Proceedings of the XVlll International Conference
KO15
2007
This page intentionally left blank
Proceedings of the XVlll International Conference
ICOLS
2u07
Telluride, Colorado, USA
24 - 29 June 2007
editors
Leo Hollberg, Jim Bergquist NIST-Boulder, USA
Mark Kasevich Stanford University, USA
KS World Scientific
N E W JERSEY
- L O N D O N - SINGAPORE
*
BElJlNG
- SHANGHAI
*
H O N G KONG
TAIPEI
*
CHENNAI
Published by World Scientific Publishing Co. Pte. Ltd. 5 Toh Tuck Link, Singapore 596224 USA ofice: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
U K office: 57 Shelton Street, Covent Garden, London WC2H 9HE
British Library Cataloguing-in-PublicationData A catalogue record for this book is available from the British Library.
LASER SPECTROSCOPY Proceedings of the Eighteenth International Conference Copyright 0 2008 by World Scientific Publishing Co. Pte. Ltd. All rights reserved. This book, or parts thereoJ may not be reproduced in any form or by any means, electronic or mechanical, including photocopying, recording or any information storage and retrieval system now known or to be invented, without written permission from the Publisher.
For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy is not required from the publisher.
ISBN-13 978-98 1-281-3 19-0 ISBN-10 981-281-3 19-5
Printed in Singapore by World Scientific Printers
PREFACE The eighteenth International Conference on Laser Spectroscopy was held 24-29 June, 2007 in Telluride, Colorado. Telluride, which is nestled in a box canyon near the south-west corner of the state, proved to be an exceptionally beautiful site for ICOLS-07. Although its remote location and high altitude (2909 m) did present some challenges for our participants, all in all, the setting and facilities were well worth the trip. We were also fortunate to experience warm, clear weather for the entire week of the conference. In keeping with its rich tradition, ICOLS-07 was truly an international gathering with 173 delegates and 34 accompanying guests from 21 countries (Australia, Austria, Canada, China, Denmark, France, Germany, Ireland, Israel, Italy, Japan, Netherlands, New Zealand, Poland, Russia, South Africa, Sweden, Switzerland, Taiwan, United Kingdom, and the United States). The technical program consisted of 34 invited talks arranged in the general topic areas of degenerate quantum gases, quantum information and control, precision measurements, fundamental physics and applications, ultra-fast control and spectroscopy, novel spectroscopic applications, spectroscopy on the small scale, cold atoms and molecules, single atoms and quantum optics, optical atomic clocks. We are indebted to the ICOLS-07 program committee for their time and effort in putting together an exceptional and broad technical program; but most importantly, we are indebted to all of the ICOLS-07 attendees whose participation and vibrant exchange of ideas provided the real strength and foundation of the Conference. In addition to the daily oral sessions, Monday and Tuesday evenings were spent in active discussions around 200 outstanding contributed posters. With only limited time available for the oral and poster sessions, ICOLS-07 necessarily focused on but a few areas of the ever expanding field of laser spectroscopy. Unfortunately, with the time demands imposed by the rich technical program, we could only block out a few hours Wednesday afternoon to sample the abundant outdoor activities and recreational opportunities around Telluride. Wednesday evening we joined forces with two local science education organizations in Telluride, the Pinhead Institute and Telluride Science Research Center, for a town talk on quantum computing presented by Rainer Blatt. After the technical sessions on Thursday, we gathered in the Telluride Town Park for an informal banquetlpicniclparty, complete with music, good food, hula-hoops and slack-lines. V
VI
Following the lead of the 2005 conference, we have continued the webbased presence and archive for ICOLS at www.lasersDectroscoDy.org. This year, in addition to this publication by World Scientific of the manuscripts submitted by our invited speakers, we have compiled a DVD that contains pdf images of most of the ICOLS posters and some of the oral presentations. Hopefidly, these records will provide a useful reference and at least a snap-shot of some major research activities in Laser Spectroscopy, circa 2007. The DVD and website also contain some memorable photos from ICOLS-07. Not surprisingly, there were also a few “issues” along the way. Serious delays were caused by the late loss of a major lodging provider which negatively impacted attendance, caused logistic problems, and resulted in the use of condominium-style lodging for most attendees. It was also most unfortunate and sobering that some delegates (at least 5 from China and one from India) were forced to cancel plans to attend ICOLS-07 because of excessive delays in obtaining entry visas to the U.S. The very slow and selective security procedures now in place continue to dampen international scientific exchanges; this is all too reminiscent of the cold-war days of the early ICOLS meetings. Many people and organizations contributed to the ultimate success of ICOLS-07. The conference was only possible because of the generous support of our industrial and organizational sponsors listed below. For local organization, we are especially indebted to Svenja Knappe, Ying-Ju Wang, Lisa Barnes, Lindsey Wilson, Viki Bergquist and Andrew Novick. A very special thanks goes to Bill Fairbank and Siu-Au Lee for assistance with logistics that went well beyond the call of duty. The staff of the Telluride Conference Center and the excellent audio-video services of Curt Rousse allowed us to concentrate on the science rather than on the facilities, meals and hardware. Wayne Itano deserves special recognition for his selfless contribution of time and energy in constructing and maintaining the ICOLS-07 website. Didi Leibfried did an outstanding job with all aspects of the poster sessions. Our sincere thanks also go Erling Riis, Allister Ferguson and Ed Hinds, the organizers of ICOLS-05, for invaluable advice and carryover support. Last but not least, we were extremely fortunate to have connected with the Telluride Science Research Center for organizational arrangements and conference operations. It was a real pleasure to work with Nana Naisbitt, Kari Koch and others of TSRC who did a splendid job in bring all the pieces together efficiently and in addressing numerous issues. Our sincere thanks to all involved with ICOLS-07.
Leo Hollberg and Jim Bergquist, NZST-Boulder Mark Kasevich, Stanford University October 2007
In memoriam Herbert Walther (1935 - 2006)
An impassioned physicist Laser spectroscopy in Germany and throughout the world is closely connected with Professor Herbert Walther, who organized the 4th International Conference on Laser Spectroscopy (ICOLS) held in Rottach-Egern in 1979. A giant in the field, he was a widely recognized and well respected participant of most of the bi-annually held ICOLS until his untimely death last year. On the occasion of the first ICOLS without Professor Walther, it is appropriate and fitting that we look back at the scientific oeuvre of a scientist who produced world-class results throughout his career. Remarkably, his research covered a surprisingly wide spectrum of different topics with great scientific depth as evidenced by the large number of citations his publications have attracted. The story of his scientific career tells in large part also the history of high resolution spectroscopy. For his PhD thesis in 1962, Professor Walther analyzed the Doppler-free transverse fluorescence from an atomic beam with a plane parallel Fabry-Perot interferometer to learn about the nuclear quadrupole moment of the manganese isotope 55Mn. Lifetime measurements and level crossing experiments followed. His publication in 1970 with Dr. John L. Hall that describes the development and performance of a narrow-band dye laser marked the beginnings of tunable, high-resolution, laser spectroscopy and was an important milestone in Professor Walther's scientific life. Dye lasers became a central component in many of his experiments and in 1991, on the occasion of the 25th anniversary of the first dye laser, at a meeting held at his favorite retreat in Ringberg castle, he is pictured with those representing the Who's Who of dye laser spectroscopists of that period. The year he passed away marked the 40th anniversary of the dye laser,
Vlll
but by then it had largely been phased out, often replaced with semiconductor lasers or semiconductor-laser-pumped solid-state lasers. We have seen the spectacular development towards single-cycle pulses of ultrafast lasers, and Professor Walther again contributed with innovative experiments. He demonstrated the carrier-envelope phase dependence in the spatial distribution of photoelectrons comprising an attosecond double slit experiment. In addition to his many beautiful experiments in high-resolution spectroscopy and metrology, Professor Walther conducted a large number of experiments with notable results on resonance fluorescence, Rydberg-atoms, the one-atom maser, the spectroscopy of excimer molecules, radiation pressure, ion trapping, delayed choice experiments, above-threshold ionization, as well a whole series of early atmospheric LIDAR measurements, .. . the list is longer than the space on this page. Almost needless to say, Professor Walther was awarded a large number of prestigious prizes in recognition of his monumental contributions. Professor Walther believed great benefit could be derived from bringing together the world’s best scientists and he strove successfully to make the MaxPlanck Institute of Quantum Optics an international meeting place of laser scientists and quantum opticians. He himself collaborated with many internationally-renowned scientists, some over long periods of time. Professor Walther was also an outstanding teacher and mentor. Many of his former students now hold distinguished positions in academia and industry. With his passing, the fields of laser science and spectroscopy lost a great champion whom we will gratefully and fondly remember. We are decidedly pleased with the joint establishment of the Herbert Walther Award by the Optical Society of America and the Deutsche Physikalische Gesellschaft, which recognizes and will annually remind us of Professor Herbert Walther’s outstanding scientific contributions and his exceptional leadership. There is no better way to show our community’s appreciation. Gerd Leuchs September 2007
ICOLS-07 gratefully acknowledges the generous support of our corporate sponsors:
TQPTICA
fHOtOHICS
1
of
IX
Our thanks go to the following institutions and agencies that provided administrative, logistical and financial assistance to ICOLS-07 ||
ARO
OARPA
Conference Services Provided by;
The views and opinions, and/or findings contain in this report are those of the author(s) and should not be construed as the official policy or decision of any of the organizations listed above.
ICOLS-07 Program Committee E. Arimondo V. Bagnato K. Baldwin R. Ballagh V. Balykin J.C. Bergquist I. Cirac W. Ertmer A. Ferguson H. Fielding L. Hollberg
M. Inguscio W-H. Jhe H. Katori M. Leduc C. Salomon P. Schmidt F.T. Shimizu S. Svanberg W. Ubachs M-S. Zhan
Italy Brazil Australia New Zealand Russia U.S. Germany Germany Scotland England
Italy Korea Japan France France Austria Japan Sweden Netherlands China
us.
ICOLS Steering Committee members serving in recent years F.T. Arecchi E.A. Arimondo H.-A. Bachor R. Blatt N. Bloembergen C.J. Borde R.G. Brewer S. Chu W. Demtroder M. Ducloy M.S. Feld A. Ferguson J.L. Hall P. Hannaford T.W. Hansch S. Haroche
S.E. Harris E.A. Hinds L. Hollberg M. Inguscio V.S. Letokov A. Mooradian E. Riis Y.R. Shen F.T. Shimizu T. Shimizu K. Shimoda B.P. Stoicheff S. Svanberg H. Walther Y.Z. Wang Y.R. Shen
Italy Italy Australia Germany U.S. France U.S. U.S. Germany France U.S. Scotland U.S. Australia Germany France
U.S. England U.S. Italy Russia U.S. Scotland U.S. Japan Japan Japan Canada Sweden Germany China U.S.
Participants in the ICOLS-07 planning committee meeting J.C. Bergquist F.T. Shimizu J.L. Hall A. Ferguson R. Blatt D. Liebfried P. Hannaford S. Haroche
E.A. Hinds E. Riis T.W. Hansch K. Baldwin S-A. Lee H. Yoneda M. Ritsch-Marte D. Leibfried
U.S. Japan U.S. Scotland Germany U.S. Australia France xi
England Scotland Germany Australia U.S. Japan Austria U.S.
This page intentionally left blank
CONTENTS Degenerate Gases
1
Probing Vortex Pair Sizes in the Berezinskii-Kosterlitz-Thouless Regime on a Two-Dimensional Lattice of Bose-Einstein Condensates V. Schweikhard, S. Tung, G. Lumporesi and E.A. Cornell
3
Interacting Bose-Einstein Condensates in Random Potentials P. Bouyer, L. Sanchez-Palencia, D. Cle'ment, P. Lugan and A. Aspect
11
Towards Quantum Magnetism with Ultracold Atoms in Optical Lattices I. Bloch
23
Precision Measurement and Fundamental Physics
37
T-Violation and the Search for a Permanent Electric Dipole Moment of the Mercury Atom E.N. Fortson
39
Quantum Information and Control I
51
Quantum Information Processing and Ramsey Spectroscopy with Trapped Ions C.F. Roos, M Chwalla, T. Monz, P. Schindler, K. Kim, M. Riebe and R. Blatt
53
Quantum Non-Demolition Counting of Photons in a Cavity S. Haroche, C. Guerlin, J . Berm, S. Deleglise, C. Sayrin, S. Gleyzes, S. Kuhr, A4 Brune and J.-M. Raimond
63
Ultra-fast Control and Spectroscopy
73
Frequency-Comb- Assisted Mid-Infrared Spectroscopy P. de Natale, D. Mazzotti, G. Giusfredi, S. Bartulini, P. Cancio, P. Mudduloni, P. Malara, G. Gagliardi, I. Gulli and S. Borri
75
xiii
XIV
Precision Measurement and Applications
87
Precision Gravity Tests by Atom Interferometry G.M. Tino, A. Alberti, A . Bertoldi, L. Cacciapuoti, M. De Angelis, G. Ferrari, A. Giorgini, V. Ivanov, G. Lamporesi, N. Poli, A4 Prevedelli and F. Sorrentino
89
Novel Spectroscopic Applications
101
On A Variation of the Proton-Electron Mass Ratio W. Ubachs, R. Buning, E.J. Salumbides, S. Hannemann, H.L. Bethlem, D. Bailly, M. Vewloet, L. Kaper and M. T. Murphy
103
Quantum Information and Control I1
111
Quantum Interface between Light and Atomic Ensembles H. Krauter, J.F. Sherson, K. Jensen, T. Fernholz, J.S. NeergaardNielsen, B.M. Nielsen, D. Oblak, P. Windpassinger, N. Kjaergaard, A.J. Hilliard, C. Olausson, J.H. Miiller and E.S. Polzik
113
Degenerate Fermi Gases
125
An Atomic Fermi Gas Near a P-Wave Feshbach Resonance D.S. Jin, J.P. Gaebler and J. T. Stewart
127
Bragg Scattering of Correlated Atoms from a Degenerate Fermi Gas R.J. Ballagh, K.J. Challis and C. W. Gardiner
138
Spectroscopy and Control of Atoms and Molecules
151
Stark and Zeeman Deceleration of Neutral Atoms and Molecules S.D. Hogan, E. Vliegen, D. Sprecher, N. Vanhaecke, A4 Andrist, H. Schmutz, U. Meier, B.H. Meier and F. Merkt
153
Generation of Coherent, Broadband and Tunable Soft X-Ray Continuum at the Leading Edge of the Driver Laser Pulse A. Jullien, T. Pfeijer, M.J. Abel, P.A4 Nagel, S.R. Leone and D.M Neumark Controlling Neural Atoms and Photons with Optical Conveyor Belts and Ultrathin Optical Fibers D. Meschede. W. Alt and A. Rauschenbeutel
167
175
xv
Spectroscopy on the Small Scale
185
Wide-Field Cars-Microscopy C. Heinrich, A. Hofer, S. Bernet, and M Ritsch-Marte
187
Atom Nano-Optics and Nano-Lithography KI. Balykin, P.N. Melentiev, A.E. Afanasiev, S.N. Rudnev, A.P. Cherkun, V.S. Letokhov, P. Yu Apel, V.A. Skuratov and V.V. Klimov
195
Pinhead Town Talk, Public Lecture and Mountainfilm
205
The Quantum Revolution - Towards a New Generation of Supercomputers R. Blatt
207
Cold Atoms and Molecules I
217
Ultracold & Ultrafast: Making and Manipulating Ultracold Molecules with Time-Dependent Laser Fields C.P. Koch, R. Koslofl E. Luc-Koenig, F. Masnou-Seeuws and R. Moszynski
219
Bose-Einstein Condensates on Magnetic Film Microstructures M Singh, S. Whitlock, R. Anderson, S. Ghanbari, B. V. Hall, M Volk, A. Akulshin, R. McLean, A. Sidorov and P. Hannaford
228
Cold Atoms and Molecules I1
241
Ultracold Metastable Helium-4 and Helium-3 Gases W. Vassen, T. Jeltes, J.M. McNamara, A.S. Tychkov, W. Hogeworst, K.A.H. Van Leeuwen, V. Krachmalnicofl M. Schellekens, A. Perrin, H. Chang, D. Boiron, A. Aspect and C.I. Westbrook
243
Single Atoms and Quantum Optics I
257
Recent Progress on the Manipulation of Single Atoms in Optical Tweezers for Quantum Computing A. Browaeys, J. Beugnon, C. Tuchendler, H. Marion, A. Gaetan, Y. Miroshnychenko, B. Darquik, J. Dingjan, Y.R.P. Sortais, A.M. Lance, M.P.A. Jones, G. Messin and P. Grangier
259
xvi
Progress in Atom Chips and the Integration of Optical Microcavities E.A. Hinds, M. Trupke, B. Darquik, J. Goldwin and G. Dutier
27 1
Single Atoms and Quantum Optics I1
283
Quantum Optics with Single Atoms and Photons H.J. Kimble
285
Optical Atomic Clocks
295
Frequency Comparison of Al' and Hg' Optical Standards T. Rosenband, D.B. Hume, A. Brusch, L. Lorini, P.O. Schmidt, T.M. Fortier, J.E. Stalnaker, S.A. Diddams, N.R. Newbury, W.C. Swann, W.S. Oskay, K M Itano, D.J. Winelandand J. C. Bergquist
297
Sr Optical Clock with High Stability and Accuracy A. Ludlow, S. Blatt, M. Boyd, G. Campbell, S. Foreman, M. Martin, M H. G. De Miranda, T. Zelevinsky and J. Ye
303
Author Index
317
DEGENERATE GASES
This page intentionally left blank
PROBING VORTEX PAIR SIZES IN THE BEREZINSKII-KOSTERLITZ-THOULESS REGIME ON A TWO-DIMENSIONAL LATTICE OF BOSE-EINSTEIN CONDENSATES V. SCHWEIKHARD, S. TUNG, G. LAMPORESI, and E. A. CORNELL
J I L A , National Institute of Standards and Technology and University of Colorado, and Department of Physics, University of Colorado, Boulder, Colorado 80309-0440, USA jilawww. colorado. edu/bec/ We present results of a study of vortex proliferation in the BerezinskiiKosterlitz-Thouless (BKT) regime on a two-dimensional (2D) array of Josephson-coupled Bose-Einstein condensates. In our lattice system, tunneling between nearest-neighbor condensates provides a Josephson coupling J which acts t o keep the condensates' relative phases locked. A cloud of uncondensed atoms, on the other hand, interacts with the condensates and induces thermal phase fluctuations, which we observe as vortices. As long as the Josephson energy J exceeds the thermal energy T, the array is vortex-free, while with decreasing J / T , thermally activated vortices appear. We give an extended description of a time-to-length mapping technique that allows us t o obtain information on the size of vortex pairs as J/T is varied.
Keywords: Vortices; Bose-Einstein condensates; Josephson-junction array.
1. Introduction
Two dimensional (2D) superfluids undergo a thermal phase transition to a normal state, which proceeds through the unbinding of vortex-antivortex pairs, i.e. pairs of vortices of opposite circulation. Our theoretical understanding of this transition is due to work by Berezinskii' and Kosterlitz and Thouless2 (BKT). The BKT picture applies to a wide variety of 2D systems, among them Josephson junction arrays (JJA), i.e. arrays of superfluids in which phase coherence is mediated via a tunnel coupling J between adjacent sites. Placing an isolated (free) vortex into a J J A is thermodynamically favored if its free energy F = E - T S 5 0. In an array of period d the vortex energy diverges with array size R as E M J l ~ g ( R / d ) , ~ but may be offset by an entropy gain S M log(R/d) due to the available
3
~ R2/d? sites. This leads to a critical condition (J/T)crit sa 1 independent of system size, below which free vortices will proliferate. In contrast, tightly bound vortex-antivortex pairs are less energetically costly and show up even above (J/T)crn. The overall vortex density is thus expected to grow smoothly with decreasing J/T in the BKT crossover regime. The BKT transition in ultracold gases has been the subject of much experimental4"6 and theoretical7"9 work, following the observation of concurrent thermal phase decoherence and vortex formation4 in a continuous 2D Bose gas. Our work is focused on a more detailed understanding of vortexformation, collected in a 2D array of Bose-Einstein condensates (BECs) with experimentally controllable Josephson couplings. Parts of our results have been published previously.5
2. Experimental System and Procedure
Fig. 1. (a) Experimental 2D optical lattice system. In the white-shaded area a lattice of Josephson-coupled BECs is created. The central box marks the double-well potential shown in (b). The barrier height VOL and the number of condensed atoms per well, Nweii, control the Josephson coupling J, which acts to lock the relative phase A<£. A cloud of uncondensed atoms at temperature T induces thermal fluctuations and phase defects in the array when J < T. (c) data showing thermally activated vortex formation. (i) condensate without optical lattice applied, (ii) no vortices form after application of a weak lattice with J > T, whereas in (iii) for J < T vortices (dark spots) appear as remnants of the thermal fluctuations in the array.
5
W e create an array of Josephson-coupled BECs by adiabatically loading a partially Bose-condensed sample of 87Rb atoms into a 2D hexagonal optical lattice of period d = 4.7pm in the x-y plane, as shown in Fig.l(a). The resulting potential barriers between adjacent sites [Fig.l(b)] rise above the condensate's chemical potential, splitting it into an array of condensates which now communicate only through tunneling. Each of the central wells contains NzuellM 7000 condensed particles. By varying the optical lattice depth VOLin a range between 500 H z and 2 k H z we tune J, the collective Josephson coupling," between 1.5p K and 5 n K . The temperature T of the array can be adjusted between 30 - 7 0 n K . The "charging" energy E, due to repulsive mean field interactions, defined in Ref. 11, is on the order of a few p K , much smaller than both J and T . These parameters place our array in the Josephson regime," where J >> E, but E, >> J/Niezl. In this regime the Josephson coupling energy J(l - cos(A4)) acts to lock the relative phases Ag5, and if dominant will ensure at least local phase coherence in the array. A cloud of uncondensed atoms at temperature T on the other hand induces thermal fluctuations of the relative phases of order Aq5~hM The charging energy iE,(ANw,11)2 disfavors population imbalances between sites. In the Josephson regime however, with J >> E,, the resulting quantum fluctuations of the relative phase are quite negligible," of order A& M (E,/4J)1/4. After allowing time for thermalization we probe the array. Because we do not have direct experimental access to the condensate phases in the array, we turn down the optical lattice on a time-scale t,, which is fast enough to trap phase winding defects, but slow enough to allow neighboring condensates to merge, provided their phase difference is small. Phase fluctuations are thus converted to vortices in the reconnected condensate, We then expand as has been observed in the experiments of Scherer et the condensate and take a destructive image in the x-y plane.
m.
3. Earlier Results
Figure l ( c ) illustrates our observations: When J/T < 1 vortices occur in the BEC, as remnants of the thermal fluctuations in the array. In an earlier publication5 we proved thermal activation as the origin of these phase fluctuations. We studied vortex activation while varying J at distinct temperatures T , and showed that vortex proliferation is controlled almost exclusively by the ratio J/T, with a steep rise of vortex number around J/T 1, just as suggested by the free energy arguments presented above.
-
6 4. Inferring Vortex-Antivortex Pair Sizes
Fig. 2. Vortex-antivortex pairs, imaged just prior to their annihilation. Following the optical lattice ramp-down, tightly bound pairs annihilate faster than loosely bound pairs, providing a time-to-length mapping that allows to extract information on vortex pair sizes.
Here we describe a technique that allows us to infer vortex antivortex pair sizes. As in our earlier work we use as a robust vortex-density surrogate the "roughness" T> of the condensate images (see Fig. 1) caused by the vortex cores. This vortex density T> by itself provides no distinction between bound vortex-antivortex pairs and free vortices. In the following we exploit our time-dependent control of the optical potential to distinguish free or loosely bound vortices from tightly bound vortex-antivortex pairs. We make use of the fact that, once the optical lattice potential has been turned off, vortices and antivortices annihilate in the bulk condensate over a w 100 ms timescale. Figure 2 shows an example image of pairs of vortices just prior to their annihilation. It is intuitively obvious that tightly bound vortex pairs will annihilate on a much faster timescale than loosely bound pairs. A "slow" optical lattice ramp-down therefore allows time for tightly bound pairs to annihilate before they can be imaged. By slowing down the rampdown duration r [inset of Fig. 3 (a)], we can thus selectively probe vortex pairs of increasing size. Figure 3 shows vortex activation curves, probed with two different rampdown times.13 A slow ramp compared to a fast one shows a reduction of the vortex density Z>< in arrays with fully randomized phases at low J/T. The difference directly shows the fraction of tightly bound pairs that have
time ramp-down tirrescale T o r=6ms T = 36rYB
•*>> 10 Fig. 3. Vortex density £> probed at different optical lattice ramp-down timescales T. A slow ramp provides time for tightly bound vortex-antivortex pairs to annihilate, allowing selective counting of loosely bound or free vortices only, whereas a fast ramp probes both free and tightly bound vortices. A fit to the vortex activation curve determines its midpoint (J/T) 50% , its 27% - 73% width A(J/T) 2 7-73, and the limiting values T>< (£>>) well below (above) (J/T) 50% .
annihilated on the long ramp, but not on the fast one. To map the experimental ramp-down time-scale to theoretically more accessible vortex-antivortex pair sizes, we compare the observed number of vortices in fully randomized arrays at low J/T to simulations of vortex distributions in a hexagonal array with random phases. In these simulations, following Ref. 12, we count a vortex if all three phase differences in an elemental triangle of junctions are E (0, TT), or if all are G (—7r,0). A snapshot of a simulated vortex distribution is shown in Fig. 4(a). Within the central 20 lattice sites, comparable to the experimental region of interest5 we find, on average, a total of 10 vortices. 6 vortices occur in nearest-neighbor vortex-antivortex pairs [configuration I in Fig. 4(b)], 1.7 (0.4) occur in configuration II (III) respectively, and 1.9 occur in larger pairs or as free vortices. To relate these time-independent simulations to the experiment, we show in Fig. 4(d) the relevant cumulative vortex distributions, i.e. all vortices occurring in pairs larger than a given lower cutoff size. For a given experimental ramp-down duration, we expect only those vortex configurations to survive which are above a lower cutoff pair size imposed by the ramp-down rate. In Fig. 4(e) we compare the simulated cumulative vortex distributions to experimentally measured vortex numbers as a function of ramp down
• 4 » a * »¥! a^i), * «,,»* * , • '.orcco. TO.
n\ vy
» # , « "V^.* * • * .» • * 404. "b"j v
* * * * *A* * * * ^» » *„ »+. .G, . V .„. , . '4.V&
U5
rTY//t*9Laj/t*.V,0.' ^ 2-1 *0* *+'?*a ' to«°* ' '?•%* H' 1 * * *o* i *o*^ •+* *o' * ' »
»
* , * A.#
9
» . # * * * + »
r m X4. /d$?v£S/' ^ ° ° ">" vortex ^) 1 •• configuration 10
O antivortex
b)A;
/ «4-'>\
e)
1A/3 1
s
:s|i?eI .] i i i i i 8 |S24| pHni '€ i (•• •• l/A/3 1 2A/3 >l-5 ^ smallest surviving pair size [d] 2A/3
0.14Ti ii '" in
>1.5
rpair size [d]
">"
-~———i
smallest surviving pairs
12 55
0.12^
1
.°i
oe
I °- i x 0.06 IIIIII
C
.
, T
„i §
^< (J/T <<: 1)
^ T3
A £) (J/T»1) -«™ * 0.02 III * Pun T , iilaa |liiipii iii|pi iliilai 0 0 F *ggia^^^J^B^j^^AJiiP -A----t— .----A
^ £ ® 1 _ -JO 60
IIIIIII ___
§ 0.04||lpl
0
20
40
ramp-down timescale i [ms]
2
Fig. 4. Time-to-length mapping based on vortex-antivortex annihilation, (a) simulation of vortices and antivortices in array with random phases, (b) smallest possible pair sizes in a hexagonal array, I: d/\/3, II: d, III: 2d/v/3. (c) simulated vortex pair size distribution (">": free vortices or pairs larger config. Ill), (d) cumulative distribution, (e) Mapping between ramp-down timescale T and estimated size of the smallest pairs surviving the ramp (upper axis). The difference 2?< — X>> measures the number of observed vortices surviving the ramp (right axis). Comparison to the simulated vortex distribution yields a size estimate of the smallest surviving pairs (upper axis).
9 timescale, to obtain the desired time-to-length mapping. Downward triangles show the decrease of the experimentally measured saturated (low-J/T) vortex density V < with increasing ramp timescale r. The right axis shows the inferred number of vortices that survived the ramp. M 11 vortices are observed for the fastest ramps, in good agreement with the total number of vortices expected from the simulations (indicated as grey bars). For just somewhat slower ramps of T M 5 m s , only 3 vortices survive, consistent with only vortices in configuration I1 & I11 or larger remaining (indicated in Fig. 4(e), top axis). For T 2 30ms ramps less than 2 vortices remain, according to our simulations spaced by more than 2 d / & Thus we infer that ramps of r M 30ms or longer allow time for bound pairs of spacing 5 2 d / a to decay before we observe them.
1.'
0h1y free vortices or pairs > config. I11 contribute
All vortices
j i 0.d
0
t
20 40 ramp-down timescale T [ms]
I '0
Fig. 5. A downshift in the midpoint ( J / T ) 5 0 %of vortex activation curves such as in Fig. 3 is seen for slow ramp-down times, consistent with the occurrence of loosely bound or free vortices at lower J I T only.
With this time-to-length mapping we now return to the observations in Fig.3. For the slower ramp we observe vortex activation at lower ( J / T ) 5 0 % , confirming that free or very loosely bound vortices occur only at higher T (lower J ) . In Fig. 5 we plot the midpoint ( J / T ) 5 0 %of vortex activation curves versus the applied ramp-down time. The data quantitatively show a shift of (J/T)So%from 1.4 for fast ramp times when all vortices are expected to contribute to the signal, to 1.0 for slow ramp times when only loosely bound vortices survive. The data therefore reveal that loosely bound pairs of size larger than 2 d / a , or indeed free vortices, do not appear in quantity
10
until J / T 5 1.0, whereas more tightly bound vortex pairs appear in large number already for J / T 5 1.4. This result clearly illustrates the mechanism of vortex-antivortex unbinding with increasing temperature or decreasing superfluid coupling, which underlies BKT theory.
Acknowledgments We acknowledge illuminating conversations with Leo Radzihovsky and Victor Gurarie. This work was funded by NSF and NIST.
References 1. V. Berezinskii, Sov. Phys.-JETP 32,493 (1971); 34,610 (1972). 2. J. Kosterlitz, D. Thouless, J. Phys. C 6, 1181 (1973). 3. M. Tinkham, Introduction t o Superconductivity, McGraw-Hill, Inc., New York (1996). 4. Z. Hadzibabic et al., Nature 441,1118 (2006). 5. V. Schweikhard et al., Phys. Rev. Lett. 99, 030401 (2007). 6. P. Kriiger et al., Phys. Rev. Lett. 99, 040402 (2007). 7. A. Polkovnikov et al., Proc. Natl. Acad. Sci. U. S. A. 103,6125 (2006). 8. T. Simula and P. Blakie, Phys. Rev. Lett. 96, 020404 (2006). 9. L. Giorgetti et al., Phys. Rev. A. 76, 013613 (2007). 10. J is obtained from 3D numerical simulations of the Gross-Pitaevskii equation for the central double-well system, self-consistently including meanfield interactions of both condensed and uncondensed atoms. A useful approximation for J in our experiments is: J ( V ~ L , N , , ~ ~ , T M ) Nwell x 0.315nKexp[Nw,ll/3950 - V o ~ / 2 4 4 H z ] ( l +0.59T/100nK). 11. A. Leggett, Rev. Mod. Phys. 73,307 (2001). 12. D. Scherer et al., Phys. Rev. Lett. 98, 110402 (2007). 13. Within a dataset, the ramp-down rate is kept fixed, t , = r x V o ~ 1 1 . 3 kHz.
INTERACTING BOSE-EINSTEIN CONDENSATES IN RANDOM POTENTIALS P. Bouyer, L. Sanchez-Palencia, D. ClBment, P. Lugan, A. Aspect Laboratoire Charles Fabry de l’lnstitut d’optique, CNRS and Univ. Paras X I , Campus Polytechnique, 2 av Fresnel, 91128 PALAISEAU cedex, France We investigate the transport properties of an interacting Bose-Einstein condensate in a speckle random potential. At equilibrium in a trapping potential and for the considered small disorder, the condensate shows a Thomas-Fermi shape modified by the disorder. When the condensate is released from the trap, a strong suppression of the expansion is obtained as observed in recent experiments. For the parameters of the experiment, it is shown to result from the competition between the disorder, the interactions and the kinetic energy. A scenario for disorder-induced trapping is proposed and analyzed. Numerical calculations performed in the mean-field approximation agree with the analytical results derived on the basis of this scenario. Keywords: Anderson Localisation, Random Potential, Bose-Einstein Condensate
1. Introduction : disorder and ultracold atoms Disorder can dramatically change the properties of quantum systems and result in a variety of non-intuitive phenomena, many of which are not yet fully understood. Striking examples are Anderson localization,’ percolat i o q 2 disorder-driven quantum phase transitions and the corresponding B o s e - g l a ~ sor ~ >spin ~ glass5 phases. It is known from the Bloch theory of solid state6 that all eigenstates of non-interacting particles in a periodic potential extend over the full system (as in free space). In contrast, it has been shown by Anderson‘ that the single-particle eigenstates in a random potential can be localized in regions significantly smaller than the size of the system. This effect is particularly dramatic in one-dimensional (1D) systems as it can be rigorously established that almost all eigenstates are locali~ed.~,~ Quantum disordered systems are of practical interest in modern condensed matter physics (CM). Indeed, since ‘Nature is never perfect’, the main periodic structure of real solids has to be completed by additional 11
12
quenched r a n d o m potentials. Understanding of quantum transport in amorphous solids is thus one of the main issues in this context, related to electric and thermal conductivities. The basic knowledge is that contrary to Bloch's theory which predicts a (frictionless) transport of non-interacting particles6 as a consequence of the extension of all eigenstates in a periodic crystal, localization effects in disordered potentials result in a strong suppression of the electronic transport in amorphous solids.' On the experimental front, the persistence and the stability of the superfluid phase have been studied g in systems such as 4He on Vycor glasses and dirty electronic materials." Ultracold atomic gases are now widely considered to revisit standard problems of CM with unique control possibilities. Dilute atomic BoseEinstein condensates (BEC) and degenerate Fermi gases (DFG) are currently produced taking advantage of the recent progress in cooling and trapping of neutral atoms. In particular, periodic potentials (optical lattices) with no defects can be designed in a wide variety of geometries." For example, in periodic optical lattices, transport has been widely investigated. l2 Controlled disordered potentials can also be produced by a variety of techniques, for instance speckle optical fields,'3p16 the use of magnetic traps designed on atomic chips with rough wires, localized impurity atoms," or radio-frequency fields.18 Optical speckle ptentials are of interest as both amplitudes and correlation functions with submicron correlation length14 can be controlled at will. For instance, the atom-atom interactions can be treated almost exactly in a tractable mean-field approximation or in other many-body theories. In addition of providing priviledge playgrounds for textbook models, ultracold gases in random potentials are also of fundamental interest as they introduce novel viewpoints related to finite size effects, inhomogeneities and, probably the most important, the possibility of investigating non-equilibrium phenomena and dynamical response funct ions. l9 Within the context of quantum gases, many recent theoretical efforts have considered disordered or quasi-disordered optical lattices. In these systems, one expects a large variety of phenomena, such as the Bose-glass phase transition, localization and the formation of Fermi-glass and quantum percolating and spin glass phases (for a recent review, see27).Localisation properties in interacting Bose gases a t equilibrium in speckle potentials (without an underlying lattice potential) have been discussed in." Further effects have also been addressed in connection to superfluid flows through disordered media. In particular, the reduction of the superfluid fraction
13
and a significant shift as well as the damping of sound waves has been calculated in ref^.^^)^' More recently, the coherent transport of a BEC ~ ' the propagation of a soliton in a BEC in along a disordered g ~ i d e ~ ' )and the presence of disorder have been in~estigated.~' The interplay between the kinetic energy, the atom-atom interactions and disorder is a challenging question that is relevant for interacting matterwaves in random potentials. We investigate here the transport properties of an interacting one-dimensional (1D) Bose-Einstein condensate in a speckle random potential. We focus on a regime where the interatomic interactions strongly dominate over the kinetic energy (hydrodynamic or Thomas-Fermi regime), a situation that significantly differs from the textbook Anderson localization problem and that is relevant for almost all current experiments with BECs (see for instance recent works on disordered BECs13-16). 2. Suppression of expansion of a condensate in a speckle random potential
The question of the coherent dynamics of interacting matterwaves in random media is currently attracting significant experimental attentionl4)l5 mainly related to the search for a suppression of transport similar to that related to Anderson localization.' In this section, we consider the transport of an interacting BEC in a random potential in a tight binding 1D guide. We assume (i) that the chemical potential of the BEC is larger than the depth of the additional potential, p > VR,and (ii) that the correlation length oR of the potential V is much larger than the healing length of the BEC and much smaller than the initial size of the BEC, E << oR << L T F . We present numerical results that show the suppression of the transport of the BEC in a speckle random potential and discuss a scenario to explain this phenomenon. To this end, we distinguish the behaviors of the BEC in the center and in the wings which turn out to be radically different. 2.1. Expansion of a n interacting BEG'
As the initial condition for the BEC wavefunction, we use the TF profile as obtained at equilibrium in the combined potential32 random potential (characterized by uR and V) and trap (of frequency w ) in a regime where p >> VR.The subsequent evolution of the BEC is governed by the time dependant Gross-Pitaevski Equation (GPE) with w = 0 and V # 0. This situation corresponds to switching off abruptly the confining harmonic potential a t time t = 0 while keeping unchanged the random potential, as in
14
recent experiments.14'15 The root mean square (rms) size of the BEG, Az(i) = \7{z2} — (z}2, measures the spreading of the matterwave. Its time evolution is reported in Fig. 1 for several amplitudes VR of the random potential.
theory, VR=0-
15 cot
10
20
25
30
Figure 1. (color online) Time-evolution of the rms-size of the BEG wavefunction evolving in the random potential V for several values of the amplitude VR . The (black) dashed line is the theoretical prediction of the scaling theory (2) with a vanishing random potential.
2.1.1. Absence of disorder In the absence of disorder, the interacting hydrodynamic BEG expands self-similarly as predicted by the scaling approach:36 exp
''mz 2 b(t) 2Kb(t)
rt
(1)
where 6 is the scaling parameter which is governed by the scaling equation36 b = iv2/b2 with initial conditions b(t = 0) = 1 and b(t = 0) = 0. Integrating these equations, we find
that asymptotically reduces to a linear expansion at large time, b(t) ~ \/2wi. Our numerical calculations agree well with this expression (see Fig. 1). 2.1.2. Non-vanishing disorder The situation is significantly different in the presence of disorder. For small enough amplitudes of the random potential, the initial BEG wavefunction
15
is the usual Thomas-Fermi inverted parabola perturbed by the random potential.32 For t 5 l/w,the scaling form (1) is still a good solution of the GPE so, according to the scaling theory36 the BEC wavefunction expands. For larger times and small amplitudes of the random potential (VR 5 O.lp), the effect of disorder on the transport is small and the BEC expands by about one order of magnitude for wr = 10. For larger amplitudes of the disorder (VR2 0 . 1 5 ~the ) ~ expansion of the BEC stops after the initial expansion stage described above. The fluctuations of Az observed in Fig. 1 are due to small contributions in the wings of the BEC wavefunction that still evolve while the core of the wavefunction is localized. This is confirmed in Fig. 2 where the density profiles of the stopped BEC are plotted at two different times: the comparison between the two density profiles at wr = 10 and wr = 20 shows a static dense core and fluctuating dilute wings. Further details are presented in section 2.2. It is worth noting that the stopping of the BEC expansion already occurs for amplitudes of the disorder significantly smaller than the typical energy per particle in the initial BEC:VR < p.
2.2. Scenario f o r the suppression of the transport of the BEG' i n the presence of a speckle random potential This suppression of transport could be related to what is expected from Anderson localization.' However, we stress that the presence of predominant inter-atomic interactions may change the picture.14 Strictly speaking, Anderson localization is related to the existence of localized single-particle eigenstates in a random potential and to the subsequent absence of diffusion.' Here, repulsive interactions are expected to reduce the localization effect.33 During the initial expansion of the BEC, the interaction energy greatly dominates over the kinetic energy in the center so that no Andersonlike localization effect is expected in this region. We now introduce the scenario for disorder-induced trapping that we have sketched in a previous work.14 The dynamics of the BEC in the random potential is governed by three different forms of energy: (i) the amplitude of the random potential, (ii) the interaction energy and (iii) the kinetic energy. It is thus useful to evaluate and compare the kinetic and interaction energies to understand the behavior of the BEC in the random potential. For this, note first that it follows from the initial expansion of the BEC that the fast atoms populate the wings of the expanding BEC whereas the slow atoms remain close to the center. In addition, note that, except for very small amplitudes of the random potential and subsequent long expansion times,
16
the density in the core of the BEC remains large whereas it drops to zero in the wings (see Fig. 2). As a consequence, the interaction energy exceeds the kinetic energy in the core of the BEC while inversely the kinetic energy is dominant over the interactions in the wings. We thus propose a scenario for disorder-induced trapping of the BEC that shows a crucial influence of mean-field interactions. More precisely, we identify two regions of the BEC that participate in the suppression of transport. The first region corresponds to the center, where the trapping results from a competition between the interactions and disorder. The second region corresponds to the wings of the BEC where almost free particles are multiply reflected from and transmitted through random barriers. There, localization is rather due to the competition between the kinetic energy and the random potential and crucially depends on the details of the random potential. In this case, the suppression of the expansion results from classical reflections on large modulations of the random potential. 2.2.1. Disorder-induced trapping in the core of the BEC For the sake of clarity, let us arbitrarily define the core of the BEC as half the size of the initial condensate: -LTF/2 < z < LTF/2 and call
1
fLTF/2
fi,(t) = LTF
-Lm/2
dz If%t)l2
,
(3)
the average density in the center. In particular, at t = 0, due to the parabolic envelope resulting from the harmonic trap, one finds
nc(t= 0 ) = -- . l2g1D
(4)
During the initial expansion stage, the average density in the core n, slowly decreases and the parabolic envelope disappears. Since the interaction energy significantly exceeds the kinetic energy, we expect that the local density l$(z, t )1’ follows almost adiabatically the instantaneous value of nc(t)approximately in the Thomas-Fermi regime. Using a local approximation in the Thomas-Fermi regime, we find a profile lq!1(z,t)1~ N nc(t) - V(Z)/SlD. In order to check this prediction, we plot in Fig. 2 the density profile as obtained from the integration of the GPE. This shows the central region of the BEC during the evolution in the disordered potential at two different times w t = 10 and w t = 20 together with a plot of the analytical expression
17
of \ i p ( z , t ) \ 2 . The two results correspond to a single evolution and to the period when the BEG is trapped by the random potential. We observe in Fig. 2c that the time-dependent fluctuations of the local density profile are significantly smaller than to the modulations of the random potential V/g1D.
-0.4
-0.2
Figure 2. (color online) Above : Density profiles of the BEG for VR = 0.2/j,, CTR = 0.015LTF and £ = 8 X 10~4I/TF (solid red lines) for two different values of the expansion time T and expected Thomas-Fermi profiles in the absence of a random potential (dashed green lines). The random potential, normalized to directly compare to the interaction energy density (V/gm), is also represented (dotted blue line). Below : Magnification of the profile in the core (c) and the wings (d) of the BEG. The solid (red online) and dashed (blue online) lines correspond to the times ujt = 10 and ut = 20 of the same evolution and the dotted (purple online) line corresponds to the calculation of |i/>(z,i)| 2 with nc as a fitting parameter. Note the different scales in the two figures.
The expansion of the core of the BEG in the random potential stops when the condensate starts to fragment, i.e. when the effective chemical potential in the center of the BEG ("p = ncg1D) decreases down to the value of typically two large modulations of the random potential. At this time, the energy per particle in the core of the BEG becomes too small to over-pass the potential barriers and the core of the BEG gets trapped between these large modulations. The density in the center of the BEG after the trapping has occurred corresponds to the maximum value of nc below which two modulation peaks of amplitude V in average are present in the central part of the BEG. If Npeaks(V) is the number of maxima of the random potential in the central part of the BEG (—LxF/2 < z < £
18
solving Npeaks(V= n,glD) = 2. If we consider the case of a random speckle
(%)
potential37 with V, > 0, Npeaks(V)21 Q: exp [-/3$] where Q: Y 0.37 and /3 Y 0.75. From this, we easily find that the final density of the core of . This is of course only valid if the final the BEC is n, Y In
2 [ 1 O
(i ( 2 ) [
In ?&] 5 p/glD).In density cannot exceed the initial density the opposite situation, the BEC is fragmented at t = 0 and the final density L [see Eq. (4)]. In Fig. 3, we plot n, as a function of saturates at n - u 12 91D VR for several correlation lengths gR together with the result of numerical simulations. The results show that n, is a good estimate of the final density in the core of the BEC.
-
0.8 1
1
0.7
0.6 0.5
0.4 0.3 0.2
0.1
.,
n
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
VRh
Figure 3. (color online) Average density nc in the core of the BEC trapped by disorder versus the amplitude of the random potential VR for different values of the correlation length OR and comparison to the calculations. The horizontal (red online) line corresponds to the saturation limit nc = 11p/12gI~.
It is worth noting that our scenario of disorder-induced trapping is expected to apply also to the case of a random potential bounded from above (VR < 0) as well. This situation has been investigated experimentally in Ref.13>15In this case, the fragmentation occurs when = lV~lindependently of the correlation length of disorder (if E << cR).Then, the fragmented BEC is trapped in the small wells of the random potential with a typical size cR and with a central density n, Y lVR1/glD (independent of
4.
19
2.2.2. Behavior in the wings of the BEC The situation is completely different in the wings of the BEC. Due to the small atomic density, the kinetic energy now dominates over the interaction energy. The wings are populated by fast moving weakly interacting atoms that undergo multiple reflections and transmissions from the modulations of the random potential. Ultimately, the trapping of these atoms results from almost total classical reflection on a single large modulation of the random potential with an amplitude exceeding the typical energy of a single p a r t i ~ 1 e . lThis ~ scenario is supported by the density profiles plotted in Fig. 2a and Fig. 2b where one can observe a sharp drop of the atomic density a t the edges of the BEC ( i e . at positions Zmin 21 - ~ L T Fand Zmaxn 2 ~ L T F in Fig. 2a and Fig. 2b). Note that the significant drops correspond either (i) to modulations of the random potential larger than the initial chemical potential p (e.g. at z,in 2 -7LTF) or (ii) to a concentration of weaker barriers (e.g. a t zmin 2 -3.5LTF). The stationary dense core of the BEC creates a barrier that traps the atoms on the other side of the wing. This barrier is however smooth as the atoms can penetrate the intermediate region where the kinetic and interaction energies are of the same order of magnitude. This behavior strongly contrasts with the situation in the core of the BEC. Actually, it is expected that (i) the density profile should not show a Thomas-Fermi shape and (ii) the local density should not be stationary. Both these properties agree with our numerical results as shown in Fig. 2d In particular, the smaller modulations of the wavefunction observed in Fig. 2d are due to the kinetic energy of the particles in the wings. To show this, we calculate the energy per particle of the BEC E = J d z l $ ( z ) l z [ V ( Z ) I $ ( ~ ) 1 ~ / 2 g Due ~ ~ ] .to the conservation of the energy, it is enough to compute this energy a t the initial time" t = 0. We find that
+
E=-
[ ;(I;x)2]
2p 1 - -
5
(5)
Remarkably, the random potential perturbes the energy per particle only at . for VR << p as considered in this work, E 0: p second order in V R / ~ Thus, and we expect the typical wavelength A of the fluctuations in the wings to "The initial kinetic energy is neglected here as we are considering a BEC in the hydrodynamic regime where it is much smaller than the interaction energy.
20 be of the order of the healing length in the initial condensate :
This is confirmed numerically by the properties of the momentum distribution of the BEC which shows two sharp peaks located at p Y &Ti/[. 3. Relation to recent experiments
It is instructive to compare our results to our recent experiments14 on the suppression of transport of BECs in very elongated geometries. We have restricted our analysis to a 1D-BEC as it is expected to show the strongest effects of disorder. This also approximates experimental 3DBECs in very elongated magnetic traps and 1D random potential. Consider a BEC confined in a cylindrically-symmetric 3D-harmonic trap with frequencies wl in the radial directions and w, in the axial direction and subjected to a 1D potential V along z . Assuming tight radial confinement (tiWl > fiw,, p , l c ~ Twhere T is the temperature), the radial degrees of freedom of the BEC are 'frozen' to zero-point oscillations and the dynamics is effectively 1D. In most of the current experimental situations with atomic BECs, the characteristic radius of inter-atomic interactions Re is small compared to the extension of the radial wavefunction 1 1 = so that the interactions are 3D and are parametrized a t low energy by the single scattering length usc.Now if usc<< 1 1 the effective 1D interaction parameter that appears in the GPE is easily computed:34 gl,, = 2haScw1. The regime of large interactions ( p >> VR and E << a,) that we have discussed in section 2 is relevant for the experiments reported in Refs.13-16 In Refs,14,15the transport has been studied in a 1D waveguide by abruptly switching off the axial confinement while keeping unchanged the radial confinement and the random potential. In addition, the speckle potential was almost translation invariant in the radial plane.14 We thus expect that the situation under study in this paper is very similar to these experiments. Our results show a strong suppression of transport of the BEC in the random potential in agreement with experimental observations. Note in addition that qualitative and quantitative agreement are reported in our previous work.14 Because the speckle potential used in14>15has the same amplitude and correlation length as in this paper, we thus expect that the scenario for disorder-induced trapping that we have proposed and analyzed in section 2 is relevant for interpreting the experimental results. In particular, it is interesting to study two characteristics of our model: (i) the behavior of the central density after trapping and (ii) the characteristic
JK,
21
energy in the wings. The first has been investigated in a very recent experiment by Clkment et di4and show an excellent agreement with our prediction. The latter characteristics should be measurable in experiments using time-of-flight techniques or Bragg s p e c t r o ~ c o p y . ~ ~
4. Conclusion Recent experimental techniques in A M 0 offer plenty of possibilities to further study disordered quantum systems.27In the future, it would be interesting to investigate the behavior of interacting Bose gases at equilibrium in a random potential, where smoothing of a BEC,32 new lifshitz glass states and anderson localisation of bogoliubov quasiparticules can be observed.? Another very promising direction would be to extend the experimental studies ofl4>l5to weaker speckle potentials where Anderson localization occurs and the BEC acquires an exponentially decaying shape, owing to multiple quantum reflections and transmissions from the modulations of the random p ~ t e n t i a l . ~Finally, ’ this work could be extended to higher dimensions: for 2D or 3D speckle fields, the disorder-induced trapping described in this work is expected to be much weaker.
Bibliography 1. P. W. Anderson, Phys. Rev. 109,1492 (1958); Y. Nagaoka and H. Fukuyama (Eds.), Anderson Localization, Springer Series in Solid State Sciences 39 (Springer, Berlin, 1982); T. Ando and H. Fukuyama (Eds.), Anderson Localization, Springer Proceedings in Physics 28 (Springer, Berlin, 1988); B. van Tiggelen, in W a v e Diffusion in Complex Media, lectures notes at Les Houches 1998, edited by J. P. Fouque, NATO Science (Kluwer, Dordrecht, 1999). 2. A. Aharony and D. Stauffer, Introduction t o Percolation Theory (Taylor & Francis, London, 1994). 3. T. Giamarchi and H.J. Schulz, Phys. Rev. B 37,325 (1988). 4. M.P.A. Fisher, et al., Phys. Rev. B 40, 546 (1989). 5. M. Mhzard, G. Parisi, and M.A. Virasoro, Spin Glass and Beyond (World Scientific, Singapore, 1987). 6. N.W. Ashcroft and N.D. Mermin, Solid State Physics (Saunders College Publishing , New York, 1976 ). 7. N.F. Mott and W.D. Towes, Adv. Phys. 10,107 (1961). 8. R. Borland, Proc. Roy. SOC.(London), ser A 274, 529 (1963). 9. B.C. Crooker, et al., Phys. Rev. Lett. 51, 666 (1983); J.D. Reppy, Physica 126B,335 (1984); P.A. Crowell, F.W. Van Keuls, and J.D. Reppy, Phys. Rev. Lett. 7 5 , 1106 (1995). 10. Y. Shapira and G. Deutscher, Phys. Rev. B 27, 4463 (1983); R.C. Dynes, et al., Phys. Rev. Lett. 53,2437 (1984); A.T. Fiory and A.F. Hebard, Phys.
22
11. 12.
13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37.
38. 39.
Rev. Lett. 52, 2057 (1984); J.M. Graybeal and M.R. Beasley, Phys. Rev. B 29, 4167 (1984). G. Grynberg and C. Robilliard, Phys. Rep. 355, 335 (2000). S. Burger, et al., Phys. Rev. Lett. 86, 4447 (2001); M. Kramer, L.P. Pitaevskii, and S. Stringari, Phys. Rev. Lett. 88, 180404 (2002); C.D. Fertig, et al., Phys. Rev. Lett. 94, 120403 (2005); A. Trombettoni and A. Smerzi, Phys. Rev. Lett. 86, 2353 (2001); Th. Anker, et al., Phys. Rev. Lett. 94, 020403 (2005). J.E. Lye, et al., Phys. Rev. Lett. 95, 070401 (2005). D. Clkment, et al., Phys. Rev. Lett. 95, 170409 (2005); D. Clkment, et al., New Jour. Phys. 8, 165 (2006). C. Fort, et al., Phys. Rev. Lett. 95, 170410 (2005). T. Schulte, et al., Phys. Rev. Lett. 95, 170411 (2005). U. Gavish and Y. Castin, Phys. Rev. Lett. 95, 020401 (2005). Ph.W. Courteille, et al., J. Phys. B: At. Mol. Opt. Phys. 39 1055 (2006) . L.P. Pitaevskii and S. Stringari, Bose-Einstein Condensation (Clarendon press, Oxford, 2004). P. Lugan et al., Phys. Rev. Lett. 98, 170403 (2007); P. Lugan et al., Phys. Rev. Lett. in press (2007), ArXiv:0707.2918. B. Damski, et al., Phys. Rev. Lett. 91, 080403 (2003). A. Sanpera, et al., Phys. Rev. Lett. 93, 040401 (2004). L. Sanchez-Palencia, et al., J. Phys. B: At. Mol. Opt. Phys. 39, S121 (2006). L. Sanchez-Palencia and L. Santos, Phys. Rev. A 72, 053607 (2005). A. De Martino, et al., Phys. Rev. Lett. 94, 060402 (2005). H. Gimperlein, et al., Phys. Rev. Lett. 95, 170401 (2005). V. Ahufinger, et al., Phys. Rev. A 72, 063616 (2005). K. Huang and H.F. Meng, Phys. Rev. Lett. 69, 644 (1992). S. Giorgini, L.P. Pitaevskii, and S. Stringari, Phys. Rev. B 49, 12938 (1994). T. Paul, et al., Phys. Rev. A 72, 063621 (2005). N. Bilas and N. Pavloff, Phys. Rev. Lett. 95, 130403 (2005). L.Sanchez-Palencia,Phys.Rev.A74,053625(2006). D.L. Shepelyansky, Phys. Rev. Lett. 73, 2607 (1994). M. Olshanii, Phys. Rev. Lett. 81, 938 (1998); D.S. Petrov, G.V. Shlyapnikov, and J.T.M. Walraven, Phys. Rev. Lett. 85, 3745 (2000). Yu. Kagan, E.L. Surkov, and G.V. Shlyapnikov, Phys. Rev. A 54, R1753 (1996); Y. Castin and R. Dum, Phys. Rev. Lett. 77, 5315 (1996). J.W. Goodman, Statistical Properties of Laser Speckle Patterns in Laser Speckle and Related Phenomena, J.-C. Dainty ed. (Springer-Verlag, Berlin, 1975). S. Richard, et al., Phys. Rev. Lett. 91, 010405 (2003). L. Sanchez-Palencia, et al., Phys. Rev. Lett. 98, 210401 (2007)
23
TOWARDS QUANTUM MAGNETISM WITH ULTRACOLD ATOMS IN OPTICAL LATTICES I. Bloch* Institut fur Physilc, Johannes Gutenberg- Universitat, 55118 Mainz, Germany *E-mail:
[email protected] www. quantum.physik.uni-mainr. de Ultracold bosonic and fermionic quantum gases in optical lattices can be used as versatile model systems for the investigation of fundamental condensed matter physics hamiltonians. We report on a novel detection techniques using quantum noise correlations in expanding bosonic and fermionic atom clouds, which has allowed us t o observe quantum statistical Hanbury-Brown and Twiss type bunching and antibunching of the atoms. Furthermore, this method can be used to reveal the ordering of the particles in the lattice. One of the next frontiers in the field is t o observe quantum magnetism with ultracold atoms in optical lattices. Such quantum magnetism relies on a spin-spin superexchange interaction between neighboring atoms, which we demonstrate in this work.
Keywords: Optical lattices, Ultracold Quantum Gases, Noise Correlations, Quantum Magnetism
1. Introduction Ultracold quantum gases in optical lattices have proven to be promising new synthetic quantum materials for the investigation of fundamental condensed matter physics problems and beyond (see e.g.1-3). One of the most prominent examples in this respect are the Bose-Hubbard and Hubbard Hamiltonians, which describe interacting bosons or fermions in periodic potentials. Ultracold bosonic and fermionic atoms in optical lattices are almost perfectly described by such model hamiltonians and in addition offer unique manipulation, control and detection possibilities. One of the next frontiers of research in the field is to realize quantum spin systems in the strongly correlated regime with ultracold atoms. Such systems on a lattice have served for decades as paradigms for condensed matter and statistical physics, elucidating fundamental properties of phase transitions and
24
acting as models for the emergence of quantum magnetism in strongly correlated electronic media. In all these cases the underlying systems rely on a spin-spin interaction between particles on neighboring lattice sites, such as in the Ising or Heisenberg As initially proposed for electrons by Dirac7i8 and Hei~enberg,~,’ effective spin-spin interactions can already arise due to the interplay between the spin-independent Coulomb repulsion and exchange symmetry and do not require any direct coupling between the spins of the particles. The nature of such spin-exchange interactions is typically short-ranged, since it is governed by the wavefunction overlap of the underlying electronic orbitals. In several topical insulators, such as ionic solids like e.g. CuO and MnO, however, antiferromagnetic ordering arises even though the wave function overlap between the magnetic ions is practically zero. In this case a ” superexchange” interaction can be effective over large distance, as introduced in the seminal works of Kramers and Anderson. Here, the spin-spin interactions are mediated by higher-order virtual hopping processes, which in the case of bosons (fermions) leads to an (anti)-ferromagnetic coupling between particles on neighboring lattice sites.6 Such superexchange interactions are believed to play an important role in the context of high-T, superconductivity.12 Furthermore, they can form the basis for the generation of robust quantum gates similar to recent work in electronic double quantum-dot systems,13 and can be employed for the efficient generation of multi-particle entangled states,14 as well as for the production of many-body quantum phases with topological order.l5>l6 For ultracold quantum gases, the characterization of the correlations present in such quantum states requires novel detection techniques. A very powerful detection method has recently been proposed by Altman et al.17 that is based on the analysis of noise-correlations in expanding atom cloud^.^^^^^ In the cases where these are released from optical lattices, such correlations in the spatial fluctuations of the atomic gas after time of flight can be traced back to the fundamental Hanbury-Brown and Twiss type bunching effect for bosons or antibunching effect for fermions. Here we present results on both cases for Mott insulating clouds of bosonic 87Rb atoms and degenerate Fermi gases of 40K atoms.
2. Quantum Noise Correlations For now almost 10 years, absorption imaging of released ultracold quantum gases has been a standard detection method for revealing information on the macroscopic quantum state of the atoms in the trapping potential. For strongly correlated quantum states in optical lattices, however, the average
25
signal in the momentum distribution that one usually observes, e.g. for a Mott insulating state of matter, is a featureless Gaussian wave packet. From this Gaussian wave packet one cannot deduce more about the strongly correlated quantum states in the lattice potential apart from the fact that phase coherence has been lost. Recently, however, the widespread interest in strongly correlated quantum gases in optical lattices as quantum simulators has lead to the prediction of fascinating new quantum phases for ultracold atoms, e.g. with anti-ferromagnetic structure, spin waves or charge density waves. A theoretical proposal by Altman et al.17 has shown that noise correlation interferometry can be a powerful tool to directly visualize such quantum states.
Paths A
\ Paths B _ \. * \t
click! \ I t :
>f
1 click!
Fig. 1. Hanbury Brown-Twiss correlations in expanding quantum gases from an optical lattice. For bosonic particles that are detected at distances d (e.g. on a CCD camera), an enhanced detection probability exists due to the two indistinguishable paths the particles can take to the detector. This leads to enhanced fluctuations at special detection distances d, depending on the ordering of the atoms in the lattice. Detection of the noise correlation can therefore yield novel information on the quantum phases in an optical lattice.
The basic effect relies on fundamental Hanbury-Brown and Twiss (HBT) correlations20"22 in the fluctuation signal of an atomic cloud.18'23"25 For bosons e.g. a bunching effect of the fluctuations is predicted to occur at special momenta of the expanding cloud, which directly reflects the ordering of the atoms in the lattice. Such bunching effects in momentum space can be directly revealed as spatial noise correlations in the expanding atom cloud. Our goal therefore is to reveal correlations in the fluctuations of the expanding atomic gas after it has been released from the trap. Such correlations in the expanding cloud at a distance d can be quantified through
26
the second order correlation function
+
J(n(x d/2) n(x - d/2))d2x J(n(x d/2))(n(x- d/2))d2x’ Here n(x) is the density distribution of a single expanding atom cloud and the angled brackets (.) denote a statistical averaging over several individual images taken for different experimental runs. C(d) =
+
2 . 1 . Noise correlations for bosons - HBT type bunching
In order to understand why, HBT type bunching can be observed when the atoms are released from the lattice, let us for simplicity consider two detectors spaced at a distance d below our trapped atoms and furthermore restrict the discussion to only two atoms trapped in the lattice potential (see Fig. 1). As the trapping potential is removed and the particles propagate to the detectors, there are two possibilities for the particles to reach these detectors, such that one particle is detected at each detector. First, the particles can propagate along paths A in Fig. 1to achieve this. However, alternative propagation paths exists, which are equally probable, paths B in Fig. 1. If one cannot fundamentally distinguish which paths the particles have followed to the detectors, one has to form the sum (for bosons) of the two propagation amplitudes and square the resulting value to obtain the two particle detection probability at the detectors. As one increases the separation between the detectors, the phase difference between the two propagation paths increases, leading to constructive and destructive interference effects in the two-particle detection probability. The length scale of this modulation in the two particle detection probability of the expunding atom clouds depends on the original separation of the trapped particles at a distance slat and is given by the characteristic length scale:
where t is time of flight. 2 . 2 . Noise correlations f o r fermions - HBT type
antibunching For fermionic particles, the same arguments hold as for the case of bosons. However, in the case of fermions, the Pauli principle dictates the twoparticle wave function to be antisymmetric and therefore the difference
27
Fig. 2. Single shot absorption image including quantum fluctuations and associated spatial correlation function for bosonic atoms, (a) 2D Column density distribution of a Mott insulating atomic cloud of bosonic 87Rb containing 6 X 10s atoms, released from a 3D optical lattice potential with a lattice depth of 50-Er. The white bars indicate the reciprocal lattice scale i denned in eq. 2. (b) Horizontal cut through the centre of the image in a and Gaussian fit to the average over 43 independent images each one similar to a. (c) Spatial noise correlation function obtained by analyzing the same set of images, which shows a regular pattern revealing the lattice order of the particles in the trap, (d) Horizontal profile through centre of pattern, containing the peaks separated by integer multiples of i. (from ref.18)
instead of the sum between the two propagation paths has to be taken in order to calculate the two-particle detection probability C(d). This results in destructive interference and instead of correlation peaks one expects correlation dips, corresponding to an antibunching of the fermionic particles. Antibunching for fermionic particles has been observed only in a few cases so far for electrons26"28 and very recently for neutrons.29 It is interesting to note that in strong contrast to the widespread bunching effect for bosons, which can still be explained through classical fields with fluctuating phases, the corresponding fermionic antibunching does not have a classical analogue. In the experiment we release a degenerate Fermi gas of 40K atoms at temperatures of T/TF « 0.23(3) from a 3D optical lattice potential and record the resulting single shot absorption images after a time of flight of
28
-400
-200
0
200
400
-200
O
Fig. 3. Single shot absorption image including quantum fluctuations and associated spatial correlation function for fermionic atoms, (a) Single absorption image of a fermionic 40 K atom cloud after 10 ms of free expansion, (b) ID cut through the same picture together with a Gaussian fit. (c) Spatial noise correlations obtained from an analysis of 158 independent images, showing an array of eight dips. The positions of these dips are spaced at integer multiples of t. (d) Horizontal profile through the correlation image.
t = 10 ms. After accumulating more than 100 of such images, a spatial correlation analysis is carried out on these images, according to eq. 1. The results of such a correlation analysis, together with the single shot images and profiles can be seen in Fig. 3. Instead of peaks, we can indeed see from the analysis that one observes correlation dips spaced by distances of i. This unambiguously demonstrates fermionic antibunching with neutral atoms, which has eluded direct experimental observation so far. Note that bosonic and fermionic antibunching has recently also been observed using advanced single atom detectors and the bosonic and fermionic isotopes of mestable helium.25'30 The anticorrelations obtained from the shot noise analysis of standard absorption images have allowed us to reveal the quantum statistics and furthermore identify the ordering and temperature of the atoms in the periodic potential. These measurements show that noise correlations are a robust tool for further studies of degenerate bosonic and fermionic quantum gases. Especially, this method could allow the unambiguous detection of fermionic quantum phases such as an antiferromagnetically ordered Neel
29 phase,31 which is strongly connected to models of High-T, superconductivity.32,33Further complex orders could be revealed by applying this method ~~,~~ quantum phases to low dimensional quantum s y s t e m ~ , mixture^,^^>^^ with disorder38 or supersolid phases.39
3. Superexchange interactions
Here, we report on the first direct observation of superexchange interactions with ultracold atoms in optical lattices. Previous experiments with ultracold atoms in optical lattices have shown that spin-spin interactions between atoms on neighboring lattice sites can be generated through controlled collis i o n ~ or ~ on-site ~ - ~ exchange ~ interaction^,^^ by bringing neighboring atoms together on a single site. A repeated application of such controlled interactions between neighboring atoms can then allow one to simulate quantum spin systems in discrete time evolution The superexchange interactions demonstrated here, however, are derived by first principles from the Hubbard model and directly implement long ranged spin interactions in the many-body system, allowing for a continuous simulation of spin-interactions between neighboring atoms. We probe the superexchange interactions after first preparing two atomic spin states of 87Rb in an king type antiferromagnetic N6el order, and subsequently recording the time evolution of the spins of neighboring atoms in isolated double well potentials45p47for weak and strong onsiterepulsion between the particles. For dominating onsite interactions over the tunnel coupling between lattice sites, we find pronounced sinusoidal spin-oscillations due t o an effective Heisenberg-type superexchange Hamiltonian, whereas for weaker interactions a more complex dynamics emerges.
