zoller, P. (2000). Phys. Rev.
A 61, 022304.
Cassettari, D., Chenet, A., Denschlag, J., Schneider, S., and Schmiedmayer, J. (1998). In "Technical
Digest, EQEC 98, Glasgow," September.
Cassettari, D., Chenet, A., Denschlag, J., Folman, R., Hessmo, B., Haase, A., Krfiger, P., Schneider, S.,
and Schmiedmayer, J. (1999). In "Towards Mesoscopic Physics with Cold Atoms," XIVth
International Conference on Laser Spectroscopy, June 7-11 (R. Blatt, J. Eschner, D. Leibfried and
F. Schmidt-Kaler, Eds.). World Scientific, Singapore, p. 324.
Cassettari, D., Hessmo, B., Folman, R., Maier, T., and Schmiedmayer, J. (2000). Phys. Rev. Lett.
84, 1124.
IX] MICROSCOPIC ATOM OPTICS" FROM WIRES TO ATOM CHIP 353

Cheng, C.-C., and Raymer, M.G. (1999). Phys. Rev. Lett. 82, 4807.
Cook, R.J., and Hill, R.K. (1982). Opt. Commun. 43, 258.
C6t6, R., Dalgarno, A., and Jamieson, M.J. (1994). Phys. Rev. A 50, 399.
Davis, T.J. (1999). J. Opt. B 1,408.
Dekker, N.H., Lee, C.S., Lorent, V., Thywissen, J.H., Smith, S.P., Drndi6, M., Westervelt, R.M., and
Prentiss, M. (2000). Phys. Rev. Lett. 84, 1124.
Denschlag, J. (1998). Ph.D. Thesis. Universit~it Innsbruck.
Denschlag, J., and Schmiedmayer, J. (1997). Europhys. Lett. 6, 405.
Denschlag, J., Umshaus, G., and Schmiedmayer, J. (1998). Phys. Rev. Lett. 81,737.
Denschlag, J., Cassettari, D., Chenet, A., Schneider, S., and Schmiedmayer, J. (1999a). Appl. Phys.
B 69, 291.
Denschlag, J., Cassettari, D., and Schmiedmayer, J. (1999b). Phys. Rev. Lett. 82, 2014.
DiVincenzo, D.E (2000). Fortschr. Phys. 48, 771.
Domokos, P., Gangl, M., and Ritsch, H. (2000). Opt. Commun. 185, 115.
Dowling, J.E, and Gea-Banacloche, J. (1996). Adv. At. Mol. Opt. Phys. 37, 1.
Drndi6, M., Johnson, K.S., Thywissen, J.H., Prentiss, M., and Westervelt, R.M. (1998). Appl. Phys.
Lett. 72, 2906.
Duan, L.-M., Cirac, J.I., Zoller, P., and Polzik, E.S. (2000). Phys. Rev. Lett. 85, 5643.
Duan, L.-M., Lukin, M.D., Cirac, J.I., and Zoller, P. (2001). Nature 414, 413.
Ekstrom, C.R., Keith, D.W., and Pritchard, D.E. (1992). Appl. Phys. B 54, 369.
Esry, B.D., Greene, C.H., and Burke Jr, J.E (1999). Phys. Rev. Lett. 83, 1751.
Fedichev, P.O., Reynolds, M.W., and Shlyapnikov, G.V. (1996). Phys. Rev. Lett. 77, 2921.
Fleischhauer, M., and Lukin, M.D. (2002). Phys. Rev. A 65, 022314.
Folman, R., Krfiger, P., Cassettari, D., Hessmo, B., Maier, T., and Schmiedmayer, J. (2000). Phys.
Rev. Lett. 84, 4749.
Fortagh, J., Grossmann, A., and Zimmermann, C. (1998). Phys. Rev. Lett. 81, 5310.
Fortagh, J., Ott, H., Grossmann, A., and Zimmermann, C. (2000). Appl. Phys. B 70, 701.
Fortagh, J., Ott, H., Kraft, S., and Zimmermann, C. (2002). cond-mat/0205310.
Frisch, R., and Segr~, E. (1933). Z. Phys. 75, 610.
Fujita, J., Mitake, S., and Shimizu, E (2000). Phys. Rev. Lett. 84, 4027.
Gauck, H., Hartl, M., Schneble, D., Schnitzler, H., Pfau, T., and Mlynek, J. (1998). Phys. Rev. Lett.
81, 5298.
Gehm, M.E., O'Hara, K.M., Savard, T.A., and Thomas, J.E. (1998). Phys. Rev. A 58, 3914.
Gerton, J.M., Sackett, C.A., Frew, B.J., and Hulet, R.G. (1999). Phys. Rev. A 59, 1514.
Giulini, D., Joos, E., Kiefer, C., Kupsch, J., Stamatescu, I.-O., and Zeh, H.D. (1996). "Decoherence
and the Appearance of a Classical World in Quantum Theory." Springer, Berlin.
G6rlitz, A., Vogels, J.M., Leanhardt, A.E., Raman, C., Gustavson, T.L., Abo-Shaeer, J.R.,
Chikkatur, A.P., Gupta, S., Inouye, S., Rosenband, T.P., and Ketterle, W. (2001). Phys. Rev.
Lett. 87, 130402.
Gott, Y.V., Ioffe, M.S., and Tel'kovskii, V.G. (1962). Nucl. Fusion Suppl. 3, 1045.
Gov, S., Shtrikman, S., and Thomas, H. (2000). J. Appl. Phys. 87, 3989-3998.
Greiner, M., Mandel, O., Esslinger, T., H~insch, T.W., and Bloch, I. (2002). Nature 415, 39.
Grimm, R., Weidemfiller, M., and Ovchinnikov, Y.B. (2000). Adv. At. Mol. Opt. Phys. 42, 95.
Gustavson, T.L., Chikkatur, A.E, Leanhardt, A.E., G6rlitz, A., Gupta, S., Pritchard, D.E., and
Ketterle, W. (2002). Phys. Rev. Lett. 88, 020401.
Guth6hrlein, G.R., Keller, M., Hayasaka, K., Lange, W., and Walther, H. (2001). Nature 414, 49.
Haase, A. (2000). Diploma Thesis. University of Innsbruck.
Haase, A., Cassettari, D., Hessmo, B., and Schmiedmayer, J. (2001). Phys. Rev. A 64, 043405.
H~insel, W., Hommelhoff, P., H~insch, T.W., and Reichel, J. (2001a). Nature 413, 498.
H~insel, W., Reichel, J., Hommelhoff, P., and H~nsch, T.W. (2001b). Phys. Rev. Lett. 86, 608.
H~insel, W., Reichel, J., Hommelhoff, P., and H~insch, T.W. (2001c). Phys. Rev. A 64, 063607.
354 R. F o l m a n et al. [IX

Hau, L., Burns, M., and Golovchenko, J. (1992). Phys. Rev. A 45, 6468.
Hau, L., Golovchenko, J., and Burns, M. (1995). Phys. Rev. Lett. 74, 3138.
Hau, L.V., Harris, S.E., Dutton, Z., and Behroozi, C.H. (1999). Nature 397, 594.
Henkel, C., and Gardiner, S.A. (2002). Spatial decoherence of a trapped Bose condensate, in
preparation.
Henkel, C., and P6tting, S. (2001). Appl. Phys. B 72, 73.
Henkel, C., and Scheel, S. (2002). in preparation.
Henkel, C., and Wilkens, M. (1999). Europhys. Lett. 47, 414.
Henkel, C., P6tting, S., and Wilkens, M. (1999). Appl. Phys. B 69, 379.
Henkel, C., Joulain, K., Carminati, R., and Greffet, J.-J. (2000). Opt. Commun. 186, 57.
Henkel, C., Krfiger, P., Folman, R., and Schmiedmayer, J. (2002). quant-ph/0208165.
Hinds, E.A., and Eberlein, C. (2000). Phys. Rev. A 61, 033614. Erratum: 2001, 64, 039902.
Hinds, E.A., and Hughes, I.G. (1999). J. Phys. D 32, R119.
Hinds, E.A., Vale, C.J., and Boshier, M.G. (2001). Phys. Rev. Lett. 86, 1462.
Hood, C.J., Lynn, T.W., Doherty, A.C., Parkins, A.S., and Kimble, H.J. (2000). Science 287, 1447.
Imry, Y. (1987). "Introduction to Mesoscopic Physics." Oxford University Press, Oxford.
Jaksch, D., Briegel, H.-J., Cirac, J.I., Gardiner, C.W., and Zoller, P. (1999). Phys. Rev. Lett. 82, 1975.
Jaksch, D., Cirac, J.I., Zoller, P., Rolston, S.L., C6t6, R., and Lukin, M.D. (2000). Phys. Rev. Lett.
85, 2208.
Julienne, P.S., Mies, EH., Tiesinga, E., and Williams, C.J. (1997). Phys. Rev. Lett. 78, 1880.
Julsgaard, B., Kozhekin, A., and Polzik, E.S. (2001). Nature 413, 400.
Ketterle, W. (1999). Phys. Today 30(December).
Ketterle, W., and Pritchard, D. (1992). Appl. Phys. B 54, 403.
Key, M., Hughes, I.G., Rooijakkers, W., Sauer, B.E., Hinds, E.A., Richardson, D.J., and Kazansky, P.G.
(2000). Phys. Rev. Lett. 84, 1371.
Kokkelmans, S.J.J.M.E, Boesten, H.M.J.M., and Verhaar, B.J. (1997). Phys. Rev. A 55, R1589.
Landau, L.D., and Lifshitz, E.M. (1976). "Mechanics." Pergamon Press, Oxford.
Leanhardt, A.E., Chikkatur, A.P., Kielpinski, D., Shin, Y., Gustavson, T.L., Ketterle, W., and
Pritchard, D.E. (2002). Phys. Rev. Lett. 89, 040401.
Lee, K.I., Kim, J.A., Noh, H.R., and Jhe, W. (1996). Opt. Lett. 21, 1177.
Littlejohn, R.G., and Weigert, S. (1993). Phys. Rev. A 48, 924.
Liu, C., Dutton, Z., Behroozi, C.H., and Hau, L.V. (2001). Nature 409, 490.
Luiten, O.J., Reynolds, M.W., and Walraven, J.T.M. (1996). Phys. Rev. A 53, 381.
Lukin, M., Fleischhauer, M., C6t6, R., Duan, L.M., Jaksch, D., Cirac, J.I., and Zoller, P. (2001).
Phys. Rev. Lett. 87, 037901.
Mabuchi, H., Armen, M., Lev, B., Loncar, M., Vuckovic, J., Kimble, H.J., Preskill, J., Roukes, M.,
Scherer, A., and van Enk, S.J. (2001). Quantum lnf Comput. 1, 7.
Majorana, E. (1932). Nuovo Cimento 9, 43.
Menotti, C., Anglin, J.R., Cirac, J.I., and Zoller, P. (2001). Phys. Rev. A 63, 023601.
Meschede, D., Bloch, I., Goepfert, A., Haubrich, D., Kreis, M., Lison, E, Schfitze, R., and Wynands, R.
(1997). SPIE Proc. 2995, 191.
Moerdijk, A.J., and Verhaar, B.J. (1996). Phys. Rev. A 53, R19.
Moerdijk, A.J., Boesten, H.M.J.M., and Verhaar, B.J. (1996). Phys. Rev. A 53, 916.
Miiller, D., Anderson, D.Z., Grow, R.J., Schwindt, P.D.D., and Cornell, E.A. (1999). Phys. Rev. Lett.
83, 5194.
Mfiller, D., Cornell, E.A., Prevedelli, M., Schwindt, P.D.D., Zozulya, A., and Anderson, D.Z. (2000).
Opt. Lett. 25, 1382.
Miiller, D., Cornell, E.A., Prevedelli, M., Schwindt, P.D.D., Wang, Y., and Anderson, D.Z. (2001).
Phys. Rev. A 63, 041602(R).
Myatt, C.J., Burt, E.A., Ghrist, R.W., Cornell, E.A., and Wieman, C.E. (1997). Phys. Rev. Lett.
78, 586.
IX] MICROSCOPIC A T O M O P T I C S " F R O M W I R E S TO A T O M C H I P 355

Nowak, S., Stuhler, N., Pfau, T., and Mlynek, J. (1998). Phys. Rev. Lett. 81, 5792.
O'Dell, D., Giovanazzi, S., Kurizki, G., and Akulin, V.M. (2000). Phys. Rev. Lett. 84, 5687.
Olshanii, M. (1998). Phys. Rev. Lett. 81,938
Opat, G.I., Wark, S.J., and Cimmino, A. (1992). Appl. Phys. B 54, 396.
Orzel, C., Tuchman, A.K., Fenselau, M.L., Yasuda, M., and Kasevich, M.A. (2001). Science
291, 2386.
Osnaghi, S., Bertet, P., Auffeves, A., Maioli, P., Brune, M., Raimond, J.M., and Haroche, S. (2001).
Phys. Rev. Lett. 87, 037902.
Ott, H., Fortagh, J., Schlotterbeck, G., Grossmann, A., and Zimmermann, C. (2001). Phys. Rev. Lett.
87, 230401.
Palm, T., and Thyl6n, L. (1992). Appl. Phys. Lett. 60, 237.
Paul, W. (1990). Rev. Mod. Phys. 62, 531.
Paz, J.P., Habib, S., and Zurek, W.H. (1993). Phys. Rev. D 47, 488.
Pellizzari, T., Gardiner, S.A., and Zoller, P. (1995). Phys. Rev. Lett. 75, 3788.
Petrov, D.S., Shlyapnikov, G.V., and Walraven, J.T.M. (2000). Phys. Rev. Lett. 85, 3745.
Pfau, T. (2001). private communication.
Pfau, T., and Mlynek, J. (1996). In "Ultracold Atoms and Bose-Einstein-Condensation," Proceedings
of the European Quantum Electronics Conference, September 1996, Hamburg, Germany (K.
Burnett, Ed.). Optical Society of America, Washington, DC, p. 33.
Pfau, T., Gauck, H., Schneble, D., Hartl, M., and Mlynek, J. (1997). In "Quantum Electronics Conf.,"
Vol. 17. Optical Society of America, Washington, DC.
Phillips, D.E, Fleischhauer, A., Mair, A., Walsworth, R.L., and Lukin, M.D. (2001). Phys. Rev. Lett.
86, 783.
Pinkse, P.W.H., Fischer, T., Maunz, P., and Rempe, G. (2000). Nature 404, 365.
Pritchard, D. (1983). Phys. Rev. Lett. 51, 1336.
Pron'kov, G.P., and Stroganov, Yu.G. (1977). Soy. Phys. JETP 45, 1075.
Reichel, J. (2002). Appl. Phys. B 74, 469.
Reichel, J., H/insel, W., and H~nsch, T.W. (1999). Phys. Rev. Lett. 83, 3398.
Reichel, J., H/insel, W., Hommelhoff, R, and H~nsch, T.W. (2001). Appl. Phys. B 72, 81.
Richmond, J.A., NicChormaic, S., Cantwell, B.E, and Opat, G.I. (1998). Acta Phys. Slov. 48, 481.
Roach, T.M., Abele, H., Boshier, M.G., Grossman, H.L., Zetie, K.R, and Hinds, E.A. (1995). Phys.
Rev. Lett. 75, 629.
Rooijakkers, W., Vengalatorre, M., and Prentiss, M. (2001 ). In "ICOLS proceedings." Optical Society
of America, Washington, DC.
Rowlands, W.J., Lau, D.C., Opat, G.I., Sidorov, A.I., McLean, R.J., and Hannaford, E (1995). In
"Laser Spectroscopy, XIIth International Conference, Capri, Italy" (M. Inguscio, M. Allegrini,
A. Sasso and L. Capri, Eds.). p. 134.
Rowlands, W.J., Lau, D.C., Opat, G.I., Sidorov, A.I., McLean, R.J., and Hannaford, E (1996a). Opt.
Commun. 126, 55.
Rowlands, W.J., Lau, D.C., Opat, G.I., Sidorov, A.I., McLean, R.J., and Hannaford, E (1996b). Aust.
J. Phys. 49, 577.
Saba, C.V., Barton, P.A., Boshier, M.G., Hughes, I.G., Rosenbusch, E, Sauer, B.E., and Hinds, E.A.
(1999). Phys. Rev. Lett. 82, 468.
Sauer, J.A., Barrett, M.D., and Chapman, M.S. (2001). Phys. Rev. Lett. 87, 270401-1.
Schmiedmayer, J. (1992). In "Technical Digest, IQEC 92: XVIIIth International Conference on
Quantum Electronics" (G. Magerl, Ed.). Technische Universit~t Wien, Vienna, p. 284.
Schmiedmayer, J. (1995a). Appl. Phys. B 60, 169.
Schmiedmayer, J. (1995b). Phys. Rev. A 52, R13.
Schmiedmayer, J. (1998). Eur. Phys. J. B 4, 57.
Schmiedmayer, J., and Scrinzi, A. (1996a). Phys. Rev. A 54, R2525.
Schmiedmayer, J., and Scrinzi, A. (1996b). Quantum Semiclass. Opt. 8, 693.
356 R. F o l m a n et al. [IX

Schmiedmayer, J., Folman, R., and Carlarco, T. (2002). J. Mod. Opt. 49, 1375.
Schneble, D., Gauck, H., Hartl, M., Pfau, T., and Mlynek, J. (1999). In "Bose-Einstein Condensation
in Atomic Gases," Proceedings of the International School of Physics 'Enrico Fermi,' Varenna,
1998 (M. Inguscio, S. Stringari and C. Wieman, Eds.). lOS Press, Amsterdam, p. 469.
Schneble, D., Hasuo, M., Anker, Th., Pfau, T., and Mlynek, J. (2001). In "Technical Digest,
Postconference Edition, QELS 2001." Optical Society of America, Washington, DC.
Sekatskii, S.K., and Schmiedmayer, J. (1996). Europhys. Lett. 36, 407.
Sekatskii, S.K., Riedo, B., and Dietler, G. (2001). Opt. Commun. 195, 197.
Shapere, A., and Wilczek, E, eds. (1989). "Geometric Phases in Physics." World Scientific,
Singapore.
Shimizu, E, and Morinaga, M. (1992). Jpn. J. Appl. Phys. 31, L1721.
Shimizu, E, Shimizu, K., and Takuma, H. (1992). Jpn. J. Appl. Phys 31, L436.
Sidorov, A.I., McLean, R.J., Rowlands, W.J., Lau, D.C., Murphy, J.E., Walkiewicz, M., Opat, G.I.,
and Hannaford, P. (1996). Quantum Semiclass. Opt. 8, 713.
S6ding, J., Gu6ry-Odelin, D., Desbiolles, P., Ferrari, G., and Dalibard, J. (1998). Phys. Rev. Lett.
80, 1869.
S6ding, J., Gu6ry-Odelin, D., Desbiolles, P., Chevy, E, Inamori, H., and Dalibard, J. (1999). Appl.
Phys. B 69, 257.
Spreeuw, R.J.C., Voigt, D., Wolschrijn, B.T., and van Linden van den Heuvell, H.B. (2000). Phys.
Rev. A 61, 053604.
Stern, A. (1992). Phys. Rev. Lett. 68, 1022.
Stern, A., Aharonov, Y., and Imry, Y. (1990). Phys. Rev. A 41, 3436.
Sukumar, C.V., and Brink, D.M. (1997). Phys. Rev. A 56, 2451.
Thywissen, J.H., Olshanii, M., Zabow, G., Drndi6, M., Johnson, K.S., Westervelt, R.M., and
Prentiss, M. (1999a). Eur. Phys. J. D 7, 361.
Thywissen, J.H., Westervelt, R.M., and Prentiss, M. (1999b). Phys. Rev. Lett. 83, 3762.
Timmermans, E., and C6t6, R. (1998). Phys. Rev. Lett. 80, 3419.
Turchette, Q.A., Myatt, C.J., King, B.E., Sackett, C.A., Kielpinski, D., ltano, W.M., Monroe, C., and
Wineland, D.J. (2000). Phys. Rev. A 62, 053807.
van Enk, S.J., Cirac, J.I., and Zoller, P. (1998). Science 279, 205.
van Enk, S.J., Kimble, H.J., Cirac, J.l., and Zoller, P. (1999). Phys. Rev. A 59, 2659.
Vardi, A., and Anglin, J.R. (2001). Phys. Rev. Lett. 86, 568.
Varpula, T., and Poutanen, T. (1984). J. Appl. Phys. 55, 4015.
Vengalattore, M., Rooijakkers, W., and Prentiss, M. (2001). physics/0106028.
Vitali, D., Fortunato, M., and Tombesi, P. (2000). Phys. Rev. Lett. 85(2), 445.
Vladimirskii, V.V. (1961). Soy. Phys. JETP 12, 740.
Voronin, A.I. (1991). Phys. Rev. A 43, 29.
Vuletic, V., H~nsch, T.W., and Zimmermann, C. (1996). Europhys. Lett. 36, 349.
Vuletic, V., Fischer, T., Praeger, M., H/insch, T.W., and Zimmermann, C. (1998). Phys. Rev. Lett.
80, 1634.
Weiner, J., Bagnato, V.S., Zilio, S., and Julienne, P.S. (1999). Rev. Mod. Phys. 71, 1.
Weinstein, J.D., and Libbrecht, K.G. (1995). Phys. Rev. A 52, 4004.
Wesstr6m, J.-O.J. (1999). Phys. Rev. Lett. 82, 2564.
Wing, W. (1984). Prog. Quantum Electron. 8, 181.
Zeh, H.D. (1970). Found. Phys. 1, 69.
Zokay, O., and Garraway, B.M. (2000). Opt. Commun. 93, 93.
Zurek, W.H. (1991). Phys. Today 44(October), 36.
Zurek, WH., Habib, S., and Paz, J.P. (1993). Phys. Rev. Lett. 70, 1187.
A D V A N C E S IN A T O M I C , M O L E C U L A R , A N D O P T I C A L P H Y S I C S , V O L . 48

ME THOD S OF MEA S URING


ELE C TR ON-A TOM COLLISION
CROSS SECTIONS WITH A N ATOM TRAP
R.S. SCHAPPE
Department of Physics, Lake Forest College, Lake Forest, Illinois 60045

M.L. KEELER
Department of Physics, Wesleyan University, Middletown, Connecticut 06459

TODD A. Z I M M E R M A N and M. L A R S E N
Department of Physics, University of Wisconsin, Madison, Wisconsin 53706

PAUL FENG
Department of Physics, University of St. Thomas, St. Paul, Minnesota 55105

RENEE C. NESNIDAL
New Focus, Inc., Middleton, Wisconsin 53562

J O H N B. BOFFARD, THAD G. WALKER, L. W A N D E R S O N and C H U N C. LIN


Department of Physics, University of Wisconsin, Madison, Wisconsin 53 706

I. Introduction ................................................ 357


II. General E x p e r i m e n t O v e r v i e w . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 359
A. T r a p p e d A t o m s as a Target ................................... 359
B. A p p a r a t u s ............................................... 362
III. M e t h o d s for M e a s u r i n g Cross Sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367
A. Loss Rate M e a s u r e m e n t s ..................................... 368
B. E x p e r i m e n t s B a s e d on Direct D e t e c t i o n o f Final State ................. 376
IV. C o n c l u s i o n s ................................................ 386
V. A c k n o w l e d g m e n t s ............................................ 387
VI. A p p e n d i x . N u m e r i c a l M o d e l for R e s i d u a l Polarization ................... 387
VII. R e f e r e n c e s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389

I. I n t r o d u c t i o n

Measurements of electron-atom collision cross sections are a subject of great


interest. Until recently, a typical experiment consisted of an electron beam
incident on target atoms in the form of either an atomic beam or a static gas

357 Copyright 9 2002 Elsevier Science (USA)


All rights reserved
ISBN 0-12-003848-X/ISSN 1049-250X/02 $35.00
358 Schappe et al. [I

target. In this chapter we describe a new type of atomic target: ultra-cold trapped
atoms (Schappe et al., 1995, 1996; Keeler et al., 2000).
An atom trap has three features that make it desirable for use as a target in
scattering experiments. First, the atoms in a trap can be confined to a very small
volume (diameter < 1 mm). Hence, an atom-trap target can be treated practically
as a point target of atoms. This property greatly simplifies the calculation of the
overlap of the target with the electron beam; absolute calibration can be done
without undertaking the difficult task of determining the profiles of both the
electron beam and atomic beam and then finding their convolution as must be
done in a crossed beam experiment (Mark and Dunn, 1985). Second, atoms in
a typical optical trap are cold, with temperatures on the order of hundreds of
gK or atomic speeds on the order of ten cm/s. Thus the atoms in a trap are for
all practical purposes at rest. Consequently, an atom that has undergone even
a weak electron-atom collision can pick up a sufficient atomic recoil velocity
to differentiate it from atoms that have not undergone a collision. Furthermore,
for the investigation of excited states, the linewidth of an atomic transition for
the trapped atoms is determined by the natural linewidth of the transition, and
not a Doppler profile. This vastly reduces the laser intensity needed to saturate
a transition. Third, atoms in the trap are constantly scattering photons from
the trapping laser beams and thus the trap fluorescence is directly proportional
to the number of atoms in the target. Because this signal is large ( - g W ) it
can be measured easily with excellent time resolution using even a simple
photodiode.
The high rate of photon scattering, Doppler-free transition width, and small
target size can be combined to yield a target with a large laser-excited state
fraction using only low-power lasers. For example, to study electron-impact
excitation out of the Na(3P) level with a conventional crossed beam target,
Stumpf and Gallagher (1985) needed -100 mW of laser power to create an
excited state fraction of~20%, while an atom-trap target can yield ~40% excited
state fraction with only ~10 mW of laser power (Keeler et al., 2000).
Collisions between cold atoms in a trap have been extensively studied since
the start of atom trapping (Walker and Feng, 1994; Weiner et al., 1999). To
our knowledge, however, the first application where all of the advantages of
using a trap as a separate target for measuring cross sections was the work of
Dinneen et al. (1992) on photoionization of Rb(5P). In their work, trapped atoms
were ionized (and thus lost from the trap) by a Kr-ion laser. Atom traps have
subsequently been used to study photoionization of the Rb(5D) level (Duncan
et al., 2001), and of the Li(2P) (Wippel et al., 2001), Na(3P) (Wippel et al.,
2001), and Cs(6P) (Marag6 et al., 1998) levels. Atom traps have also been used
in the study of heavy ion collisions (Flechard et al., 2001; Turkstra et al., 2001;
van der Poel et al., 2001).
In this chapter we describe the techniques for using trapped atoms for
II] ELECTRON-ATOM COLLISION CROSS SECTIONS 359

measuring a wide variety of electron-atom cross sections including total,


excitation, and ionization cross sections, from both the ground state and a laser
excited state of trapped atoms. In Sect. II we briefly describe the technique of
atom trapping in the context of electron-atom collision studies. In Sect. III we
discuss two major ways to use atom traps to measure cross sections, the trap-loss
technique and the direct detection technique.

II. General Experiment Overview

A. TRAPPED ATOMS AS A TARGET

A.1. Basics of magneto-optical trapping


In our experiments we use a standard Magneto Optical Trap (MOT) (Chu et al.,
1985) as the target for our collision experiments. The general principles of atom
trapping are described elsewhere (Foot, 1991; Metcalf and van der Straten, 1999;
Wieman et al., 1995); in this section we briefly touch upon those subjects that
are relevant to the use of atom traps as a target in electron-impact studies.
An atom trap requires two forces: a velocity-dependent viscous force to cool
atoms and a spatially-dependent restoring force to confine atoms. In an optical
trap both forces are derived from modulating the photon scattering rate, and thus
the resultant momentum transfer rate, between an atom and an opposing laser
beam. A standard MOT consists of (a) three pairs of mutually perpendicular,
circularly-polarized, counter-propagating laser beams tuned to a frequency
slightly below resonance ("red-detuned") and (b) a quadrupole magnetic field
(formed using a pair of anti-Helmholtz coils) with zero (minimum) field located
at the intersection of the laser beams. An atom moving opposite the direction of
a red-detuned laser beam is Doppler-shifted closer into resonance, increasing
the photon scattering rate, which produces the velocity-dependent force that
cools the atoms (i.e., an optical molasses). For an atom situated away from the
magnetic field zero of the trap, the Zeeman effect shifts one of the mj sub-levels
of the atom closer into resonance with the corresponding circularly polarized
laser beam. This beam pushes the atom back towards the center of the trap,
resulting in the restoring force that confines the atoms.
In order to make our discussion concrete we analyze the particular trap we use
that functions on the 5281/2 --+ 52P3/2 (780nm) transition in 85Rb (see Fig. 1).
The 5281/2 level of 85Rb has two hyperfine levels with F = 2 and F = 3. The
trapping laser is red-detuned from 5281/2 F = 3 --~ 52p3/2 F ~ = 4 recycling
transition. On the order of 1 in 102 to 103 absorption cycles, however, will be into
the 52p3/2 F ~ = 3 level which can decay into the 5281/2F - 2 level. Atoms in this
F = 2 "dark" level are non-resonant with the primary trapping laser beam and
are free to drift out of the trapping region. To prevent this, a second laser beam
360 [II

FIG. 1. Hyperfine structure of the 52S1/2 and 52P3/2 levels of 85Rb.

[the "hyperfine (re)pump"] is needed to excite atoms to either the 52p3/2 F t = 2


or F ~ - 3 levels which can both decay back to the 52S1/2 F - 3 trapped level.

A.2. Using a M O T as a target

There are a number of issues involved in using a MOT as a target for collision
studies. Some of these issues are challenges; for example, a MOT relies on a
quadrupole magnetic field that interferes with the propagation of an electron
beam. On the other hand, the very complexity of atom trapping allows the
experimentalist many options for achieving a given task. For instance, the trap
(i.e., the confining and cooling forces) can be turned on/off in many ways:
(a) by turning on/off the magnetic field, (b) by modulation of the trapping laser,
or (c) by modulation of the hyperfine repump laser. Each method has its own
advantages and disadvantages in terms of complexity, cost, and time response. In
fact, there are a number of different natural time scales involved in the trapping
of atoms as listed in Table I. For example, if the trapping forces on a cloud of
trapped atoms are turned off for 5 ms and then restored, few if any atoms will
be lost from the trap. But, if the trapping forces are left off for 30 ms, all of the
atoms will be lost from the trap due to ballistic expansion. This sets a natural
upper bound on the time the trap can be left off. Alternatively, the decay time of
the magnetic field sets a lower bound on the time the trap must be turned off to
propagate an electron beam. Finally, we note that trap parameters and controls are
richly interconnected. For instance, to increase the fraction of trapped atoms in
the 5P excited level one might increase the photon scattering rate, which depends
on both the intensity and frequency detuning of the trapping laser. This will,
however, also affect the forces on the trapped atoms, which causes changes in
II] ELECTRON-ATOM COLLISION CROSS SECTIONS 361

Table I
Approximate times for processes in a typical Rb MOT

Process Time

Lifetime o f 52P3/2 excited level 27 ns


Transfer time by Raman scattering into 52S1/2F = 2 dark state ~25 gs
Decay time of magnetic field (L/R) ~0.5 ms
Time to escape the trap region after the trap is turned off ~20 ms
Time to load trap (trap lifetime) seconds

the loading rate of atoms into the trap, the trap depth, and even the total number
of atoms in the trap.

A.3. Polarization o f atoms in the trap


One problem of interest in electron-scattering experiments is whether or not the
target atoms have some preferred polarization relative to the electron-beam axis.
When the trap is on, the three sets of counter-propagating laser beams assure
that average light polarization taken over the entire trap is zero (Metcalf and van
der Straten, 1999). At any particular location in the trap, however, the atoms can
be polarized along the local magnetic field direction due to the optical pumping
of the atoms; only the spatial average is zero 1.
We also consider what happens when the magnetic field and hyperfine
repumping lasers are both turned off, allowing atoms to fall into the F = 2
dark state. When the current through the anti-Helmholtz coils is turned off, the
magnetic field gradually dies out, and only the local earth's magnetic field is left.
Since the magnetic field decay time (~500 gs) is much longer than the precession
time of an atom, the polarization of the atoms will adiabatically follow the
local magnetic field direction so that the atom cloud may have a small residual
polarization along the earth's field 2.
To estimate the magnitude of these effects, we have developed a simple
rate equation model for a one-dimensional atom-trap model as described in
Appendix 1. For a typical trap, the model yields a polarization at the edge of the
atom cloud of about 0.16, and a value of 0.1 averaged over the entire volume of
trapped atoms. This value is an overestimate since it ignores the other four laser
beams that can depolarize the atoms via absorption followed by spontaneous

1 Intensityimbalances in the counter-propagatinglaser beams could also lead to polarization effects.


2 This potential problem could be greatly reduced by leaving the hyperfine repump on until the
trapping magnetic field has completely decayed away.
362 Schappe et al. [II

emission. We conclude that the polarization of the cloud of trapped atoms (both
before and after the trap's anti-Helmholtz coils are turned off) is small (P < 0.1).

B. APPARATUS

In the following four sub-sections we describe the components of our atom trap
common to all our measurements: the laser system, the vacuum system, the
magnetic field coils, and the electron gun. Further details on the components
unique to each type of cross-section measurement are described in Sect. III.

B. 1. Laser system
The optics associated with the frequency stabilization of the trapping laser are
shown in Fig. 2, while Fig. 3 illustrates the optics associated with the trap. We
use a commercially available 30 or 70mW, 780nm diode laser in an external
cavity arrangement (MacAdam et al., 1992; Arnold et al., 1998) using a Littrow-
mounted 1200 groove/mm holographic grating for feedback. Tuning of the laser
frequency is achieved by varying the laser diode current and temperature, and
coarse adjustment of the grating angle. A piezo-electric stack is installed on the
horizontal grating adjustment to provide fine-control of the laser wavelength by
changing the effective cavity length.
The linearly-polarized, elliptical output beam from the diode laser is converted
into a circular profile with an anamorphic prism pair, and passed through a 40 dB
optical isolator to prevent disruptive optical feedback from retro-reflected laser
light. A Doppler-free saturated absorption spectrometer is used to determine the
laser wavelength relative to the Rb (52S1/2 F = 3 + 52p3/2 F t = 4) trapping
transition. Since we want to lock the laser to a frequency A less than this
trapping transition, we use an acoustic-optical modulator (AOM) to downshift
the frequency of the beam entering the saturated absorption spectrometer relative
to the trapping beam. Feedback from the saturated absorption spectrometer
(Wieman et al., 1995) is used to stabilize the laser frequency at the peak of
the crossover feature between the F = 3 ---+ F ' = 2 and F - 3 --+ F ' = 4
transitions, located 92 MHz below 52S1/2 F = 3 ---+ 52p3/2 F ' = 4 transition.
A single mode optical fiber is used to decouple the laser alignment from the
trap optics alignment and to provide spatial filtering.
A beam expanding telescope is used to convert the output of the optical fiber
into a collimated beam approximately 1 cm in diameter. The linearly polarized
laser beam is split into three beams using half-wave plates and polarizing
beamsplitter cubes. Since the magnetic field gradient is twice as large along
the axis of the anti-Helmholtz coils (the z-axis) as along the x and y axes,
the power in the beams are set in the ratio 2:2:1 so as to roughly equalize the
x , y , z restoring forces. The linearly polarized beams are converted into circularly
polarized beams with zero-order quarter-wave plates. Since the direction of the
II] E L E C T R O N - A T O M C O L L I S I O N CROSS SECTIONS 363

Fie. 2. Trapping laser. 780 nm light is generated by an external cavity, grating-stabilized 70 mW


diode laser. The portion of the beam directed into the Doppler-free saturated absorption spectrometer
is downshifted by -80 MHz by an AOM to allow locking to the large 2-4 crossover peak (see insert
for sample spectrum). The purpose of the LCVR is described in Sect. III.B.2.

magnetic field gradient is opposite for the z-axis relative to the x&y axes, the
beam incident on the trap along the z-axis is left circularly polarized, while the
x&y beams are right circularly polarized.
The hyperfine repump beam is obtained by coupling a 2 . 9 1 5 G H z mi-
crowave modulation signal into the primary trapping diode laser drive current
(Feng and Walker, 1995). This modulation creates a sideband approximately
1% the intensity of the primary laser transition at the frequency of the
5281/2F = 2 ~ 5 2 p 3 / 2 F t -- 3 transition (see Fig. 1). Since only -1 in
200 photons result in a Raman scattering event that populates the dark state,
this weak sideband intensity is enough to keep the dark state essentially empty.
If, however, the hyperfine repump is turned off (by shifting the microwave
364 Schappe et al. [II

FIG. 3. Top view of the atom trap chamber (and optics) used in electron-impactexcitation studies
(Sect. III.B.4), but contains all the elements used in our earlier trap-loss measurements(Sect. Ill.A).
Photodiode and CCD camera are used to monitor the trap. For clarity, the z laser beam and associated
optics are not shown.

modulation frequency -`200 MHz off-resonance), it takes only --400 photon


scattering cycles (< 50 ~ts) to transfer all of the atoms in the trap into the dark
state. This is the primary way we switch the optical forces on the atom on/off
and convert the atoms in the trap from a mixture of atoms in the 52S1/2 F = 3 and
52p3/2 F ~ = 4 levels when the hyperfine repumping is on, to only atoms in the
52S1/2 F = 2 level when the hyperfine repumping is off. Thus it is straightforward
to produce a target of all ground level atoms or a mixed target of ground and
excited level atoms.
Typically we have 30 mW of power directly out of the external-cavity diode
laser. The total laser power (before being split into the three trapping beams) is
10 mW. Most of the losses occur in coupling the light into the optical fiber. The
II] ELECTRON-ATOM COLLISION CROSS SECTIONS 365

laser frequency is very stable (< 1 MHz), typically remaining locked >48 hours
at a time, with an observed variation in trap fluorescence over 24 hours (due to
changes in laser intensity and frequency) of ~< 5%.

B.2. Vacuum chamber


The atom trap is located at the center of the vacuum chamber as shown in Fig. 3.
This chamber has fourteen ports in the horizontal plane, and two ports in the
vertical plane. Four of the horizontal ports and the two vertical ports are for
the three pairs of counter-propagating laser beams used for trapping. Three of
the horizontal ports are used by components of the electron beam: the electron
gun, the Faraday cup, and the translating wire used to measure the electron beam
profile. These are described in Sect. II.B.4. Two ports are used to monitor the
trap fluorescence: a CCD camera provides a qualitative picture of what is going
on, while a photodiode is used to quantitively measure the relative number of
atoms in the trap. Also shown in Fig. 3 are the optics used in the electron-impact
excitation experiment described in Sect. III.B.4.
The ultra-high vacuum chambers we use for trapping are constructed from
non-magnetic stainless steel with an electropolish finish. The chambers are
initially evacuated with a turbomolecular pump and baked out at 150~ for two
to six days. The turbo-pump is then valved off with an all-metal UHV valve
and the chamber is pumped by a 201/s ion-pump. Typical base pressure in the
chamber is 8 x 10-l~ Torr.
A reservoir with a one gram ampule of 99.95% purity Rb is connected to the
main chamber by an all-metal valve. The thermal-velocity Rb atoms diffusing off
the walls of the vacuum chamber provide the source of Rb for the trap. A higher
Rb number density yields more trapped atoms due to an increased trap-loading
rate but also results in shorter trap lifetimes due to increased collisions with
background atoms. By periodically adding only enough Rb to maintain a good
number of trapped atoms (by opening the valve to the reservoir and heating it
to ~50~ with a heating tape), long trap lifetimes can be achieved (~4 s) at the
price of variations in trap size on the time-scale of days. Alternatively, for a more
constant number of trapped atoms but shorter trap lifetimes (~ 1.5 s) the reservoir
can be left open at a reduced temperature (~45~

B.3. Magnetic field


The quadrupole magnetic field necessary for the trap is generated by two
air-cooled coils located above and below the chamber. Each coil consists of
40 turns of 16 gauge high temperature magnet wire. The diameter of each coil
is approximately 7.2 cm, with a coil-to-coil distance of 5.3 cm (which is slightly
larger than the 3.6 cm spacing for true Helmholtz coils). The coils are wired in
series and run in an 'anti-Helmholtz' configuration to provide an approximately
366 Schappe et al. [II

uniform field gradient. Typically, we run 4.0 A through each coil to generate
a measured magnetic field gradient of 9Gauss/cm along the z-axis, although
trapping is possible over a wide range of currents (2 to 18 A).
Since the trapping magnetic field deflects the electron beam, the magnetic field
must be switched off during electron beam pulses. We have used two circuits to
perform this function, one using a power MOSFET as a switch, and one using a
commercial solid-state relay. We measure the effective decay rate of the magnetic
field by monitoring the temporal distortion in the shape of an electron beam
current pulse. The minimum measured delay between the time the magnetic field
is switched off and the start of the electron beam pulse that does not distort the
electron current is 500 ~ts, which is comparable in performance to slightly more
advanced switching circuits with eddy-current compensation (Dedman et al.,
2001).
The center of the trap is determined by the location of the minimum in the
magnetic field, while the optimum trap fluorescence and trap-loading rate are
located at the intersection of the six laser beams. We use the vacuum chamber
to align the two; the coils are attached to the top&bottom viewports, and the
laser beams are centered on each viewport. The most uniform trap, however, is
achieved by using magnetic field shim coils to steer the magnetic field 'zero'
to the intersection of the laser beams. The quality of the trap alignment can be
monitored by observing the dispersal of the atom cloud when the magnetic field
is turned off. For a well aligned trap the atoms disperse isotropically ("poof"),
but for a poorly aligned trap the atoms move off in a directional jet.