3.1. From the Hubbard model to an effective spin-spin interaction
An isolated system of two coupled potential wells constitutes the simplest conceptual setup for the investigation of superexchange mediated spindynamics between neighboring atoms. In the following, we consider a single double well potential occupied by a pair of bosonic atoms with two different spin-states denoted by IT) and II). If the vibrational level splitting in each well is much larger than all other relevant energy scales and intersite interactions are neglected, the system can be described in a two mode
30 approximation by the Hubbard type Hamiltonian
+u(fitLhlL + 'f'tRfiIR)
(3)
1
where the operators iiLL,R and i j , , , ~ , ~create and annihilate an atom with spin 0 in the left and right well respectively, h,,L,R count the number of atoms per spin-state and well, J is the tunnel matrix element, A the potential bias or tilt along the double well axis and U = U T =~ g x w : , ~ ( x ) ~ x the onsite interaction energy between two atoms in I T) and 11). Here, g = ( 4 7 r h 2 a A L ) / m R b is the effective interaction strength with aBL being the (positive) scattering length for the spin states used in the experiment and m R b the rubidium mass, and WL,R(X) denote the wavefunctions for a particle localized on the left or right side of the double well. The state of the system can be described as a superposition of the Fock states 1 T, I), 11,T), ITJ,O) and l O , T l ) , where the left and right side in the notation represent the occupation of the left and right well, respectively, and the states lTJ,O) and 10,TL) are spin triplets. In the following, we will focus on the dynamical evolution of the population imbalance LC = n~ - nR and the spin R ~ T -Rn 1 ~ ) / starting 2 with double wells iniimbalance m = ( n t L T L ~ tially prepared in IT, I). Here n t , L ; L , R = ( f i t , l ; L , ~ denote ) the corresponding quantum mechanical expectation values and TLL,R = ~ T L , R ~ L L , R .
+
+
I
bJ
Fig. 4. Schematics of superexchange interactions. Second order hopping via l T l , 0) and 10, TL) mediates the spin-spin interactions between atoms on different sides of the double well.
In the limit of strong interactions (U >> J ) , when starting in the sub-
31 space of singly occupied wells spanned by I T, I) and I 1,T), the energetically high lying states ITJ, 0) and 10, TI) can only be reached as "virtual" intermediate states in second order tunneling processes. Such processes lead to a long-ranged (super) spin-exchange interaction, which couples the states I T, I) and I I,T' )' (see Fig. 4). The effective coupling strength for this superexchange can readily be evaluated by perturbation theory up to quadratic order in the tunneling operator and yields J,, = 2 J 2 / U . More generally, for an arbitrary spin configuration with equal interaction energies Urr = Url = ULL, the second order hopping events are described by an isotropic Heisenberg type effective spin H a m i l t ~ n i a n : ~ > l ~ > ~ *
where 92,~ = IT)(JIL,R and S ~ , ,=RIL)(TIL,R and S:,R = (%L,R - f i L L , R ) / 2 denote the corresponding spin operators of the system, with S:,R = Sz z t
is,. 3.2. Detecting superexchange interactions The eigenstates of the effective spin Hamiltonian in Eq. 4 are singlet and triplet spin states, separated in energy by the superexchange coupling energy 2Jex. In order to detect the presence of superexchange interactions, we initially prepare a single double well of an optical superlattice in a N6el ordered state I T , 1) and subsequently monitor its time evolution in spinand population imbalance. The results of such a measurement are shown in Fig. 5 where the spin- and population imbalance are plotted over time for different interaction energies U relative to the tunnelling rate J . For weak interactions, one finds a beating between single particle tunnelling and the effective spin-spin superexchange interaction, which are both comparable in coupling strength (see Fig. 5A). For the case of dominating interactions, however, the first order single particle tunnelling becomes more and more suppressed by the strong interactions between the particles and the remaining dynamical evolution is governed by the superexchange coupling alone. In this case, we find a purely sinusoidal dynamical evolution between the spin-states IT, 1) II, T) in the double well (see Fig. 5C). In all cases, we find however, that when starting from an initial state in a single double well, with single occupancy per well, the population imbalance always vanishes for all times.
32
200
Fig. 5. Spin and population dynamics in symmetric double wells. The time evolution of the mean spin m(t) (blue circles) and population imbalance x(t) (brown circles) are shown for three barrier depths within the double well potential: (A) Vs^OIt = 6Er, J/U = 1.25, (B) V^rt = UEr,J/U = 0.26 and (C) Vshort = l7Er,J/U = 0.048. The measured traces for the spin imbalance are fitted to the sum of two damped sine waves (blue lines). The population imbalance x(t) stays flat for all traces.
4. Outlook
In conclusion, we have demonstrated quantum noise correlations in expanding atom clouds and time resolved measurements of superexchange spin interactions between ultracold atoms on neighboring lattice sites and have shown how to to control and detect such interactions with optical superlattices. It is useful to compare the strength of the superexchange interactions demonstrated here, to other long ranged interactions between atoms on neighboring lattice sites. Although for deep optical lattice, superexchange interactions become exponentially suppressed, for lattice depths of around 12 — 15 Er the superexchange coupling strength Jex/h can be several hundred Hertz and thus almost a factor thousand larger than the direct mag-
33
netic dipole-dipole interaction of Rb atoms on neighboring lattice sites. One order of magnitude larger coupling strengths than the ones shown here, however, could still be achieved using electric dipole-dipole mediated spin interactions between ground state polar molecules.4g The loading scheme used here, can also be used to engineer valencebond solid type spin states using optical superlattices and spin-changing collisions. Such VBS states can be viewed as a large array of highly robust entangled Bell pair^,^'>^' whose entanglement may be connected t o a multiparticle entangled state using e.g. superexchange interaction between the initially disconnected pairs. In the context of quantum information, such large entangled quantum states have been shown to be powerful resources for quantum computing.14 The control of superexchange interactions along different lattice directions also offers novel possibilities for the generation of topological many-body states for quantum information p r o ~ e s s i n g . ~ ~ ~ ~ ~
References 1. D. Jaksch and P. Zoller, Annals ofphysics 315,p. 52 (2005). 2. I. Bloch, Nature Physics 1, 23 (2005). 3. M. Lewenstein, A. Sanpera, V. Ahufinger, B. Damski, A. S. De and U. Sen, Adv. Phys. 56,243 (2007). 4. E. Ising, Zeitschrift fiir Physik A 31,p. 253 (1925). 5. W. Heisenberg, Zeitschrzft fiir Physik A 38, 411 (1926). 6. A. Auerbach, Interactzng Electrons and Quantum Magnetism (Springer, 2006). 7. P. A. M. Dirac, Proc. Roy. SOC.,Ser. A 112,661 (1926). 8. P. A. M. Dirac, Proc. Roy. SOC.,Ser. A 123,714 (1929). 9. W. Heisenberg, Zeitschrift fiir Physik 49,619 (1928). 10. H. A. Kramers, Physica 1, 182 (1934). 11. P. Anderson, Phys. Rev. 79,350 (1950). 12. P. Lee, N. Nagaosa and X.-G. Wen, Rev. Mod. Phys. 78,17 (2006). 13. J. R. Petta, A. Johnson, J. Taylor, E. Laird, A. Yacoby, M. Lukin, C. Marcus, M. Hanson and A. Gossard, Science 309,2180 (2005). 14. H. J. Briegel and R. Raussendorf, Phys. Rev. Lett. 86,910(Jan 2001). 15. A. Kitaev, Ann. Phys. 321,2 (2006). 16. L.-M. Duan, E. Demler and M. D. Lukin, Phys. Rev. Lett. 91, p. 090402 (2003). 17. E. Altman, E. Demler and M. D. Lukin, Phys. Rev. A 70,p. 013603 (2004). 18. S. Folling, F. Gerbier, A. Widera, 0. Mandel, T. Gericke and I. Bloch, Nature 434,481 (2005). 19. M. Greiner, C. Regal, J . Stewart and D. Jin, Phys. Rev. Lett. 94,p. 110401 (2005). 20. R. H. Brown and R. Q. Twiss, Nature 177,27 (1956). 21. R. H. Brown and R. Q. Twiss, Nature 178,1447 (1956).
34 22. G. Baym, Act. Phys. Pol. B 29, 1839 (1998). 23. M. Yasuda and F. Shimizu, Phys. Rev. Lett. 77, 3090 (1996). 24. A. Ottl, S. Ritter, M. Kohl and T. Esslinger, Phys. Rev. Lett. 95, p. 090404 (2005). 25. M. Schellekens, R. Hoppeler, A. Perrin, J. V. Gomes, D. Boiron, A. Aspect and C. I. Westbrook, Science 310, 648 (2005). 26. W. D. Oliver, J. Kim, R. C. Liu and Y . Yamamoto, Science 284, 299 (1999). 27. M. Henny, S. Oberholzer, C. Strunk, T. Heinzel, K. Ensslin, M. Holland and C. Schonenberger, Science 284, 296 (1999). 28. H. Kiesel, A. Renz and F. Hasselbach, Nature 418, 392 (2002). 29. M. Iannuzzi, A. Orecchini, F. Sacchetti, P. Facchi and S. Pascazio, Phys. Rev. Lett. 96, p. 080402 (2006). 30. T. Jeltes, J. McNamara, W. Hogervorst, W. Vassen, V. Krachmalnicoff, M. Schellekens, A. Perrin, H. Chang, D. Boiron, A. Aspect and C. Westbrook, Nature 445, 402 (2007). 31. F. Werner, 0. Parcollet, A. Georges and S. Hassan, Phys. Rev. Lett. 95, p. 056401 (2005). 32. P. Lee, N. Nagaosa and X.-G. Wen, Rev. Mod. Phys. 78, 17 (2006). 33. W. Hofstetter, J. I. Cirac, P. Zoller, E. Demler and M. D. Lukin, Phys. Rev. Lett. 89, p. 220407 (2002). 34. L. Mathey, E. Altman and A. Vishwanath, cond-mat/0507108 (2005). 35. A. Polkovnikov, E. Altman and E. Demler, PNAS 103, 6125 (2006). 36. D.-W. Wang, M. D. Lukin and E. Demler, Phys. Rev. A 72, p. 051604 (2005). 37. V. Ahufinger, L. Sanchez-Palencia, A. Kantian, A. Sanpera and M. Lewenstein, Phys. Rev. A 72, p. 063616 (2005). 38. A. M. Rey, I. I. Satija and C. W. Clark, cond-mat/0601307 (2006). 39. V. W. Scarola, E. Demler and S. D. Sarma, Phys. Rev. A 73, p. 051601(R) (2006). 40. D. Jaksch, H. J. Briegel, J . I. Cirac, C. W. Gardiner and P. Zoller, Phys. Rev. Lett. 82, 1975 (1999). 41. A. Sgirensen and K. Mgilmer, Phys. Rev. Lett. 83, 2274(Sep 1999). 42. 0. Mandel, M. Greiner, A. Widera, T. Rom, T. W. Hansch and I. Bloch, Nature 425, p. 937 (2003). 43. M. Anderlini, P. J. Lee, B. L. Brown, J. Sebby-Strabley, W. D. Phillips and J. V. Porto, Nature 448, 452 (2007). 44. E. Jan6, G. Vidal, W. Dur, P. Zoller and J. Cirac, Quantum Inf. Comput. 3, 15 (2003). 45. J. Sebby-Strabley, M. Anderlini, P. Jessen and J. Porto, Phys. Rev. A 73, p. 033605 (2006). 46. J. Sebby-Strabley, B. Brown, M. Anderlini, P. Lee, P. Johnson, W. Phillips and J. Porto, Phys. Rev. Lett. 98, p. 200405 (2007). 47. S. Folling, S. Trotzky, P. Cheinet, M. Feld, R. Saers, A. Widera, T. Muller and I. Bloch, Nature 448, 1029 (2007). 48. A. B. Kuklov and B. V. Svistunov, Phys. Rev. Lett. 90, p. 100401 (2003). 49. A. Micheli, G. K. Brennen and P. Zoller, Nat. Phys. 2, 341 (2006). 50. C. F. Roos, G. P. T. Lancaster, M. Riebe, H. Haffner, W. Hansel, S. Gulde,
35 C. Becher, J. Eschner, F. Schmidt-Kaler and R. Blatt, Phys. Rev. Lett. 92, p. 220402 (2004).
51. C. Langer, R. Ozeri, J. D. Jost, J. Chiaverini, B. DeMarco, A. Ben-Kish, R. B. Blakestad, J. Britton, D. B. Hume, W. M. Itano, D. Leibfried, R. Reichle, T. Rosenband, T. Schaetz, P. 0. Schmidt and D. J. Wineland, Phys. Rev. Lett. 95, p. 060502 (2005).
This page intentionally left blank
PRECISION MEASUREMENT AND FUNDAMENTAL PHYSICS
This page intentionally left blank
T-VIOLATION AND THE SEARCH FOR A PERMANENT ELECTRIC DIPOLE MOMENT OF THE MERCURY ATOM E. N. FORTSON* Department of Physics, University of Washington, Seattle, W A 98195, USA *E-mail:
[email protected]. edu There has been exciting progress in recent years in the search for a permanent electric dipole moment (EDM) of an atom, a molecule, or the neutron. An EDM along the axis of spin can exist only if time reversal symmetry ( T ) is violated. Although such a dipole has not yet been detected, mainstream theories of possible new physics, such as Supersymmetry, predict the existence of EDMs within reach of modern experiments. After a brief survey of current and planned EDM searches worldwide and the implications of current results for the existence of new T-violating (and hence CP-violating) interactions, I review recent work on our own EDM experiment with mercury atoms, describing the newest version of this experiment and discussing current measurements. We have instituted a fixed blind offset that permits us t o test for systematic errors while insuring that any cuts in the data are made objectively. Compared with our 2001 result l d ( l g 9 H g ) / < 2.1 x 1OP2*ecm, an improvement by a factor of 5 t o 10 should be forthcoming, thereby probing yet further for expected new physics.
1. Introduction Half a century ago, a precise search for an EDM of the neutron was undertaken by Smith, Purcell, and Ramsey.1>2They set what seemed at the time a remarkably small upper limit, ld(n)I < 5 x 10W2'ecm. Since then, there have been many further searches for an EDM of the neutron, with ever increasing precision. Likewise there have been continually improved searches for an EDM of an atom or a molecule - including experiments sensitive to an intrinsic EDM of the electron. Thus far, all EDM experiments have yielded a null result. Nevertheless, elementary particle theories that attempt to go beyond the Standard Model,3 most notably Supersymm e t r ~ predict ,~ that EDMs should exist and be large enough to detect by experiments now underway or soon to begin.5,6 The existence of an EDM of any non-degenerate quantum system would 39
40
imply a breakdown of time-reversal symmetry (T), and through the CPT theorem, a violation of CP-symmetry as ell.^>^ (C is charge conjugation, or particle/antiparticle symmetry, and P is parity, or space-inversion symmetry.) CP-violation was first discovered in the decays of KO mesons 40 years ago,7 and has recently been confirmed in B meson decay^.^)^ For many years after the initial discovery, the search for a neutron EDM provided an exacting test of theories put forward to account for the &, and ruled out most of them as the experimental upper limit on the neutron moment steadily decreased to its current value." Atomic and molecular EDM experiments made equally striking advances as well, starting in the 1960s" and leading up to recent work that includes measurements on thallium" and mercury,13 with a host of new experiments now planned or underway. At current accuracies, the atomic and neutron experiments set comparable and complementary bounds on Supersymmetry and other theories of new physics. The rest of this paper is divided into the following sections: 2. Underlying Theory; 3. Survey of EDM Experiments; and 4. The lg9Hg EDM Measurement in Seattle. The reader interested mainly in experimental details first can skip to section 4, and afterwards read about the significance and the relation to other work in sections 2 and 3 if desired. 2. Underlying Theory
It is now generally accepted that a satisfactory explanation of the observed CP-violation (and equivalently, T-violation) in the KO and Bo systems is given by the Standard Model, in which CP-violation occurs as a complex phase factor (the KM phase) in the interaction of quarks with W bosons. Particle EDMs can be calculated from this mechanism, but due to cancellations at the lowest orders, the Standard Model gives negligibly small prediction^.^>^ (In the Standard Model there is one other phase which can lead to physically observable effects, the OQCD term of the QCD Lagrangian. However, the limits on the neutron and atomic EDMs indicate that O Q c D < lo-', and the usual assumption now is to take OQCD = 0 in connection with the existence of a postulated new light particle, the axion.5) Thus the Standard Model by itself predicts EDMs far too small to be observed in current or contemplated experiments. If an EDM is found, it will be compelling evidence for the existence of some sort of physics beyond the Standard Model. There is no shortage of theories of such new physics, but by far the most cherished among particle theorists is Supersymmetry (SUSY).4>14
41
SUSY incorporates quantum gravity consistently, and also solves the gauge hierarchy problem, i.e. it protects the huge energy gap between grandunification/quantum-gravity at 10l6 - lo1' GeV and the electroweak scale at 100GeV. Another reason to expect that new sources of CP violation, as in SUSY, will eventually be found is that Standard Model CP violation is too small to explain the matter-antimatter asymmetry of the universe.15 A feature of SUSY and most other models of new physics that is of great importance for EDMs is the existence of many new particles with CP-odd phase angles that do create EDMs in lowest order and have no natural reason to be small, just as the KM phase angle is about 7r/4 in the Standard Model. SUSY requires that for every particle there exist a superpartner particle, with spin that differs by one half. The emission and reabsorption of virtual spin-0 superpartners tends to generate EDMs in lowest order,5 which will automatically be of observable size if the lowest superpartner mass scale is in the 100 - 1000 GeV range required for SUSY to protect the gauge hierarchy. A number of authors have pointed out that EDM searches therefore have a good chance of being the first experiments to discover SUSY or whatever new physics does lie beyond the Standard Model.
Fig. 1. Allowed values of C P violating phases for the MSSM, assuming a superpartner mass scale of M= 500 GeV. For SUSY to protect the guage hierarchy, M should be in the range 100 - 1000 GeV. The EDM sensitivity scales as M-', so if M= 1000 GeV the angle bounds would be four times larger. The figure is adapted from Ref. 16, updated by M. Pospelov (2003).
As shown in Fig. 1, EDM predictions from SUSY models are already
42
worrisomely large when compared to experiment. The Minimally Supersymmetric Standard Model (MSSM), with “natural” values (of order unity) for its two additional C P violating phases, gives EDMs that are between 10 and 100 times larger than current experimental limits. Fig. 1 shows the allowed phase values in the MSSM when the neutron,” electron (determined from the atomic thallium EDM limit12), and mercury13 EDM limits are considered. The combined limit constrains both phases to be very near zero, which indicates that the MSSM requires some degree of “fine tuning” to be a valid model. Further improvements in the precision of EDM experiments will continue to inform SUSY models, and in general can be considered a sensitive method of probing for C P violating new ~ h y s i c s . ~
3. Survey of EDM Experiments The way T(or CP)-violation at the fundamental elementary particle level would generate an observable EDM depends upon the system under study. The neutron is sensitive almost exclusively to T-violation in the quark sector, while atoms and molecules have bound electrons and are therefore sensitive to T-violation in the lepton sector as well as the quark sector. In atoms and molecules there are actually a number of ways that Tviolating interactions at the particle level could give rise to an EDM, and all are enhanced considerably in heavy atoms.6 Calculations have been made of the atomic EDM due to an EDM distribution in the nucleus, to a T violating force between electrons and nucleons, and to an intrinsic EDM of the electron itself, corresponding respectively to hadronic (quark-quark), semi-leptonic (electron-quark), and purely leptonic interactions as the chief source of T-violation. Which of the possible effects will predominate in a given atom or molecule depends upon the net electronic angular momentum J . In systems with J = 0 (i.e. systems with only closed electronic shells, such as Hg, Xe, and Ra), the EDM vector points along the nuclear spin I, and the greatest sensitivity is to purely hadronic T-violation inside the nucleus. In this case, the important quantity is the nuclear Schiff M ~ m e n t ,which ~ > ~ measures the part of the nuclear EDM that is not completely shielded from the outside world by the atomic electrons. Although shielding does reduce the size of EDMs in closed shell atoms, it turns out that this loss can be more than compensated by the extra experimental EDM sensitivity attained in these atoms. Another source of an EDM along I could in principle be a tensor-pseudotensor form of electron-nucleon T - ~ i o l a t i o n . ~ > ~ In systems with non-zero J (i.e. paramagnetic systems such as Cs, T1
43 or open-shell molecules) the EDM has a component parallel to J , and the greatest sensitivity is to an intrinsic electron EDM, or to a scalarpseudoscalar form of electron-nucleon T - ~ i o l a t i o n .The ~ > ~great atomic theory discovery here, made in the 1960s by Sandars,17 is that the effect of an electron EDM is actually enhanced in a heavy atom, by over a factor of 100 in cesium and considerably more in thallium and other heavier atoms. An additional enhancement, also discovered by Sandars," takes place in polar molecules due to the large internal electric field in these molecules that can couple to an EDM. This field can be of order lo4 - lo5 times available laboratory fields, yielding a corresponding enhancement. The field axis of a polar molecule can generally be aligned in a relatively modest laboratory field. New experiments, some of which are shown in Table 1, are expected to improve over current EDM sensitivity by factors of 10 - 100. Table 1. Some EDM experiments underway or planned Spin
System
Method
Location
Nuclear
lggHg lZ9Xe Ra Neutron
4-cell vapor Liquid cell Optical trap Superfluid He bath Neutron cell
Seattle Princeton Argonne Los Alamos, SNS Grenoble, ILL, PSI
Electron
YbF PbO Other molecule
Beam Cell Optical and ion traps Optical lattice traps Macroscopic B or E
Imperial College Yale Oklahoma, Boulder Penn St, Austin Amherst, Yale, Indiana
133cs
Magnetic Crystal
All experiments are based on what should happen when a spinning elementary particle, atom or molecule having an EDM is placed in the electric field that exists between two oppositely charged parallel plates. In the manner of a spinning top, the spin will precess about the electric field axis due to the electric torque on the dipole. The longer the spin remains in the electric field without being otherwise disturbed, i.e. the longer the spin relaxation time Tz, the larger will be its angle of precession due to an EDM and the more sensitive will be the experiment. When the electric field direction is reversed by reversing the sign of the voltage between the plates, the sense of spin precession about the field axis also reverses. This behavior helps distinguish the precession due to an EDM from that due to
44
other torques. 4. The lg9Hg EDM Measurement in Seattle
lg9Hg has a 6 '5'0 ground state electronic configuration, and a nuclear spin I = Because the ground state carries no electronic angular momentum, an EDM search in mercury is primarily sensitive to T-violation associated with the quarks in the nucleus. The T-violating nature of an EDM is apparent from the Hamiltonian describing the interaction of the mercury spin with external magnetic and electric fields:
i.
H
=
-(dE
+ pB) . I/I,
(1)
where d is the electric dipole moment and p is the magnetic dipole moment. Under time reversal, H must change since I and B change sign while E does not. A search for an EDM of lg9Hg has been underway in our laboratory at the University of Washington for over 20 years. Our last experiment,13 which used a frequency-quadrupled laser diode on the 254 nm mercury absorption line to orient the lg9Hg nuclear spins, yielded the 2001 EDM result: d(lggHg)= -(1.06 f 0.4gStatf 0.4OsYst)x 10-28ecm, which set an upper bound on the EDM of Id(199Hg)I< 2.1 x 10-28e cm (95% confidence level)
As shown in Fig. 1 above, the leading theoretical extension to the Standard Model, Supersymmetry, is expected to generate a lg9Hg EDM comparable to our experimental limit. By increasing the precision of our result, we could provide important information about the model parameter space of Supersymmetry and other theories or of course possibly observe a nonzero EDM. ~
4.1. 4-cell Experiment With such motivations in mind, upon completion of our 2001 measurement we undertook a major improvement in the lg9Hg EDM experiment. We began with a study of the spin relaxation in our vapor cells, which led us to construct new cells that on average have 1.5 times longer spin coherence times. However, the main improvement to the experiment was the construction of an apparatus that incorporates a stack of four vapor cells (See the cutaway view in Fig. 4 below). Previous versions of the experiment have all compared the spin precession frequency between two vapor cells, where the
45
cells are in a common magnetic field and oppositely directed electric fields. In the current experiment the two additional cells are at zero electric field and are used as magnetometers above and below the EDM sensitive cells. They help to improve our statistical sensitivity by allowing magnetic field gradient noise cancellation, and they are also used to cancel out possible magnetic systematic effects (See Fig. 5 below).
Fig. 2. Simplified diagram of the 199Hg EDM apparatus.
As before, to search for an EDM, we measure the Larmor spin precession frequency of 199Hg. A common magnetic field produces Larmor precession in a vapor of spin polarized mercury in each cell, and a strong electric field applied in opposite directions in the middle two cells modifies the precession frequency by an amount proportional to the electric dipole moment. From Eq. (1), an EDM would cause a frequency shift of 2Ed/h, with opposite sign in the two cells; so the magnitude of the EDM is given by d = h8v/(^E), where 5v is the difference in precession frequency between the two cells. The current version of the experiment is shown in Fig. 2. We spin polarize the 199Hg nuclei by optical pumping on the 253.7 nm absorption line in mercury. Strong laser beams line up the nuclear spins, and weaker probe beams monitor the free-precession frequency in each cell. The pump-probe pattern is shown in Fig. 3. Since the light beam is transverse to the precession axis, the circularly polarized pumping light is modulated at the
46
0
»
>
16
_a
'S "2 4 1
0
20
40
60
SO
100
120
Time (sec) Fig. 3. Pump-probe sequence showing the Larmor precession frequency expanded in the inset.
Larmor frequency to synchronously pump the precessing spins. The probe beam is made linearly polarized and the back and forth optical rotation at the Larmor frequency is used to measure the frequency. The ultraviolet light for this transition is obtained by quadrupling the output of an infrared diode laser. Our laser system produces several milliwatts of stable, tunable UV radiation with good spatial characteristics. This system has operated continuously and problem-free for several years, and requires only occasional maintenance. We lock the laser frequency to the absorption line in a separate Hg vapor cell. The cells are held as shown in Fig. 4 inside a sealed vessel filled with about I bar of SFg or Nj gas to reduce the leakage currents. The vessel and electrodes are constructed of conductive polyethylene, which we found had exceptionally low magnetic impurity content. The vapor cells have been altered slightly since our last publication, containing a 100% CO buffer gas, instead of the 95% N2 / 5% CO mixture used for the 2001 measurement. Our studies of spin relaxation in mercury vapor cells19 indicated that the wax coating on the interior of the cells could be damaged by collisions with excited metastable mercury atoms. The CO buffer gas efficiently quenches these metastable states and thus helps prevent damage to the coating. The end result is that we can achieve polarization lifetimes that are a factor of 1.5 longer than was possible with the old vapor cells. With these improvements, we are now sensitive to spin precession frequency shifts on the 10"10 Hz scale. We have made an extensive effort to assess the noise performance,
47
with the goal of improving the sensitivity still further. The shot noise contribution is modeled using computer simulations, which show we are within a factor of three of the shot noise limit. While the modeling also shows that further improvements to reduce the shot noise itself are possible, we must first eliminate the current extraneous noise limiting the experiment. We are pursuing these goals while at the same time we are accumulating EDM data with the present sensitivity.
Fig. 4. Cutaway view of the EDM cell-holding vessel. High voltage (± 10 kV) is applied to the middle two cells with the ground plane in the center, so that the electric field is opposite in the two cells. The outer two cells are enclosed in the HV electrodes (with light access holes as shown here for the bottommost cell), and are at zero electric field. A uniform magnetic field is applied in the vertical direction.
In order to reach the 1x10 2g e cm. level we must place tight bounds on any systematic effects in the measurement. The most dangerous effects are those which generate magnetic fields that are correlated with the direction of the applied electric field. Leakage currents across the cell when high voltage is applied are one prime example. We continually monitor all leakage currents, and with careful cleaning and preparation we limit such currents to the pA level. Our measurements continue to suggest that this is below the level that could cause a problem at the present level of sensitivity. An important new safeguard is possible now that we have 4 cells. As one example, the "leak-test" combination of individual cell measurements, as shown in Fig. 5, is sensitive to leakage current fields while canceling any
48
EDM effect. Another possible problem could be high voltage sparks which might change the field of a trace magnetic impurity located near the cells or electrodes. We have constructed the apparatus from materials that are as free of such impurities as possible. Again, some combinations of cell frequencies will be sensitive to such local fields, and can reveal the presence of impurities. 4.2. Blind Analysis Because of the need to cut some data (for example, when magnetic impurities do appear), we initiated a blind analysis procedure for all data taken after March 2006. The analysis program adds a fixed, blind HV correlated offset to the middle cell fitted frequencies, +6/2 to the middle top cell and -6/2 to the middle bottom cell, which gives an artificial EDM-like signal e cm (our previous upof size 6, randomly generated between f 2 x per bound). This range is large enough to insure the analysis is blind, but small enough to reveal any large spurious signals that might appear due to the changes made when the blind analysis began. Once selected, the blind offset remains fixed throughout all data, and therefore does not interfere with tests for systematic effects (e.g. correlations with leakage currents, etc). And of course it guards against human bias in decisions about making data cuts, etc. We are now taking data for a new measurement of the lg9HgEDM. Thus far the accumulated statistical error is f 1 . 5 x 10-29ecm, over a factor of 10 below the upper limit of our 2001 measurement. It remains to be seen how small a systematic error will emerge from this measurement. 4.3. The Ig9Hg Stark Interference Eflect
A static electric field applied to an atom with an El (electric dipole) optical transition induces M1 (magnetic dipole) and E 2 (electric quadrupole) transitions. The presence of these additional transitions leads to an interference effect of a particular vector character. For a F = 4 F = El transition, such as the one we use in the lggHg EDM search, the fractional change in the absorptivity a is of the form,
;
;
where a is a factor denoting the strength of the effect, E is the direction of the electric field vector of the light driving the transition, k is the propagation direction of the light, Es is the static electric field, and B is the atomic spin
49
Frequency combinations Middle cell difference:
(
0- ~
~
cancels common made noisc cqnivaknt to 2001 mcasurrmcnt
Anti-symmetric combination 1 (EDMcombo): (%-qd$I%-%)
.
cancels np to 2nd order grndicnt noisc same EDM rcsponsc ils middic =n d i & r ~ o c ~
Symmetric combination (LeakTest combo): ((DMT + %)-
( ( ~ a+ r
cancels linear giadicnt noise
givcszcfoforahucEDM scmitivc to lcakagc currents and other -tic systematics
Other combinations can also help reveal the presence of spurious magnetic effects
Fig. 5.
Frequency combinations with 4 cells.
polarization direction of the ground state. The factor a has been calculated t o be -6.6 x lo-' (kV/cm)-120 for the 254 nm El transition in lg9Hg. This Stark interference effect is of interest for the EDM search because it can lead to a light shzft (also called an ac-Stark shift), an apparent Larmor frequency shift that is linear in the strength of the applied electric field; in other words, it can mimic an EDM. The effect can be measured with the present EDM apparatus with only minor modifications, and a preliminary result agrees in order of magnitude with the calculated effect. A satisfying feature of the result is the confirmation that we see an effect that, like an EDM, is linear in Es while using the same apparatus with almost the identical procedure and analysis as used in the EDM experiment itself. A more precise measurement is currently underway. It is crucial to guard against the Stark interference appearing as a systematic effect. One way we have exploited to control the problem is t o use the probe laser at two different wavelengths where the Stark interference light shift has opposite sign, and average the results to cancel out the Stark interference. A way to completely eliminate the Stark interference problem is to evaluate the Larmor frequency 'in the dark' between two probe laser pulses (which establish the Larmor phase at the beginning and end of the dark period). We are currently implementing such a scheme.
50
Acknowledgments
I wish to thank Clark Griffith, Blayne Heckel, Tom Loftus, Mike Romalis, Matthew Swallows, a n d m y other colleagues on the mercury EDM experiment over the years. This work was supported by NSF Grant PHY 0457320. References 1. 2. 3. 4. 5. 6. 7. 8.
E. M. Purcell and N. F. Ramsay, Physical Review 78,p. 807 (1950). J. H. Smith, E. M. Purcell and N. F. Ramsey, Phys. Rev. 108,120 (1957). S. M. Barr, International Journal of Modern Physics A (1993). G. L. Kane, Perspectives on Supersymmetry (World Scientific, Singapore, 1998), p. xv. N. Fortson, P. Sandars and S. Barr, Physics Today 56,33(June 2003). I. P. Khriplovich and S. K. Lamoreaux, C P Violation Without Strangeness (Springer, Berlin, 1997). J. H. Christenson, J. W. Cronin, V. L. Fitch and R. Turlay, Physical Review Letters 13, 138 (1964). Babar Collaboration, B. Aubert, et al., Physical Review Letters 87, 091801/1
(2001). 9. Belle Collaboration, K. Abe, et al., Physical Review Letters 87, 091802/1 (2001). 10. P. G. Harris, C. A. Baker, K. Green, P. Iaydjiev, S. Ivanov, D. J. R. May, J. M. Pendlebury, D. Shiers, K. F. Smith and M. van der Grinten, Phys. Rev. Lett. 85,904 (1999). 11. P. G. H. Sandars and E. Lipworth, Physical Review Letters 13, 718 (1964). 12. B. C. Regan, E. D. Commins, C. J. Schmidt and D. DeMille, Phys. Rev. Lett. 88 (2002). 13. M. V. Romalis, W. C. Griffith, J. P. Jacobs and E. N. Fortson, Phys. Rev. Lett. 86,2505 (2001). 14. J. H. Schwarz and N. Seiberg, Reviews of Modern Physics 71,S112 (1999). 15. M. Trodden, Reviews of Modern Physics 71, 1463 (1999). 16. T. Falk, K. A. Olive, M. Pospelov and R. Roiban, Nuclear Physics B 560,3 (1999). 17. P. G. H. Sandars, Physical Review Letters 14, 194 (1964). 18. P. G. H. Sandars, Physical Review Letters 19, 1396 (1967). 19. M. V. Romalis and L. Lin, Journal of Chemical Physics 120, 1511 (2004). 20. S. K. Lamoreaux and E. N. Fortson, Physical Review A 46,7053 (1992).
QUANTUM INFORMATION AND CONTROL I
This page intentionally left blank
QUANTUM INFORMATION PROCESSING AND RAMSEY SPECTROSCOPY WITH TRAPPED IONS C. F. ROOS, M. CHWALLA, T. MONZ, P. SCHINDLER, K. KIM, M. RIEBE, and R. BLATT Institut fur Experimentalphysik, Universitat Innsbruck, Technikerstr. 25, A-6020 Innsbruck, Austria and Institut fur Quantenoptik und Quanteninformation, Osterreichische Akademie der Wissenschaften Otto-Hittmair-Platz 1, A-6020 Innsbruck, Austria High-resolution laser spectroscopy and quantum information processing have a great deal in common. For both applications, ions held in electromagnetic traps can be employed, the ions’ quantum state being manipulated by lasers. Quantum superposition states play a key role, and information about the experiment is inferred from a quantum state measurement that projects the ions’ superposition state onto one of the basis states. In this paper, we discuss applications of Ramsey spectroscopy for quantum information processing and show that techniques developed in the context of quantum information processing find useful applications in atomic precision spectroscopy. Keywords: Trapped ions, quantum information processing, precision spectroscopy, Ramsey spectroscopy, entanglement
1. Introduction
Single trapped and laser-cooled ions held in radio-frequency traps constitute a quantum system offering an outstanding degree of quantum control. The ions’ internal as well as external quantum degrees of freedom can be controlled by coherent laser-atom interactions with high accuracy. At the same time, the ions are well isolated against detrimental influences of a decohering environment. The combination of these two properties have enabled spectacular ion trap experiments aiming at building better atomic c l o c k ~ , l - ~ creating entangled state^^)^ and processing quantum i n f ~ r r n a t i o n . ~ > ~ At a first glance, the construction of a quantum computer and of an atomic clock might not seem to have much in common. However, there are close ties linking the two fields of research. The implementation of en-
53
tangling quantum gatesg-I3 with ultra-high fidelity necessitates a precise knowledge of the Hamiltonian governing the dynamics of the atomic system and its interaction with the laser beams applied for steering it. Equally important are the characterization of decohering or dephasing mechanisms arising from the interaction of the atoms with fluctuating electromagnetic fields. For this task, Ramsey spectroscopy turns out to be an extremely important tool. In an atomic clock, the transition frequency between two atomic levels is measured by exciting the atom with laser pulses. If the excitation is done in a Ramsey experiment, probing of the clock transition can be described in the language of quantum information processing as a phase estimation algorithm. For this purpose, the use of multi-particle entangled states has been shown to be of interest.14>15In addition, entangling interactions have found applications in atomic clock measurements for quantum state detection of a system that is otherwise difficult to measure" and for the detection of small energy level shifts by preparing a system of two ions in a manifold of entangled states that are part of a decoherence-free subspace. l7 In the first part of this paper, generalized Ramsey experiments investigating ion-ion couplings which are important in the context of high-fidelity quantum gates will be presented. In its second part, experiments aiming at making quantum information processing more robust against environmental noise will be discussed. We will show how to apply (quantum mechanically) correlated states of two ions for precision measurements of atomic constants. These ion-trap experiments demonstrate high-precision spectroscopy in a decoherence-free subspace using a pair of calcium ions for a determination of energy level shifts and transition frequencies in the presence of phase noise. For the measurement, maximally entangled ions are advantageous for achieving a good signal to noise ratio. As the preparation of these states is more involved than single-ion superposition states, we explore the possibility of using classically correlated ions for achieving long coherence times. 2. Experimental setup In our experiments, two 40Ca+ ions are confined in a linear Paul trap with radial trap frequencies of about w1/27r = 4 MHz. By varying the trap's tip voltages from 500 to 2000 V, the axial center-of-mass frequency w, is changed from 860 kHz to 1720 kHz. The ions are Doppler-cooled on the Sl12 H P112 transition. Sideband cooling on the Sl/2 H D5/2 quadrupole transition18 prepares the stretch mode in the motional ground state Simultaneous cooling of stretch and rocking modes is accomplished
55 by alternating the frequency of the cooling laser exciting the quadrupole transition between the different red motional sidebands. Motional quantum states are coherently coupled by a laser pulse sequence exciting a single ion on the IS)= Sl12(rn = -1/2) c-) ID) = D 5 p ( r n = -1/2) transition with a focused laser beam on the carrier and the blue sideband. Internal state superpositions (IS) ei@ID)) 10)can be mapped to motional superpositions lD)(lO)+ei@Il))by a 7r pulse on the blue motional sideband and vice versa. We discriminate between the quantum states Sl/2and D512 by scattering H Pllz dipole transition and detecting the presence or ablight on the Sl/2 sence of resonance fluorescence of the individual ions with a CCD-camera. A more detailed account of the experimental setup is given in
+
3. Ramsey spectroscopy techniques for quantum information processing
In trapped ion quantum computing, continuous quantum variables occur in the description of the joint vibrational modes of the ion string. The normal mode picture naturally appears when the ion trap potential is modelled as a harmonic (pseudo-)potential and the mutual Coulomb interaction between the ions is linearized around the ions’ equilibrium positions.20 In this way, the collective ion motion is described by a set of independent harmonic oscillators with characteristic normal mode frequencies. The normal modes are of vital importance for all entangling quantum gates as they can give rise to effective spin-spin couplings in laser-ion interaction^.^ All entangling ion trap quantum gates demonstrated so far use laser beams that intermittently entangle the internal states of the ion with a vibrational mode of the ion string. At the end of the interaction, the vibrational mode returns t o its initial state and the propagator describing the entangling gate operations is an operator acting only on the ions’ internal degrees of freedom. In most gate operations, the fidelity of the gate suffers if the vibrational state of the ion string couples to an environment that heats or dephases the ion motion. In previous experiments investigating the coherence of the center-ofmass mode of a two-ion crystal, we had observed heating rates of about 100 ms/phonon and coherence times for superpositions (0) 11) of vibrational states of about the same order of magnitude. For the coherence measurement, a Ramsey experiment was performed where first a , by carrier 7 ~ / 2pulse was applied to the ions in state IS)(S)lO)followed a 7r-pulse on the blue sideband of the center-of mass mode in order to create the state IS)lD)(lO) 11)).After a variable delay T , the inverse pulse sequence mapped the state IS)lD)(lO) ei@ll))onto a superposi-
+
+
+
56 tion IS)(cos(+ - +o)lS) +sin(+ - +o)lD))IO).A Ramsey fringe pattern was recorded by scanning the phase $0 which was achieved by switching either the phase of the second blue sideband pulse or the phase of the second carrier pulse with respect t o the phase of the corresponding first pulse. Surprisingly, when this kind of Ramsey experiment was applied to investigate the coherence of the stretch mode of the two-ion crystal, the measured coherence time was found to be nearly two orders of magnitude shorter than for the center-of-mass mode. Fig. 1 (a) shows the contrast C(T)of the Ramsey fringe pattern as a function of the delay time. In this experiment, a coherence time of less than 2 ms was measured at a trap frequency wl(27r) = 1486 kHz. We found that the loss of contrast could be attributed to the nonlinear terms in the Coulomb interaction between the ions giving rise to a cross-coupling between the normal modes.21 For a two-ion crystal, this leads to a dispersive coupling between the stretch mode and the rocking mode where the ions oscillate out of phase in the transverse direction. As a result, the bare stretch mode frequency v$j is lowered slightly by an amount that is proportional t o the number of rocking mode phonons so that (0) v,t, = vst, - xn,,,k. After cooling the rocking modes to the ground state
02-
Fig. 1. Experiments probing the coherence of the stretch mode. (a) Ramsey experiment. (b) Spin echo experiment.
before repeating the experiment shown in Fig. 1 (a), we observed coherence times similar to the ones found for the center-of-mass mode. The fact that the stretch mode frequency is varying from experiment to experiment for ( n r O c k ) # 0 but constant within a single experiment is also revealed by a spin echo experiment probing the stretch mode coherence. For the experiment shown in Fig. 1 (b), a pulse sequence similar to the sequence for (a)
57
is used, but with additional pulses in the middle of the sequence that swap the population of the two lowest quantum states of the stretch mode. This makes the experiment insensitive against small changes of the stretch mode frequency so that the contrast decays to half of its initial value only.after 100 ms. To confirm that the observed spread in vibrational frequencies is indeed due to the postulated mechanism, we could even measure the shift induced by a single rocking mode phonon by performing a spin echo experiment and increasing the rocking phonon number by exactly one at the beginning of the second spin echo period by a blue sideband pulse on the rocking mode. The extra phonon shifts the Ramsey fringe pattern by an amount that can be related t o the strength x of the cross mode coupling. In our experiments, we find frequency shifts of up to 20 Hz per phonon.21 These shifts dramatically reduce the fidelity of Cirac-Zoller gate operations making use of the stretch mode as long as the rocking modes are cooled only to the Doppler limit.22 For the realization of high-fidelity quantum gate operations, this observation points to the necessity of either cooling the rocking modes to the ground state or using the center-of-mass mode for mediating the ion-ion coupling. 4. Quantum information processing techniques for precision spectroscopy
In atomic high-resolution spectroscopy, dephasing is often the most important factor limiting the attainable spectral resolution. Possible sources of dephasing are fluctuating electromagnetic fields giving rise to random energy level shifts but also the finite spectral linewidth of probe lasers. Under these conditions, two atoms located in close proximity to each other are likely to experience the same kind of noise, i.e. they are subject to collective decoherence. The collective character of the decoherence has the important consequence that it does not affect the entangled two-atom state
as both parts of the superposition are shifted by the same amount of energy by fluctuating fields. Here, for the sake of simplicity, 19) and le) denote the ground and excited state of a two-level atom. Because of its immunity against collective decoherence, the entangled state Q+ is much more robust than a single-atom superposition state L(1g) le)). This properties makes Jz states like Q+ interesting candidates for high-precision spectroscopy. In the following, we will first discuss how t o use Bell states for the measurement
+
58
of energy level shifts. Then, it will be shown that certain unentangled twoatom states can have similar advantages over single atom superposition states albeit at lower signal-to-noise ratio.
4.1. Spectroscopy with entangled states In a Ramsey experiment, spectroscopic information is inferred from a measurement of the relative phase q5 of the superposition state Ig) +eZ41e)).
&(
The phase is measured by mapping the states &(lg) f) .1 to the measurement basis { lg),) .1 by means of a 7 r / 2 pulse. In close analogy, spectroscopy with entangled states is based on a measurement of the relative phase 4 of the Bell state Q4 = &(lg)le) e'4le)Ig)). Here, the phase is determined by applying 7r/2 pulses t o both atoms followed by state detection. 7r/2 pulses with the same laser phase on both atoms map the singlet state 1 (1g)le) - 1e)Ig)) to itself whereas the triplet state (1g)le) 1e)lg)) is fi mapped to a state (1g)Ig) ez"le)le)) with different parity. Therefore,
+
-&
&
+
+
oL1)op)
yields information about the measurement of the parity operator = cos 4. If the atomic transition frequencies relative phase since (oL1)oL2)) are not exactly equal but differ by an amount 6, the phase will evolve as a function of time 7 according to q5(7)= q50 ST. Then, measurement of the phase evolution rate provides information about the difference frequency 6. To keep the notation simple, it was assumed that in both atoms the same energy levels participated in the superposition state of eq. (1). In general, this does not need to be the case and the phase evolution is given , ~ the atomic by q5(-r) = ( ( w A ~- w h l ) f ( W A ~- w h 2 ) ) 7 . Here, W A ~ denote ~ fretransition frequencies of atom 1 and atom 2, and w ~ are~the, laser quencies used for exciting the corresponding transitions. The minus sign applies if in the Bell state the ground state of atom 1 is associated with an excited state of atom 2 and vice versa. If the Bell state is a superposition of both atoms being in the ground state or both in the excited state, the plus sign is appropriate.
+
4.2. Spectroscopy with unentangled states of two atoms
One may wonder whether entanglement is absolutely necessary for observing long coherence times in experiments with two atoms. In fact it turns out that the kind of measurement outlined above is applicable even to completely unentangled atoms.23 If the atoms are initially prepared in the
59 product state
this state will quickly dephase under the influence of collective phase noise. The resulting mixed state
appears to be composed of the entangled state 9+ with a probability of 50% and the two states Igg) and lee) with 25% probability each. If the state @+ is replaced by the density operator p p in the measurement procedure described in subsection 4.1, the resulting signal will be the same apart from a 50% loss of contrast. The states 1g)Ig) and 1e)Ie) do not contribute to the signal since they become equally distributed over all four basis basis states by the 7r/2 pulses preceding the state detection. Their only effect is t o reduce the signal-to-noise ratio by adding quantum projection noise since only half of the experiments effectively contribute to the signal.
'
-1 0
I 50
100
150
200
Time (ms) Fig. 2. Parity oscillation caused by the interaction of a static electric field gradient with the quadrupole moment of the D512 state of 40Ca+. The first data point significantly deviates from the fit since the quantum state has not yet decayed to a mixed quantum state.
1 h
2
0.5-
2
0
(b)
T
v)
0 v)
of
0 t
*
L
T
-
-
-0.5I
10
20 30 Field gradient (V/rnrn2)
40
50
Fig. 3. Electric quadrupole shift measured with a pair of atoms in a product stat,e. (a) The shift varies linearly with the applied electric field gradient. (b) Residuals of the electric quadrupole shift measurements. The plot shows deviations of the data points measured with unentangled ion (open circles) and entangled ions (filled circles) with respect to the fit obtained from the entangled state data.
4.3. Measurement of a n electric quadrupole m o m e n t We applied the method outlined in subsection 4.2 to a measurement of the quadrupole moment of the D5l2 state. For this, we prepared the state 1 9, = + 5 / 2 )
+ I-W)
8 (1+3/2)
+I-1~))
(4)
61
and let it decohere for a few milliseconds. Here, Im) = m) denotes the Zeeman sub-level of D512 with magnetic quantum number m. After a waiting time ranging from 0.1 to 200 ms, 7r/2 pulses were applied and a parity measurement performed. Fig. 2 shows the resulting parity oscillation whose contrast decays over a time interval orders of magnitude longer than any single atom coherence time in 40Ca+.A sinusoidal fit to the data reveals an initial contrast of 48(6)% and an oscillation frequency v = 38.6(3) Hz. For the fit, the first data point at t=O.lps is not taken into account. At this time, the quantum state cannot yet be described by a mixture similar to the one of eq. (3) as some of the coherences persist for a few milliseconds and thus affect the parity signal. The parity signal decays exponentially with a time constant Td = 730(530) ms that is consistent with the assumption of spontaneous decay being the only source of decoherence (in this case, one would have T d = T D , / , / ~ FZ 580 ms where T D ~ is/ ~the lifetime of the metastable state). The quadrupole moment is determined by measuring the quadrupole shift as a function of the electric field gradient El. The latter is conveniently varied by changing the voltage applied to the axial trap electrodes. For a calibration of the gradient, the axial oscillation frequency of the ions is measured. Further details regarding the measurement procedure are provided in ref.17 Fig. 3(a) shows the quadrupole shift AvQs as a function of the field gradient (the small offset at El = 0 is caused by the second-order Zeeman effect). By fitting a straight line to the data, the quadrupole moment can be calculated provided that the angle between the orientation of the electric field gradient and the quantization axis is known. Setting AvQs = aE‘, the fit yields the proportionality constant ~1 = 2.977(11) Hz/(V/mm2). The quadrupole shift had been previously measured using a pair of ions in an entangled state. Both measurement give consistent results and thus confirm the validity of the approach based on correlated, unentangled atoms. 5 . Conclusion
Techniques developed for atomic clock measurements turn out to be very useful for precisely characterizing quantum interactions in a system of trapped ions dedicated to quantum information processing. For the realization of ultra-high fidelity quantum gates even small effect like the crosscoupling between vibrational modes that is not predicted by the simple normal mode picture become important. Quantum gates based on nonresonant excitations of vibrational sidebands are less affected than those relying on a resonant excitation. Still, to approach the precision required
62 for fault-tolerant quantum operations might require cooling all modes to the ground state. On the other hand, precision spectroscopy itself can profit from concepts developed for processing of quantum information by making use of more advanced detection schemes.
Acknowledgments We acknowledge support by the Austrian Science Fund (FWF), the European Commission (SCALA, CONQUEST networks), and by the Institut fur Quanteninformation GmbH. K. K acknowledges funding by the LiseMeitner program of the FWF.
References H. S. Margolis et al., Science 306, 1355 (2004). T. Schneider, E. Peik, and Chr. Tamm, Phys. Rev. Lett. 94, 230801 (2005). W. H. Oskay et al., Phys. Rev. Lett. 97, 020801 (2006). T. Rosenband et a]., Phys. Rev. Lett. 98, 220801 (2007). 5. D. Leibfried et al., Nature 438, 639 (2005). 6. H. Haffner et al., Nature 438, 643 (2005). 7. M. Riebe et al., Nature 429, 734 (2004). 8. M. D. Barrett et al., Nature 429, 737 (2004). 9. D. Leibfried et al., Nature 422, 412 (2003). 10. F. Schmidt-Kaler et al., Appl. Phys. B 77,789 (2003). 11. M. Riebe et a]., Phys. Rev. Lett. 97, 220407 (2006). 12. P. C. Haljan et al., Phys. Rev. A 72, 062316 (2005). 13. J. P. Home et al., New J. Phys. 8, 188 (2006). 14. J. J. Bollinger, W. M. Itano, D. J . Wineland, and D. J. Heinzen, Phys. Rev. A 54, R4649 (1996). 15. D. Leibfried et al., Science 304, 1476 (2004). 16. P. 0. Schmidt, T. Rosenband, C. Langer, W. M. Itano, J. C. Bergquist, and D. J. Wineland, Science 309, 749 (2005). 17. C. F. Roos, M. Chwalla, K. Kim, M. Riebe, and R. Blatt, Nature 443, 316 (2006). 18. C. F. Roos et al., Phys. Rev. Lett. 83, 4713 (1999). 19. F. Schmidt-Kaler et al., J. Phys. B: At. Mol. Opt. Phys. 36, 623 (2003). 20. D. F. V. James, Appl. Phys. B 66, 181 (1998). 21. C. F. Roos, T. Monz, K. Kim, M. Riebe, H. Haeffner, D. F. V. James, and R. Blatt, preprint, arXiv:0705.0788. 22. F. Schmidt-Kaler et al., Nature 422, 408 (2003). 23. M. Chwalla, K. Kim, T. Monz, P. Schindler, M. Riebe, C. F. Roos, and R. Blatt, preprint, arXiv:0706.3186.
1. 2. 3. 4.
QUANTUM NON-DEMOLITION COUNTING OF PHOTONS IN A CAVITY S.HAROCHE*, C. GUERLIN, J. BERNU, S. DELEGLISE, C. SAYRIN, S.GLEYZES, S.KUHRt , M. BRUNE, and J.-M.RAIMOND Laboratoire Kastler Brossel, ENS, CNRS, UPMC 24 rue Lhomond, 75005 Paris, France Collbge de France 11 place Marcelin Berthelot, 75005 Paris, France *E-mail: haroche0lkb. ens.fr t Permanent adress: Johannes Gutenberg Universitat, Institut fur Physik, Staudingerweg 7, 55128 Maint, Germany The photons of a microwave field stored in a high-Q cavity are detected nondestructively by a beam of circular Rydberg atoms crossing the cavity one by one. The field collapses into a Fock state as information is progressively extracted by the atoms. The photon number subsequently decays through a succession of quantum jumps under the effect of cavity damping. The QND detection of photons could be used for the preparation and study of various kinds of non-classical fields localized in one or two cavities. Keywords: Cavity Quantum Electrodynamics, Quantum Non Demolition Measurement
1. Introduction Counting photons is generally a destructive process, since light is usually absorbed by photo-sensitive materials. It does not have to be so, however, and it has been known for a long time that light intensity can in principle be measured by photon non-absorbing quantum non-demolition (QND) methods.' Such QND procedures have been successfully used to analyze the fluctuations of relatively intense light beams containing many light quanta,2 but have so far been unable to pin-down discrete photons numbers. Taking advantage of the very strong light-matter coupling provided by Cavity Quantum electrodynamic^,^ we have recently been able to count photons in a high-Q cavity, in a way which fulfills all the conditions of an ideal QND measurement. The breakthrough for realizing these experiments has been the development of a very high-Q superconducting Fabry-Prot resonator 63
64
made out of precisely machined copper mirrors sputtered with a thin layer of Niobium.4 A microwave field is trapped between these mirrors for 0.13 s on average, a time long enough to let thousands of circular Rydberg atoms of Rubidium cross the cavity and extract progressively information from the field. The measurement induces the field to collapse into a Fock state with a well-defined number of photons. The field remains in this state for a while, until relaxation makes the photon number cascade down to zero by undergoing successive quantum jumps at random times. This QND procedure has allowed us to detect - for the first time in a real experiment the staircase-like field-intensity signals that were previously exhibited only by Monte Carlo simulations of quantum field evolution.5 We briefly present here the results of these experiments which have been recently published in two papers,6'7 and discuss the perspectives they open for the study of non-classical states of light. 2. Principle of experiment: atoms as clocks to read out the number of photons stored in a box
S
Fig. 1. Schematic view of the circular Rydberg atom-superconducting cavity set-up for photon QND counting (from ref. 6)
The measurement is based on the detection of the phase-shift induced by the field on the atomic coherence between the circular Rydberg states
65 Ie) and 19) of rubidium atoms crossing one by one the cavity (1e)and 19) have principal quantum numbers 51 and 50 respectively). Fig. 1 presents an artist's view of our experimental set-up. A stream of Rydberg atoms, prepared in state Ie) in box B , are sent one at a time through the superconducting cavity C. This cavity sustains a Gaussian transverse profile mode at 51 GHz, nearly resonant with the transition between the states le) and 19). The atoms are subjected to classical pulses of microwave emitted by the pulsed source S and applied in the auxiliary cavities R1 and Rz sandwiching C. The combination of these two pulses constitute a Ramsey interferometer. After leaving Rz, the atoms are detected by a state selective field-ionization detector D . The velocity of each atom and its preparation time are controlled through a pulsed optical pumping process involving properly tuned laser beams. The central part of the set-up, from B to Rz, is cooled to a temperature of 0.8 K by a He cryostat. This ensures the good operation of the superconducting cavity and suppresses most of the thermal radiation background. More experimental details about this set-up can be found in refs. 3,6 and 8. The atoms and the field in C are slightly off-resonant, the frequency offset 6 being at least of the order of the atom-cavity vacuum Rabi frequency R (R/27r = 50 kHz). In spite of this relatively small detuning, no real atomic transitions can occur, because the atom-field coupling varies adiabatically as the atoms travel across the Gaussian profile of the cavity mode. The method is thus truly quantum non-destructive for the field. Precise tuning of the atomic transition is achieved by applying across the cavity mirrors a small electric field which Stark-shifts the circular states. This tuning field is added to a constant directing field applied across the mirrors to protect the circular states from unwanted transitions towards non-circular levels. The necessity to apply these fields to the atoms while they cross C precludes for this experiment the use of closed cavity structures. Due to the very strong coupling of the Rydberg atoms to microwaves, the phase-shift per photon accumulated by the atomic coherence during cavity crossing reaches values of the order of 7 r , for an atom-field detuning 6/27r = 70 kHz and an atomic velocity 'u = 250 m/s. Smaller phase shifts are easily obtained by merely increasing 6. In order to analyse our QND procedure, which is a variant of a method we had proposed in the early 1 9 9 0 ~ , it~ 1is~convenient ~ to describe each atom crossing C as a spin 1/2, the circular levels Ie) and Ig) corresponding to the spin states I+)z and I-)z respectively, along the direction Oz. The atomic states can then be represented as Bloch vectors whose tips are
66 on a Bloch sphere. Just before entering C, the atomic Bloch vector, ini), is rotated by the pulse R1 along the tially prepared along Oz (state 1.) transverse direction Ox [the corresponding state is the linear superposition = ).1( Ig))/fi]. The Bloch vector then starts to rotate in the equatorial plane of the Bloch sphere, in full analogy with the ticking of a clock’s hand. Due to the light-shifts, this clock is delayed by the presence of the field in the cavity, so that the spin ends up in different directions depending upon the photon number. If the atomic phase shift per photon is adjusted to r/q (q integer), the spin’s hand when the atom leaves the cavity points in 2q directions spanning 360 degrees for photon numbers ranging from n = 0 to n = 2q - 1. The QND method consists in reading out these directions. For n 2 2q, the spin’s positions repeat themselves, so that the method is measuring n modulo 2q. Detecting a single atom provides in general only partial information, since the 2q final spin states are non-orthogonal (a notable exception is q = 1, see section 3 ) . Suppose that we decide to detect the spin component along the direction Ou(p) making the angle 4 ( p ) = p r / q ( p integer) with Ox.If the cavity contains p photons, the spin ends up in the state I+) u ( p ) and the probability for finding the result - l / 2 for the spin component along Ou(p) is zero. Conversely, if one finds the spin in this state, the probability that C contains p photons must obviously vanish. If, on the other hand, the spin is found in the +1/2 state along O u ( p ) , it is the probability for finding p+q photons which cancels. In other words, the atom detected along the direction Ou(p) provides information enabling us to suppress either the value p or the value p q from the photon number distribution. This logical argument, which allows us to infer the change in the photon probability distribution due to an acquisition of knowledge on the final spin’s state, is an expression of Bayes law in probability theory. By choosing for the next atom another detection direction Ou(p’), different photon numbers are decimated. With q different detection directions adjusted for successive atoms crossing C, we find out which photon number survives out of 2q initially possible values. In practice, the transverse spin is detected along a given direction making an angle 4with Ox by mapping out this direction onto Oz with the pulse Ra, applied to the atom after cavity exit, before performing a measurement of the atom in the energy basis. The angle 4 is fixed by properly choosing the phase of the R2 pulse. A measurement is thus ideally constituted by a sequence of q atoms crossing C, providing each a +1/2 or -1/2 reading, associated to one out of q different detection angles, i.e. different phases of the R2 pulse. In a real
+
+
67
situation, some redundancy is necessary and more atoms are required to pin down n without ambiguity, because of -Ri and R-2 pulses imperfections and of unread atoms due to limited detection efficiency. 3. A simple situation: counting single photons and detecting field quantum jumps
0.0
0.5
1.0 1.5 Tim e (s)
2.0
2.5
Fig. 2. QND detection of a single photon. Upper and lower bars show the signal, a sequence of atoms detected in e) and \g) respectively. The photon observed here is exceptionally long-lived (about three cavity damping times). Erroneous counts (\e) detections in vacuum and \g) detection when 1 photon is present) are due to the imperfections of the Ramsey interferometer (adapted from ref. 6).
We have first applied this method to the measurement of the residual field produced in C by the thermal excitation of the mirrors.6 According to Planck's law, the cavity contains on average 0.05 photons at T = 0.8 K, this mean value resulting from random fluctuations of the number of light quanta between zero and one. The probability that C contains more than one photon is negligible at this low temperature. We thus have merely to distinguish between two photon number values (0 and 1). We choose in this case q = 1, which corresponds to a TT phase-shift per photon. There is then only one detection direction and the jf?2 pulse has an unique phase, mapping the transverse Bloch vectors corresponding to 0 and 1 photon onto the + and — directions along Oz respectively. An atom detected in \g} thus signals 0 photon and an atom detected in e) one photon. Fig. 2 shows a sequence of 2200 atomic detections recorded over a 2.5 second interval. The upper and lower vertical bars correspond to atoms found in \e) and \g) respectively. A long sequence of atoms detected mostly in \g) indicates that the field is in vacuum. Then, around t = 1.05 s, the telegraphic signal suddenly changes. The atoms are then detected mostly in e), signalling the appearance of one photon. The photon number has undergone a quantum jump from n = 0 to n = 1, followed
68 about half of a second later by a jump in the opposite direction, marking the annihilation of the photon. Thousand of similar signals have been recorded, whose statistical analysis is in complete agreement with the predictions of Planck’s law and quantum electrodynamics theory. In another test, we have first prepared a photon in C by having a first resonant atom emit it, then detected this photon by sending across C a long sequence of QND-detector atoms. Repeating the experiment many times, we have analysed the statistical distribution of these single-photon survival times and obtained an exponential distribution, with a mean life time equal to the cavity damping time T, (with a small well-understood correction due to the effect of residual radiative thermal processes). 4. Progressive field state collapse and stochastic evolution of the photon number
In order to count larger photon number^,^ we inject in C a coherent field produced by a microwave source. This field is coupled in the cavity via diffraction on the mirrors edges. Its photon number has a Poisson distribution, with an average no = 3.84. The probability for finding more than 7 photons is 3.5%. The task of the QND procedure is thus to distinguish between 8 consecutive values of n comprised between 0 and 7. For this, we . detection choose q = 4 and adjust the phase shift per photon to ~ / 4 The phase is, from one atom to the next, adjusted to four different angles corresponding to the directions of the Bloch vectors associated to n = 6 , 7 , 0 and 1. The atomic data are processed by exploiting Baysian logic, extracting information from a long sequence of detection events. The atomic detection rate is about 5 atoms per milliseconds. At time t = 0, the photon probability distribution is flat, since no a priori knowledge is assumed, except that the photon number is bounded by seven. Then, as atoms are successively detected, photon numbers are decimated, until the distribution has converged to a single integer value. This corresponds to the collapse of the field state, induced in a step-by-step process by the progressive acquisition of information provided by the atomic readings. The measuring sequence corresponds to 110 atoms, detected within 26 milliseconds. This number is a compromise. It is large enough to let the photon number converge on most sequences, and small enough for the measurement time to remain short compared to T,. At the end of this collapse stage, the procedure is resumed. We drop information provided by the first atom and add information extracted from the lllth one and so on.. . In this way, the data are decoded continuously, N
69
using at each time information provided by the last 110 atoms. An example of signal is shown in Figure 3, as a 3-D histogram. The photon number is plotted along one horizontal direction and the atom number along the other. Photon numbers from 0 (foreground) to seven (background) are represented by channels of different shades whose heights (representing the corresponding probabilities) evolve from left to right. Out of the initially uniform distribution, a single channel (n = 5) surges, as the others decay to zero. This is the state-collapse process. The selected channel remains at a plateau-level for a while, illustrating the repeatability of a QND measurement. Cavity damping then takes over. The n = 5 channel suddenly drops to zero while the n = 4, 3, 2,1 ones successively and transiently surge. This describes a photon-number cascade towards zero occuring through sudden quantum jumps. The evolution ends with a steady n = 0 channel (field in vacuum). For clarity, the time scale - and hence the calibration of the atomic axis in Figure 3 - are non-linear.
Fig. 3. Three dimensional histogram showing the evolution of the photon number distribution under repeated QND measurement. Note the non-linear calibration of the atomic axis which makes visible the fast field collapse stage.
We have observed thousands of such field trajectories. The histograms of the n-values obtained at the end of the collapse stage reproduce, to an excellent approximation the Poisson distribution of the initial coherent field.
70
This illustrates the quantum postulate about the statistics of measurement outcomes. This experiment has generated for the first time Fock states of radiation with photon numbers larger than 2.
5. Perspectives for the study of non-classical field states in one or two cavities This QND measurement opens novel perspectives for the generation of nonclassical states of light. If the initial photon number distribution spans a range of ns larger than 2q, the decimations induced by successive atoms do not distinguish between n and n 2q. The field then collapses in a coherent superposition of the form Encn+zqln 2q). For instance, ~ 0 1 0 ) c2,12q) represents a field coherently suspended between vacuum and 2q photons. This superposition of states with energies differing by many quanta is a new kind of Schrodinger cat state of light. Other kinds of Schrodinger cat states are produced during the QND sequence. As the photon number is pinned-down, its conjugate variable, the field’s phase, gets blurred. After the first atom’s detection, the initial state collapses into a superposition of two coherent states with different phases.”>” Each of its components is again split into two coherent states by the next atom and so on, leading to complete phase uncertainty when the photon number has converged.” The evolution of the Schrodinger cat states generated in the first steps of this process could be studied by measuring the field Wigner function.12 De~oherence’~ of superpositions of coherent states containing many photons could be monitored in this way. Finally, we intend to extend these experiments t o the generation and study of field states belonging to two high-Q cavities, successively crossed by a beam of circular Rydberg atoms.14>15We could for instance prepare the field in a superposition of the form ( a ,0) 10, a ) representing a coherent field of complex amplitude a which is “at the same time” in the first cavity and in the second.16 If a beam of circular Rydberg atoms is used to measure the global photon number of the two cavities in a QND way, this field will collapse into a two-cavity Fock state of the form In, 0)+ 10, n ) ,corresponding to n photons being in a superposition of the state in which they all belong to the first cavity with the state in which they all belong to the second. These strange non-classical states, which have been recently generated in different c o n t e ~ t s , ~ will ’,~~ be very interesting to investigate in this Cavity QED situation.
+
+
+
+
71
Acknowledgements
We acknowledge funding by Agence Nationale pour la Recherche (ANR), by the Japan Science and Technology Agency (JST), by the EU under the I P projects “SCALA and ‘CONQUEST. C.G and S.D are funded by a grant from Dklkgation Gknkrale B 1’Armement (DGA). JMR is a member of Institut Universitaire de France (IUF) References 1. V. B. Braginsky and Y. I. Vorontsov, Usp. Fiz. Nauk, 114,41 (1974) [Sov. Phys. Usp. 17,644 (1975)]; K. S. Thorne, R. W. P. Drever, C. M. Caves, M. Zimmerman and V. D. Sandberg, Phys. Rev. Lett. 40,667 (1978). 2. P. Grangier, J. A. Levenson and J.-P. Poizat, Nature 396,537 (1998). 3. S. Haroche and J. M. Raimond, Exploring the Quantum: Atoms, Cavities and Photons (Oxford Univ. Press, Oxford, UK, 2006). 4. S. Kuhr et al., Appl. Phys. Lett. 90,164101 (2007). 5. H. Carmichael, An open system approach to quantum optics (Springer, Berlin, 1993). 6. S. Gleyzes et al., Nature 446,297-300 (2007). 7. C. Guerlin et al., Nature in press (2007). 8. J. M. Raimond, M. Brune and S. Haroche, Rev. Mod. Phys. 73,565 (2001). 9. M. Brune, S. Haroche, V. Lefkvre, J. M. Raimond and N. Zagury, Phys. Rev. Lett. 65,976-979 (1990). 10. M. Brune, S. Haroche, J. M. Raimond, L. Davidovich and N. Zagury Phys. Rev. A. 45,5193 (1992). 11. M. Brune et al, Phys. Rev. Lett. 77,4887 (1996). 12. P. Bertet et al, Phys. Rev. Lett. 89,200402 (2002). 13. W. H. Zurek, Rev. Mod. Phys. 75,715 (2003). 14. L. Davidovich, M. Brune, J. M. Raimond and S. Haroche, Phys.Rev.A, 53, 1295 (1996). 15. P. Milman, A. Auffeves, F. Yamagushi, M. Brune, J. M. Raimond and S. Haroche, Eur. Phys. J . D., 32,233 (2005). 16. L. Davidovich, M. Brune, J.-M. Raimond and S. Haroche, Phys. Rev. Lett. 71,2360 (1993). 17. M. W. Mitchell, J. S. Lundeen and A. M. Steinberg, Nature, 429,161 (2004). 18. P. Walther, J. W. Pan, M. Aspelmeyer, R. Ursin, S. Gasparoni and A. Zeilinger, Nature, 429,158 (2004).
This page intentionally left blank
ULTRA-FAST CONTROL AND SPECTROSCOPY
This page intentionally left blank
FREQUENCY-COMB-ASSISTED MID-INFRARED SPECTROSCOPY P. DE NATALE*, D. MAZZOTTI, G. GIUSFREDI, S. BARTALINI and P. CANCIO Istituto Nazionale di Ottica Applicata (INOA) - CNR and European Laboratory for Nonlinear Spectroscopy (LENS) Via Carrara 1 , 50019 Sesto Fiorentino FI, Italy * e-mail: paolo.
[email protected], web: http://www.inoa.it
P. MADDALONI, P. MALARA and G. GAGLIARDI Istituto Nazionale di Ottica Applicata (INOA) - CNR and European Laboratory for Nonlinear Spectroscopy (LENS) Via Campi Flegrei 34, 80078 Pozzuoli NA, Italy I. GALL1 and S. BORRI Dipartimento di Fisica, Universitb di Firenze and European Laboratory for Nonlinear Spectroscopy (LENS) Via Sansone 1 , 50019 Sesto Fiorentino FI, Italy A new class of IR coherent sources and IR frequency combs, that combine optical frequency-comb synthesizers (OFCSs) and optical parametric up/downconversions, is already available and still progressing at a very fast pace. Peculiar features for IR radiation produced by difference-frequency-generation (DFG) set-ups or quantum-cascade lasers (QCLs) can he achieved when they are phase and frequency controlled by the OFCS. Indeed, their frequency is accurately known against the primary frequency standard and their linewidth is highly narrowed thanks to the transferred OFCS coherence even for laser sources whose frequencies are several THz apart. These features, together with their wide tunability and their small intensity fluctuations (down to the shotnoise limit), make these IR sources well suited for a wide range of applications, in particular for spectroscopic ones. Very high sensitivity for trace-gas detection has been achieved when combined with enhancement absorption techniques as high-finesse Fabry-Perot cavities or multipass cells. Moreover, the large number of fundamental ro-vibrational transitions of many stable and transient molecular species accessible with this spectrometers, make them particularly attractive for environmental applications, especially considering their compactness and ruggedness when a fiber-based set-up is chosen. Their unique capabilities in terms of achievable precision for absolute frequency measurements can he used to create a “natural” grid of secondary frequency standards of IR molecular absorptions, frequency measured with these high-resolution spectrometers. More important, we have directly generated an IR frequency comb around
75
76 3 pm by DFG conversion of an OFCS. The generated comb can be employed both as a frequency ruler and as a direct source for molecular spectroscopy.
Keywords: mid IR, optical frequency comb; difference-frequency generation; quantum-cascade laser; molecular spectroscopy.
1. Introduction The molecular “fingerprint” region, roughly located in the 2.5-10 ,urn interval of the IR spectrum, can be considered the natural “gateway” for any molecular-based spectroscopic study or sensing. Indeed, the strongest rovibrational transitions generally lie in this range, for most simple molecules, thus guaranteeing a high detection sensitivity. In addition, Doppler-limited linewidths are narrower than in the visible/near-IR range, thus providing a better selectivity. However, very few tunable sources have been available until recently, thus favoring overtone-transitions-based investigations, relying on relatively cheap and compact near-IR telecom sources. Nonetheless, high-quality frequency standards have been developed, as He-Ne/CH4 laser, relying on fortuitous coincidences between laser lines and molecular transitions. The introduction, only a few years ago, of the OFCS and the parallel development of high-Q frequency standards more and more aiming at UV frequencies, with a more favorable A u / u value, has suddenly revolutionized frequency metrology. Molecular standards have been quickly replaced by trapped-ions-based ones and nowadays neutral atoms trapped in optical lattices are also under study and very promising. All this progress is quickly pushing frequency standards towards values of 10-16-10-17 for A u / u and even better values are already foreseen. If new ideas and novel optical technologies have suddenly changed the world and the perspectives of frequency metrology, also the community of people using spectroscopy has literally boomed. In particular, trace molecular sensing is becoming the primary tool for any quantitative assessment in environmental sciences, like greenhouse effect studies, atmospheric studies, anthropogenic as well as natural release of gases in the atmosphere, or also homeland security problems, just to mention some. The common requirements for these and other spectroscopic applications are always high resolution (to get high selectivity) and high sensitivity (to get highly accurate concentration values). For all such applications, as explained above, the spectral window of choice is the IR, and the primary concern is not the absolute frequency determination. On the other hand, a steep development has also been undergone by IR technologies. More specifically, at least two classes of new coherent sources, emitting in this spectral window, have emerged: sources
77 based on nonlinear generation in periodically-poled crystals and QCLs. In the first class are included optical parametric oscillators (OPOs) and DFG radiation sources. They are both very widely tunable sources, OPOs being mainly limited by the nonlinear crystal transparency range and DFG sources by the crystal or by the overall tunability range of pump/signal lasers. Instead, QCLS’ are not widely tunable but proper design of the quantum well structure may have them emitting in the range that roughly goes from 3.5 to several hundreds microns. Continuous-wave (CW) midIR QCLs need generally to be operated around liquid-N2 temperatures, but also room-temperature operation has recently been demonstrated.2 All such sources are very species-selective because linewidths are generally several orders of magnitude narrower than mid-IR Doppler profiles and often sufficient to detect saturated-absorption line shape^.^^^ Moreover, very high ~ > ~the very molecular detection sensitivities have been d e m ~ n s t r a t e dand wide tunability range allows to freely move throughout this rich portion of the spectrum containing fundamental ro-vibrational bands. It is also worth noticing that the mid-IR spectral window is “naturally” endowed with an ultra-wide comb of lines, represented by the manifolds of ro-vibrational transitions that can be easily saturated and that often have natural widths of a few tens of Hertz. Therefore, whatever the OFCS IR extension is realized, once the lines of interest are measured, they can be directly used as secondary frequency standards, similarly to what has been done until now with I2 lines. In this completely new situation, concerning the midIR spectral coverage, moderate-quality standards, easy to realize wherever is required in the IR, are probably the right choice for the ever increasing community of spectroscopy end-users. Partly to this purpose, several groups have been working, in the last few years, to an IR extension of OFCSs. So far, direct broadening of the spectrum of fs mode-locked lasers through highly-nonlinear optical fibers has succeeded in extending combs up to a 2.3 pm ~ a v e l e n g t hFor . ~ longer wavelengths, a few alternative schemes have been devised, essentially based on parametric generation processes in nonlinear crystals. A 270-nm-span frequency comb a t 3.4 pm has been realized by DFG between two spectral peaks emitted by a single uniquely-designed Ti:sapphire fs laser.8 In our group we have focused the attention onto the development of DFG- and QCL-based spectrometers and we will review, in the next paragraphs, the main results achieved. We report three different schemes which exploit a nonlinear optical process to transfer the metrological performance of a visible/near-IR OFCS to the mid IR. In the first scheme (Sec. 2), the
78
metrological performance of a Ti:sapphire OFCS is extended to the mid IR by phase-locking the pump and signal lasers of a DFG source to two nearIR teeth of an optical comb. Then, the generated IR radiation is used for high-resolution spectroscopy providing absolute frequency measurements of molecular lines at 4 pm. However, a drawback of this approach is the impossibility of comb-referencing for laser sources directly emitting in the mid-IR, such as QCLs. In Secs. 3 and 4 we demonstrate two novel schemes that overcome this limitation, based respectively on optical parametric upand down-conversion. In Sec. 3 a QCL at 4.43 pm has been used for producing near-IR radiation at 858 nm by means of sum-frequency generation with a Nd:YAG source in a periodically-poled LiNbO3 (PPLN) nonlinear crystal. The absolute frequency of the QCL source has been measured by detecting the beat note between the sum frequency and a diode laser at the same wavelength, while both the Nd:YAG and the diode laser were referenced to the OFCS. Vice versa, in Sec. 4 a frequency comb is directly created at 3 pm by nonlinear mixing of a near-IR fiber-based OFCS with a CW laser.g Possible applications for the generated comb are as a clockwork to transfer IR-frequency standards to other spectral regions, as a frequency ruler for high-precision molecular spectroscopy or telecommunications, and as a direct source for molecular spectroscopy. lo 2. DFGat 4 p m
Our OFCS-referenced DFG source at 4 pm is described About 200 p W of idler radiation at 4.2 pm is generated by nonlinear frequency mixing in a PPLN crystal of about 130 mW from a diode laser (pump laser) operating between 830 and 870 nm, and about 5 W from a fiber-amplified Nd:YAG laser (signal laser) at 1064 nm. We follow the scheme13 shown in detail in Fig. 1, to control the frequency and phase of the generated IR radiation against our mode-locked Ti:sapphire-laser-based OFCS, which covers an octave in the visible/near-IR region (500t1100 nm). Both pump and signal lasers are beaten with the closest tooth of the OFCS ( N p and N,), with residual R F beat notes Aupc and Ausc respectively. The contribution of the OFCS carrier-envelope-offset (CEO) frequency u, is canceled out of these beat notes by standard RF mixing, yielding Aupc- u, and Au,, - u, respectively. A low bandwidth (- 10 Hz) phase-locked-loop (PLL1) is used to control the long-term frequency fluctuations and drifts of the Nd:YAG laser. As a result, the signal-laser frequency is u, = N,u, + y o ,where y o is the R F frequency of the local oscillator used
79
v,=Nsvr+vLO To DFG/SFG
To
Fig. 1.
Apo3
Schematic of the OFCS-DFG/SFG phase lock
in the PLL1 loop, and vr = 1 GHz. In order to control the frequency and phase of the diode laser against the Nd:YAG one, a direct-digital synthesis (DDS) multiplying the A^sc — v0 frequency by a factor Np/Ns is used as a local oscillator in a second PLL circuit (PLL2) with a wide bandwidth (~ 2 MHz). Then the pump frequency is vp = (Np/Ns)vB, without any contribution from the OFCS parameters (v0 and vr). As a consequence, the absolute frequency of the generated idler radiation is given by Vi-vp-vs
= (Np -
N,
(1)
with a precision and accuracy limited only by the reference oscillator of our OFCS. The latter consists of a Rb/GPS-disciplined 10-MHz quartz with a stability of 6 • 10~13 at 1 s and a minimum accuracy of 2 • 10~12. Moreover, continuous scans of i/j can be performed by properly sweeping vr. Because in the above DDS-PLL scheme, the pump laser linewidth Az/p is a factor Np/Ns higher than the narrow Nd:YAG linewidth Az^ s , a residual (Np/Ns — l)A^ 5 idler linewidth is expected. We have measured a idler linewidth AI/J ~ 11 kHz by coupling the OFCS-locked 4-/zm beam to a high-finesse optical cavity (FSR=150 MHz, finesse > 17000), as shown in Fig. 2. It is more than 30 times narrower than the DFG without any OFCS
80
Voigt fit with fixed Lorentz
5 *
1
C
._
2
‘E C
-
+F e
z n LL
Frequency (20 kHz/div)
Fig. 2. High-finesse Fabry-Perot transmission of the OFCS-DFG source at 4 pm and Voigt function fit. Total acquisition time 5 ms. The fit that takes into account a Lorentz contribution from the Fabry-Perot with a fixed linewidth (9 kHz measured with CRDS) and a Doppler contribution from the idler radiation. The linewidth of the OFCS-DFG source extrapolated by the fit is 10.8(1) kHz.
control. Such a narrow linewidth can satisfy most of the spectroscopic needs and can still be improved with a proper choice of signal/pump lasers. This idler radiation has been used both for high-precision“ and highsensitivity” spectroscopy of CO2 molecular transitions around 4 pm. In the former case, we have performed saturated-absorption spectroscopy with a medium-finesse optical cavity (FSR=1.3 GHz, finesse > 500) of even verylow-populated rotational levels ( J > 80). Absolute frequency measurements of these transitions with an accuracy of about are proposed as a “natural” grid of secondary frequency standards in this spectral region. For trace-gas detection, we perform cavity-ring-down spectroscopy (CRDS) by coupling the idler beam to a high-finesse optical cavity (FSR=150 MHz, finesse > 17000), filled with the COz gas. In this case, the OFCS control of the DFG spectrometer helps not only t o get a narrow-linewidth idler radiation (thus increasing the cavity-coupling), but also to interrogate the molecular absorption at the same resonant frequency for long times, due to the high reproducibility of the OFCS-referenced IR frequency. In this way, minimum absorbances a L of the order of few parts in lo8 (i.e. linestrengths of about cm) can be achieved with few hours of integration. The power of this technique can be used, e.g., to detect C02 isotopologues with very low natural abundance, or to search for highly-forbidden C02 transitions, as those due t o the wave-function symmetry under the bosonic l60exchange.
81
3. QCL-based spectrometer QCLs operating in the mid-IR region can represent a valid alternative to DFG systems, especially when the application requires high emission power and very compact designs. It makes QCLs very appealing not only for highsensitivity spectroscopy experiments, but also for a huge variety of industrial and commercial applications. On the other hand, their capabilities cannot be fully exploited at present, due to the lack of precise references in most of the IR region used to control their absolute frequency. Here, we illustrate an experiment14 which overcomes this problem: the frequency of a 4.43-/Ltm QCL was measured against our Ti:sapphire OFCS by means of a parametric up-conversion process. The set-up is shown in Fig. 3. The QCL is a CW, liquid-N2-cooled,
1||P Oetedtor Nd-.YAO @ 1C84 nm
Dsfectal Grating $ • = OFB4SG Sgssr
Sum-Frequency Frequency-Comb
I -'"" •to Filter JSSSBrar
Fig. 3. Schematic of the experimental apparatus, focused on the SFG generation process providing the optical link between the QCL and the OFCS.
distributed-feedback (DFB) device at 4.43 /mi. The collimated QCL beam is split into two parts: 1 mW is used for CC-2 Doppler-absorption spectroscopy and 2 mW are used for the nonlinear up-conversion process. The latter is achieved by mixing the QCL and a fiber-amplified Nd:YAG laser in a PPLN crystal for a sum-frequency generation (SFG) process to produce 858-nm radiation. With about 1.2 W of Nd:YAG power and only 2 mW of QCL
82 radiation incident on the nonlinear crystal, about 10 p W of SFG radiation has been obtained. This radiation is beaten with an external-cavity diode laser (ECDL) working at the same wavelength, yielding a beat note A U + ~ that can be easily counted (40 dB S/N at 500 kHz resolution bandwidth). Both the ECDL and the Nd:YAG lasers are the same of the DFG source at 4 pm, and are phase-locked to the OFCS following the DDS-PLL scheme described in Sec. 2 . Then, the QCL absolute frequency can be expressed as:
where the symbols have the same meaning as in Sec. 2 . Moreover, since is the beat note between the sum frequency v+ = vi v, and up, it can be used t o measure the phase/frequency noise of the QCL similarly to what has been described for the DFG source at 4 pm. We used this OFCS-referenced QCL for several absolute frequency measurements of two 13C02 Doppler-broadened ro-vibrational transitions, the (OOol-00'0) P(30) and the (Ol11-0l1O) P(17). A recording of the latter, weaker line is shown in Fig. 4. Each point of the trace results from the
+
0.50
'
'
67683.5
67683.6
67683.7
67683.8
67683.9
Absolute frequency [GHz]
Fig. 4. Absolute frequency measurement of the 1 3 C 0 2 (Ol11-0l1O) P(17) line. The gas pressure in the cell was 3 mbar. A Voigt fit of the data is also shown.
simultaneous measurements of the amplitude of the absorption signal, and the QCL frequency measured by counting with a spectrum analyzer. The acquisition time for each point is 500 ms, during which an average on both the amplitude and frequency measurements is performed. The uncertainty associated to the absolute frequency measurement of each point
83
is due to the frequency fluctuations of the free-running QCL during the 500 ms single-point acquisition time. For our measurements we obtain a frequency uncertainty of about 2 MHz. Several spectra have been acquired for the same transition, even at slightly different gas pressures. Each set of data has been fitted t o a Voigt profile (Fig. 4) to determine the corresponding line-center frequency. The final precision of these absolute frequency measurements is 3 . mainly limited by the above mentioned QCL jitter. This number can be heavily improved (at least 3 orders of magnitude) with proper frequency stabilization of the QCL. This upgrade will match the implementation of high-precision spectroscopic techniques such as Doppler-free detection. 4. 3-pm comb generation
The apparatus devised to create the 3-pm frequency comb,g is shown in Fig. 5. The nonlinear down-conversion process occurs in a PPLN crystal ECDL
I
I
I
4 Fig. 5. Layout of the optical table. A 3-pm frequency comb is created by DFG in a PPLN crystal between a near-IR OFCS and a CW laser. A fast, 100-pm-diameter HgCdTe detector is used to characterize the generated mid-IR comb.
(with a period around 30 pm) between a near-IR OFCS and a CW tunable laser. The generated mid-IR frequency comb covers the region from 2.9 to 3.5 pm in 180-nm-wide spans with a 100-MHz mode spacing and
84
keeps the same metrological performance as the original comb source. Such a scheme can be easily implemented in other spectral regions by use of suitable pumping sources and nonlinear crystals. The near-IR OFCS is an octave-spanning (1050-2100 nm) fs modelocked fiber-laser-based system, but for the DFG process only the OFCS fraction (25 mW) covering the 1500-1625 nm interval is used as combsignal laser. The power of this comb-signal beam is enhanced by amplifying it with an external Er-doped fiber amplifier (EDFA). The amplified combsignal beam has an overall power of 0.7 W and spans from 1540 to 1580 nm with a 100 MHz spacing, corresponding to nearly 50000 teeth (i.e. about 14 p W per tooth). The pump beam is an external-cavity diode laser (ECDL) at 1030-1070 nm, amplified up to 0.7 W by an Yb-doped fiber amplifier (YDFA). Both pump and comb-signal lasers are mixed in a PPLN crystal, whose period and temperature are chosen depending of the pump wavelength satisfying the quasi-phase-matching (QPM) condition for the center wavelength (1560 nm) of the comb-signal laser. The comb-idler beam is detected by filtering out the unconverted near-IR light and focusing it on to a liquid-Nz-cooled, 150-MHz-bandwidth HgCdTe detector. In this way, a RF beat note at v, = 100 MHz is recorded by a spectrum analyzer, which is the sum of the beat signals between all pairs of consecutive teeth in the generated DFG comb. The latter has a span of 180 nm (5 THz), limited by the comb-signal coverage, and is centered in the 2.9 to 3.5 pm interval, depending on the pump wavelength used. The teeth on both sides of the near-IR comb are involved in many DFG processes, with a conversion efficiency decreasing according to the well-known sinc2 law.15 The overall measured power of the 5-THz-spanning radiation is about 5 pW. This value corresponds to a power of nearly 100 pW per mode of the mid-IR comb. We phase-lock the pump laser to the closest tooth of the near-IR OFCS, in order to cancel out the CEO vo frequency in the generated mid-IR comb and to fix the beat-note frequency Avpcbetween the pump and the closest near-IR comb frequencies. In this way the frequency of the generated IR modes is
and, hence, with the same metrological performance of v,. In the following we discuss the application of the mid-IR comb as an absolute frequency ruler at 3 pm. For this purpose, we have beaten this comb with a CW laser at 3 pm. The CW radiation is generated by a second DFG process which uses most of the same set-up used to produce the
85 mid-IR comb. Indeed, a CW extended-cavity diode laser at 1520-1570 nm, is amplified by the EDFA simultaneously with the fraction of the near-IR OFCS used to generate the 3-pm comb. As described above, this amplified CW laser is DFG mixed with the 1-pm pump laser to generate mW-powerlevel CW idler radiation around 3 pm, co-propagating with the DFG comb. A beat note Aucwc between the CW DFG radiation and the mid-IR comb can be detected by sending directly the generated light to the HgCdTe fast detector. As the frequency of each tooth of the mid-IR comb is well known (Eq. 3), the mid-IR CW laser frequency can be measured by counting Aucwc. Furthermore, because the 1-pm pump laser is comb-locked, Aucwc can be used to phase-lock the mid-IR CW radiation to the DFG comb by feeding back proper phase corrections to the 1.5-pm laser. The S/N ratio of Aucwc was measured when the ECDL wavelength was tuned from 1540 to 1570 nm ( ( X ~ ) C W from 3.22 to 3.35 pm), in order to characterize the effective DFG-comb span suitable for use in phase-locked systems and frequency counting. Such value reaches a maximum of 40 dB at the center wavelength, while decreases almost symmetrically down to less than 20 dB at the upper and lower edges, limiting in principle to about 130 nm the interval in which a mid-IR source can be locked. Actually, the 180-nm span can be fully exploited, as stronger beat notes are expected when two different DFG apparata are used for the mid-IR comb and CW lasers. Moreover the beat-note S/N can be improved if filtering of the comb modes not contributing to the beat is done before and after the nonlinear conversion. The described mid-IR DFG comb has demonstrated to be a suitable absolute frequency ruler in this spectral window and may be strategic for future metrological applications with direct mid-IR lasers as QCLS." On the other hand, the generation of a frequency comb in the mid IR leads straightforwardly to consider its use as a direct spectroscopic source. In this sense, several schemes involving coherent coupling to high-finesse cavities, as well as Fourier-transform molecular spectroscopy schemes have already been demonstrated in the near IR, and may be now extended, taking advantage of the detection sensitivities achievable in the fingerprint region.
References 1. J. Faist, F. Capasso, D. L. Sivco, C. Sirtori, A. L. Hutchinson and A. Y. Cho, Science 264, p. 553 (1994).
2. M. Beck, D. Hofstetter, T. Aellen, J. Faist, U. Oesterle, M. Ilegems, R. Gini and H. Melchoir, Science 295, p. 301 (2002). 3. J. T. Remillard, D. Uy, W. H. Weber, F. Capasso, C. Gmachl, A. L. Hutchin-
86
4. 5. 6. 7. 8.
9. 10. 11. 12. 13. 14. 15. 16.
son, D. L. Sivco, J. N. Baillargeon and A. Y. Cho, Opt. Express 7,p. 243 (2000). A. Castrillo, E. De Tommasi, L. Gianfrani, L. Sirigu and J. Faist, Opt. Lett. 31, p. 3040 (2006). Y. A. Bakhirkin, A. A. Kosterev, R. F. Curl, F. K. Tittel, D. A. Yarekha, L. Hvozdara, M. Giovannini and J. Faist, Appl. Phys. B 82, p. 146 (2006). S. Borri, S. Bartalini, P. De Natale, M. Inguscio, C. Gmachl, F. Capasso, D. L. Sivco and A. Y. Cho, Appl. Phys. B 85, p. 223 (2006). I. Thomann, A. Bartels, K. L. Corwin, N. R. Newbury, L. Hollberg, S. A. Diddams, J. W. Nicholson and M. F. Yan, Opt. Lett. 2 8 , p. 1368 (2003). S. M. Foreman, A. Marian, J. Ye, E. A. Petrukhin, M. A. Gubin, 0. D. Mucke, F. N. C. Wong, E. P. Ippen and F. X. Kartner, Opt. Lett. 30, p. 570 (2005). P. Maddaloni, P. Malara, G. Gagliardi and P. De Natale, New J . Phys. 8, p. 262 (2006). M. J. Thorpe, K. D. Moll, R. Jason Jones, B. Safdi and J . Ye, Science 311, p. 1595 (2006). D. Mazzotti, P. Cancio, G. Giusfredi, P. De Natale and M. Prevedelli, Opt. Lett. 30, p. 997 (2005). D. Mazzotti, P. Cancio, A. Castrillo, I. Galli, G. Giusfredi and P. De Natale, J . Opt. A 8, p. S490 (2006). H. R. Telle, B. Lipphardt and J. Stenger, Appl. Phys. B 74, p. 1 (2002). S. Bartalini, P. Cancio, G. Giusfredi, D. Mazzotti, P. De Natale, S. Borri, I. Galli, T. Leveque and L. Gianfrani, Opt. Lett. 32, p. 988 (2007). P. Maddaloni, G. Gagliardi, P. Malara and P. De Natale, Appl. Phys. B 80, p. 141 (2005). F. Capasso, C. Gmachl, R. Paiella, A. Tredicucci, A. L. Hutchinson, D. L. Sivco, J. N. Baillargeon, A. Y. Cho and H. C. Liu, IEEE J . Sel. Top. Quantum Electron. 6, p. 931 (2000).
PRECISION MEASUREMENT AND APPLICATIONS
This page intentionally left blank
PRECISION GRAVITY TESTS BY ATOM INTERFEROMETRY G. M. TINO*, A. ALBERTI, A. BERTOLDI, L. CACCIAPUOTI~,
M. DE ANGELISt, G. FERRARI, A. GIORGINI, V. IVANOV, G. LAMPORESI, N. POLI, M. PREVEDELLIf, F. SORRENTINO
Dipartimento di Fisica and L E N S Laboratory - Universitb d i Firenze Istituto Nazionale d i Fisica Nucleare, Sezione d i Firenze via Sansone 1, Polo Scientifico, I-5001 9 Sesto Fiorentino (Firenze), Italy We report on experiments based on atom interferometry to determine the gravitational constant G and test the Newtonian gravitational law at micrometric distances. Ongoing projects to develop transportable atom interferometers for applications in geophysics and in space are also presented.
1. Introduction
Advances in atom interferometry led to the development of new methods for fundamental physics experiments and for applications. In particular, atom interferometers are new tools for experimental gravitation as, for example, for precision measurements of gravity acceleration [l]and gravity gradients [2], for the determination of the Newtonian constant G [3,4], for testing general relativity [5,6] and l / r 2 law [7-lo], and for possible applications in geophysics. Ongoing studies show that future experiments in space will allow to take full advantage of the potential sensitivity of atom interferometers [ll].The possibility of using atom interferometry for gravitational waves detection was also investigated (see [12] and references therein). In this paper, we report on experiments we are performing using atom interferometry t o determine G and test the Newtonian gravitational law a t *E-mail:
[email protected] - Web: www.lens.unifi.it/tino tpermanent address: ESA Research and Scientific Support Department, ESTEC, Keplerlaan 1- P.O. Box 299, 2200 AG Nordwijk ZH, The Netherlands $On leave from: Istituto Cibernetica CNR, 80078 Pozzuoli (Napoli), Italy f Permanent address: Dipartimento di Chimica Fisica, Universita di Bologna, Via del Risorgimento 4, 40136 Bologna, Italy
89
90
micrometric distances. We also present ongoing projects to develop transportable atom interferometers for applications in geophysics and in space. 2. Determination of G by atom interferometry
The Newtonian constant of gravity G is one of the most measured fundamental physical constants and at the same time the least precisely known. The extreme weakness of the gravitational interaction and the impossibility of shielding the effects of gravity make it difficult t o measure G, while keeping systematic effects well under control. Despite the numerous experiments, the uncertainty on G has improved only by one order of magnitude in the last century [13]. Many of the experiments performed to date are based on the traditional torsion pendulum method, direct derivation of the historical experiment performed by Cavendish in 1798. Recently, many groups have set up new experiments based on different concepts and with completely different systematics. However, the most precise measurements available today still show substantial discrepancies, limiting the accuracy of the 2006 CODATA recommended value for G t o 1 part in lo4. From this point of view, the realization of conceptually different experiments can help to identify still hidden systematic effects and therefore improve the confidence in the final result. We use atom interferometry to perform precision measurements of the differential acceleration experienced by two samples of laser-cooled rubidium atoms under the influence of nearby tungsten masses. In our experiment, specific efforts have been devoted to the control of systematic effects related to atomic trajectories, positioning of source masses, and stray fields. In particular, the high density of tungsten and the distribution of the source masses are crucial in our experiment to compensate for the Earth gravity gradient and reduce the sensitivity of the measurement to the initial position and velocity of the atoms. The measurement, repeated for two different configurations of the source masses, is modeled by a numerical simulation which takes into account the mass distribution and the evolution of atomic trajectories. The comparison of measured and simulated data provides the value of the Newtonian gravitational constant G. Proof-of-principle experiments with similar schemes using lead masses were already presented in [3,4]. In our interferometer, laser pulses are used to stimulate s7Rb atoms on the two-photon Raman transition between the hyperfine levels F = 1 and F = 2 of the ground state [14]. The light field is generated by two counter-propagating laser beams with wave vectors kl and kz E -kl aligned along the vertical direction. The laser frequencies v1
91
and 1/2 match the resonance condition v\—vi = v§, where hi/o is the energy associated to the F = 1 —)• F = 2 transition. The atom interferometer, obtained with a 7T/2 — TT—7T/2 sequence of Raman pulses, drives the atoms on two independent paths along which the quantum mechanical phases of the atomic wavepackets independently evolve. In the presence of a gravity field, atoms experience a phase shift $ = k-g T2, where k = kl—k2, 2T is the duration of the pulse sequence, and g is gravity acceleration. A measurement of the phase $ is equivalent to an acceleration measurement. The gravity gradiometer consists of two absolute accelerometers operated in differential mode. Two spatially separated atomic clouds aligned along the vertical direction are simultaneously interrogated by the same ?r/2 — ir — ?r/2 pulse sequence. The difference of the phase shifts detected on each interferometer provides a direct measurement of the differential acceleration induced by gravity on the two atomic samples. This method has the major advantage of being highly insensitive to noise sources appearing in common mode on both interferometers. In particular, any spurious acceleration induced by vibrations or seismic noise on the common reference frame identified by the vertical Raman beams is efficiently rejected by the differential measurement technique. Figure 1 shows a schematic of the experiment. The gravity gradiometer set-up and the configurations of the source masses (Ci and C^) used for the G measurement are visible. The atom interferometer apparatus and
upper gravimeter
/
1 ^
lower gravimeter
1 1
detection beams
Fig. 1. Schematic of the experiment showing the gravity gradiometer set-up with the Raman beams propagating along the vertical direction. For the measurement of G, the position of the source masses is alternated between configuration C\ (left) and C-2 (right).
92
the source masses assembly are described in detail elsewhere [3,15]. In the vacuum chamber at the bottom of the apparatus, a magneto-optical trap (MOT) collects rubidium atoms from the vapor produced by getters. After turning the MOT magnetic field off, the atomic sample is launched vertically along the symmetry axis of the vacuum tube by using the moving molasses technique. The gravity gradient is probed by two atomic clouds moving in free flight along the vertical axis of the apparatus and simultaneously reaching the apogees of their ballistic trajectories a t 60 cm and 90 cm above the MOT. Such a geometry, requiring the preparation and the launch of two samples with high atom numbers in a time interval of about 100 ms, is achieved by juggling the atoms loaded in the MOT. The interferometers are realized at the center of the magnetically shielded vertical tube shown in Fig. 1. The three-pulse interferometer has a duration of 2T = 320ms. The 7r pulse lasts 48 ps and occurs 5 ms after the atomic clouds reach their apogees. In this configuration, only one pair of counterpropagating laser beams with frequencies ul and v2 and crossed linear polarizations is able to stimulate the atoms on the two-photon transition. At the end of their ballistic flight, the population of the ground state is measured by selectively exciting the atoms in both hyperfine levels of the ground state and detecting the light-induced fluorescence emission. We typically detect lo5 atoms on each rubidium sample at the end of the interferometer sequence. Even if the phase noise induced by vibrations washes out the atom interference fringes, the signals simultaneously detected on the upper and lower accelerometer remain coupled and preserve a fixed phase relation. Therefore, when the trace of the upper accelerometer is plotted as a function of the lower one, experimental points distribute along an ellipse. The differential phase shift is then obtained from the eccentricity and the rotation angle of the ellipse fitting the data [16]. The Allan deviation shows the typical behavior expected for white noise. The instrument has a sensitivity of 140 mrad at 1s of integration time, corresponding to a sensitivity to differential accelerations of 3.5 . lop8 g in 1s. The source masses [15] are composed of 24 tungsten alloy (INERMET IT180) cylinders, for a total mass of about 516 kg. They are positioned on two titanium platforms and distributed in hexagonal symmetry around the vertical axis of the tube (see Fig. 1). The value of G was determined from a series of gravity gradient measurements performed by periodically changing the vertical position of the source masses between configuration C1 and C2 while keeping the atomic trajectories fixed. Because of the high density of tungsten, the gravitational
93 field produced by the source masses is able to compensate for the Earth gravity gradient. As a consequence, the acceleration becomes less sensitive to the positions of the atomic clouds around extremal points, allowing for a better control of systematic effects and a relaxation of the requirements on the knowledge of atomic trajectories. Figure 2 shows a data sequence used for the measurement of G. Each phase measurement is obtained by fitting a 24-point scan of the atom interference fringes t o an ellipse. After an analysis of the error sources affecting our measurement, we obtain a value of G = 6.667.10-11 m3 kg-l s - ~ with , a statistical uncertainty of f O . O 1 l . l O - l l m3 kg-' s-' and a systematic uncertainty of 3~0.003. m3 kg-' s - ~ .Our measurement is consistent with the 2006 CODATA value within one standard deviation. The main contribution to the systematic error on the G measurement derives from the positioning accuracy of the source masses. This error will be reduced by about one order of magnitude by measuring the position of the tungsten cylinders with a laser tracker. Eventually, uncertainties below the 10 ppm level could be reached with this scheme using for the source mass a material with a higher density homogeneity.
"Q.3
0.4
0.5
0 F'
interletence signal (lower ,&d)
j 0.3 0 4 0.5 0.6;' jlnterlerence signal (lower,cioud)
19
13
0
5000
10000
Time (s)
Fig. 2. Typical data set showing the modulation of the differential phase shift measured by the atomic gravity gradiometer when the distribution of the source masses is alternated between configuration C1 and C Z .Each phase measurement is obtained by fitting a 24-point scan of the atom interference fringes to an ellipse.
94
3. Precision gravity measurements at pm scale with
laser-cooled Sr atoms in an optical lattice The extremely small size of atomic sensors can lead to applications for precision measurements of forces at micrometer scale. The investigation of forces at small length scales is indeed a challenge for present research in physics for the study of surfaces, of the Casimir effect, and in the search for deviations from Newtonian gravity predicted by recent theories beyond the standard model. We showed [lo] that using laser-cooled **Sr atoms in an optical lattice, persistent Bloch oscillations can be observed for a time 10 s, because of remarkable properties of immunity of this atom from perturbations due to stray fields and interatomic collisions. This system can reach an unprecedented sensitivity as sensor to measure gravity acceleration on micrometer scale with ppm precision opening the way to the investigation of small-scale gravitational forces in regions so far unexplored. The experiment starts with trapping and cooling 5 x lo7 *'Sr atoms at 3 mK in a magneto-optical trap (MOT) operating on the lSo-lP1 blue resonance line at 461 nm. The temperature is then further reduced by a second cooling stage in a red MOT operating on the 1S0-3P1 narrow transition at 689 nm and finally we obtain 5 x lo5 atoms at 400 nK. After this preparation phase, the red MOT is switched off and a one-dimensional optical lattice is switched on adiabatically in 50 ps. The lattice potential N
N
N
+
L----/--d 10
12
1016 1018
3020 3022
7020 7022
t (W
Fig. 3. Bloch oscillation of "Sr, atoms in the vertical 1-dimensional optical lattice under the effect of gravity. Two quantities are extracted from the analysis of the data: The vertical momentum of the oscillating atoms (a) and the width of the atomic momentum distribution (b).
95
is originated by a single-mode frequency-doubled Nd:YV04 laser (XL = 532 nm). The beam is vertically aligned and retro-reflected by a mirror producing a standing wave with a period X L / ~= 266 nm. We obtain a diskshaped sample of lo5 atoms at T 400 nK with a vertical rms width of 12 pm and a horizontal radius of 150 pm. We observe Bloch oscillations in the vertical atomic momentum by releasing the optical lattice at a variable delay, and by imaging the atomic distribution after a fixed time of free fall. Figure 3 shows the signal recorded for 7 s, corresponding t o 8000 oscillations. The coherence time for the Bloch oscillation is 12 s. These values are the highest ever observed for Bloch oscillations in atomic systems. Measuring the oscillation frequency we determine the vertical force on the atoms, that is, Earth gravity with a resolution of 5 x lop6. In the effort to increase the sensitivity, recently we investigated strontium atoms trapped in phase-modulated optical lattices. We found that we can induce a broadening of the atomic distribution in the lattice potential with a phase modulation of the lattice at frequencies multiple of Bloch frequency. We observed a resonant broadening up to the 5th harmonic which corresponds to a hop through 5 lattice sites (Fig. 4). All the resonance spectra exhibit a Fourier-limited width for excitation times as long as 2 s. The resulting high-resolution measurement of Wannier-Stark levels of the atomic wavefunction in the gravity potential allows to determine the local gravity acceleration with a relative precision lop6. When studying atom-surface interactions, one key point is the precision of sample positioning close to the surface. In our experiment, the optical lattice is also used for an accurate positioning of the sample close to the
-
--
N
N
-
2
. ii4n
i'
s., iiso
2298
23W
modulation frequency (Hz)
Fig. 4. Spatial width of the atomic distribution for 2 s phase modulation of the optical lattice at the lst, 2nd and 4th harmonics of the Bloch frequency.
96
surface. We translate the atomic sample along the lattice axis by applying a relative frequency offset to the counterpropagating laser beams producing the lattice. In this way, we place the atoms close to a transparent test surface placed ~ 45 mm far from the MOT. We measure the atoms number and the phase of the Bloch oscillation with absorption imaging after bringing the atoms back to the original position. In Fig. 5.a), we show the number of atoms recorded after an elevator round-trip, as a function of the travelled distance. A sudden drop, corresponding to the loss of atoms hitting the test surface, is clearly visible. The plot provides a measurement of the vertical size of the atomic sample. This scheme directly applies to transparent materials. In order to study metallic surfaces, the atomic sample in the optical lattice is displaced by means of optical components mounted on micrometric translation stages. In that case, the optical lattice is produced by retroreflecting the laser beam on the test surface itself. The minimum attainable atom-surface distance is limited by the vertical size of the atomic distribution. For experiments at distances below 10 /zm, we compress our sample using an optical tweezer. This is obtained with a strongly astigmatic laser beam with the vertical focus centered on the atoms. Figure 5.b) shows an image of the atoms trapped in the optical tweezer. Deviations from the Newtonian gravity law are usually described assuming a Yukawa-type potential with two parameters, a giving the relative strength of any new effect compared to Newtonian gravity and A its range. Experiments searching for possible deviations have set bounds for the parameters a and A. Recent results using microcantilever detectors lead to extrapolated limits a ~ 104 for A ~ 10 /zm [17,18]. The small size and high sensitivity of the atomic probe allow a direct, model-independent measure-
Fig. 5. a) Number of atoms in the lattice versus vertical displacement. The inset shows the region close to the test surface. The vertical displacement is varied by changing the duration of the motion at uniform velocity, b) In-situ absorption image of the atomic sample trapped in the optical tweezer. Untrapped falling atoms are also visible.
97 ment at distances of a few pm from the source mass with no need for modeling and extrapolation as in the case of macroscopic probes. This allows to directly access unexplored regions in the Q - X plane. Also, in this case quantum objects are used to investigate gravitational interaction. If we consider a thin layer of a material of density p and thickness d , the Newtonian gravitational acceleration due to the source mass is a = 2nGpd; for d 10 pm and p 2: 20 g/cm3, as for gold or tungsten, the resulting acceleration is a 10-l' msP2. Measuring v~ at a distance of 10 pm from the surface will then provide a direct test of present constraints on a [18]. For smaller distances, around 5 pm, it is possible to improve present limits on (Y by more than two orders of magnitude in the corresponding X range. Even shorter distances could probably be accessed, also considering a related scheme based on a Sr lattice clock [19]. Non-gravitational effects (Van der Waals, Casimir forces), also present in other experiments, can be reduced by using a conductive screen and performing differential measurements with different source masses placed behind it. For this experiment we developed a source mass made of a sandwich of A1 and Au layers covered by an Au layer. This layer acts as the mirror t o produce the optical standing wave and as the conductive screen. Placing the atoms in proximity of the different materials with different mass densities, mass-dependent effects can be investigated. Also, by performing the experiment with different isotopes of Sr, having different masses but the same electronic structure, gravitational forces can be distinguished from other surface interactions.
-
-
4. Transportable atom gravimeters for geophysics applications and future experiments in space Inertial and rotational sensors using atom interferometry display a potential for replacing other state-of-the-art sensors for e.g. geophysics and space applications. The intrinsic benefits making direct use of fundamental quantum processes promise significant advances in performance, usability, and efficiency, from the deployment of highly optimized devices on satellites in space or from the use of ground based transportable devices. 4.1. Geophysics applications
The knwoledge of the value of g and its time and space variations is of interest to a wide field of physical sciences connected to geophysics and geodesy. Surface gravity measurements on local scale can help in under-
98
standing active tectonics areas dynamics, faulting during the interseismic phase, fluid migration/diffusion due to stress changes. A variety of maninduced physical and chemical processes are known to produce substantial vertical displacements of the Earth’s surface and are related to mass or fluid extraction or to subsurface pore fluid flows. Gravity measurements are used in acquifier and reservoirs monitoring, monitoring of fluid infiltration and water table rise near nuclear repositories and monitoring of subsidence and mining effects. Recently, micro-gravimetric observations have found an important field of application in volcanoes monitoring. Variations in the local gravity field were observed prior and during eruptive episodes of variable size at a number of volcanoes worldwide. It appears that crucial mass redistribution in geodynamics occur over time scales spanning the 1 - lo6 s interval and have amplitudes ranging from 10 microgals up to hundreds of microgals (I Gal = 1 cm/s2). Some of these phenomena are observed very early and occur before other phenomena as strain deformations or seismic signals. These considerations indicate that the continuous observation of the local gravity field using sensitive instrumentation with comparable accuracy on long periods is a major goal to be attained towards a better understanding of active volcanic systems and prediction of eruptive activity. Prototypes of transportable gravimeters based on atom interferometry have been realized by different labs. In Stanford, a transportable system was developed and used to measure the components of the gravity tensor. At JPL, a compact gravity gradiometer is being developed. A transportable gravimeter based on atom interferometry was developed for metrological applications at SYRTE in Paris. Our group is involved in these activities in the frame of a European STREP/NEST project (FINAQS) and with support from the Istituto Nazionale di Geofisica e Vulcanologia (INGV). We are developing a transportable atomic gravimeter that will be used for geophysics applications. Our main interest is in vulcanoes monitoring t o investigate the possibility of predicting eruptions. The high sensitivity and long term stability achievable with an atom gravimeter are important characteristics for this application. 4.2. Space applications
Future experiments in space will allow to take full advantage of the potential sensitivity of cold atom interferometers as acceleration or rotation sensors. Indeed, atomic quantum sensors can reach their ultimate performances if operated in space because of the extremely long achievable interaction time and the vibration-free environment.
99
In Europe, the interest on application of atomic quantum sensors in space is demonstrated by the activities initiated by ESA and by national space agencies, CNES, ASI, DLR. The study conducted by ESA on HYPER mission, proposed in 2000 in response to the call for the second and third Flexi-missions (F2/F3), showed the feasibility of cold atom interferometry in space both for inertial sensing and fundamental physics studies. SAI, a new project funded by ESA, started in 2007 [20]. The project intends t o exploit the potential of matter-wave sensors in microgravity for the measurement of acceleration, rotations, and faint forces. SAI aims to push present performances to the limits and t o demonstrate this technology with a transportable sensor which will serve as a prototype for the space qualification of the final instrument. The atom interferometer will be used to perform fundamental physics tests and t o develop applications in different areas of research (navigation, geodesy). These activities are financed by the HME Directorate in the frame of the ELIPS2 program. Several pieces of technology for this activity are common to those developed for other ESA projects based on cold atom clocks, namely, ACES (Atomic Clocks Ensemble in Space) and SOC (Space Optical Clocks). Different proposals based on the utilization of matter-wave interferometers and atomic clocks for fundamental physics studies were submitted to ESA in the context of the ”Cosmic Vision 2015-2025” program. The applications of atomic quantum sensors in space are interdisciplinary, covering diverse and important topics. In fundamental physics, space-based cold atom sensors may be the key for new experiments, e.g., accurate tests of general relativity, search for new forces, test of l / r 2 law for gravitational force at micrometric distances, neutrality of atoms. Possible applications can be envisaged in astronomy and space navigation (inertial and angular references), realization of SI-units (definition of kg, measurements of Newtonian gravitational constant G, h/m measurement), GALILEO and LISA technology, prospecting for resources and major Earth-science themes.
Acknowledgments For the experiment on G, G.M.T. acknowledges seminal discussions with M.A. Kasevich and J. Faller and useful suggestions by A. Peters. M. Fattori, T. Petelski, and J. Stuhler contributed to setting up the apparatus. This work was supported by INFN (MAGIA experiment), EU (contract RII3CT-2003-506350 and FINAQS STREP/NEST project), INGV, ESA, ASI.
100
References 1. A. Peters, K. Y. Chung and S. Chu, Nature 400,p. 849 (1999). 2. J. M. McGuirk, G. T. Foster, J. B. Fixler, M. J. Snadden and M. A. Kasevich, Phys. Rev. A 65,p. 033608 (2002). 3. A. Bertoldi, G. Lamporesi, L. Cacciapuoti, M. D. Angelis, M. Fattori, T. Petelski, A. Peters, M. Prevedelli, J. Stuhler and G. M. Tino, Eur. Phys. J . D 40,p. 271 (2006). 4. J. B. Fixler, G. T. Foster, J. M. McGuirk and M. Kasevich, Science 315, p. 74 (2007). 5. S. Fray, C. A. Diez, T. W. Haensch and M. Weitz, Phys. Rev. Lett. 93,p. 240404 (2004). 6. S. Dimopoulos, P. Graham, J. Hogan and M. Kasevich, Phys. Rev. Lett. 98, p. 111102 (2007). 7. G.M. Tino, in 2001: A Relativistic Spacetime Odyssey - Proceedings of J H Workshop, Firenze, 2001 (I. Ciufolini, D. Dominici, L. Lusanna eds., World Scientific, 2003). Also, Tino G. M., Nucl. Phys. B 113, 289 (2003). 8. S. Dimopoulos and A. A. Geraci, Phys. Rev. D 68,p. 124021 (2003). 9. D. M. Harber, J. M. Obrecht, J. M. McGuirk and E. A. Cornell, Phys. Rev. A 72,p. 033610 (2005). 10. G. Ferrari, N. Poli, F. Sorrentino and G. M. Tino, Phys. Rev. Lett. 97,p. 060402 (2006). 11. G. M. Tino, L. Cacciapuoti, K. Bongs, C. J. Bordk, P. Bouyer, H. Dittus, W. Ertmer, A. Gorlitz, M. Inguscio, A. Landragin, P. Lemonde, C. Laemmerzahl, A. Peters, E. Rasel, J. Reichel, C. Salomon, S. Schiller, W. Schleich, K. Sengstock, U. Sterr and M. Wilkens, Nucl. Phys. B (Proc. Suppl.) 166, p. 159 (2007). 12. G. Tino and F. Vetrano, Class. Quantum Grav. 24,2167 (2007). 13. P. J. Mohr and B. N. Taylor, Rev. Mod. Phys. 77-1,42 (2005). 14. M. Kasevich and S. Chu, Appl. Phys. B 54,p. 321 (1992). 15. G. Lamporesi, A. Bertoldi, A. Cecchetti, B. Dulach, M. Fattori, A. Malengo, S. Pettorruso, M. Prevedelli and G. Tino, Rev. Sci. Instrum. 78,p. 075109 (2007). 16. G. T. Foster, J. B. Fixler, J. M. McGuirk and M. A. Kasevich, Opt. Lett. 27, p. 951 (2002). 17. J. C. Long, H. W. Chan, A. B. Churnside, E. A. Gulbis, M. C. M. Varney and J. C. Price, Nature 421,p. 922 (2005). 18. S. J. Smullin, A. A. Geraci, D. M. Weld, J. Chiaverini, S. Holmes and A. Kapitulnik, Phys. Rev D 72,p. 122001 (2005). 19. P. Wolf, P. Lemonde, A. Lambrecht, S. Bize, A. Landragin and A. Clairon, Phys. Rev. A 75,p. 063608 (2007). 20. G. M. Tino et al., Space Atom Interferometers (SAI) (AO-2004-64), Proposal in response to ESA Announcement of Opportunity in Life and Physical Sciences and Applied Research Projects, ESA-AO-2004.
NOVEL SPECTROSCOPIC
APPLICATIONS
This page intentionally left blank
ON A VARIATION OF THE PROTON-ELECTRON MASS RATIO W. UBACHS', R. BUNING', E. J. SALUMBIDES', S. HANNEMANN', H. L. BETHLEM', D. BAILLY2, M. VERVLOET3, L. KAPERIs4,M. T. MURPHY' 'Laser Centre Vrije Universiteit Amsterdam, The Netherlands 2Laboratoire de Photophysique Mol&culaire, Universitg de Paris-Sud, Orsay, France 3Synchrotron Soleil, Gif-sur-Yvette, France 41nstituut Anton Pannekoek, Universiteit van Amsterdam, The Netherlands 'Institute of Astronomy, Cambridge University, UK Recently indication for a possible variation of the proton-to-electron mass ratio p=mp/me was found from a comparison between laboratory H2 spectroscopic data and the same lines in quasar spectra. This result will be put in perspective of other spectroscopic activities aiming at detection of variation of fundamental constants, on a cosmological as well as on a laboratory time scale. Furthermore the opportunities for obtaining improved laboratory wavelength positions of the relevant H2 absorption lines, as well as the prospects for obtaining a larger data set of HZabsorptions at high redshift will be presented. Also an experiment to detect Ap on a laboratory time scale will be discussed.
1. Introduction Recently the finding of an indication for a decrease of the proton-to-electron mass ratio p=mp/m by 0.002% in the past 12 billion years was reported [l]. Laser spectroscopy on molecular hydrogen, using a narrow-band and tunable extreme ultraviolet laser system resulted in transition wavelengths of spectral lines in the B-X Lyman and C-X Werner band systems at an accuracy of 5 x for the best lines. This corresponds to an absolute accuracy of 0.000005 nm. A database of 233 accurately calibrated H2 lines is produced for future reference and comparison with astronomical observations. Recent observations of the same spectroscopic features in cold hydrogen clouds at redshifts z=2.5947325 and z=3.0248970 in the line of sight of two quasar light sources (Q 0405-443 and Q 0347-383) resulted in 76 reliably determined transition wavelengths of H2 lines at accuracies in the range 2 x to Those observations were performed with the Ultraviolet and Visible Echelle Spectrograph at the Very Large Telescope of the European Southern Observatory at Paranal, Chile [2]. A third ingredient in the analysis is the calculation of an improved set of sensitivity coefficients Ki, a parameter 103
104
associated with each spectral line, representing the dependence of the transition wavelength on a possible variation of the proton-to-electron mass ratio. Details of the methods are reported in Ref. [3]. A statistical analysis of the data yields an indication for a variation of the for a weighted fit proton-to-electron mass ratio of Adp = (2.45 f 0.59) x for an unweighted fit. This result has a and A d p = (1.98 5 0.58) x statistical significance of 3.50. The redshifts of the hydrogen absorbing clouds can be converted into look-back times of 11.7 and 12 billion years with the age of the universe set to 13.7 billion years. Mass-variations as discussed relate to inertial or kinematic masses, rather than gravitational masses. The observed decrease in p corresponds to a rate of change of per year, if a linear variation with time is assumed. This dlngdt = -2 x remarkable result should be considered as no more than an indication for a possible variation of p. Only a very limited data set is available: two quasar systems with a total of 76 spectral lines.
In the following we put these results in perspective of other spectroscopic activities concerning variation of fundamental constants, and present possibilities to obtain confirmation of the findings in the near future by producing improved laboratory data for H2 and extend the data set of H2 astronomical observations.
2. Variation of dimensionlessfundamental constants: a and p Renewed interest in the possibility of temporal variation of fundamental constants arose through the findings of Webb et al. [4]. Based on highly accurate spectroscopic observations of atomic and ionic resonance lines at high redshift (from the HIRES-Keck telescope at Hawaii) a variation of the fine structure constant a was deduced. This breakthrough could be made through implementation of the so-called Many-Multiplet method, which allows for using many transition wavelengths in the analysis [5], rather than just separations between fine structure lines, as in the alkali-doublet method. By this means the sensitivity to detect A a is improved. These findings on a lower value of a in the past were disputed by competing teams who found essentially a null result on A a from data obtained with the UVES-VLT on the southern hemisphere [6,7]. Meanwhile the Webb-Murphy-Flambaum team extended their data set to some 150 quasar systems, obtaining a more than 5 0 effect with Aala = (-0.574 f 0.102) x [8]. The discrepancy in the findings by different teams on Aala were resolved by the recent reanalysis of the UVES-VLT data set by Murphy et al. [9]; flaws in the fitting procedures were uncovered and a reanalysis yields a revised value of A d a = (-0.44 f in agreement with the values of [8]. 0.16) x
105
The invention of frequency comb lasers has immensely increased the accuracies in atomic spectroscopy, to the extent that absolute precision at the can be obtained, in fact limited by the Cs time and frequency level of standard. From atomic precision experiments on various systems boundaries to the rate of change in the fine structure constant dln aldt were set to the per year by performing laboratory laser spectroscopic level of 2 x studies with time intervals of one or a few years. The NET-Boulder group set a limit of 1.2 x lo-'' per year from measurements on a singly trapped '99Hg+ ion [lo]. The Munich group deduced a similarly small rate from calibrating the H-atom (1s-2s) transition against the Paris portable Cs fountain clock [ l l ] , as did the PTB-Braunschweig team from 171Yb+ions [12]. Very recently the NIST-Boulder group pushed the boundary on dln p/dt to 1.3 x 10-16per year by comparing Hg' against Cs [ 131; at the ICOLS 07 even a tighter limit was presented at the 2 x level from a comparison of Hg' and Al' clock transitions (see this book). It has been hypothesized that the changes in the proton-electron mass ratio p and the fine structure constant a are linked and that p would change faster by an order of magnitude or more; this hypothesis [14] is based on Grand Unification Theories. From a recent analysis of microwave spectra from the astrophysical object B0218+357 at redshift z = 0.68 Flambaum and Kozlov [15] put a limit to the variation of the mass-ratio at ANp = (0.6 5 1.9) x Data on the inversion motion of ammonia (23 GHz) were compared to microwave transitions in other molecules. Hence there is evidence for a variation of a , and some indication for a variation of p at high redshifts (z > l), while the laboratory studies seem to put strict boundaries on Aa. At the same time the recent findings of high redshift ammonia put a strict boundary to Ap at a redshift of z = 0.68. In this context the hypothesis of a phase transition occurring in the history of the universe, going from a matter-dominated (dust era) to a dark energy dominated (curvature era) universe may play a role [16]. Barrow hypothesized that only before this transition, which may have occurred near z=0.5, the fundamental constants may have changed.
3. Extension of the database of molecular hydrogen at high redshift There exists only a limited data set of H2 absorptions at high redshift. Of the tens of thousands identified quasar sources some 600 are known to be associated with a damped Lyman-a system; such systems are characterized by a fully saturated and broad L-a absorption feature from a relatively dense cloud of atomic hydrogen with a column density of N(H) > 2 x 10'' ern-'. Such systems display metal absorptions and in some cases also H2
106
absorptions. For the investigations probing A a some 200 systems have been spectroscopically analyzed (from metal lines Mg, Si, Zn, etc), but in only 14 of them H2 has been detected [3]. Lyman-a of quasar emission
I
3795
3500
'I
3800
I 4000
I
I
4500
I
i
5ooo
I
I
5500
6ooo
1
I 6500
Fig. 1. Typical spectrum of a damped Lyman-a quasar system, in this case Q2348-011, as recorded by Noterdaeme et al. with the Ultraviolet-Visible Echelle Spectrometer at the Very Large Telescope at Paranal, Chile [18]. The large emission peaks can be assigned to H-La and to C IV, and give the redshift of the quasar emission (z=3.0236). Spectra are recorded with a certain setting of a dichroic, detecting blue light on a single CCD, and the red part on two distinct CCDs. A damped L a absorption line is found at z = 2.426 and also H2 absorption is found at this redshift. Hence the Lyman and Werner absorption lines (in the laboratory at 90-112 nm) are shifted to below 380 nm. Part of the H2 window is enlarged in the left upper corner, displaying the complicated velocity structure of the H2 cloud: at least 7 velocity sub-components are visible for each absorption line (LOR0 is shown).
Obtained spectra in existing databases have been surveyed and besides the two systems used in our previous analysis (Q0347 and Q0405) three others have a potential to play a role in detecting Ap if spectra at sufficient SNR and resolution, with optimum wavelength calibrations were to be obtained. The system Q2348-011 at z= 2.426 (an archived spectrum shown in Fig. 1) will be observed under such conditions at UVES-VLT in August 2007. Other appropriate systems would be Q0528-250 at z=2.81 and Q1443+272 at z=4.22; the latter is the system with H2 detected at the highest redshift [17]. Of course there should be many damped L-a systems with H2 at high redshift
107
in the universe that can be implemented in Ap analyses. They 'just' need to be found and subjected to high resolution observation; as a figure of merit, at current dish sizes of 8 m typical observation times in access of 20 h on target (depending on brightness of the quasar background source) are needed to obtain spectra at resolutions of R = 60000 and SNR of 50. In view of the importance of the subject the data set will be extended in coming years; currently a number of observation stations, HIRES-Keck (Hawaii), UVESVLT (Chile), and HDS-Subaru (Japan), are suitable for the purpose. In 2009 the PEPSI-LBT system in Arizona, equipped with two 8 m dishes, will become available for detection of H2 at high redshift.
4. Improving the laboratory accuracy of the Lyman and Werner lines The prominent electronic absorption systems, also detected in quasars, are the B'C,+- X'C; or Lyman and the C'l3,- X'C; Werner band systems. At zero redshift these lie in the difficult to access wavelength range of 90-112 nm. With the use of a narrowband and tunable extreme ultraviolet (XUV) laser source the lines could be calibrated to an accuracy of 5 x [ 191.
0 7840
0.7860
0.7880
0.7900
Wavenumber - 991 64.0 ( 6 ' )
Fig. 2. Recording of the Qo two-photon line in the EF-X (0,O) band of H2 at 99164.78691(11) cm-I. Note that the resonance width is 36 MHz, determined by the linewidth of the laser at its 8'h harmonic.
We have devised an alternative spectroscopic scheme to derive the wavelengths in the B-X and C-X systems via combination differences. This method is based on two independent spectroscopic measurements. First the lowest energy levels in the EF'C;, v=O state are determined via a Dopplerfree two-photon-excitation scheme in the deep-UV at h=202 nm, that was previously described [20]. Using various advanced techniques, such as calibration against a frequency comb laser, a Sagnac configuration to avoid Doppler shifts, and on-line frequency chirp evaluation for each of the laser
108
pulses an absolute accuracy of 3.5 MHz on the resonances (see Fig. 2 for a spectrum of the Q0 line) was obtained, which translates to a relative accuracy of A A A = l x 10"9.
6778 8780
8782
8784 67S8 8788 S7W W92 cm
its
dull 6450
6500
..JaUk^jujylluJj
lltjiL ijLJli.lLw-uJi
6550
6600
6650
6700
6750
6800
6850
cm'
Fig. 3. A portion of the FT infrared emission spectrum of H2 in the range 64006900 cm"1 displaying lines in the EF-B (0,2) and (1,4) bands as indicated.
A second experiment entails Fourier-Transform infrared and visible emission spectrocopy performed on a low pressure electrodeless discharge in H2. In the near infrared domain ranging between 0.5-4 jam many lines in the EF-B (v',v") are observed (a portion of the spectrum is shown in Fig. 3). Although the spectral lines are Doppler-broadened (0.02 - 0.2 cm"1), the high SNR and the fact that each energy level is connected to 10 or more other quantum levels produces a consistent framework of energy levels at accuracies in the 10"3 - 10"4 cm"1 range. Level energies in the C state are determined, somewhat less accurate, through transitions in the l'ng-C, j'Ag-C, H1Sg+-C and GK'lg*C systems. Systematic effects are addressed by absolute wavelength calibration in a wide range using Ar and CO lines. This work in progress will yield relative level energies that can be combined with the level energies of the lowest EF, v=0 levels; from the combined set transition wavelengths of most of the relevant Lyman and Werner lines can be calculated at accuracies in the range AAA = 1-5 x 10"9. Bearing in mind that the uncertainties in the current quasar absorption data are at the level of AX/X = 2 x 10"7, these
109
studies provide a laboratory or zero-redshift data set for H2 which is exact for the purpose of comparison. 5. A molecular fountain for precision studies and detection of A|i Variation of the proton-electron mass ratio may be detected from comparisons on a cosmological time scale, but also on a laboratory time scale. Since the intervals are reduced to years the required spectroscopic precision has to be much higher in the latter case. In view of the fact that quantum tunneling phenomena scale exponentially with mass, the inversion splitting in ammonia is extremely sensitive for Au.. This also underlies the tight constraint to A|i from ammonia spectra at z=0.68 discussed above [15].
mijrowave cavity
ion detector
skimmer ffi
nozzle
Fig. 4. Design of the molecular fountain under construction in Amsterdam.
In order to obtain long effective measurement times in a Ramsey-type scheme a molecular fountain is under construction, in which NH3 molecules will be launched, after deceleration by the Stark technique [21].
110
Acknowledgement We thank Prof. G. Meijer (Fritz Haber Institut, Berlin, Germany) for the collaboration on the molecular fountain project.