B.4. Electron beam

The electron gun used in this work consist of five grids and a cathode assembly.
The stainless steel grids are 1.78 cm square with alumina spacers between grids.
The cathode consists of an indirectly heated BaO cathode. One of the grids can
be biased negative relative to the cathode to chop the beam on and off. The
electron beam current is measured by a deep Faraday cup (L/D ,~ 5). The back
plate of the Faraday cup is conical so that specularly reflected secondary electrons
are not reflected back towards the collision region. The back plate is also biased
at +18 V to prevent the escape of low energy secondary electrons. The electron
gun and Faraday cup are separated by distance of 4.5 cm to allow the trapping
laser beams access to the collision region. Space charge expansion of the electron
beam over this distance limits the low energy performance of the electron gun.
To determine the electron beam current density, J, we translate a thin (0.19 mm
diameter) tungsten wire across the electron beam at the location of the atom trap.
The current measured on the wire produces a series of line integrals of the beam
current density. Assuming the beam is cylindrically symmetric, the measured
beam spatial profile is converted into the current density J ( x , y ) using an Abel
transform (Hansen and Law, 1985). Due to secondary electron emission from
III] E L E C T R O N - A T O M C O L L I S I O N CROSS SECTIONS 367

Fl~. 4. Profile of the electron beam (100 eV) at the location of the trap. Measured current values as
the wire is traversed along the x-axis are deconvoluted with an Abel transform assuming cylindrical
symmetry. The current density is essentially constant over the size of the atom cloud (~< 0.1 cm).

the wire, we only use the translating wire measurements to find the shape o f
the current density. We put J on an absolute scale using the total electron beam
current measured by the Faraday cup. At high energies (>100 eV) the beam is
approximately Gaussian, with a F W H M of 2 mm at high energies (see Fig. 4).
The value of the peak current density ranges from 0.1 mA/cm 2 at low energies
to 3 mA/cm 2 at high energies.
Due to the small size of the atom trap, the electron beam is carefully aimed
at the atom cloud by using a set o f gimbals (particularly for the trap-loss
measurements described in Sect. III.A). The decrease in trap intensity due to
electron-atom scattering collisions can be used to aim the electron beam at
the trap. Due to residual magnetic fields and the build-up of surface charge
on insulators, the center of the electron beam has some dependence on the
electron beam energy, requiring that this alignment must be repeated for different
energies.

III. Methods for Measuring Cross Sections

We have measured electron-atom collision cross sections using trapped atoms


with two different classes of experiments. The first class of experiments (Schappe
et al., 1995, 1996), which is described in Sect. III.A, monitor the time
dependence of the number of atoms in a trap to deduce total-scattering and
ionization cross sections. This trap-loss method relies on the advantages of an
atom trap listed in Sect. I: the trap fluorescence provides a relative measure o f
368 Schappe et al. [III

the number of atoms in the trap; the low initial velocity of atoms in the trap
differentiates electron-scattered atoms, and the small size of the trap simplifies
the absolute calibration. On the other hand, the second class of experiments
(Keeler et al., 2000) uses the high excited state fraction in the trapped atom
target to measure ionization and excitation cross sections from the laser excited
5P level and is described in Sect. III.B.

A. Loss RATE MEASUREMENTS

A.1. Rate equations


Before analyzing a particular type of electron-atom collision process, we first
derive the general relation between the number of atoms in a trap and a collision
cross section. Consider an atom trap that is exposed to a periodically pulsed
electron beam. Since the trapping magnetic field interferes with the electron
beam propagation, the trap is turned off for a short period of time before the start
of the electron beam pulse. The number of atoms in the trap, N, as a function
of time is described by the differential equation

dN
- A L - r0N - f G N , (1)
dt

where L is the loading rate of atoms into the trap from the background vapor, F0
is the loss rate of atoms out of the trap for all causes other than electron-atom
collisions, Fe is the loss rate of atoms out of the trap due to electron collisions,
J~ is the fraction of time the trap is on, and f is the fraction of time the electron
beam is on. For a vapor loaded trap, the loading rate, L, depends on the Rb partial
pressure in the chamber and the laser detuning and intensity. Atoms can only be
loaded into the trap during the fraction of time, J~, when both the trapping lasers
and magnetic field are 'on'.
The loss rate of atoms out of the trap when the electron beam is off,
F0, is due to a number of causes. At very high trap densities, atom-atom
collisions within the trap are the largest loss mechanism, while at lower densities
collisions with hot background gas atoms become important (Walker and Feng,
1994) 3. Additionally, when the trap is turned off, all of the formerly trapped
atoms ballistically expand away from the center of the trapping region with a
velocity dependent on the temperature of atoms in the trap. If the trap is turned
back on after only a very short delay, these atoms are retrapped with almost
100% efficiency. For longer delay times, this ballistic expansion allows atoms to

3 Since the asymptotic number of atoms in the trap is different with and without the electron beam,
to eliminate any possibility of density-dependent changes in F0, one should vary fL so that N~ is
the same in both cases.
III] ELECTRON-ATOM COLLISION CROSS SECTIONS 369

escape the physical limits of the trapping laser beams, thus depleting the number
of atoms in the trap. The average (rms) speed of Rb atoms in a 150 ~tK trap is
21 cm/s. For a laser beam of radius 0.5 cm, atoms leave the trapping volume in
about 24 ms. Since the atoms have a Maxwell-Boltzmann velocity distribution,
and the laser beam has a Gaussian profile, this is not a well-defined time limit.
Gravity will also eventually pull atoms out of the trapping region. The time
it takes to fall 0.5 cm is 32 ms, which is only slightly longer than the ballistic
expansion time for Rb.
Solving Eq. (1) with the initial condition that there are no atoms in the trap,
the number of atoms in the trap as a function of time (a loading curve) is simply

N(t) = N~ (1 - e-(r~ , (2)

where N ~ is the steady state number of atoms in the trap, and is equal to
j~L/(F0 +fFe). If the electron beam is always off, the form of the solution is the
same, except Fe = 0. By fitting a rising exponential to two sets of experimental
d a t a - one with the electron beam on, and one with the electron beam o f f - F~
can be separated from the background loss rate.
The loss rate due to electron collisions, F~, is directly related to the
corresponding electron-atom collision cross section, o, by

oJ
r~ - , (3)

where J is the current density of the electron beam at the location of the trap,
and e is the magnitude of the electron charge. Note that since the size of the
trapped atom cloud is small compared to the size of the electron beam, we can
safely assume J is constant over the volume of the trap. Furthermore, in contrast
to the general difficulties of crossed beam experiments, for an atom-trap target
the only measurements needed to find absolute cross sections are the electron
beam current density (see Sect. II.B.4) and the change in the loss rate with the
electron beam on/off. There is no need to measure the absolute number of target
atoms.
By varying the delay time between the electron beam pulse and the time the
trap is turned back on, T, it is possible to measure different types of collision
cross sections. To measure total scattering cross sections (Sect. III.A.2), we use
a long delay. With a long delay, any atom that has gained any excess recoil-
momentum due to an electron-atom collision will have enough time to leave the
trapping region. On the other hand, if the trap is turned on immediately after
a short electron gun pulse, recoiling atoms will not have had enough time to
leave the trapping region, and will be retrapped. Ions formed via electron-impact
ionization, though, are not resonant with the trapping lasers, and are lost. Thus
only ionizing collisions result in trap loss, allowing us to measure ionization
cross sections as is described in Sect. III.A.3.
370 Schappe et al. [III

Hyperfine~ o f f - - I ton = 30 ms

Magnetic
Field
one, ~ i ~'"
off/ ~,.
i"~
-IT= 0- 18 ms
a

Electron
Beam
off n 0.8 - 4 ms

FIG. 5. Timing diagram for total scattering experiment. The electron beam pulse is delayed from
the the start of the trap-off phase to allow time for the magnetic field to decay (dashed lines).

A.2. Total scattering cross section

Total scattering cross sections (Schappe et al., 1995) are obtained by monitoring
the time dependence of the trap fluorescence as the trap is periodically hit with
an electron beam pulse. A timing diagram of one electron beam pulse cycle is
shown in Fig. 5. Atoms are loaded'into the trap for 30 ms, at which time both the
magnetic field and hyperfine repump are turned off. With no repumping, Raman
scattering shifts atoms into the F = 2 dark state of Rb, which is non-resonant with
the primary trapping laser. Before pulsing the electron beam, however, a delay of
1 ms is needed for the decay of the magnetic field. After a short electron beam
pulse (0.8 to 4 ms long), the trap is left off for a variable time of 0 to 18 ms.
At the end of this delay, the trap is turned back on (i.e., hyperfine repumping
is resumed and the magnetic field is turned on) and the number of atoms in
the trap is recorded. After acquiring trap fluorescence data for approximately
12 s, the number of atoms in the trap reaches the asymptotic value, N ~ . To
obtain another loading curve, the magnetic field is turned off for 2 s to empty
the trap, and the above process is repeated. A loading curve is also obtained
with the exact same timing structure, but without the electron beam to obtain
the background trap-loss rate F0. Typically, the results from five pairs of loading
curves are averaged together to reduce the statistical noise in the data.
The relative number of atoms in the trap, for a fixed laser intensity and laser
detuning, is proportional to the trap fluorescence. This fluorescence is collected
by a lens and detected with a photodiode. The -50 nA signal is amplified by a
current to voltage amplifier with a gain of 6 x 107 V/A and recorded by either a
digital storage oscilloscope or a data acquisition computer. The computer is also
used to control the timing of the electron beam, magnetic field, and hyperfine
repumping.
In Fig. 6 we show a pair of sample loading curves with and without an electron
beam. The trap fluorescence demonstrates a slow exponential rise due to atoms
being loaded into the trap via Eq. (2), superimposed on a rapid modulation due
III] ELECTRON-ATOM COLLISION CROSS SECTIONS 371

' ! ' ! 9 i ' ! ' ! , ]

no e l e c t r o n b e a m .= ..,--,,..:~.

9
" b ~ - _-~r"or'-=~.,L
lb--
l
- ~
l
9 --o
I

.Q

v 3

~
e"
2 e l e c t r o n beam
m
e~

0.1 0.2 0.3 0.4

0
0 2 4 6 8 10 12

Time (s)

FIa. 6. Sample loading curves for an electron beam energy of 50 eV with a 2 ms electron beam
pulse and 13 ms delay time. The data presented in the main plot has been averaged over one full
pulse cycle (46 ms). The raw data for the start of the electron beam off loading curve (shown in the
insert) demonstrates the rapid time dependence of trap fluorescence on the modulation of hyperfine
repumping.

to the modulation of the hyperfine repumping. Since there is no trap fluorescence


when the atoms are in the dark state, this portion o f the raw signal is removed
from the signal prior to fitting. Since Eqs. (1) and (2) include only the time-
averaged loading rate and electron-impact induced loss rate, they do not model
our data for time scales less than the electron-beam pulsing period 4. Thus, before
fitting the raw data in Fig. 6 to Eq. (2), the data needs to be time averaged over
one timing cycle. Numerical simulations have shown that a direct fit to both the
raw data and the time averaged data (over a very wide range of averaging times)
produce equivalent fitted loss rates 5
Note that we measure the loss rate of atoms out o f the trap by monitoring the
time dependence of atoms being loaded into the trap. It would seem to be more
natural, however, to monitor the induced decay rate o f atoms out o f a fully loaded
trap. Indeed, this was the approach of Dinneen et al. (1992) to measure photo-
ionization o f trapped Rb atoms. The general difficulty with this later approach
is the presence of the loading term in Eq. (1), i.e., the n u m b e r o f atoms in the
trap does not decrease as a simple exponential since new atoms are also being
loaded into the trap from the background vapor whenever both the trapping lasers
and magnetic field are on. Dinneen et al. (1992) overcame this limitation by

4 Technically F0 is not fully time independent, since the background loss rates are different when
the trapping lasers are on or off. While some high-frequency residuals in the fit can be removed by
including a two parameter background loss rate, only the time averaged value is needed to extract
the electron-impact induced loss rate.
5 Only data averaged over integer multiples of a timing cycle will produce fits with no spurious
high-order frequency components and valid Z2 values.
372 S c h a p p e et al. [III

20 eV
g..
E 6

T--
v
e- 4
0

co
0
2 250 eV
o
. i . i , i . i , I l I , I , i
0
0 2 4 6 8 10 12 14 16
Delay Time toff (ms)

Fie. 7. Variation of loss rate, and thus cross section, with delay time.

using a laser-slowed atomic beam that could be switched off during loss rate
measurements. Alternatively for a vapor loaded trap, the decay can be measured
for a short time interval if the trapping lasers are turned off so that there is no
loading term. The presence of the loading term in Eq. (1) also complicates the
extraction of loss rates from monitoring only the asymptotic number of atoms
in the trap, N ~ . Relative measurements of loss rates (and thus cross sections)
can be obtained by monitoring the equilibrium number of atoms in the trap with
and without an electron beam. But since N ~ = f L L / ( F o + f F e ) , knowledge of
the loading rate is necessary to place these results on an absolute scale. Hence,
while being less intuitive, we have found the loading curve method to be easier
to implement and analyze experimentally.
The measured electron-induced loss rate varies with the delay time as is shown
in Fig. 7. For short delay times, only atoms with a very large recoil velocity have
sufficient time to leave the retrapping region before the trap is turned back on.
Kinematically, the relation between the recoil velocity of the atom, V, and the
scattering angle of the electron, 0, is (McDaniel, 1989)

2E0 - AE M(m+M) V 2
COS 0 = 2m , (4)
2 ~Eo(Eo - AE - 89M V 2)

where E0 is the energy of the incoming electron, AE is the excitation energy of


the collision, and m and M are the masses of the electron and atom, respectively.
As small electron scattering angles correspond to low atomic recoil velocities 6,

6 At 10 eV all inelastically scattered atoms via 5S-5P excitation have recoil velocities in excess of
100 cm/s, and are thus included for delay times in excess of 5 ms. For elastic scattering (at 10 eV),
a delay time of 18 ms corresponds to 0mi n ~1.6~
III] ELECTRON-ATOM COLLISION CROSS SECTIONS 373

longer delay times allow more of these slow moving atoms to escape, which is
equivalent to probing electron-scattering angles closer to 0 ~ Thus the measured
trap loss corresponds to the integral of the relevant differential cross sections

2~ (0) sin 0 dO, (5)


,n(r) d ~

where T is the delay time, d o / d ~ is the "total" differential cross section, and
Omin(T)is found from Eq. (4) assuming V = r/T with r being the effective radius
of the retrapping zone. Since the data in Fig. 7 converges to an asymptotic value
well before the maximum delay time, we can be assured that our total scattering
cross section measurements include the contribution down to all non-negligible
scattering angles.
In principle, since there is a functional relationship between delay time and
the electron scattering angle, the derivative of the trap loss vs. delay time curve
can be used to measure differential cross sections,

do d0min do
d---~(T) = -2:r dT (T)~-~(0mi,) sin 0min. (6)

For electron energies below the first excitation energy (-2eV), this would
correspond to the elastic scattering differential cross section. On the other
hand, at very high energies the elastic cross section is negligible, leaving
only contributions from excitation into all bound levels because the ionization
component does not contribute to the variation of trap loss with the delay time as
the ions are never retrapped regardless of the delay (see Sect. III.A.3). For alkali
atoms the nS ~ nP excitation cross section dominates all the other nS ~ n~L
excitation cross sections (Phelps et al., 1979). It may be possible to utilize a
detailed measurement of the trap loss as a function of the delay time to obtain
information about the 5S ~ 5P differential excitation cross section at very small
scattering angles, which is difficult with conventional methods. The effective
angular resolution of such a measurement is limited by the length of the electron
beam pulse, the number of data points, and the quality of the trap model used
to relate the atomic recoil velocity (V) to the delay time (T).
As indicated by the simplicity of Eq. (3), there are only two measured
quantities, and hence only two major sources of uncertainty, that enter into
the determination of the cross section. The measurement of the peak current
density which is obtained by taking the Abel transform of the translating thin-
wire electron beam profile has an estimated uncertainty of 7%. The uncertainty
in the extraction of the electron-induced trap-loss rate Fe from fitting the loading
curves, and finding the asymptotic delay time value is estimated to be 6%.
Results obtained over a wide variety of trap parameters (laser intensity, detuning,
laser beam diameters, trap size) and experimental parameters (electron beam
pulse length, electron beam spatial width, trap on time) give consistent results.
374 Schappe et al. [III

Trap
off
1 ton=-10-20 ms J I
Oil

Electron ' -~1 ms


~ 0.167 - 2 ms
Beam
off
FIG. 8. Timing diagram for ionization measurements. Both the trapping magnetic field and
hyperfine repump laser are turned off during the trap-off phase.

Thus, in the apparent absence of any secondary effects, we believe the total
uncertainty in our measurements is about 9%.
One very important secondary effect that can complicate measurements made
with the trap-loss technique is electron stimulated desorption (ESD) of Rb atoms
from the Faraday cup. The background trap loss is dominated by collisions
of trapped atoms with background gas atoms in the trap. Any increase in the
background gas pressure synchronous with the electron beam will appear to
be due to electron-atom collisions and will be erroneously included in Fe.
Due to the low operating pressure of the trap chamber, the small number of
atoms/molecules liberated from the Faraday cup by ESD can significantly change
the background gas number density. For example, we have observed the pressure
in the chamber rise from 10 -9 to 10 -7 Torr when the electron beam first hits
the Faraday cup in a chamber that was pumped down after being opened. Two
actions minimize the ESD-induced gas load: the chamber is baked out at 250 ~
for 48 hours to remove as much contamination from the chamber as possible,
and before data is collected the electron beam is left on at a high current
(100 to 400 gA) for an extended period of time. Data is only acquired after the
pressure in the chamber is the same with the electron beam gated on or off.

A.3. Ionization cross section

As described in Sect. II1.A, if the trap is turned on immediately after the electron
gun pulse, most of the trap losses will be due to ionization since the ions are
non-resonant with the trapping laser which is tuned to an atomic transition. As
seen by the timing diagram in Fig. 8, the ionization experiment is very similar
to the previously described total scattering experiment. The trap is periodically
turned off by turning off the magnetic field and hyperfine repump. After the delay
needed to allow the magnetic field to decay, the electron beam is turned on for
a short pulse (0.167 to 2 ms). The trap is turned on immediately following the
electron beam pulse and the trap fluorescence is recorded. New atoms are loaded
into the trap for 20 ms and the process is repeated. Note that the cross section
measured with this technique is related to the number of ions created (or the
III] ELECTRON-ATOM COLLISION CROSS SECTIONS 375

1.75 . . . . . ! . . . . ! . . . . ! 9 , , ~ 1-

1.50
50 eV
1.25

,~ 1.00

IT" 0.75
2 5 0 eV
co
o~
0
--J 0.50
G)
._
0.25
n'-
0.00 | | | , l i i i | i , ,

0.0 0.5 1.0 1.5 2.0


Electron B e a m Pulse W i d t h (ms)

FIG. 9. Variation in trap-loss rate versus the width of the electron beam pulse. At high electron
energies, collisions are mainly small-angle electron scattering (low atom recoil). At low energies
(50eV), however, large-angle (fast atom recoil) scattering is more evident. Measurements at widths
less than 0.2 ms are dominated by trap-depth effects.

number of atoms lost from the trap), and does not depend on the charge state of
the ion produced, i.e., we measure ~rion
"count defined as

ount = ~
O-ci o n
ok+, (7)
k

where o k+ is the cross section for producing a Rb k+ ion.


In addition to the loss of atoms via ionizing collisions, the finite duration of
the electron beam pulse allows atoms scattered with a sufficiently high recoil
velocity to leave the retrapping zone before the trap is turned on at the end of the
electron beam pulse, i.e., Eq. (5) with T approximated by half the electron beam
pulse width. For example, with a 1 ms electron beam pulse, an atom receiving
an electron-impact recoil velocity in excess of 69 cm/s will escape the retrapping
zone. Since the total scattering cross section is dominated by 5 S - 5 P excitation
(Schappe et al., 1995), this velocity corresponds to an electron scattering angle
of 1.5 ~ for a 50eV electron. By taking measurements with shorter electron
beam pulses, this angle can be increased. In principle it appears that taking
measurements with ever shorter electron beam pulse widths would completely
eliminate this effect. There is, however, an upper limit o n 0mi n which is set
by the finite trap depth [estimated to be 3 • 10-5 eV based on measurements
of Hoffmann et al. (1996)]. Since the differential cross sections are sharply
peaked at low scattering angles, few electrons undergo large angle scattering,
so the non-ionizing collisions with scattering angles between 0mi n and 180 ~ do
not significantly affect our results. This can be demonstrated by taking loss rate
measurements at a variety of electron-beam pulse widths and extrapolating to
zero pulse width as shown in Fig. 9.
376 Schappe et al. [III

Due to the inclusion (and subtraction) of secondary effects in the ionization


cross-section measurements versus the total-scattering measurements described
in the last section, the estimated uncertainty in our ionization cross sections
(Schappe et al., 1996) are slightly higher. At 50 eV, the uncertainty is estimated
to be 14%, but due to the diminished large-angle scattering at high energies, the
uncertainty drops to 9% at 500 eV.