References E. Reinhold, R. Buning, U. Hollenstein, A. Ivanchik, P. Petitjean, W. Ubachs, Phys. Rev. Lett. 96, 151101 (2006). 2. A. Ivanchik, P. Petitjean, D. Varshalovich, B. Aracil, R. Srianand, H. Chand, C. Ledoux, P. BoisseC, Astron. Astroph. 440,45 (2005). 3. W. Ubachs, R. Buning, K. S. E. Eikema, E. Reinhold, J. Mol. Spectr. 241, 155 (2007). 4. J. K. Webb, V. V. Flambaum, C. W. Churchill, M. J. Drinkwater, J. D. Barrow, Phys. Rev. Lett. 82, 884 (1999). 5. V. A. Dzuba, V. V. Flambaum, J. K. Webb, Phys. Rev. Lett. 82, 888 (1999). 6. R. Srianand, H. Chand, P. Petitjean, B. Aracil, Phys. Rev. Lett. 92, 121302 (2004). 7. R. Quast, D. Reimers, S. Levshakov, Astron. Astroph. 415, L7 (2004). 8. M. T. Murphy, J. K. Webb, V. V. Flambaum, Mon. Not. Roy. Astr. SOC. 345,609 (2003). 9. M.T. Murphy, J. K. Webb, V. V. Flambaum, arXiv:astro-ph/0612407vl. 10. S. Bize, et al. Phys. Rev. Lett. 90, 150802 (2003). 11. M. Fischer et al. Phys. Rev. Lett. 92,230802 (2004). 12. E. Peik, B. Lipphardt, H. Schnatz, T. Schneider, C. Tamm, S. G. Karshenboim, Phys. Rev. Lett. 93,170801 (2004). 13. T. M. Fortier et al. Phys. Rev. Lett. 98,070801 (2007). 14. X. Calmet, H. Fritsch, Eur. J. Phys. C 24, 639 (2002). 15. V. V. Flambaum, M. G. Kozlov, Phys. Rev. Lett. 98,240801 (2007). 16. J. D. Barrow, H. B. Sandvik, J. Magueijo, Phys. Rev. D. 65, 063504 (2002). 17. C. Ledoux, P. Petitjean, R. Srianand, Astroph. J. 640, L25 (2006). 18 P. Noterdaeme, P. Petitjean, R. Srianand, C. Ledoux, F. Le Petit, Astron. Astroph. 469,425 (2007). 19. J. Philip, J. P. Sprengers, P. Cacciani, C. A. De Lange, W. Ubachs, E. Reinhold, Can. J. Chem. 82, 713 (2004). 20. S. Hannemann, E.J. Salumbides, S. Witte, R. T. Zinkstok, E.-J. Van Duijn, K. S. E. Eikema, W. Ubachs, Phys. Rev. A 74,062514 (2006). 21. H.L. Bethlem, G. Berden, G. Meijer, Phys. Rev. Lett. 83, (1999). 1.
QUANTUM INFORMATION AND CONTROL 11
This page intentionally left blank
QUANTUM INTERFACE BETWEEN LIGHT AND ATOMIC ENSEMBLES Hanna Krauter, Jacob F. Sherson, Kasper Jensen, Thomas Fernholz, Jonas S. Neergaard-Nielsen, Bo Melholt Nielsen, Daniel Oblak, Patrick Windpassinger, Niels Kjzergaard, Andrew J . Hilliard, Christina Olausson, Jorg Helge Miiller, Eugene S. Polzik Quantop -Danish Research Center f o r Quantum Optics, Niels Bohr Institute, Copenhagen University, Denmark
1. Introduction
Recent years have witnessed astounding progress in the ability to control quantum systems making the vision to create working quantum networks more realistic than ever. A key component in any quantum network is certainly the interface between stationary and flying carriers of information. One avenue towards a reliable interface makes use of macroscopic atomic ensembles to distribute fragile quantum states over many particles. We review here our experiments and recent progress with atomic samples in different temperature regimes and with non-classical light sources of suitable spectral characteristics for efficient coupling to atomic ensembles. 2. Quantum interface between Cesium atoms at room
temperature and light 2.1. Canonical variables
In the first group of experiments we study a quantum interface between an ensemble of lo1' Caesium atoms at room temperature and a coherent light beam. We use dispersive atom light interaction as a versatile tool for quantum communication protocols and as a method to read out the atomic state via light. We describe the quantum interface in the language of canonical variables for light and atoms.' The atomic ensemble is characterized by the collective spin of the Cesium atoms J = j i , where j i represents the spin 113
114
of a single atom. In the experiment, the atoms are oriented along the xdirection achieved by optically pumping the atoms into the F=4, m=4 state of the 6S1/2 ground state of Cesium. Following the commutation relation for the spin components and with J, being a large classical number, the Heisenberg uncertainty principle reads V a r ( j , ) . V a r ( j , ) 2 For light we consider the polarization state, characterized by the Stokes-operators Sz, Sy and S,,with [Sy,SZ]= is,, where S, is treated classically for a strong beam with a large polarization in y-direction. In order to have a common language for the light and the atoms, we introduce canonical operators: 2 = A, a Ij = A and y = 6'ij = L, 6 where each set of operators follows the commutation relation of canonical operators. Initially atoms and light will be in a minimum uncertainty state where the variables have a Gaussian probability distribution with variance $. A magnetic field is added in direction of the macroscopic spin leading to a Larmor precession of the transverse spin components around the x axis. The relevant atomic variables will then be the spins in the rotating frame. For light the cosine and sine modes at the Larmor frequency R will be of interest: GC = :J SZ(t)c o s ( ~ t ) d t ,ijs = :J S z ( t )sin(Rt)dt, = ...
$.
a
&
&
cc
For the interaction of light and atoms, we consider a beam of light coupled off-resonantly to the 6 S 1 / 2 , ~ = 4+ 6P3/2 transition. Via the Faraday interaction the polarization of light is rotated proportional to the spin component in the propagation direction. At the same time the atomic spin is rotated due to the angular momentum of light. The atomic quadratures after an interaction of duration T become:2
-
with the coupling-parameter IF. d m . The two spin components rotate in and out of the interaction direction and are thus affected by the cosine and sine modes of S z . For the light, the equations look a little more complicated:
Here and ts,l are higher order temporal modes, which enter the equations because of the back-action of light on itself mediated by the precessing atoms. From these equations it is evident, that light and atoms leave an
115
imprint on each other — they become entangled. In a modified experimental setting two cells with oppositely oriented macroscopic spins are used; J% = —J.2. = Jx. By introducing non-local J 1 —j 2 * jl+j2 canonical operators such as: X = v,n , v and P — *,„ , * the back-action V2Jz
V2Jx
terms cancel out and the input output relations simplify to: •\rout
v\
yin
— .A.
put = .in
pout __ pi q
out
in
= q .
(4) (5)
Here one of the input quadratures of light is directly mapped on the atoms (eq. 4), while one of the atomic quadratures is read out by the light (eq. 5). 2.2. Teleportation of a quantum state of light onto atoms
Fig. 1. In (a) the teleportation setup of the experiment is shown. The first stage of the teleportation is the entanglement. For this a strong entangling pulse, seen on the left side, is sent through the atomic ensemble (eqs. 1, 2 and 3), which is kept by Bob, while the light pulse is sent to Alice. Alice also has an unknown input state created with an electrooptical modulator (EOM) and characterized by Y and Q. Then the joined measurement (also known as Bell measurement) is performed, where her entangled beam and the input beam are mixed at a beamsplitter and the two quadratures are measured via polarization homodyning at the two output ports. The outcomes of those measurements are sent to Bob, who uses RF-coils to perform feedback on his atoms, thus recreating Alice's input state. In (b) the gain was extracted by comparing the first pulse measurements to the second pulse measurements.
The concept of teleportation of continuous variables was introduced by Vaidman,3 where the canonical variables Y and Q of a quantum system (held by Alice) are transferred onto another (Bob's) system with the help of a shared Einstein-Podolsky-Rosen (EPR) entangled pair. To complete
116
the state transfer, a joint measurement on the input state and Alice’s part of the entangled pair is followed by classical communication and a local transformation on Bob’s remaining part of the EPR pair. This general procedure can be applied to teleport a quantum state of light onto atoms4 The principle of teleportation in our setup2 is illustrated in Fig.1 (a). Since the states we consider here are Gaussian, we only have to verify the mean value transfer and find the variances of the outcomes of the atomic states to characterize the quality of the teleportation. For the read-out a second pulse is sent through the atoms after the completion of all teleportation steps (for the timing see inset in Fig.l(a)). Now one can compare the mean value of the first pulse measurements, which bear the mean value of the input state, with the mean value of the second pulse, with which the atomic state was read out via light (eq. 2, 3 ) . In Fig.l(b) one can see the calibration for one of the two quadratures, for which the input was scanned. The gain = < z t e l e p o r t e d > can thus be extracted and set to one.
The remaining task is to detect the variances of the final state. Again the second pulse is used as the read out of the atomic state, but this time the input is not varied but held constant. The variance of the atomic state can be retrieved from the light measurement with help of eqs. 2 and 3 . In Fig.2(a) one can see the outcomes of a light measurement. From this the atomic state can be reconstructed as shown in 2 (b). In the case of the displayed measurement, where the displacement corresponds to a photon number of fi = 5, the fidelity is F = 0.58 f .02, which clearly lies above the classical limit of 0.5.5 For a limited range of input states the fidelity is maximized with a gain different from one. For input distributions with widths of < n >= 2,5,10,20,200 fidelities of F 2 = 0.64 f 0.02, F 5 = 0.60 f 0.02, F ~= o 0.59 f 0.02, F ~ = o 0.58 f 0.02 and I7200 = 0.56 f 0.03 can be extrapolated from our measurement^,^ which should be compared with the achievable classical fidelities5 F t l a s s i c a l - 0.60, ~ g c l a s s i c a l= 0.545, F $ a s s i c a l - 0.52, F ; m s i c a l 0.51 and F;&jssical= 0 .50 . There are different possibilities to improve the performance of this experiment. The protocol as it is suffers from residual noise introduced by the initial entangling interaction. By including higher order temporal modes and utilizing squeezed light in the entangling arm those extra noise contributions can be lessened and the fidelity increased.
117 (a)
(b) probability density . •!.
I
"'... ..' 'I ' ' J. - i.'-'t- i- '•- .
verifying pulse: ^
Fig. 2. From the light measurement (a) the atomic state can be reconstructed (b). There are two probability distributions indicated. The one without coloring is the best possible classical teleportation. It has been shown,5 that the best possible achievable classical fidelity F is 0.5. Compared to the best classical teleportation (uncolored graph) the teleported state (colored graph) is narrower.
2.3. Single atom squeezing In section 2.1 the two-cell setup was discussed briefly. From eqs. 4 and 5 the possibility of a mapping protocol of light onto atoms arises, where after the interaction one of the light quadratures is automatically written on the atoms. Then y of the light is measured and fed back to the atoms, so that: Pout = Pout + g • yout = -yin, if the gain is adjusted properly. This experiment has been conducted successfully.6 However, the fidelity of the mapping protocol is limited by the residual input noise from the other atomic quadrature (see eq. 4). This can be partly overcome by squeezing Xin. Here the multilevel structure of the Cesium atoms is utilized to reach this goal. By creating a suitable superposition of the even magnetic sublevels m = 4, 2,0,... a spin squeezed state7 can be achieved, at the expense of a decrease in the macroscopic spin Jx. As a result, one of the normalized transversal spin components x or p has a variance smaller than ^. Experimentally such a superposition state can be obtained by inducing Raman transitions with two light beams. The important features of the experiment are sketched in Fig.3(a). Figure 3(b) shows preliminary results for the experiment. The crosses are measured variances of the squeezed quadrature, obtained for 10000 measurements with equal pulse duration, power and detuning. For first tests a single cell setup was used and the Faraday interaction utilized for the read out of the two atomic quadratures. A noise reduction of (30 ± 10)% compared to the standard quantum limit is achieved. The dotted line shows the maximum achievable squeezing with
118
Fig. 3. In (a) the squeezing experiment is shown, where two beams, which are off resonantly coupled to the Dl line (A = —550 MHz), are sent through the atoms along the macroscopic atomic orientation to create a superposition of m = 4, 2,0,.... The two beams are detuned by twice the magnetic sublevels splitting, given by the Larmor frequency Ci = 322 kHz. The Faraday interaction is used to determine the noise reduction of the atomic quadratures. Figure (b) shows measurement results. The solid line indicates the noise level of a coherent spin state (CSS) with the equivalent macroscopic spin. The crosses and the circles show the measured squeezed and anti-squeezed variance, whereas the dotted line gives the calculated optimum squeezing.
the used interaction in the absence of decoherence predicting approximately 70% noise reduction. Different effects limit the performance of the experiment. The decay of the created state during the preparation and the read-out introduces constraints. Furthermore, the second order Zeeman and Stark shifts lead to the fact that the atomic spin can not be described by a Larmor precession with a single frequency. Experimentally, one needs to strike a balance between the limiting effects by choosing an optimal power and timing, as well as detuning, of the Raman beams to achieve optimal noise reduction.
3. Dispersive measurements on dipole trapped cold Cs atoms In a different experimental setup we investigate non-destructive probing of laser cooled Cs atoms confined by a focussed laser beam and restricted to populate the two ground level clock states (F — 3,mp = 0) and (F = 4, mp = 0) and superpositions thereof. As is well known such a two level system is formally equivalent to spin 1/2 system and we can form a collective pseudo-spin for the atomic ensemble in much the same way as introduced in section 2.1. The system is conveniently illustrated in the Bloch sphere picture Fig. 4.
119
fz ]F=4,mF=0>
Spin Squeezed State
1
!F=3,m,=0> Uncertainty disk prior to (2ND measurent
Fig. 4. Bloch sphere representation of collective pseudo- spin. Beginning from situation with all the atoms in one of the clock states (a pole on the sphere) and applying a Tf/2pulse produces an equal superposition of both clock states (the equatorial plane - here assumed to point in the a;-direction) with an uncertainty disk associated to it. Such a so-called coherent spin state will give rise to projection noise when measuring the z-axis spin projection. Performing a QND measurement produces a spin squeezed state.
3.1. Atom-light interaction A probe laser beam propagating through the trapped atoms will experience an index of refraction according the population of atoms in either clock state. This leads to a state dependent phase shift of light which we can measure by comparing to a reference beam in free space using a Mach Zehnder interferometer.8 Again, light in the two-path interferometer can be described using an angular momentum algebra in a similar way as for the polarization states in section 2.1. It turns out that the interaction between the collective angular momenta for light and atom leads to a quantum nondemolition (QND) measurement of an atomic pseudo spin projection when the phase shift of light is detected. The outcome of this measurement can in principle be used to infer spin squeezing, i.e. a reduction of uncertainty in one spin component (Sz) on the expense of an increased uncertainty in another component (Sy)9'w (See Fig. 4). 3.2. Rabi Oscillations As a first step towards the observation of projection noise and production of spin squeezed states on the clock transition, we have studied the coherent evolution of a spin polarized sample.11 The atoms are prepared in the (F = 3, mp = 0) state and Rabi oscillations on the clock transition are driven with a resonant 9.1 GHz microwave field. During Rabi flopping
120
the atoms are probed with short (200 ns) off resonant (150 MHz blue detuned from 6Sri/2(-F = 4) —> 6P3/2(F' = 5) transition.) laser light pulses. The probe laser beam will be shifted in phase proportionally to the number of atoms in the (F = 4, mF = 0) state. Figure 5 shows an example of non-destructive probing of Rabi oscillation using our Mach Zehnder interferometer. Using this method makes it possible to follow the coherent evolution of the ensemble quantum state in "real time". Having established that the atomic ensemble can be controlled coherently, we have the tools at hand to produce the coherent spin state Fig. 4. The challenge is now to detect the atomic quantum fluctuations in the recorded phase shift for independently prepared ensembles and that this variance can be reduced by using information gained in a previous QND measurement.
(b)
A A A A A A A A * A
=yvy\/vvv^vv '"%
MO
~4C8
' CM ~
flea
ioaa
X50\
Fig. 5. Rabi Oscillations measured as a state dependent phase shift using our interferometer, (a) An average over 50 experimental runs, (b) Single experimental run.
4. Non-classical states of light 4.1. Gaussian states With the purpose of being a resource for the experiments presented in sections 2 and 3, we have a setup for generating various non-classical states of light. The heart of the experiment is an optical parametric oscillator (OPO) pumped below threshold. Employing a nonlinear PPKTP crystal (periodically poled potassium titanyl phosphate), the blue pump beam at frequency 2o>o is down-converted into several longitudinal cavity modes centered around the frequencies ujk = w0 +
k = . . . , -2, -1, 0, +1, +2, . . . ,
121
Fig. 6. Results of homodyne measurement and reconstruction of the two different nonGaussian states; Wigner function and density matrix in number state representation. Left: Photon subtracted squeezed vacuum (kitten state). Note the predominance of odd photon numbers in the density matrix diagonal. Right: Single photon state.
where WA is the free spectral range of the cavity. The field emitted from the OPO in the degenerate mode U>Q is in a squeezed vacuum state with one of the field quadratures being less noisy than the vacuum level. We are currently able to produce a very pure state with -6.5 dB squeezing versus 10 dB anti-squeezing. As mentioned in section 2.2, the squeezed vacuum can be used to improve the fidelity of the memory protocol. Apart from the u>o mode, the non-degenerate longitudinal modes are pairwise correlated such that e.g. cj_i and w+i are in a two-mode squeezed state. In the OPO the two modes are produced in the same spatial and polarization modes, but since they have different frequency they can be separated via a cavity resonant on w_ 12 4.2. Non-Gaussian states The single mode and two-mode squeezed states are indeed non-classical states, but they are still Gaussian. As demonstrated by Wenger et al,13 it is possible to de-gaussify the states by conditioning on detection of a photon in a part of the field. In the case of the squeezed vacuum, we reflect on a beam splitter a small fraction towards a single photon counter. When this photon counter clicks, we have effectively subtracted a photon from the remaining transmitted field, turning it into a superposition of odd-photon number states (the initial squeezed vacuum is ideally an even-photon number superposition), sometimes referred to as a 'Schrodinger kitten'.14 If we instead focus on the two-mode squeezed state, detection of a photon in the w-\ mode will de-gaussify the correlated u}+\ mode. If the pump intensity is sufficiently low, the generated state in w+i will be a single pho-
122
detector Fig. 7. Experimental configuration for the detection of backscattered light from a trapped BEC. The inset illustrates momentum detection after time of flight.
ton state.15 This scheme for a single photon source is similar to the standard down-conversion scheme of triggering on one photon from a down-converted pair, except that they are usually performed by single pass pulsed pumping. In that case the bandwidth of the state will be several GHz or nm wide. Due to the cavity enhancement by the OPO, we can reduce the bandwidth to the order of 10 Mhz while still keeping a high production rate of ~10,000 s"1. This property - which is shared by the kitten states - is very important for future perspectives, where storage of such non-Gaussian states in atomic memories is a possibility. The wavelength of the source is frequency tunable around the 852 nm Cs line. In order to characterize the non-Gaussian states, we measure them in a broadband homodyne detection setup. The conditional states appear in a temporal mode centered around the trigger time and with a shape roughly determined by a double-sided exponential decay with the decay constant of the OPO. Hence it is necessary to temporally filter the continuous homodyne detection signal to properly measure the non-Gaussian state and not the background squeezed state. After acquiring several thousand quadrature points at different phase angles we can reconstruct the density matrix and Wigner function of the state. The results, presented in Fig. 6, shows clearly non-Gaussian Wigner functions which even have deep negative regions around the origin. The purity of the states are between 60% and 70%. 5. Atom-Light interface with quantum degenerate atoms
The dispersive coupling between Cs atoms and light discussed in the previous sections can, of course, also be applied to different atomic species. The coupling strength K introduced in Sec.2 between the collective variables of atoms and light can be conveniently expressed as K2 = ao?7, the product of on-resonance optical depth ao and time integrated spontaneous emission rate r\. For a cold atomic sample optical depth is monotonous in phase
123
Fig. 8. Left: atomic momentum distribution measured after 45 ms of free expansion with the depleted original condensate to the right and the recoiling atoms to the left surrounded by an isotropic halo, populated by light scattering and by s-wave collisions during expansion; Right: Background corrected photodetector signal showing the time dependent reflectivity of the atomic sample;
space density, so that quantum degenerate bosonic samples (BEG) offer the ultimate coupling strength for a given amount of dissipation by spontaneous emission. Furthermore, the absence of Doppler broadening allows to resolve the excited state hyperfine structure in the optical excitation, which gives rise to a different effective interaction between atomic internal states and light polarization states with more complex input/output relations.16 For atomic samples far below the recoil temperature the momentum degree of freedom becomes accessible as an additional quantum system with a well denned initial state, where the coupling between light and atomic momentum states can be formulated in a language analogous to the polarization/angular momentum case. In first experiments we try to assess the achievable coupling strength by studying super-radiant Rayleigh scattering off EEC's.17 Experimentally we prepare EEC's of 87Rb atoms by standard rf evaporation techniques inside a magnetic trap of the loffe-Pritchard type.18 The cigar shaped clouds, containing ~ 5 • 105 atoms with no discernible thermal fraction, are probed along the long axis with IfjW pulses of circularly polarized light, focused to a waist of 20fj,m and detuned by —2.6GHz from the 5si/2(F = 1) —> 5pi/2(F' = 2) transition (see Fig.7). The elongated geometry of the sample favors repeated scattering and super-radiant gain in the direction of the long axis. We detect light scattered in the backward direction and measure the atomic momentum distribution by shadow imaging after time-of-flight expansion (see Fig.8) clearly demonstrating that the
124
super-radiant regime is reached. Ongoing experiments aim to verify the strict correlations between scattered photons a n d recoiling atoms implied by momentum conservation.
6. Acknowledgements
This research is funded by DG and through the EU projects COVAQUIAL,
QAP, a n d EMALI. References 1. J. F. Sherson, B. Julsgaard and E. S. Polzik, Deterministic atom-light quantum interface, in Adv. At. Mol. Opt. Phys. 54, eds. P. Berman, C. Lin and E. Arimondo (Elsevier, 2006) pp. 82-131. 2. K. Hammerer, E. S. Polzik and J . I. Cirac, Phys. Rev. A 72, 064301 (2005). 3. L. Vaidman, Phys. Rev. A 49, 1473 (1994). 4. J. F. Sherson, H. Krauter, R. K. Olsson, B. Julsgaard, K. Hammerer, I. Cirac and E. S. Polzik, Nature 443, 557 (2006). 5. K. Hammerer, M. M. Wolf, E. S. Polzik and J. I. Cirac, Phys. Rev. Lett. 94, 150503 (2005). 6. B. Julsgaard, J. Sherson, J. I. Cirac, J. Fiurasek and E. S. Polzik, Nature 432, 482 (2004). 7. M. Kitagawa and M. Ueda, Phys. Rev. A 47, 5138 (1993). 8. P. G. Petrov, D. Oblak, C. L. Garrido Alzar, N. Kjaergaard and E. S. Polzik, Phys. Rev. A 75,033803 (2007). 9. A. Kuzmich, N. P. Bigelow and L. Mandel, Europhys. Lett. 42, 481 (1998). 10. D. Oblak, P. G. Petrov, C. L. Garrido Alzar, W. Tittel, A. K. Vershovski, J. K. Mikkelsen, J. L. Sorensen and E. S. Polzik, Phys. Rev. A 71,043807 (2005). 11. P. Windpassinger et al., in prep. (2007). 12. C. Schori, J. L. S~rensenand E. S. Polzik, Phys. Rev. A 66, 033802 (2002). 13. J. Wenger, R. Tualle-Brouri and P. Grangier, Phys. Rev. Lett. 92, 153601 (2004). 14. J. S. Neergaard-Nielsen, B. M. Nielsen, C. Hettich, K. Mdmer and E. S. Polzik, Phys. Rev. Lett. 97, 083604 (2006). 15. J. S. Neergaard-Nielsen, B. M. Nielsen, H. Takahashi, A. I. Vistnes and E. S. Polzik, Opt. Express 15,7940 (2007). 16. 0. S. Mishina, D. V. Kupriyanov, J. H. Miiller and E. S. Polzik, Phys. Rev. A 75,042326 (2007). 17. S. Inouye, A. P. Chikkatur, D. M. Stamper-Kurn, J. Stenger, D. E. Pritchard and W. Ketterle, Science 285,571 (1999). 18. T. Esslinger, I. Bloch and T. W. Hansch, Phys. Rev. A 58,R2664 (1998).
DEGENERATE FERMIGASES
This page intentionally left blank
AN ATOMIC FERMI GAS NEAR A P-WAVE FESHBACH RESONANCE D . S. JIN, J. P. GAEBLER, AND J. T. STEWART JILA, Quantum Physics Division, NIST and Department of Physics, University of Colorado, Boulder, 440 UCB Boulder, CO 80309-0440, USA
Abstract: Atomic scattering resonances, called Feshbach resonances, have been used to create molecular Bose-Einstein condensates and Fermi superfluids. Past work has focused on s-wave, or non-rotating, pairs created from two fermionic atoms. Here we report on investigations of pair creation in an ultracold Fermi gas of 40Katoms near a p-wave Feshbach resonance.
1. A p-wave Feshbach Resonance 1.1. Introduction and Motivation Ultracold gases of atoms are powerful model systems for exploring many-body quantum phenomena. A unique feature of these systems is that the experimenter can actually control the interactions between the particles through the magic of a Feshbach resonance. By going to the strongly interacting regime in a Fermi gas of atoms, it is now possible to create a Fermi superfluid state [l]. This state results from the pairing of atoms; a pair of correlated fermions is itself a composite Bose particle, which can form a Bose condensate and thus give rise to superfluidity. This basic phenomenon can be seen in many other Fermi systems including superconductors, superfluid liquid 3He, and nuclear matter. The simplest type of pairing is s-wave pairing. In this case, the pairing is isotropic in space and does not involve orbital angular momentum. The Fermi superfluid state realized in ultracold gases, using either 40K atoms or 6Li atoms, is an s-wave superfluid. For ultracold gases, unlike for dense Fermi systems, we understand extremely well the microscopic origin of the interactions. Indeed, we control those interactions in order to form the Fermi superfluid state. If, as in the case of current experiments, the interactions are controlled using an swave Feshbach resonance: then the resulting pairs are obviously s-wave. Now, let us consider the possibility of using a non-s-wave Feshbach resonance to create non-s-wave pairs. We know that non-s-wave pairing occurs in high T, superconductors (d-wave) and in superfluid liquid 3He (p-wave). These condensed matter systems have some unique properties because of the non-s-wave pairing. Non-s-wave pairing is anisotropic and can give rise to an a
An s-wave Feshbach resonance, as discussed here, couples an s-wave scattering state of two atoms to an s-wave (non-rotating) molecule. The p-wave Feshbach resonance discussed in this paper couples a p-wave scattering state of two atoms to a p-wave (rotating) molecule.
127
128
anisotropic pairing gap. In addition, there are now different possible quantum numbers describing the pairs; these correspond to the different projections of the orbital angular momentum of the pair. For example, in the case of p-wave pairing, with one quanta of orbital angular momentum (L=l), the projection, mL, can be - 1 , 0, or 1. This allows for multiple superfluid states, and the opens the possibility of quantum phase transitions between distinct superfluid states. In addition to providing access to these intriguing features in a uniquely controllable model system, p-wave superfluidity in an ultracold Fermi gas has also been discussed as an interesting system for topological quantum computing.
1.2. Not One, But Two Resonances In ultracold Fermi gases, magnetic-field tunable p-wave Feshbach resonances have been observed for both 40Katoms and 6Li atoms. In 2003 our group reported the observation of a p-wave resonance between spin-polarized 40K atoms [3]. As a first step toward pursuing the possibility of using this resonance to create correlated fermion pairs and p-wave superfluidity, we discuss here experiments studying weakly bound p-wave molecules created using this Feshbach resonance. More details about these experiments can be found in Ref. [3].
iij E
8x1O5
1
6x1O5
i
1
fi 4 ~ 1 -0 ~
3 2X1O5 a1
I
198
199
200
B (Gauss) Figure 1. Loss in atom number observed near a p-wave Feshbach resonance between atoms. spin-polarized 40K
A Feshbach resonance arises from the coupling of two-particle scattering states to a bound state. And, therefore, it turns out that one can use a variety of
experimental techniques near a magnetic-field tunable Feshbach resonance to very efficiently convert atoms painvise into weakly bound molecules. Perhaps the simplest technique is simply to set the magnetic-field strength to a value near the resonance position and wait. Figure 1 shows the loss in the number of atoms that we observe for such an experiment near the 40Kp-wave Feshbach resonance. A striking feature in this data is the splitting of the p-wave resonance into two distinct resonances. This was first observed in Ref. [2] and explained in Ref. [4]. The two resonances correspond to p-wave scattering with different projections (onto the direction of the magnetic field) of the pair orbital angular momentum mL. The lower field resonance corresponds to mL = +1 or mL= -1, while the higher field resonance corresponds to mL=O. These two resonances are separated by about 0.5 Gauss; this splitting is due to a small energy difference that comes from the magnetic dipole interaction between atoms. The loss in the observed number of atoms seen in Figure 1 is consistent with the conversion of atoms into weakly bound Feshbach molecules. Since we measure the atom number using resonant absorption imaging of the gas, and since molecules, in general, do not absorb the same color light as free atoms, we expect that the creation Feshbach molecules would appear as a loss of atoms in our measurment. More direct evidence of p-wave Feshbach molecule creation is presented later in section 2.2. 2. p-wave Molecules
2.1. Molecule Energy
Another technique that has been used to create s-wave Feshbach molecules is to set the magnetic-field strength to a value near the resonance position and then add a small-amplitude sinusoidal modulation to the magnetic field. This oscillating magnetic field can resonantly couple free atoms pairs to the Feshbach molecule state[5][6]. By varying the frequency of the magnetic-field modulation and looking for the resonant loss of atoms, we can determine the resonant frequency for a particular magnetic-field detuning from the Feshbach resonance. Then, repeating the measurement for a variety of magnetic-field values, we mapped out the energy difference between free atom pairs and Feshbach molecules. This is shown in Figure 2. Compared with similar plots of molecule energies for an s-wave resonance, Figure 2 illustrates several new features of the p-wave resonance.
130 The p-wave resonance is split into two distinct resonances, as discussed in the previous section. Here, we directly measure the energy splitting of the two p-wave Feshbach molecule states. 2. For magnetic fields above the Feshbach resonance, there exists a metastable state with a well-defined energy. Such a state is not seen for the s-wave resonances that have been used for Fermi superfluidity. 3. The Feshbach molecule has a binding energy that depends linearly on the magnetic-field. In contrast, for the s-wave resonances that have been used for Fermi superfluidity, the s-wave Feshbach molecule energy depends quadratically on the magnetic-field detuning from resonance.
1.
-1.0
-0.5
0
0.5
1.0
1.5
Figure 2. Energy of the Feshbach molecule relative to free atoms vs. magnetic-field detuning from the (lower) Feshbach resonance. This plot is taken from Ref. [3]. A negative energy corresponds to a true bound-state while a positive energy corresponds to a metastable “quasi-hound” resonance state.
2.2. “Seeing”p-wave Molecules The existence of the quasi-bound state for fields above the Feshbach resonance provides a way to more directly probe the p-wave Feshbach molecules. After creating molecules (by holding the magnetic-field for a few ms near the resonance) we can increase the magnetic-field to a value above the resonance. The molecule state then adiabatically becomes a quasi-bound state with an energy greater than that of free atoms. This quasi-bound state is metastable
131
because of the p-wave centrifugal barrier. The height of this centrifugal barrier, in units of frequency, is about 5.8 MHz. This is much larger than typical kinetic energies in the gas (the Fermi energy is typically about 10 kHz) and also much larger than the quasi-bound state energies explored here (see Figure 2). However, the quasi-bound state will eventually dissociate into free atoms by quantum mechanical tunneling through the centrifugal barrier. The resulting free atoms will have a relative kinetic energy that is defined by the quasi-bound state energy.
Figure 3. Images of dissociated p-wave Feshbach molecules. A linear grayscale indicates the optical depth, with white corresponding to more absorption of the resonant probe light. The images are taken after ballistic expansion from the trap and therefore show the velocity distribution of the atoms resulting from dissociation of the Feshbach molecules. The left image corresponds to the mL = ±1 resonance and the right image corresponds to the mL = 0 resonance. The image plane is transverse to the quantization axis, which is defined by the external magnetic field. The images show the expected angular distributions for p-wave pairs.
Figure 3 shows images of p-wave molecules taken using this technique. After the Feshbach molecules are created, we remove all remaining free atoms using resonant laser light. Since the molecules do not absorb this light, they remain unperturbed. We then increase the magnetic-field strength to a value above the resonance and allow the now-quasi-bound molecules to dissociate. The resulting free atom clouds are shown in Figure 3. The left and right images are taken using the mL = ±1 resonance and the mL = 0 resonance, respectively. Both images are taken after ramping the magnetic-field to the same positive detuning from the relevant resonance and for the same expansion time after
132 turning the trap off. The difference in the clouds’ sizes and shapes then reflects the different angular distribution of atoms in the p-wave pairs. 2.3. Creating p-wave Molecules
As mentioned in section 1.2. it is possible to produce p-wave molecules simply by setting the magnetic field to a value near the resonance and waiting. With our method to see the molecules we could dynamically probe this process. It should be noted that we do not currently understand this molecule creation process. Moreover, the data presented in this section represent our first investigations, for the p-wave resonance, of the region where many-body effects can be expected to play an important role in determining the behavior of the gas.
t
Time (ms) Figure 4. (a) Measured atom number as a function of time that the magnetic field is held at the mL = i 1 resonance. (b) Measured molecule number for the same hold time on resonance. The plot is taken from Ref. [3]. The inset shows the timing sequence for this experiment. The number of molecules is measured using the dissociation technique described in the text (solid line.) The number of atoms not in molecules is measured by ramping the field below the resonance (dashed line).
133 Figure 4 shows the atom and molecule populations as a function of time. In this experiment, we quickly changed the magnetic field from a value far from the resonance to a value near the resonance and held for a variable amount of time. The data was taken for the magnetic-field value where we observe the highest conversion efficiency to molecules. It can be seen that the molecule population quickly reaches its maximum value in about 1 ms, and then slowly decays on a timescale of order 10 ms. The atomic population monotonically decays, indicating the presence of an inelastic decay process.
I " " ' " " "
Fermi Enemy -013.0k H t . 9.8kHz
12000
3L
-
8000
(u
i?
z
4000 0
-50
0
50
I 0 0 150 200
Figure 5. Number of molecules created for a 1 ms hold vs. magnetic-field detuning from the resonance. The data suggest that molecule creation occurs only when the p-wave resonant state has a positive energy that is less than the maximum collision energy between atoms in the Fermi gas.
If we set the hold time at the near resonant magnetic field to be 1 ms and then vary the value of the field, we observe that molecule creation occurs only over a small range of magnetic-field values near the resonance; this can be seen in Figure 5 . The molecule creation feature is observed to be asymmetric with a width that scales with the Fermi energy of the gas. This suggests that the measured width of the resonance feature reflects the atomic kinetic energy distribution rather than an intrinsic energy width of the resonance. In other words, the energy width of the resonance is narrower than the distribution of kinetic energies in the gas. The s-wave resonances that have been used to
134
realize super fluidity in Fermi gases have been in the opposite, broad resonance limit where the resonance simultaneously affects all collision energies in the gas.
2.4. Molecule Lifetimes Using the technique to see the p-wave molecules we could measure their lifetimes. Note that our imaging technique is state selective and we only detect the p-wave Feshbach molecules (and not other molecule states or free atoms). We could also measure the quasi-bound molecule lifetimes. The result of these measurements is shown in Figure 6.
-200
-100
0
I0
Energy (kHz) Figure 6 Lifetimes of the p-wave molecules as a function of their energy This plot is taken from Ref [3] Negative energies (left) correspond to bound molecules, while positive energies (nght) correspond to quasi-bound pairs Data for the mL = 0 resonance are shown in open symbols, while mL= *1 are closed symbols The two dotted lines on the left indicate the averages of the measured bound state lifetimes The solid line on the nght is a theory curve for the quasi-bound lifetimes
For the bound molecules (negative energy relative to the free atoms), we find that the lifetime does not depend strongly on the binding energy (and therefore does not depend strongly on the magnetic-field detuning from
135 resonance). This is very different from the case of the s-wave resonances that have been used for Fermi superfluidity. For those s-wave molecules, the lifetime has been seen to change by orders of magnitude over a similar range of binding energies. The dotted lines in Figure 6 indicate the average measured lifetimes for bound molecules created at each resonance. On the right side of Figure 6 we show the measured lifetimes for the quasibound p-wave molecules. Here, we see a very strong dependence on the pair energy. The solid line in Figure 6 shows a zero-free-parameter theory prediction by John Bohn [ 3 ] for the lifetime due to quantum mechanical tunneling through the p-wave centrifugal barrier. The expected (and measured) dependence of the lifetime on energy is a power law with a power of -312. The good agreement between the experiment and the theory show that the dominant decay mechanism for the quasi-bound p-wave pairs is dissociation into free atoms by tunneling through the centrifugal barrier. It should be noted that this is not an inelastic process and so does not cause any heating of the gas. 2.5. Future Prospects
One motivation for exploring a p-wave Feshbach resonance in an ultracold Fermi gas is the possibility of creating a p-wave superfluid state and exploring the many-body behavior of this new state. The measured lifetimes of the pwave Feshbach molecules tell us something about how difficult this will be to achieve. For bound p-wave molecules, our measured lifetimes of 1 or 2 ms are short compared to thermalization times for the trapped gas. This short lifetime is a serious problem for future prospects for creating Bose-Einstein condensates of these p-wave molecules. There are two different decay mechanisms for our pwave molecules. One is collisional vibrational quenching, where a molecule collides with a free atom or with another molecule. This inelastic collision produces a more tightly bound, and therefore lower energy, molecule state and releases a relatively large amount of energy. The second decay mechanism is not due to collisions, but rather can occur for single molecules in isolation. This decay is due to dipolar spin relaxation, where the molecule decays into a pair of free atoms whose hyperfine spin states have lower energy than the original internal states of the two atoms paired in the molecule. This process only exists when the original atom states are not the lowest energy states in a magnetic field. This is the case for our 40K p-wave resonance. It is possible to consider creating p-wave molecules using a different pwave resonance where dipolar spin relaxation of the molecules would not be
136
time (ms) Figure 7. P-wave Feshbach molecule lifetime after removal of free atoms. The molecules were created at the mL=Oresonance. The solid line is a fit to an exponential decay, which gives a lifetime of 7 1 ms.
*
possible. While such a resonance does not exist for 40K atoms, it does for fermionic 6Li atoms. However, collisional vibrational quenching of the molecules would remain as a possible inelastic decay channel. To see if collisional decay of the p-wave molecules was important we measured the molecule lifetime after removing all remaining free atoms. (The conversion efficiency from atoms to molecules was about 20%, so the remaining free atoms would be the most likely collision partner for the molecules.) Free atoms were removed using a pulse of resonant light to heat them out of the trap. The molecule lifetime was then measured as described previously. The results of this measurement are shown in Figure 7. The measured molecule lifetime without atoms present was 7 + 1 ms. This is significantly longer than the previously measured molecule lifetime, but still relatively short. This measured lifetime without atoms present is consistent with a prediction by John Bohn for the molecule lifetime due to dipolar relaxation [3]. This result suggests that atom-molecule collisions do play an important role in limiting the lifetime of the p-wave Feshbach molecules. For future work it will be important to understand if it is possible to create non-s-wave molecules with better stability against collisional decay. Arguably, the most interesting many-body physics will occur on the quasibound molecule side of the resonance. Here, there lifetimes shown on the right side of Figure 6 are even shorter, but this is not necessarily a bad thing. The
137 measured lifetime of the quasi-bound molecules comes from dissociation into free atoms by tunneling through the centrifugal barrier. This is an elastic process and, in fact, this tunneling is the process that can establish the pairing required for Fermi superfluidity. Very close to the resonance, where one would like to explore the many-body physics of the Fermi gas, inelastic collisional decay may become important. Exploring this behavior very close to resonance will be important in assessing the possibility for achieving p-wave Fermi superfluidity.
2.6. Conclusion Using a Fermi gas of 40K atoms, we have been able to create and detect pwave Feshbach molecules. We have explored several novel aspects of this resonance, such as the existence of a quasi-bound state that dissociates by tunneling through the p-wave centrifugal barrier. Measurements of the molecules lifetime suggest that inelastic collisional decay presents a serious challenge for future experiments attempting to investigate equilibrium manybody physics with this system.
Acknowledgments
This work was supported by the National Science Foundation and by NASA. References
1. C. A. Regal and D. S. Jin, Adv. Atom. Mol. Opt. Phys. 54, 1 (2007). 2. C. A. Regal, C. Ticknor, J. L. Bohn, and D. S. Jin, Phys. Rev. Lett. 90, 053201 (2003). 3. J. P. Gaebler, J. T. Stewart, J. L. Bohn, and D. S. Jin, Phys. Rev. Lett. 98, 200403 (2007). 4. C. Ticknor, C. A. Regal, D. S. Jin, and J. L. Bohn, Phys. Rev. A 69,042712 (2004). 5. M. Greiner, C. A. Regal, and D. S. Jin, Phys. Rev. Lett. 94,070403 (2005). 6. S. T. Thompson, E. Hodby, and C. E. Wieman, Phys. Rev. Lett. 95, 190404 (2005).
BRAGG SCATTERING OF CORRELATED ATOMS FROM A DEGENERATE FERMI GAS R. J. BALLAGH, K. J. CHALLIS, and C. W. GARDINER Jack Dodd centre for Photonics and Ultra-Cold atoms Physics Department, University of Otago, Dunedin, New Zealand *E-mail: [email protected] www.physics. otago. ac. nz/research/jackdodd We formulate a treatment of Bragg scattering of a Fermi gas in a BCS state, based on the time-dependent Bogoliubov de Gennes equations. We solve these equations in three dimensions to present a quantitative analysis of the scattering. We find that, in addition to the expected scattering of atoms by the Bragg momentum transfer tzq, a comparable fraction of atoms is scattered into a spherical shell in momentum space, centered at &/2. The atoms in the scattered shell are pair correlated, and we present an analytic model that provides an interpretation of the correlated scattering mechanism, and explains the key parameter dependencies.
1. Introduction Superfluidity in fermion systems arises due to momentum correlations between pairs of particles (the phenomenon of Cooper pairing). Techniques for probing these pair correlations in ultra-cold atomic gases have concentrated primarily on observations of the energy gap associated with the pairing (e.g., Ref. 1).Bragg scattering has been suggested in a number of theoretical studies (e.g see Refs. 2-8) as a means for characterising Fermi gases. In this paper we show that Bragg scattering can be used t o coherently probe and manipulate a degenerate Fermi gas, and that it reveals a unique signature of the pairing correlation. Bragg scattering, with a light potential of the form Acos(q . r - w t ) , has proven to be a versatile tool for manipulating and characterising BoseEinstein condensates. A long Bragg pulse is used in the so-called momentum spectroscopyg , where a narrow momentum range of the condensate is selected and incremented in momentum by an amount hq. Short Bragg pulses 138
139 on the other hand can coherently divide the condensate spatial wavefunction, giving one of the packets a momentum kick of fiq. This has enabled the implementation of atomic beam-splitters and mirrors, and the exploitation of the large scale spatial coherence of condensates in atom interferometry" . The correlations in a Fermi gas are of a different nature: they are due to attractive collisional interactions, and reside primarily on the Fermi surface. Observations of the associated energy gap using single photon processes destroy the correlation. However, Bragg scattering, which is a coherent two-photon process, has the potential to manipulate the correlation into a different form which, as we show in this paper, may be directly observed. The regime we will consider is for weak collisional interactions, for which the initial equilibrium state and its pair correlations have been successfully described by the BCS model" . This provides an appropriate starting point for our treatment, but we find the Bragg scattering is sensitive to the value of the collisional interaction at the Fermi surface, and a better treatment of the collisional interaction is required than for the conventional BCS theory, as we outline in section 2. The dynamical behaviour of the atom field under the influence the Bragg potential is calculated using the formalism of time-dependent Bogoliubov de Gennes equations. We sketch the derivation of these equations in section 2.2, and present numerical solutions for the case of a homogeneous three dimensional gas. As expected, the Bragg potential induces scattering of some of the atoms by momentum fiq. However, the key result of our calculations is that a new phenomenon of Bragg scattering atoms occurs, giving rise to a scattered spherical shell of atoms centered at momentum fiq/2. There is a threshold Bragg frequency for this shell t o form, and its radius increases with w . The atoms in the shell are correlated spin-up spin-down pairs, scattered from Cooper pairs on the Fermi surface. In the final section of the paper, we develop an analytic model that describes the underlying mechanisms for this new signature for Cooper pairing in a degenerate Fermi gas. We show in detail how it results from laser mediated transfer of the initial correlation to a different region of momentum space.
2. Time-dependent Bogoliubov de Gennes equations 2.1. Treatment of the BCS state
Our treatment is based on a BCS-like approach for the description of a degenerate Fermi gas, but with an improved treatment of the interatomic
140
collisions. In the conventional BCS approach collisions are described by a contact interaction potential, which has an infinite momentum-space range, and leads to the divergence of the pairing mean-field potential (a central object of the theory). The pair potential is typically renormalised to a finite value by introducing a momentum-space cutoff. The BCS ground state is cutoff independent, but in the limit that the cutoff tends to infinity, the renormalisation process requires that the collisional interaction strength becomes arbitrarily small. The renormalisation process described above is not sufficient for the case of Bragg scattering. We find that the correlated-pair scattering depends quantitatively on the initial pair correlations at the Fermi surface (see section 4.2), and therefore on the value of the collisional interaction potential at that momentum. We use a more sophisticated treatment of the interatomic collisions, in which the momentum-space range of the collisional interaction potential appears explicitly and is determined from experimental measurements. We assume a homogeneous Fermi gas with equal numbers of spin ‘up’ and spin ‘down’ atoms (denoted here as and -, respectively). In the absence of the Bragg field, the many-body Hamiltonian has the form
+
4cy
where is the field operator for spin state a , ‘Ho = ( - f i 2 / 2 M ) V2, M is the atomic mass, and p is the chemical potential. The interaction Hamiltonian f i c o l l n is the most general translationally invariant two-body collisional Hamiltonian for s-wave interactions between fermions of opposite spins12 ;
x
4-(Y (R
-
Here R is the centre-of-mass coordinate, r and r‘ are the relative coordinates of the two particles, and V ( r , r’) is a non-local collisional interaction potential. Our approach is to approximate V ( r , r’) by the separable potentia1l3 ,
V(r, r’) = g F ( r ) F * ( r ’ ) , (3) where F(r) has a finite range 0,even parity, and is normalised to 1. For a particular form of F(r),the interaction V(r,r’) is characterised by the interaction strength g and the range 0.
141
For a given range CT,the interaction strength g can be determined by solving the Lippman Schwinger equation for the separable potential. At low energy this gives
where T ~ B = 47rh2a/M is the two-body T-matrix, with a the s-wave scattering length. The parameter y is given by -
if(k)i2d3k
M ( 2 4 3 ~
k2
'
(5)
and depends on the range c7 through the function f ( k )= F(r)e-"k"d3r. Szymariska et al. assume a Gaussian form for the function F(r),and then determine g and c7 for a particular pair of colliding atoms, far from a Feshbach resonance. For computational convenience, we approximate the collisional interaction potential using the step function f ( k ) = Q ( k c - lkl), and then choose the wave-vector cutoff k , to match the parameters g and y with those used by Szymariska et al. By this method, we find that for 40K atoms prepared in the ( F = 9 / 2 , m ~= -9/2) and ( F = 9 / 2 , m ~= -7/2) Zeeman states, the wave-vector cutoff is k , = 0.0154/a~,h,. Therefore, for a typical Fermi momentum of k~ 0.5 x 10-3/ag0h,14 , we find that k, 30k~. In the spirit of BCS, we introduce mean-field potentials and approximate kcolln by the sum of the dominant single-particle interaction terms, which are of the form (GiGa)G',&a and ( ~ ~ ~ ~ a Assuming ) ~ a ~ that p a the BCS correlation length greatly exceeds the range of the collisional interaction potential, we obtain the Heisenberg equations of motion
-
-
The quantities W and X contain the usual Hartree and pair potentials, but with 'smearing' functions that sample the field operators over the collsional range, i.e.,
X ( r , r ' , t ) = A ( r , t ) F * ( r- r'),
(7)
.
142
with the Hartree and pair potentials given by
The prefactors in these mean-field definitions (i.e., T ~ B and g, respectively) are chosen here in order to agree with the standard renormalised BCS theory. 2.2. The Bragg formalism
We assume that the Bragg field does not change the particle spin, and describe the effect of the Bragg field on the Fermi gas by adding V B = ~ ~ ~ Acos(q.r-wt) to 7-10. The new single particle Hamiltonian 7-1 = 7 - 1 o + V ~ ~ ~ ~ ~ is used in the dynamic field equations (6). To solve those equations we introduce a time-dependent form of the well-known Bogoliubov transformation,l1>l5i.e.,
In equation (8), the 9 k e are t i m e - i n d e p e n d e n t quasi-particle annihilation operators defined to obey the standard fermion commutation relations. The dynamic evolution of the gas is described by the evolution of the amplitudes U k ( r , t ) and Z)k(r,t ) . The quasi-particle modes are populated according to the equilibrium mean-value rules (?ka?k,p) t = bkk’bapfik and (?ka?k,p) = 0. The Fermi function is f i k = l/[exp(Ek/kBT) 11, with T the temperature and the quasi-particle energies Ek of the gas in the ground state being measured relative to the chemical potential p. The equations of motion for the amplitudes U k ( r , t ) and ?&(I-, t ) are derived from equation (6) giving
+
+
1
W(r, r’, t ) U k ( r ’ ,t)d3r’
+
J
X(r,r’, t ) U k ( r ’ , t)d3r’,
and
-
J W(r, r’,
t)’Uk(r’,t)d3r’
+
J
X*(r,r’, t ) U k ( r ‘ ,t)d3r’,
in which W and X act as projectors in momentum space, with ranges 2k, and k , respectively. We solve the coupled sets of equations (9) and (10) in three dimensions, for a large set of modes k,within the regime k ~ l u l< 1, for which the BCS theory is valid.
143 The initial state is a standard BCS ground state, found from the time-independent form of equations (9) and (10) with the Bragg field off. Figure l(a) shows the momentum space column density /n(k, t)dkz, where the number density of the gas at momentum Kk is n(k, t) = C(^(k, £)> a (k, t ) } . The momentum space field operator is defined as (/>a(k, t) = /^ Q (r,i)e~ l k ' r d 3 r/L 3 / 2 , where L3 is the computational volume, and the normalisation constant C is chosen such that J n(k, t)d?k = 1. The unit of energy is chosen to be the Fermi energy EF = hujp, and the unit of wavevector is the Fermi wavevector defined in f f k p / Z M = Ep. In this paper we consider only zero temperature, in which case the Fermi energy is related to the chemical potential via EF = /z + (1 — gjT^B^U . The momentum radius of the cloud is approximately hk'F, where the effective Fermi wave vector k'F is defined in terms of the effective chemical potential by fj,' = EF - U = H2k'F2/2M (e.g., Ref. 16). Column Density
|
(a)
(b)
O -4
-4
kx
kx
Fig. 1. Column density, and pair correlation function |<7o(k)| in the kz = 0 plane, of a three-dimensional homogeneous Fermi gas at zero temperature. Parameters are fee = 30.0/CF and fcFo = -0.427 [i.e., [/(O) = -0.256.EF and A(0) = 0.0491EF]. In general, we define the correlation function between a pair of atoms with momenta fi(k' + k) and ?i(k' — k) as
k)},
(11)
Figure l(b) shows the pair correlation function <7o(k) for the initial state i.e., the initial pair correlation between atoms with momenta fik and — ftk, in the kz = 0 plane. It takes the maximum possible value of 1/2 at |k| « k'F. 3.
Results of Bragg scattering
The result of applying a Bragg pulse of length tp to an initial BCS state is shown in Fig. 2, where q is in the x direction, and for clarity we have
144
chosen the case of scattering well outside the Fermi sea, i.e., q > 2kp. Column Density u> — 9.2wF
(b)
(a)
^jjjjp'
^SP
uj = 15.6u>F
J
(c)
0* 8
-
4
0
MM
4
-
4
(d)
f 0
4
MM
Fig. 2. Column density, and pair correlation function |G q / 2 (k)| in the kz = 0 plane, of a Bragg scattered three-dimensional homogeneous Fermi gas at zero temperature. The plotted quantities are saturated on the chosen scales in order to clearly observe the features. Parameters are A = l.SOEf, q = 4.80fcF, t = 8.22/ojF, kc = 30.0fcF, kFa = -0.427, and (a), (b) u; = 9.2wF, and (c), (d) ui = 15.6wF.
In the momentum space column density plots (left-hand column), the initial state is a spherical atom cloud centered at k = 0. A fraction of the initial population has been scattered by momentum /iq, by the usual singleparticle Bragg scattering (e.g., Ref 17). This is understood in terms of the well known resonance condition
(12) which equates the net (photon) energy HUJ supplied by the Bragg field to the change in kinetic energy of a particle scattered from momentum fik' to H(k' + q). For a given u> and q, the particular momentum components in the initial distribution that are resonant with the Bragg field have a momentum ftk', where k' satisfies equation (12). The momentum-space width of the transition is S w ^/(
145
Of more interest is the correlated-pair scattering," which gives rise to a spherical shell centered at momentum hq/2 in Figs. 2(a) and (c). The correlated-pair scattering has a marked dependence on the Bragg frequency w,appearing as a tight cloud a t a threshold value Wthres [see Fig. 2(a) and equation (4.1) below] but with a radius that increases as w increases, as seen in Fig. 2(c). In the right-hand column of Fig. 2 we plot the correlation function Gq12(k),which shows that the atoms in the spherical shell are correlated as spin-up spin-down pairs, with momenta h(q/2+k) and h(q/2k). The maximum value of lGq,2(k)l in Fig. 2(d) is 0.394. We have investigated the dependence of the correlated-pair scattering on a range of system parameters. For the example shown in Fig. 2 , approximately 0.2% of the atoms are scattered into the pair-correlated spherical shell. The number of scattered pairs grows linearly with the length of the Bragg pulse (until t 7 0 / w ~ )The . number of pairs scattered increases quadratically with the initial pair function A (0) , as shown in Fig. 3. Approximately 6% of the initial population is scattered into the pair-correlated shell for A(0) = 0.293 ( k ~ = a -0.689).
-
AIEFl
Fig. 3. Number of scattered pairs vs the initial pairing field A(0) (on a log-log scale). Parameters are A = 1 . 8 0 E ~q, = 4 . 8 0 k ~ ,t = 8 . 2 2 / w ~ ,Ic, = 3 0 . 0 k ~ and , w =1 0 . 8 ~ ~ .
The number of scattered pairs also increases as the Bragg amplitude A increases, or as the Bragg wavevector q decreases. However, if lql becomes too small the correlated-pair scattering is supressed because the scattered pairs fall within the Fermi sea. At finite temperatures, the correlated scattering decreases roughly in proportion to the decrease in the initial pair function, i.e., as A2 ( 0 ) .
146
4. Mechanism for Correlated-Pair Scattering
The fundamental mechanism for correlated-pair scattering is the formation by the Bragg fields of a moving grating in the pair function, which we can describe as A(r, t ) M A, ( t ) A, ( t )exp [i (q . r - w t ) ] . The correlated pairs then scatter from this grating to produce the spherical shell illustrated above.
+
4.1. Geometry of correlated scattering An explanation can be given for the geometric behaviour of the correlatedpair scattering, noting that in a correlated Bragg scattering event the photons provide net energy hu and net momentum hq to the initial pair of correlated atoms. Before the event, the individual momenta are (hk’, -hk’) and total momentum 0. After scattering, the pair has total momentum hq, distributed amongst the individual atoms as (ri (q/2 krel) h ( q / 2 - krel)). Energy conservation requires the photon energy to equal the change in energy of the scattered atoms, i.e.,
+
Equation (13) describes a sphere of momentum radius krel centred on q/2 and we have used the fact that initially, the most strongly pair correlated atoms reside on the Fermi surface (i.e.] lk’l = k ~ )At . the correlated scattering threshold frequency, the radius of the scattered sphere is krel = 0, giving
4.2. Analytic Model A more detailed understanding of the correlated scattering can be obtained by noting that due t o the periodic nature of the Bragg potential, the solutions to equation (9) and (10) must have Bloch form, and can be expressed
147
as
*hZ
dt
=t..;(k)a;(t)
1 + ;A
[&-l(t)
+ C Urn(t)If(k/2 + nq/2
+~k+,(t)] -
mq/2)12alc-m(t)
m
+ C~m(t)f*(k+nq-mq)blc-m(t),
(18)
m
where b z ( k ) = Ek+nq-p-nb, and E k = h 2 k 2 / 2 M .A similar equation is found for dbk/dt by substituting Eqs. (14) - (17) into Eq.(lO). By analysing resonance conditions and coupling terms, we can show that for each mode k, the dominant coefficients are uk, u f , bk,, bk, as we have also confirmed from numerical simulations of the full time-dependent Bogoliubov de Gennes equations. In terms of the dominante coefficients, the pair potential grating amplitude is
In the initial BCS state, the amplitude A, is zero because the only nonzero coefficients are uk (for Ikl 2 kf.) and bk (for Ikl 5 kf.). The process of correlated scattering begins by first creating A, by single-particle scattering of atoms on the Fermi surface, with Ikl M kf.. This process can be approximately described by a truncated version of equation (B), namely
which converts u i to uy , creating a non-zero pair-potential grating amplitude, namely A, M -9 CFermi Surface uy(t)b:* ( t ) .We emphasize that this is a coherent process which transforms the initial correlated pair (hk, -hk) to
148
a new correlated pair (hk+ q, -tik). This process forms part of the correlation Gq,2(k),i.e. that which is represented by the outer circles in Fig. 2(c) and (d), whose radii are independent of the Bragg frequency w . Once A, is formed, it provides a coupling from uk to b k l , which is resonant for the region of k in the correlated shell (i.e., at q / 2 f krel). The population in the correlated shell is given by lbk112. We can follow the evolution of b k l ( t )approximately by truncating equation (18), and the corresponding equation for d b k l d t , to yield the set of equations
ih& dt
[
] [ AT@) =
Ek
-
p’
bkl(t)
I):;$[
+ + h]
-Ek--q A1 ( I-1’ t)
3
(21)
which can be understood in terms of the well known Rabi problem. The coupling frequency is A, [which is seeded as discussed above, but must be calculated self consistently according to equation (19)],but b k , is generated only if (Ek
-
P’)
-
(-Ek-q
+ p’ + h) 0. M
(22)
Equation (22) can be rewritten as
which is equivalent to equation (13). Alternatively it can be written w Wthres = li2k:el/M, which shows that above the threshold frequency, any additional energy provided by the Bragg field (i.e., h) is distributed equally between the two atoms of the scattered pair as kinetic energy. From equations (19) and ( 2 l ) , we find that the total amount of correlated scattering g 2 . Therefore, the parameter g must be accurately known in order to quantify the correlated scattering, which was a motivation for the detailed collisional formalism of section 2.1.
-
References 1. C. Chin, M. Bartenstein, A. Altmeyer, S. Riedl, S. Jochim, J. Hecker Denschlag and R. Grimm, Science 305,p. 1128 (2004). 2. J. Ruostekoski, Phys. Rev.A 61, p. 033605 (2000). 3. A. Minguzzi, G. Ferrari and Y. Castin, Eur. Phys. J . D 17, p. 49 (2001). 4. M. Rodriguez and P. Torma, Phys. Rev. A 66, p. 033601 (2002). 5 . H. P. Buchler, P. Zoller and W. Zwerger, Phys. Rev. Lett. 93, p. 080401 (2004). 6. Bogdan Mihaila, Sergio Gaudio, Krastan B. Blagoev, Alexander V. Balatsky, Peter B. Littlewood and Darryl L. Smith, Phys. Rev. Lett. 95, p. 090402 (2005).
149 7. R. Combescot, S. Giorgini and S. Stringari, Europhys. Lett. 75 (2006). 8. Bimalendu Deb, J . Phys. B 39 (2006). 9. J. Stenger, S. Inouye, A. P. Chikkatur, D. M. Stamper-Kurn, D. E. Pritchard and W. Ketterle, Phys. Rev. Lett. 82, p. 4569 (1999). 10. M. Kozuma, L. Deng, E. W. Hagley, J. Wen, R. Lutwak, K. Helmerson, S. L. Rolston and W D Phillips, Phys. Rev. Lett. 82, p. 871 (1999). 11. P. G. de Gennes, Superconductivity of metals and alloys (W. A. Benjamin, Inc., New York, 1966). 12. G. Bruun, Y. Castin, R. Dum and K. Burnett, Eur. Phys. J . D 7, p. 433 (1999). 13. M. H. Szymanska, K. G6ra1, T Kohler and K Burnett, Phys. Rev. A 72, p. 013610 (2005). 14. C. A. Regal, M. Greiner, S. Giorgini, M. Holland and D. S. Jin, Phys. Rev. Lett. 95, p. 250404 (2005). 15. J. B. Ketterson and S. N . Song, Superconductivity (Cambridge University Press, Cambridge, 1999). 16. N. Nygaard, G. M. Bruun, C. W. Clark and D. L. Feder, Phys. Rev. Lett. 90, p. 210402 (2003). 17. P. B. Blakie and R. J. Ballagh, J . Phys. B 33, p. 3961 (2000). 18. K. J. Challis, R. J . Ballagh and C. W. Gardiner, Phys. Rev. Lett. 98, p. 0930002 (2007).
This page intentionally left blank
SPECTROSCOPY AND CONTROL OF ATOMS AND MOLECULES
This page intentionally left blank
STARK AND ZEEMAN DECELERATION OF NEUTFCAL ATOMS AND MOLECULES
s. D. HOGAN, E. VLIEGEN, D. SPRECHER, N. VANHAECKE*, M. ANDRIST, H. SCHMUTZ, U. MEIER, B. H. MEIER AND F. MERKT Laboratorium fur Physikalische Chemie, ETH Zurich CH 8006 Zurich, Switzerland Argon and hydrogen atoms excited to Rydberg Stark states in supersonic expansions have been decelerated using inhomogeneous electric fields. In the case of hydrogen, the atoms have been decelerated from an initial velocity of 700 m/s to zero velocity in the lab frame using time-dependent inhomogeneous electric fields and subsequently stored in two- and three-dimensional traps. The dynamics of the Rydberg atoms in the traps and the phasespace characteristics of the decelerated atoms have been characterized by pulsed field ionization and imaging techniques. Multi-stage Zeeman deceleration of ground state H and D atoms has been demonstrated. Using this technique H atoms, traveling at 420 m/s, have been decelerated to half of their initial kinetic energy.
1. Introduction The development of general methods with which to control the translational motion of atoms and molecules in the gas phase is of interest in high-resolution spectroscopy and in studies of reactive collisions at very low collision energies. Laser cooling which has been very successful in stimulating research on (u1tra)cold atoms is not suited to generate cold samples of molecules in the gas phase. Over the past years, many promising techniques have been introduced that enable the production of such samples: photoassociation of cold atoms [ 11, Stark deceleration of polar molecules [ 2 ] , buffer gas cooling [3], low velocity filtering from effusive beams of polar molecules [4], billiard collisions [5]. Several methods rely on the use of cold samples formed in supersonic expansions and aim at stopping them in the laboratory frame. The methods employed so far include multi-stage Stark deceleration of polar molecules [6], optical Stark deceleration [7], and Rydberg Stark deceleration [8]. Recently, the
*
Permanent address: Laboratoire Aime Cotton du CNRS, UniversitC de Paris-Sud, 91405 Orsay, France
153
154 Zeeman equivalent of the multi-stage Stark deceleration method has also been realized experimentally [9]. We summarize here recent progress in Rydberg Stark deceleration and in multi-stage Zeeman deceleration of free radicals with emphasis on the work carried out in Zurich on atomic samples of argon and hydrogen. The methods of Rydberg Stark deceleration and of Zeeman deceleration can also be applied to molecular samples: Because all atoms and molecules possess Rydberg states, Rydberg Stark deceleration can in principle be applied to any particle in the gas phase. Zeeman deceleration is restricted to atoms and molecules subject to an electron Zeeman effect at low magnetic fields. Fig. 1 illustrates schematically the Stark effect in Rydberg states of atomic hydrogen (panel a), and of a nonhydrogenic atom or a molecule (panel b), and the Zeeman effect in the ground state of atomic hydrogen (panel c). Quantum systems undergoing a linear Stark or Zeeman effect can be thought of as having an electric belec) or magnetic (,urn,,) dipole moment which directly couples to the applied external fields according to equation (1).
A quantum state the energy of which is lowered (raised) in the presence of a field is often referred to as a "red-shifted" (blue-shifted) state, or a "high-field seeking" (low-field seeking) state. Particles undergoing a Stark or a Zeeman effect are subject to a force f in inhomogeneous fields which is proportional to the field gradient
In inhomogeneous fields, particles in low-(high-)field seeking states are accelerated in the direction of decreasing (increasing) field strength. If the force is directed perpendicularly to the beam propagation axis, it can be exploited to deflect an atomic beam or split it into different components, as in the wellknown experiments of Gerlach and Stem. If the acceleration vector points in the direction parallel or antiparallel to the beam propagation axis, the particles in the beam are accelerated or decelerated in the longitudinal dimension.
155
Figure 1. The Stark effect in Rydberg states of the H atom (a) and of nonhydrogenic atoms (Na) (b). (c) represents the Zeeman effect in the ground state of atomic hydrogen (adapted from Refs. [9,10 and 11]).
The energy levels of the H atom in an electric field of strength F are given to first approximation, in atomic units, by [10] E = E{- H2n2 + (3/2)nkF,
(3)
where n is the principal quantum number, E\~ 1/2 Hartree the ionization energy and A: is a quantum number which runs from - (n - I - \mf\) to + (n - 1 - \mt\) in steps of two. At zero electric field, the potential is Coulombic so that the energy eigenvalues do not depend on -6 or mt and have a degeneracy factor of n2. When an electric field is applied, the energy level structure fans out in linear manifolds of Stark states. States with k < 0 are red-shifted or high-field seeking, states with k > 0 are blue-shifted or low-field seeking, and states with k = 0 have, to a good approximation, a field-independent energy. Atoms and molecules excited to k = 0 Rydberg Stark states thus do not decelerate nor accelerate in an inhomogeneous electric field. If a H atom excited to a high-field seeking state in an electric field F flies out of this field, it decelerates, whereas an atom in a lowfield seeking state accelerates. The loss/gain in kinetic energy is equal to the Stark energy AE = (3/2)nkF.
(4)
The Stark effect in other atoms and molecules differs from that of the H atom in that (i) the penetrating low •£ states undergo a quadratic Stark shift at low fields and (ii) the crossings between Stark states at high field, which are exact in H, are avoided as a result of the non-Coulombic nature of the potential at short range (see Fig. 1). These qualitative differences between hydrogenic and nonhydrogenic systems must be considered when planning Rydberg Stark deceleration experiments and analyzing their results [12]. Nevertheless, the
linear manifolds of the nonpenetrating Stark states are sufficiently similar to those of H that Rydberg Stark deceleration is applicable to any particle in the gas phase. A property of Rydberg atoms and molecules particularly attractive for atom and molecule optics experiments is the fact that the dipole moment of the extreme components of the Stark manifolds scales as n2 and becomes very large at high n values (at n = 20 the dipole moment of a Ikl n state amounts to more than 1000 Debye). A qualitative understanding of the Zeeman deceleration of radicals can be gained from similar considerations taking into account the fact that the Zeeman shifts are proportional to the magnetic quantum number ms (or mJ) and the magnetic field strength. However, the Zeeman shifts of a radical are very small compared to the Stark shifts of a Rydberg atom or molecule, even if magnetic fields of more than 1 Tesla are used. For the Zeeman shift of a radical to be on the order of 1 ern-', a magnetic field in the range 1-2 Tesla is required. To achieve a similar Stark shift in the ground electronic state of a polar molecule, electric fields of 100 kV/cm or more must be applied. In contrast, a field of only 25 V/cm can induce Stark shifts of more than 1 cm-' in Rydberg states with n > 25. Given that the kinetic energy of a particle of mass number A4 moving at a velocity of 400 m/s in a supersonic beam of xenon amounts to
-
Ekin /
(hc cm-' )= 6.7 M
(5)
and that pulsed electric fields beyond 150 kV/cm and pulsed magnetic fields beyond 1-2 Tesla are difficult to generate experimentally, one can immediately conclude that while a single deceleration stage may be sufficient to completely decelerate a beam of Rydberg particles, the complete deceleration of a radical beam in a Zeeman decelerator will necessitate many (of the order of 10 to 100) successive deceleration stages, i.e., approximately as many as required for the complete Stark deceleration of a polar molecule in its ground electronic state. The Zeeman deceleration of radicals thus possesses many common features with the method of multi-stage Stark deceleration of polar molecules developed and successfully used to slow down and even trap polar molecules [2,6].
2. Stark deceleration and trapping of Rydberg atoms Breeden and Metcalf [13] and Wing [14] were the first to propose experiments in which inhomogeneous electric fields would be used to control
157 the translational motion of Rydberg atoms. Selected sentences from their abstracts "A method is proposed to decelerate neutral atoms in a thermal beam for the purposes of precision spectroscopy, atomic clocks, or loading into shallow traps. Rydberg atoms exhibit large dipole moments..." [13] and "Excited neutral atoms having positive Stark energies may be trapped near magnitude minima of electrostatic fields. At moderate field strengths Rydberg atoms have trap depths comparable to ambient kT [...]. Sustained trapping and cooling should be possible, allowing Doppler-free few-atom spectroscopy and novel collision studies" [14] make it clear that they had envisaged a wide range of applications. Their suggestions were first implemented in the laboratory of Tim Softley at Oxford where time-independent inhomogeneous fields were used to deflect beams of krypton Rydberg atoms [15] and both deflect and decelerate beams of hydrogen Rydberg molecules [8]. In these early proof-of-principle experiments, the change of kinetic energy did not exceed a few percent.
(a)
Photolysis La«*r (193nm}
JIL
Etectrottes
'"/'
Phosphor Screen
ceo Camera
NHjArjlrtO) 4
vuvauv /
mcp Quarts Capfflary
(c)
Figure 2. (a) Schematic representation of a Rydberg Stark decelerator. The capillary and photolysis laser are specific to experiments on radicals. The electrode configurations with equifield lines used to decelerate and trap Rydberg atoms are shown in panels (b) and (c), respectively (adapted from Refs. [16] and [20]).
158
The work in Zurich, which is summarized below, has focused so far on the deceleration of argon [ 12,16,17] and on the deceleration, mirroring and trapping of H atoms [18-201. A schematic illustration of the experimental set up is depicted in Fig. 2a, and Figs. 2b and 2c show electrode configurations used to decelerate (2b), and trap the Rydberg atoms in two dimensions (2c). Fig. 2c also displays the electric field distribution along the y (symmetric distribution) and z (asymmetric distribution) directions. The atomic beams are produced in a supersonic expansion at the exit of a pulsed nozzle. To generate the H atoms, ammonia in a lO:l, Ar:NH3 gas mixture is photolyzed in a quartz capillary mounted at the exit of the nozzle [ 181. The atoms move at a velocity in the range 650-700 m l s and are photoexcited to Rydberg Stark states in the electric field generated by a set of electrodes depicted in Figs. 2b-c. The atoms then propagate through the electrode setup in the direction of a microchannel plate detector the anode of which is connected to a phosphor screen. The Rydberg atoms ionize in the field of the detector and the ionization signal is converted into both a time-of-flight profile (extracted on the back plate of the MCP) and an image recorded by a charged-coupled-device (CCD) camera located behind the phosphor screen. When the Rydberg atoms are reflected back or trapped in the electrode set up, they do not reach the MCP detector. However, they can be field ionized after an adjustable delay by applying large pulsed voltages on the electrodes. The polarity of the pulsed voltages is chosen so as to extract the ions toward the MCP detector. The TOF of the ions can be used to determine the position of the Rydberg atoms at the time of field ionization. Indeed, ions produced close to the exit electrodes (3 and 4 in Figs. 2b and 2c) arrive later than ions produced close to the entrance electrodes (1 and 2 in Figs. 2b and 2c). In the first deceleration experiments on argon Rydberg atoms [ 12,161, the main limitation associated with the use of time-independent fields was identified as being related to the adiabatic traversal of the avoided crossings in the Stark map that occur at fields beyond the Inglis-Teller field, i.e., the field at which the Stark states of neighboring n values first cross (See Fig. 1). The adiabatic traversals of these crossings make it impossible to gain (or lose) more kinetic energy than the Stark energy at the Inglis-Teller field [12]. Much more kinetic energy can be removed using time-dependent fields [12] and an optimal deceleration experiment is one in which the voltages applied to the electrodes in the decelerator are constantly adjusted to the instantaneous positions of the Rydberg atoms so that the deceleration takes place at constant electric field and maximal field gradient [ 16,171. The results of an experiment using timedependent electric fields are summarized in Fig. 3 and demonstrate that the time
159 dependence of the applied electric fields (here an exponentially decaying function, see Ref. [16] for more details) can be tailored to enhance the deceleration efficiency and that a kinetic energy more than three time larger than the Stark shift at the operating field can be removed in a single deceleration stage. Under these conditions, a small acceleration of the particles in the transverse dimensions is unavoidable. However, the translation temperature of the decelerated sample can be kept well below 1 K as was demonstrated in Ref. [ 171 by imaging the cloud of decelerated atoms and comparing the experimental images with images calculated with a trajectory simulation program. The Rydberg Stark deceleration experiments on atomic hydrogen have led to the development of a mirror [ 191 in which a cloud of H Rydberg atoms initially moving at a velocity of 700 m / s is stopped over a distance of 1.8 mm in less than 5 ps, and reflected back to the initial spatial distribution. The mirroring process is illustrated in Fig. 4, in which panel (c) shows the results of measurements of the central position of the atom cloud during the mirroring process, panel (a) the simulated distributions of atoms in the yz plane at selected times during the mirroring process and panel (b) images of the actual distributions at times of 4 and 6 ps.
Figure 3 Comparison of the time-of-flight (TOF) profiles measured after decelerationiacceleration of n = 15 argon Rydberg atoms with time- independent and time-dependent inhomogeneous electric fields (a) and their simulation (b). The dotted line corresponds to a reference measurement with k = 0 Rydberg states which do not accelerate or decelerate. The inner dashed and full profiles represent measurements on blue- and red-shifted states with time-independent electric fields for which the change of kinetic energy corresponds closely to the Stark shift. The outer dashed and full TOF profiles represent the corresponding measurements with time-dependent fields in which the changes of kinetic energy are approximately three times the Stark shifts. The fact that the red-shifted states are decelerated when using time-independent fields and accelerated using time-dependent fields is a consequence of the opposite signs of the field gradients that result from the different voltage configurations (adapted from Ref. [ 161).
160
Recently these experiments have culminated in the development of two[20] and three- [21] dimensional traps for Rydberg atoms in selected Stark states. The electrode configuration used for the three-dimensional trapping experiments is depicted in Fig. 5a. In the experiments, the deceleration to zero velocity in the laboratory frame is achieved using time-dependent inhomogeneous electric fields generated by applying exponentially decaying voltages to the electrodes as described above for the deceleration of argon. The voltages decay to small nonzero final values corresponding to the electric field distributions displayed in Figs. 5b and 5c. The trap is narrower in the y and z dimension than in the x dimension so that a large breathing motion of the trapped Rydberg atom cloud can be observed by imaging techniques, as depicted in Fig. 5d. Typically around 104-105 atoms can be trapped at each measurement cycle in a volume of about 1x1x6 mm3, as determined from the amplitude of the field ionization signals and the size of the images. The trap losses result primarily from radiative processes although collisions with atoms in the trailing part of the gas pulses and transitions induced by black-body radiation also play a role (see Refs. [20,21] for a more complete discussion). These results represent the first demonstration of a three-dimensional trap for state-selected Rydberg atoms.
Figure 4. Operation principle of a Rydberg atom mirror. The electrode configuration used for deceleration and mirroring corresponds to that displayed in Fig. 2b or 2c. (a) Simulated evolution of the Rydberg atom cloud during the mirroring process which shows the focusing of the cloud taking place at around 4-5 us. (b) Experimental ion images at times of 4 and 6 us illustrating the focusing effect apparent in the simulations presented in panel (a), (c) Measured positions (full dots) of the center of the Rydberg atom cloud during the mirroring process. The open circles correspond to a reference measurement for which the electrode voltages were kept at 0 V (adapted from Ref. [19]).
Because all atoms and molecules have Rydberg states, the Rydberg Stark deceleration and trapping techniques outlined above have, at least in principle, the potential of being universal methods to generate translationally cold samples
161
of atoms and molecules. The atoms and molecules in these samples are in electronically excited states and exhibit very large dipole moments, which may turn out to be an attractive feature for quantum information processing. The radiative lifetimes of Rydberg Stark states at n values beyond 20 exceeds 100 (is and can be made much longer (i.e. > 1 ms) by optically preparing Stark states with mt > 3. The possibility of decelerating and trapping H atoms may offer opportunities for precision measurements of the Rydberg constant and may lead to new experimental schemes [22] for research on anti-hydrogen which is currently generated in Rydberg states [23]. Our future plans include the deceleration and trapping of hydrogen and NO molecules.
Figure 5. Electrode configuration (a) with equifield lines in the yz (b) and xz planes (c) used to trap hydrogen atoms in n — 30, k = 25 Rydberg states in three dimensions, (d) Breathing motion of the Rydberg atom cloud in the x dimension observed experimentally by ion imaging (adapted from Ref. [21]).
3. Zeeman deceleration of hydrogen and deuterium Atoms and molecules with unpaired electrons (so-called free radicals) tend to be particularly reactive. Several chemical reactions involving radicals are barrierless and exothermic processes and are expected to still take place at very low temperature [24]. By virtue of their unpaired electrons, free-radicals are subject to an electron Zeeman effect (see Fig. Ic). The translational motion of free radicals can thus be controlled by inhomogeneous magnetic fields and the
162
use of such fields in atom optics experiments has a long and rich history in physics starting with the experiments of Gerlach and Stern which enabled the measurements of magnetic moments and led to the experimental observation of space quantization [25-27]. A natural way to generate cold samples of free radicals at rest in the laboratory frame consists of using a supersonic expansion in combination with a multi-stage Zeeman decelerator of the type depicted in Fig. 6 which was recently constructed and tested at ETH [9]. The decelerator is placed along the axis of a supersonic beam of radicals generated using a versatile radical source developed in our laboratory [28]. The decelerator consists of several copper solenoids (length 7.8mm, inner diameter 5 mm, 42 windings, diameter of the copper wire 300 |j,m, operating current up to 250 A, maximal field on axis 1.4 Tesla) through which current pulses with rise and fall times, 5/r and 8tf, of less than 5 j^s can be applied (see Fig. 7b). Figure 7 [panel (a)] shows the magnetic field distribution in and around a solenoid, and a magnetic field pulse measured using a small pick-up coil placed inside the solenoid [panel (b)] [29]. s»I»a-ilf3ai9i
Lc
Quartz ««piilay
To
Pulsed
Figure 6. Schematic representation of the Zeeman decelerator developed at ETH. The pulsed valve delivering the supersonic beam is separated from the solenoids by a skimmer. At the end of the array of solenoids the supersonic beam is intersected at right angles by vacuum ultraviolet and/or ultraviolet laser beams between two parallel metallic plates used to extract the ions produced by photoionization toward a microchannel plate detector (adapted from Ref. [29]).
As the free radicals in the beam approach each solenoid in the decelerator, they are subject to an increasing magnetic field and the atoms in magnetic sublevels with positive Zeeman shifts are decelerated. Before the atoms reach the position of maximal field strength, the magnetic field is rapidly turned off to prevent the atoms from being accelerated when leaving the solenoid. The deceleration cycle is repeated at each successive solenoid in the decelerator. In optimizing the sequences of pulses to be applied to the solenoids
163
care has to be taken to (i) preserve phase stability during the whole deceleration process both in the transverse and longitudinal dimensions, and (ii) prevent the magnetic field from returning to zero in order to avoid transitions to other magnetic sublevels (see Ref. [29] for more details). At the end of the decelerator, the atoms are photoionized between two metallic plates and the ions are extracted by an electric field pulse to an MCP detector. Time-of-fiight profiles are recorded by monitoring the ion signal as a function of the time delay between the production of the radicals and the laser pulse used to photoionize them. (a)
(b)
Figure 7. (a) Magnetic field distribution in and around one of the solenoids calculated for an operating current of 250 A. (b) Temporal profile of a typical magnetic field pulse measured using a pick-up coil placed in the middle of the solenoid (adapted from Refs. [9,29]).
To demonstrate the operational principle of the Zeeman decelerator, Fig. 8a shows a set of deceleration measurements carried out on hydrogen atoms in their ground state. The lowest trace shows a time-of-flight profile corresponding to a reference measurement with the decelerator turned off and provides information on the velocity distribution of the H atoms in the beam. Comparing this reference measurement with the time-of-flight profiles obtained with pulse sequences designed to decelerate atoms of selected initial velocities, one observes two main differences: First, a late peak, corresponding to the decelerated atoms, is observed and its position shifts to longer times of flight as the selected initial velocity of the atoms to be decelerated decreases, as expected. Second, a clear increase of the signal at the edge of the main time-offlight peak is observed. This signal increase arises from the lateral guiding effect of the decelerator that results from the fact that the magnetic field in the solenoids increases as one moves away from their symmetry axis (see Fig. la). The profiles presented in Fig. 8b were obtained in full three-dimensional
164
simulations of the trajectories of the H atoms in the beam. The simulations involved no adjustable parameters. The analysis of these simulations and the good agreement between experimental and simulated data enable one to determine the final velocities of the decelerated atoms which are given above each trace in Fig. 8b. From the comparison between experimental and simulated time-of-flight profiles, one can conclude that close to half of the initial kinetic energy can be removed with only seven stages. The last panel (c) of Fig. 8 compares two deceleration measurements carried out on H and D atoms under otherwise identical experimental conditions (a pulse sequence optimized for the same initial velocity, the same carrier gas (krypton), etc.). As expected from the larger mass, the decelerated peak in the case of D appears at earlier times than in the case of H. We attribute the better signal-to-noise ratio of the D measurement to the fact that the H atoms are located closer to the edge of the gas expansion cone than the heavier D atoms and that, consequently, fewer H atoms pass through the skimmer placed at the entrance of the decelerator (see Fig. 6) and eventually reach the detector.
Figure 8. Zeeman deceleration of H atoms using a seven-stage Zeeman decelerator. (a) Measured time-of-flight profiles after deceleration with magnetic field pulse sequences optimized for atoms moving at selected initial velocities. (b) Simulations of the experimental time-of-flight profiles for comparison with the experiments. The velocities indicated in (a) represent the velocities of the atoms for which the magnetic field pulse sequence was calculated. Those indicated in (b) represent the final velocities of the decelerated bunches of atoms. (c) Comparison of time-of-flight profiles of H and D atoms obtained after Zeeman deceleration (adapted from Ref. [29])
These measurements represent the first demonstration of multi-stage Zeeman deceleration. This method may become useful to generate cold samples of free radicals. A 100-stage decelerator is currently under construction with which we plan to decelerate, and ultimately trap, molecules such as NH and oxygen.
165
4. Conclusions Two new methods have been described with which translationally cold samples of atoms and molecules can be generated. Both methods rely on the cooling of internal and translational degrees of freedom that takes place in supersonic expansions and involve decelerating the particles in the supersonic beam from a high velocity (between 400 and 2000 m/s depending on the carrier gas) to zero velocity in the laboratory frame. Rydberg Stark deceleration exploits the very large electric dipole moments of atoms and molecules in Rydberg states. With inhomogeneous electric fields of moderate strength, Rydberg atoms and molecules can be decelerated to zero velocity over only a few millimeter and in a few p, and loaded in two- or three-dimensional electrostatic traps. Zeeman deceleration relies on the linear electron Zeeman effect exhibited by most atoms and molecules with unpaired electrons. We hope that both methods will turn out to be useful tools in atomic and molecular physics.
Acknowledgments We thank Prof. T. P. Softley, Dr. S. Willitsch and Dr. M. Tomaselli for many discussions and their precious input to the work presented here. This work is supported financially by ETH Zurich and the Swiss National Science Foundation under Project Number 200021-1 13886.
References 1. A. Fioretti, D. Comparat, A. Crubellier, 0. Dulieu, F. Masnou-Seeuws, and P. Pillet, Phys. Rev. Lett. 80,4402 (1998). 2. H. L. Bethlem, G. Berden, and G. Meijer, Phys. Rev. Lett. 83, 1558 (1999). 3. J. M. Doyle, B. Friedrich, J. Kim, and D. Patterson, Phys. Rev. A 52, 2515(R) (1995). 4. S. A. Rangwala, T. Junglen, T. Rieger, P. W. H. Pinkse, and G. Rempe, Phys. Rev. A . 67,043406 (2003). 5. M. S. Elioff, F. F. Valentini, and D. W. Chandler, Science 302, 1940 (2004). 6. H.L. Bethlem and G. Meijer, Znt. Rev. Phys. Chem. 22,73 (2003). 7. R. Fulton, A. I. Bishop, and P. F. Barker, Phys. Rev. Lett. 93, 243004 (2004).
166 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22.
23. 24. 25. 26. 27. 28. 29.
S. R. Procter, Y. Yamakita, F. Merkt, and T. P. Softley, Chem. Phys. Lett. 374,667 (2003). N. Vanhaecke, U. Meier, M. Andrist, B. H. Meier, and F. Merkt, Phys. Rev. A. 75,03 1402(R) (2007). T. F. Gallagher, Rydberg Atoms, (Cambridge University Press, Cambridge, 1994). M. L. Zimmerman, M. G. Littman, M. M. Kash and D. Kleppner, Phys. Rev. A. 20,225 1 (1979). E. Vliegen, H. J. Worner, T. P. Softley, and F. Merkt, Phys. Rev. Lett. 92, 033005 (2004). T. Breeden and H. Metcalf, Phys. Rev. Lett. 47, 1726 (1981). W. H. Wing, Phys. Rev. Lett. 45,631 (1980). D. Townsend, A. L. Gooodgame, S. R. Procter, S. R. Mackenzie, and T. P. Softley, J. Phys. B. 34,439 (2001). E. Vliegen and F. Merkt, J. Phys. B. 38, 1623 (2005). E. Vliegen, P. Limacher, and F. Merkt, EPJD. 40,73 (2006). E. Vliegen and F. Merkt, J. Phys. B. 39, L241 (2006). E. Vliegen and F. Merkt, Phys. Rev. Lett. 97,033002 (2006). E. Vliegen, S. D. Hogan, H. Schmutz, and F. Merkt, Phys. Rev. A 76, 023405 (2007). S . D. Hogan and F. Merkt, submitted for publication. A. Kellerbauer et al., submitted to NIM B for publication as contribution to the proceedings of the Low-Energy Positron and Positronium Workshop, Reading, 2007 M. Amoretti et al., Nature 419,456 (2002). I. R. Sims and I. W. M. Smith, Ann. Rev. Phys. Chem. 46, 109 (1995). W. Gerlach and 0. Stem, 2. Phys. 9,349 (1922). W. Gerlach and 0. Stern, 2. Phys. 9,353 (1922). W. Gerlach and 0. Stern, Ann. Phys. 16,673 (1924). S. Willitsch, J. M. Dyke and F. Merkt, Helv. Chim. Acta 86, 1152 (2003). S. D. Hogan, D. Sprecher, M. Andrist, N. Vanhaecke, and F. Merkt, Phys. Rev. A. 76,023412 (2007).
GENERATION OF COHERENT, BROADBAND AND TUNABLE SOFT X-RAY CONTINUUM AT THE LEADING EDGE OF THE DRIVER LASER PULSE AURELIE JULLIEN, THOMAS PFEIFER, MARK J. ABEL, PHILLIP M. NAGEL, STEPHEN R. LEONE AND DANIEL M. NEUMARK Departments of Chemistry and Physics, University of California, Berkeley, CA 94720, and Chemical Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720 Coherent soft x-ray continuous spectra are experimentally generated at the leading edge of the driving laser pulse by temporal confinement of the high-order harmonics emission. The analysis relies on the measurement of individual half-cycle cutoffs [ l ] and the proposed interaction regime opens the route to the production of spectrally tunable isolated attosecond pulses.
1
Basics of attosecond pulses production
Recent developments in attosecond sources [2-41 are motivated by the desire to study electronic dynamics and electron-electron correlations on the atomic time scale. In particular, pioneering studies demonstrate the direct probing of ultrafast electronic motions following the creation of an inner-shell hole [5, 61. In another experiment, a subfemtosecond XUV pulse probes the motion of an electronic wave packet under the influence of the electric field of an infrared pulse [7]. Very recently, the real-time observation of electron tunnelling in an atom exposed to a strong laser field with attosecond resolution is reported [8]. The ability to produce coherent attosecond laser pulses is opening the way to a wide field of investigation devoted to the measurement of various electronic processes in atoms, clusters and molecules. The continuous trend towards shorter laser pulses has made major progress because of Kerr-lens modelocking, self-phase modulation broadening, and dispersion control, but is limited by the intrinsic barrier of the visible optical period to the femtosecond timescale. To go beyond this femtosecond barrier, laser sources with higher carrier frequencies are needed. The generation of high-order harmonics (HHG) of intense laser pulses focused into a gas provides a tabletop coherent soft x-ray source. The HHG process can be described by a three step semiclassical model [9]. In the first step 167
of the interaction, an electron is released by tunnel ionization of an atom or molecule and, in the second step, accelerated by the laser field. The electron is driven back by the reversing electric field and can recombine with its parent ion leading to the emission of a photon whose energy depends on the recombining electron kinetic energy. The recombination occurs on an ultrashort time scale and leads to a broad spectral emission. The process is repeated every intense half-cycle for a multicycle laser pulse. Therefore, high-order harmonic generation is considered theoretically as a potential solution for attosecond pulse production [lo]. An experimental breakthrough was achieved when it was confirmed that the HHG process can emit a train of attosecond pulses [ l 11, the duration of which can reach 130 as [12]. So far, to extract a single attosecond pulse from this train, few-cycle carrierenvelope phase (CEP) stabilized laser pulses have been required [7, 131. When the CEP value is close to 0 (cosine-shaped pulse), the maximum of the electric field coincides with the maximum of the pulse envelope; the most energetic harmonic burst of XUV light is emitted for the highest field strength and consequently only once per pulse for this particular phase value. Then, adequate spectral filtering selects this highest energetic part of the harmonic spectrum and an attosecond pulse can be isolated. This method has led to the production of isolated pulses with a duration of 250 as (10 eV bandwith around 100 eV), about 1/10 of the optical period of a Ti:sapphire laser [7]. Another approach has combined few-cycle phase-stabilized pulses with a polarization gating technique [ 141 to confine emission of harmonic radiation to one optical cycle, enabling the generation of 130 as single pulses [ 151. Here, we present experimental evidence for the confinement of the emission of harmonic radiation to the leading edge of the driving laser pulse due to an ionization gating mechanism, without the need for polarization control [ 161. We demonstrate that this interaction regime allows the generation of quasicontinuum harmonic spectra, with a broad bandwidth and a tunable central frequency. These spectra are strongly indicative of an isolated attosecond pulse.
2
Confinement of HHG production at the leading edge of the driver pulse: towards tunable isolated attosecond pulse generation
The proposed analysis method relies on the recent work of Haworth et al. [l]: when the electric field of the driver pulse is composed of several half-cycles whose extrema vary significantly from one to the other, which corresponds to a rapid change for the pulse envelope on the timescale of an optical cycle (particularly true for few-cycle pulses), the harmonic cutoff emission of each individual half-cycle is visible on the harmonic spectrum as a local maximum.
169
Measurement of the positions of these individual half-cycle cutoffs (HCO) allows the precise knowledge of the absolute phase value of the laser pulse to be obtained. Extending this analysis, we show that it is also possible to determine conditions when the harmonic radiation is confined to the leading edge of a few cycle driver pulse. Our experiments have been carried out with a commercial Ti:sapphire femtosecond laser delivering 800 pJ, 25 fs pulses at 3 kHz repetition rate. A hollow-core fiber filled with neon (1.7 bar) and a chirped mirror compressor provide external pulse compression to the few-cycle regime (about 6 fs) [17, 181. A pair of glass wedges is used to finely adjust the pulse dispersion. Laser pulses are CEP stabilized (root mean square fluctuations rms 250 mrad). Highorder harmonics are produced by focusing the driver pulses into a 2 mm long gas cell filled with 180 mbar of neon. Two Zr filters (200 nm thickness each) remove the near infrared light and low order harmonics. The remaining radiation is sent to a soft-x-ray spectrometer composed of a Si3N4transmission grating and a soft-x-ray-sensitive CCD.
60
70
80
90
100
130
120
130
140
Photon energy (eV) Figure 1. High-harmonic spectra obtained for different CEP values (the initial offset w value is unknown). On the first graph, the transmission of the Si3N4 (grating material) is shown by the dashed line.
Figure 1 shows spectra registered when varying the CEP value around an unknown initial offset. The overall minimum transmission occurring beyond 100
170
eV corresponds to the absorption edge of Si in the spectrometer grating. The measured spectra extend up to 130 eV. In the whole spectral range explored (60 eV - 130 eV), a slow modulation carrying the harmonic peaks and local spectral maxima is visible. This modulation is CEP-dependent and the energy of the maxima (HCO) shifts towards higher photon energy when the relative CEP value is increasing. As the energy of each HCO is dependent on the amplitude of the corresponding electric-field extremum, it can provide interesting information about the absolute phase and the intensity envelope of the laser pulse. In order to extract this information from the HCO behavior, we measured harmonic spectra for a relative CEP varying from 0 to 2co in a/8 steps. To determine more easily the HCO energy, the harmonic modulation (at 2ojlaser) is removed by Fourier analysis and the spectra are normalized by the CEP-averaged spectrum. Fig. 2 shows the resulting spectra plotted as a function of CEP. Positions of the HCOs are indicated. 130 0 Norm, spsctrat
1
Intensity
1
2 3 4 5 Relative CEP value (rad)
6
Figure 2. Harmonic spectra as a function of the CEP. The harmonic modulation is removed and all the spectra are normalized by a CEP-average spectrum. The HCOs positions (local maxima) for each spectrum are marked with white points. One can observe the quasi linear dependence of the HCOs positions with the CEP.
Experimental data are fitted by a simple calculation of the HCOs positions as a function of the carrier-envelope phase using the following approximation: EHCO = Ip +3.17(E/2o)iaser)2 (in atomic units, E is the extremum field value of the half cycle, Ip is the ionization potential of the target gas) [1]. If we assume that the last cutoff is emitted at the maximum pulse intensity, the best result
171 corresponds to a 4.5 fs driver pulse and the HCO prediction clearly does not match the experimental behavior (fig. 3a). Indeed, for a fixed value of the phase, the gap between two successive HCOs is not decreasing when the harmonic photon energy is increasing, as would be expected for half-cycles reaching the maximum of a Gaussian intensity envelope. The HCO energy evolves linearly with the phase and does not exhibit any saturation. 130
(b)
x i •' ^ it .
- 1 0 1 3 3 4 Absolute CEP value (rad)
5
-
1
t
.
i
.+J 1^1 <
<
0 1 2 3 4 5 AbsclutB CEP value
Figure 3, Ex peri men tally recorded HCO energy as a dinclion of CEP (black squares). Considering a driver pulse of 4.5 & and the last cutoff corresponds to the maximum pulse intensity (a, dotted line), the fit does not describe the experimental HCO behavior. A fit to the leading edge of a 9 fs pulse assuming an ionization gate {b, dotted line) results in a good agreement with the experimental data.
The observations are understood if we include ionization of the medium in the HCO analysis. Cao et al theoretically demonstrated that a very intense pulse (in the range of 1015 Wcm"3) can create a Fast depletion of the ground state population, occurring on the leading edge of the laser pulse [19]. In our measurements, the experimental intensity in the harmonic cell reaches this high value, since the cell is positioned very close to the laser focus. In that case, the efficiency of harmonic generation decreases when the ionization yield increases and stops when the ground state is depleted. Ionization acts as a gate, as the harmonic emission is shut down not only on the trailing edge of the pulse, but also before the maximum pulse intensity can interact with the harmonic generation gas. We note that the gate could also be achieved by phase mismatching due to ionization, even without a complete depletion of the ground state. To include this gating phenomenon in the HCO calculation, we consider that the efficiency of the HHG process depends on the ground state population and drops rapidly to 0 on the leading edge. Fig. 3b shows a fit of the experimental HCO energy when the harmonic emission is restricted to the leading edge of a 9 fs pulse whose peak intensity would allow, without
172
ionization gating, emission of harmonic radiation up to 144 eV. Good agreement with experimental measurements is obtained. The duration of the driver pulse during the interaction is longer than expected. But, ultrashort laser pulses are extremely sensitive to excessive or insufficient dispersion compensation, which can be induced by the wedge positions, compressor setting or ionization-induced pulse reshaping due to the gas in the harmonic cell. As a consequence, the HCO analysis can be a useful tool to optimize in situ the pulse duration (see below). Notice also that a fit to the leading edge of an 8 fs or 8.5 fs pulse (not shown here) can also give convincing results. This uncertainty of the laser pulse duration prevents an extraction from our data of the absolute CEP value, as was done in [l]. This can be understood by the fact that the CEP is defined as the offset between the electric field extremum and the maximum of the pulse intensity envelope. In our case, the ionization gating mechanism prevents HCO investigations after the closure of the gate, and assumptions about the pulse duration and the pulse shape would be necessary to estimate the absolute CEP value. However, as we show below, in the ionization-gating regime, the CEP stabilization is required but a knowledge of the absolute phase value is no longer essential to produce isolated attosecond pulses. The direct consequence of this particular regime is the ability to spectrally isolate the contribution of adjacent half-cycles if the leading edge is sufficiently steep to provide significant intensity discrepancy from one half-cycle to another. In our experiment this condition is achieved by making the pulse shorter (about 6 fs by optimizing the compression parameters) and consequently more intense. Two resulting representative spectra are shown in figure 4(a). 250
.zc 150
.. 0
fe loo L
3P: 50 0
60
70
80
90
Photon energy (eV)
100
110
Time (fs)
Figure 4. (a) Tunable and broadband quasi continuum harmonic spectra for two CEP values. These spectra are obtained when the pulse is shorter and consequently the rising edge is steeper (neon pressure : 240 mbar). The remaining modulation is weak and would lead to satellite attosecond pulses with a relative intensity of less than 10% (b). The Fourier transform is calculated assuming a flat spectral phase for the solid line spectrum of (a).
173
As only one maximum is visible, we estimate that a single half-cycle is mainly responsible for the harmonic emission in the considered spectral range. As a consequence, the spectra exhibit a quasi-continuous shape, strongly indicative of an isolated attosecond pulse. The remaining weak spectral modulation results from a weaker attosecond pulse generated by the previous or the next half-cycle. The relative intensity of this (these) parasitic pulse(s) is estimated to be less than 10% by Fourier transformation of the experimental spectrum (fig. 4b). Another noteworthy feature is the width of these quasi-continuous spectra (about 15 eV), which is broader than the usual continuum generated at the peak of the pulse with the same duration. This is the result of the large difference in intensity between two consecutive half-cycles on the leading edge, which becomes smaller near the peak of the pulse. It also becomes clear from figure 4(a) that in this regime the CEP is a parameter to tune the central wavelength of the continuum. The CEP adjusts the intensity of the most intense half-cycle before the closure of the ionization gate and therefore the photon energy of the last HCO. One is no longer restricted to work with a CEP stabilized to 0, but the absolute phase can be used to spectrally tune the attosecond pulse to the preferred energy for possible experiments. 3
Conclusion
To conclude, we have experimentally demonstrated that broadband continuum soft x-ray continuum spectra can be generated at the leading edge of the driver pulse, by means of an ionization gating mechanism. This regime enables the isolation of part of the XUV emission of an individual half-cycle on the leading edge, opening the way to produce isolated attosecond pulses with a large bandwidth and a tunable central frequency controlled by the carrier-envelope phase. Acknowledgments
This work is supported by a MUM program from the Air Force Office of Scientific research, contract No. FA9550-04-1-0242. Portions of the laboratory were supported by the Director, Office of Science, Office of Basic Energy Sciences, of the US Department of Energy under contract DE-AC0205CH11231. T. P. acknowledges support of a Feodor Fellowship of the Alexander von Humboldt Foundation. References
1. 2.
C. A. Haworth, L. E. Chipperfield, J. S. Robinson, P. L. Knight, J. P. Marangos and J. W. G. Tisch, Nature Physics 3,52 (2007). P. Agostini and L. F. DiMauro, Rep. Prog. Phys. 67,813 (2004).
174
3. 4. 5.
6.
7.
8.
9. 10.
11. 12.
13.
14.
15.
16.
17. 18. 19.
A Scrinzi, M Yu Ivanov, R Kienberger and D M Villeneuve, J. Phys. B: At. Mol. Opt. Phys. 39, R1 (2006). P.B. Corkum and F. Krausz, Nature Phys. 3,381 (2007). M. Drescher, M. Hentschel, R. Kienberger, M. Uiberacker, V. Yakovlev, A. Scrinzi, T. Westerwalbesloh, U. Kleineberg, U. Heinzmann and F. Krausz, Nature 419,803 (2002). M. Drescher, M. Hentschel, R. Kienberger, M. Uiberacker, Th. Westenvalbesloh, U. Kleineberg, U. Heinzmann, and F. Krausz, J. Electron. Spectrosc. 137,259 (2004). R. Kienberger, E. Goulielmakis, M. Uiberacker, A. Baltuska, V. Yakovlev, F. Bammer, A. Scrinzi, T. Westerwalbesloh,U. Kleineberg, U. Heinzmann, M. Drescher and F. Krausz, Nature 427,817 (2004). M. Uiberacker, T. Uphues, M. Schultze, A. J. Verhoef, V. Yakovlev, M. F. Kling, J. Rauschenberger, N. M. Kabachnik, H. Schroder, M. Lezius, K. L. Kompa, H. G. Muller, M. J. J. Vrakking, S. Hendel, U. Kleineberg, U. Heinzmann, M. Drescher and F. Krausz, Nature 446,627 (2007). P. B. Corkum, Phys. Rev. Lett. 71, 1994 (1993). M. Y. Ivanov, P. B. Corkum, T. Zuo and A. Bandrauk, Phys. Rev. Lett. 74, 2933 (1995). P. M. Paul, E. S. Toma, P. Breger, G. Mullot, F. Auge, P. Balcou, H. G. Muller and P. Agostini, Science 292, 1689 (2001). Y. Mairesse, A. de Bohan, L.J. Frasinski, H. Merdji, L.C. Dinu, P. Monchicourt, P. Brtger, M. Kovacev, R. Taieb, B. Carre, H.G. Muller, P. Agostini, P. Salieres, Science 302, 1540 (2003). A. Baltuska, T. Udem, M. Uiberacker, M. Hentschel, E. Goulielmakis, C. Gohle, R. Holwarth, V. S. Yakovlev, A. Scrinzi, T. W. Hansch and F. Krausz, Nature 421,611 (2003). 0. Tcherbakoff, E. Mevel, D. Descamps, J. Plumridge and E. Constant, Phys. Rev. A 68,043804 (2003). G. Sansone, E. Benedetti, F. Calegari, C. Vozzi, L. Avaldi, R. Flammini, L. Poletto, P. Villoresi, C. Altucci, R. Velotta, S. Sagira, S. De Silvestri and M. Nisoli, Science 314,443 (2006). T. Pfeifer, A. Jullien, M. J. Abel, P. M. Nagel, L. Gallmann, D. M. Neumark and S. R. Leone “Generating coherent broadband continuum soft-x-ray radiation by attosecond ionization gating”, submitted to Opt. Exp. (2007). M. Nisoli, S. De Silvestri and 0. Svelto, App. Phys. Lett. 68,2793 (1996). L. Gallmann, T. Pfeifer, P. M. Nagel, M. Abel, D. M. Neumark and S. R. Leone, App. Phys. B 86,56 1566 (2007). W. Cao, P. Lu, P. Lan, X. Wang and G. Yang, Phys. Rev. A 74, 063821 (2006).
CONTROLLING NEUTRAL ATOMS AND PHOTONS WITH OPTICAL CONVEYOR BELTS AND ULTRATHIN OPTICAL FIBERS Dieter Meschedet , Wolfgang Alt , and Arno Rauschenbeutel* Insitut fur Angewandte Physik, Universitat Bonn, 591 15 Bonn, Germany t E-mail: meschede0uni-bon71.de www.iap. uni- bonn. de We store Caesium atoms in a 1D standing wave optical dipole trap with no more than one atom per site and sufficient spacing t o individually control the properties of each atom. The Caesium ground state hyperfine levels are used as qubit states, and all single-atom qubit operations have been realized in the past already, thereby realizing a neutral atom quantum register for a bottom-up approach towards quantum information processing with neutral atoms. With a second optical dipole trap we extract atoms from their trapping sites and insert them into to other micropotentials. This allows us to regularize the spacing of stored atoms, to insert two atoms into one and the same micropotential in order to induce interactions, and to insert atoms in a controlled way into high finesse optical resonators. In a new line of research we are exploring the potential of light confined by micro-structured thin optical fibers for controlled light-matter interaction in quantum optics. We have constructed fully tunable bottle micro-resonators and we have found that we can very efficiently detect and manipulate atoms or molecules in the immediate vicinity of fibres with sub-wavelength diameters. Keywords: Quantum information with neutral atoms, ultrathin fibres.
Introduction One great promise of controlling microscopic particles at the ultimate quantum level draws on the perspective of studying complex many particle systems with fully controlled interactions, for instance for applications in quantum information processing or quantum simulations.' For neutral atoms, there are two complementary approaches currently pursuing this attempt: The bottom-up and the top-down approach. The top-down approach relies 'Nvw address: lnsritiir fiir Physik, Illliversitat ?rlainz, hlaiiiz, C;criiiany
175
176
on 2D- and 3D- optical lattices filled with exactly one atom per lattice site. Such systems have become available with the realization of the superfluid Mott insulator transition. It has been shown already that highly correlated many-particle states can be generated with this system.2 In this report we describe in the first part advances with the bottom-up approach which for neutral atoms may be symbolized with Fig. 1 showing small strings of neutral Caesium atoms stored in a standing wave optical dipole trap (DT, ID optical lattice) which can be operated as an optical conveyor belt.3 In the second part we give preliminary results about a new concept of controlling light matter interaction with ultrathin optical fibres. 1. Controlling atoms in optical conveyor belts
Fig. 1. Fluorescence image of sorted strings of 1 to 7 neutral Caesium atoms stored in an optical conveyor belt. The conveyor belt consists of two counterpropagating laser beams at 1.03 /^m wavelength forming a ID optical lattice with ~ 0.5 /urn periodicity. Here, equidistant strings of 1-7 atoms were created by rearranging an inital string of atoms with random distributions in the ID lattice.4 The elongated shape of the spots is due to the pancake shaped micropotentials which are observed edge-on.
At the level of single atoms we have shown earlier (for an overview see ref. 5) that we can tightly control the internal dynamics of every atom stored in the conveyor belt. We have found that the quantum states of the well known atomic clock microwave transition connecting the hyperfine ground states F=3 —> F=4 form an excellent qubit system offering long coherence
177
times of tens of ms in the optical dipole trap. Individual addressability with a resolution of about 2.5 pm has been achieved, too, with a magnetic field gradient method. With strong experimental command realized for all single particle (or qubit) operations, the next important step is to prepare small many particle systems (typically 2-10 atoms) and to deterministically induce interactions between individual particles. We have taken first steps towards this goal with an accelerated experimental routine for the generation of a desired small number of atoms (Sect. l.l),by showing that we can deterministcally induce atom-atom interactions (Sect. 1.2), and by inserting atoms into and removing them from an optical high finesse cavity (Sect. 1.3). 1.1. Number-triggered loading of atoms 100-
9080; 70c 60-
-
P
2
2
50-
a 40-
302010-
. . . . . . . . . . . . . . . . . . . . . .
Our 1D lattice is always loaded from a magneto-optical trap (MOT) providing a small number of atoms with Poissonian distribution. In Fig. 2 we show the strongly enhanced preparation efficiency for generating a desired number of neutral atoms in our present system. Since the loading process is fast compared with all other experimental procedures it is simply interrupted and rapidly repeated until the desired number of atoms is found ("number-triggered loading"). One cycle takes typically 0.2 s. The success probability for larger atom numbers drops as a consequence of 1and 2-atom losses from the trap during preparation. We have found that during the loading routine atoms undergo light induced collisions favouring single occupation of the lattice sites. Details are found in ref. 6.
178
1.2. Induced atom-atom interaction The sorting procedure used to regularize the atom strings shown in Fig. 1 is realized by extracting atoms from the ID lattice with an optical tweezers and re-inserting the atoms at evenly spaced locations which are separated by more than 10 /im in the example. At such large separations an atom already residing in the DT is not affected by the second optical tweezers. The spatial resolution of the placement procedure is indicated by the distribution of atoms over the micropotentials shown in Fig. 3. At zero separation, i.e. insertion into exactly the same micropotential one expects from the data in Fig. 3 a probability of about 20% that the second atom is indeed inserted into exactly the same micropotential as the first one. To experimentally detect the operation we have used light induced collisional loss of the atom pair which occurs if and only if two atoms are present in exactly the same site. We have found a success rate of 16% in agreement with theoretical expectations. While this interaction - which actually resembles a chemical process - is still highly dissipative, it demonstrates that interactions can be deterministically induced at the level of exactly two atoms.
04-
23
25
27
28
31
33
35
Final distance between two atoms I)vml2]
Fig. 3. Spatial distribution of the placement routine with two atoms.7 The modulation of the distribution reflects the periodicity of the ID optical lattice. The resolution of the procedure can be improved by tightening the focus of the optical tweezers and by lowering atom temperature.8
1.3. Conveyor belts and optical high finesse cavities With cold atoms stored in dipole traps and interacting with the field of a high finesse optical cavity in the strong couling regime of cavity QED many impressive experimental results have been achieved already.9 The
179
strong coupling regime is realized with large atom-field coupling constants, <7/7 = \/14tom/Knode & and low field relaxation rates re, i.e. with large g2/(2K^) and g/K values. Controlled transport of atoms within10 and into11 a high finesse cavity has been reported previously. We have realized an apparatus that allows to not only simultaneously insert 1-10 atoms prepared by the methods described above into a cavity but also to take them out again and detect their quantum state (Fig. 4).
conveyor belt, fast transport
cavity parameters: finesse 1.1-106 Qvalue4-10 8 mode volume 1.1 -10s A,3
Fig. 4. Shuttling of atoms across an optical high finesse cavity in the strong coupling regime. At constant velocity the temporal dependence translates into a spatial profile of the atom-field coupling strength. The average resonant photon number was 0.5 here. The signal on the right has been averaged over 16 individual transits.
Atoms are prepared and detected about 4.6 mm outside the cavity with parameters indicated in Fig. 4. They are then transported into the cavity where interactions in the strong coupling regime of cavity QED take place. After the interaction atoms can be taken back to the original position where their presence as well as their quantum state can be detected. In a high finesse cavity driven by resonant laser radiation the transmission is expected to strongly drop when an atom is inserted with transmission due to vacuum Rabi splitting.9 In a preliminary experiment we have observed strong reduction of transmission (Fig. 4) when an atom is shuttled forward and backward across the cavity mode, mapping the spatial profile of the atom-field coupling strength. We observe long residence times of our atoms in the cavity of several s indicating that cavity mediated laser cooling is active.12 a
7 is the spontaneous emission rate, Vmo^e is the cavity mode volume, and V^tom = c
180
A characteristic of our cavity is a very high finesse (F ~ 1.1 • 106) corresponding to a photon life time of T = (2/t)"1 = 0.19 yus and for the Cs transition at A = 852 nm with natural line width 27/(2?r) = 5.2 • 106 s^1 to a ratio 7/K = 6. Thus damping in our system is dominated by atomic relaxation due to spontaneous emission. For comparison we have displayed in Fig. 5 the characteristic parameters of recent cavity-QED experiments using different set-ups. relaxation photon dominated
atom dominated
flj -o o
I 1
D5
<5
1 0.01
0.1
1
10
7/K Fig. 5. Overview: Relevant strong coupling parameters of cavity QED in recent experiments. Data points are explicited in Tab. 1.3.
#
Laboratory
(all in 106 1 2 3 4 5 6 7
8
Optical cavities with dipole traps 2. 6 Caltech15 16* 3.8 13 16* MPI Quantum Optics16 1.4 3 30 1.25 3 MPI Quantum Optics17 3.1* 1.3 3 Georgiatech11 17* 3.5 7 2. 6 18* 0.43 University of Bonn 145 Optical cavities on atom chips 215* 53 3 137 ENS Paris18 3 Imperial College19 97* 2180 .75 Microwave cavities, with superconductors and Rydberg atoms 1 105 .125 4.4-103 Yale University14 ENS Paris20 .05 6-10-7 0..003t 7-105
9 *. calculated for the strongest transition; t : from atom transit time.
4.2
11 2.5 4.8 42 3.9
0.05 840
9-10 4
The dividing line here is chosen to be 7/K = 1 where atomic relaxation
181
equals cavity relaxation. In a simplified manner of speaking one can say that physical information about the dynamics of the coupled atom-cavity system is gained to the left of this line by observing photons emitted from the system. On the right side it is more profitable to derive information about the status of the system by monitoring the dynamic evolution of long lived and coherent atomic quantum states. This is well known from the microwave domain which has been deeply explored by experiments at the ENS Paris.13 More recently it has been shown that this regime can be realized experimentally also with superconducting circuit systems.’ With our system it should be possible to venture into this regime also with an optical cavity-QED set-up which has the specialty of controlled insertion of several atoms simultaneously into the system. 2. Thin and ultrathin fibres
Having conquered communication technologies already optical fibres have begun to revolutionize also other areas of controlled propagation of light waves. For instance micro structured fibres, so called photonic crystal fibres, are now widely applied for efficient applications of nonlinear processes.” On of the simplest micro structuring methods is to reduce the thickness of standard 125 pm diameter optical fibres. We have begun to explore the properties of such thinned fibres with modulated cross section for applications in quantum optics. Here we report on an advantageous concept for ”bottle micro resonators” (Sect. 2.1), and on ultrathin fibre applications as an ”infinite focus” (Sect. 2.2). 2.1. Bottle resonators - tunable micro cavities
As outlined in Sect. 1.3 the realization of cavity QED processes for e.g. quantum information processing requires large atom-field coupling strengths g/y = & & o m / V m o d e . Built from discrete mirror substrates the performance (Vmode, K ) of the cavities used in experiments 1-5 in Tab. 1.3 seems to be limited by technical imperfections, and no clear route towards substantially improved values is in sight. An alternative is offered by micro resonators reyling on total internal reflection such as glass spheres or toroids.” Very high Q values in the lo8 to lo1’ range with small damping rates K rivaling the performance of conventional cavites have been demonstrated. More importantly, micro resonators promise further enhancement of the g-factor, since their mode volume is only of order 103X3 instead of 105X3,resulting in an order of magnitude enhancement of g.
182
We have found a novel geometry which can be fabricated with a fibre pulling apparatus where the optical fibres are thinned while they are heated with a flame or with a CO2-laser.23 The resonators look like little
photodiode
I °o
bottle resonator
•
excitation! laser Fig. 6. "Bottle resonators". The light ray travels along the circumference of 16 /urn diameter optical fibres. Due to angular momentum conservation turning points occur for identical diameters where the caustics generate strong field enhancement and provide efficient coupling ports. Light is coupled in and out of the bottle resonator by coupling to the evansecent field of an ultrathin fibre, see Sect. 2.2. On the right side we show a microscope image of a test fibre drawn by our appartus. For practical bottle resonators the diameter modulation is only a few % and not visible.24
bottles (Fig. 6) which may be considered an extremely prolate deformation of spherical resonators. The conceptual advantages of such resonators24 include a small free spectral range (100 GHz compared to I THz for spherical resonators), straightforward strain tuning over more than one free spectral range, and two input/output ports for convenient coupling by ultrathin fibres. Preliminary results in our laboratory indicate that such "bottle resonators" behave precisely as theoretically expected.
2.2. An "infinite" focus For coupling of external light fields to the bottle resonator described above the evanscent light field of ultrathin fibres provides a very convenient coupling mechanism (Fig. 6). The thin fibres are drawn on our pulling appartus, and Fig. 7 shows both the taper section transferring the light field of a conventional, weakly guiding fibre into the strongly guided mode of an ultrathin fibre. The electron microscope image on the lower left shows the
183 0,6
0,4
0,2
Fig. 7. The infinite focus. Upper left: Fibre taper section; lower left: ultrathin fibre, electron microscope image. Right: Calculation of the effective mode area of the guided wave (Aeff) and the quantity Aeff/R which determines the sensitivity for surface spectroscopy.
ultrathin section with diameter ~ A/2. For an intuitive understanding of the properties of the evanescent light field in this limit it is useful to define an effective area of the guided mode by Aes(X) = P(A)// s u r f(A) where P(X) is the total power travelling in the guide, /surf is the surface field strength.25 The area AeS is a universal function26 of R/X (right side of Fig. 7, R the ultrathin fibre radius). It shows a pronounced maximum of order A2 near _R/A=0.25. Thus all material particles in the immediate vicinity of the surface including adsorbed molecules or atoms bound in surface trapping states are subject to a light field resembling a very tightly focused laser beam. In contrast to free space focusing, however, this condition is extended from the short depth of focus (~ A) along the entire stretch of the ultrathin fibreL, resulting in an enhancement factor of order L/X ~ 103 — 104. We have carried out an initial experiment with organic molecules adsorbed on the surface of an ultrathin fibre and found that we obtain very high sensitivity due this "infinite focus" situation.25 Acknowledgments
We wish to thank the students that have worked on this project: Y. Miroshnychenko, I. Dotsenko, M. Khudaverdyan, S. Reick, T. Kampschulte, A. Stiebeiner, E. Vetsch, F. Warken, and our colleague M. Sokolowski. Support by the Deutsche Forschungsgemeinschaft and the European Commission is gratefully acknowledged. 3. References References
1. ARDA roadmap: http://qist.lanl.gov; http://qist.ect.it/Reports/reports.htm.
EU
QIPC
Strategic
Report:
184
2. I. Bloch, Nature Physics 1,23 (2005). 3. D. Schrader, S. Kuhr, W. Alt, M. Mueller, V. Gomer, D. Meschede, Appl. Phys. B 73 819 (2001). 4. Y. Miroshnychenko, W. Alt, I. Dotsenko, L. Forster, M. Khudaverdyan, D. Meschede, D. Schrader, A. Rauschenbeutel, Nature (London), 442,151(2006). 5. D. Meschede and A. Rauschenbeutel, in: Adv. A t. Mol. Phys. 53, 76 (2006), G. Rempe and M. 0. Scully, eds. 6. L. Forster, W. Alt, I. Dotsenko, M. Khudaverdyan, D. Meschede, Y. Miroshnychenko, S. Reick, A. Rauschenbeutel, New. J . Phys. 8,259 (2006). 7. Y.Miroshnychenko, W. Alt, I. Dotsenko, L. Forster, M. Khudaverdyan, D. Meschede, S. Reick, A. Rauschenbeutel, Phys. Rev. Lett. 97,243003 (2006). 8. Y.Miroshnychenko, W. Alt, I. Dotsenko, L. Forster, M. Khudaverdyan, A. Rauschenbeutel, D. Meschede, New J . Phys. 8,191 (2006). 9. R. Miller, T.E. Northup, K. M. Birnbaum, A. Boca, A. D. Boozer and H. J. Kimble, J. Phys. B: A t. Mol. Opt. Phys. 38,551, (2005). 10. S. Nuamann, M. Hijlkema, B. Weber, F. Rohde, G. Rempe, and A. Kuhn, Phys. Rev. Lett. 95, 173602 (2005). 11. K. M. Fortier, S. Y. Kim, M. J. Gibbons, P. Ahmadi, M. S. Chapman, arXiv:quant-ph/0703156vl;Phys. Rev. A , in print (2007). 12. Stefan Nuamann, Karim Murr, Markus Hijlkema, Bernhard Weber, Axel Kuhn, and Gerhard Rempe, Nature Physics 1, 122 (2005). 13. S. Haroche and J. M. Raimond, Exploring the Quantum (Oxford Univ. Press 2006). 14. D. I. Schuster et al., Nature 445,515 (2007). 15. A. D. Boozer, A. Boca, R. Miller, T. E. Northup, and H. J. Kimble, Phys. Rev. Lett. 98,193601 (2007). 16. T.Puppe, I. Schuster, A. Grothe, A. Kubanek, K. Murr, P. W. H. Pinkse, and G. Rempe, Phys. Rev. Lett. 99,013002 (2007). 17. T. Wilk, S. C. Webster, H. P. Specht, G. Rempe, and A. Kuhn, Phys. Rev. Lett. 98,063601 (2007). 18. Y . Colombe, T.Steinmetz, G. Dubois, F. Linke, D. Hunger and J. Reichel, arXiv:0706.1390~1 [quant-ph] 19. M. Trupke, J. Goldwin, B. Darqui, G. Dutier, S. Eriksson, J. Ashmore, and E.A. Hinds, Phys. Rev. Lett. 99,063601 (2007). 20. S. Gleyzes, S. Kuhr, C. Guerlin, J. Bernu, S. Delglise, U. Hoff, M. Brune, J. M. Raimond, S. Haroche, Nature 449,297 (2007). 21. P. Russell, Science, 299,358 (2003). 22. K. Vahala, Nature, 424,839, (2003). 23. T.A. Birks and Y. W. Li, J. Lightw. Tech., 10,432 (1992). 24. Y. Louyer, D. Meschede, A. Rauschenbeutel, Phys. Rev. A , Rapid Comm, 72,031801(R) (2005). 25. F. Warken, E.Vetsch, D. Meschede, M. Sokolowski, and A. Rauschenbeutel, Opt. Exp., in print, (2007). 26. F. Le Kien, J. Q. Liang, K. Hakuta, V. I. Balykin, Opt. Commun., 242,445 (2004).
SPECTROSCOPY ON THE
SMALL SCALE
This page intentionally left blank
WIDE-FIELD CARS-MICROSCOPY CHRISTOPH HEINRICH, ALEXANDER HOFER, STEFAN BERNET, MONIKA RITSCH-MARTE Division f o r Biomedical Physics, Innsbruck Medical University, Innsbruck, Austria * E-mad: monikarztsch-marteoi-med.ac.at We have developed a non-scanning version of Coherent anti-Stokes Raman Scattering (CARS) microscopy. In contrast to the more common confocal implementation, where one scans tightly focussed femto- or picosecond pulses over the sample, we use nanosecond pulses and a special excitation geometry that is designed to satisfy the phase matching condition over the whole field of view. This allows fast image acquisition, even snapshots with a single shot of nanosecond laser pulses are possible.
Keywords: CARS microscopy, vibrational imaging, Raman spectroscopy
1. Introduction Making use of nonlinear effects in optical microscopy requires intense electromagnetic waves, i.e. pulsed laser beams. Continuous technical advances in the development of laser sources (and highly sensitive detectors) have led to laser systems that are ideally suited for the task. Depending on the specific multiphoton setup, femtosecond (fs), picosecond (ps), or even nanosecond (ns) laser sources are used. Coherent anti-Stokes Raman scattering (CARS) is a third order nonlinear process that is suitable for nonlinear microscopy. CARS mixes two laser pulses at frequencies wp and ws which generate a blue shifted signal at frequency W A S = 2 wp - ws.This signal is resonantly enhanced, if the frequency difference wp - ws approaches the resonant frequency of molecules inside the medium. This effect was first demonstrated by Maker and Terhune in 1964 with discrete frequencies and became a practical tool in the 1970s when tunable dye laser for continuous scanning were developed. At that time CARS was mainly used as a spectroscopic tool, especially for the analysis of combustion processes. 187
188
The first ” proof-of-principle” CARS-microscope (CARSM) was set-up by Duncan et al. in 1982 The authors used dye lasers delivering ps-pulses in the visible region. This resulted in a rather strong nonresonant CARS background that almost overwhelmed the spectrally selective signal, Thus today nearly all CARSM setups utilize laser pulses in the near-infrared (NIR) region.
’.
2. Features of a CARS-microscope CARSM is a spectrally (chemically) selective method that allows functional mapping of the targeted substances in the unstained sample. Recently it has attracted much interest in the biomedical sciences, since the method enables imaging of cells components within living cells with three dimensional sectioning capability. This interest was triggered by the pioneering work of Xie and coworkers in 1999. Since then it has conquered a ”niche” in vibrational microscopy, as on the one hand CARS is more sensitive than spontaneous Raman scattering due to stimulated emission, and on the one hand it offers better resolution than IR microscopy, because the emitted signal is blue-shifted with respect to the excitation laser pulses. CARS imaging of living cells can be done at video frame rate and even faster ’. On the ”negative side” CARS microscopy has to deal with an inherent nonresonant, i.e. spectrally independent, background, which negatively influences the signal-to-noise ratio. Thus it is crucial to adapt the microscopic CARS setup in a way to suppress this nonresonant background. Many approaches have been developed to this end. For instance, detecting the CARS signal in the backward direction (epi-CARS) delivers high contrast for scatterers smaller than the wavelength of light An ensemble of dipoles lined up in a plane orthogonal to the propagation of the excitation fields emit a symmetrical radiation pattern in axial direction. Coherent addition of the radiation fields from an ensemble of dipoles lined up in axial direction, however, generates a large signal in forward and a weak signal in backward direction, due to constructive and destructive interference. Thus detecting the CARS signal in the backward direction avoids the large forward-going signal from bulk water, but at the cost of a lower signal. Another possibility is to utilize the polarization difference between the resonant and nonresonant CARS fields Adjusting the analyzer in front of the detector orthogonally with respect to the nonresonant polarization, the background can be suppressed efficiently. Unfortunately, the absolute strength of the Raman resonant CARS signal is likewise reduced, which limits the sensitivity for the detection of weak Ranian resonances. This
‘.
’.
189
dilemma can be circumvented by the interference of a weak CARS signal with an intense nonresonant field acting as a local oscillator, an approach that has been named heterodyne-CARS '. Other approaches utilizing interferometry ', modulation techniques with lock-in amplifiers or timedomain techniques by means of pulsed-sequenced detection have been demonstrated. It is also possible to combine epi-CARS with polarization sensitive detection and correlation spectroscopy to monitor dynamic processes within biological samples
'',
''
'.
3. WIDE-FIELD CARS MICROSCOPY
A standard CARS microscope utilizes a confocal setup working with tightly focussed beams of NIR laser pulses and with oil-immersion objectives of high numerical aperture (NA = 1.4). The confocal setup has the advantage of high spatial resolution. The tight focussing in space leads to an increased spread in momentum space and thus "relaxes" the phase matching condition. Image formation is carried out by scanning the sample (or the laser) which, however, limits the acquisition time. Wide-field CARS microscopy avoids scanning taking an image the whole sample "all at once". The schematic setup of our wide-field CARS setup13 is sketched in Fig.1. The idea is to use a non-collinear beam geometry, distributing the intensity of ns pulse trains homogeneously over the whole sample region of interest. We actually use a markedly non-collinear beam geometry ("extremely folded box-CARS" ), which utilizes an ultra-dark field condenser to deliver a cone of light from above, while the Stokes beam is coupled through the objective of the inverted microscope from below. This gives rise to an anti-Stokes beam that counter-propagates with respect to the Stokes beam, which can conveniently be read out through the microscope objective. Note that the high numerical aperture of the dark-field condenser provides a narrow sheet of light illumination which very effectively suppresses the nonresonant CARS background. Based on the axial confinement of the nonlinear interaction zone and the intrinsic nonlinear intensity dependence, this system offers good optical sectioning capability. We have also investigated using a structured illumination technique based on laser speckles, Dynamic Speckle Illumination (DSI), which was recently invented by J. Mertz l4 and coworkers, to improve the sectioning capability even further. Our laser system consists of a frequency doubled and tripled diodepumped Nd:YAG laser and and an optical parametric oscillator (GWU Lasertechnik). The 1064 nm line of the laser system is currently used as the
190
Fig. 1. Schematic setup of the wide-field CARS-microscope: The laser pulses enter the microscope from different directions. The pump pulse is coupled in from above through a dark-field condenser whereas the Stokes pulse comes from the backside of the inverted microscope and travels through the objective. As a consequence of the phase matching condition the generated anti-Stokes signal counter-propagates with respect to the Stokes beam and is imaged by the same objective on an intensified CCD-camera. Multimodefibers are used for beam delivery in order to destroy the spatial coherence.
Stokes beam, whereas the 355 nm line drives the optical parametric oscillator, which emits ns-pulses that are continuously tunable between 410 nm and 2600 nm and are used as the pump beam. The sample is typically immersed in a liquid (e.g. water) and sandwiched between two glass coverslips.
(B)
(C)
(D)
Fig. 2. Spectrally selective imaging of a mixture of polymer particles. (A) presents a reference dark-field image of the test sample. (B) and (C) show the acquired CARS images obtained at 3052cm"1 (yellow) and 2953 cm^ 1 (red), respectively. (D) displays an overlays of the images from (B) and (C).
The performance of the wide-field CARS-microscope is demonstrated on test samples consisting of two kinds of polymer beads of 1.5/mi size. One
191
species is polystyrene (PS) and the other polymethylmethacrylat (PMMA). Both substances exhibit strong Raman resonances that are close to each other, namely PMMA at 2953cm-1 and PS at 3052cm-1. Fig.2 demonstrate the high spatial resolution as well as the spectral selectivity of our CARS setup. Fig.2 (A) shows a reference dark-field image of the test sample clearly indicating the position of the various polymer beads. Besides in Fig.2 (B) one can see the detected CARS signal when the difference of the laser pulses is tuned to the Raman resonance of PS at 3052 cm"1. Fig.2 (C) present the CARS signal obtained by changing the beating frequency of the laser pulses to 2953cm-1. Fig.2 (D) overlays both CARS signals encoded with different colors. A comparison of the overlay with the reference darkfield image shows that the brighter beads in the dark-field image perfectly match with the beads that light up when tuning the excitation laser frequencies to 3052cm-1 indicating PS beads, as it should be, since the higher refractive index of PS (nps = 1.55) compared to PMMA (UPMMA = 1.49) results in a stronger signal under dark-field illumination.
Oark-ffeW iHuminalon 1 CARS fi> 5850cm"1
Fig. 3. (color) CARS image of an example of a living alveolar type II lung cell. Left: Dark-field illumination. Middle: Detected CARS signal at 2850 cm^ 1 exciting aliphatic CH2 stretching vibration. Right: Overlay of both images.