A.4. Trap-loss measurements f o r collisions involving laser excited atoms


When the trap is on (or at least the trapping laser and hyperfine repump), a
significant fraction (fe ~ 0.4) of atoms in the trap are in the 52p excited
state. Hence, if the hyperfine repump laser is left on for the duration o f the
electron beam pulse, it would appear to be possible to repeat the measurements o f
Sect. III.A.2 and Sect. III.A.3 to find total-scattering and ionization cross sections
from the 52p excited state. Unfortunately, even without the trapping magnetic
field, the laser beams alone set up the optical molasses viscous force that slows
the escape of electron-scattered atoms. Accounting for this effect may entail non-
trivial modifications in the measurements o f the total scattering cross section 7,
but only slightly complicates ionization measurements which can be taken in the
limit of very short pulses (Schappe et al., 1996; Keeler et al., 2000). Ionization
results with this technique are compared to the more straightforward direct
detection measurements in Sect. III.B.3.

B. EXPERIMENTS BASED ON DIRECT DETECTION OF FINAL STATE

B. 1. Ratio measurements f o r Rb 5P/5S processes


In the trap-loss measurements described in the preceding sections, the electron
beam pulse occurred when the hyperfine repump was off and the trapped atoms
were all in the 52S1/2 F = 2 dark state. If the electron beam and hyperfine repump
are on simultaneously, the resulting signal is the weighted sum of cross sections
out of both the 52S and 52p levels. If both the excited state fraction o f the target
and the cross section from the 52S ground level are known one can derive the
corresponding cross section from the 52p level. It is difficult, however, in trap-
loss measurements to fully isolate the 52p signal contribution from changes in
the 52S signal due to the lasers being on (Sect. III.A.4). We therefore find it
advantageous to directly detect the ions or excited atoms formed by collisions.
In Sect. III.B.3 we describe measurements for ionization out o f the 52p level,

7 It may be possible to overcome this limitation on measuring 5P total scattering cross sections,
however, by shifting the red-detuned trapping laser beam directly onto resonance. On resonance, the
absorption and emission of photons does not provide any net force to the atoms - it will, however,
rapidly heat them, leading to increased background trap-loss rates which may pose a new problem.
III] ELECTRON-ATOM COLLISION CROSS SECTIONS 377

while excitation cross sections out of the 52p level are described in Sect. III.B.4.
Both of these measurements, however, depend upon an accurate measurement of
the excited state fraction.

B.2. Measurement o f excited state fraction

Generally, it has been the practice to calculate the expected excited state fraction
based upon the known laser intensity (Walker and Feng, 1994). If we assume the
trapped atoms can be treated as a simple two level atom, stimulated emission
limits the maximum excited state fraction, fe, to be 0.5 or less in steady state.
As a function of the laser intensity, I, the fraction of excited atoms in the trap
is (Metcalf and van der Straten, 1999),

1 I/Is
fe - 2 1 + I/I, + 4A2/F 2' (8)

where Is is the saturation intensity of the transition, A is the detuning of the


trapping laser from the atomic transition, and F is the natural linewidth of the
transition (5.89 MHz for Rb 5 3 - 5P). We generally set A to a value between
7 to 10 MHz.
For a two level atom, the saturation intensity can be calculated from

;rhc
/" 3/Pr (9)

where h is Planck's constant, c is the speed of light, A is the wavelength of the


trapping laser (780nm), and r is the lifetime of the 52p level (27ns), which
yields a value for the saturation intensity of 1.64 mW/cm 2. This value, however,
is not directly applicable to atoms in a MOT since it assumes the atoms are in a
closed two-state system, i.e., the 5231/2 F = 3, MF = 3 and 5 2 p 3 / 2 F ' = 4, M~ = 4
states. This corresponds to the target being completely polarized; however, as
described in Sect. II.A.3, atoms in a MOT are almost completely unpolarized.
Assuming equal populations in all 7 MF states (no polarization), we calculate a
saturation intensity of 3.6 mW/cm 2. This value is slightly larger than that found
in a MOT, since the atoms in a MOT are moving slowly enough that they are
optically pumped into an internal state consistent with the local intensity and
polarization of the standing wave created by the counter-propagating laser beams.
Using a realistic population distribution among the MF states, we calculate
a saturation intensity of 3.1 mW/cm 2. Additional complications can arise in
calculating the saturation intensity for trapped atoms due to spatial variations of
intensity, polarization, and magnetic field across the trap (Javanainen, 1993).
The measurement of the laser intensity, I, at the location of the trap also has
a degree of uncertainty associated with it. Typically one measures the effective
laser intensity at the position of the trap by measuring the laser power that passes
378 Schappe et al. [III

through a small aperture approximately the size of the trap (d = 1 mm). If the
laser beams are well aligned, the total laser intensity is simply the sum of the
peak intensities of the six counter propagating laser beams. Any misalignment of
the beams, or unaccounted-for losses at windows, etc., would introduce an error
into this value. Additionally, for our laser not all of the measured laser power
is available for excitation into the 52p3/2F' = 4 level since some of the laser
power is in the hyperfine repumping sidebands.
Due to these complications, we instead determine the excited state fraction of
the trap by fitting a surrogate measure of the excited state fraction to Eq. (8)
using Is as a parameter. In steady-state, the 5P ~ 5S trap fluorescence is equal
to the number of 5P atoms in the trap times the transition probability for the
5P ---. 5S transition. The number of 5P atoms in the trap is simply feN where
N is the total number of atoms in the trap. Thus, for a fixed number of trapped
atoms, the trap fluorescence is directly proportional to the excited state fraction.
The number of atoms in the trap, however, generally depends upon the laser
intensity, detuning, and a vast variety of other parameters. To keep the number of
atoms in the trap constant, the trap is always loaded with the same laser intensity,
/Load. A fast variable attenuator is then used to rapidly switch to a new laser
intensity, Im, and the trap fluorescence is recorded. These relative measurements
of the excited state fraction as a function of the easily measured laser power are
then fitted to the shape of Eq. (8) with I~ as the only free parameter (i.e., the
asymptotic value is forced to 0.5).
A liquid crystal variable retarder (LCVR) followed by a linear polarizer is used
to rapidly attenuate the trapping laser intensity. The response time of the LCVR
we use is on the order of 10ms. Few atoms are lost out of the trap in 10 ms,
even with a full attenuation of the laser intensity. However, ballistic expansion
of the atom cloud in this time may affect the overlap of the laser beam and
atom cloud for very low attenuation powers. This effect should be minimal for
Im >~ 0.05ILoad. The trap fluorescence is monitored with a photodiode in the
same way as that used in our trap-loss measurements. Since the photodiode also
detects scattered laser light (and fluorescence from background atoms), we take
the difference of two measurements: one with the trap on, and one with the trap
off obtained by turning off the hyperfine repump.
A sample plot of excited state fraction versus laser intensity is shown in
Fig. 10 for a laser detuning of 9.9 MHz and maximum laser power of 6 mW (total
power before being split into separate beams). The fitted value of the saturation
intensity, 3.5 mW/cm 2 falls within the range of expected values. Interestingly,
the horizontal scaling of Fig. 10 is irrelevant, e.g., we plot fe versus laser power
instead of laser intensity as in Eq. (8). If we assume the calibration of the
horizontal scale is off by some scale factor (e.g., not properly accounting for
window losses), our fitted Is value will be off by the same factor. However, the
vertical scale, and thus the extracted f~ value, is unaffected.
III] ELECTRON-ATOM COLLISION CROSS SECTIONS 379

0.5 . , . , . , . , . , . , . , . , . , .

0.4
.o
0.3
LL

0.2

._

x
UJ
0.1

0.0 . . . . . . . . . .

L a s e r P o w e r (mW)

FIG. 10. T y p i c a l p l o t o f e x c i t e d s t a t e f r a c t i o n v e r s u s laser power with a detuning A = 9.9 MHz.


One can obtain higher excited state fractions (~0.45) at a given laser power by decreasing the
detuning or focusing the laser beams to a smaller diameter.

The one underlying (and possibly problematic) assumption made by using the
trap fluorescence as a surrogate forf~N is that the atoms in the trap are assumed
to be in either the 52SI/2 F = 3 or 52P3/2 F ~ = 4 levels. However, if the hyperfine
repump laser is at a low intensity, a substantial number of atoms in the trap will
be in the 5 2 3 1 / 2 F = 2 dark state. Assuming the microwave modulation induced
sidebands for hyperfine pumping are 3% of the intensity of the primary trapping
laser intensity, we calculate <~ 5% of the atoms are in the F = 2 dark state at any
one time. This has a negligible effect on our f~ measurements. If the intensity
of the sidebands is reduced to 0.5%, the fraction in the dark state soars to 22%.
To verify that there are essentially no atoms in the dark state we increase the
size of the microwave-modulation sideband intensity until the trap fluorescence
is essentially independent of the repump intensity.

B.3. Ionization cross section out o f the 5P level


In this experiment trapped atoms ionized via electron-atom collisions are directly
detected using a channel electron multiplier (CEM) detector (Keeler et al., 2000).
A pulsed electric field is used to extract the ions, allowing time of flight (TOF)
analysis to separate out the individual Rb +, Rb++, ... channels. The ionization
cross section of the 52p level is determined relative to the known 52S ionization
cross section (Schappe et al., 1996) using three different measurements: (1) the
ion signal recorded when no trap is present, Coff; (2) the ion signal from a trap
target with the hyperfine repump off at the time of the electron beam pulse (only
52S atoms), Css; and (3) the signal with the hyperfine repump on (a mixed target
of 5 2 S and 52P atoms), C m i x . The Coff signal is the sum of dark counts (D) and
ions created by ionizing the background Rb gas in the vacuum chamber (BG).
The Css count is equal to the Coff signal plus ions created by electron-atom
collisions with the 52S atoms in the trap (Sss). In addition to the background
380 S c h a p p e et al. [III

Magnetic Fiel'~']' 4.0 ms

I trapoZ.
Hypermfi;e~' 5"~on'~Ysigna 1
I
~ 5S&5P mixed signal

i_ .I
Electron Beam [0.6 ms-]

Ion Extractor

Ion Detection Gate ]

FIG. 11. Timing diagram for measuring ionization out of the Rb(5P) level. The target is controlled
by turning the hyperfine pumping laser on and off: no trapped atoms are present when it is always
off, a 5S target is achieved by turning it off only during the electron beam pulse, and a mixed
5S&5P target is obtained by leaving it on during the electron beam pulse.

sources, the Cmix signal includes a contribution from the fraction of atoms in
the excited state, fe Ssp, and a contribution from the fraction of atoms in the
ground state, (1 - f e ) Sss. Hence,

Cmix = D + B G + feS5p + ( 1 - f~) $5s. (10)

In general, the electron-impact signal rates (BG, S5s, S5p) depend upon a vast
collection of experimental parameters including: the electron beam current, the
number of target atoms (in the trap and in the background vapor), and the 52S
and 52p ionization cross sections. Assuming a constant number of atoms in the
trap for both the 52S and mixed 52S & 52p targets, and a constant overlap of
the electron beam with the trapped atoms, it is possible to find the ratio of the
52p to the 52S ionization cross sections from the expression

o+(5P) (Cmix - Coff) - (1 - f e ) ( C 5 s - Coff)


(11)
a+(ss) L(Css -Co~)
Note that only measurements of the three signal rates and of the excited state
fraction are needed. Quantities such as the electron beam current, number of
atoms in the trap, and ion collection efficiency are essentially the same for the
ground state and mixed targets and thus divide out.
A timing diagram of the three phases of the experiment is shown in Fig. 11.
After a delay to allow the magnetic field to decay, the trap is hit with electrons
for 0.6 ms. After a 100 ~ts delay to make sure there are no longer any electrons
in the collision region, an ion extractor electrode (used to pull ions into the
III] ELECTRON-ATOM COLLISION CROSS SECTIONS 381
t ' i | | ! !

16 Rb++ Rb +

.mr 12

~ 8

r- 4
._~

o 0
m 2 3 4 5 6 7 8
Time (gs)

Fie. 12. Time of flight spectra for detection of Rb ions at 150 eV. The dotted line is background
only (no trap), the solid line is with an atom trap target (53 only).

CEM detector assembly) receives a high voltage pulse. After the ion signal
is recorded, the magnetic field is turned back on. To replace atoms lost in
electron-atom collisions, additional atoms are added to the trap for 100 ms. We
obtain reasonable signal rates by averaging counting rates for 300 pulses, which
corresponds to 30 s.
To obtain a signal from only trapped atoms in the 5231/2 F = 2 dark state,
the hyperfine repump is turned off for the duration of the electron-beam pulse.
A mixed 523 & 5 2 p target is obtained by leaving the hyperfine repump on during
the electron beam pulse. To maintain a constant number of atoms in the trap for
the two cycles the hyperfine pumping is turned off for 0.6 ms shortly before the
electron beam pulse to maintain a constant repump duty cycle. A background
run (no trap) is obtained by leaving the hyperfine pump (or magnetic field) off.
When switching from the trap off data run to one of the trap signal runs, we first
reload the trap for 10 seconds before acquiring data.
A time of flight spectra at an electron energy of 150 eV is shown in Fig. 12.
Singly ionized Rb atoms from the trap arrive at the detector at (4.5+0.5)~ts. As
seen in Fig. 12, ions formed by ionization of background gas atoms arrive at the
detector with a much broader range of arrival times, approximately (3.5+2)~ts.
To maximize the SNR ratio, the ion signal is recorded with a box-car integrator
with a signal gate centered 4.5 ~ts after the start of the ion-extractor pulse with
a total width of 1 ~ts. Based upon the arrival time of the Rb + ions, the Rb ++ ions
are expected to arrive at the detector at 3.2 pts, versus an observed peak arrival
time of 2.9 pts. This illustrates another advantage of the trapped atom target in
that the Rb + and Rb ++ yields can be determined separately by means of a simple
time-of-flight detection system.
The largest single source of ions in this experiment is from ionization of
background gas atoms. The number of ions produced by the electron beam is
proportional to f n(F)J(F)dY, where n is the number density of atoms, and J is
the electron beam current density. For ionization of trapped atoms, this is equal
to NJo, where N is the number of atoms in the trap, and J0 is the current density
382 Schappe et al. [Ill

at the trap location. For ionization of atoms in the Rb background vapor, ions are
created along the entire length of the electron beam from the electron gun to the
Faraday cup. The number of background ions created is proportional to nscleL,
where nBc is the number density of background atoms, Ie is the total electron
beam current, and L is the effective length of the electron beam from which ions
are extracted into the detector. Using typical values (N ~ 106, J0 ~ 2 mA/cm 2,
nsa ~ 109 cm -3, Ie ,~ 100gA, L ~ 1 cm), the ion yield from the background
vapor is a factor of fifty larger than the ion yield from the trap. Two factors help
minimize the contribution from background gas ionization: (1) the ion-extractor
is designed to minimize the effective range L from which ions are collected (since
ions from the trap are well localized at the trap center), and (2) TOF selection is
used to separate the 'cold' trap ions from the 'hot' ions created by ionization of
room temperature background gas atoms. Details about the design and operation
of the ion extractor are given in Keeler et al. (2000). Computer modeling of
ion trajectories indicate that the transmission of the ions to the detector is
essentially constant for ions within +0.25 cm of the extraction axis. In principle,
the extractor could be designed to have a smaller extraction volume (i.e., the size
of the ball of trapped atoms) to further eliminate the inclusion of background gas
ions, but this would require the placement and alignment of the extractor relative
to the cloud of trapped atoms to tolerances beyond our control. To maximize the
alignment of the trap to the fixed location of the ion-extractor, shim coils are used
to move the zero of the trapping magnetic field. The contribution from ionization
of background gas atoms can also be improved by switching from a vapor loaded
trap to one fed by a collimated, slowed atomic beam.
The estimated uncertainty in our o+(5P)/o+(5S) measurements is on the order
of 28%. The 52p ionization cross section is obtained by multiplying the measured
ratio in Eq. (11) by the ionization cross section out of the 52S ground state. Note
that the total ionization cross section measured in Sect. III.A.3 using the loss-
rate technique (Schappe et al., 1996) is the sum of the Rb +, Rb++, ... ionization
cross sections, i.e., Eq. (7). Thus the contribution from higher ionization levels
must be removed before these values can be used to place the 52p results on an
absolute scale. Combining the uncertainties in the ratio measurement, the ground
state total ionization cross section, and the correction of Ocount to o +, the total
uncertainty in our 52p cross sections (Keeler et al., 2000) is estimated to be
+33%.
It is interesting to compare the 52p ionization cross-section results obtained by
direct detection of the ions with those using the trap-loss technique as described
in Sect. III.A.4. The loss-rate measurements must be interpreted with some
trepidation since the method forces us to assume that loss rate from atoms in the
trap during a brief electron beam pulse is the same if the hyperfine repumping
is on or off. That is, the recapture rate of atoms is the same for 52S atoms with
no optical molasses and for a mixture of 52S and 52p with an optical molasses.
III] ELECTRON-ATOM COLLISION CROSS SECTIONS 383

Nonetheless, at 50 eV, the loss rate measurement (Keeler et al., 2000) yields a
total ionization cross section of 2.2x 10-15 cm 2 in remarkably good agreement
with the o+(5P) direct detection value of (2.5+0.7)x 10-15 cm 2.

B.4. Excitation cross sections out o f the 5P leoel

Measurement of electron-impact excitation cross sections into discrete levels of


Rb from atoms starting in the 52p level can be accomplished in much the same
way as the ionization cross sections described in the last subsection except that
one detects photons from the decay of excited atoms rather than ions themselves.
The general methods for measuring electron-impact excitation cross sections
by optical detection with a static or atomic-beam target have been discussed
extensively in the literature (Filippelli et al., 1994). Generally the determination
of excitation cross sections requires difficult absolute measurements of both the
radiation from the excited atoms and the target-atom number density. Similar to
what was done to measure the 52p ionization cross section, we bypass these
difficult steps by using the trap to measure the ratio of the 52p --+ n2L to
52S ~ nZL excitation cross sections and then using previously measured 52S
ground state excitation results to place the 52p results on an absolute scale.
Excitation cross sections from the Rb(52S) ground level into the 52p, 72S, 82S,
52D, and 62D levels are available in the literature (Chen and Gallagher, 1978;
Wei et al., 1993).
The apparatus for the electron-impact excitation work is shown in Fig. 3. The
electron gun and Faraday cup are almost identical to those used in our earlier
loss-rate measurements described in Sect. II.B.4. The component unique to this
experiment is the optical system to detect the fluorescence from atoms excited by
the electron beam. An f~ 1.33 lens located within the trapping chamber collimates
light emitted from the electron-impact excitation of trapped atoms. To cut down
on the large amount of scattered light at the wavelength of the trapping lasers,
we use two holographic notch filters each having an effective optical density of
4 at 780 nm. A narrowband interference filter (1 nm FWHM) is used to provide
spectral isolation of an individual transition line. Scattered laser light is further
blocked from reaching the PMT by using a 2 mm field stop that acts as the
limiting aperture of the optical system. Photons are detected with a Hamamatsu
R943-02 (GaAs photocathode) PMT operating in photon-counting mode.
The large background signal that comes from the intense trapping laser beams
represents one of the greatest challenges to this experiment. To illustrate the
difficulties due to this background, consider an attempt to measure the electron-
impact excitation cross section into the 52p1/2 level which decays to the ground
state with transition wavelength of 795 nm. To simplify our illustration, we
consider light from only three sources: electron-impact excitation signal at
795 nm, trap fluorescence at 780nm, and scattered laser light at 780 nm. There
are approximately 106 atoms in the trap (assumed to have a radius of 0.5 mm).
384 S c h a p p e et al. [III

Accounting for the solid angle of the optical system, and assuming a typical
cross section value of 10- 1 6 cm 2, and an effective electron beam current of 10 gA
passing through the trap volume, the electron-impact excitation signal is expected
to be approximately 3 x 10-14 W. If we are measuring the excitation cross section
out of the laser-excited level, the atoms in the trap will be fluorescing at the
same time we are looking for electron-impact excitation signal. The 780 nm trap
fluorescence radiated into the solid angle of the optical system is approximately
2 x 10-6 W. Note that as this light is co-located in space and time with the electron
excitation signal, the only way to remove it is by wavelength selection. For a 1 nm
FWHM filter centered at 795 nm, the transmission at 780 nm is ~ 10--4. With only
10 -4 attenuation, the trap fluorescence is four orders of magnitude larger than
the expected excitation signal, hence the need for the holographic notch filters
which have a high transmission (>80%) at all wavelengths except the 780nm
laser wavelength. Another source of scattered laser light arises from a beam
reflecting off of surfaces near the trap region. For example, to keep the space
charge spreading of the electron beam to a minimum, the electron gun is located
within 12 mm of the collision region. For a laser beam with a 1/e 2 diameter of
1 cm, we might expect ~1% of the beam to be reflected off of the electron gun.
For an average beam power of 3 mW, this corresponds to 3 x 10-s W of scattered
780nm light. Two additional measures beyond the wavelength-selection filters
are used to minimize this light source. First, surfaces close to the viewing region
are coated with colloidal graphite which has a measured reflection coefficient of
<1% at 780 nm. Second, since this light does not originate at the exact location
of the trap, most of the scattered light can be removed by limiting the viewing
region to the trap volume.
Data is taken in three phases similar to those used in the 5P ionization
experiment (see Sect. III.B.3). However, in comparison to the ionization
experiment where there were four major sources of ion counts (two signal sources
and two background sources), there are seven major sources of photon counts
(two signal sources and five background sources). A detailed description of these
sources and typical count rates from each source are listed in Table II.
The signal is extracted from these numerous background sources using the
timing cycles shown in Fig. 13. In the first phase two counting gates are used.
Within each of these gates the electron gun is on half the time and the hyperfine
repump is on half the time. In the A-gate the timing of electron beam and laser
pulses are synchronous, while in the B-gate the two pulses are asynchronous.
When the pulses are synchronous, the resulting signal is from a mixed 52p &
52S target, while the asynchronous gating produces excitation from only the
52S level. This leads to a very simple determination of the relative difference of
the 52p and 52S signal rates. Using the definitions provided in Table 2:

A1 = 2 S L + 2L + B G + B F + T F + (1 - f e ) E s + f e E p , (12a)
B 1 = 2 S L + 2L + B G + B F + T F + E s , (12b)
III] ELECTRON-ATOM COLLISION CROSS SECTIONS 385

Table II
Light sources in an atom-trap experiment measuring the optical emission cross section for the i ~ j
transition

Name Process /l Size a

Excitation from the laser excited 52p level into level i /l~j 100
Excitation from the 52S ground state into level i /lij 130
BG Excitation of background gas (52S) atoms into level i /1#- 6
TF Trap fluorescence ~rap 1
BF Fluorescence from background gas /1trap 1
L Scattered laser light /1trap 2
SL Other stray light sources (i.e., cathode blackbody emission) all/1 20

a Relative signal sizes are based upon measurements of the 52P1/2 ~ 52S1/2 optical emission
cross section, after taking all measures to reduce scattered laser light at/1trap = 780 nm.