To demonstrate the non-invasiveness of the method we applied it to visualize phospholipid-rich vesicles inside living alveolar type II cells from rats. These cells are specialized to generate specialized vesicles, the so-called "lamellar bodies", that contain pulmonary surfactant. These vesicles are released from the cells by exocytosis to provide the monomolecular lipid film of dipalmitoyl phosphatidylcholine (DPPC) on the surface of lung alveoli that is required to lower the surface tension to the level necessary for optimal gas exchange. Lamellar bodies contain various lipid and protein components of varying composition, according to the special requirements of the organ. The lamellar bodies of the lung epithelium typically have a diameter range of 1 - 2,5 //m. Fig.3 is an example for imaging these cell
192
components by means of wide-field CARS microscopy. The left image in Fig.3 presents an example of a lung cell under dark-field illumination. Next one can see the CARS signal when tuned to the aliphatic CH2 stretching vibration at 2850cm-l generating a strong signal for all lipids and phospholipids. The five white circles in the dark-field image now light up bright and show that these structures contain DPPC. An overlay of both images is shown on the right. Recently Toytman et al. l5 have presented another type of wide-field microscopy where the excitation geometry of the laser beams does not satisfy the phase-matching condition. Thus, in a homogenous medium such as the solvent alone, no CARS signal is generated effectively. In this approach, somewhat reminiscent of the idea of dark-field microscopy, only light that is scattered from structures in the sample to satisfy the phase matching can contribute to the CARS signal. This avoids the nonlinear background from the bulk, but a disadvantage seems to be the fact that one has to rely on "accidentally" satisfying phase matching in regions of the sample, which might give rise to only weak CARS signals, or might make the interpretation of images of unknown structure and orientation difficult. Nevertheless the authors have successfully implemented their scheme and given proofof-principle. 4. APPLICATIONS A N D OUTLINE
Since lipids give particularly strong CARS signals, it is very likely that CARSM will become an established tool in lipid metabolism research, which is a very "hot" topic at present. Many groups are currently setting up CARS experiments with living cells, e.g. trying to visualize changes in the location, distribution, or concentration of fatty acids. It has to be stressed, however, that CARSM is not a single-molecule method. Currently the sensitivity typically is in the order of lo6 vibrating modes per focal volume, which allows the imaging of single lipid bilayers l6 and single cellular membranes, but not the monitoring of single biomolecules and their dynamics. The fact that CARSM is a labelling-free method could be a strong advantage in an area of biomedical research that is presently growing very rapidly: There is growing concern about possible unknown risks of nanoparticles, which are increasingly used in the emerging "nanomedicine", e.g. as contrast agents or for drug delivery. To date there often exists only limited or unsatisfactory information on the "fate" of the nanoparticles in the human body. Being an optical method, diagnostics deep within the body is impossible - but CARSM might be able to assist cell or tissue studies on
193
the up-take or clearance the nanoparticles and their degradation products, for instance, whether they aggregate or where they accumulate. The fact that there is no need for staining or any other treatment of the particles which could possibly change their behavior in the tissue - can give CARSM a special advantage here. This work was supported by the Austrian Science Fund FWF (Project P16658-N02). References 1. P. Maker, R. Terhune. ”Study of optical effects due to an induced polarization third order in the electric field strength.” Phys. Rev. A 137, 801 (1965). 2. M. Duncan, J . Reintjes, and T. Manuccia. ”Scanning coherent anti-Stokes Raman microscope.” Opt. Lett. 7,350 (1982). 3. A. Zumbusch, G.R. Holtom, X.S. Xie. ”Three-dimensional vibrational imaging by coherent anti-Stokes Raman scattering.” Phys. Rev. Lett. 82, 4142 (1999). 4. C.L. Evans, E. 0. Potma, M. Puoris’haag, D. Cbte, C.P. Lin, and X.S. Xie.”Chemical imaging of tissue in vivo with video-rate coherent anti-Stokes Raman scattering microscopy. Proceedings of the National Academy of Sciences 102, 16807 (2005). 5. C. Heinrich, S. Bernet, and M. Ritsch-Marte. ”Nanosecond microscopy with spectroscopic resolution.” New J. Phys. 8, 36 (2006). 6. A. Volkmer, J.X. Cheng, and X.S. Xie. ”Vibrational imaging with high sensitivity via epidetected coherent anti-Stokes Raman scattering microscopy.” Phys. Rev. Lett. 87, 023901 (2001). 7. J.X. Cheng, L.D. Book, and X.S. Xie. ”Polarization coherent anti-Stokes Raman scattering microscopy.” Opt. Lett. 26, 1341 (2001). 8. E.O. Potma, C.L. Evans, X.S. Xie. ”Heterodyne coherent anti-Stokes Raman scattering (CARS) imaging.” Opt. Lett. 31, 241 (2006). 9. C.L. Evans, E.O. Potma, and X.S. Xie. ”Coherent anti-Stokes Raman scattering spectral interferometry: determination of the real and imaginary components of nonlinear susceptibility chi(3) for vibrational microscopy.” Opt. Lett. 29, 2923 (2004). 10. F. Ganikhanov , C.L. Evans, B.G. Saar, X.S. Xie. ”High-sensitivity vibrational imaging with frequency modulation coherent anti-Stokes Raman scattering (FM CARS) microscopy. Opt. Lett. 31, 1872 (2006). 11. A. Volkmer, L.D. Book, and X.S. Xie. ”Time-resolved coherent anti-Stokes Raman scattering microscopy: Imaging based on Raman free indiction decay.” Appl. Phys. Lett. 80, 1505 (2002). 12. T. Hellerer, A. Schiller, G. Jung, and A. Zumbusch. ”Coherent anti-Stokes Raman scattering (CARS) correlation spectroscopy.” 3, 630 (2002). 13. C. Heinrich, S. Bernet, and M. Ritsch-Marte. ”Wide-field Ccherent antiStokes Raman scattering microscopy.” Appl. Phys. Lett. 84, 816 (2004). 14. C. Ventalon, J. Mertz. ”Quasi-confocal fluorescence sectioning with dynamic speckle illumination”, Opt. Lett. 30, 3350 (2005).
194
15. I. Toytman, K. Cohn, T. Smith, D. Simanovskii, and D. Palanker. ” Widefield coherent anti-Stokes Raman scattering microscopy with non-phasematching illumination.” Opt. Lett. 32, 1941 (2007). 16. E.O. Potma and X.S. Xie. ”Detection of single lipid bilayers with coherent anti-Stokes Raman scattering (CARS) microscopy.” Journal of Raman Spectroscopy 34,642 (2003).
ATOM NANO-OPTICS A N D NANO-LITHOGRAPHY V.I. Balykin, P.N. Melentiev, A.E. Afanasiev, S.N. Rudnev, A.P. Cherkun, and V.S. Letokhov Institute of Spectroscopy, Russian Academy of Sciences, Troitsk, Moscow region, 142190 Russia P.Yu. Apel, and V.A. Skuratov Flerov Laboratory of Nuclear Reactions, Joint Institute for Nuclear Research, Dubna, Moscow region, 141980 Russia V.V. Klimov Lebedev Physical Institute, Russian Academy of Sciences, Leninski pr. 53, Moscow, 11 9991 Russia
1. Introduction Semiconductor devices are currently constructed on a microscopic scale, predominately using optical lithography. The fabrication of structures at scales smaller than the current limits is a technological goal of great practical, but also fundamental interest: material structures with dimensions in the 10 nm range represent a bridge between the classical and the quantum mechanical world. In practice, it is desirable to have the ability to build any nanostructures with atomic precision using any atomic species. To date, no single approach meets this demand. Rather, there are a number of techniques, each of them possessing some advantages and having some drawbacks. In particular, there are known difficulties for further development of well-known techniques: Conventional optical lithography is diffraction limited down to 100 nm; charged particle beam lithography suffers from the serial nature of patterning and Coulomb repulsion; scanning probes, as they manipulate single atoms, are generally too slow; and self-assembled fabrication still requires a better understanding of the physical processes. Micro and nanofabrication of material structures is denoted generically as atom lithography. 195
196
We report here on the new approaches in atom lithography that is based on the use of (1) the spatially on nanometer scale localized laser fields; (2) the atom nanopencil; (3) the atom pinhole camera, and (4) the laser induced quantum atom adsorption on a surface.
2. Nanometer scale localized laser fields The ultimate goal in the field of atom nano-optics is the formation of nanometer-sized ensembles and beams of neutral atoms. It is known that laser fields capable of forming such atomic ensembles must be well localized. Standing light waves - the best known laser fields - are finding widespread use in atom optics. The first application of standing light waves in atom nano-optics was the localization of atoms in a nanometer-sized space and their placement in channels. The idea of channeling atoms in a standing light wave has led to the development of techniques of atom lithography that allow periodic one- and two-dimensional structures to be produced on a surface.' On the basis of general physical considerations, it is evident that the use of spatially localized fields (and, accordingly, spatially localized atomic potentials) can offer new possibilities for the construction of atomoptical elements. To date, we have knowledge of only two types of laser fields that are well enough localized in space to make this possible: surface (evanescent) light waves and light that arises in the vicinity of structures with a characteristic size of less than the wavelength. Evanescent waves find widespread use in atom optics for the reflection, localization and cooling of atoms. A drawback of evanescent waves is that they are localized exclusively on a 2D surface. The three dimensional localized light field is that which results after diffraction by an aperture the size of which is small in comparison to the wavelength of light. In this case, a local 3D maximum of field intensity is formed near the aperture. The magnitude of the maximum is governed mainly by the size of the aperture. A substantial drawback of the field structures considered above is the accompanying standing or traveling light wave: when atoms move in a standing light wave they can undergo spontaneous decay processes that, in many cases, undesirable. The light nanofield configurations that are free from the abovementioned shortcomings are the following.' Two plane conductive plates spaced a distance d of the order of or smaller than the wavelength of light apart form a plane waveguide for the laser radiation coupled into it from one side. If the electric field strength vector of the laser radiation is normal to the plane of the waveguide, the radiation can propagate through the
197
waveguide, no matter how thin it is. Now let two small coaxial apertures be made in the conductive screens that form the waveguide. If the diameters a of these apertures are smaller than the wavelength of the radiation coupled into the waveguide, little radiation escapes from the waveguide through the apertures, but the light field near them is strongly modified. Figure 1 shows the light field intensity distribution near the apertures inside and outside the waveguide when the thickness of the waveguide is equal to the radius of the apertures. As can be seen in Fig. 1, there is a light field intensity minimum in the direction normal to the plane of the waveguide. Such a light field configuration can be called as a photon hole. Its characteristic size is determined by the size of the apertures, the thickness of the waveguide and its volume V ~ cfd
Fig. 1. Electromagnetic-field intensity for a photon hole.
198
If two conductive screens with coaxial apertures are spaced a distance of d = A/2 apart and the electric field strength vector is parallel to the waveguide plane than the intensity distribution in such a light field is different than in the previous case, Fig. 2. As can be seen, the field drops off rapidly outside the waveguide in the direction normal to the waveguide and has its maximum at the center of the waveguide. This maximum is caused by the constructive interference of the fields scattered by the apertures. Such a light field configuration can be called as a photon dot. The maximum intensity at the photon dot is twice as high as the field intensity in the absence of apertures.
x/a
a
Fig. 2. Electromagnetic-field intensity for a photon hole.
The characteristic size of a photon dot or a photon hole is in the nanometer region, which allows for nanometer-sized atomic ensembles to be formed. Let us consider as an example two possible uses for photon dots and photon holes: the focusing and localization of atoms. When an atom is exposed to laser light, electric field of the laser light induces in the atom an oscillating dipole moment. If the light field amplitude at the atom position is spatially nonuniform, a gradient force on an atom develops. It is precisely the gradient force that is used to modify the trajectory of the atoms. When the detuning of the laser radiation frequency relative to the atomic transition
199
frequency is positive, an atom in the laser light configuration is drawn into the weak field region. In the case of a photon hole, the nanometer- sized weak field region is surrounded by the strong field inside the waveguide; if the light field frequency detuning is positive, the atoms that fly through the apertures in the waveguide walls will be attracted to the axis of the system-in other words, they will be focused. A photon dot draws atoms in negative frequency detuning and also provides focusing for the atomic beam that passes through the apertures in the waveguide walls. It has been ~ h o w nthat ~ > ~an atomic beam can be focused to a spot of the order of the de Broglie wavelength, which for a thermal beam amounts to a few Angstroms. The photon dot and photon hole light-field configurations have extreme points at which the gradient force is zero. Such light-field configurations are naturally considered as the possible atom trap c~nfigurations.~ The photon dot light-field configuration is stable and is truly three dimensional, while in a photon hole light-field configurations either the axial or the radial motion of the atom will be infinite since no matter what the sign of the frequency detuning. The extreme photon hole point is a saddle point. A number of schemes can be used to make the atomic motion in the photon hole region finite. One is based on the use of frequency detuning that varies in sign and over time, making it possible to localize atoms dynamically in a way similar to the localization of ions in high-frequency electromagnetic traps. 3. Atom nanopencil "Atom nanopencil" is an aperture in a thin material foil with a diameter ranging from 1 nm to 1 pm. As atomic beam passes through the aperture arbitrary patterns can be written by moving the aperture relative to a suitable substrate behind the foil. The aperture size defines the feature size of the written atom structures. A quantitative theoretical analysis shows that the van der Waals forces acting on the motion of atoms in the individual tapered channels of the aperture and the diffraction of the atomic wave to determine the final size of the deposited nano~tructures.~ To produce a set of appropriate microholes for use as atom nanopencils, nanosieves have been built and characterized by using the atomic force microscope. The nanosieve has multiple openings of equal diameter, randomly situated in a dielectric foil. We have nanosieves with openings ranging from 40 nm to 1000 nm. In our experiment, a mica or glass substrate was positioned at a distance around 10 pm behind the nanosieves. By moving the nanosieves across an atomic beam, we are able to create atom nanostruc-
200
tures on the various substrates. Fig. 3 shows an example of such nanostructures: we create chromium nanolines with heights in the range of 1-10 nm, widths of 180 nm and lengths of 1200 nm. The total amount of nanolines produced in one process exceeds 107 and occupy an area of 2 x 2 mm on the substrate surface.
Atomic Nanostructure (Cr) 2.5}itri
3 atomic layers
Fig. 3. Nanostructure of Cr atoms built with the atom nanopencil.
4. Atom pinhole camera
The most difficult problem in atom optics is the problem of high-resolution focusing of neutral atoms, which is promising for the nondestructive method for probing the surface at the atomic level, as well as for the creation of nanostructures on the surface. The main difficulty is the creation of the interaction potential of the atom with the electromagnetic field that is close to an "ideal" lens for atoms. We experimentally implement another approach to the problem of focusing and construction of an image in atom optics, which is based on a well known idea of "optical pinhole camera". The pinhole camera in optics is a camera without lens. Light forming an image passes through a small hole. In our experiment with the atom pinhole camera the atomic beam passes through a set of holes in a metal mask and thereby forms, by analogy with optics, a "glowing" atomic object of a given geometry. The atoms pass through the mask, propagate in vacuum along
201
rectilinear trajectories, similar to light rays, and are incident on a pinhole. The great many pinholes were created in a thin plastic film.6 As a thin film, we used a track membrane of an asymmetric structure.6 The initial material for the track membrane was a Hostaphan RES (Hoechst AG) polyethylene terephthalate film 5-10 /um thick. The film was irradiated by a 253-MeV accelerated krypton ion beam at the U-400 accelerator (Flerov Laboratory of Nuclear Reactions, Joint Institute for Nuclear Research). Then, the film was irradiated by ultraviolet radiation from one side; after that, chemical etching was performed. The thin film with a large number (n =107 — 10s cm~ 2 ) of nanoholes (50 nm diameter) is placed at a distance of 90 mm from the mask. Each hole of the film is a pinhole camera for atoms, which forms its individual image of the object on the substrate surface placed at a distance of I = 5 jum behind the film. Figgure 4(a) shows 2x2 p,m surface section with the image of the "cross" object. In addition to the almost completely formed cross images, the figure also exhibits structures with images of only its part. This occurs owing to the partial blocking of atoms forming the image of the cross of atoms and because the axes of various holes of the track membrane are nonparallel to each other.
Fig. 4. Nanostructures of Cr atoms on the glass surface that are obtained using the pinhole camera and the "cross" atomic object. Surface sections of sizes (a) 2x2 /nm and (b) 800 x 800 nm are shown. The nanostructures are measured by means of an atomic force microscope.
Figure 4(b) shows the detailed image of one of the crosses. As seen in the figure, the cross consists of partially resolved subnanostructures that are images of separate holes of the object mask. The width of the nanostructures is approximately equal to 70 nm, which well corresponds to the passage of
202
the atomic beam through the holes of pinhole cameras and is determined by the sum of the hole diameter d = 50 nm and the mask image diameter do = 0.5 mm.
5. Laser-Induced Quantum Adsorption of Atoms on a Surface The adsorption of an atom (molecule) on a surface is a natural process of the trapping of the atom (molecule) in a surface potential well. The probability of trapping of the particle on the surface is determined both by the electronic structures of the particle and surface and by the thermodynamics of the collision of particles with the surface. In addition to the general physical interest in the process of the adsorption of particles on the surface, the adsorption of particles underlies the modern industry of microand nanoelectronics based on the methods of molecular beam epitaxy or gas phase epitaxy. For this reason, the control of processes of the adsorption of particles on the surface are of both fundamental and large applied importance. We propose a new mechanism of loading atoms into the surface potential well (i.e., their adsorption on a surface) and demonstrate the implementation of this scheme for Rb atoms adsorbed on the surface of a YAG We also show the possibility of producing micro- and nanostructures of arbitrary shape that consist of atoms localized on the dielectric surface. The proposed mechanism of the loading of atoms into the surface trap is based on the energy-pooling effect, i.e., inelastic collision of two excited atoms followed by the transition of one of them to the ground state and the other one to the highly excited state.g The defect of the internal energy is compensated by the kinetic energy of the atoms. When the atomic collision occurs inside the surface potential well, an atom can be trapped in this potential well. Figure 5a shows the scheme of the low-lying levels of the Rb atom. The 5 2 P 3 p level of the Rb atom is populated due to the absorption of 780-nm laser light. The energy of the 5 0 level is close to the 5 P 5 P asymptotic energy of the Rb2 molecule and, therefore, can be populated owing to the energy-pooling collisions of excited atoms:
+
Rb(5P)
+ Rb(5P) + A E
4
Rb(50)
+ Rb(5S),
(1)
where A E = 93 K is the difference between the total kinetic energies before and after collision (energy defect). As a result of collision process (1) of two atoms, one atom passes to the ground state and the other atom passes to
203
(a)
.
7-
5
.sp+sp
--
m
0
-
i
- . ~ . ~2.0 . - 3.0. ~ . - . 0.5
1.0
1.5
2.5
Atom to surrace distance, nm
Fig. 5. (a) The energy level diagram for Rb atom. (b) The mechanism of loading of Rb atom in the surface trap.
the 5D excited state. The energy defect is compensated by the energy of the translational atomic motion; i.e., the kinetic energy of two atoms after collision event (1) decreases by 93 K. If the collision of two excited atoms occurs near the bottom of the surface potential well, the loss of the kinetic energy in the atomic collision can lead to the localization of an atom in the surface potential, i.e., to its laser induced adsorption, as schematically shown in Fig. 5b. The quantum adsorption of Rb atoms is experimentally implemented on the surface a YAG crystal. The temperature of the surface can be varied from room temperature to 240 'C. Laser radiation is tuned t o resonance with the Dz line of Rb atom and a laser beam passing through and perpendicular to the surface. The laser beam diameter is varied between 0.5 and 2 mm and the maximum power of laser radiation is equal to 70 mW. The energy-pooling process is identified by detecting the blue fluorescence of Rb atoms (A= 420.2 and 421.6 nm) from the laser beam.g Laser induced quantum adsorption opens the possibility of creating atomic micro- and nanostructures with a given geometry on the dielectric surface. Such a possibility is illustrated in Fig. 6 on the example of creation of the three letters P, R, and L on the surface. Atomic structure
204
a
Fig. 6. (a) The mask consists of 35 holes in a thin metal foil, (b) The microstructure of Rb atoms in the form of three letters P R L that are obtained by the laser induced quantum adsorption.
of the shape PRL was produced in a following way: (1) 35 openings in a metal foil were done to produce a mask for these letters; (2) the mask was illuminated by the laser light; (3) the image of the mask was projected onto the surface and during the time interval t = 5 min were produced 35 microtraps on the surface reproducing the letters P, R, and L of size about 50 j,m. 6. Acknowledgments This work was supported by the Russian Foundation for Basic Research (project nos. 06-02-16301-a, 06-08-01299-a, and 05-02-16370-a). References 1. J.J. McClelland et al., Science 262, 877 (1993). 2. V.I. Balykin, V.V. Klimov, V.S. Letokhov, Atom Nanooptics, In "Handbook of Theoretical and Computational Nanotechnology", eds., M. Rieth and W. Schommers, American Scientific Publishers (2006). 3. V.I. Balykin, V.V. Klimov and V.S. Letokhov, Atom Nano-Optics, Optics & Photonics News, 16, 33 (2005). 4. V. Balykin, V. Klimov, Letokhov, JETP Lett., 78, 8 (2003). 5. A.E. Afanasiev, P.N. Melentiev, and V.I. Balykin, JETP, (2007) (to be published). 6. V. I. Balykin, et all., JETP Lett., 84, 466 (2006). 7. A.E. Afanasiev, P.N. Melentiev, and V.I. Balykin, JETP Lett., 86, 198 (2007). 8. A.E. Afanasiev, P.N. Melentiev, and V.I. Balykin, Phys.Rev.Lett., 2007 (to submitted). 9. Z.J. Jabbour, et.al., Phys. Rev. A 54, 1372 (1996).
PINHEAD TOWNTALK, PUBLIC LECTUREAND MOUNTAINFILM
This page intentionally left blank
THE QUANTUM REVOLUTION TOWARDS A NEW GENERATION OF SUPERCOMPUTERS
R. BLATT Institut f u r Experimentalphysik, Universitat Innsbruck, Technakerstr. 25, A-6020 Innsbruck, Austria, and Institut fur Quantenoptik und Quanteninformation (IQOQI), Osterreichzsche Akademie der Wissenschaften, Otto-Hittmair-Platz 1, A-6020 Innsbruck, Austria Computers that operate with quantum processes promise unprecedented computational power for some algorithms, much more than could be obtained by classical machines. The implementation of such a quantum computer requires the precise control and manipulation of individual quantum systems, a task that can be achieved by quantum optical means and the use of precision laser spectroscopy.
Keywords: Quantum computing, entanglement, trapped ions
1. Introduction
Computing machines have always been a dream of mankind and many mechanical tools have been developed to allow for faster and more complicated calculations. While the tools of the 19th century were mostly of mechanical nature, the technical evolution of the 20th century allowed the construction of fast electronic switches and hence the development of powerful computing machines. This led eventually to the revolution of technology by the ubiquitous availability of personal computers and its corresponding use in all areas of our daily life. 2. Computers and technology
The technical advance in computer technology is well illustrated by Moore's law' that states that nowadays computing power, as indicated by the number of transistors in a processor or the available memory on a chip, doubles approximately every 24 months. Surprisingly enough, this empirical law 207
208 holds since more than 40 years and provides a guideline for the manufacturing industries. Also, the number of atoms that are required, for example, to store the information of one bit, decreases in a similar way.2 While the storage of a single bit required about 10’’ atoms in 1962, only lolo atoms were necessary in 1988 and this trends still holds. Assuming an ongoing development, a simple extrapolation reveals that around 2020 only a single atom would carry the information of one bit. Clearly, a t this stage the laws of quantum mechanics will govern the storing and retrieving of information; in practice, however, quantum mechanical effects will become important a long time before that. Therefore, it seems natural to investigate whether using quantum physics for information processing can actually be used to our advantage. Already in the 1980s Deutsch and Feynman have discussed information processing using quantum physic^,^,^ however, at that time this subject was more an academic exercise since it was neither known how quantum information processing could be particularly useful nor how it could be implemented. This situation changed in 1994 when Peter Shor came up with an algorithm5 to factor large numbers that required only polynomial efforts (in terms of the digits of the number in question) while a classical computer needs an exponential overhead to solve this problem. This application, which is of enormous impact for cryptology, and the later found fast data base search by Lov Grover‘ led to an intense search for physical systems that allow one to really implement quantum computing. Thus, the mid 1990s mark the start of the quest for a quantum computer that is still ongoing and inspiring wide fields in physics today. Meanwhile, quantum information processing has matured as a multidisciplinary area in physics and computer science that ranges nowadays from most fundamental theoretical concepts to implementations using technologies from laser physics and laser spectroscopy to solid state and condensed matter physics and the field is still growing and widening. 3. Quantum bits, registers and gate operations
The smallest unit of classical information is the bit, usually represented by a switch either in ”up”/”down” or ”0”/”1” position. More generally, the quantum system comprised of the two levels 10) and 11) can be written also as the superposition
I+)
= COlO)
+Clll)
(1)
209
and this two-level system is commonly known as a quantum bit, or short, qubit. While a classical bit is best visualized as a switch, the qubit must be described as an arrow pointing somewhere on the surface of a sphere as is indicated in Fig. 1. For a quantum register, a row of several qubits is
T
Fig. 1. Classical information (a,b) and quantum information (c-f). (c,d) show the system in state |0), |1) respectively, (e) describes a superposition of equal weights |co|2 = |ci|2 = l/-\/2 and (f) shows a general, usually unknown superposition.
formed and the corresponding quantum state of the entire register must be considered. Clearly, this state can be, and generally is, in a superposition of all the pertaining states of the individual qubits. This generalizes the superposition concept that we commonly encounter in individual quantum systems to the entire system that seemingly consists of separate two-level systems. However, that point of view is no longer valid since even a manipulation of any one of the qubits in a row will change the quantum state of the entire system. Hence, the quantum register must usually be considered as entangled. Superposition and entanglement are the two features that generally distinguish a classical computer from a quantum computer and their control and manipulation will eventually enable one to speed up computational processes. Analogous to a classical computer, a set of gate operations is required that allow one to formulate a calculation step by step. It has been shown in 1995 that two operations are sufficient to provide a universal set of quantum operations, i.e. that allow one to implement arbitrary computations.7 Aside from the single qubit operation, which implements an arbitrary positioning and rotation of the arrow (see Fig. 1) given on the sphere, it is necessary to allow for conditional operations equivalent to the classical XOR operation in computer science. With such operations, the state of a target qubit can
210
be changed depending on the state of a control qubit. The corresponding truth table is shown in Fig. 2 and looks exactly like the classical XOR operation, however, it needs to hold for all coherent superpositions. Thus,
|0) 1}
11)11} |1)|0)
Fig. 2. Truth table of a CNOTgate operation. The state of a target bit is flipped if and only if the control bit is in state " 1". The notation indicates that this must hold for all superpositions.
I*
the concept of a quantum computer can be visualized as shown in Fig. 3. Starting from an arbitrary input state
E
ne{0,l}"
cn\ni,.. .,
El
(2)
the computation works as a series of one and two-qubit operations according to the specific algorithm under consideration and the outcome is just another superposition f ( \ x ) } with an operation f that can be described by a unitary operator. To this point a quantum computation is completely reversible. The outcome of the calculation is then obtained by a measurement that projects the system on its eigenstates and yields classical information in terms of zeros and ones for the individual qubits. The realization concepts of a specific quantum computer and its implementation vary widely depending on the quantum system considered, the quantum operations that are performed and the measurements that are taken to obtain classical information. On the other hand, there are a few building blocks and requirements that can be generally defined and described and that are common to all quantum computers. The requirements for a system to be considered for the implementation of a quantum computer are currently known as the so-called DiVincenzo criteria.8 Irrespective of the specific system, a quantum computer clearly needs (1) storage sites for the quantum information, i.e. qubits that can be arranged to form a (scalable) quantum register, (2) the possibility to initialize the qubits to
211
Quantum Processor
Fig. 3. Scheme of a circuit model quantum computer illustrated with 3 qubits. An arbitrary input superposition state x) is processed with a series of one and two-qubit operations described by a unitary operation F(\x)). (0,if) denote single qubit rotations, where (6, if) may vary from pulse to pulse depending on the algorithm.
arbitrary states, (3) long unperturbed computation (coherence) times, (4) a universal set of gate operations to allow for universal computations and (5) a highly efficient measurement of the qubit states to read out the result of the computation. In order to achieve large scale quantum computation, (6) the system under consideration should allow for a conversion between the stored, i.e. static and flying qubits that (7) can be faithfully transmitted between two quantum computer nodes. During the last decade a large number of systems has been and still is investigated for its suitability to implement a quantum computer.9 In particular, quantum optical systems of atoms and ions in traps, using the tools of laser spectroscopy, are among the most promising candidates for such a device. 4. Quantum computer with trapped ions
One of the first and seminal schemes to implement a scalable quantum computer was proposed by I. Cirac and P. Zoller in 1995.10 They considered a string of laser cooled trapped ions in a Paul trap as quantum register and formulated how a CNOT-gate operation can be realized with laser pulses that individually address the ions. The crucial idea was that the harmonic motion of ions in the trap can be used as a quantum bus. That allows one
212 to map the excited state of the controlling qubit to the motion and thus enables state manipulation of the target qubit conditioned on the motion and thus the controlling qubit. However, this requires that the ion string is optically cooled to the ground state of the harmonic oscillator which can be achieved in ion traps by sideband cooling.11 Moreover, the state of laser cooled ions in traps can be detected with nearly 100% efficiency using the "shelved electron technique", where the absorption of a single photon results in a lack of a huge number of fluorescence photons on a monitoring transition.11 As is well known from precision spectroscopy, trapped ions also offer extremely long coherence times12 and therefore provide an ideal system for the implementation of a quantum computer. Fig. 4 shows schematically
70 pm Fig. 4. Sketch of a linear ion trap holding a string of laser cooled ions that can be individually addressed for quantum state manipulation. With a CCD camera the fluorescence light is detected and allows for efficient state detection. how a string of laser cooled ions is manipulated with focussed laser beams. State detection is achieved by observing resonance fluorescence with a CCD camera. With such setups, single qubit operations were performed, and CNOT-gate operations have been implemented with trapped ions13"16 using the Cirac-Zoller idea as well as other proposals based on geometric phase changes.17 5. Simple quantum computations With a small trapped ion quantum computer, both the Boulder13 and the Innsbruck14 groups have demonstrated simple quantum algorithms. As a
213
basic building block, the CNOT gate operation (or an equivalent phase gate operation) was used to generate Bell states at the push of a These subsequently could be used as a resource for a demonstration of the teleportation of an atomic ~ t a t e . Non-classical ’ ~ ~ ~ ~ entangled states of three particles, in particular the GHZ and W states were deterministically and analyzed using quantum state t~mography.’~ A quantum Fourier transform was implementedz4 and error correction was demonstrated by the Boulder Multi-partite entanglement was achieved and analyzed for a 6-ion GHZ statez3 and an 8-ion W state,z6 the latter demonstrating for the first time a quantum byte. This experiment for the first time also manifested the power of quantum computation: While the creation of an eight-ion entangled state takes just about 1 ms to achieve, its analysis via quantum state tomography required a lot more efforts. In order to obtain all density matrix elements (i.e. 256x256 entries) more than 6500 different qubit rotations had to be applied that took more than 10 hours of uninterrupted running time of the quantum computer. Moreover, the data analysis on a classical computer required a raw computing time of several days on a computer cluster. This clearly demonstrates that even with relatively few qubits highly complex states can be created. 6. Future developments
With the ion trap quantum computer quantum information processing clearly has become reality, albeit on a small scale yet. On the other hand, there exist already architectures to scale such a system up.28 Of course, this still requires enormous efforts in physics and technology, especially in order to meet the requirements for error correction. While this is technically involved, there is however not a real roadblock in sight at this time. Therefore, current efforts are dedicated especially towards the development of yet smaller, so-called segmented ion traps that allow one to move the ions and thus the information around.2g With such chip traps available, there is good hope for an even broader application of quantum information processing. In particular, the use of entanglement for precision ~ p e c t r o s c o p y seems ~~>~~ a very promising avenue for further enhanced measurements and sensor technology.
7. Conclusion While the ion trap quantum computer seems farthest advanced at this time, it is quite foreseeable that other quantum optics and laser spectroscopy
214
technologies, e.g. based on atoms in lattices, quantum dots, or superconducting qubits will become very strong alternatives soon. The steadily growing field of quantum information has enormously profitted from the fields of laser spectroscopy, laser cooling and precision measurements. On the other hand, quantum information processing provides us with new tools that allow us to further enhance our understanding of fundamental physics as well as to develop future devices. References 1. G. Moore, Electronics 38, Nr. 8, April 19, (1965); see also http://www.intel.com/technology/mooreslaw/. 2. R. W. Keyes, IBM J. R&D. V32, N1, Jan 1988, p.26 3. D. Deutsch, Proc. R. Soc. London A 400, 97 (1985). 4. R. Feynman, Int. J. Theor. Phys. 21, 467 (1982). 5. P. W. Shor, Proceedings of the 35th IEEE Symposium on Foundations of Computer Science, pages 124-134. IEEE, 1994. 6. L. Grover, in Proceedings of the 28th Annual ACM Symposium on the Theory of Computation (ACM Press, New York, 1996), pp. 212219. 7. A. Barenco, C.H. Bennett, R. Cleve, D. P. DiVincenzo, N. Margolus, T. Sleator, J . A. Smolin, and H. Weinfurter, Phys. Rev. 52, 3457 (1995). 8. D. DiVincenzo, Quant. Inf. Comp. 1 (Special), 1 (2001). 9. for the US roadmap see http://qist.lanl.gov/ and for the European roadmap see http://qist.ect.it/. 10. I. Cirac and P. Zoller, Phys. Rev. Lett. 74, 4091 (1995). 11. D. Leibfried, R. Blatt, C. Monroe, and D. Wineland, Rev. Mod. Phys. 75, 281 (2003) and references therein. 12. D. J. Wineland, C. Monroe, W. M. Itano, D. Leibfried, B. E. King, and D. M. Meekhof, J. Res. Natl. Inst. Stand. Technol. 103, 259 (1998) and references therein. 13. D. Leibfried, B. DeMarco, V. Meyer, D. Lucas, M. Barrett, J. Britton, W. M. Itano, B. Jelenkovic, C. Langer, T. Rosenband, and D. J. Wineland, Nature 422, 408-411 (2003). 14. F. Schmidt-Kaler, H. Haffner, M. Riebe, S. Gulde, G. P. T. Lancaster, T. Deuschle, C. Becher, C. F. Roos, J. Eschner and R. Blatt, Nature 422, 412415 (2003). 15. J. P. Home, M. J. McDonnell, D. M. Lucas, G. Imreh, B. C. Keitch, D. J . Szwer, N. R. Thomas, S. C. Webster, D. N. Stacey and A. M. Steane, New Journal of Physics 8, 188 (2006). 16. P. C. Haljan, P. J . Lee, K-A. Brickman, M. Acton, L. Deslauriers, and C. Monroe, Phys. Rev. A72, 062316 (2005). 17. K. M ~ l m e rand A. S~rensen,Phys. Rev. Lett. 82, 1835 (1999); G. J. Milburn, S. Schneider, and D. F. V. James, Fortschr. Phys. 48, 801 (2000). 18. Q. A. Turchette, C. S. Wood, B. E. King, C. J. Myatt, D. Leibfried, W. M. Itano, C. Monroe, and D. J. Wineland; Phys. Rev. Lett. 81, 3631 (1998).
215 19. M. Riebe, H. Haffner, C. F. Roos, W. Hansel, J. Benhelm, G. P. T. Lancaster, T. W. Korber, C. Becher, F. Schmidt-Kaler, D. F. V. James, R. Blatt, Nature 429, 734 (2004). 20. M. D. Barrett, J. Chiaverini, T. Schaetz, J. Britton, W. M. Itano, J. D. Jost, E. Knill, C. Langer, D. Leibfried, R. Ozeri and D. J. Wineland, Nature 429, 737 (2004). 21. C. F. Roos, M. Riebe, H. Haffner, W. Hansel, J. Benhelm, G. P. T. Lancaster, C. Becher, F. Schmidt-Kaler, R. Blatt, Science 304, 1478 (2004). 22. D. Leibfried, M. D. Barrett, T. Schaetz, J. Britton, J . Chiaverini, W. M. Itano, J. D. Jost, C. Langer, and D. J . Wineland, Science 304, 1476 (2004). 23. D. Leibfried, E. Knill, S. Seidelin, J . Britton, R. B. Blakestad, J . Chiaverini, D. B. Hume, W. M. Itano, J . D. Jost, C. Langer, R. Ozeri, R. Reichle and D. J . Wineland, Nature 438, 639-642 (2005). 24. J. Chiaverini, J. Britton, D. Leibfried, E. Knill, M. D. Barrett, R. B. Blakestad, W. M. Itano, J . D. Jost, C. Langer, R. Ozeri, T. Schaetz,ddagger D. J. Wineland, Science 308, 997-1000 (2005). 25. J . Chiaverini, D. Leibfried, T. Schaetz, M. D. Barrett, R. B. Blakestad, J. Britton, W. M. Itano, J. D. Jost, E. Knill, C. Langer, R. Ozeri, D. J. Wineland, Nature 432, 602-605 (2004). 26. H. Haffner, W. Hansel, C. F. Roos, J. Benhelm, D. Chek-al-kar, M. Chwalla, T. Korber, U. D. Rapol, M. Riebe, P. 0. Schmidt, C. Becher, 0. Guhne, W. Dur and R. Blatt, Nature 438, 643-646 (2005). 27. C. F. Roos, G. P. T. Lancaster, M. Riebe, H. Haffner, W. Hansel, S. Gulde, C. Becher, J. Eschner, F. Schmidt-Kaler, R. Blatt, Phys. Rev. Lett. 92,220402 (2004). 28. T. S. Metodi, D. D. Thaker, A. W. Cross, F. T. Chong, I. L. Chuang, arXiv:quant-ph/0509051vl. 29. D. Kielpinski, C. Monroe, D. J . Wineland, Nature 417, 709-711 (2002). 30. P. 0. Schmidt, T. Rosenband, C. Langer, W. M. Itano, J. C. Bergquist, and D. J . Wineland, Science 309, 749-752 (2005). 31. C. F. Roos, M. Chwalla, K. Kim, M. Riebe, R. Blatt, Nature 443, 316-319 (2006).
This page intentionally left blank
COLD ATOMSAND MOLECULESI
This page intentionally left blank
ULTRACOLD & ULTRAFAST: MAKING AND MANIPULATING ULTRACOLD MOLECULES WITH TIME-DEPENDENT LASER FIELDS C. P. KOCH
Institut fur Theoretische Physik, Freie Universitat Berlin, Arnimallee 14, 14195 Berlin, Germany E-mail: christiane.koch@physik. fu- berlin. de
R. KOSLOFF The Fritz Haber Research Center, T h e Hebrew University of Jerusalem, Jerusalem 91904, Israel
E. LUC-KOENIG and F. MASNOU-SEEUWS Laboratoire Aim6 Cotton, C N R S and Universite' Paris Sud, B i t i m e n t 505, 91405 Orsay, France
R. MOSZYNSKI Department of Chemistry, University of Warsaw, Pasteura 1, 02-093 Warsaw, Poland While ultracold matter brought quantum effects onto the macroscopic scale, ultrafast lasers made quantum dynamical phenomena observable in real-time. Bringing the two together seems natural and holds the promise of employing quantum interferences in an unprecedented way. Photoassociation provides an optimal framework for the merger since in principle it relies only on the presence of optical transitions. Combining it with coherent control where the potential energy surfaces governing the dynamics can be 'shaped', a general route toward stable ultracold molecules is obtained.
1. Introduction
The formation of ultracold (T 5 100 pK) molecules along with the creation of quantum degeneracy in molecular gases offers exciting perspectives for investigating ultracold chemistry, implementing new schemes for quantum computing, building a molecule laser, and performing high precision measurements.' Moreover, since at ultralow temperatures only a single or very few quantum states contribute to the dynamics of a process, the ultracold 219
220 regime holds the promise of achieving coherent control in an unprecedented way. Coherent control relies on the superposition principle of quantum mechanics, and in particular on the ability to manipulate the relative phases of different quantum pathways. In a theoretical approach one can formulate control as an inversion problem, asking for the 'potential' (containing the potential energy surfaces as well as the interaction of the molecule with an external field) which generates the desired dynamics. For state-tostate transfer, i.e. when starting from a given initial state with the goal of reaching a certain target state, efficiencies close to 100% can be obtained. This was demonstrated for the example of stabilizing molecules in a highly excited vibrational state just below the dissociation limit to the vibrational ground state.2 The experimental approach is based on feedback loops where a genetic algorithm steers a pulse shaping device and the outcome of each measurement is used to create the next generation of pulses. At high temperatures prevailing in experiments to date, averaging over many thermally populated states is implied such that quantum interference effects are washed out or hidden. Thermal averaging can be avoided by going to very low temperatures, and it thus seems natural to combine the techniques of coherent control with ultracold gases. However, first experiments aimed a t photoassociating ultracold atoms to form molecules with femtosecond laser pulse^^,^ had to face a number of difficulties due to the high peak intensities of short lasers, interference with the trapping light, and the slow timescales of ultracold collisions. Nevertheless, coherent photoassociation could provide a means to finally realize control of a binary reaction. In the following, a brief overview over the theory of coherent photoassociation will be given. The concept of the pump-dump scheme is introduced in section 2, followed by examining in section 3 which potentials are required to make this scheme work. Section 4 will present a new proposal to engineer favorable potentials by an external field in order to ensure a successful route toward ground state molecules, and section 5 concludes. 2. The concept of pump-dump photoassociation
In photoassociation, two colliding ground state atoms are transferred to an electronically excited state due to the interaction with a laser which is red-detuned with respect to the atomic resonance f r e q ~ e n c y .Depending ~)~ on the shape of the excited state potential, molecules in the electronic ground state can be obtained by spontaneous emission or by applying a
221
11900 11800 11700 3
i3 h
f W
100 0
-100 -200
20
40 60 80 100 internuclear distance [ Bohr radii ]
120
Fig. 1. Pump-dump photoassociation, shown here for Cs2: A first pulse (pump) excites part of the initial ground state population, creating a wavepacket in the excited state. A second pulse (dump), delayed in time and blue detuned with respect to the first pulse, transfers part of the excited state wavepacket into bound levels of the ground state.
second laser field, blue-detuned with respect to the first one. In order to render this molecule formation scheme coherent, two requirements need to be met: (i) the timescale of the process should be shorter than that of spontaneous emission, and (ii) the time-reversal symmetry needs to be broken in order to avoid back-stimulation. Both conditions can be fulfilled by applying picosecond pulses, possibly chirped, in a pump-dump sequence, cf. figure This concept was first applied to the formation of Cs2 and R b ~ , ~ 1two ' systems that had extensively been studied in experiments using continuous wave lasers. The effect of the pump pulse shall be examined first.g Due to the finite bandwidth of the pump pulse, the resonance condition is fulfilled for a range of internuclear distances around the Condon radius, giving rise to the concept of the photoassociation The pulse intensity can be chosen such as to deplete the initial population within the photoassociation window, creating a hole in the ground state wavefunction. This is depicted in the lower panel of figure 2. The corresponding pulse energies are fairly low, amounting to a few nano-Joule.' The pump pulse creates a wavepacket in the excited state whose vibrational distribution is shown in the upper panel of figure 2. The width of this distribution is given by the spectral bandwidth of the pulse (inversely
222 8x10
•a
a! 4xlOH 'o OH
10 15 20 binding energy [ cm ]
25
20 40 60 80 100 internuclear distance R [ Bohr radii 1 Fig. 2. Applying a photoassociation pump results in a hole in the ground state distribution (bottom) and an excited state wavepacket, represented by its vibrational distribution (top), shown here for Rb2.
proportional to the duration of a transform-limited pulse rp as used in figure 2). The detuning Ap of the pulse determines the peak position of the distribution.8 The small feature close to zero detuning should be noted: It appears despite the spectral amplitude of the laser field being almost zero, and it is due to the huge free-bound Franck-Condon factors which increase by several orders of magnitude as the detuning, or, respectively, binding energy, becomes smaller. Thus it is imperative to employ red-detuned narrowbandwidth pulses8 or, alternatively, to remove spectral components exciting the atomic resonance from a broad-bandwidth pulse.3'4 Once transferred to the excited state, the wavepacket starts to evolve under the influence of the excited state potential, cf. figure 1. The oscillation period for binding energies of one to a few wavenumbers is on the order of 200 ps for the heavy alkali dimers Rb2 and Cs2- The time delay between the pump and dump pulses should be chosen such that the overlap between the time-dependent wavepacket and the bound levels of the ground electronic state becomes maximal. This is the case when the wavepacket reaches the inner turning point, cf. figure 1. Two different regimes for the dump pulse are then conceivable:7 a narrow-bandwidth pulse populates a single ground state level, while a broad-bandwidth pulse achieves maximum electronic ground state population which is distributed over several vibrational levels. In both cases, the dump pulses are short compared to the vibrational period of the excited state potential, and the dynamics can be understood in terms of an instan-
223 taneous two-level s y ~ t e m . ~ Since, > ~ > lin~ analogy to the pump pulse, weak fields and no 'shaping' of the dump pulse beyond a linear chirp is assumed, the transition probabilities are determined solely in terms of spectral amplitudes and Franck-Condon factors. Therefore, how much of the excited state wavepacket can actually be transferred to molecular levels of the ground state, depends crucially on the shape of the excited state potential.
3. The role of the excited state potentials The pump-dump scheme involves an excitation and a stabilization step, and the overall efficiency is determined by those of both steps combined. Figure 3 illustrates possible underlying physical mechanisms: Efficient excitation is afforded by the long-range attractive -1/R3 behavior of the excited state potential (upper panel, shown here is the Cs2 O;(p3/2) state), yielding large free-bound Franck-Condon factors. The -1/R3 behavior is caused by the resonant dipole interaction between two identical atoms where one is in its electronic ground state, while the other one is in the first excited state. Hence this behavior is observed in almost all excited state potentials of homonuclear dimers correlating to the n s n p asymptote (with the exception of cases where the 1/R3-term is accidentally cancelled such as in
+
O,(Pl/2)).
--
11900
5
y
11800
Y
8 11700 20
40 60 80 I00 internuclear distance Ia,u -1
120
~
--
2mo 1 1900
5
' & 16000 I
-,llS00 >
p 12000
D 8 11700
B
so00
1 100
20
40 60 80 100 internuclear distance [ a. ]
120
20
40 60 80 internuclear distance [ a, ]
100
Fig. 3. Excited state potentials with -l/R3 long-range behavior provide an efficient mechanism for the pump step (upper panel), while the dump step (lower panels) can be facilitated by a softly repulsive +l/R3 potential wall at intermediate distances (left) or by resonantly coupled excited state potentials (right).
224
x10
Timefps] 600_^g Energy f cm ]
[cm l]
Fig. 4. Overlap of the excited state wavepacket with ground state levels \{
In general, stabilization is difficult. Two mechanisms which lead to efficient stabilization are shown in the lower panel of figure 3. A repulsive + l/ J R 3 -term at intermediate range, shown for the double-well 0~(j»3/2) potential of Cs2 slows the wavepacket down, leading to a piling up of probability amplitude at the inner turning point of the long-range well.7 Due to the softly repulsive nature of the potential wall, the wavepacket stays close to this turning point for a few picoseconds such that it can easily be caught by the dump pulse. The wavepacket dispersion can be minimized, i.e. the wavepacket can be focussed at the inner turning point, by applying an appropriately chosen negative chirp to the pump pulse.7'9'10 About 20% of the overall excited state population can be transferred to bound levels of the ground state potential, the remaining part, if addressed by the dump pulse, is dissociated.7 The dissociation can be attributed to the high-momentum components which the wavepacket acquired while propagating on the 1/-R3 potential and which cannot be projected onto bound levels of the ground state with its l/R6 long-range behavior. Two excited state potentials which are coupled resonantly by e.g. spinorbit interaction provide a second mechanism for stabilization to bound levels by the dump pulse (lower right panel of figure 3).8'11 Some of the vibrational wavefunctions of the coupled potentials show then a doublepotential structure with four turning points. While the outermost turning point implies good free-bound Franck-Condon factors for the pump pulse, the next turning point at intermediate distances leads to greatly improved bound-bound Franck-Condon factors for the dump pulse.12 If none of the two stabilization mechanisms is present which is true for most excited state
225
potentials, it is not possible to form ground state molecules by the dump pulse except in the very last vibrational levels. This is demonstrated in figure 4 which compares the time-dependent overlap of the excited state wavepacket with bound ground state levels for the resonantly coupled 0+ states and the generic lg states of Rb2 (the binding energy of the excited state wavepacket is identical in both cases). 4. Engineering favourable potentials
The potentials which govern the dynamics can be 'shaped' by applying laser fields to a molecular system, thus steering the dynamics in the desired direction. This is the fundamental message to be learnt from coherent control. However, if this concept is applied straightforwardly to two generic potentials coupled by a laser field, all work needs to be done by the field. In an example where molecules in levels close to the dissociation limit of the electronic ground state potential shall be transferred to their vibronic ground state, very high pulse energies and many Raman steps are then required.2 The high intensity of a short pulse can be avoided by employing a train of pulses over a longer time leading to coherent population accumulation.13 The number of necessary Raman steps can be minimized by joining the concepts of resonant coupling and of shaping the potentials, i.e. by inducing resonant coupling by an external field. This concept is visualized in figure 5 with potentials corresponding wavepacket motion
25000 20000
coupling
I"115000
pump
'fi 1000 _!i ° So 5000 O) 0) 500 0 -500 -1000
dump
5
10
15
20
25
30
50 100 150
internuclear distance R [ Bohr radii ] Fig. 5. Engineering resonant coupling by a strong off-resonant laser field.
226 to the calcium dimer. A strong off-resonant infrared laser field couples two excited state potentials, while the pump and dump pulses couple the ground state to one of the excited states. The coupling laser modifies the wavepacket dynamics in the excited state such that dumping to deeply bound levels of the ground state, including u = 0, becomes fea~ib1e.l~ Depending on the target level, the frequency of the coupling laser field should be in the infrared to near infrared with intensities on the order of lo8 - lo9 W/cm2. The coupling laser field can be realized using nanosecond pulses, while the pump and dump pulses should be picosecond pulses as discussed before in section 2 for the alkali dimers.
5 . Conclusions
The formation of stable molecules from ultracold atoms with short laser pulses in a pump-dump scheme was presented. Picosecond pump pulses are expected to be the most efficient as they can best make use of large free-bound F'ranck-Condon factors occuring close t o atomic resonance while still being much shorter than the excited state lifetimes. In order to obtain ground state molecules with binding energies larger than one wavenumber, special stabilization mechanisms are required. Softly repulsive potential walls a t intermeditate distance or resonantly coupled excited states lead to appreciable overlap of the time-dependent excited state wavepacket with ground state levels a t the time when the wavepacket is close to its inner turning point. However, most excited potentials do not show such a favorable topology. Pump-dump photoassociation creates then only molecules in the very last vibrational levels which are extremely weakly bound. If these molecules shall be transferred to their vibronic ground state, a large number of Raman transitions is required. However, building on the coherent control idea of shaping the potentials, an additional strong off-resonant laser field can be employed which modifies the excited state dynamics such that transitions to target levels including 'u = 0 become feasible. The 'inverse' concept of de-coupling (instead of coupling) two excited state potentials with a strong off-resonant infrared laser pulse has already been realized successfully.l5 Our work demonstrates the potential of combining ultracold molecular processes and coherent control concepts.
Acknowledgements
This work has been supported by the Deutsche Forschungsgemeinschaft (C.P.K.) and by the European Commission in the frame of the Cold Molecule Research Training Network under contract HPRN-CT-200200290.
References 1. J. Doyle, B. Friedrich, R. V. Krems and F. Masnou-Seeuws, Eur. Phys. J . D 31, 149 (2004). 2. C. P. Koch, J. P. Palao, R. Kosloff and F. Masnou-Seeuws, Phys. Rev. A 70, p. 013402 (2004). 3. W. Salzmann, U.Poschinger, R. Wester, M. Weidemiiller, A. Merli, S. M. Weber, F. Sauer, M. Plewicki, F. Weise, A. Mirabal Esparza, L. Waste and A. Lindinger, Phys. Rev. A 73, p. 023414 (2006). 4. B. L. Brown, A. J. Dicks and I. A. Walmsley, Phys. Rev. Lett. 96, p. 173002 (2006). 5. F. Masnou-Seeuws and P. Pillet, Adv. in At., Mol. and Opt. Phys. 47, 53 (2001). 6. K. M. Jones, E. Tiesinga, P. D. Lett and P. S. Julienne, Rev. Mod. Phys. 78, p. 483 (2006). 7. C. P. Koch, E. Luc-Koenig and F. Masnou-Seeuws, Phys. Rev. A 73, p. 033408 (2006). 8. C. P. Koch, R. Kosloff and F. Masnou-Seeuws, Phys. Rev. A 73, p. 043409 (2006). 9. E. Luc-Koenig, R. Kosloff, F. Masnou-Seeuws and M. Vatasescu, Phys. Rev. A 70, p. 033414 (2004). 10. E. Luc-Koenig, F. Masnou-Seeuws and M. Vatasescu, Eur. Phys. J . D 31, 239 (2004). 11. C. M. Dion, C. Drag, 0. Dulieu, B. Laburthe Tolra, F. Masnou-Seeuws and P. Pillet, Phys. Rev. Lett. 86, 2253 (2001). 12. H. K. Pechkis, D. Wang, Y. Huang, E. E. Eyler, P. L. Gould, W. C. Stwalley and C. P. Koch, Phys. Rev. A 76, p. 023425 (2007). 13. A. Pe’er, E. A. Shapiro, M. C. Stowe, M. Shapiro and J. Ye, Phys. Rev. Lett. 98, p. 113004 (2007). 14. C. P. Koch and R. Moszyriski, in preparation. 15. B. J. Sussman, D. Townsend, M. Y. Ivanov and A. Stolow, Science 314, 278 (2006).
BOSE-EINSTEIN CONDENSATES ON MAGNETIC FILM MICROSTRUCTURES M. SINGH, S. WHITLOCK, R. ANDERSON, S. GHANBARI, B. V. HALL, M. VOLK, A. AKULSHIN, R. McLEAN, A. SIDOROV AND P. HANNAFORD ARC Centre of Excellence for Quantum-Atom Optics, and Centrefor Atom Optics and Ultrafast Spectroscopy, Swinburne University of Technology, Melbourne, Australia 3122 Ernail: phannaford@swin. edu.au We report on recent experiments with BECs and ultracold atoms in magnetic microtraps created by magnetic film microstructures on an atom chip. We describe the use of RF spectroscopy plus absorption imaging of ultracold atoms as a sensitive high resolution technique to map the magnetic field topology of the magnetic film. We find that the distribution of condensate atoms between the wells of an asymmetric double-well potential on the magnetic film can provide a highly sensitive technique for determining potential gradients, such as gravitational gradients, using an atom chip. Finally, we report on the use of periodic magnetic film microstructures to generate a magnetic lattice for manipulating BECs and ultracold atoms.
1. Introduction
At the 2001 Laser Spectroscopy conference in Snowbird the Tubingen and Munich groups reported the realization of a Bose Einstein condensate (BEC) in a magnetic microtrap created by current-carrying micro-wires on a substrate [ 1, 21. Such ‘atom chips’ can produce tightly confining magnetic potentials using modest electric currents, thus simplifying and speeding up the production of a BEC. In addition, atom chips allow precise control of the motion and position of the condensate, and they are compact and relatively robust, making them attractive for BEC-based devices and applications. Since Snowbird more than 20 groups around the world have now produced BECs om an atom chip. The performance of atom chips based on current-carrying conducting wires has, however, certain limitations. Current instabilities and current noise can limit the lifetime and coherence properties of a condensate, and high current densities can lead to excessive heating and breakdown of the conductors [3]. In addition, small imperfections in the wires can lead to tiny spatial deviations in the current flow which can fragment the atom cloud [4], while thermal fluctuations associated with Johnson noise in the conducting wires can lead to spin flips and a loss of atoms close to the surface of the chip [5,61. 228
229 At the last Laser Spectroscopy conference, in Aviemore, we reported the observation of a BEC on a permanent magnetic film atom chip [7, 81 based on perpendicularly magnetised TbGdFeCo magneto-optical films previously developed for our atom optics experiments [9-111. Magnetic film microstructures can produce highly stable, tightly confining magnetic potentials without ohmic heating, and they can produce very fine-structured magnetic potentials, with periodicities down to about 1 pm [l 11, and complex magnetic potentials such as ring-shaped structures. Although the potentials from magnetic film microstructures are necessarily static, they can be combined with magnetic potentials from current-carrying conductors when time-dependent magnetic fields are required, e.g., when loading the atom chip. BECs on permanentmagnet atom chips have also recently been produced using videotape [12], a CoCrPt hard disk [ 131 and FePt thin films [14]. In this paper we report on recent experiments with BECs and ultracold atoms on perpendicularly magnetised magnetic film microstructures on an atom chip. We describe a sensitive high resolution technique based on RF spectroscopy plus absorption imaging of ultracold atoms to map the magnetic topology of the magnetic film. 1151. We find that the distribution of condensate atoms between the wells of an asymmetric double-well potential on the magnetic film can provide a highly sensitive technique for determining potential gradients, such as gravitational gradients, using an atom chip 1161. Finally, we report on the use of periodic magnetic film microstructures to generate a magnetic lattice for manipulating BECs and ultracold atoms [ 17-191 2.
The Magnetic Film Atom Chip
Our atom chip consists of a perpendicularly magnetised 1 pm-thick TbGdFeCo magneto-optical film deposited on a 300 pm-thick glass substrate mounted on a silver foil ‘circuit’, which provides time-dependent magnetic fields for initial trapping and loading of ultracold atoms in the magnetic film trap. The silver circuit has U- and Z-shaped wires for a surface magneto-optical trap (MOT) and a Ioffe-Pritchard magnetic trap. The magnetic film microtrap is produced by the field above an edge of the film plus a bias field Bblasto provide radial confinement and two end-wires to provide weak axial confinement [8] (Fig. 1). The magneto-optical films are prepared in-house by magnetron sputtering at substrate temperatures of 100-200°C 1111. We typically use films comprising six 150 nm layers of Tb10Gd6Fe8&04separated by 100 nm layers of Cr. The films have high perpendicular anisotropy with excellent magnetic homogeneity, high remanent magnetization (- 3 kG), high coercivity (- 2 kOe) and a
230
Figure 1. Schematic of the magnetic film atom chip. Long range inhomogeneity in the magnetic film leads to fragmentation of the trapped atom cloud when close to the surface. After [15].
nominally high Curie temperature (~ 300°C). The magneto-optical films are coated with a reflecting gold film for use in a mirror MOT. The atom chip is mounted in the vacuum chamber with the magnetic film facing down. Initially, about 2 x 108 87Rb atoms are collected in the mirror MOT located 5 mm below the surface, optically pumped into the |F=2, mF=+2> tapping state and subsequently transferred to the Z-wire microtrap. After a preliminary RF evaporative cooling stage, the atoms are transferred to the magnetic film microtrap by adiabatically reducing the current in the Z-wire to zero. Forced RF evaporative cooling of the atom cloud in the magnetic film microtrap leads to a BEC with about 105 atoms [7, 8]. Temperature measurements on a dilute cloud of ultracold atoms indicate a remarkably low heating rate of 3 nK s"1 in the magnetic film trap, compared with 270 nK s"1 in the Z-wire trap, and a trap lifetime of about 5 s, which is limited by time-varying stray magnetic fields. 3. Spatially Resolved RF Spectroscopy to Probe Magnetic Field Topology The small kinetic energies and small spatial extent of the cloud of ultracold atoms can provide a sensitive high resolution (~ 5 urn) probe of corrugations in the magnetic potential [20]. Here, we employ precision RF spectroscopy plus high resolution imaging of tapped ultacold atoms to probe the magnetic field topology along the edge of the magnetic film [15]. After allowing the atom cloud to expand along the 5 mm edge of the magnetic film by reducing the current in the end-wires, a ramped RF field is applied perpendicular to the trap axis to resonantly out-couple atoms to untapped magnetic states at positions where the RF frequency matches the Zeeman splitting of the atoms. At the end of the ramp the resonant frequency approaches a final cut-off frequency vf corresponding to the bottom of the tap, and the atom distribution is described by a truncated Boltzmann distribution which is characterised by a spatially dependent truncation parameter [15].
231
j~ 0.8 P ' >, 0.4 L
I
I.
-S.S
-2
-1.5
I
- 1
I
-0.5
i
0
I
O.B
U_
1
J
J
1.6
2
2.8
longitudinal position z (mm) Figure 2. Absorption images for a 87Rb cold atom cloud in the magnetic microtrap located 67 urn below the edge of the magnetic film, for a RF cut-off frequency vf of (a) 1238 kHz, (b) 890 kHz, (c) 766 kHz and (d) 695 kHz. Trap parameters: Sbias = 5.7 G, field offset B0 = 0.82 G, cor = 27i x 1070 Hz. Initial cloud temperature T= 10 uK. After [15]. (3 (z, Vf) = [mFh vf - mpgfUB \ Bz(z) + B0 \ }lkBT,
where B0 is an offset field and T is a fit parameter that characterises the nonequilibrium distribution during truncation. Figure 2 shows absorption images for a cold atom cloud in the magnetic microtrap located 67 um below the magnetic film edge for four values of the RF cut-off frequency Vf. For Vf < 1.3 MHz significant fragmentation of the atom cloud is observed, and for Vf < 0.9 MHz well separated regions appear corresponding to atoms in the lowest potential wells. Using an iterative procedure we extract values of I B^y(z) \ and obtain the reconstructed magnetic field profiles shown in Fig. 3 for a number of trap heights. We find that the amplitude of the observed magnetic corrugations falls off with distance y as ^,-1.85 ±0.3 yj^ characteristic period of the corrugations is about 390 um for j > 1 0 0 u m with additional higher frequency components appearing as the atom cloud is brought closer to the film. The amplitude of the corrugated potential is about three orders of magnitude larger than that estimated from the roughness (~ 50 nm) of the polished film edge, but is consistent with the y'2 behaviour predicted for long range inhomogeneities in the magnetisation of the film [15]. Investigating this further, measurements taken with the atom cloud positioned below the magnetic film at x= 100 um from the edge show significant fragmentation while those recorded below the non magnetic film at x = -100 um from the edge show almost no corrugation, thus confirming that the fragmentation results from spatial variations of the magnetisation in the body of the magnetic film rather than from fluctuations along the edge of the film.
232
I -2.5
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
2.5
longitudinal position z (mm) Figure 3. Magnetic field profiles B,(z) measured using spatially resolved RF spectroscopy (solid lines) for various distances y below the edge of the magnetic film. The dotted lines correspond to measurements of the magnetic film edge using a scanning magneto-resistance probe, which measures the By component of the corrugated magnetic field. The relative longitudinal offset has been adjusted for optimal agreement. After [15].
After removing the magnetic film atom chip from the vacuum chamber, the magnetic corrugation was further characterised using a home-built magnetoresistance microscope, which allows measurements over a wide range of distances with a spatial resolution of about 50 pm. The magneto-resistance measurements, indicated by the dotted lines in Fig. 3 , show a remarkable correlation with the FW spectroscopy cold atom data, thus verifying the validity of the RF spectroscopy cold atom technique. The observed magnetic inhomogeneity in the magnetic films is attributed to deterioration during vacuum bake out of the magnetic film atom chip (140°C for 4 days). Assuming the inhomogeneity originates from reversal of magnetic domains, we conclude that the mean magnetisation of the film is about 90% of the saturation value. After remagnetising the film the level of inhomogeneity was reduced by about a factor of 10. 4. Dynamic Splitting of a BEC in an Asymmetric Double Well
In the reconstructed magnetic field profiles in Fig. 3 , we identify a double-well potential in the region near z = 0 when the magnetic trap is brought close to the surface of the film by increasing Bbiasin the x direction [16]. The double well originates from higher spatial frequency components of the magnetisation inhomogeneity at distances close to the magnetic film. Figure4 (left) shows a series of high resolution absorption images as a BEC is slowly brought (in 0.5 s) from about 170 pm below the film, where the
233
Figure 4. Left: dynamic splitting of a BEC in a double-well potential. High resolution absorption images are shown as the BEC is slowly moved from about 170 (xm below the film to about 57 ^m below the film. Right: characterisation of the double-well potential as a function of trap-surface separation is performed using two-component clouds. The well separation A,, barrier height (3 and trap asymmetry A are shown schematically in (c). After [16].
potential behaves as a single well, to 57 um below the film, where it behaves as a double well. Under these conditions the BEC dynamically splits into two, with about equal numbers of atoms in each well. The double well is characterised by taking absorption images for various trap heights y and determining the well separation A, and barrier height (3 as a function of y (Fig. 4, right) [16]. The trap frequencies of the two wells, obtained from dipolar oscillations, are identical within a few percent. The asymmetry A in the double well can result in a marked difference in the number of atoms in the two wells when the barrier height P < the chemical potential u. Slight tilts of the atom chip with respect to gravity, e.g., from tiny movements of the optical table, were found to have a marked effect on the distribution of the condensate between the two wells. The gravitational gradient is cancelled by applying a small magnetic field gradient provided by a current imbalance ± 57 between the two end-wires on the chip. The trap asymmetry A can be calibrated in situ against 87 by adiabatically splitting a condensate in a symmetric double well with barrier height P > ^ and using spatially resolved RF spectroscopy, as described in Sect. 3. A condensate with a relatively small number of atoms is initially prepared in a single well at y = 170 um below the film and then Bbias is slowly increased to produce a tailored double well at y = 155 um, where the well separation A, ~ 70 um and the barrier height p = V* u. The splitting time ts is chosen to be larger than A'1 to ensure the condensate remains in the ground state of the potential well.
234
'-2S3
-2CQ
-ISO
-100
-50
0
50
100
150
2S&
25S
Asymmetry A (Hz) Figure 5. Fractional number difference versus asymmetry A of a double-well potential. The asymmetry is varied by changing the end-wire current imbalance 87. The experimental points and line of best fit (dashed line) compare well with the simple analytic result (solid line). After [21].
Absorption images of the condensate are then recorded as 87 is ramped over ±117 mA, and the fractional atom number difference (NR + NL)/(NR -NL) = AN/N is recorded against 87, and hence A. From the distribution of the data (Fig. 5) we estimate a single-shot sensitivity ASNR = 16 Hz, which is limited by the shot-to-shot variation in the number of atoms in the condensate. From this result we infer a single-shot sensitivity to gravity gradients of 8g/g ~ 2 x 10"4. Using a simple model in which the double-well potential is represented by two uncoupled 3D harmonic oscillators we obtain [16] AAMVw 1.65 (47r)(a0/15aA02/5 A/TO, where a is the scattering length, ao is the ground state harmonic oscillator length, and TO is the geometric mean trap frequency. Inserting experimental parameters into this expression yields a straight line in good agreement with the experimental data (Fig. 5). The sensitivity could be further enhanced by decreasing the trap frequency TO, thereby lowering the chemical potential n, or by using multiple double-well potentials on a single chip. For 100 double wells and reasonable values for the parameters we estimate a possible single-shot sensitivity of ASNR ~ 0.04 Hz, or 8g/g ~ 5 x 10~7. 5. Periodic Magnetic Lattices Optical lattices produced by the interference of intersecting laser beams are widely used to manipulate BECs and clouds of ultracold atoms, e.g., in quantum tunnelling experiments. Here, we report on the use of periodic magnetic lattices based on permanent magnetic films as an alternative to optical lattices [17-19].
235
(a)
(b)
Figure 6. (a)-(c) ID periodic array of parallel magnets with perpendicular magnetisation, (d)—(f) Calculated contour plots of the magnetic field in the central region in the yOz plane with (d) no bias fields, (e) bias field Bly = -15 G, and (f) bias fields Blx=-20G, B,y = -l5 G. 1001 magnets, a = 1 urn, t = 0.05 urn, /, = 1000.5 urn, and 4nM, = 3.8 kG. 7 G contour spacing. After [17].