L 1.35 ms .L 37.8 ms J
Magnetic Field [ i
i
i
T i
! i i
i
i
I
! r
Hyperfine Pump

Electron Beam ilj !ll


Photon Counting Gate

~0.50 ms'],,
9 !

Magnetic Field I i il
[
.ypo ne um I iil
Electron Beam l !

Photon Counting Gate ~-~


.~, ,~.-

0.25 ms

Magnetic Field I j! jJI

Hyperfine Pump !

Electron Beam [ i
|

Photon Counting Gate

FIG. 13. Timing diagram for 5P electron-impact excitation experiment. Phase-1 measures
excitation from the mixed target. Phase-2 measures excitation from the 5S target, and Phase-3
measures the background. We accumulate counts in each phase for 2000 cycles (~100 s).
386 Schappe et al. [IV

SO
A1 - B1 : fe(Ep - Es). (~3)
In phase two, we record the electron-impact excitation signal from the trap, but
with the hyperfine repumping off during the entire counting gate; hence signal
arises only from excitation of the 52S level. Phase three is the same as phase
two, except the hyperfine repump is always off, so we never have a trap. The
difference in these two counts gives the signal rate for excitation of trapped atom
in the 52S ground state,

A2 - A3 : (SL + L + BG + Es) - (SL + L + BG) : Es. (14)

The signal rates Es and Ep are proportional to the cross sections for the
respective 52S ~ n2L and 52P ---. nZL electron-impact excitation processes.
In addition, they are also proportional to the total number of atoms in the trap,
the electron beam current density at the location of the trap, the optical system
detection efficiency, and the branching ratio of the particular nL ~ n~L ' transition
observed. All of these other factors, however, are the same for excitation from
both the initial levels, so the ratio of cross sections can be found from the ratio

o(SeP--~neL)_ 1 (A1-B1)
o(52S~n2L) fe A 2 - A 3 +1. (15)

Thus the unknown o(52P -~ n2L) excited state excitation cross section can
be found from the four photon counter readings, the measured excited state
fraction (see Sect. III.B.2) and the known o(52S -~ n2L) ground state excitation
cross section.
Due to the large amount of background light, the signal to noise ratio of
these measurements is very low. Long counting times (-2 days per energy)
are needed to reduce the statistical uncertainty in our measurements to less
than 10%. We estimate the total uncertainty in our measurements including
that from the f~ measurement and from the uncertainty in the ground state
excitation cross section to be 35%. Preliminary measurements for excitation
into the 72S indicate that the 52p ~ 72S cross section is significantly
larger than the 52S -+ 72S cross section as predicted by earlier theoretical
calculations (Krishnan and Stumpf, 1992). A further discussion of the results
of this experiment is outside the scope of this chapter and will be published
separately.

IV. Conclusions

The unique characteristics of trapped-atom targets makes them a powerful tool


for measuring collision cross sections. The small size of the target minimizes
VI] ELECTRON-ATOM COLLISION CROSS SECTIONS 387

the need for beam profile and overlap calculations. The low temperature (low
velocity) of the trapped atoms makes them very sensitive to collisions so
that even small angle scattering (low-recoil) events can be detected. Accurate
measurements of the relative trap population are made by simple measurements
of the trap fluorescence and suitable experimental design can eliminate the need
to determine the absolute target density. These distinctive features allow trap-loss
rate measurements which produce very accurate total scattering cross sections
(uncertainty <10%) that include scattering events down to very shallow angles.
Trap-loss measurements also produce ionization cross sections out of the ground
state with uncertainties of < 14%. Also, it is possible to create a trap target with
a significant and well-known fraction of atoms in an excited state. By directly
detecting the collisionally excited atoms (or ions), it is also possible to measure
ionization and excitation cross sections out of the 5 2 p laser-excited level.
Our present discussion has mainly been limited to measuring Rb cross sections
with an atom trap. However, a wide variety atomic species can be trapped,
including all of the alkali atoms, alkaline-earth, and the noble gases (in
metastable levels) (Metcalf and van der Straten, 1999). Work is also ongoing
to cool and trap molecules (Bethlem et al., 2000). In comparison to the low
polarization of a typical trap (see Sect. II.A.3), it is also possible to create spin-
polarized (P ~ 0.7) atom traps (Walker et al., 1992), which could be used as
targets for a new class of electron collision experiments.

V. Acknowledgments

This work was supported by the National Science Foundation and the U.S. Air
Force Office of Scientific Research.

VI. Appendix. Numerical Model for Residual Polarization

In this appendix we derive a numerical value for the residual polarization of


atoms in a simplified model atom trap using an idealized alkali atom with
no nuclear spin. We further simplify the system by considering only a one-
dimensional trap with two oppositely directed incident laser beams as shown
in Fig. 14. The polarization in the target is proportional to the difference in
the populations of the m j = +~1 and m j - - 1 sublevels in the 2S~/2 level.
For a stationary atom located at the center of the trap where the magnetic field
is zero, the atom is equally likely to absorb photons from either the o + beam
approaching from the left, or the a - beam incident from the right, resulting in
no net optical pumping, and thus no polarization. If the atom is displaced to
the right, however, the Zeeman splitting caused by the non-zero magnetic field
in this region shifts the A m j = --1 transitions closer into resonance, causing a
388 Schappe et al. [VI

(a) (3.+ (y-

I
f

(b) m= -3/2 - 1/2 1/2 3/2

3 4 5 6

1 2

FIG. 14. (a) A 1D model atom trap is located at zero magnetic field. The spins of the photons
entering the trap are anti-parallel to the local direction of the magnetic field. (b) We denote the
2S1/2 m=- 89and+l states by the labels l&2, and the 2P3/2 m - ~,- 89 and +3 states by
the labels 3 through 6.

slight excess in scattering from the o - beam approaching from the right. Besides
pushing the atom back towards the center of the trap, this will also tend to
optically pump atoms into the mj -_ - ~ 1 sublevel The local force on the atoms in
the trap due to the trapping lasers depends on the local polarization. The force
is increased by optical pumping which pumps the atoms into states with a high
differential scattering rate between o + and o - light.
We denote the states in the 2S1/2 ground level with m -- - ~i a n d m = + l by the
3
labels 1 and 2, respectively, and the states in the 2p3/2 excited level with m - 2,
l ~,
~., I and ~3 by the labels 3 , 4, 5 , and 6, respectively (see Fig. 14). At a given
location in space the rate of change of the population NI of state 1 is given by
dN2
dN1 _ 2 Rz4N2 - 2RlsN1 - , (16)
dt 3 3 dt
where RI5 and R24 are the absorption rates for the levels 1 and 2 into states
5 and 4, respectively. The absorption of laser light between levels 2 and 6 and
between levels 1 and 3 do not affect the ground level populations since the
2
excited levels all decay back to the same initial state. The factors 5 are the
branching ratios out of the excited states 3 and 4 to return to the opposite
spin ground state. The population of state 2 is given by N2 = 1 - N1. The
population of state 2 in the steady state is given by N2 = R15/(R15 + R24).
The local polarization of the atoms at a particular location in space is given
by P = N2 - N 1 = (R15 - R z 4 ) / ( R 1 5 + R24).
The rate of absorption of light between state i and state j is given by
16~ij 2
(17)
Rij = [09- o90 + (gumu -gtml)ff2] 2 + F2/4 '
where f2 = ~ B is the Larmor precession frequency of the atoms in the magnetic
field, gu = 4 and gt = 2 are the Land6 g-values of the 2p3/2 and 2S1/2 levels,
VII] E L E C T R O N - A T O M C O L L I S I O N CROSS SECTIONS 389

F is the natural linewidth of the atomic transition, o00 is the natural frequency o f
the transition, and o) is the frequency o f the laser light with a Rabi frequency e~.
If we define the detuning of the laser as A = o) - o)0, in the approximation where
A >> ~ , the denominator can be expanded, resulting in

Rij ~ A2 + F2/4
I ]
1 - A2 + F2/4 (gumu - g t m t ) 9 (18)

Inserting the definition o f the oscillator strength o f the transition, f u , we obtain

2A~ )
Rij = R o f u 1 - A2 + F2/4 (gumu - gzml) 9 (19)

From this we can obtain the individual absorption rates of the m sublevels,

R26 = R0(1 - b), (20a)


R15 :
R0(1 (20b)
R13 = R0(1 + b), (20c)
l
R24 = ~R0(1 + 5 b), (20d)

where b = 2Aff2/(A 2 + F2/4).


The polarization of the atoms at a particular location in space is

R15-R24 _ 5b" (21)


P = Rl5 + R24 3

Typical values for an atom trap used in our work are a detuning A/2Jr - 12 MHz,
a magnetic field gradient of 9G/cm, and an atom-trap size of approximately
0.5 mm diameter. Thus the magnetic field at the edge of the atom cloud is approx-
imately 0.2 G, which yields a precession frequency o f f2 =/z0B - 6.3 x 105 Hz.
Inserting the F/2sr = 5.9 MHz natural linewidth of the Rb resonance transition
along with these values into Eq. (21) yields a polarization value at the edge o f
the atom cloud of 0.16. We have extended this model to the F = 3 case of interest
here; numerical solution gives P = - 1 . 2 9 b , implying a polarization at the edge
of the cloud of 0.21.

VII. References

Arnold, A.S., Wilson, J.S., and Boshier, M.G. (1998). Rev. Sci. Inst. 69, 1236.
Bethlem, H.L., Berden, G., van Roij, A . J . A . , Crompvoets, F.M.H., and Meijer, G. (2000). Phys. Rev.
Lett. 84, 5744.
Chen, S.T., and Gallagher, A.C. (1978). Phys. Rev. A 17, 551.
Chu, S., Hollberg, L., Bjorkholm, J., Cable, A., and Ashkin, A. (1985). Phys. Rev. Lett. 55, 48.
390 S c h a p p e et al. [VII

Dedman, C.J., Baldwin, K.G.H., and Colla, M. (2001). Rev. Sci. Inst. 72, 4055.
Dinneen, T.E, Wallace, C.D., Tan, K.Y.N., and Gould, EL. (1992). Opt. Lett. 17, 1706.
Duncan, B.C., Sanchez-Villicana, V., Gould, EL., and ~adeghpour, H.R. (2001). Phys. Rev. A
63, 043411.
Feng, E, and Walker, T. (1995). Am. J. Phys. 63, 905.
Filippelli, A.R., Lin, C.C., Anderson, L.W., and McConkey, J.W. (1994). Adv. At. Mol. Opt. Phys.
33, 1.
Flechard, X., Nguyen, H., Wells, E., Ben-Itzhak, I., and DePaola, B.D. (2001). Phys. Rev. Lett.
87, 123203.
Foot, C.J. (1991). Contemp. Phys. 32, 369.
Hansen, E.W., and Law, E-L. (1985). J. Opt. Soc. Am. A 2, 510.
Hoffmann, D., Bali, S., and Walker, T. (1996). Phys. Rev. A 54, R1030.
Javanainen, J. (1993). J. Opt. Soc. Am. B 10, 572.
Keeler, M.L., Anderson, L.W., and Lin, C.C. (2000). Phys. Rev. Lett. 85, 3353.
Krishnan, U., and Stumpf, B. (1992). At. Data Nucl. Data Tables 51, 151.
MacAdam, K.B., Steinbach, A., and Wieman, C. (1992). Am. J. Phys. 60, 1098.
Marag6, O., Ciampini, D., Fuso, E, Arimondo, E., Gabbanini, C., and Manson, S.T. (1998). Phys.
Rev. A 57, R4110.
Mark, T.D., and Dunn, G.H., eds. (1985). "Electron Impact Ionization." Springer, New York.
McDaniel, E.W. (1989). "Atomic Collisions: Electron and Photon Projectiles." Wiley, New York.
Metcalf, H.J., and van der Straten, P. (1999). "Laser Cooling and Trapping." Springer, New York.
Phelps, J.O., Solomon, J.E., Korff, D.E, Lin, C.C., and Lee, E.T.P. (1979). Phys. Rev. A 20, 1418.
Schappe, R.S., Feng, P., Anderson, L.W., Lin, C.C., and Walker, T. (1995). Europhys. Lett. 29, 439.
Schappe, R.S., Walker, T., Anderson, L.W., and Lin, C.C. (1996). Phys. Rev. Lett. 76, 4328.
Stumpf, B., and Gallagher, A. (1985). Phys. Rev. A 32, 3344.
Turkstra, J.W., Hoekstra, R., Knoop, S., Meyer, D., Morgenstern, R., and Olson, R.E. (2001). Phys.
Rev. Lett. 87, 123202.
van der Poel, M., Nielsen, C.V., Gearba, M.-A., and Andersen, N. (2001). Phys. Rev. Lett. 87, 123201.
Walker, T., and Feng, P. (1994). Adv. At. Mol. Opt. Phys. 34, 125.
Walker, T., Feng, P., Hoffmann, D., and Williamson, R.S. (1992). Phys. Rev. Lett. 69, 2168.
Wei, Z., Flynn, C., Redd, A., and Stumpf, B. (1993). Phys. Rev. A 47, 1918.
Weiner, J., Bagnato, V.S., Zilio, S., and Julienne, P.S. (1999). Rev. Mod. Phys. 71, 1.
Wieman, C., Flowers, G., and Gilbert, S. (1995). Am. J. Phys. 63, 317.
Wippel, V., Binder, C., Huber, W., Windholz, L., Allegrini, M., Fuso, E, and Arimondo, E. (2001).
Eur. Phys. J. D 17, 285.
Index

A generation of attosecond pulses, 89


Absolute configuration of enantiomers HATI (high-order ATI), 50, 51, 59
D and L notation, 227 history, 36
definition, 225 intensity-dependent enhancements in
determination, 230 plateau, 63-65
R and S notation, 230 interference
Absolute phase of few-cycle laser pulses, between direct and rescattered elec-
90 trons, 71
ADK (Ammosov-Delone-Krainov) tun- between direct electrons, 47
neling rate in strong fields, 40 KFR amplitudes for elliptical polariza-
Amino acids tion, 49
chirality, 225 lobes in energy-resolved angular distri-
Murchison meteorite, 237 butions, 52
homochirality, origin of, 227 need for femtosecond laser, 38
parity-violating energy difference origin of staircase structure for elliptical
(PVED), 241
polarization, 70
polymerization, 254
plateau, 39, 52, 59
racemization, 256
plateau enhancement
stereogenic centers, 225
caused by channel closing, 65
ATI (above-threshold ionization), 7, 35-
dependence on late returns, 67
92
resonance versus channel closing, 66
angle-resolved electron spectra, 38
quantum-mechanical description
application of COLTRIMS, 38
of direct electrons, 44
applications, 86
few-cycle laser pulses, 90 of rescattering, 53
measurement of pulse durations, 88 quantum orbits two-dimensional for
boundary conditions in a focused pulse, elliptical polarization, 70
43 quasi-energy formalism, 61
classical model of direct ionization, 40 relativistic regime, 73-76
classical orbits, 41, 51 rescattering
cutoff, 52 for zero-range potential, 54
dependence of spectrum on quantum in the relativistic regime, 75
orbits, 61, 68 into arbitrary angle, 51
direct cutoff, 42 mechanism, 50-71
direct ionization, 40 in MPI, 27
discrete energies due to interference, 46 model, 37
electric field versus vector potential, 42 relation to Feynman's path integral,
elliptical polarization described by 57
saddle-point equations, 68 relation with closed-orbit theory, 58
experimental methods, 38 role of the binding potential, 59

391
392 INDEX

ATI (above-threshold ionization) (cont'd) heat loss problems, 330-335


saddle-point method for calculating current fluctuations in chip wires, 333
ionization amplitudes, 46 fluctuations in trap potential, 331
saddle-point solution light heating, 334
for elliptical polarization, 48 Lamb-Dicke parameter, 334
for rescattering at high intensity, 55 overview, 334
simple-man model thermal fluctuations in trap potential,
compared to QM solution, 56 332
in relativistic regime, 73 heating problems, 324
not applicable to elliptical polarization, interferometers, 278
68 loading, 304, 307-313
strong-field approximation, 45 into sequentially smaller traps, 309
strong-field approximation to Feynman's magneto-optical trap, 307-309
path integral, 57 observing the atoms, 313
structure in energy-resolved angular transfer from bigger to smaller wires,
distributions, 71 310
TDSE solutions, 39 trap loss problems, 324, 330
theoretical methods, 39 mechanisms, 324-330
time-of-flight detection methods, 38 collisions between trapped atoms,
zero-range potential, 59 328
Atom chips, 264, 303-323 current fluctuations in chip wires,
arrays of traps, 275 327
Bose-Einstein condensate (BEC), 273 dependence on Larmor frequency,
cold atoms on a warm surface, 264 327
decoherence (dephasing) problems, 324 dependence on skin depth, 327
decoherence problems, 335-342 Majorana flips, 325
definition, 335 noise-induced spin flips, 325
fringe contrast, 335 stray light, 329
longitudinal decoherence, 337 thermally induced noise, 326
loss of coherence between internal tunneling, 329
states of a trapped atom, 335 reduction utilizing permanent mag-
motional decoherence, 337 nets, 328
interferometers, 337-342 wire guides, 271
transverse decoherence, 339 Atom optics
design problems, 324-342 applications of miniaturization, 264
experiments, 314-323 atom chips, 264
beam splitters, 319 atom wire, 264
Bose-Einstein condensate (BEC), beam splitters, 276
320-323 guidance of atoms, 264
guiding and transport, 316-318 arrays of traps, 275
microtraps, 314 atom chips, 303-323
fabrication, 304-307 combining magnetic and electric
nanofabrication, 305-307 fields, 286-289
thin-film hybrid production technol- electric interactions, 283-285
ogy, 304 free-standing structures, 292-303
future developments, 342-351 Ioffe-Pritchard trap, 268
guiding and transport, 316-318 Kepler guide, 266, 293
INDEX 393

magnetic interactions, 265-283 Atomic beam splitter using HLB, 175


magnetic traps, 273 Atomic fountain, efficiency improved with
moving potentials, 275 HLB, 176
simple traps, 272 Atomic Hamiltonian
strong-field seekers, 265 dressed-atom model, 158
two-wire guides, 269 kinetic theory of cold atoms in laser
U- and Z-trap, 273 light, 154
weak-field seekers, 266 Axicon
Weinstein-Libbrecht traps, 274 use in dark optical traps, 113
HOF for atomic lens, 169 use in hollow laser beam (HLB), 106
importance for transfer of trapped cold
atoms, 154
B
interferometers made of wire guides,
Beam splitters, 276
278
Ioffe-Pritchard configuration, 277
manipulation of cold atoms in hollow
X-beam splitters, 278
laser beams (HLBs), 153-188
Y-beam splitters, 276
microscopic, 263-351
BEC (Bose-Einstein condensate), 273
miniaturization, 289-292
compression of trap to improve cooling
permanent magnets, 281
rate, 132
Van der Waals interaction, 290
created in CO2 laser trap, 103
Atom-optics billiards, 141
created in optical trap, 131
Atom traps
investigation of movement in billiard,
advantages for scattering experiments,
146
358
on atom chips, 320-323
coherence time of dark optical traps,
preserving coherence in manipulation,
107
171
dark optical traps for cold atoms, 99-
produced by evaporative cooling, 131
147
propagating through orifice made by
depth of potential in single-beam trap,
HLB, 182
119
quantum information processing, 320
difference between red- and blue-
Bessel laser field mode
detuned dipole traps, 103
high-order beams are HLB, 162
general experimental principles, 359
Beta decay, asymmetry in, 243-252
measurement of scattering rates, 117
Blue-detuned traps, s e e Dark optical traps
storage capacity of dark optical traps,
Born approximation in ATI rescattering
108, 114
mechanism, 54
storage time of dark optical traps, 108
Bose-Einstein condensate, s e e BEC
use as target in collision experiments,
Bright optical traps, s e e Red-detuned
358-389
traps
use of HLBs, 178-188
B-splines used in TDSE calculations of
use of light sheets, 178
ATI, 39
volume of single-beam dark optical
traps, 117
Atom wires, 264 C
loss rate depends on angular momentum, Carbohydrates
302 chirality, 227
Mach-Zehnder interferometer, 278 enantiomers, 227
394 INDEX