Magnetic lattices based on magnetic films have a number of distinctive characteristics: (i) They are highly stable with low technical noise; (ii) There is no spontaneous emission; (iii) The atoms need to be prepared in low magnetic field-seeking states, allowing RF evaporative cooling in situ and the use of spatially resolved RF spectroscopy; (iv) They can have large and controllable barrier heights and large trap curvature leading to high trap frequencies; (v) They can be constructed with a wide range of periods, down to about 1 jam, and they can have complex potential shapes; (vi) They are necessarily static, precluding use in time-dependent lattice experiments; and (vii) 2D and ID magnetic lattices, but not 3D, can be constructed. Thus magnetic lattices can be considered as complementary to optical lattices, in much the same way as magnetic traps are complementary to optical dipole traps. We first consider a simple ID magnetic lattice produced by a single periodic array of parallel magnets of thickness t with perpendicular magnetisation, period a, and bias fields B\x and B\y parallel and perpendicular to the grooves, respectively (Fig. 6) [17]. For B\x = B\y = 0 (Fig. 6(d)), the magnetic field falls off exponentially with distance z above the surface, representing the case of a magnetic mirror [9, 10]. For B\x = 0, B\y = -15 G (Fig. 6(e)), the magnetic field develops 2D magnetic traps with zero potential minima; this configuration can give rise to spin flips and is not suitable as a
236
Figure 7. (a) 2D periodic array consisting of two crossed layers of parallel magnets with perpendicular magnetisation, (b) 3D plot of the magnetic field. 1001 magnets, a = 1 urn, ?, = 0.322 urn, <2 = 0.083 urn, /, = /,= 1000.5 urn, 47tMz = 3.8kG, B,,= -4.08G, BU, = -6.05G, S,2 = -0.69 G. S min = 2.7 G, zmin = 1.22 urn, AU* = ALP' = 485 uK, AU Z = 307 nK, to, = CD, = 2n x232 kHz,
lattice. For Blx = -20 G, B\y = -15 G (Fig. 6(f)), the magnetic field has 2D magnetic traps with non zero potential minima. For an infinite ID magnetic lattice, the potential minimum, trap height, and barrier heights are given by [17] Bmm = Blx\; zmm = (a/2n) In [B0yl\Bly\] B,
where B0y = B0 (1 - e kt) ekl and k = 2n/a. The potential minimum, trap height and barrier heights can be controlled by means of the bias fields 5 u and B\y, We now consider a 2D magnetic lattice produced by two crossed periodic arrays of parallel magnets, thicknesses t\ and ti, separated by distance s (< a/2n), and with bias fields B]x, B}y (Fig. 7(a)). Figure 7 (b) shows the calculated magnetic field for the parameters given in the caption. For an infinite symmetrical 2D magnetic lattice, the trap minimum, trap height, and barrier heights are given by [17]
Bl
with the constraint B\y = cQB\x for a symmetrical lattice. The c\s are dimensionless constants that involve geometrical constants a, s, t\ and t2 of the magnetic arrays and B0x = B0 (1 - e~*'2) e^+'1+'2). The potential minimum, trap position and barrier heights can be controlled by varying BIX and Biy. Other configurations of 2D magnetic lattices have also been proposed [17, 19].
237
Figure 8. Absorption images, after 8 ms time of flight, of a 87Rb condensate after it has been brought to various distances below the 10 urn-period magnetic microstructure, ranging from z ~ 6 um (first image) to z ~ 150 \an (final image). Bv = 0 (perpendicular to grooves).
6. Realisation of a Permanent-Magnet Lattice for Ultracold Atoms We have constructed a ID magnetic lattice with period a = 10 (am and dimensions 10 mm x 10 mm using a six-layer structure of perpendicularly magnetised TbGdFeCo magneto-optical films deposited on a grooved silicon wafer. The magnetic microstructure is coated with a reflecting gold film for use in the mirror MOT. The magnetic microstructure is mounted on an atom chip, similar to that described in Sect. 2, but with a 30 mm long Z-wire, in addition to the 5 mm long Z-wire, perpendicular to the grooves for effective loading of ultracold atoms into the magnetic lattice. To ensure minimal demagnetisation during vacuum bakeout care was taken not exceed a temperature of 100°C. First, a BEC with ~ 2 x 105 87Rb atoms is prepared in the 5 mm Z-wire trap 250 um below the chip surface where the magnetic microstructure has a negligible effect. The current in the Z-wire is then slowly ramped down to bring the condensate closer to the surface so that it interacts with the magnetic microstructure for less than 5 ms before being released. In this initial experiment the bias field perpendicular to the grooves was switched off, so the microstructure is not yet a trapping magnetic lattice. Absorption images are taken after a time of flight (TOP) of 8 ms. As the condensate is brought closer (from ~ 150 um to ~ 6 (am) to the magnetic microstructure, the TOP images change from a tiny cloud to sharply defined crescent-shaped structures, indicating that the condensate has interacted with the magnetic microstructure. The crescent-shaped structures are attributed to reflection of the atoms from a periodic corrugation in the magnetic potential caused largely by a magnetic field produced by residual current in the Z-wire. When the temperature of the atom cloud is raised above the BEC transition the crescent-shaped structures broaden markedly due to the larger trap volume. Bias fields of B\x = —15 G, B\y = -30 G are then turned on to create a magnetic lattice of 2D traps with a barrier height of about 1.2 mK. Absorption images taken after loading about
238
Holding Time (ms) Figure 9. Left: Absorption image, after 14 ms time of flight, of a thermal cloud of ~ 5 x 10s 87Rb atoms at 15 uK trapped in a 10 urn-period ID magnetic lattice at 6 - 8 urn from the surface of the atom chip. Bx = —15 G, By — -30 G. The image is taken in a direction parallel to the grooves. Right: Decay of the thermal cloud after release from the magnetic lattice. The lifetime is 0.34 s.
5 x 105 87Rb atoms in a 15 uK thermal cloud from the 5 mm Z-wire trap into the 10 mm x 10 mm magnetic lattice and after 14 ms TOP indicate that atoms are trapped in about 100 of the 1000 lattice sites at 6 - 8 urn from the surface (Fig. 9, left). By releasing the atoms from the magnetic lattice, the lifetime of the atoms trapped in the magnetic lattice is found to be 0.34 s (Fig. 9, right). With our current pixel image resolution of 9 urn it was not possible to resolve atoms in individual lattice sites of the 10 um-period magnetic lattice. We also note that at this resolution there is little evidence of fragmentation even though the atoms are trapped very close (6-8 urn) to the magnetic microstructure. In future it should be possible to populate a large fraction of the 1000 lattice sites, by loading from the 30 mm long Z-wire trap, and to resolve atoms in the individual lattice sites using spatially resolved RF spectroscopy. We also plan to implement RF evaporative cooling on the trapped atom clouds to create condensates on individual sites of the magnetic lattice and to study the diffraction of the condensate atoms when released from the magnetic lattice. We also plan to implement a 2D magnetic lattice on an atom chip and to construct magnetic lattices with smaller periods (1-5 um) in order to study quantum tunnelling in ID and 2D magnetic lattices. One of the major challenges will be to fabricate periodic magnetic microstructures having sufficiently smooth magnetic potentials very close to the surface to produce high-quality small-period magnetic lattices that preserve quantum coherence. During the preparation of this manuscript a preprint appeared in which ultracold rubidium atoms were trapped in > 30 lattice sites of a 2D FePt magnetic lattice with period 20 urn [22].
239 7. Summary We have implemented perpendicularly magnetised magnetic film microstructures to trap and manipulate BECs and ultracold atoms on an atom chip. We have developed a sensitive high resolution technique, based on RF spectroscopy plus absorption imaging of ultracold atoms, to map the magnetic field topology of the magnetic film. We have used a double-well potential close to the magnetic film to dynamically split a BEC. We find that the distribution of condensate atoms between the two wells can provide a highly sensitive technique for determining potential gradients, such as gravitational gradients, on an atom chip. Finally, we have loaded and trapped small clouds of ultracold atoms in a 1D permanent-magnet lattice on an atom chip. Acknowledgments
This project is supported by the ARC Centre of Excellence for Quantum-Atom Optics and a Swinburne University Strategic Initiative grant. References
1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22.
H. Ott et al., Phys. Rev. Lett. 87,230401 (2001). W. Hansel et al., Nature 413,498 (2001). D. C. Lau et al., Eur. Phys. J. D 5, 193 (1999). D. W. Wang et al., Phys. Rev. Lett. 92,076802 (2004). M. P. A. Jones et al., Phys.Rev. Lett. 91,080401 (2003). J. Esteve et al., Phys. Rev. A 70, 043629 (2004). B. V. Hall et al., Laser Spectroscopy XVII, (World Scientific, 2005), 275. B. V. Hall et al., J. Phys. B: At. Mol. Opt. Phys. 39,27 (2006). A. Sidorov et al., Compt. Rend. 2, Series IV, 565 (2001). A. Sidorov et al., Acta Phys. Pol. B 33,2137 (2002). J. Y. Wang et al., J. Phys. D: Appl. Phys. 38,4015 (2005). C. D. J. Sinclair et al., Phys. Rev. A 72,031603(R) (2005). M. Boyd et al., preprint cond-mat/0608370 (2006). T. Fernholz et al., preprint cond-mat/07052569 (2007). S. Whitlock et al., Phys. Rev. A 75, 043602 (2007). B. V. Hall et al., Phys. Rev. Lett. 98, 030402 (2007). S. Ghanbari et al., J. Phys. B: At. Mol. Opt. Phys. 39, 847 (2006). R. Gerritsma and R. J. C. Spreeuw, Phys. Rev. A 74,043405 (2006). S. Ghanbari et al., J. Phys. B: At. Mol. Opt. Phys. 40, 1283 (2007). S. Wildermuth et al., Nature 435,440 (2005). S. Whitlock, PhD Thesis, Swinburne University of Technology, 2007. R. Gerritsma et al., preprint physics/07061170 (2007).
This page intentionally left blank
COLD ATOMS AND MOLECULES11
This page intentionally left blank
ULTRACOLD METASTABLE HELIUM-4 AND HELIUM-3 GASES W. VASSEN, T. JELTES, J.M. MCNAMARA, A.S. TYCHKOV, W. HOGERVORST Laser Centre Vrije Universiteit Amsterdam, The Netherlands K.A.H. VAN LEEUWEN Dept. of Applied Physics, Eindhoven Univ. of Technologv, The Netherlands V. KRACHMALNICOFF, M. SCHELLEKENS, A. PERRIN, H. CHANG, D. BOIRON, A. ASPECT, C.I. WESTBROOK Lab. Charles Fabry de 1’Institut d’Optique, Univ. Paris-Sud, Palaiseau, France We discuss our work to obtain a condensate containing more than lo7 atoms and the first degenerate Fermi gas in a metastable state. Sympathetic cooling with Helium-4 is used to cool lo6 Helium-3 atoms to a temperature T/TF< 0.5. The ultracold bosonic and fermionic gases have been used to observe the Hanbury Brown and Twiss effect for both isotopes, showing bunching for the bosons and antibunching for the fermions. A proposal for high resolution spectroscopy at 1.557 pm, connecting both metastable states directly, is discussed at the end.
1. Introduction Helium in the metastable 2 3SI state has been Bose condensed in 2001 by two French groups [ 1,2]. Since then research on ultracold metastable helium gases has been concentrated on determining the scattering length, measuring loss processes, studies of the hydrodynamic regime and producing longrange molecules. In 2005 two other groups realized BEC in metastable helium [3,4]. Studies of atom-atom correlations in ultracold clouds of 4He bosons were started around that time as well, exploiting the unique detection properties of metastable atoms [5]. In this contribution we will discuss experiments performed in Amsterdam. We will discuss the setup in which a BEC containing over 10’ atoms and a degenerate Fermi gas (DFG) of metastable 3He containing over lo6 atoms was realized. Next we discuss experiments where the Amsterdam and OrsayRalaiseau groups joined forces measuring the Hanbury Brown and Twiss effect both for a gas of ultracold 243
244
4
bosons and a cloud of ultracold fermions, demonstrating atom bunching for He and atom antibunching for 3He.
2. Helium level structure and relevant parameters Relevant energy levels are shown in Fig. 1. The helium atom is strongly LScoupled effectively leading to two spectra, that of parahelium, where the electron spins are antiparallel and that of orthohelium where the spins are parallel. The ground state of orthohelium (ls2s 3S1 or He*) can be laser cooled with laser light of 1083 nm exciting the ls2s 3SI - ls2p 3P (2 3S - 2 3P) transition. The 2 'S1 state, populated in a DC discharge, has a lifetime of -8000 s which is infinite for all practical purposes. Its 20 eV internal energy provides the unique detection possibilities: (almost) everything a He* atom hits gets ionized and the released electron can be counted with high efficiency using electron multipliers or microchannel plate (MCP) detectors. IONIZATION LIMIT
24
-
23 22 -
metastable state: is2s 3S, (4He*) F= i/2,3/2(3He*)
4.8 eV
Figure 1: Level scheme showing the relevant energy levels for laser cooling and trapping of 4He in the 2 3SIstate (M,=+l in a magnetic trap). The 2 'S state is split in a F=3/2 and a F=1/2 state (AE=6740 MHz) in the case of 'He. The F=3/2 state has the lowest energy and is used for cooling and trapping (M~=+3/2in the magnetic trap).
245
Losses in ultracold He* gases, either in a magneto optical trap (MOT) or a magnetic trap, are mainly due to two-body inelastic Penning ionization where one ion and one ground state atom are produced in a collision. The loss rate constant for this process in an unpolarized gas is quite large: -1x10"'° cnrVs in the dark and ~4xlO" 9 cm3/s in the light of a MOT [6,7]. When the atoms are in a fully stretched state (as in a magnetic trap) these losses are suppressed in first order and the rate constant drops to ~lxlO~ 1 4 crnVs. In the fully stretched state of 4He* (|J,Mj>= !,+!>) Penning losses are therefore acceptable and a EEC can be realized. The 4He*-4He* scattering length a44 that determines many properties of a EEC has recently been measured with high accuracy: a44(exp)= +7.5105 (25) nm [8]. Recent ab initio calculations of the molecular potential have allowed an astonishing accuracy in the calculation of a44 as well and approach the experimental accuracy: a44(theory)= +7.562 (28) nm [9]. Mass scaling of the potential then provides a theoretical value of a34(meory)= +27.1 (5) nm for the scattering length determining the collision rate between spin-polarized 3He* and 4He* atoms [9].
3. The experimental apparatus Laser cooling and trapping of He* atoms is performed at a wavelength of 1083 nm. As the isotope shift is 33 GHz we use two separate laser systems, each laser locked to the cycling transition of the relevant isotope in a DC discharge cell. The experimental setup is shown in Figs. 2 and 3. We use a liquid nitrogen cooled DC discharge source, which uses recycling of helium atoms from the turbo pump exhaust via molecular sieves.
jA*.
**».
Jtom».
MOT
dc discharge source laser collimator
Zeeman slower
+1° deflection Figure 2: Schematic or the experimental setup used to trap 3He* and 4He* atoms. The collimation section (two-dimensional) significantly improves the MOT loading. A deflection zone (in the horizontal plane) prevents ground state atoms entering the 2 m long Zeeman slower.
246
Figure 3: Schematic of one-dimensional Doppler cooling in the cloverleaf magnetic trap (left) and UHV chamber (right; the arrow indicates the absorption imaging beam). The coils are located in plastic boxes positioned inside the re-entrant window flanges.
We typically load 2 x 1Q9 4He* atoms or 1 x 1093He* atoms in 1 s in a MOT at a temperature of 1 mK. After a short spin polarization pulse the atoms are loaded into a spherical cloverleaf trap matching the MOT cloud. The high (24 G) bias magnetic field in this geometry allows efficient 1-dimensional Doppler cooling to a temperature of 0.15 mK (in 2 s), shrinking the cloud considerably and increasing the phase space density about a factor 600 to a value around 10"4 - 10"5 [3]. In another 2 s the cloud is compressed at small bias field (~3 G). The trap lifetime is 2 - 3 minutes (either 3He or 4He). To produce a BEC we perform a 15 s rf evaporative cooling ramp realizing a BEC containing (1.5-4) x 1Q7 4He* atoms [3]. We monitor the BEC applying absorption imaging at 1083 nm (not very efficient due to the low efficiency of CCD cameras at 1083 nm) or on an MCP detector mounted directly under the trap (see Fig. 4). The latter technique is very sensitive allowing us to monitor a BEC after up to 75 s trapping inside the magnetic trap. To monitor the growth of a condensate we use a second MCP detector that attracts all ions produced by Penning ionization. As soon as the condensate starts to grow a dense cloud forms and the ion signal suddenly increases [3]. The lifetime of our condensate is about 1 s. We studied this as a function of the thermal fraction and found that a large thermal fraction reduces the lifetime of the condensate considerably. We attribute this to transfer of condensate atoms to the thermal cloud during condensate decay. A simple model that assumes thermal equilibrium during the decay confirms this and allows extraction of the two- and three-body loss rate constants [3,10].
247
210
170
180
190
200
210
200
210
250
180 190 Tims of flight (ms)
Figure 4: Time-of-flight pictures of a EEC (upper figure), a DFG (middle figure) and a mixture (lower figure). The upper figure shows three plots, each with a slightly different end rf frequency, showing a thermal cloud above the EEC temperature, a mixture of BEC and thermal cloud, and a pure BEC with the typical inverted parabola shape. In the middle figure a fit to a Fermi-Dirac distribution is shown from which we extract a temperature T=0.45 TV. In the lower figure the dashed-dotted line shows the BEC contribution to the signal and the dashed line the DFG contribution.
248 70 (XI
2 50 e
:
44,
c
8
i"
XI 10 120
110
160
IHI
im
I-,,
IW
llmE 1mj
I
Figure 5: He* atom laser, realized by repeatedly output coupling of a fraction of the BEC in 250 MHds ramps. The time-of-flight signal shows the effects of mean-field repulsion.
We can couple out a small fraction of the atoms from a condensate applying an rf ramp as shown in Fig. 5. We applied repeatedly a 250 MHz/s rf ramp to couple atoms out of the condensate showing a typical atom laser pulse shape, revealing mean-field repulsion. To produce a degenerate Fermi gas a we load a MOT with both isotopes [3,6,7,11,12]. With an equal number of both isotopes in our gas reservoir we trap 1 x lo94He* atoms and 7 x 10' 3He*atoms in a MOT. However, we can not cool so many fermions and therefore reduce the number of fermions to 110% of the number of bosons. The large heteronuclear scattering length provides almost ideal conditions for sympathetic cooling and we obtain a DFG containing 2 x lo6 3He* atoms at T/TF=0.45 (Fig. 4) as well as a degenerate mixture (also Fig. 4). Three-body losses probably limit the lifetime of this mixture to -1 0 ms [ 121.
4. Hanbury Brown and Twiss experiments In 2005 the atomic analogon of the Hanbury Brown and Twiss (HBT) effect was demonstrated for metastable 4He* atoms released from a cloverleaf magnetic trap very similar to the one in Amsterdam [5]. The HBT effect, first demonstrated for light in the fifties of the 20th century [ 131, represents the measurement of the two-body second-order correlation function
displaying the joint probability of detecting two atoms (or photons) at locations r and r'. For incoherent sources this function equals 1 for large separations and will tend to 2 for bosons and 0 for fermions, for detector
249 separations smaller than the correlation length /: bosons bunch and fermions antibunch. In the Amsterdam experiments we demonstrated bunching for 4 He* and antibunching for 3He* trapped in identical traps at the same temperature [14]. For this purpose the position sensitive detector (PSD) of the Orsay (now Palaiseau) group was transported to Amsterdam and mounted under the Amsterdam UHV chamber (which is shown in Fig. 3). The detection part of the setup is shown in Fig. 6, showing on the left the PSD and on the right a schematic of the measurement. The correlation function for a measurement of atoms released from a harmonic trap is given by [14] 2 T( )
(Ax, Ay, Az) = 1 ± r/exp
with correlation lengths in each of the three spatial directions /; = ms where sf — ^ kBT /mcof , m the mass of the atom, T the temperature, t the drop time and co, the trap frequency. The contrast 77 (0 - 1) depends on the detector resolution in all three spatial dimensions. This two-particle detector resolution is 6 nm in the vertical (z-) direction and 500 |am in the horizontal plane of the detector. As the correlation length in the x-direction is much smaller than the detector resolution in that direction (we have a cigar-shaped trap with long axis in the x-direction) 77 is on the order of 1/15.
Figure 6: Position-sensitive MCP detector and schematic of the experiment. The inset of the figure on the right conceptually shows the two 2-particle amplitudes that interfere to give bunching or antibunching.
250 As the detector resolution is very good in the vertical plane, we plot the measured correlation function along the z-direction in Fig. 7. This plot was obtained by averaging over about 1000 clouds per isotope, with -lo4 detector points per shot. The temperature was 0.5 pK. We also measured the correlation function along the y-direction, however with substantial broadening along the horizontal axis due the 500 pm resolution in the x-y plane [14]. Fig. 7 clearly demonstrates antibunching for fermions and bunching for bosons, over a macroscopic distance of -1 mm. The measured ratio of the correlation lengths is 1.3 0.2, which is as expected, realizing that the cloud sizes of both isotopes in the trap are equal, the drop times are equal and the mass ratio is 4/3. Also the individual correlations lengths agree very well with the formula 1, =ht/ms, We observe a small discrepancy in the amount of contrast for both isotopes. There are two possible experimental imperfections that affect the bosons and fermions differently and that may explain this, i.e. the non-sudden trap switch-off and the determination of the detector resolution function.
-1 0
2
3
Separation Az (mm)
Figure 7: Normalized correlation functions for 4He* in the upper plot and 'He* in the lower plot, measured in the same trap at the same temperature of 0.5 pK.Error bars correspond to the square root of the number of pairs in each time bin. The line is a fit to a Gaussian function. Correlation lengths of 0.75 0.07 mm and 0.56 f 0.08 mm for fermions and bosons, respectively, are extracted.
*
251
T
0.5
1
0,5
Temperature
1.5
Temperature (^K)
Figure 8: Measured temperature dependence of correlation length and contrast. The curves are calculations and describe the measurements reasonably well.
We took data for fermions at 0.5 uK, 1 .0 |^K and 1 .4 uK. The fit results for /z and 77 are shown in Fig. 8 together with the calculated results. The limited resolution in the horizontal plane can be circumvented by using a negative atom lens, reducing the apparent size of the cloud in the horizontal plane. This does not affect the correlation length in the vertical plane but does affect the contrast via the correlation length ly, which increases. To implement this we focused a 300 mW laser beam, with elliptical waist and 300 GHz blue detuned from the 1083 nm transition, for 0.5 ms through the expanding 3He* cloud after switching off the magnetic trap. The results are shown in Fig. 9 and qualitatively agree with expectations. 1JS-,
«^i**-M^M* ,&f~r*V
1 %, § emO
0
1
1
3
Separation Az (mm) Figure 9: Normalized correlation functions along the z (vertical) axis for 3He*, demonstrating the effect of a diverging atomic lens in the x-y plane. The light circles are without lens and the dark squares with lens. The dip is deeper with the lens because the correlation length in the ydirection is increased which affects the contrast (and not the correlation length in the zdirection).
252
a unbundling
Pair separation (ms) Figure 10: Raw data of an atomic lens experiment with fermions. The data show the numbers of pairs measured for a 25 us time bin (1 ms corresponds to a vertical separation of 3.5 mm). They are averaged over 1000 shots.
We wish to emphasize that the antibunching effect in Fig. 7 is big enough that significant processing of the data is not necessary. Fig. 10 shows the raw data of one of our runs before normalizing the data. Normalization (to get the data of Figs. 7 and 9) is done by dividing the raw data by the autoconvolution of the sum of the 1000 single-particle distributions. Therefore the raw data show the antibunching effect as well as the Gaussian spatial distribution of the cloud.
5. Proposed metrology experiment As one of the possible applications of samples of ultracold 4He* and 3He* atoms we envision to excite the "forbidden" 2 3S| - 2 'S0 transition at 1.557 urn [15]. This transition is only magnetic-dipole (Ml) allowed with an Einstein A-coefficient of 6.1 x 10"8 s"1 [16]. The high-resolution potential is very large as the natural linewidth of the transition is 8 Hz, fully determined by the 20 ms lifetime of the 2 'S0 metastable state due to two-photon (2E1) decay to the ground state. Measuring this transition will directly link the orthohelium and parahelium system (see also Fig. 1). The absolute frequency (for both isotopes) will provide sensitive tests of atomic theory for twoelectron systems, measuring QED contributions which are large for S-states. Measuring the isotope shift, the main theoretical inaccuracy is in the difference in mean nuclear charge radius of the 4He and 3He nucleus.
253
To measure the transition we theoretically investigated two setups: (1) spectroscopy with a laser-cooled and collimated atomic beam as provided from the Zeeman-slower and (2) spectroscopy in a one-dimensional optical lattice, loaded from a magnetically trapped and evaporatively cooled cloud. For both situations we solved the optical Bloch equations simplifying the model neglecting the decay from 2 'S to 2 3S as well as the decay from 2 3S to 1 'S (see Fig. 11). For the beam experiment we assume a 100 m/s Zeemanslowed atomic beam, transversely cooled to twice the Doppler limit. With an atomic beam intensity of 1011 atoms per second, a 2 W 100 kHz laser beam (Icmxlcm excitation region) will provide an on-resonance Rabi frequency Q= 61 s"1, which for an excitation time of 100 us will lead to an excited state population P22=4.6 x 10"7. A considerable flux of 5 x 103 atoms per second in the 2 'S state then may be expected.
Figure 1 1 : Relevant energy levels for 2 Si — 2 S0 spectroscopy and decay rates.
The expected spectroscopic linewidth is 400 kHz in this case, limited by Doppler broadening. Using 1083 nm light resonant with the 2 3Si - 2 3P2 transition the nonexcited atoms can simply be deflected realizing a (hopefully) zero-background signal on a metastable atoms detector. A schematic view of the proposed setup, with three laser-atomic-beam crossings depicted, is shown in Fig. 12.
2 ifycdSmuSani deffeoion IWBnm
2-'S,rallimat dsfieaion tSBrni!
Fig. 12: Schematic of the proposed setup to measure the 2 3Si — 2 'S0 transition in a beam.
254
The one-dimensional lattice has more potential to achieve high accuracy. In the experiments described in Section 4 an ultracold cloud at temperatures around 1 pK of 3He*, 4He*, or a mixture was presented. Such a cloud may be transferred to a far off-resonance dipole trap (FORT). When a single retroreflected beam is used a one-dimensional lattice results which, when a wavelength of 1.557 pm is used, can easily trap a cloud from our cloverleaf trap at a temperature of a few pK. The experiment may be performed using the same 2 W laser as discussed in the beam experiment as the trapping potential is insensitive to the wavelength on a scale of an experiment on the forbidden transition. A narrower linewidth laser, though, is useful here as the width of the spectroscopy signal will now be limited by the laser linewidth as Doppler broadening is suppressed by Lamb-Dicke narrowing. Detection is now more challenging as the 2 'S atoms are anti-trapped and slow. Photo ionization, or detection of ions due to Penning ionization with trapped 2 3S atoms, are then options to consider. The accuracy of the spectroscopy in the lattice will be limited by how well the AC Stark shift due to the lattice potential can be controlled and measured. Calculations predict a 2 MHz shift of the transition at the maximum of the potential (the atoms are trapped in the light as the 1.557 pm wavelength is red detuned from the 1083 nm transition). The best experiment therefore will be to trap at a magic wavelength (where the AC Stark shifts of the metastable states are equal), which will not be easy as the highest magic wavelength available is 410 nm at which wavelength very high laser power is required to trap He* atoms. A second magic wavelength occurs at 35 1 nm and may be more suitable for spectroscopy on this forbidden transition.
References 1. A. Robert, 0. Sirjean, A. Browaeys, J. Poupard, S. Nowak, D. Boiron, C. Westbrook, A. Aspect, Science 292,461 (2001). 2. F. Pereira Dos Santos, J. LConard, J. Wang, C.J. Barrelet, F. Perales, E. Rasel, C.S. Unnikrishnan, M. Leduc, C. Cohen-Tannoudji, Phys. Rev. Lett. 86, 004359 (2001). 3. A.S. Tychkov, T. Jeltes, J.M. McNamara, P.J.J. Tol, N. Herschbach, W. Hogervorst, W. Vassen, Phys. Rev. A 73,031603(R) (2006). 4. R.G. Dall, A.G. Truscott, Opt. Comm. 270,255 (2007). 5. M. Schellekens, R. Hoppeler, A. Perrin, J.V. Gomes, D. Boiron, A. Aspect, C.I. Westbrook, Science 310, 648 (2005).
255
6. R.J.W. Stas, J.M. McNamara, W. Hogervorst, W. Vassen, Phys. Rev. A 73,032713 (2006). 7. J.M. McNamara, R.J.W. Stas, W. Hogervorst, W. Vassen Phys. Rev. A 75,06271 5 (2007). 8. S. Moal, M. Portier, J. Kim, J. DuguC, U.D. Rapol, M. Leduc, C. CohenTannoudji, Phys. Rev. Lett. 97, 023203 (2006); for a small correction to a44see S. Moal, thesis ENS Paris (2006). 9. B. Jeziorski, Private communication (2007). 10. P. Zin, A. Dragan, S. Charzynski, N. Herschbach, P. Tol, W. Hogervorst, W. Vassen, J. Phys. B 36, L149 (2003). 11. R.J.W. Stas, J.M. McNamara, W. Hogervorst, W. Vassen Phys. Rev. Lett. 93,053001 (2004). 12. J.M. McNamara, T. Jeltes, A.S. Tychkov, W. Hogervorst, W. Vassen Phys. Rev. Lett. 97,080404 (2006). 13. R. Hanbury Brown, R.Q. Twiss, Nature 177,27 (1956). 14. T. Jeltes, J.M. McNamara, W. Hogervorst, W. Vassen, V. Krachmalnicoff, M. Schellekens, A. Perrin, H. Chang, D. Boiron, A. Aspect, C.I. Westbrook, Nature 445,402 (2007). 15. K.A.H. van Leeuwen, W. Vassen, Europhys. Lett. 76,409 (2006). 16. K. Pachucki, Private communication (2006).
This page intentionally left blank
SINGLE ATOMS AND QUANTUM OPTICS I
This page intentionally left blank
RECENT PROGRESS ON THE MANIPULATION OF SINGLE ATOMS IN OPTICAL TWEEZERS FOR QUANTUM COMPUTING A. Browaeys*, J. Beugnon, C. Tuchendler, H. Marion, A. Gaetan, Y.Miroshnychenko, B. DarquiB, J. Dingjan, Y.R.P. Sortais, A.M. Lance, M.P.A. Jones, G. Messin, P. Grangier
Laboratoire Charles Fabry de l’lnstitut d’optique, CNRS, Univ. Paris-sud, Campus Polytechnique, R D 128, 91 I27 Palaiseau cedex, France *E-mail: antoine. browaeysQinstitutoptique.f r This paper summarizes our recent progress towards using single rubidium atoms trapped in an optical tweezer t o encode quantum information. We demonstrate single qubit rotations on this system and measure the coherence of the qubit. We move the quantum bit over distances of tens of microns and show that the coherence is preserved. We also transfer a qubit atom between two tweezers and show no loss of coherence. Finally, we describe our progress towards conditional entanglement of two atoms by photon emission and twophoton interferences.
Quantum computing has been proposed to solve certain classes of computational problems, such as factoring and searching, faster than using a classical computer’ . In addition, one could engineer these quantum computers in such a way that they could perform simulations of quantum systems. From a fundamental point of view, a quantum computer can be thought as a collection of two-level systems, well isolated from the environment, which can interact with each other in a controlled way. Building such a quantum computer may therefore help to understand decoherence of a macroscopic quantum system towards a classical system. The practical implementation of a quantum computer relies on a physical system that constitutes a good approximation of a two-level system. Among all the systems proposed so far, neutral atoms present the advantage of well controlled manipulations of the internal and external degrees of freedom. Furthermore, neutral atoms offer built-in scalability when the atoms are trapped in periodic potentials. Following this route, we have chosen to encode the quantum information
259
260 on two hyperfine states of a single rubidium atom trapped in an optical tweezer. Using this system, it has been demonstrated that several tweezers can be arranged in an array, with each sites well localized and adressable by optical methods2i3 . This paper describes how we trap and observe a single atom in an optical tweezer created by focusing a far-off resonant laser down to a micrometer size waist. We then show the coherent manipulation of the two-level system and characterize the coherence of this quantum bit. As a first step towards the controlled interaction between two atoms trapped in an array of tweezers, we demonstrate a scheme where the qubit is transfered between two tweezers, with no observed loss of coherence and no change in the external degrees of freedom of the atom. Additionnally, we move the atom over distances that are typical of the separation between atoms in an array of optical traps, and show that this transport does not affect the coherence of the qubit. Finally, we are working towards the conditional entanglement of two atoms trapped in tweezers separated by 10 microns. We have shown two key ingredients of this protocol: the controlled emission of a single photon by a single atom and the two-photon interference of photons emitted by two atoms. 1. Diffraction-limited optics for single-atom manipulation
Our optical tweezer is a far off-resonance dipole trap, with a size of about one micrometer. We produce this tweezer by focusing a laser beam down to the diffraction limit of a large numerical aperture aspherical lens. We use a lens manufactured by LightPath Technologies with a numerical aperture of 0.5. The working distance between the lens and the focal point is 6 mm, large enough to allow a good optical access around the lens. We have separately tested the lens using a wavefront analyzer. We have found a RMS deviation of the resulting wavefront of less than X/30 over the whole numerical aperture, thus demonstrating that the aspherical lens is diffraction limited. The optical layout, consisting of the aspherical lens and standard optical elements, is shown in figure 1. The dipole trap laser beam, produced by a 850 nm laser diode, is sent throught a single mode optical fiber and is shaped a t the output of the fiber by a triplet lens. The beam goes through the viewport of the vacuum chamber inside which the aspherical lens is placed. Before putting this system together, we have checked on a separated bench that the whole optical system is diffraction limited (see reference4 for more details).
261
! Dichroic *!^S plate Doublet £
tes
i FocaS
1p«|
Pinhole
Singlet
f
<
Vacuum chamber
>
• optical tweezer (850 nm) imaging system (780 nm}
Fig. 1. Optical setup of the single-atom trapping (solid line) and imaging systems (dotted lines).
The same lens is used to collect the fluorescence light (780 nm) emitted by the atom trapped in the tweezer. This fluorescence light is collected outside the vacuum chamber using a confocal imaging system, as represented in figure 1. The light is sent onto an avalanche photodiode, used in counting mode, and a CCD camera. This imaging system is also diffraction limited and has a spatial resolution of 1 micrometer4 . We use this optical system to trap and observe single atoms in the tweezer. The tweezer is loaded from the cloud of rubidium atoms cooled in an optical molasses. Figure 2 shows an example of the signal obtained on the avalanche photodiode versus time. The steps in the signal correspond to the fluorescence of a trapped atom at 780 nm, induced by the molasses laser. The absence of double steps is an indication that only individual atoms are trapped. This single-atom trapping is made possible by a "blockade mechanism" studied in detail in reference5 . 2. Single-atom quantum bit
Our quantum bit is encoded on the |0) = \F — 1,M = 0), |1) = \F = 2, M = 0} hyperfine sublevels of a rubidium 87 atom. This choice of levels for the qubit provides the advantage of zero first-order sensitivity to magnetic fields. We initialize the qubit in state |0) by optical pumping, with an efficiency of 85 %. We read the state of the qubit using a state selective measurement limited by the quantum projection noise. For this purpose, we send a laser on resonance with state |1) and the state F' = 3 connected by a transition
262 50000-
41000
41500
time (ms)
42000
histogram
Fig. 2. (a) Fluorescence of a single atom measured by the avalanche photodiode. (b) Histogram of the measured fluorescence recorded over 100 sec. The two well separated peaks in this histogram correspond to the absence of atom and the presence of exactly one atom.
at 780 nm. Radiation pressure expels the atom out of the trap, if the atom is initially in state |1). Otherwise, if the atom is in state |0), the laser leaves the atom unaffected. We then check for the presence of the atom. This method therefore maps the internal state of the atom on the presence or the absence of the atom at the end of the sequence. We drive the qubit transition with two Raman lasers, one of them being the dipole trap. The two beams are colinear and sent through the same optical fiber and the large numerical aperture lens. Due to the tight focusing of the two lasers, we observe a Rabi frequency of the two-photon transition as high as 2?r x 6.7 MHz. Figure 3 shows the population of state |0) for two durations of the Raman pulse6 . We study the coherence of the quantum bit using Ramsey spectroscopy. We apply a first Tr/2 Raman pulse to prepare the atomic state (|0} + |l})/-\/2We let the system evolve and we apply a second ?r/2 pulse. The signal exhibits oscillations at a frequency given by the detuning of the Raman lasers with respect to the qubit transition. The decay of the contrast of the oscillations as a function of the time interval between the two pulses is the signature of the loss of coherence. We attribute this decay to the residual motion of the atom in the trap, together with the fact that the two states of the qubit experience a slightly different trapping potential, leading to a dephasing of the quantum bit. Our best 1/e dephasing time is 630 //sec.
263
a"
oi...
50 '
100
150
200
250
300
Pulse length (ps)
Pulse length (ps)
Fig. 3. Single-atom Rabi oscillations. Fraction of atoms in F = 1, as a function of the Raman pulse length, at low (a) and high (b) intensity. Details can be found in reference6 .
For details, see reference6 . We rephase this dephasing by inserting a 7r pulse between the two 7r/2 pulses. Using this spin-echo technique, we observe a revival of the oscillations after a time as long as 40 ms, corresponding to a 70 fold improvement with respect to the coherence time of the qubit, as shown in figure 4.
3. Transport and transfer of atomic qubits Neutral atoms are promising candidates for the realization of a large-scale quantum register. To perform a quantum computation, a key feature is the ability to perform a gate between two arbitrary qubits of the register. As a first step, we have demonstrated a scheme where the atomic qubit is transfered between two tweezers, and then transported over several tens of micrometers. We show that these manipulations of the external degrees of freedom preserve the coherence of the qubit, and do not induce any heating. The distance travelled is typical of the separation between atoms in an array of dipole traps. These techniques can also be useful to position an atom at the node of the electromagnetic field in an optical cavity for QED experiments7 . To show that the transfer preserves the coherence of the quantum bit, we prepare a superposition ( 1 0) Il))/JZin a first tweezer using a first 7r/2 pulse. We then decrease the depth of the first tweezer while increasing the
+
264
60 40
s?
0
100
004
002
006
008
01
996
998
1000
1002 1004
1006
v 7
80
II
60 ._
6
._ 40 L
m -
2
20
0
a
0 0
1
2
4
3
5
6
7
8
9
10
Time (ms)
0
’.
0
,
I
I
10
20
30
40
Time between xi2 pulses (ms) Fig. 4. Example of the spin-echo signal. Figure (a) shows the revival of the oscillations after the 7r pulse has been applied. Figure (b) shows the amplitude of the echo signal for different durations of the spin-echo sequence.
depth of a second tweezer superimposed with the first one. After a dwelling time of 200 psec, we transfer the atom back to the initial tweezer and apply a second 7r/2 pulse. We vary the time between the two pulses around 200 psec and observe the corresponding Ramsey oscillations. We have shown that the amplitude of the Ramsey signal remains unchanged whatever the depth of the second trap is. We have also checked that when the depths of the two traps are identical, the “temperature” of the atom is unchanged. We have also moved the tweezer when the qubit is prepared in a superposition (10) Il))/fi. We have transported the atom up to f9 pm
+
265
away from the axis of the focusing lens, and brought it back to its initial position8 . The entire round trip takes 6 ms. As this time is longer than the dephasing time of the quantum bit, we apply a TT pulse when the tweezer is at its turning point, as shown in figure 5. We have measured that the ampli-
(a)
Initial position
Target position
-§
(b) 1.0oj
0.8 -
"5. 0.6 10
I °-4 ai
I °-2 ^ 0.0-1 - 8 - 6 - 4 - 2 0 2 4 6 8 Trap displacement {pm)
Fig. 5. (a) Principle of the moving qubit experiment, with the position of the tweezer when the various pulses of the sequence are applied, (b) Amplitude of spin-echo signal versus the amplitude of the displacement.
tude of the spin-echo signal remains constant when we varied the amplitude of the displacement. This indicates that the coherence of the quantum bit
266 is preserved during the transport. Finally, we have measured the phase of the Ramsey sequence during the transfer experiment and the phase of the spin-echo signal during the transport. We have modelled this phase evolution and found a good agreement with the result of the experiment. This understanding of the phase is crucial for a possible implementation in a quantum computer where qubit phases need to be controlled.
4. Towards conditional entanglement of two atoms The ability to generate entanglement is a key feature that any physical system should exhibit to be useful for any quantum information processing task. For example, entanglement is essential in teleportation protocols between two parties. Entanglement is also a necessary ingredient in the two approaches to quantum computing. In the quantum circuit approach' , entanglement is generated during the course of the implementation of the algorithms. In the cluster state approachg , it even constitutes the starting point of the calculation. It is therefore important to be able to generate and control the entanglement in this quantum system. Among all the methods proposed to entangle two neutral atoms, conditional entanglement based on photon emission is promising, as it does not require any direct interaction between the atoms. A simplified scheme is presented in figure 6 (see for example referenceslo>l1).We isolate three levels in the atom. The upper level is connected t o the two logic levels by an optical transition, with frequencies v1 and 7 4 . Each atom can decay to level 10) and 11) by emitting a photon with equal probability. The scheme entangles the internal state of the atom with the frequency of the emitted photon, generating for atom A the state (IOA, v1) 11A, v 2 ) ) / a .The state of the two photons and two atoms A and B is therefore
+
1
1
16)= -(IOA,
Jz
v1)
+ 11A, v2))8 -Jz (\oB,
v1)
+ IlB, v2)) .
Re-arranging the terms, this state can be re-written as
267
When the two photons are recombined on a beam-splitter, a two-photon interference prevents coincidences on the two detectors when the two-photon state before the beam-splitter is z'l, z'l) and |z/2, z^}- The absence of simultaneous coincidence is also true when the two-photon state is v\, v^} +11/2, v\), as the two-photon amplitudes of each component exactly cancel out (see for example reference11). Therefore a simultaneous detection event on the two detectors heralds the preparation of the entangled atomic state (\QA-, IB) — |lA>O.B))/V / 2- In this scheme, two-photon interferences acts as a "filter", and the preparation is heralded by the double detection. AtomB
Atom A
-
Rg
Detector 1 ,>
Detector 2
Fig. 6. Principle of a conditional entanglement of two atoms A and B based on the entanglement between each atom and an emitted photon followed by the recombination of the two photons on a 50/50 beam-splitter (BS).
Three ingredients are required to implement this protocol. The first one is the triggered emission of a single photon by a single atom. The second one is the observation of two photon interferences, and the third one is the ability to entangle an atom with an emitted photon. In the following sections we describe our implementation of the two first steps. We note that the third step has been realized by two groups in the recent years12'13 . 5. Single atom as a single-photon source
We control the emission of single photons by a single atom placed at the focal point of a large numerical aperture lens by sending -K pulse on the optical transition connecting (F = 2,M = 2) and (F' = 3,M' = 3), see figure 7. The duration of the pulses is 4 ns, and the separation between the pulses is 200 ns. The emitted photons all have the same
268
We collect 0.6 % of the emitted photons on an avalanche photodiode. We have characterized the single-photon nature of the source by measuring the coincidences on two photodetectors placed in the imaging system14 . The resulting curve is shown in figure 7(c). The absence of coincidence at zero delay is the signature that single photons are emitted by the atom. A careful analysis of this curve shows that the probability that the source emits two photons following the same excitation pulse is 1.8%. This number is in good agreement with a calculation taking into account the 4 ns duration of the excitation pulse, which is not negligible with respect to the 26 ns lifetime of the upper state.
(b) Atom
Counter K- pulse // 780 nm //
lens Excitation pulse (C)
500
g 400 •g 300 "3 .£ 200 o ° 100
0
-1000 -500 0 500 1000 1500 Delay (ns)
Fig. 7. (a) Principle of the single-photon source based on an atom trapped at the focal point of a large numerical aperture lens, (b) Relevant hyperfine levels of rubidium 87 used in the experiment, (c) Histogram of the coincidences measured on the two single-photon counters.
6. Interference of two photons emitted by two atoms
When two indistinguishable single photons are fed into the two input ports of a beam-splitter, the photons will leave together from the same output
269
port. This is a quantum interference effect, which occurs because the two possible paths, where the photons leave in different output ports, interfere destructively. This effect was first observed in parametric downconversion by Hong, Ou and Mandel15 . We have shown the interference of two photons emitted by two atoms trapped in two tweezers separated by 6 microns (see details in reference16). The two atoms are excited by the same 2 ns-laser pulse following the procedure described in the previous section. We recombine the two photons of same polarization on an optical setup equivalent to a 50/50 beam-splitter and we measure the coincidences at zero delay on two photodiodes placed in the outpout ports of the beam-splitter, as represented in figure 8(a). If the two-photon interference were perfect, one should not measure any coincidence as the two photons must leave in the same output port.
(a)
•BS
Pulsed excitation -100
r
T
0
i
100
Translation (\im)
Fig. 8. (a) Principle of the two-photon interference experiment. The photons emitted by two atoms A and B are recombined on a 50-50 beam-splitter (BS), and the coincidences are measured using two single-photon detectors (b) Influence of the wavefront matching. We plot the amplitude of the residual coincidence signal at zero delay for various relative distance between the two photon modes, translated parallel to each other. For perfect interference, the signal should go all the way down to zero when the two modes are not translated.
To analyze the visibility of the interferences, we varied the spatial overlap between the photons propagating in free space. In order to do so, we translated the spatial mode of one photon with respect to the other one. The result is represented in figure 8(b). For our best overlap, we find a visibility of the interferences of 60%, coming from the difficulty to mode-match the two photons. This number is a measure of the indistinguishability of the two interfering photons. This mode-matching can be improved by coupling the photons to single mode fibers.
270
7. Conclusion We have demonstrated some basic manipulations of a single quantum bit encoded on a n atom trapped in an optical tweezer. We have shown internal state rotation at the single atom level and we have measured the internal coherence of the qubit state. We have also demonstrated two necessary ingredients of a conditional entanglement protocol. Our current estimate gives an efficiency t o produce one entangled pair on the order of 10-6-10-5. We anticipate a rate of entangled pair production of one every 100 seconds.
Acknowledgments: We acknowledge financial support from IFRAF, ARDA/DTO and the European Integrated project SCALA. LCFIO is CNRS UMR8501. M.P.A. Jones and A.M. Lance are supported by Marie Curie Fellowships. A. Gaetan is supported by a DGA Fellowship.
References 1. M.A. Nielsen, & I.L. Chuang Quantum Computation and Quantum Information. Cambridge University Press (2000). 2. S. Bergamini, et al., J . Opt. SOC.A m . B 21, 1889 (2004). 3. R. Dumke, et al., Phys. Rev. Lett. 89, 097903 (2002). 4. Y.R.P. Sortais, et al., Phys. Rev. A 75, 013406 (2007). 5. N. Schlosser, G. Reymond and P. Grangier, Phys. Rev. Lett. 89, 023005 (2002). 6. M.P.A. Jones, J. Beugnon, A. Gaetan, J. Zhang, G. Messin, A. Browaeys, P. Grangier, Phys. Rev. A 75, 013406(R) (2007). 7. S. Nussmann, M. Hijlkema, B. Weber, F. Rohde, G. Rempe, and A. Kuhn, Phys. Rev. Lett. 95, 173602 (2005). 8. J. Beugnon, et al., in press Nature Physics (2007); arXiv:0705.0312 [quantPhl . 9. R. Raussendorf and H. Briegel, Phys. Rev. Lett. 86, 5188 (2001). 10. C. Simon, W.T.M. Irvine, Phys. Rev. Lett. 91, 110405 (2003). 11. D. L. Moehring, M. 3. Madsen, K. C. Younge, R. N. Kohn, Jr., P. Maunz, L.-M. Duan, C. Monroe and B.B. Blinov, J . Opt. SOC.A m . B 24, 300 (2007). 12. B.B. Blinov, D.L. Moehring, L.-M. Duan, C. Monroe, Nature 428,153 (2004). 13. J. Volz, M. Weber, D. Schlenk, W. Rosenfeld, J. Vrana, K. Saucke, C. Kurtsiefer and H. Weinfurter, Phys. Rev. Lett. 96, 030404 (2006). 14. B. Darqui6, et al., Science 309, 454-456 (2005). 15. C.K. Hong, Z.Y. Ou, and L. Mandel, Phys. Rev. Lett. 59, 2044-2046 (1987). 16. J. Beugnon, et al., Nature 440, 779 - 782 (2006).
PROGRESS IN ATOM CHIPS AND THE INTEGRATION OF OPTICAL MICROCAVITIES E. A. HINDS,* M. TRUPKE, B. DARQUIE, J. GOLDWIN and G. DUTIER Center for Cold Matter, Imperial College London, Prince Consort Road, London, SW7 ZAZ, UK *E-mail: ab-ed. hindsQimperial.ac.uk http://www3. imperial. ac. uk/ccm/ We review recent progress at the Centre for Cold Matter in developing atom chips. An important advantage of miniaturizing atom traps on a chip is the possibility of obtaining very tight trapping structures with the capability of manipulating atoms on the micron length scale. We recall some of the pros and cons of bringing atoms close to the chip surface, as is required in order to make small static structures, and we discuss the relative merits of metallic, dielectric and superconducting chip surfaces. We point out that the addition of integrated optical devices on the chip can enhance its capability through single atom detection and controlled photon production. Finally, we review the status of integrated microcavities that have recently been demonstrated at our Centre and discuss their prospects for future development. Keywords: Atom chip; Integrated optics; Cavity QED; Quantum information.
1. Introduction
Atom chips offer a powerful way to miniaturize experiments in atomic physics.' Microstructures on the surface of the chip produce magnetic and electric fields which can be used to confine and manipulate cold atoms above the substrate surface. Experiments have shown that it is possible to transport and split cold neutral atom clouds or BoseEinstein condensates in tight traps near the surface of a chip using either microscopic patterns of permanent magnetization in a film or microfabricated wire structures carrying currents or charges. Large field gradients can be generated close to a microstructured surface, creating tight traps with oscillation frequencies up to 1MHz at the micron scale where tunneling and coupling between traps become important. This makes atom chips suitable for applications in matter-wave interferometry,' quantum ~ e n s i n g atomic ,~ quantum N
271
272 information p r o ~ e s s i n gand , ~ the study of low-dimensional quantum gases.6 However many of these potential applications require the ability to detect a small number of atoms, to create versatile microscopic trapping configurations, and to manipulate the atomic quantum state, both internal and motional. Optical methods exist to accomplish these tasks, but are only just now starting to be miniaturized and integrated into atom chips. One microscopic optical device we are studying is an atom detector based on a pair of optical fibres facing each other. These single mode fibres have been machined at their ends, providing small beam waists a t the foci. Light transmitted through one fibre and collected by the other should allow detection of 5 to 10 atoms placed in the gap,7 either through their absorption (on-resonance imaging) or by the optical phase shift (dispersive imaging). The fibres can also be used to generate a one-dimensional standing-wave pattern, making an optical lattice in which the atoms can be manipulated by the optical dipole force. However, in order to bring the optical sensitivity down to the single atom level, we require an optical cavity to increase the atom-light interaction strength. The integration of microcavities with the trapping and guiding capabilities of atom chips opens new possibilities for experiments in atomic physics. In an optical microresonator, the interaction strength of an atom with a photon in the cavity mode increases with decreasing mode volume, as 1/fi.8 This is evident in the expression for the vacuum Rabi frequency in the cavity,
where p is the dipole moment of the atomic transition and wc is the resonance frequency of the cavity. This interaction is damped by the decay rate of the excited atomic population, which is 27 in free space, and of power in the cavity, which is 2 6 = m / ( L F ) .Here L is the cavity length and F is the finesse of the resonator. The driving and damping rates are succinctly compared in the single atom cooperativity, C = g2/2&,y. The total spontaneous emission rate of an atom in a cavity can be made to differ greatly from the free-space value, depending on the cooperativity. With the cavity tuned to the atomic transition frequency and with 6 < g, the natural decay rate becomes 2y(l 2C). The second term represents emission of photons into the cavity mode at a rate 4Cy = 2g2/r;. The resonators used successfully in other experiments have reached single-atom cooperativity values exceeding 50,’ but are not ideal for atom chip experiments as they consist of comparatively large components, mak-
+
273 ing it impossible to bring the intense part of their mode close t o the trapping structures on the chip surface. For this reason, research is being carried out on a variety of integrable microcavity designs. These include microtoroids, fibre-fibre microcavities and our fibre-chip cavities. Microtoroids offer the highest quality factors among these, and are well-suited for scalable atomchip integration as they can be produced by standard silicon microfabrication techniques. Strong coupling between single atoms and the field of a microtoroid resonator has been demonstrated" with atoms passing through the evanescent field in free-fall. To couple light into and out of these resonators, a tapered fibre is used. The fibre position needs to be adjusted so that the evanescent field at the taper overlaps with the evanescent component of the microtoroid mode. This is the only manual procedure required to make the devices operational. While the strong-coupling condition has been experimentally fulfilled for the first time for such a device, the challenge of reliably positioning and trapping atoms in the evanescent field of these devices with the required accuracy has yet to be surmounted. By contrast, atoms can be placed directly and accurately into the region of highest field strength of Fabry-Perot-type resonators. For this reason, the efforts of several research groups are currently focussed on this type of resonator. Several experimental groups have already succeeded in positioning atoms accurately within the mode of an optical resonator using optical, electrostatic and magnetic transport techniques. The positioning of a Bose-Einstein condensate in a microcavity on a chip by means of current-carrying wire guides has also been recently demonstrated using a fibre-coupled microcavity.ll The resonators used in that work are however not ideally suited for scalable integration, as they are constructed in a sequence of mostly manual steps. Here we report on recent experiments at our Centre on the detection of atoms with on-chip micocavities, whose design aims to combine the advantages of microfabrication and fibre-coupling with the open access of a Fabry-Perot resonator. In the next section we describe two typical atom chips that are in use in our laboratory to magnetically trap atoms either with current carrying wires or permanent magnet structures. Section 3 deals with detection of atoms and enhanced emission of photons using on-chip microcavities. We conclude with an outlook in Sec. 4.
274
2. Magnetic atom chips
To give a concrete example of an atom chip, we show in Fig. 1 an interferometer chip7 currently being used in our laboratory. The reflective gold surface is used to form a mirror magneto-optical trap (MOT) as a reservoir of pre-cooled atoms. The coils used for the MOT are seen in the figure. The chip consists of a 3 /zm layer of gold, thermally evaporated onto a silicon substrate, in which wires have been lithographically defined by ion-beam milling. Currents in the wires allow cold 87Rb atoms to be magnetically trapped and split, in a matter-wave analogue of an optical beam splitter.2'3 The inset in Fig. 1 is a microscope image showing the four parallel wires at the heart of the interferometer, which have a center-to-center distance of 300 /zm between the thick outer pair.
Fig. 1. Current carrying wire atom chip and chip mount used at Imperial College by Eriksson et al.7 The inset shows a high resolution microscope image of the central region of the chip where atoms are trapped and manipulated.
With metallic chips such as this one, the homogeneity and stability of cold atom clouds can be compromised close to the room temperature surface. Two main phenomena have been identified: (i) spatial imperfections of the wire cause the current to flow non-uniformly and make the atom trap rough, leading to fragmentation of the atom clouds,12'13 and (ii) thermal fluctuations of the magnetic field near the surface, caused by current noise in the conductor, drive spin flips of the atoms, thereby inducing loss and destroying quantum coherences.14 The fragmentation can be controlled by using fabrication methods that give very homogeneous wires with smooth
275 edges and uniform thickness. The magnetic noise field can be altered by adjusting the thickness of the surface and by changing the material in order to alter the complex c ~ n d u c t i v i t y . ~ ~ ~ ~ ~ We have also been using a chip based on patterns of permanent magnetization, written on commercial videotape. This exhibits very much longer spin flip times because the thermal noise currents are suppressed by the high resistivity of the videotape. A pattern of sinusoidal magnetisation (in plane), with a period of about 100pm, allows the confinement of atoms at distances less than 100 pm from the surface of the chip in an array of long, thin traps. These have a very high aspect ratio (> lo3), with tight transverse confinement that can bring ultra-cold atoms into the one-dimensional regime.l7>l8As well as offering a long spin-flip lifetime for BECs trapped near its surface,17 the permanent magnet atom chip has the benefit of low power dissipation. Some fragmentation of these videotape traps has been observed and is due primarily to inhomogeneity of the magnetic layer. Preliminary experiments on Pt/Co multilayer magnetic thin filmslg indicate that these are a promising alternative t o videotape. Other permanent magnet materials are being studied in the group of P. Hannaford, also reporting at this confererence. Superconductors offer another way to reduce the magnetic noise level and increase the spin-flip lifetimes by many orders of magnitude. Indeed, atoms have already been trapped near superconducting surfaces in two laboratories.20,21A recent calculation22 shows that rubidium atoms trapped 1 pm away from a superconducting niobium surface at liquid helium temperature should have a long spin flip lifetime, in excess of 1000s. Under these conditions, the cold atoms offer a new way to probe the superconducting surface because they are very sensitive to the local magnetic field. For example, cold atoms would be able to image the vortices of a thin superconductor close to the Kosterlitz-Thouless transition or more generally to study vortex dynamics in a type II s u p e r c o n d ~ c t o r . ~ ~ 3. Chips with optical micro-cavities
We have recently tested a new type of optical microcavity, which combines the advantages of microfabrication and fibre-coupling with the open access of a Fabry-Perot resonator. The plano-concave resonator is formed between an isotropically etched hemisphere on the surface of a silicon chip and the plane end of single-mode fibre. Both surfaces are coated with a high-reflectivity multilayer dielectric film. Our collaborators at the University of Southampton have fabricated large arrays of concave mirrors on
276
silicon chips and we have used them to build cavities24 with F > 5000 and Q > 106. More recently, we have demonstrated both the detection of atoms and the enhanced emission of photons into the mode of the cavity.25 The apparatus is shown in Fig. 2. Two fibres with plane dielectric mirrors on their ends are held in grooves on a glass-ceramic substrate, facing two of the mirrors in the silicon array. The silicon mirror chip is mounted on a piezo-electric translator (PZT), which is adjusted to tune the upper cavity to the free-space atomic resonance. In this work the lower cavity was not used. A mirror is used to form a reflection MOT directly above the cavities.
APD
Probe Fig. 2.
Fibre
Mirror chip
Schematic diagram of apparatus used in the experiment.
For these experiments, we used cavities with F = 280 and L = 133 //m, giving K = 27r x 2 GHz. The beam waist of the mode is we = 4.6 /mi, giving a vacuum Rabi frequency of 2g = 1-n x 200 MHz for 85Rb atoms at an antinode of the cavity mode, driven with circularly polarised light on the closed transition \F — 3,mp = ±3) —> \F' = 4,mp/ — ±4). This transition has a natural lifetime 27 = 2?r x 6 MHz, yielding a single-atom cooperativity of C = 0.8. The cooperativity increases linearly with the number A^A of atoms in the cavity. Since these are not generally at an antinode, the total cooperativity is Ctot = jC^,n=i A r «) where /(r n ) is the fraction of peak intensity at the position of the n th atom. The factor | accounts for an average of Clebsch-Gordan coefficients over all the F = 3 Zeeman sublevels. The effective number of atoms interacting with the cavity field is then defined as Nf = C to t/(f C).
277 3.1. Atom detection
z , n
320
'v)
m
0
--
v
300
a, 280
c
3
0
0 260 10
20
30
40
50
60
Time (ms) Fig. 3. Change in reflected probe light as atoms traverse the cavity. Laser, cavity, and atom frequencies are equal: W L = wc = W A . Average time-of-flight signal of 34 MOT 0.7 atoms in the cavity on releases, with an integration time of 2 5 0 p . There are average at the peak.
-
-
To put atoms into the microcavity, we release a cloud of 2 x lo7 85Rb atoms from the MOT, 7mm above. Weak resonant probe light (1 p W ) is incident on the cavity through a beam splitter and the reflected photons are counted by an avalanche photodiode (APD, see Fig. 2). When the cavity is far from resonance with the light, the reflected intensity is I,,, = 419 x lo3 s-l, which drops to I m i n = 272 x lo3 s-l at resonance. As displayed in Fig. 3, this rises to a peak of Iatoms = 315 x lo3 s-l when the atoms pass through the cavity. From these three count rates we derive the peak cooperativity of the atom-cavity system using the relation
+
where Ptot= 2CtOt 1 is an effective Purcell factor, and v = 1- ,/is the empty cavity fringe visibility. From our experiments, we determine a total cooperativity of Ctot 0.23 at the peak of the atom signal, corresponding to an effective atom number of only Nzff = 0.7. We have measured the change in the peak reflected intensity when we scan the detuning of the probe light ALA= (WL - w ~ ) / ykeeping , the cavity frequency equal to the atomic resonance frequency w c = W A . Figure 4
-
278
0.701, -10
,
, 0
,
,I
10
E -10
0
10
Detuning (MHz)
Fig. 4. Peak fraction of power reflected versus detuning ALA when w c = W A for two different initial MOT numbers. The fits (solid lines) yield average atom numbers of (a) (Niff)= 1.1 (b) (Niff)= 0.6.25
displays, for two different values of the initial atom number in the MOT, Iatoms/Imax versus ALA,for which we expect
assuming g/y >> ( ~ , A L A The ) . solid lines in Fig. 4 are fits to the data, taking into account fluctuations of Ctot and the probe laser l i n e ~ i d t h . ' ~ These lineshapes confirm that we have better than single atom sensitivity.
3.2. Noise suppression The arrival of atoms in the cavity is signaled not only by an increase in the reflected light level, but also by a decrease in the intensity noise. To measure this, we dropped the cloud 48 times, recording the reflected intensity in 10 ps bins. We then determined for each bin the ratio of variance to mean, taken over the 48 drops. After correcting our count rate for the measured APD dead-time of 4 4 n ~ we , ~found ~ the correct variance-to-mean ratio f ( n ) V a r ( n ) / ( n at ) each time bin ( n being the counts already corrected for the dead-time) by including the 10% loss at the beamsplitter of Fig. 2 and the 60% loss due to the quantum-efficiency of the APD, according to fcorr(n)1 = [ f ( n )- 1]/0.54. The resulting values, plotted in Fig. 5, show a strong
279
20
30
40
50
Time (ms) Fig. 5. Ratio of variance to mean for photon count rate reflected from the cavity during passage of atom cloud (corrected for APD dead-time and losses). Vertical dashed line: arrival time of the peak number of atoms. Horizontal line at 1.0: shot noise level. The sample size is 48 atom cloud drops.
reduction of the noise during the passage of the atom cloud. In addition to the photon shot noise, intensity fluctuations are due to small shifts of the cavity length away from its resonance. The effect of these mechanical fluctuations is expected to be less pronounced when atoms are present, and this partially explains the decrease. However many of our experiments seem to show additional noise reduction, the origin of which requires further investigation.
3.3. Photon generation
A further effect observed with our microcavity is the enhancement of spontaneous emission. For this experiment, we once again release a cloud of atoms into the resonator. When the atom number in the cavity reaches its highest value, we switch on a resonant excitation beam, aimed at the atoms from below the cavity. This beam pushes the atoms out of the cavity and also pumps them into the dark F = 2 state. The loss of interacting atoms produces a sharp drop in the cavity reflection signal, as shown in Fig. 6 (a). We then perform the same experiment but without any probe light. This time, a photon peak is recorded at the moment the excitation laser is turned on, as shown in Fig. 6 (b). These are fluorescence photons
280 I
I I
,
,
,
,
’
1.o 0 10
20
30
40
50
60
Time (ms) Fig. 6. (a) Sharp drop in the cavity reflection signal due to atom loss and optical pumping when the excitation laser is turned on. (b) Cavity-enhanced spontaneous emission collected via the fiber when the atom cloud is excited. The dotted line shows coincidence of the photon pulse with the turn-on of the excitation laser.
emitted into the cavity mode at the Purcell-enhanced rate 4CtOty,as mentioned in the introduction, and transported to the APD through the fibre. This is confirmed by the fact that no photons are collected when the cavity is detuned from the atomic transition or when there are no atoms in the cavity. 4. Outlook
These experiments have shown that this type of microresonator is suitable for the detection of numbers of atoms smaller than one, and can be used t o generate photons by enhanced spontaneous emission. The latter is of particular interest for quantum-cryptography and quantum-computation schemes. These cavities have by no means yet reached their full potential. In the laboratory, with materials of the same type, we have created cavities with F > 500 and I m i n / l m a x < 15%. Both values are improvements by approximately a factor 2 compared to the published results, which will considerably increase the detection signal, as can be deduced from Eq. (2). We had previously shown that we can even achieve finesse values of F > 5000,
281
albeit with a low fringe-extinction of I m i n / l m a x 5 At present, these values are limited by scattering losses caused by the surface roughness of the silicon mirror-substrate. Working with the group of M. Kraft at the university of Southampton, we expect in the near future to have smoother cavity mirrors that can operate well within the strong-coupling regime of cavity quantum-electrodynamics, in which the coherent exchange of quantum information between atoms and photons in the cavity becomes possible. These microcavities have the virtues that they are open, giving atoms direct access to the peak of the mode pattern, they couple the light in and out in a simple way through an integrated fiber, and they can be fabricated in large numbers. Being integrated into a silicon substrate, they can be readily combined with other micromachined components on atom chips. Some examples include the magnetic trapping and guiding structures described in Sec. 2, which allow the deterministic transport of atoms in and out of the cavity, integrated actuators for tuning the resonator,26 or pyramidal micro-mirrors to realize an array a of single-atom sources from an array of microscopic M O T S . Work ~ ~ is in progress to produce a fully integrated device above which neutral atoms can be trapped, guided, detected and manipulated in situ, using magnetic and optical fields.
Acknowledgments This work was supported by EU networks Atom Chips, Conquest, and SCALA, by the Royal Society, and by EPSRC grants for QIPIRC, CCM programme and Basic Technology. The atom chip devices in use at Imperial College were fabricated by the nanosystems group of Professor M. Kraft a t the University of Southampton.
References 1. J. Fortagh and C. Zimmermann, Rev. Mod. Phys. 79,235 (2007). 2. T. Schumm, S. Hofferberth, L. M. Andersson, S. Wildermuth, S. Groth, I. Bar-Joseph, S. J. and P. Kriiger, Nut. Phys. 1, 57 (2005). 3. E. A. Hinds, C. J . Vale and M. G. Boshier, Phys. Rev. Lett. 86, 1462 (2001). 4. P. Treutlein, P. Hommelhoff, T. Steinmetz, T. W. Hansch and J. Reichel, Phys. Rev. Lett. 92, 203005 (2004). 5. T. Calarco, E. A. Hinds, D. Jaksch, J. Schmiedmayer,J. I. Cirac and P. Zoller, Phys. Rev. A 61, 022304 (2000). 6. L. Pricoupenko, H. Perrin and M. Olshanii (eds.), Quantum Gases in Low Dimensions, J. Phys. IV France Vol. 116, (EDP Sciences, 2004). 7 . S. Eriksson, M. Trupke, H. F. Powell, D. Sahagun, C. D. J. Sinclair, E. A.