Carbohydrates (cont 'd) group theory, 230


homochirality in naturally occurring homochirality, 221
carbohydrates, 230 of molecules, 219
CAT (conical atom trap), 181 homochirality of amino acids in cellular
CGH (computer-generated hologram) chemistry, 227
construction, 161 influence on taste, 235
use in dark optical traps, 119 mirror symmetry breaking, advantage
CH4 (methane), symmetry properties, 221 factor, 232
Chaotic motion non-linear production, 254
dependence on wall softness, 145 of biomolecules, 219-257
in intensity-modulated HLB, 146 of electrons and neutrons, 232
of cold atoms in 'gravitational wedge' of fields, 233
billiard, 143 of galaxies, 233
Characterization of high-order harmonics of pharmaceuticals, 235
utilizing ATI, 87 of plants, 234
Charged-particle impact of right and left circularly polarized
ionization mechanisms, 24 light, 232
two-step picture of double ionization, optical activity, 225
24 measurement of, 223
CHBrC1F (bromochlorofluoromethane), wavelength dependence, 223
222 optical activity of chiral molecules, 219
chirality, 222
parity-violating energy difference
symmetry properties, 222
(PVED), 241,252-257
CH3F (fluoromethane)
permanent versus instantaneous, 224
symmetry properties, 222
polymerization of amino acids, 254
Chirality
production of chiral molecules, 224
amino acids, 225
racemization, 256
Murchison meteorite, 237
relation to parity, 231
amplification processes, 247-252
relation to time reversal, 230
autocatalysis, 248-251
kinetic PVED effect, 251 stereogenic centers, 223,225
optically active alcohol, 250 two stereogenic centers, 228
photochemistry, 248 stereoisomers, 223
seeding, 247 thalidomide, 236
archaeological dating, 256 theories of origin, 236-243
Barron's definition, 230 beta decay, 240
beta decay, 243-252 circularly polarized light, 237
carbohydrates, 227, 230 extraterrestrial origin, 237
CHBrC1F, 222 true and false, 230-233
crystal growth, 253 Hund's paradox, 231
definition, 220 of nonstationary objects, 230
degradation, 255 Circularly polarized light
diastereomers, 229 influence on chirality, 237
effects of stirring, 247 natural sources, 239
effects on drugs, 236 Classical orbits, see ATI
electroweak interaction, 240 Cold atom manipulation using HLB, 170-
enantiomers, 223 188
INDEX 395

COLTRIMS (cold-target recoil-ion- compression of scanning wave trap, 132


momentum spectroscopy), 3-6 cooling mechanisms, 127
applied to ATI, 38 dark optical lattices, 111
experimental setup for MPI, 3 definition of "darkness factor", 125
target density for MPI experiments, 4 density-dependent heating, 133, 134
Continuous Stern-Gerlach effect on effect of cooling on phase-space density,
atomic ions, 191-216 128
applied to C 5+, 194 effects of gravity, 127
detection of spin flip, 206-209 evanescent-wave traps, 109
Cooling in magnetic versus optical traps, evaporative cooling, 111, 131
128 generated by multiple laser beams, 106-
Correlated electron momenta in MPI, 20- 113
30 heating and loss mechanisms, 128, 133-
dependence on photon polarization, 20- 136
23 loading atoms into, 103
influence of electron repulsion, 22 loss mechanisms, dependence on
parallel and transverse momentum in density, 134
double ionization of Ar, 22 losses due to light-assisted collisions,
relation between electron and ion recoil 135
momenta, 20 measurement of two-photon transition
S-matrix calculations, 27, 28 in Rb, 140
time-dependent calculations, 28 measurements of weak optical transi-
CR-BPE (concentric-rings binary phase tions, 140
element), use in dark optical traps, multiple light beams added incoherently,
119, 184 107
Cross-correlation normalized properties, 107
ATI applied to measurement of pulse Jr-phase plate trap, 116
duration, 87 polarization-gradient cooling, 128
ATI applied to phase measurement, 88 precision measurements, 136
Cross sections of ground-state hyperfine splitting in
measurements with an atom trap, 357- Na, 136
389 Raman cooling, 129
reflection (Sisyphus) cooling, 130
D scanning single-beam trap, 122
Dark optical traps scattering rate
efficiency of loading single-beam traps, dependence on detuning, 133
119 for CR-BPE trap, 120
application of HLB, 108 for LG HLB trap, 126
applications, 127-147 for scanning beam trap, 124
atom-optics billiards, 141 single beams combined with
based on single laser beams, 113-124 diffractive optical elements, 115
coherence time, 107 refractive optical elements, 113
comparing blue-detuned with red- storage capacity
detuned traps, 126 of axicon trap, 114
comparing different designs, 124 of CR-BPE trap, 120
comparing dynamics in different bil- of evanescent-wave trap, 110
liards, 142 of scanning beam trap, 123, 124
396 INDEX

Dark optical traps (cont'd) measured for He, Ar, Ne and Xe in


storage time MPI, 19
in dark optical lattices, 111 in MPI, 2
of axicon trap, 115 mechanisms, 9
of evanescent-wave trap, 110 rescattering mechanism, 10
of Jr-phase plate trap, 117 in MPI, 25
of scanning beam trap, 123, 124 S-matrix calculations, 14
trap designs using a single laser beam, sequential ionization, 9
113-124 shake-off mechanism, 9
use of atomic trampoline, 109 single-photon absorption, 23
use of axicons, 113, 119 TS 1 (two-step-one) mechanism, 10
volume of TS2 (two-step-two) mechanism, 9
combined CR-BPE/'axicon telescope' Double peak momentum structure in MPI
trap, 122 approximate calculations, 15
CR-BPE trap, 120 consistent with rescattering mechanism
Jr-phase plate trap, 116 in MPI, 18
scanning beam trap, 123 dependence on polarization, 18
single-beam traps, 117,. 119 in CTMC calculations, 18
Dark state, 359 in nonsequential ionization, 11
S-matrix calculations, 14, 15
Density matrix, 335
Wannier theory, 16
time evolution of an atom in a laser
Dressed-atom model
field, 154
density matrix for HLB, 159
Diastereomers
energy levels in HLB, 158
definition, 229
for cold atoms in laser field, 157
properties, 229
Dyson equation, use in ATI, 45, 60
Dipole interaction
between atom and laser field, 154
dipole force in dressed atom model, E
160 EAD (energy-resolved angular distribu-
dipole potential for two-level atom, 102 tion), 71
expression for gradient force, 156 EDM (electric dipole moment)
gradient force in HLB, 156 advantages of measurements in dark
in dressed-atom model, 158 optical traps, 139
momentum transfer, 101 error analysis
multi-level atom, 103 for Cs atoms in a dark optical lattice,
optical potential, 156 138
repulsive force used in dark optical traps, of experiment in dark optical traps,
101 138
rotating-wave approximation, 155 precision measurements
Directional quantization observed as in dark optical traps, 137
Stern-Gerlach effect, 191 using dark optical lattices, 113
Double ionization ee, see Enantiomeric excess
by charged particle impact, 23, 24 Electron mass: continuous Stern-Gerlach
double peak momentum structure, 14 experiment, 214
electron energy distribution Electron scattering in intense laser fields
in MPI, 19 relation to HATI, 55
INDEX 397

Electron-atom collisions extension of measurements to heavier


cross sections for Rb, 357-389 ions, 214
measurement of cross sections with an for hydrogen-like carbon
atom trap, 357-389 calculated, 194
differential scattering, 373 measured, 213
excitation, 383 measured using Stern-Gerlach effect,
ionization, 374, 379 191
total scattering, 370 measurement in Penning trap, 206
Electroweak interaction of free electron, 192
in atoms, 240 Gouy phase shift, 119, 161
in molecules, 240 Guidance of atoms, 264
Z ~ interactions, 242 arrays of traps, 275
Enantiomeric excess (ee), 224, 237 atom chips, 303-323
Enantiomers charged optical fibers, 285
absolute configuration, 227 charged-wire experiments, 300
definition, 225 cold gas, 302
carbohydrates, 227 interferometry, 301
definition, 223 combining magnetic and electric fields,
diastereomers, 229 286-289
Fischer projection, 226 current-carrying wire, 299
properties, 223 electric AC traps, 284
racemic mixture, 223 electric interactions, 283-285
stereogenic centers "electric motor" for ac(de)celeration of
number of, 229 atoms, 288
two stereogenic centers, 228 free-standing structures, 292-303
Evaporative cooling in dark optical traps, Ioffe-Pritchard trap, 268
128, 131 Kepler guide, 266, 293
Excited state fraction, 377 magnetic, 265
magnetic interactions, 265-283
F magnetic traps, 273
FEL (Free Electron Laser) experiments, moving potentials, 275
two-photon double ionization of He, relation between traps and guides, 272
30 state-selective combined electric and
Feynman's path integral, in rescattering magnetic traps, 287
mechanism, 57 storage ring, 300
Fischer projection, 226-228 strong-field seekers, 265
Fokker-Planck equation for cold atoms in tip trap, 298
laser light, 155 two-wire guides, 269
Frequency metrology, impact of phase U- and Z-trap, 273
control of femtosecond laser, 91 weak-field seekers, 266
Frequency standard, use of atomic in magnetic structures, 295, 296
fountain, 176 Weinstein-Libbrecht traps, 274

G
g-factor, 192 It
electron bound in neutral atom, 192 H- photodetachment in constant electric
electron bound to ion, 193 field, 50
398 INDEX

HATI (high-order ATI), s e e u n d e r ATI of cold atoms, 104


Helicity of elementary particles, 221 of Cs, 173
HHG (high-order harmonic generation) of metastable Ne, 105, 172
cutoff, 36 of Rb, 170
in two-color bicircular light, 80 guiding efficiency for cold atoms, 176
experimental facts, 36 horizontal confinement in gravito-optical
in elliptically polarized light, 78 surface trap, 180
in the relativistic regime, 84 LG laser beam as dipole trap, 178
in two-color bicircular light, 78 LG modes, 105
Lewenstein model, 77 methods of generation, 160-169
with linearly polarized light, 77 minimizing loss of atomic coherence,
multiplateau structure, 86 177
origin via rescattering, 50 optical dipole trap made by two HLBs,
plateau, 36 182
in two-color bicircular light, 80 Rabi frequencies in three level model,
quantum orbits, 76 155
relativistic effects on plateau, 76 single HLB used as three-dimensional
relativistic versus non-relativistic results, trap, 183
85 splitting of an atomic beam, 175
saddle-point approximation, 77 storage time in three-dimensional dipole
semiclassical model, 37 trap, 184
sequence of attosecond pulses, 84 storage time of optical dipole trap, 179
HLB (hollow laser beam) storage times in CAT limited by rest gas
axicon used to form three dimensional collisions, 181
trap, 185 theoretical models for cold atoms, 154-
CGH used to form three dimensional 160
trap, 185 three-dimensional trap constructed from
cold atom manipulation, 153, 170-188 rotating laser beam, 186
definition, 104, 160 three-dimensional trap made by two
definition of detuning, 154 Gaussian beams, 185
definitions of laser beam parameters, three-dimensional trap RODiO, 187
156 three-level atom model, 154
dependence of guiding efficiency on transfer efficiency from MOT to CAT,
detuning, 171 181
dougnut shape using HOF, 168 use as atomic trap, 178-188
effect of dipole gradient force, 156 used as two-dimensional trap for cold
efficiency of guiding cold atoms, 171 atoms, 104
efficiency of loading HLB trap, 183 used for creating an atomic fountain,
generated by axicon, 106, 164 176
generated using double-cone prism, 165 used for levitation, 174
generation by phase-plate method, 163 used to produce CAT, 181
generation of Bessel beams, 162 waist size, 156
generation using geometric optics, 163- HOF (hollow-core optical fiber)
165 applications to constructing hollow laser
generation using hologram, 161 beam (HLB), 106, 166-169
guidance cold atom manipulation, 153
of BEC, 105, 173 diffraction pattern, 167
INDEX 399

electric field structure, 167 generated from HG modes, 105, 161


near-field diffraction, 166 generation by hologram, 162
Homochirality, s e e Chirality generation by phase-plate method, 163
intensity distribution, 105
I waist size, 105
Interferometers, 278, 346 LOPT (lowest-order perturbation theory),
decoherence problems, 342 applied to MPI, 36
in the spatial domain, 278 Low-dimensional systems, 347
effect of excitation, 280
in the time domain, 280 M
Ioffe-Pritchard trap, 273 Mach-Zehnder interferometer, 278
field potential, 268 Magnetic moment determination utilizing
relation to two-wire guides, 269 Stern-Gerlach effect, 191
Ionization in strong fields, 40 Magnetic traps
3-dimensional wire, 273
K magnetic quadrupole trap, 298
Keldysh parameter, 41 Magneto-optical trap, s e e MOT
Kepler guide Mechanisms of double ionization in MPI,
comparison with 2-dimensional hydro- 9
gen, 266 Methods for measuring cross sections in
experimental setup, 293 MOT, 367-389
experiments, 293-295 Microscopic trapping potentials, 287
strong-field seeking state, 293 Miniaturization, 290-292
use of strong-field seekers, 266 application in atom "motors" or
KFR (Keldysh-Faisal-Reiss) ionization "conveyor belts", 275
amplitudes in strong-field approxima- finite size effects, 290
tion, 45 limiting factors, 291
Klein-Gordon equation nanofabrication, 289
in relativistic ATI, 75 permanent magnets, 281
quantum orbits in relativistic HHG, 84 Van der Waals interaction, 290
Momentum distribution
L for multiphoton absorption, 6, 8
Larmor frequency, 327 for single photon absorption, 6, 8
Laser field Momentum transfer in interaction between
elliptically polarized, 42 light and atom, 101
Gouy phase shift for Gaussian beams, s e e a l s o Dipole interaction
161 MOT (magneto-optical trap), 292, 298,
HG (Hermite-Gaussian) mode, 160 301
in HLB, 154 atom-chip loading, 307-309
intensity distribution in HLB, 156 basic principles, 359
LG (Laguerre-Gaussian) mode, 160 comparison with crossed-beam experi-
propagation of HG and LG modes, 161 ments, 369
LG (Laguerre-Gaussian) laser field cooling and confining in, 359
modes dark state, 359
angular momentum of, 161, 163 differential cross sections, 373
constituting a HLB, 161 ESD (electron stimulated desorption),
generated from CGH, 105 374
400 INDEX

MOT (magneto-optical trap) (cont'd) target for collision experiments, 360


excitation cross sections for excited timing cycles
levels, 383 for excited-state ionization measure-
experimental setup, 362-367 ments, 380
electron beam, 366 for fluorescence measurements, 384
laser system, 362 for ground-state ionization measure-
magnetic field, 365 ments, 374
avoiding interaction with electron for measuring total cross sections,
beam, 366 370
vacuum chamber, 365 TOF measurements of ions for excited
final-state measurements of cross level cross sections, 379
sections, 376-386 transfer efficiency into CAT, 181
hyperfine repump, 360 MPI (multiphoton ionization)
ionization cross section, 374 comparison to single-photon and
for excited states, 379 charged particle double ionization,
ionization of background gas atoms, 23
381 laser described as classical electromag-
load rates versus loss rates, 371 netic field, 2
loading efficiency of dark optical traps, role of electron correlation, 2
114, 115, 120 two-step picture, 8
loading microscopic guidance structures, use of COLTRIMS, 2
293 Multi-well potentials, 347
loss rate measurements, 368-376 Multilevel atoms: dipole interaction with
measurement of excited state fraction, light beam, 103
377 Multiple ionization in MPI, 2
measurements on mixed ground and Murchison meteorite, 237, 256
excited state population, 379
measuring total cross sections, 370 N
minimization of contribution from Nanofabrication, 289
background ions, 382 atom chips, 303, 305-307
natural time scales, 360 Nonlinear optics applications of ATI, 86
optical molasses, 359 Nonsequential ionization, 2
photoionization experiments, 358 correlated electron momenta, 20
polarization of atoms in trap, 361,387 double peak momentum structure, 11
pyramid MOT, 307 mechanisms for double peak momentum
rate equations involving cross sections, structure, 13
368 NSDI (nonsequential double ionization),
rate loss measurement for excited states, origin via rescattering, 37, 50
376
red-detuning, 359 O
reflection MOT, 307 Optical activity, 219
signal-to-noise ratio in excitation designation of rotation, d and 1 notation,
experiments, 386 225
source of cold atoms, 103, 170, 174, measurement of, using polarimeter, 223
176, 180, 186 origin of, 225
sources of photons in fluorescence Optical billiards
experiments, 384, 385 effect of soft-wall potential, 145
INDEX 401

for cold atoms, 123 PNC (parity nonconservation)


loading scheme for dark optical traps, measurement for Fr in dark optical trap,
145 140
survival probability for cold atoms as tests in dark optical traps, 137
measure of dynamics, 144 Polarimeter for measurement of optical
Optical Bloch equations, 102 activity, 223
Optical bottle beam Polarization-gradient cooling, s e e PGC
HLB, 185 Ponderomotive energy, 41
use as atomic trap, 118 Precision measurements
Optical molasses, 359 ground-state hyperfine splitting in Na
Optical rotary dispersion (ORD), wave- measured in dark optical trap, 136
length dependence of optical activity, limiting accuracy in dark optical traps,
223 136
O R D , s e e Optical rotary dispersion weak optical transitions measurable in
dark optical traps, 140
Precision spectroscopy using dark optical
P traps, 127
Parity-violating energy difference (PVED), PVED, s e e Parity-violating energy differ-
241, 251-257 ence
chirality, 241
experiments, 242 Q
for molecules, 240 QED (quantum electron dynamics)
Z 6 dependence, 242 corrections for C 5+, 195
Penning trap expansion in Za for bound electrons,
calibration of magnetic field, 201 193
detection of spin flip, 206 g-factor of bound electron, 193
double-trap technique, 209-212 g-factor of free particle, 192
error analysis, 212, 213 test of, using magnetic moments, 214
for a single ion, 195 Quantum information processing, 287,
for continuous Stern-Gerlach experi- 348
ment, 195 using BEC, 320
influence of magnetic field inhomo- using dark optical lattices, 113
geneities on motional frequencies, Quantum orbits
205 applied to finite pulses, 78
influence of trap imperfections on in elliptically polarized light, 78
motional frequencies, 205 in HHG versus ATI, 76
ion motion, 197 in two-color bicircular light, 79, 82
ion temperature, 201 interference, 61
loading the trap, 198 quantum-mechanical description of
measurement of oscillation frequencies, HHG and ATI, 37
201 specifying parameters, 61
Permanent magnets, 281 Qubit manipulation, 348
PGC (polarization-gradient cooling) in
dark optical traps, 127, 129 R
Phase-space density, dependence on Racemic mixture, definition, 223
adiabatic change in potential shape, s e e a l s o Chirality
132 Rainbow scattering in ATI, 52, 71
402 INDEX

Raman cooling low energy electrons in ATI of Ar,


applied to dark optical lattices, 130 27
in dark optical traps, 128, 129 Wannier theory, 16
Raman scattering, used to determine in- linear polarization, 50
teraction between atoms and trapping quantum-mechanical description, 53
light in optical trap, 117 saddle-point methods, 55
Ramsey spectroscopy use of Born-approximation, 54
for EDM measurement, 139 use of zero-range potential, 54
for precision measurement of ground- Rescattering model of HHG and ATI, 37
state hyperfine splitting in Na, 136 "Resistive cooling" in Penning traps, 198
Recoil ion momentum ROBOT (rotating beam optical trap), 186
in MPI, 11-19 RODiO (rotating off-resonant dipole
measured for He, Ne and Ar, 11 optical trap), 187
origin of double peak structure, 13-19 storage time, 187
parallel and perpendicular distribu- Rotating wave approximation (RWA), 102
tions in He and Ne, 17
S
saddle potential, 17
S-matrix calculations
S-matrix calculations, 14
are time-independent, 17
integration of one-dimensional
double peak momentum structure, 14
Schr6dinger equation, 18
of correlated electron momenta in MPI,
Red-detuning, 359
27, 28
Reflection cooling
SAE (single-active-electron) approxima-
in dark optical traps, 128
tion
in evanescent-wave trap, 110
applied to ATI, 39
cooling rate, 130
in MPI, 39
unique to blue-detuned optical traps, Scattering length: importance for spin-
130 exchange collisions in atom traps,
Relativistic effects 328
in ATI, 73 Scattering rate: dependence on detuning,
ponderomotive energy, 73-75 133
radiation pressure, 74 Sequential ionization, 2, 9
in HHG, 84 correlated electron momenta, 20
magnetic-field-induced drift, 76 single peak momentum structure, 12
"relativistic effective mass", 75 Side guide
suppression of rescattering effects, 76 experiments, 295
v xB drift in HHG, 86 Side guide, weak-field-seeking state, 267
Rescattering mechanism Simple-man model of HHG and ATI, 37
classical theory in ATI, 50 Simple traps
connection with closed-orbit theory, 58 bent guide wire, 273
in HHG versus ATI, 76 crossed wires, 273
in MPI, 10, 25-27 straight guide wire and bias field, 272
dependence on polarization, 10 U- and Z-trap, 273
double ionization, 25, 26 Single ionization
double peak structure, 13 in MPI, 2
electron energy distribution, 19 momentum distribution in single- and
influence of photon polarization, 26 multiphoton absorption, 6-9
INDEX 403

Single-photon absorption of correlated electron momenta in MPI,


influence of electron repulsion, 23 28-30
Standard Model, tests by tabletop experi- recoil ion momentum distribution, 18
ments, 138 Time-independent calculations; solution
Stereogenic centers, 227 of dressed-atom model for cold atoms
absolute configuration, 230 in laser field, 158
amino acids, 225 TOF (time-of-flight) analysis for ATI
definition, 223 electrons, 38
number of, 229 Two-step picture
two stereogenic centers, 228 of MPI, 8
Stereoisomers, 228 single peak momentum structure, 12
chiral molecules, 223 Two-wire guides, 269
diastereomers, 229
Stern-Gerlach effect, s e e Continuous U
Stern-Gerlach effect U-trap (quadrupole trap), 273
Strong-field approximation in HHG, 77
Strong-field-seeking state, 265, 292, 293, V
328 Van der Waals interaction, 290
Sub-Doppler cooling in MOT, 170 Volkov methods, alternative to TDSE for
ATI, 40
Volkov states for free electron in laser
T field, 44
TDSE (time-dependent Schr6dinger
equation) W
HHG analyzed in terms of quantum Weak-field-seeking state, 266, 267, 292,
orbits, 77 295, 328
numerical solution in one or more Weinstein-Libbrecht traps, 274
dimensions for ATI, 39
solution by split-operator method, 40 Z
Thalidomide, chirality, 236 Z-trap (Ioffe-Pritchard trap), 273
Tilted Bunimovich stadium billiard, 142 Zero-range potential
chaotic motion found in dark optical bound states, 59
trap, 142 calculation of wave function, 60
Time-dependent calculations in ATI rescattering, 54
CTMC (Classical Trajectory Monte in HHG, 78
Carlo) approach, 18
for three-body systems, 18
This Page Intentionally Left Blank
Contents of Volumes in This Serial

Radiofrequency Spectroscopy of Stored


Volume 1
Ions I: Storage, H.G. Dehmelt
Molecular Orbital Theory of the Spin
Properties of Conjugated Molecules, Optical Pumping Methods in Atomic
G.G. Hall and A.T. Amos Spectroscopy, B. Budick
Electron Affinities of Atoms and Energy Transfer in Organic Molecular
Molecules, B.L. Moiseiwitsch Crystals: A Survey of Experiments,
H. C. Wolf
Atomic Rearrangement Collisions,
B.H. Bransden Atomic and Molecular Scattering from
The Production of Rotational and Solid Surfaces, Robert E. Stickney
Vibrational Transitions in Encounters Quantum, Mechanics in Gas Crystal-
between Molecules, K. Takayanagi Surface van der Waals Scattering,
The Study of Intermolecular Potentials E. Chanoch Beder
with Molecular Beams at Thermal Reactive Collisions between Gas and
Energies, H. Pauly and J.P Toennies Surface Atoms, Henry Wise and
Bernard J. Wood
High-Intensity and High-Energy
Molecular Beams, J.B. Anderson,
R.P Andres and J.B. Fen
Volume 4
Volume 2
H.S.W M a s s e y - A Sixtieth Birthday
The Calculation of van der Waals Interac-
Tribute, E.H.S. Burhop
tions, A. Dalgarno and W.D. Davison
Electronic Eigenenergies of the Hydrogen
Thermal Diffusion in Gases, E.A. Mason,
Molecular Ion, D.R. Bates and
R.J Munn and Francis J. Smith
R.H. G. Reid
Spectroscopy in the Vacuum Ultraviolet,
Applications of Quantum Theory
W.R.S. Garton
to the Viscosity of Dilute Gases,
The Measurement of the Photoionization R.A. Buckingham and E. Gal
Cross Sections of the Atomic Gases,
James A.R. Samson Positrons and Positronium in Gases,
PA. Fraser
The Theory of Electron-Atom Collisions,
R. Peterkop and V. Veldre Classical Theory of Atomic Scattering,
A. Burgess and I.C. Percival
Experimental Studies of Excitation in
Collisions between Atomic and Ionic Born Expansions, A.R. Holt and
Systems, EJ. de Heer B.L. Moiseiwitsch
Resonances in Electron Scattering by
Mass Spectrometry of Free Radicals,
Atoms and Molecules, P G. Burke
S.N. Foner
Relativistic Inner Shell Ionizations,
Volume 3 C.B. 0. Mohr
The Quantal Calculation of Photoioniza- Recent Measurements on Charge Transfer,
tion Cross Sections, A.L. Stewart J.B. Hasted

405
406 C O N T E N T S OF V O L U M E S IN THIS SERIAL

Measurements of Electron Excitation The Diffusion of Atoms and Molecules,


Functions, D. W.O. Heddle and R. G. W E.A. Mason and T.R. Marrero
Keesing Theory and Application of Sturmian
Some New Experimental Methods in Functions, Manuel Rotenberg
Collision Physics, R.F. Stebbings Use of Classical mechanics in the
Atomic Collision Processes in Gaseous Treatment of Collisions between Massive
Nebulae, M.J. Seaton Systems, D.R. Bates and A.E. Kingston
Collisions in the Ionosphere, A. Dalgarno
Volume 7
The Direct Study of Ionization in Space,
R.L.E Boyd Physics of the Hydrogen Maser, C. Audoin,
JP. Schermann and P. Grivet
Molecular Wave Functions: Calculations
Volume 5 and Use in Atomic and Molecular
Flowing Afterglow Measurements of Processes, J. C. Browne
Ion-Neutral Reactions, E.E. Ferguson, Localized Molecular Orbitals, Harel
EC. Fehsenfeld and A.L. Schmeltekopf Weinstein, Ruben Pauncz and Maurice
Experiments with Merging Beams, Cohen
Roy H. Neynaber General Theory of Spin-Coupled Wave
Radiofrequency Spectroscopy of Stored Functions for Atoms and Molecules,
Ions II: Spectroscopy, H.G. Dehmelt J. Gerratt
Diabatic States of Molecules -
The Spectra of Molecular Solids,
Quasi-Stationary Electronic States,
0. Schnepp
Thomas EO'Malley
The Meaning of Collision Broadening of
Selection Rules within Atomic Shells,
Spectral Lines: The Classical Oscillator
B.R. Judd
Analog, A. Ben-Reuven
Green's Function Technique in Atomic
The Calculation of Atomic Transition and Molecular Physics, Gy. Csanak,
Probabilities, R.JS. Crossley H.S. Taylor and Robert Yaris
Tables of One- and Two-Particle Coeffi- A Review of Pseudo- Potentials with
cients of Fractional Parentage for Con- Emphasis on Their Application to
figurations sZs'Upq, C.D.H. Chisholm, Liquid Metals, Nathan Wiser and
A. Dalgarno and ER. Innes A.J. Greenfield
Relativistic Z-Dependent Corrections to
Atomic Energy Levels, Holly Thomis Volume 8
Doyle
Interstellar Molecules: Their Formation
and Destruction, D. McNally
Volume 6 Monte Carlo Trajectory Calculations of
Dissociative Recombination, JN. Bardsley Atomic and Molecular Excitation in
and M.A. Biondi Thermal Systems, James C. Keck
Analysis of the Velocity Field in Plasmas Nonrelativistic Off-Shell Two-Body
from the Doppler Broadening of Coulomb Amplitudes, Joseph C.Y. Chen
Spectral Emission Lines, A.S. Kaufman and Augustine C. Chen
The Rotational Excitation of Molecules Photoionization with Molecular Beams,
by Slow Electrons, Kazuo Takayanagi R.B. Cairns, Halstead Harrison and
and Yukikazu Itikawa R.I. Schoen
CONTENTS OF VOLUMES IN THIS SERIAL 407

The Auger Effect, E.H.S. Burhop and Role of Energy in Reactive Molecular
W.N. Asaad Scattering: An Information-Theoretic
Approach, R.B. Bernstein and
R.D. Levine
Volume 9
Inner Shell Ionization by incident Nuclei,
Correlation in Excited States of Atoms, Johannes M. Hansteen
A. W. Weiss
Stark Broadening, Hans R. Griem
The Calculation of Electron-Atom Exci-
tation Cross Sections, M.R.H. Rudge Chemiluminescence in Gases, M.E Golde
and B.A. Thrush
Collision-Induced Transitions between
Rotational Levels, Takeshi Oka Volume 12
The Differential Cross Section of Nonadiabatic Transitions between Ionic
Low-Energy Electron-Atom Collisions, and Covalent States, R.K. Janev
D. Andrick
Recent Progress in the Theory of
Molecular Beam Electric Resonance
Atomic Isotope Shift, J. Bauche and
Spectroscopy, Jens C. Zorn and
R.-J. Champeau
Thomas C. English
Topics on Multiphoton Processes in
Atomic and Molecular Processes
Atoms, P Lambropoulos
in the Martian Atmosphere,
Michael B. McElroy Optical Pumping of Molecules, M. Broyer,
G. Goudedard, J.C. Lehmann and
J Vigu~
Volume 10
Highly Ionized Ions, Ivan A. Sellin
Relativistic Effects in the Many-Electron
Time-of-Flight Scattering Spectroscopy,
Atom, Lloyd Armstrong Jr. and Serge
Wilhelm Raith
Feneuille
Ion Chemistry in the D Region,
The First Born Approximation, K.L. Bell
George C. Reid
and A.E. Kingston
Photoelectron Spectroscopy, W.C. Price Volume 13
Dye Lasers in Atomic Spectroscopy, Atomic and Molecular Polarizabilities-
W. Lange, J. Luther and A. Steudel Review of Recent Advances,
Recent Progress in the Classification of Thomas M. Miller and Benjamin
the Spectra of Highly Ionized Atoms, Bederson
B. C Fawcett Study of Collisions by Laser Spectroscopy,
A Review of Jovian Ionospheric Paul R. Berman
Chemistry, Wesley T. Huntress Jr. Collision Experiments with Laser-Excited
Atoms in Crossed Beams, I. V. Hertel
and W. Stoll
Volume 11 Scattering Studies of Rotational and
The Theory of Collisions between Vibrational Excitation of Molecules,
Charged Particles and Highly Excited Manfred Faubel and J. Peter Toennies
Atoms, I.C. Percival and D. Richards Low-Energy Electron Scattering
Electron Impact Excitation of Positive by Complex Atoms: Theory and
Ions, M.J Seaton Calculations, R.K Nesbet
The R-Matrix Theory of Atomic Process, Microwave Transitions of Interstellar
P G. Burke and W.D. Robb Atoms and Molecules, WB. Somerville
408 CONTENTS OF VOLUMES IN THIS SERIAL

Atomic Collision Processes in Controlled


Volume 14
Thermonuclear Fusion Research,
Resonances in Electron Atom and H.B. Gilbody
Molecule Scattering, D.E. Golden
Inner-Shell Ionization, E.H.S. Burhop
The Accurate Calculation of Atomic
Properties by Numerical Methods, Excitation of Atoms by Electron Impact,
Brian C. Webster, Michael J. Jamieson D. W.O. Heddle
and Ronald F. Stewart Coherence and Correlation in Atomic
(e, 2e) Collisions, Erich Weigold and Collisions, H. Kleinpoppen
Ian E. McCarthy Theory of Low Energy Electron-Molecule
Forbidden Transitions in One- and Collisions, PG. Burke
Two-Electron Atoms, Richard Marrus
and Peter J. Mohr Volume 16
Semiclassical Effects in Heavy-Particle
Atomic Hartree-Fock Theory, M. Cohen
Collisions, M.S. Child
and R.P. McEachran
Atomic Physics Tests of the Basic Experiments and Model Calculations
Concepts in Quantum Mechanics, to Determine Interatomic Potentials,
Francis M. Pipkin R. Diiren
Quasi-Molecular Interference Effects in Sources of Polarized Electrons,
Ion-Atom Collisions, S. V. Bobashev R.J. Celotta and D. T. Pierce
Rydberg Atoms, S.A. Edelstein and Theory of Atomic Processes in Strong
T.E Gallagher Resonant Electromagnetic Fields,
UV and X-Ray Spectroscopy in S. Swain
Astrophysics, A.K. Dupree Spectroscopy of Laser-Produced Plasmas,
M.H. Key and R.J. Hutcheon
Relativistic Effects in Atomic Collisions
Volume 15 Theory, B.L. Moiseiwitsch
Parity Nonconservation in Atoms: Status
Negative Ions, H.S.W. Massey
of Theory and Experiment, E.N. Fortson
Atomic Physics from Atmospheric and and L. Wilets
Astrophysical Studies, A. Dalgarno
Collisions of Highly Excited Atoms,
Volume 17
R.E Stebbings
Collective Effects in Photoionization of
Theoretical Aspects of Positron Collisions
Atoms, M. Ya. Amusia
in Gases, J. W. Humberston
Nonadiabatic Charge Transfer,
Experimental Aspects of Positron
D.S.E Crothers
Collisions in Gases, T.C. Griffith
Atomic Rydberg States, Serge Feneuille
Reactive Scattering: Recent Advances and Pierre Jacquinot
in Theory and Experiment,
Richard B. Bernstein Superfluorescence, M.F.H. Schuurmans,
Q H . E Vrehen, D. Polder and
Ion-Atom Charge Transfer Collisions at
H.M. Gibbs
Low Energies, J.B. Hasted
Applications of Resonance Ionization
Aspects of Recombination, D.R. Bates Spectroscopy in Atomic and Molecular
The Theory of Fast Heavy Particle Physics, M.G. Payne, C.H. Chen,
Collisions, B.H. Bransden G.S. Hurst and G. W. Foltz
CONTENTS OF VOLUMES IN THIS SERIAL 409

Inner-Shell Vacancy Production in The Reduced Potential Curve Method


Ion-Atom Collisions, C.D. Lin and for Diatomic Molecules and Its
Patrick Richard Applications, E Jen6
Atomic Processes in the Sun, PL. Dufion The Vibrational Excitation of Molecules
and A.E. Kingston by Electron Impact, D.G. Thompson
Vibrational and Rotational Excitation in
Molecular Collisions, Manfred Faubel
Volume 18
Spin Polarization of Atomic and Molecular
Theory of Electron-Atom Scattering in a
Photoelectrons, N.A. Cherepkov
Radiation Field, Leonard Rosenberg
Positron-Gas Scattering Experiments,
Talbert S. Stein and Walter E. Kauppila Volume 20
Nonresonant Multiphoton Ionization Ion-Ion Recombination in an Ambient
of Atoms, J.. Morellec, D. Normand Gas, D.R. Bates
and G. Petite Atomic Charges within Molecules,
Classical and Semiclassical Methods G. G. Hall
in Inelastic Heavy-Particle Collisions, Experimental Studies on Cluster Ions, T
A.S. Dickinson and D. Richards D. Mark and A. W. Castleman Jr.
Recent Computational Developments Nuclear Reaction Effects on Atomic
in the Use of Complex Scaling in Inner-Shell Ionization, WE. Meyerhof
Resonance Phenomena, B.R. Junker and J.-F Chemin
Direct Excitation in Atomic Collisions: Numerical Calculations on Electron-
Studies of Quasi-One-Electron Systems, Impact Ionization, Christopher
N. Andersen and S.E. Nielsen Bottcher
Model Potentials in Atomic Structure, Electron and Ion Mobilities,
A. Hibbert Gordon R. Freeman and
Recent Developments in the Theory David A. Armstrong
of Electron Scattering by Highly On the Problem of Extreme UV and
Polar Molecules, D.W. Norcross and X-Ray Lasers, I.I. Sobel'man and
L.A. Collins A. V. Vinogradov
Quantum Electrodynamic Effects in Few-
Radiative Properties of Rydberg States
Electron Atomic Systems, G. W.E Drake
in Resonant Cavities, S. Haroche and
J.M. Raimond
Rydberg Atoms: High-Resolution
Volume 19
Spectroscopy and Radiation Interaction-
Electron Capture in Collisions of Rydberg Molecules, J.A. C. Gallas,
Hydrogen Atoms with Fully Stripped G. Leuchs, H. Walther, and H. Figger
Ions, B.H. Bransden and R.K. Janev
Interactions of Simple Ion Atom Systems,
JT. Park Volume 21
High-Resolution Spectroscopy of Stored Subnatural Linewidths in Atomic
Ions, D.J. Wineland, Wayne M. Itano Spectroscopy, Dennis P O'Brien, Pierre
and R.S Van Dyck Jr. Meystre and Herbert Walther
Spin-Dependent Phenomena in Inelastic Molecular Applications of Quantum
Electron-Atom Collisions, K. Blum and Defect Theory, Chris H. Greene and
H. Kleinpoppen Ch.Jungen
410 CONTENTS OF VOLUMES IN THIS SERIAL