282
8.
9. 10.
11. 12. 13. 14.
15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27.
Curtis, B. E. Sauer, E. A. Hinds, Z. Moktadir, C. 0. Gollasch and M. Kraft, Eur. Phys. J D 35,135 (2005). P. R. Berman (ed.), Cavity Quantum Electrodynamics Advances in Atomic, Molecular and Optical Physics, Supplement 2, (Academic Press, New York, 1994). K. J. Vahala, Nature 424,839 (2003). T. Aoki, B. Dayan, E. Wilcut, W. P. Bowen, A. S. Parkins, T. J. Kippenberg, K. J . Vahala and H. J . Kimble, Nature 443,671 (2006). Y. Colombe, T. Steinmetz, G. Dubois, F. Linke, D. Hunger and J. Reichel, arXiv:0706.1390 (2007). M. P. A. Jones, C. J. Vale, D. Sahagun, B. Hall, C. C. Eberlein, B. E. Sauer, K. Furusawa, D. Richardson and E. A. Hinds, J . Phys. B 37,L15 (2004). Z. Moktadir, B. Darqui6, M. Kraft and E. A. Hinds, J . Mod. Opt. 54,2149 (2007). M. P. A. Jones, C. J. Vale, D. Sahagun, B. V. Hall and E. A. Hinds, Phys. Rev. Lett. 91,080401 (2003). P. K. Rekdal, S. Scheel, P. L. Knight and E. A. Hinds, Phys. Rev. A 70, 013811 (2004). S. Scheel, P. K. Rekdal, P. L. Knight and E. A. Hinds, Phys. Rev. A 72, 042901 (2005). C. D. J . Sinclair, E. A. Curtis, I. L. Garcia, J. A. Retter, B. V. Hall, S. Eriksson, B. E. Sauer and E. A. Hinds, Phys. Rev. A 72,031603 (2005). C. D. J. Sinclair, J . A. Retter, E. A. Curtis, B. V. Hall, I. L. Garcia, S. Eriksson, B. E. Sauer and E. A. Hinds, Eur. Phys. J . D. 35,104 (2005). S. Eriksson, F. Ramirez-Martinez, E. A. Curtis, B. E. Sauer, P. W. Nutter, E. W. Hill and E. A. Hinds, Appl. Phys. B 79,811 (2004). T. Nirrengarten, A. Qarry, C. ROUX,A. Emmert, G. Nogues, M. Brune, J.-M. Raimond and S. Haroche, Phys. Rev. Lett. 97,200405 (2006). T. Mukai, C. Hufnagel, A. Kasper, T. Meno, A. Tsukada, K. Semba and F. Shimizu, Phys. Rev. Lett. 98,260407 (2007). U. Hohenester, A. Eiguren, S. Scheel and E. A. Hinds, Phys. Rev. A 76, 033618 (2007). S. Scheel, R. Fermani and E. A. Hinds, Phys. Rev. A 75,064901 (2007). M. Trupke, E. A. Hinds, S. Eriksson, E. A. Curtis, Z. Moktadir, E. Kukharenka and M. Kraft, Appl. Phys. Lett. 87,211106 (2005). M. Trupke, J. Goldwin, B. Darqui6, G. Dutier, S. Eriksson, J. Ashmore and E. A. Hinds, Phys. Rev. Lett. 99,063601 (2007). C. Gollasch, Z. Moktadir, M. Kraft, M. Trupke, S. Eriksson and E. A. Hinds, J . Micromech. Microeng. 15,S39 (2005). M. Trupke, F. Fernandez-Ramirez, E. A. Curtis, J. P. Ashmore, S. Eriksson, E. A. Hinds, Z. Moktadir, C. Gollasch, M. Kraft, G. V. Prakash and J . J . Baumberg, Appl. Phys. Lett. 88, 071116 (2006).
SINGLE ATOMSAND QUANTUM OPTICS I1
This page intentionally left blank
QUANTUM OPTICS WITH SINGLE ATOMS AND PHOTONS H. J. Kimble Norman Bridge Laboratory of Physics MC 12-33, California Institute of Technology, Pasadena, CA 91125, USA www.its.caltech. edu/Nqoptics/ An overview of research in the Caltech Quantum Optics Group is presented. We are attempting t o harness atomic radiative processes “one-by-one” and thereby to achieve full quantum control of the optical field. Applications include the realization of a quantum interface between matter and light, and the implementation of quantum networks for quantum information science and quantum many-body physics.
1. Introduction
The general theme of research in the Caltech Quantum Optics Group is an investigation of the dynamics of open quantum systems in manifestly quantum (i.e., nonclassical) domains. The research seeks to develop the means for the deterministic control of the evolution of open quantum systems in a prescribed fashion at the level of the individual constituents. Within this setting, the investigations relate to strong coupling in optical physics whereby nonlinear interactions require only single atomic and field excitations. In qualitative terms, the goal is to move beyond traditional optical physics into a new regime with dynamical processes involving single quanta taken one by one. Three experiments are discussed, namely cavity quantum electrodynamics (cavity QED) with a conventional Fabry-Perot resonator, cavity QED with lithographically fabricated microresonators, and quantum information processing with atomic ensembles. This contribution is focussed specifically on research in the Caltech Quantum Optics Group and, with apology, does not provide an overview of the many exciting advances being made by other groups around the world. At the 2007 ICOLS, there were several presentations related to the themes discussed here, t o which the reader is referred. Included are amazing re-
285
286
sults in cavity QED presented by S. Haroche from ENS, as well as new experiments in this area described by E. Hinds from Imperial College and D. Meschede from Hannover. Likewise, landmark experiments involving entanglement and teleportation with atomic ensembles were reported by E. Polzik from the Niels Bohr Institute and future prospects by E. Giacobino from the Laboratoire Kastler Brossel. 2. Cavity QED with single trapped atoms
A long-standing ambition in the field of cavity quantum electrodynamics has been to trap single atoms inside high-Q cavities in a regime of strong coupling. A critical aspect of this research is the development of techniques for atom localization that are compatible with strong coupling and that do not interfere with cavity QED interactions.’>’ Our group first achieved this goal in 2003 for single Cs atoms stored in a state-insensitive, intracavity far-off resonance trap (FORT).3 The observed lifetime of 3 s represented an advance of lo3 - lo4 beyond the first realizations of trapping in cavity QED.4,5Unlike typical FORTS, where the AC-Stark shifts for excited and ground states are of opposite sign, our trap3 causes only small shifts to the relevant transition frequencies, thereby providing advantages for coherent state manipulation of the atom-cavity system. The capabilities reported in Ref.3 have enabled a series of advances our group, which are reviewed in Ref.‘ An essential component of this work has been the development of cooling schemes, which lead to atomic localization Ax:,,i,~ = 33 nm, Aptransverse Y 5.5 pm,’ and more recently, to AxazialN 8.3 nm.8 2.1. Photon blockade
Sufficiently small metallic and semiconductor devices at low temperatures exhibit “Coulomb blockade,” whereby charge transport through the device occurs on an electron-by-electron basis. In 1997, Imamoglu et al. proposed that an analogous effect might be possible for photon transport through an optical system by employing photon-photon interactions in a nonlinear optical ~ a v i t y . ~ In Ref.1° we reported the first observation of photon blockade, in our case, for the light transmitted by an optical cavity containing one atom strongly coupled to the cavity field.ll For coherent excitation at the cavity input, we investigated the photon statistics for the cavity output and demonstrated the nonclassical character of the transmitted field, which exhibited both photon antibunching and sub-Poissonian photon statistics. We
287
observed modulation of the cavity transmission arising from the oscillatory motion of the atom trapped within the cavity mode, which allowed us to infer a maximum kinetic energy E/ks ~ 250 ^K.
2.2. Control of the center-of-mass motion in cavity QED In recent years, great strides have made by many groups to localize atoms for strong coupling to the fields of optical cavities. However, until our work in Ref.,8 it had not previously been possible to access the quantum regime for the atomic center-of-mass motion in cavity QED. In Ref.8 we achieved localization to the ground state of motion for one atom trapped in an optical cavity in a regime of strong coupling. Resolved sideband cooling to the ground state was accomplished with a coherent pair of intracavity Raman fields. To deduce the resulting state of atomic motion, we introduced a new scheme for recording Raman spectra by way of the strong interaction of the atom with a resonant cavity probe. Our scheme is the cavity QED equivalent of state detection in free space by quantum-jump spectroscopy and achieved a confidence level for state discrimination > 98% in 100 /is. From these Raman spectra, we inferred that the lowest vibrational level n = 0 of the axial potential with zero-point energy huja/2kB = 13 fj,K was occupied with probability PQ — 0-95 for one trapped atom. We are working now to utilize this advance to investigate quantum kinematics in a quantized light field. The setting is cavity QED in a domain of strong coupling for both the internal (i.e., atomic dipole and cavity field) and external center-of-mass degrees of freedom. Examples include the transfer of quantized states of atomic motion to quantum states of light, and conversely,12 as well as measurements that surpass the standard quantum limit for sensing atomic position.13
y\r...y\r Fig. 1. Illustration of the protocol of Ref.14 for quantum state transfer and entanglement distribution from system A to system B. By expanding to a larger set of interconnected cavities, complex quantum networks can be realized.
288
a) to detectors
F=3
F=4
Fig. 2. (a) Schematic of the experiment for reversible state transfer.16 The probe A(4) resonantly drives the cavity through input mirror Min; the classical field Q(t) excites the atom transverse to the cavity axis. Photons emitted from the output mirror M0ut are directed to a pair of avalanche photodiodes. (b) Atomic level diagram. Double arrow g indicates the coherent atom-cavity coupling, and Q(t) is the classical field. The cavity and O field are blue-detuned from atomic resonance by A.
2.3. Reversible state transfer between light and a single trapped atom In the initial proposal for the implementation of quantum networks,14 atomic internal states with long coherence times serve as 'stationary' qubits, stored and locally manipulated at the nodes of the network. Quantum channels between different nodes are provided by optical fibers, which transport photons ('flying' qubits) over long distances. A crucial requirement for such network protocols is the reversible mapping of quantum states between light and matter. Cavity quantum electrodynamics (QED) provides a promising avenue for achieving this capability by using strong coupling for the interaction of single atoms and photons.6 Within this setting, reversible emission and absorption of one photon can be achieved by way of a dark-state process involving an atom and the field of a high-finesse optical cavity. The basic scheme, illustrated in Fig. 1, involves a three level atom with ground states |a) and |6) and excited state e). An optical cavity is coherently coupled to the atom on the 6 <-> e transition with rate g, and a classical field fi(t) drives the atom on the a <-> e
289
transition. If the R field is ramped adiabatically off t on, then state la, n) evolves into ( b , n I), and state ( b , n ) remains unchanged, where lu,n), Ib,n) denotes a state in which the atom is in ground state a, b and there are n photons in the cavity. Ramping R on 4 off implements the reverse transformation. This process can be used to generate single photons by preparing the atom in la)and ramping R off 4 on, thereby effecting the transfer la,0) + Ib, 1) with the coherent emission of a single photon pulse from the cavity.15 A distinguishing aspect of the protocol is that it should be re~ersible,'~ so that a photon emitted from one system A can be efficiently transferred to another system B. Furthermore, it should be possible to map coherent superpositions reversibly between the atom and the field:
+
(cop)
+ c1la)) 8 10)
Ib) c 3(
d o )+ 4 1 ) ) .
(1)
Although single-photons have been generated in diverse physical systems, most such sources are not in principle reversible, and for those that are, no experiment has verified the reversibility of either the emission or the absorption process. In Ref.16 we achieved an important advance related to the interface of light and matter by explicitly demonstrating the reversible mapping of a coherent optical field to and from the hyperfine ground states of a single, trapped Cesium atom. As illustrated in Fig. 2, we map an incident coherent state A(t) with fi = 1.1photons into a coherent superposition of F = 3 and F = 4 ground states with transfer efficiency = 0.057. We then map the stored atomic state back to a field state. The coherence of the overall process is confirmed by observations of interference between the final field state and a reference field that is phase coherent with the original coherent state, resulting in a fringe visibility w, = 0.46 f 0.03 for the adiabatic absorption and emission processes. We thereby provide the first verification of the fundamental primitive upon which the protocol in Ref.14 is based. Our experiment represents an important step towards the realization of cavity QED-based quantum networks, wherein coherent transfer of quantum states enables the distribution of quantum information across the network.
<
3. Cavity QED with microtoroidal resonators
In Ref.,17 we achieved strong coupling between single Cesium atoms and the fields of a microtoroidal cavity fabricated from SiOZ." As illustrated in Fig. 3 , laser cooled atoms are dropped from above the microtoroid, with some small number of atoms transiting through the external evanescent field
290
Fig. 3. (a) Simple schematic of our experiment17 showing a cloud of cold Cs and associated trapping lasers above an array of microtoroidal resonators. Light from the probe beam Pin is coupled into a resonator by way of the fiber taper, with the forward propagating output Pp coupled into the taper, (b) Illustration of an SiC>2 microtoroidal resonator, fiber taper, and atom cloud above. The calculated field distribution for the lowest order resonator mode is shown by the color contour plot on the right. Cold Cesium atoms fall through the external evanescent field of this mode and are thereby strongly coupled to the resonator's field.
of the resonator. By recording the dependence of the forward propagating power Pp on the frequency detuning A.AC between the atom and cavity resonances for individual atom transits, we inferred the maximum accessible coupling rate ™/27r = 50 ± 12 MHz by way of comparisons to a new theoretical model.17 Our system thereby achieved the conditions for strong coupling, namely g™ > (7, K, Ai^ 1 ), where the rate of atomic spontaneous decay 7/2?r = 2.6 MHz, the cavity field decay rate K/2ir = 18 ± 3 MHz, and At PS 2/zs is the average transit duration through the evanescent field of the toroidal resonator. Our measurements are the first to demonstrate strong coupling for single atoms interacting with an optical resonator other than a conventional Fabry-Perot cavity. Our goal with this research is to realize complex quantum networks consisting of multiple atom-resonator systems. Such capabilities could have broad impact in Quantum Optics and Quantum Information Science, including the implementation of quantum networks for quantum computation
291
and communication,14'19'20 as well as quantum metrology21 23 and quantum information processing on atom chips.24 A principal motivation is also the investigation of quantum many-body systems by way of quantum networks. 4. Quantum information with atomic ensembles
In the new science of quantum information, distributed quantum networks play an important role, including for quantum computation, communication, and metrology. An approach of particular importance has been the seminal work of Duan, Lukin, Cirac, and Zoller (DLCZ) for the realization of quantum networks based upon entanglement between single photons and collective excitations in atomic ensembles.25 In recent years, critical experimental capabilities related to the DLCZ protocol have been achieved in several laboratories around the world. However, it has not previously been possible to integrate these individual capabilities into a system with the minimal functionality required for scalable quantum networks. In Ref.,26 Fig. 4. Laboratory setup26 for entanglewe achieved an important mile- ment of two pairs of atomic ensembles for the implementation of quantum communistone in this quest. As illustrated cation by way of the protocol of Ref.28 The in Fig. 4, we created, addressed, elongated ovals represent cylinders of Ceatoms, each containing approximately and controlled pairs of atomic en- sium 105 atoms, with 1 excitation shared remotely sembles at each of two quantum at a distance of 3 meters between each pair nodes, thereby leading to the first in an entangled state. realization of entanglement distribution in a form suitable both for quantum network architectures and for entanglement-based quantum communication schemes involving independent operations on parallel chains of quantum systems.25
292
In our experiment, single collective excitations in a pair of atomic ensembles are employed to encode a coherent qubit at each of two quantum nodes ( L , R ) separated by 3 meters.27 These effective qubits are remotely entangled by suitable measurement^^^ and conditional quantum evolution.28 Entanglement is verified by mapping the stored quantum states at the ( L ,R ) nodes from the respective atomic ensembles to propagating light fields. The states of these light fields, after filtering through local operations and classical communication, lead to photoelectric counting statistics that significantly violate a Bell inequality, demonstrating entanglement between the quantum nodes. Our system is thereby capable of implementing entanglement-based quantum communication protocols, such as quantum cryptography and quantum t e l e p ~ r t a t i o n . ~ ~ Importantly, entanglement is achieved between pairs of ensembles at the L , R nodes in an asynchronous fashion, with the quantum memory providing the capability to wait independently for entanglement of the two pairs.28 This capability is critical for enabling the quantum repeater architecture of Ref.25 with parallel “chains” of ensembles, as for our L , R nodes, thereby avoiding exponential resource scaling. Combined with extensions of the protocol of Ref.25 and improvements in the coherence time of the quantum memory,2g the capabilities achieved in Ref.26 can enable the robust distribution of entanglement and thereby scalable quantum networks.
5. Acknowledgements First and foremost, I gratefully acknowledge the contributions of the Caltech Quantum Optics Group, who carried out the experiments that I have described. More information about these people and their projects can be found in the references that follow and at our group website [www.its.caltech.edu/~qoptics/]. This work is supported by the National Science Foundation and by the Disruptive Technology Office.
References 1. H. J . Kimble, C. J. Hood, T. W. Lynn, H. Mabuchi, D. W. Vernooy, and J. Ye in Laser Spectroscopy: X I V International Conference, eds. R. Blatt, et al. (World Scientific, Singapore, 1999), 80. 2. H. Katori, T. Ido, and M. Kuwata-Gonokami, J . Phys. SOC.Jpn. 68,2479 (1999). 3. J. McKeever, J.R. Buck, A.D. Boozer, A. Kuzmich, H.-C.Nager1, D.M. Stamper-Kurn, H.J. Kimble, Phys. Rev. Lett. 90, 133602 (2003). 4. J. Y e , C. J. Hood, T. Lynn, H. Mabuchi, D. W. Vernooy and H. J . Kimble, IEEE Transactions on Instrumentation and Measurement 48, 608 (1999).
5. J. Ye, D. W. Vernooy, and H. J. Kimble, Phys. Rev. Lett. 83, 4987 (1999). 6. R. Miller, T. E. Northup, K. M. Birnbaum, A. Boca, A. D. Boozer, and H. J. Kimble, J. Phys. B: At. Mol. Opt. Phys. 38, S551 (2005). 7. A. Boca, R. Miller, K. M. Birnbaum, A. D. Boozer, J. McKeever, and H. J. Kimble, Phys. Rev. Lett. 93, 233603 (2004). 8. A. D. Boozer, A. Boca, R. Miller, T. E. Northup, and H. J. Kimble, Phys. Rev. Lett. 97, 083602 (2006). 9. A. Imamoglu, H. Schmidt, G. Woods, and M. Deutsch, Phys. Rev. Lett. 79, 1467 (1997). 10. K. M. Birnbaum, A. Boca, R. Miller, A. D. Boozer, T. E. Northup, and H. J. Kimble, Nature 436, 87 (2005). 11. L. Tian and H. J. Carmichael Phys. Rev. A46, R6801 (1992). 12. A. S. Parkins and H. J . Kimble, Journal Opt. B: Quantum Semiclass. Opt. 1,496 (1999); Frontiers of Laser Physics and Quantum Optics, eds. Z. Xu et al. (Springer, Berlin, 2000), 132; Phys. Rev. A 61, 052104 (2000). 13. P. Storey, M. Collett, and D. F. Walls, Phys. Rev. Lett. 68, 472 (1992) and Phys. Rev. A47, 405 (1993); R. Quadt, M. Collett, and D. F. Walls, ibid , 74, 351 (1995). 14. J . I. Cirac, P. Zoller, H. J. Kimble, and H. Mabuchi, Phys. Rev. Lett. 78, 3221 (1997). 15. J . McKeever, A. Boca, A. D. Boozer, R. Miller, J . R. Buck, A. Kuzmich, and H. J. Kimble, Science 303, 1992 (2004). 16. A. D. Boozer, A. Boca, R. Miller, T. E. Northup, and H. J. Kimble, Phys. Rev. Lett. 98, 193601 (2007). 17. T. Aoki, B. Dayan, E. Wilcut, W. P. Bowen, A. S. Parkins, H. J. Kimble, T. J. Kippenberg, and K. J . Vahala, Nature 443, 671 (2006). 18. S. M. Spillane, T. J . Kippenberg, K. J. Vahala, K. W. Goh, E. Wilcut, and H. J. Kimble, Phys. Rev. A 71, 013817 (2005). 19. L.-M. Duan and H. J. Kimble, Phys. Rev. Lett. 92, 127902 (2004). 20. H.-J. Briegel, S. van Enk, J.I. Cirac, P. Zoller, in The Physics of Quantum Information, D. Bouwmeester, A. Ekert, A. Zeilinger, Eds. (Springer, Berlin, 2000), pp. 192-197. 21. V. Giovannetti, S. Lloyd, L. Maccone, Science 306, 1330 (2004). 22. V. Giovannetti, S. Lloyd, and L. Maccone, Phys. Rev. Lett. 96, 010401 (2006). 23. S. M. Roy and S. L. Braunstein, quant-ph/0607152. 24. P. Treutlein, T. Steinmetz, Y. Colombe, P. Hommelhoff, J. Reichel, M. Greiner, 0. Mandel, A. Widera, T. Rom, I. Bloch, T. W. Hansch, Fortschritte der Physik 54, 702 (2006). 25. L.-M. Duan, M. Lukin, J. I. Cirac, and P. Zoller, Nature 414, 413 (2001), here referred to as DLCZ. 26. Chin-Wen Chou, Julien Laurat, Hui Deng, Kyung So0 Choi, Hugues de Riedmatten, Daniel Felinto, and H. J. Kimble, Science 316, 1319 (2007); published online in Science Express, 5 April, 2007. 27. C. W. Chou, H. de Riedmatten, D. Felinto, S. V. Polyakov, S. J. van Enk, and H. J . Kimble, Nature 438, 828 (2005).
294 28. D. Felinto, C. W. Chou, J. Laurat, E. W. Schomburg, H. de Riedmatten, and H. 3. Kimble, Nature Physics 2, 844 (2006) (advanced online publication 29 October, 2006). 29. D. Felinto, C.W. Chou, H. de Riedmatten, S.V. Polyakov, and H.J. Kimble, Phys. Rev. A 72, 053809 (2005).
OPTICALATOMICCLOCKS
This page intentionally left blank
FREQUENCY COMPARISON OF Al+ AND Hg+ OPTICAL STANDARDS T. ROSENBAND*, D. B. HUME, A. BRUSCH, L. LORINIt, P. 0. SCHMIDT$, T. M. FORTIER, J. E. STALNAKERS, S. A. DIDDAMS, N. R. NEWBURY, W. C. SWANN, W. H. OSKAYT, W. M. ITANO, D. J. WINELAND, AND J. C. BERGQUIST National Institute of Standards and Technology, 325 Broadway, Boulder, CO 80305 *E-mail: [email protected] We compare the frequencies of two single ion frequency standards: 27Alf and Ig9Hgf . Systematic fractional frequency uncertainties of both standards are and the statistical measurement uncertainty is below 5 x below Recent ratio measurements show a reproducibility that is better than
Although single-ion optical frequency standards promise a potential accuracy of lo-'' or better,'>2 this long-standing goal has not yet been realized due to various technical difficulties. Here we report progress for the NIST lg9Hg+ and 27Al+ single-ion standard^,^)^ as their systematic fractional frequency uncertainty approaches In these measurements, the fourth harmonics of two clock lasers are locked t o the mercury and aluminum clock transitions at 282 and 267 nm respectively. An octave-spanning self-referenced Ti:Sapphire femtosecond laser frequency comb (FLFC)5 is phase-locked to one clock laser, and the heterodyne beat-note of the other clock laser with the nearest comb-tooth is measured. The various beat-note and offset frequencies can be combined to yield a frequency ratio, which is independent of the Cs-based definition of the second, allowing this ratio to be measured even more accurately than the fundamental unit of time can be realized. In recent comparisons of the frequencies of the two clock lasers included here, an octave-spanning selfreferenced fiber comb laser6 has provided a second independent measure of t Present address: Istituto Nazionale di Ricerca Metrologica (INRIM), Torino, Italy f Present address: Institut fur Experimentalphysik, Universitat Innsbruck, Austria §Present address: Department of Physics and Astronomy, Oberlin College, Oberlin, OH TPresent address: Stanford Research Systems, Sunnyvale, CA
297
298 the frequency ratio. The 27Al+ 'So t 3Po standard, which uses Quantum Logic Spectroscopy7 has been described previously.* Briefly, one 27Al+ ion is trapped together with a 'Be+ ion, which provides sympathetic cooling. The 27Alf clock state is mapped to 'Be+ repetitively through their coupled motion, allowing for up to 99.94% clock state detection fidelity.8 With the ability to detect the clock state comes the ability to detect state transitions, whose probability depends on the clock laser frequency. The clock laser is locked to the atomic transition by alternating between upper and lower slopes of the atomic resonance curve, and keeping the transition rates equal. At the operating field of 1 gauss, the Zeeman structure is split by several kilohertz, and the individual Zeeman components are well resolved. We alternate between the extreme states of opposite angular momentum (mF = f 5 / 2 ) , which allows compensation of magnetic field shifts up to second The accuracy of the aluminum standard is limited to 2.2 x primarily due to uncertainties in the second-order Doppler shifts. For ions confined in Paul traps, there are two types of motion: secular motion, which is the harmonic motion of the trapped particle, and micromotion, which occurs when the ion is displaced from the rf-null of the confining field. Both types of motion have rms-velocities of about 1-2 m/s, with similar uncertainties. Excess micromotion results when slowly fluctuating electro-static fields in the ion trap displace the ion from the null position. These quasi-static fields are monitored and corrected by interleaving micromotion-test experiments with the clock interrogations. Tests are performed by monitoring the strength of radial-to-axial coupling of certain normal modes via 'Be+,'' and applying compensation voltages at the ion trap to minimize this coupling. With real-time corrections, the stray electric fields are nulled to (0 f 10) V/m, allowing an estimate of the time dilation uncertainty.l' The total fractional frequency clock shift due to the extreme case of a 10 V/m field along both radial directions is 3.2 x and we estimate the fractional frequency shift caused by residual (uncompensated) stray electric fields to be (2 f 2) x when the clock is operating. Secular mode causes deviations in the secular kinetic energy from the Doppler-cooling limit. Among the 6 normal modes" of the 27Alf/gBe+ionpair, two radial modes with large amplitude on 27Al+ , and "The two axial normal-mode frequencies are 5.80 MHz and 2.62 MHz, and the frequencies of the poorly damped radial modes are u, = 3.50 MHz and vy = 4.64 MHz. The other pair of radial modes, which is well damped, has frequencies of 14.06 MHz, and 13.02
MHz.
299 small amplitude on 9Be+ are damped only weakly by the laser-cooled 9Be+ ion. Prior to clock interrogation, we enhance this damping by twisting the ion-pair with a static electric field of about 300 V/m in order to begin each experiment a t the Doppler-cooling limit. We also apply 'Be+ -Dopplercooling light continuously during each 100-ms clock interrogation, keeping the four well-damped modes at the Doppler-cooling limit. However, the two poorly damped radial modes have damping rates of y M 60 s-l, and heating rates of (&) = 400 f 300 quanta/s, and (hY)= 100 f 100 quanta/s, leading to 6.6 5 and 1.5f1.5 excess motional quanta. The fractional clock shift per radial motional quantum is 8 x for both modes, and at the Doppler-cooling limit both modes would contain 3 f 1 quanta. Conservatively we estimate that the average total motional quantum number for the two radial modes during each clock interrogation is 14 f 10 quanta, leading to a second-order Doppler shift of (-1.1 f 0.8) x Other important shifts are the blackbody radiation shift,14 which is very small in 27Al+115and the quadratic Zeeman shift, which has been accurately calibrated by varying the magnetic field, and measuring the shift in the 27Al+ /l''Hg+ ratio, together with the linear Zeeman splitting v1 between the 7r-polarization-excited m = f % lines. The resulting shift is v2 = v; x 1.0479(7) x 10W8/Hz.
*
1 i 1 -; i 1
Table 1. 27Al+ 'So + 3Po and lg9Hg+ 2 S ; + 2 D s shifts in units of lo-'' 2 of fractional frequency. Evaluations of first-order Doppler shifts, background gas collisions, and the gravitational red shift are in progress. shift Micromotion Secular motion Blackbody rad. 313 nm Stark DC Quad. Zeeman AC Quad. Zeeman Electric quadrupole TOTAL
I I
Av~i -20
I
U A ~ AvHg
20
I
UHg
-4
4
-1203 0 0 -1210
10 10 16
4 :
-685 0 0 -735
0.5 0 0.5 22
I I
limitation static electric fields Doppler cooling DC polarizability polarizability, intensity calibration trap magnetic fields B-field orientation
The lg9Hg+ ion standard is based on the 2 S + + 2D+ electric-quadrupole tran~itionA . ~ 194 nm laser cools the ion to the Doppler-cooling limit via the allowed 2S4 + 2P; transition] and a fiber laser frequency-quadrupled to 282 nm excites the clock transition. The clock state of the "'Hg+ ion is measured directly via quantum jumps in the scattering fluorescence rate of the 194 nm laser.16 Systematic uncertainties in the lg9Hg+ standard are listed in Table 1.Here the largest uncertainty is a possible quadratic Zeeman
300 shift due to unbalanced rf-currents in the ion trap. The magnitude of this shift is conservatively estimated by assuming the worst case scenario; that all of the rf-currents flowing in the ring electrode (in the middle of which the ion is held) are fully unbalanced and travel in a half loop along one side of the ring only. The maximum rms-current (IRF < 1 mA) at the site of the trap is set by the trap capacitance (C < 70 ,)?!f the rf-frequency (12 MHz) and the maximum voltage applied (1100 V). This produces an rms-field of approximately 4 x lop7 T a t the center of the ring electrode for a ring diameter of 1 mm. A field of this magnitude causes a fractional frequency shift of the order 1 x The return path for the rf-current through the endcap electrodes ideally produces no net field at the center of the trap, but if they are unbalanced, then the field produced can only help compensate the field generated by the asymmetrical flow of rf-currents through the ring electrode. Otherwise, the rms-field produced at the site of the ion is less than 4 x T. The electric-quadrupole shift, which has previously limited the accuracy of the lg9Hg+standard is constrained below by averaging over three orthogonal magnetic field direction^.'^)^^ Second-order Doppler shifts are easier to control for lg9Hgf than for 27Al+ , because the heavy mercury ion moves less in response ambient electric fields than the lighter aluminum ion does. Near the Doppler-cooling limit, the total time-dilation shift due to secular motion is ( - 3 f 3 ) x 10-lS. Micromotion is carefully compensated," leading to a similar shift of (-4 4) x The quadratic Zeeman coefficient in lg9Hg+ was calibrated in an analogous way to that of the 27Alf standard. Here the shift is u:! = -u; x 1.23564(37) x 10-ll/Hz, where u1 is the linear Zeeman splitting between the 7r-polarization-excited m F = f 2 lines. Both atomic clocks were operated simultaneously, while the FLFCs recorded their frequency ratio every second. Figure l ( b ) shows Allan deviation (an estimate of the statistical measurement uncertainty vs. measurement duration) of a typical ratio measurement. For measurement durations r greater than 100 s, the uncertainty is given by 7 x 1 0 - 1 5 / f i . A deviation from this slope at long measurement times, which would indicate fluctuating systematic shifts, has not been observed. Figure l ( a ) shows the history of measurements of the u ~ ~ + / v Hratio. ~ + - While the systematic uncertainties in Table 1 are only valid for the last data point, the consistency of the earlier measurements provides confidence in the reproducibility of this result. Currently, both standards are being evaluated for first-order Doppler shifts, which would appear if the ion motion was correlated with the probe
*
301
8 0.5
o
v
10
10
10 10 10 time [seconds]
10
0.4
Mar May Jul
Sep Nov Jan Mar
(b) Fig. 1. Fractional stability of the ratio vAl+/vHg+ (a). History of measurements of V AI+ lvHg+ O")1 Error bars are statistical, and the shaded area represents the statistical uncertainty of the weighted mean. These data points do not contain the complete set of systematic corrections listed in Table 1, and will be revised in the future.
beam. By probing alternately from opposite directions, such motion can be detected, and averaged away. Possible causes for such a shift would be stray charge buildup in the ion trap, which is correlated with the interrogation pulses, or correlated mechanical motion caused by shutters. We have not observed a direction-dependent shift for either standard. Besides their obvious application to frequency metrology and precision time-keeping, more accurate atomic clocks might also help answer a fundamental question in physics. Are the constants of nature really constant, or are they changing in time, or dependent on the gravitational potential in which they are measured? Frequency ratio measurements of dissimilar atomic clocks can help answer these questions. In particular, the ratio of atomic resonance frequencies depends on the fine-structure constant a.19'20 Repeated accurate measurements of the i^Ai+/l/Hg+ frequency ratio will provide constraints on present-day changes of the fine-structure constant.
Acknowledgements
This work is supported by ONR, DTO, and NIST. P.O.S. acknowledges support from the Alexander von Humboldt Foundation. This work is a contributions of NIST, and is not subject to U.S. copyright.
302 References 1. H. G. Dehmelt, IEEE Trans. Inst. Meas. 31,p. 83 (1982). 2. D. J. Wineland, W. M. Itano, J. C. Bergquist and R. G. Hulet, Phys. Rev. A 36,2220(Sep 1987). 3. W. H. Oskay, S. A. Diddams, E. A. Donley, T. M. Fortier, T. P. Heavner, L. Hollberg, W. M. Itano, S. R. Jefferts, M. J. Delaney, K. Kim, F. Levi, T. E. Parker and J. C. Bergquist, Phys. Rev. Lett. 97, p. 020801 (2006). 4. T. Rosenband, P. 0. Schmidt, D. B. Hume, W. M. Itano, T. M. Fortier, J. E. Stalnaker, K. Kim, S. A. Diddams, J. C. J. Koelemeij, J . C. Bergquist and D. J . Wineland, Phys. Rev. Lett. 98, p. 220801 (2007). 5. T. M. Fortier, A. Bartels and S. A. Diddams, Opt. Lett. 31,p. 1011 (2006). 6. W. C. Swann, J. J. McFerran, I. Coddington, N. R. Newbury, I. Hartl, M. E. Fermann, P. S. Westbrook, J. W. Nicholson, K. S. Feder, C. Langrock and M. M. Fejer, Opt. Lett. 31,3046 (2006). 7. P. 0. Schmidt, T. Rosenband, C. Langer, W. M. Itano, J. C. Bergquist and D. J . Wineland, Science 309,p. 749 (2005). 8. D. B. Hume, T. Rosenband and D. J. Wineland, Phys. Rev. Lett. 99, p. 120502 (2007). 9. J. E. Bernard, L. Marmet and A. A. Madej, Opt. Comm. 150,p. 170 (1998). 10. M. D. Barrett, B. DeMarco, T. Schaetz, V. Meyer, D. Leibfried, J. Britton, J. Chiaverini, W. M. Itano, B. JelenkoviC, J. D. Jost, C. Langer, T. Rosenband and D. J . Wineland, Phys. Rev. A 68,p. 042302 (2003). 11. D. J . Berkeland, J . D. Miller, J. C. Bergquist, W. M. Itano and D. J . Wineland, J. Appl. Phys. 83,p. 5025 (1998). 12. Q. A. Turchette, D. Kielpinski, B. E. King, D. Leibfried, D. M. Meekhof, C. J . Myatt, M. A. Rowe, C. A. Sackett, C. S. Wood, W. M. Itano, C. Monroe and D. J. Wineland, Phys. Rev. A 61, p. 063418(May 2000). 13. L. Deslanriers, S. Olmschenk, D. Stick, W. K. Hensinger, J. Sterk and C. Monroe, Phys. Rev. Lett. 97,p. 103007 (2006). 14. W. M. Itano, L. L. Lewis and D. J . Wineland, Phys. Rev. A 25, p. 1233 (1982). 15. T. Rosenband, W. M. Itano, P. 0. Schmidt, D. B. Hume, J . C. J. Koelemeij, J. C. Bergquist and D. J. Wineland, Blackbody radiation shift of the 27alS Is0 - 3pO transition arXiv:physics/0611125. 16. J . C. Bergquist, R. G. Hulet, W. M. Itano and D. J . Wineland, Phys. Rev. Lett. 57,1699(0ct 1986). 17. W. M. Itano, J . Res. NIST 105,p. 829 (2000). 18. W. H. Oskay, W. M. Itano and J. C. Bergquist, Phys. Rev. Lett. 94,p. 163001 (2005). 19. J . K. Webb, V. V. Flambaum, C. W. Churchill, M. J. Drinkwater and J . D. Barrow, Phys. Rev. Lett. 82,884(Feb 1999). 20. S. G. Karshenboim, V. Flambaum and E. Peik, Atomic clocks and constraints on variations of fundamental constants arXiv:physics/0410074.
SR OPTICAL CLOCK WITH HIGH STABILITY AND ACCURACY* A. LUDLOW, S. BLATT, M. BOYD, G. CAMPBELL, S. FOREMAN, M. MARTIN, M. H. G. DE MIRANDA, T. ZELEVINSKY, AND J. YE JILA, National Institute of Standards and Technology and University of Colorado Boulder, Colorado 80309-0440, USA E-mail: [email protected]
T. M. FORTIER, .I. E. STALNAKER, S. A. DIDDAMS, C. W. OATES, z. w . BARBER, AND N. P O L I ~ National Institute of Standards and Technology 325 Broadway, Boulder, Colorado 80305, USA We report on our recent evaluations of stability and accuracy of the JILA Sr optical lattice clock. We discuss precision tools for the lattice clock, including a stabilized clock laser with sub-Hz linewidth, fs-comb based technology allowing accurate clock comparison in both the microwave and optical domains, and clock transfer over optical fiber in an urban environment. High resolution spectroscopy (Q > 2 x 1OI4) of lattice-confined, spin-polarized strontium atoms is used for both a high-performance optical clock and atomic structure measurement. Using a Ca optical standard for comparison, the overall systematic uncertainty of the Sr clock is reduced to < 2 x
1. Introduction Optical atomic clocks derive their advantages in stability and accuracy from the enhanced measurement precision arising from large line quality factors (Q) that now exceed lOI4 [ l , 21. The lattice clock technique [ l , 3-71 presents an intriguing situation in which one benefits from a large signal-to-noise ratio ( S N ) provided by a neutral atom ensemble, while enjoying many of the spectroscopic features of single trapped ions. The lattice potential provides Lamb-Dicke confinement where the resonance feature of interest is not influenced by atomic motion, enabling high precision, high accuracy measurements. The weakly allowed 'So - 3P0 transition in alkaline-earth(-like) atoms is particularly well suited for such a scheme due to the existence of a magic wavelength where the ac Stark shift from the trapping laser is identical for the two clock states, and the insensitivity of the states to lattice polarization. The accuracy of a lattice clock based on "Sr is expected to reach below while *This work is supported by NIST, ONR, and NSF. talso with Dipartimento di Fisica and LENS, Universita di Firenze, INFN-Sezione di Firenze, Via N. Carrara 1 1-50019 Sesto Fiorentino (FI), Italy
303
304 the stability could reach below at 1 s when probed by a sufficiently stable optical local oscillator. Such accuracy and precision goals will require focused efforts in both laser development and precision spectroscopy. In this paper we report the present status of our clock system, including local oscillator performance, clock comparison technology using fs-combs, high resolution spectroscopy, and the current level of accuracy and stability of the lattice clock.
2.
OPTICAL CLOCKWORK
2.1. Stable Optical Local Oscillator For the purposes of operating an optical clock with the best possible stability, it is necessary for the probe laser to have an exceedingly long phase-coherence time. Development of such an oscillator to probe the Sr clock transition at 698 nm has therefore been one of the central focuses of our work. The stabilization scheme for our laser is discussed in detail elsewhere [8]. Briefly, a gratingstabilized diode laser is locked to a high-finesse ultra-low-expansion (ULE) cavity mounted in a vertical orientation to reduce fluctuations of the cavity length due to vibrations [9]. The cavity is under vacuum and mounted on a compact, passive vibration-isolation table.
IE-3
0.h
0:1
i
70
I00
time (5)
Figure 1. Allan deviation (with a linear drift removed) for a heterodyne beat between two independent clock laser systems at 698 nm. The -1 x lo-'' stability at 0.1 - 1000 s is consistent with the estimated thermal noise limit. The inset shows the beat between the lasers is well below 1 Hz for an integration time of a few seconds.
The clock laser performance has been characterized in a number of ways, including direct comparison of two similar systems at 698 nm, comparison to highly stabilized lasers at other colors using the fs-comb, and by precision atomic spectroscopy. Direct comparison between the two 698 nm systems via heterodyne beat reveals narrow linewidths often below 300 mHz for integration times of a few seconds (see the inset of Fig. 1). The Allan deviation shown in
305 Fig. 1 reveals that the fractional frequency noise of the beat i s -1 x at 0.1 1000 s, consistent with the expected limit set by thermal-mechanical noise in the cavity mirrors and mirror coatings [lo].
2.2. Optical Frequency Comb Clockwork and Precision Fiber Transfer The femtosecond optical comb plays an essential role for clock development. It provides a coherent link between high accuracy clocks operating in either the optical or microwave domains. For absolute frequency measurements, a directoctave-spanning, self-referenced frequency comb similar to that reported in [ 111 is stabilized by the clock laser at 698 nm. The comb provides a measurement of the laser frequency relative to a microwave signal derived from a hydrogen maser calibrated to the NIST primary Cs fountain clock [12]. To enable the microwave comparison, a diode laser operating near 1320 nm i s amplitudemodulated and transmitted over a 3.5-km fiber optic link from NIST to JILA [13, 141. The instability of the frequency-counting signal i s -2.5 x for a 1s integration time. The periodic stretching and compressing of the fiber, associated with daily temperature cycles, has been found to cause frequency offsets as large as Stabilization of the fiber link is thus implemented for high accuracy clock comparisons [14].
-+ 1 Hz
-40
-20
20
Offset from carrier (Hz)
40
ion
10'
lo2
!03
10'
lo5
Averaging Time (s) Figure 2. (a) Optical beat between the fiber-transferred laser and the NIST comb at 1064 nm. The fs-comb in JILA is stabilized to the 698 nm clock laser, distributing the stability across the optical spectrum. A JILA 1064 nm Nd:YAG laser is stabilized to the comb and used as a transfer laser to NIST over a 3.5 km fiber which is optical-phase stabilized. At NIST, a second fs-comb stabilized to the 1123 nm Hg+ clock laser is used to compare the two stable oscillators. The beat linewidth between the NIST comb and the transfer laser is 1 Hz. (b) Instability of the transfer systems. Open diamonds are typical passive instabilities of the 7- and 32-km fiber links. Closed diamonds (circles) are for the 32-km (7-km) fiber transfer system. Black triangles represent in-loop performance. The solid line represents 1 radian of accumulated phase noise during the averaging time.
The comb and fiber link also support transmission of stable optical frequencies between JILA and NIST. First, the comb i s used to distribute the Sr clock stability across the optical spectrum, allowing use of a transfer laser at
306 1064 nm. We have examined the precision of the distribution across the comb spectrum by comparing our 698 nm clock laser with a similar independent system operating at 1064 nm [9]. The resulting beat between the stabilized comb and the 1064 nm laser has a linewidth below 0.5 Hz, verifying the ability of the comb to hlly preserve the optical phase coherence. For transfer of the clock signal, a 1064 nm laser is phase-locked to the comb and sent through the fiber to the NIST laboratory where it is measured relative to a number of high accuracy optical clocks [15, 16, 71 using a second fs-comb [17]. To eliminate noise induced by the transfer, the 1064 nm light is partially reflected back to our lab allowing interferometric stabilization of the link [ 181. Figure 2 (a) shows a beat between the transfer laser and a NIST fs-comb that is stabilized to a local oscillator used for the Hg+ clock [19]. The beat has a resolution-bandwidthlimited linewidth of 1 Hz, demonstrating that the phase coherence of the two clock lasers is not degraded by the two fs-combs and the fiber transfer process. In addition, Fig. 2(b) summarizes results of passive and phase-stabilized optical transfer for 7 km (the round trip length of our fiber link) and 32 km fiber lengths. The transfer noise of the stabilized fiber links is -1 x at 1 s [ 181.
3.
High resolution Spectroscopy
Lattice-confined fermionic 87Sratoms are used for an accurate atomic reference. 87 Sr atoms are laser cooled in two stages [20, 41, first via a broad (32 MHz) and then a narrow (7.4 kHz) cycling transition. A vertical one-dimensional optical lattice is overlapped with the atomic cloud as the atoms are cooled in the second stage magneto-optic trap. After the cooling cycle is finished, roughly lo4 atoms remain trapped in the lattice, at a temperature of -2 pK. 3.1. Spectroscopy in the Magic Wavelength Lattice The lattice potential is sufficiently deep such that the 'So - 3P0clock transition is probed in the Lamb-Dicke regime along the lattice axis where the spatial dimension of the confinement is small compared to the interrogation wavelength. The resulting absorption spectrum is free of Doppler or recoil effects, allowing high line Q's to be achieved. When the transition is strongly saturated, spectra such as that in Fig. 3 (a) are observed. Here the narrow transition is accompanied by motional sidebands. The sideband spectrum is useful for direct measurement of the thermal-mechanical properties of the trapped atoms, including temperatures and trap motional frequencies [4]. The narrow central feature serves as the atomic reference as it excludes any motional or Stark broadenings when the magic wavelength lattice is used. In the unsaturated case, this carrier transition has a width of -5 Hz, shown in Fig. 3(b), representing a line Q of loi4.
307 3.2. Nuclear Spin Effects The limit to the achievable line Q in Fig. 3(b) is mainly due to the magnetic sensitivity of the clock transition. The hyperfine state mixing, which leads to the -1 mHz linewidth of the otherwise forbidden clock transition, increases the magnetic moment of the 3P0 state, resulting in a small differential g-factor between 3P0and 'So. The resultant sensitivity of the clock transition to magnetic fields requires careful clock design to ensure frequency stability and accuracy. I
1
v
0 .& *
cd
$0.6
e,
-VP0.4-15
-10 -5
0
5
10 15
Laser Detuning (Hz) Figure 3. High resolution spectroscopy of lattice confined "Sr atoms. (a) Resolved longitudinal sideband spectrum observed when the transition is strongly saturated. The trap frequency is determined from the carrier-sideband separation. The atom temperature can be extracted from the lineshape of the sidebands, which are asymmetrically broadened due to motion in the trap. (b) The narrow carrier transition used as the atomic reference. Here the 5 Hz linewidth is limited by residual magnetic fields.
To determine the magnitude of the differential g-factor, we have performed NMR-like measurements on various mF-sublevels of the clock transition in the presence of a bias magnetic field. Figure 4(a) shows the absorption spectrum in the presence of a bias field using linearly polarized light to drive ten possible Ktransitions (AmF= O). In Fig. 4(b), the probe light polarization is rotated by 90" from the quantization axis such that nine 0- (AmF= -1) and nine o+ (AmF= +1) transitions are observed simultaneously. While the n-transitions would be sufficient for measurement of the differential g-factor magnitude, the o+'transitions provide more information as the spectrum depends on both the magnitude and the sign of the effect (i.e. the state mixing increases the 3P0gfactor). Furthermore, calibration of the field magnitude is automatic if the ground state magnetic moment is known. Spectroscopy of the resolved sublevels gives us precise measurement of the differential g-factor, yielding 108.4(4) x mF HdG [21, 11. The resolved spectrum has also allowed experimental investigation of the tensor light shift sensitivity of the clock states,
308 demonstrating that they have a negligible impact on the current 1-D lattice clock P11. 3.3. Hz-Resolution Optical Spectroscopy With the degeneracy of the clock states lifted in the presence of a magnetic field, the resolved transitions are explored free-of-broadening from the statedependent
._
1.0
Y
-g
0.8 0.6
cd
t i
0.4
0.2
_ao0.o
-300
-200
-100
0
200
100
300
Laser Detuning (Hz)
-500
-250
0
250
500
Laser Detuning (Hz) Figure 4. Spectroscopy of resolved mF-sublevels in the presence of a magnetic bias field. (a) Ten x-transitions are observed when the probe is linearly polarized along the quantization axis set by a 0.5843 magnetic field. (b) When the probe polarization is orthogonal to the quantization axis, 18 o transitions are seen, under a bias field of 0.69 G. In both (a) and (b), peaks are labeled according to the ground state mF origin of the transition. In (b) the spectrum consists of two overlapped sets of nine evenly spaced peaks, the excitation polarization is labeled for clarity.
309 0.10
40.04 IA
30.02 -6 -4 -2 0 2 4 6 Laser Detuning (Hz)
-40 0 40 80 120 Laser Detuning (Hz)
Figure 5. High-resolution spectroscopy of the clock transition for mF = 512. (a) Ultraspectrum achieved with a 500-ms Rabi pulse. (b) Ramsey narrow (Q > 2 x spectrum for 25-ms Ramsey pulses separated by a 80-ms evolution time, resulting in a 10 Hz FWHM for the central fringe. The inset shows the spectrum for longer pulse and freeevolution times of 80 ms and 200 ms, respectively. In both (a) and the inset of (b), the spectral resolution is limited by laser frequency noise during acquisition times of -30 s.
Zeeman and Stark shifts, revealing the true spectroscopic limit of our system. An example spectrum is shown in Fig. 5(a), where the 3Po(mF= 5/2) population is measured as the laser is tuned across the resonance. The resulting resonance [I], near the -1.8 Hz Fourier limit set by the width is 2.1(2) Hz (Q > 2 x 500-ms interrogation time. Laser frequency noise prevents longer probe times as the beat between our two 698 nm lasers broadens to nearly 2 Hz when integrated over the relevant timescale for scanning the line (about 30 s). Ramsey spectroscopy can also be performed, as shown in Fig. 5(b). The Doppler-free absorption spectrum and long interaction time provided by the lattice confinement allows use of long, low-intensity pulses for the Ramsey sequence, resulting in simple spectra with a small number of fringes. 4.
Clock Performance
The high-resolution spectroscopy results establish confidence in the lattice clock technique as a tool for precision measurement. The high stability oscillator, large line Q, and relatively large S/N can result in clock stability at the 1 x 10.'' level at 1 s. Another important issue is the level of accuracy that can be achieved with the lattice system. 4.1. Accuracy Evaluation (2006): Degenerate Sublevels We first present an accuracy evaluation using degenerate sublevels without a bias magnetic field. To evaluate the accuracy of the "Sr lattice clock an interleaved scheme is used where the parameters of interest are quickly varied
31 0 during a scan of the transition, allowing extraction of the shift coefficient using the precision of the laser cavity as a reference. In this way, we explore a variety of systematic effects, with dominant contributions to the uncertainty budget being the lattice shift, density shift, and Zeeman shift. Operating at a lattice wavelength of 813.4280(5) nm, we find that the ac Stark shift for the typical For the typical density of -5 x 10" ~ m -the ~, trap depth is -2.5(6.0) x collision shift is constrained to be below 3.3 x Each of these effects is consistent with zero and limited by the statistical uncertainty associated with hundreds of measurements. The Zeeman sensitivity is found to be non-zero but < 5.3 x when the residual magnetic field is controlled to 5 mG. Due to the differential g-factor, we zero and monitor the field to this level using the transition linewidth. The total uncertainty for lattice spectroscopy under these operating conditions is 9 x [5]. As the spectroscopy systematic uncertainties are evaluated, an absolute frequency measurement is performed using the fiber link and a Cs-calibrated Hmaser. After a 24 hour run, the frequency of the 87Sr 'S0-3Po transition is determined to be 429,228,004,229,874.0 + 1.1 Hz. The total uncertainty of 2.5 x is dominated by the statistical uncertainty in the frequency comparison (1.4 x and the maser calibration (1.7 x lo-") [5]. Figure 6 summarizes the recent absolute frequency measurements made by the groups in JILA [4, 51, Paris [6, 221, and Tokyo [23]. The excellent agreement between independent groups speaks strongly for the 87Srlattice clock as a frequency standard. 8C n
7c
W
$
6C
?O
5c
>$
A
I
, -
A
761 741
4c
ni
Figure 6 . Recent absolute frequency measurements of the 87Srclock transition by JILA (circles), Paris (triangles), and Tokyo (squares). The frequency is reported relative to a set frequency vg = 429,228,004,229,800 Hz. The inset shows a close-up view of the last three points on the main plot, with the last two [5,21] differing by 7 x
311
4.2. Accuracy Evaluation: Optical Clock Comparison with SpinPolarized Samples Under its current operating parameters, the Sr lattice clock has a potential stability limit from quantum projection noise at 7 x 10"16 at 1 s. In practice, the stability is estimated to be 2 x 10"'5 at 1 s, limited by the Dick effect of the optical local oscillator [24]. Such measurement stability cannot be fully exploited by comparison to current state-of-the-art microwave standards. Rather, accuracy evaluation of optical atomic clocks greatly benefits from direct comparison against other optical clocks with similar stability. Using the optical carrier transfer system described earlier, we now evaluate systematic uncertainties of the JILA Sr lattice clock by measurement against the cold Ca optical clock at NIST [15]. Figure 7 shows the measurement potential of a clock comparison between the 87Sr lattice clock and the Ca clock. The stability at 1 s is ~ 3 x 10~15 and a precision of 3 x 10"16 is achieved in only 200 s of averaging. This frequency discrimination at relatively short averaging times facilitates high precision measurement of systematic shifts of the Sr clock. To reduce measurement sensitivity to long time-scale (> 1000s) frequency drifts of the Ca standard, we measure frequency differences between sets of 100-s data under various Sr parameters to be evaluated. By collecting many sets of these data, we characterize systematic shifts below the 1 x 10"16 level. Srvs.NlSTMaser| Srvs.MSTCa I
10-" i c o
' =I:f
10"' 10 .
Time (s)
100
Figure 7. Microwave and optical clock comparisons using the fiber link discussed in the text. The Allan deviation for the Sr - Maser comparison is shown as squares, yielding a stability of-2.5 x 10'13 T '"2. The advantage of an optical clock is apparent from the Sr Ca comparison (circles), which typically yields stability of 3 x 10'15 at 1 s.
31 2
Figure 8. Spectroscopy of the x transitions under a bias magnetic field. Blue (dark gray) is with atomic polarization to mF = + 912, and red (light gray) is without. The polarization is achieved by optical pumping using the 'SOF=9/2 - 3PIF=7/2 transition. To reduce clock sensitivity to some key experimental parameters, we spinpolarize the Sr sample to the stretched nuclear spin states, mF = & 9/2, and then perform spectroscopy with a small bias magnetic field to lift spin state degeneracy (see Fig. 8). By stabilizing the clock laser to the average of the n transitions from both spin states, we remove shifts originating from the lStorder Zeeman effect and the vector light shift due to the lattice confinement. The clock remains sensitive to 2"d order Zeeman shifts. As we vary the bias magnetic field and measure the clock frequency versus Ca, we confirm that the 1'' order Zeeman effect is eliminated, resulting in a clock shift of 1.6(2.3) x 1017 G-2, . Furthermore, we measure the 2"d order Zeeman shift at 5.8(0.8) x resulting in a 2"d order frequency shift of 2.3(0.2) x In addition to Zeeman shifts, we characterize systematic shifts resulting from: AC Stark effect from the lattice and probe lasers, atom servo error, blackbody AC Stark effect, atomic collisions, and higher order motional effects. A preliminary summary of these effects is given in Table 1. One of the largest observed shifts originates from density-dependent Sr-Sr interactions. Measurement of the clock shifts indicate that these interactions depend on both the nuclear spin state as well as the electronic excitation. We are currently investigating these effects in more detail and it is anticipated that a zero-densityshift condition will be found and implemented for the clock operation. The stated density shift and uncertainty in Table 1 are given for our current operating conditions. The total systematic shift uncertainty is < 2 x indicating the strong performance of the Sr lattice clock as a potential primary frequency standard. We expect this uncertainty to continue to decrease.
313
Contributor Lattice Stark (scalarltensor) AC Stark (probe) i,9torder Zeeman 2nd order Zeeman Servo Ekror
Correction 2.5 0.2 0.2 0.2
Uncertainty 0.5 0.2
0
0.5 1 1
Density BBR Stark Hyperpolarizability (labtice) Line Pulling 2nd order Doppler Systematic total
15 52.1 0 0 0
0.2
0.02
<0.2 0.2 KO.01 1.7
Table 1. Systematic frequency corrections for the Sr lattice clock, and their respective uncertainties. All quantities are expressed in fractional frequency of Conclusion We have presented our recent results for an optical atomic clock based on strontium atoms confined in an optical lattice. The local oscillator is now performing at the 1 x lo-’’ level at 1 - 1000 s, and the fiberkomb transfer system has been shown to support precision frequency transfer at the level of 1 x at 1 s. The lattice clock technique, combined with the high precision laser, has yielded the highest line Q (> 2 x achieved in coherent spectroscopy, with a good S i N for further enhanced stability. Nuclear-spin effects in the lattice clock have been explored, including a measurement of the differential g-factor of the clock transition. An accuracy evaluation is performed, reducing the lattice clock systematics below lo-’’, and an absolute frequency measurement with a 1 Hz uncertainty is made, which is in excellent agreement with that of other groups. Comparison of the Sr clock with the NIST at 1 s. Using the high Ca clock revealed the clock instability is < 3 x precision optical comparison and spin-polarized samples, the lattice clock systematic uncertainty has been reduced to 2 x and it will continue to decrease in the near future. Acknowledgments We gratefully acknowledge contributions from J. Bergquist, T. Parker, S. Jefferts, and T. Heavner for help with the fiber-link-based laser comparison and the absolute frequency measurements made via the Sr - Maser comparison. References 1. M. M. Boyd et al., “Optical atomic coherence at the 1-second time scale,”
Science, vol. 314, pp. 1430-1433,2006.
314
2. R. J. Rafac, B. C. Young, J. A. Beall, W. M. Itano, D. J. Wineland, and J. C. Bergquist, “Sub-dekahertz ultraviolet spectroscopy of 199Hg+,” Phys. Rev. Lett., vol. 85, p. 2462,2000. 3. M. Takamoto, F.-L. Hong, R. Higashi, and H. Katori, “An optical lattice clock,” Nature, vol. 435, pp. 321-324,2005. 4. A. D. Ludlow et al., “Systematic study of the 87Sr clock transition in an optical lattice,” Phys. Rev. Lett., vol. 96, p. 033003, 2006. 5. M. M. Boyd et al., “87Sr lattice clock with inaccuracy below 10-15,” Phys. Rev. Lett., vol. 98, p. 083002,2007. 6. R. Le Targat et al., “Accurate optical lattice clock with 87Sr atoms,” Phys. Rev. Lett., vol. 97, p. 130801, 2006. 7. Z. W. Barber, C. W. Hoyt, C. W. Oates, L. Hollberg, A. V. Taichenachev, and V. I. Yudin, “Direct excitation of the forbidden clock transition in neutral 174Yb atoms confined to an optical lattice,” Phys. Rev. Lett., vol. 96, p. 083002,2006. 8. A. D. Ludlow et al., “Compact, thermal-noise-limited optical cavity for diode laser stabilization at 1x10-15,” Opt. Lett., vol. 32, pp. 641-643, 2007. 9. M. Notcutt, L.-S. Ma, J. Ye, and J. L. Hall, “Simple and compact 1-Hz laser system via an improved mounting configuration of a reference cavity,” Opt. Lett., vol. 30, pp. 1815-1817, 2005. 10. K. Numata, A. Kemery, and J. Camp, “Thermal-noise limit in the frequency stabilization of lasers with rigid cavities,” Phys. Rev. Lett., vol. 93, p. 250602,2004. 11. T. M. Fortier, D. J. Jones, and S. T. Cundiff, “Phase stabilization of an octave-spanning Tixapphire laser,” Opt. Lett., vol. 28, pp. 2198-2200, 2003. 12. T. P. Heavner, S. R. Jefferts, E. A. Donley, J. H. Shirley, and T. E. Parker, “NIST-F 1: recent improvements and accuracy evaluations,” Metrologia, vol. 42, p. 41 1,2005. 13. J. Ye et al., “Delivery of high-stability optical and microwave frequency standards over an optical fiber network,” JOSA B, vol. 20, pp. 1459-1467, 2003. 14. S. M. Foreman, K. W. Holman, D. D. Hudson, D. J. Jones, and J. Ye, “Remote transfer of ultrastable frequency references via fiber networks,” Rev. Sci. Instrum., vol. 78, p. 021 101,2007. 15. C. W. Oates, E. A. Curtis, and L. Hollberg, “Improved short-term stability of optical frequency standards: approaching 1 Hz in 1 s with the Ca standard at 657 nm,” Opt. Lett., vol. 25, pp. 1603-1605,2000. 16. W. H. Oskay et al., “Single atom optical clock with high accuracy,” Phys. Rev. Lett., vol. 97, p. 020801,2006. 17. T. M. Fortier, A. Bartels, and S. A. Diddams, “Octave-spanning Ti:Sapphire laser with a repetition rate > 1 GHz for optical frequency measurements and comparisons,” Opt. Lett., vol. 31, pp. 1011-1013,2006.
31 5 18. S. M. Foreman et al., “Coherent optical phase transfer over a 32-km fiber with long-term instability
This page intentionally left blank
AUTHOR INDEX A Abel, M.J. 167 Afanasiev, A.E. 195 Akulshin, A. 228 Alberti,A. 89 Alt, W. 175 Anderson, R. 228 Andrist, M. 153 Apel, P. Yu. 195 Aspect, A. 11,243
Campbell, G. 303 Cancio, P. 75 Challis, K.J. 138 Chang,H. 243 Cherkun, A.P. 195 Chwalla, M. 53 Clement, D. 11 Cornell, E.A. 3 D Darquie, B. 259, 271 De Angelis, M. 89 De Miranda, M.H.G. 303 de Natale, P. 75 Deleglise, S. 63 Diddams, S.A. 297 Dingjan, J. 259 Dutier, G. 271
B Bailly, D. 103 Ballagh, R.J. 138 Balykin, V.I. 195 Bartalini, S. 75 Bergquist, J.C. 297 Bernet, S. 187 Bernu, J. 63 Bertoldi, A. 89 Bethlem, H.L. 103 Beugnon, J. 259 Blatt, R. 53,207 Blatt, S. 303 Bloch, I. 23 Boiron, D. 243 Borri,S. 75 Bouyer,P. 11 Boyd,M. 303 Browaeys, A. 259 Brune,M. 63 Brusch,A. 297 Buning, R. 103 C Cacciapuoti, L.
F Fernholz, T. 113 Ferrari, G. 89 Foreman, S. 303 Fortier, T.M. 297 Fortson, E.N. 39
G Gaebler, J.P. 127 Gaetan, A. 259 Gagliardi, G. 75 Galli, I. 75 Gardiner, C.W. 138 Ghanbari, S. 228 Giorgini, A. 89 Giusfiedi, G. 75 Gleyzes, S. 63
89 31 7
31 8
Goldwin, J. Grangier, P. Guerlin, C.
271 259 63
H Hall, B.V. 228 Hannaford, P. 228 Hannemann, S. 103 Haroche, S. 63 Heinrich, C. 187 Hilliard, A.J. 113 Hinds, E.A. 27 1 Hofer, A. 187 Hogan, S.D. 153 Hogervorst, W. 243 Hume, D.B. 297 I Itano, W.M. 297 Ivanov,V. 89
J Jeltes, T. 243 Jensen, K. 113 Jin, D.S. 127 Jones, M.P.A. 259 Jullien, A. 167
Leone, S.R. 167 Letokhov, V.S. 195 Lorini, L. 297 Luc-Koenig, E. 2 19 Ludlow, A. 303 Lugan, P. 11 M Maddoloni, P. 75 Malara, P. 75 Marion, H. 259 Martin,M. 303 Masnou-Seeuws, F. 219 Mazzotti, D. 75 McLean, R. 228 McNamara, J.M. 243 Meier, B.H. 153 Meier, U. 153 Melentiev, P.N. 195 Merkt, F. 153 Meschede, D. 175 Messin, G. 259 Miroshnychenko, Y. 259 Monz,T. 53 Moszynski, R. 2 19 Muller, J.H. 113 Murphy, M.T. 103
K Kaper, L. 103 Kim,K. 53 Kimble, H.J. 285 Kjaergaard, N. 113 Klimov, V.V. 195 Koch, C.P. 219 Kosloff, R. 219 Krachmalnicoff, V. 243 Krauter, H. 113 Kuhr, S. 63
N Nagel, P.M. 167 Neergaard-Nielsen, J.S. Neumark, D.M. 167 Newbury, N.R. 297 Nielsen, B.M. 113
L Lamporesi, G. 3, 89 Lance, A.M. 259
P Perrin,A. Pfeifer, T.
0 Oblak, D. 113 Olausson, C. 113 Oskay, W.H. 297
243 167
113
319
Poli,N. 89 Polzik, E.S. 113 Prevedelli, M. 89
R Raimond, J.-M. 63 Rauschenbeutel, A. 175 Riebe,M. 53 Ritsch-Marte, M. 187 Roos,C.F. 53 Rosenband, T. 297 Rudnev, S.N. 195 S Salumbides, E.J. 103 Sanchez-Palencia, L. 11 Sayrin, C. 63 Schellekens, M. 243 Schindler, K. 53 Schmidt, P.O. 297 Schmutz, H. 153 Schweikhard, V. 3 Sherson, J.F. 113 Sidorov, A. 228 Singh, M. 228 Skuratov, V.A. 195 Sorrentino, F. 89 Sortais, Y.R.P. 259 Sprecher, D. 153 Stalnaker, J.E. 297 Stewart, J.T. 127 Swann, W.C. 297
T Tino, G.M. 89 Trupke,M. 271 Tuchendler, C. 259 Tung, S. 3 Tychkov, A.S. 243
U Ubachs, W.
103
V Van Leeuwen, K.A.H. Vanhaecke, N. 153 Vassen, W. 243 Vervloet, M. 103 Vliegen, E. 153 Volk, M. 228 W Westbrook, C.I. 243 Whitlock, S. 228 Windpassinger, P. 113 Wineland, D.J. 297
Y Ye, J.
303
”
L
Zelevinsky, T.
303
243