Theory of Dielectronic Recombination, Photoionization and Collisional Ionization


Yukap Hahn of Excited Atoms Using Synchroton
Recent Developments in Semiclassical and Laser Radiation, F.J. Wuilleumier,
Floquet Theories for Intense-Field D.L. Ederer and J.L. Picqu~
Multiphoton Processes, Shih-I Chu
Scattering in Strong Magnetic Fields, Volume 24
M.R.C. McDowell and M. Zarcone The Selected Ion Flow Tube (SIDT):
Pressure Ionization, Resonances and the Studies of Ion-Neutral Reactions,
Continuity of Bound and Free States, D. Smith and N.G. Adams
R.M. More Near-Threshold Electron-Molecule
Scattering, Michael A. Morrison
Volume 22 Angular Correlation in Multiphoton
Positronium- Its Formation and Ionization of Atoms, S.J. Smith and
Interaction with Simple Systems, G. Leuchs
2 W. Humberston Optical Pumping and Spin Exchange
Experimental Aspects of Positron and in Gas Cells, R.J. Knize, Z. Wu and
Positronium Physics, T. C. Griffith W Happer
Doubly Excited States, Including New Correlations in Electron-Atom Scattering,
Classification Schemes, C.D. Lin A. Crowe
Measurements of Charge Transfer and
Ionization in Collisions Involving Volume 25
Hydrogen Atoms, H.B. Gilbody Alexander Dalgarno: Life and Personality,
Electron Ion and Ion-Ion Collisions David R. Bates and George A. Victor
with Intersecting Beams, K. Dolder Alexander Dalgarno: Contributions to
and B. Peart Atomic and Molecular Physics, Neal
Electron Capture by Simple Ions, Edward Lane
Pollack and Yukap Hahn Alexander Dalgarno: Contributions to
Relativistic Heavy-Ion-Atom Collisions, Aeronomy, Michael B. McElroy
R. Anholt and Harvey Gould Alexander Dalgarno: Contributions to
Continued-Fraction Methods in Atomic Astrophysics, David A. Williams
Physics, S. Swain Dipole Polarizability Measurements,
Thomas M. Miller and Benjamin
Volume 23 Bederson
Vacuum Ultraviolet Laser Spectroscopy Flow Tube Studies of Ion-Molecule
of Small Molecules, C.R. Vidal Reactions, Eldon Ferguson
Foundations of the Relativistic Theory Differential Scattering in He-He and
of Atomic and Molecular Structure, He+-He Collisions at keV Energies,
Ian P. Grant and Harry M. Quiney R.E Stebbings
Point-Charge Models for Molecules Atomic Excitation in Dense Plasmas,
Derived from Least-Squares Fitting of Jon C. Weisheit
the Electric Potential, D.E. Williams Pressure Broadening and Laser-induced
and Ji-Min Yan Spectral Line Shapes, Kenneth M. Sando
Transition Arrays in the Spectra of Ionized and Shih-I Chu
Atoms, J. Bauche, C. Bauche-Arnoult Model-Potential Methods, C. Laughlin
and M. Klapisch and G.A. Victor
CONTENTS OF VOLUMES IN THIS SERIAL 411

Z-Expansion Methods, M. Cohen On the [3 Decay of 187Re: An Interface


of Atomic and Nuclear Physics and
Schwinger Variational Methods, Deborah
Cosmochronology, Zonghau Chen,
Kay Watson
Leonard Rosenberg and Larry Spruch
Fine-Structure Transitions in Proton-Ion
Progress in Low Pressure Mercury-Rare
Collisions, R.H. G. Reid
Gas Discharge Research, J. Maya and
Electron Impact Excitation, R.J.W. Henry R. Lagushenko
and A.E. Kingston
Recent Advances in the Numerical Volume 27
Calculation of Ionization Amplitudes, Negative Ions: Structure and Spectra,
Christopher Bottcher David R. Bates
The Numerical Solution of the Equations Electron Polarization Phenomena in
of Molecular Scattering, A.C. Allison Electron-Atom Collisions, Joachim
High Energy Charge Transfer, Kessler
B.H. Bransden and D.P. Dewangan Electron-Atom Scattering, I.E. McCarthy
Relativistic Random-Phase and E. Weigold
Approximation, WR. Johnson Electron-Atom Ionization, I.E. McCarthy
Relativistic Sturmian and Finite Basis and E. Weigold
Set Methods in Atomic Physics, Role of Autoionizing States in
G. W.E Drake and S.P. Goldman Multiphoton Ionization of Complex
Dissociation Dynamics of Polyatomic Atoms, V.I. Lengyel and M.I. Haysak
Molecules, T. Uzer Multiphoton Ionization of Atomic
Photodissociation Processes in Diatomic Hydrogen Using Perturbation Theory,
Molecules of Astrophysical Interest, E. Karule
Kate P. Kirby and Ewine E van
Dishoeck Volume 28
The Abundances and Excitation of The Theory of Fast Ion-Atom Collisions,
Interstellar Molecules, John H. Black J.S. Briggs and J.H. Macek
Some Recent Developments in the
Fundamental Theory of Light,
Volume 26 Peter W. Milonni and Surendra Singh
Comparisons of Positrons and Electron Squeezed States of the Radiation Field,
Scattering by Gases, Waiter E. Kauppila Khalid Zaheer and M. Suhail Zubairy
and Talbert S. Stein Cavity Quantum Electrodynamics,
Electron Capture at Relativistic Energies, E.A. Hinds
B.L. Moiseiwitsch
The Low-Energy, Heavy Particle Colli- Volume 29
sions- A Close-Coupling Treatment, Studies of Electron Excitation of Rare-Gas
Mineo Kimura and Neal E Lane Atoms into and out of Metastable Levels
Vibronic Phenomena in Collisions of Using Optical and Laser Techniques,
Atomic and Molecular Species, V.Sidis Chun C. Lin and L. W Anderson
Associative Ionization: Experiments, Cross Sections for Direct Multiphoton
Potentials and Dynamics, John Weiner, Ionionization of Atoms, M. V. Ammosov,
Franfoise Masnou-Seeuws and Annick N.B. Delone, M. Yu. Ivanov, I.L Bondar
Giusti-Suzor and A. V. Masalov
412 CONTENTS OF VOLUMES IN THIS SERIAL

Collision-Induced Coherences in Optical Positronium Formation by Positron Impact


Physics, G.S. Agarwal on Atoms at Intermediate Energies,
Muon-Catalyzed Fusion, Johann Rafelski B.H. Bransden and C.J. Noble
and Helga E. Rafelski Electron-Atom Scattering Theory and
Calculations, PG. Burke
Cooperative Effects in Atomic Physics,
J.P Connerade Terrestrial and Extraterrestrial H~-,
Multiple Electron Excitation, Ionization, Alexander Dalgarno
and Transfer in High-Velocity Atomic Indirect Ionization of Positive Atomic
and Molecular Collisions, J.H. McGuire Ions, K. Dolder
Quantum Defect Theory and Analysis of
Volume 30 High-Precision Helium Term Energies,
Differential Cross Sections for Excitation G. W.E Drake
of Helium Atoms and Helium-Like Ions Electron-Ion and Ion-Ion Recombination
by Electron Impact, Shinobu Nakazaki Processes, M.R. Flannery
Cross-Section Measurements for Electron Studies of State-Selective Electron Capture
Impact on Excited Atomic Species, in Atomic Hydrogen by Translational
S. Trajmar and J C. Nickel Energy Spectroscopy, H.B. Gilbody
The Dissociative Ionization of Relativistic Electronic Structure of Atoms
Simple Molecules by Fast Ions, and Molecules, I.P Grant
Colin J Latimer
The Chemistry of Stellar Environments,
Theory of Collisions between Laser D.A. Howe, J.M.C. Rawlings and
Cooled Atoms, PS. Julienne, D.A. Williams
A.M. Smith, and K. Burnett
Positron and Positronium Scattering at
Light-Induced Drift, E.R. Eliel Low Energies, J. W. Humberston
Continuum Distorted Wave Methods How Perfect are Complete Atomic
in Ion-Atom Collisions, Derrick Collision Experiments?, H. Kleinpoppen
S.E Crothers and Louis J. Dub~ and H. Handy
Adiabatic Expansions and Nonadiabatic
Volume 31
Effects, R. McCarroll and D.S.F
Energies and Asymptotic Analysis for Crothers
Helium Rydberg States, G.W.E Drake Electron Capture to the Continuum,
Spectroscopy of Trapped Ions, B.L. Moiseiwitsch
R. C. Thompson
How Opaque Is a Star?, M.J. Seaton
Phase Transitions of Stored Laser-Cooled
Studies of Electron Attachment at
Ions, H. Walther
Thermal Energies Using the Flowing
Selection of Electronic States in Atomic Afterglow-Langmuir Technique, David
Beams with Lasers, Jacques Baudon, Smith and Patrik Span~l
Rudolf Diiren and Jacques Robert
Exact and Approximate Rate Equations in
Atomic Physics and Non-Maxwellian Atom-Field Interactions, S. Swain
Plasmas, MichOle Lamoureux
Atoms in Cavities and Traps, H. Walther
Volume 32 Some Recent Advances in Electron-
Photoionization of Atomic Oxygen Impact Excitation of n = 3 States
and Atomic Nitrogen, K.L. Bell and of Atomic Hydrogen and Helium,
A.E. Kingston J.E Williams and J.B. Wang
C O N T E N T S OF V O L U M E S IN THIS SERIAL 413

Classical and Quantum Chaos in Atomic


Volume 33
Systems, Dominique Delande and
Principles and Methods for Measurement Andreas Buchleitner
of Electron Impact Excitation Cross
Measurements of Collisions between
Sections for Atoms and Molecules by
Laser-Cooled Atoms, Thad Walker
Optical Techniques, A.R. FilippellL
and Paul Feng
Chun C. Lin, L.W. Andersen and
J. W. McConkey The Measurement and Analysis of Electric
Fields in Glow Discharge Plasmas,
Benchmark Measurements of Cross
J.E. Lawler and D.A. Doughty
Sections for Electron Collisions:
Analysis of Scattered Electrons, Polarization and Orientation Phenomena
S. Trajmar and J. W. McConkey in Photoionization of Molecules,
N.A. Cherepkov
Benchmark Measurements of Cross Sec-
tions for Electron Collisions: Electron Role of Two-Center Electron-Electron
Swarm Methods, R.W. Crompton Interaction in Projectile Electron
Excitation and Loss, E.C. Montenegro,
Some Benchmark Measurements of Cross
WE. Meyerhof and J.H. McGuire
Sections for Collisions of Simple Heavy
Particles, H.B. Gilbody Indirect Processes in Electron Impact
Ionization of Positive Ions, D.L. Moores
The Role of Theory in the Evaluation and
and K.J. Reed
Interpretation of Cross-Section Data,
Dissociative Recombination: Crossing and
Barry I. Schneider
Tunneling Modes, David R. Bates
Analytic Representation of Cross-Section
Data, Mitio Inokuti, Mineo Kimura,
M.A. Dillon, Isao Shimamura
Electron Collisions with N2, O2 and Volume 35
O: What We Do and Do Not Know,
Laser Manipulation of Atoms,
Yukikazu Itikawa
K. Sengstock and W. Ertmer
Need for Cross Sections in Fusion Plasma
Research, Hugh P Summers Advances in Ultracold Collisions:
Experiment and Theory, J. Weiner
Need for Cross Sections in Plasma
Chemistry, M. Capitelli, R. Celiberto Ionization Dynamics in Strong Laser
and M. Cacciatore Fields, L.E DiMauro and P Agostini
Guide for Users of Data Resources, Infrared Spectroscopy of Size Selected
Jean W. Gallagher Molecular Clusters, U Buck
Guide to Bibliographies, Books, Reviews Fermosecond Spectroscopy of Molecules
and Compendia of Data on Atomic Col- and Clusters, T. Baumer and G. Gerber
lisions, E. W McDaniel and E.J. Mansky Calculation of Electron Scattering on
Hydrogenic Targets, I. Bray and
A. T. Stelbovics
Relativistic Calculations of Transition
Volume 34 Amplitudes in the Helium Isoelectronic
Atom Interferometry, C.S. Adams, Sequence, W.R. Johnson, D.R. Plante
O. Carnal and J. Mlynek and J Sapirstein
Optical Tests of Quantum Mechanics, Rotational Energy Transfer in Small
R.Y. Chiao, P G. Kwiat and Polyatomic Molecules, H.O. Everitt
A.M. Steinberg and F.C. De Lucia
414 CONTENTS OF VOLUMES IN THIS SERIAL

Volume 36
Rydberg Ionization: From Field to Photon,
G.M. Lankhuijzen and L.D. Noordam
Complete Experiments in Electron-Atom
Collisions, Nils Overgaard Andersen Studies of Negative Ions in Storage
and Klaus Bartschat Rings, L.H. Andersen, T. Andersen
and P. Hvelplund
Stimulated Rayleigh Resonances and
Recoil-induced Effects, J.-Y Courtois Single-Molecule Spectroscopy and
and G. Grynberg Quantum Optics in Solids, WE. Moerner,
R.M. Dickson and D.J. Norris
Precision Laser Spectroscopy Using
Acousto-Optic Modulators, W.A. van
Volume 39
Wo'ngaarden
Author and Subject Cumulative Index
Highly Parallel Computational Techniques
Volumes 1-38
for Electron-Molecule Collisions, Carl
Winstead and Vincent McKoy Author Index
Quantum Field Theory of Atoms and Subject Index
Photons, Maciej Lewenstein and Li You Appendix: Tables of Contents of Volumes
1-38 and Supplements
Volume 37
Evanescent Light-Wave Atom Mirrors, Volume 40
Resonators, Waveguides, and Traps, Electric Dipole Moments of Leptons,
Jonathan P. Dowling and Julio Eugene D. Commins
Gea-Banacloche
High-Precision Calculations for the
Optical Lattices, P.S. Jessen and Ground and Excited States of the
I.H. Deutsch Lithium Atom, Frederick W. King
Channeling Heavy Ions through
Storage Ring Laser Spectroscopy,
Crystalline Lattices, Herbert E Krause
Thomas U. Kiihl
and Sheldon Datz
Laser Cooling of Solids, Carl E. Mungan
Evaporative Cooling of Trapped Atoms,
and Timothy R. Gosnell
Wolfgang Ketterle and N.J. van Druten
Optical Pattern Formation, L.A. Lugiato,
Nonclassical States of Motion in Ion
M. Brambilla and A. Gatti
Traps, J.L Cirac, A.S. Parkins, R. Blatt
and P. Zoller
Volume 41
The Physics of Highly-Charged Heavy
Ions Revealed by Storage/Cooler Rings, Two-Photon Entanglement and Quantum
P.H. Mokler and Th. St6hlker Reality, Yanhua Shih
Quantum Chaos with Cold Atoms,
Volume 38 Mark G. Raizen
Electronic Wavepackets, Robert R. Jones Study of the Spatial and Temporal
and L.D. Noordam Coherence of High-Order Harmonics,
Pascal Sali~res, Ann L'Huillier, Philippe
Chiral Effects in Electron Scattering
by Molecules, K. Blum and Antoine and Maciej Lewenstein
D. G. Thompson Atom Optics in Quantized Light
Optical and Magneto-Optical Fields, Matthias Freyburger,
Alois M. Herkommer, Daniel S. Kriihmer,
Spectroscopy of Point Defects in
Erwin Mayr and Wolfgang P. Schleich
Condensed Helium, Serguei I. Kanorsky
and Antoine Weis Atom Waveguides, Victor I. Balykin
CONTENTS OF VOLUMES IN THIS SERIAL 415

Atomic Matter Wave Amplification by Electron Impact Ionization of Organic


Optical Pumping, Ulf Janicke and Silicon Compounds, Ralf Basner, Kurt
Martin Wilkens Becker, Hans Deutsch and Martin
Schmidt
Volume 42 Kinetic Energy Dependence of Ion-
Molecule Reactions Related to Plasma
Fundamental Tests of Quantum Mechan-
Chemistry, PB. Armentrout
ics, Edward S. Fry and Thomas Walther
Physicochemical Aspects of Atomic
Wave-Particle Duality in an Atom
and Molecular Processes in Reactive
Interferometer, Stephan Diirr and
Plasmas, Yoshihiko Hatano
Gerhard Rempe
Ion-Molecule Reactions, Werner
Atom Holography, Fujio Shimizu Lindinger, Armin Hansel and Zdenek
Optical Dipole Traps for Neutral Atoms, Herman
Rudolf Grimm, Matthias Weidemiiller Uses of High-Sensitivity White-Light
and Yurii B. Ovchinnikov Absorption Spectroscopy in Chemical
Formation of Cold (T ~< 1 K) Vapor Deposition and Plasma
Molecules, J.T. Bahns, PL. Gould and Processing, L. W. Anderson, A.N. Goyette
W.C. Stwalley and J.E. Lawler
High-intensity Laser-Atom Physics, Fundamental Processes of Plasma-Surface
C.J. Joachain, M. Dorr and N.J Kylstra Interactions, Rainer Hippler
Coherent Control of Atomic, Molecular Recent Applications of Gaseous
and Electronic Processes, Moshe Discharges: Dusty Plasmas and Upward-
Shapiro and Paul Brumer Directed Lightning, Ara Chutjian
Resonant Nonlinear Optics in Phase Opportunities and Challenges for Atomic,
Coherent Media, M.D. Lukin, P Hemmer Molecular and Optical Physics in
and M. O. Scully Plasma Chemistry, Kurl Becker, Hans
Deutsch and Mitio Inokuti
The Characterization of Liquid and
Solid Surfaces with Metastable Helium
Atoms, H. Morgner
Quantum Communication with Entangled Volume 44
Photons, Harald Weinfurter Mechanisms of Electron Transport in
Electrical Discharges and Electron
Volume 43 Collision Cross Sections, Hiroshi
Plasma Processing of Materials and Tanaka and Osamu Sueoka
Atomic, Molecular, and Optical Physics: Theoretical Consideration of Plasma-
An Introduction, Hiroshi Tanaka and Processing Processes, Mineo Kimura
Mitio Inokuti Electron Collision Data for
The Boltzmann Equation and Transport Plasma-Processing Gases,
Coefficients of Electrons in Weakly Loucas G. Christophorou and
Ionized Plasmas, R. Winkler James K. Olthoff
Electron Collision Data for Plasma Radical Measurements in Plasma
Chemistry Modeling, W.L. Morgan Processing, Toshio Goto
Electron-Molecule Collisions in Low- Radio-Frequency Plasma Modeling for
Temperature Plasmas: The role of The- Low-Temperature Processing, Toshiaki
ory, Carl Winstead and Vincent McKoy Makabe
416 CONTENTS OF VOLUMES IN THIS SERIAL

Electron Interactions with Excited Atoms Molecular Emissions from the


and Molecules, Loucas G. Christophorou Atmospheres of Giant Planets and
and James K. Olthoff Comets: Needs for Spectroscopic and
Collision Data, Yukikazu Itikawa, Sang
Joon Kim, Yong Ha Kim, and Y C. Minh
Volume 45
Studies of Electron-Excited Targets
Comparing the Antiproton and Proton,
Using Recoil Momentum Spectroscopy
and Opening the Way to Cold
with Laser Probing of the Excited
Antihydrogen, G. Gabrielse
State, Andrew James Murray and Peter
Medical Imaging with Laser-Polarized Hammond
Noble Gases, Timothy Chupp and Quantum Noise of Small Lasers,
Scott Swanson J.P Woerdman, N.J van Druten and
Polarization and Coherence Analysis of M.P. van Exter
the Optical Two-Photon Radiation from
the Metastable 22S1/2 State of Atomic Volume 48
Hydrogen, Alan J Duncan, Hans
Multiple Ionization in Strong Laser Fields,
Kleinpoppen and Marian O. Scully
R. DSrner, Th. Weber, M. Weckenbrock,
Laser Spectroscopy of Small Molecules, A. Staudte, M. Hattass, R. Moshammer,
W. DemtrSder, M. Keil and H. Wenz J. Ullrich, H. Schmidt-BScking
Coulomb Explosion Imaging of Above-Threshold Ionization: From
Molecules, Z Vager Classical Features to Quantum Effects,
W. Becker, F. Grasbon, R. Kopold,
Volume 46 D.B. Milo~eviO, G.G. Paulus and
H. Walther
Femtosecond Quantum Control, T Brixner,
N.H. Damrauer and G. Gerber Dark Optical Traps for Cold Atoms,
Nir Friedman, Ariel Kaplan and Nir
Coherent Manipulation of Atoms
Davidson
and Molecules by Sequential Laser
Pulses, N.V. Vitanov, M. Fleischhauer, Manipulation of Cold Atoms in Hollow
B. W. Shore and K. Bergmann Laser Beams, Heung-Ryoul Noh, Xinye
Xu and Wonho Jhe
Slow, Ultraslow, Stored, and Frozen Light,
Continuous Stern-Gerlach Effect on
Andrey B. Matsko, Olga Kocharovskaya,
Atomic Ions, Giinther Werth, Hartmut
Yuri Rostovtsev, George R. Welch,
Hdffner and Wolfgang Quint
Alexander S. Zibrov and Marlan O.
Scully The Chirality of Biomolecules,
Robert N. Compton and Richard M.
Longitudinal Interferometry with Atomic
Pagni
Beams, S. Gupta, D.A. Kokorowski,
R.A. Rubenstein, and W.W. Smith Microscopic Atom Optics: From Wires
to an Atom Chip, Ron Folman, Peter
Kriiger, J6rg Schmiedmayer, Johannes
Volume 47 Denschlag and Carsten Henkel
Nonlinear Optics of de Broglie Waves, Methods of Measuring Electron-Atom
P Meystre Collision Cross Sections with an
Formation of Ultracold Molecules Atom Trap, R.S. Schappe, M.L. Keeler,
(T ~< 200 ~tK) via Photoassociation in a T.A. Zimmerman, M. Larsen, P Feng,
Gas of Laser-Cooled Atoms, Frangoise R.C. Nesnidal, J.B. Boffard, T.G. Walker,
Masnou-Seeuws and Pierre Pillet L. W. Anderson and C.C. Lin

Vous aimerez peut-être aussi