18.01.2013 Views

Thesis (pdf) - Swinburne University of Technology

Thesis (pdf) - Swinburne University of Technology

Thesis (pdf) - Swinburne University of Technology

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Bose-Einstein Condensation<br />

in Micro-Potentials<br />

for Atom Interferometry<br />

Dipl. Phys. Falk Scharnberg<br />

A thesis submitted for the degree <strong>of</strong><br />

Doctor <strong>of</strong> Philosophy at<br />

<strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong><br />

Melbourne, Australia<br />

Hamburg, March 06, 2006


Abstract<br />

Interferometry with atoms is a young discipline <strong>of</strong> physics. The first interfer-<br />

ometers were realised in the early 1990s, using light pulses as beamsplitters<br />

in momentum space. Recent developments in atom optics have led to new<br />

proposals for interferometers, where the splitting takes place spatially. Mag-<br />

netic and optical traps are both suited for this kind <strong>of</strong> interferometer if a well<br />

defined and highly controllable trap is realised. In this thesis results from two<br />

experiments that in principle allow the creation <strong>of</strong> such traps are presented.<br />

After reviewing the principles and techniques <strong>of</strong> atom optics necessary for<br />

the understanding <strong>of</strong> this thesis, a theoretical discussion about the spatial<br />

single atom interferometer in a double well potential follows. It is shown that<br />

within reasonable limits the system can be reduced to two levels and solved<br />

by the Bloch equations. Using the realistic case <strong>of</strong> a not perfectly symmetric<br />

double well potential allows an understanding <strong>of</strong> the physics behind such an<br />

interferometer: how phase is accumulated and how localisation <strong>of</strong> the atoms<br />

leads to the loss <strong>of</strong> the interferometric signal.<br />

Then two experiments are presented. One experiment was newly built up as<br />

part <strong>of</strong> this thesis: it uses a novel hybrid “atom chip” with a combination <strong>of</strong> a<br />

magneto-optical film and current-carrying structures to produce the magnetic<br />

trapping potentials. This experiment allowed 5·10 8 87 Rb atoms to be captured<br />

in a magneto-optical trap, where the atom-chip’s surface acts as a mirror. The<br />

atoms were then transferred into a purely magnetic trap which was created by<br />

iii


current carrying structures on the chip. From there on, RF radiation enforced<br />

evaporative cooling, so that a quantum degenerate Bose-Einstein condensate<br />

<strong>of</strong> up to 10 5 atoms was created. It was also shown that the magneto-optical<br />

film is able to hold and trap the atoms.<br />

In the second experiment atoms are trapped in the spatially varying inten-<br />

sity <strong>of</strong> light <strong>of</strong> 1.03 µm wavelength. This experiment used an existing set-up as<br />

a basis and was modified and improved for this new project. This experiment<br />

starts with 10 9 87 Rb atoms in a 6-beam magneto-optical trap. After examining<br />

and optimising the loading process, 1.5 · 10 5 atoms could be loaded directly<br />

from the Magneto-optical trap into the optical dipole potential <strong>of</strong> two crossed<br />

laser beams at 1.03 µm. Evaporative cooling was demonstrated though the<br />

phase transition to quantum degeneracy was not reached.<br />

iv


Acknowledgements<br />

It is very hard for me to order my acknowledgements as there are so many<br />

people whom I would like to thank. Fortunately, the first address <strong>of</strong> my thanks<br />

is not a question at all: I would like to thank my supervisor Pr<strong>of</strong>. Peter<br />

Hannaford for all that he has done. Without exaggeration I would like to call<br />

him a luminary, not only as a physicist and especially spectroscopist, but also<br />

as a supervisor who has a keen and generous eye for the needs <strong>of</strong> all <strong>of</strong> his<br />

students.<br />

And now it becomes difficult. I will thank location-wise, and start with<br />

<strong>Swinburne</strong>. Here someone that deserves more thanks than I could count is<br />

Shannon Whitlock. He joined after my first year and worked on the experi-<br />

ment during my stays in Hannover. Without him this experiment would never<br />

have become such a success. I enjoyed working with him, be it in the lab or<br />

discussing the atom interferometer theory. He also proved to be invaluable<br />

during my writing up in Germany, being my eyes in the lab and answering all<br />

the things I forgot to write down and now couldn’t look up quickly. Further-<br />

more I want to thank the project leader <strong>of</strong> the atom chip experiment Pr<strong>of</strong>.<br />

Andrei Sidorov for his constant endeavours to support me and promote the<br />

experiment. I enjoyed working with him in the lab. Dr. Brenton Hall I would<br />

like to thank for the fun in the lab and the expertise he brought and shared<br />

with us students. Someone who also taught me quite a few tricks in the lab<br />

was David Gough. I want to thank him that he made it possible that I could<br />

v


hear and speak German in my Hamburg accent in Queensland. I also want to<br />

thank all others who are part <strong>of</strong> the atom optics group at <strong>Swinburne</strong>, mainly<br />

Dr. Alexander Akulshin and Pr<strong>of</strong>. Russell McLean, for helping out whenever<br />

I needed a tool, advice or just another hand. Mark Kivinen I want to thank<br />

“in triplicate” for his great workshop work and Sharon Jesson I want to thank<br />

for all the fights with the administration that I didn’t have to fight.<br />

Now, more private thanks go to my friends Craig Lincoln and Ruth Plathe.<br />

It was great fun sharing a flat with Craig, and I could always count on Ruth<br />

to brighten my mood with her jokes. I also thank Heath who taught me a lot<br />

about the Australian culture and way-<strong>of</strong>-life. There are more people to thank:<br />

Dru, Craig and Grant, Jürgen and Holger, Grainne, Wayne (Rowlands), Wayne<br />

and Bob, Saeed and probably a dozen more who I should mention. I thank all<br />

<strong>of</strong> you for making my time overseas so enjoyable.<br />

On the Hannover side, the first thanks are reserved for Pr<strong>of</strong>. Gerhard Birkl<br />

and Pr<strong>of</strong>. Wolfgang Ertmer. They allowed me to work with them in Gerhard’s<br />

group as an exchange student, and made it possible for me to split my work<br />

between Australia and Germany.<br />

I would like to thank all my colleagues <strong>of</strong> the group A6 in Hannover. First<br />

Dr. Rainer Dumke, who introduced me to the system there and whose dry<br />

humour you can’t forget. Next is André Lengwenus with whom I enjoyed<br />

working during my first time in Hannover: you are doing better work than<br />

you admit to yourself. Dr. Tobias Müther I would like to thank not only for<br />

his vast knowledge on everything and the discussions we had, but also for his<br />

work in the lab, although our overlap there was rather small. Without him,<br />

the successes <strong>of</strong> the experiment would be unthinkable. I would also like to<br />

thank Johanna Nes and Anna-Lena Gehrmann, the first “women only” crew I<br />

have ever seen “man” an experiment.<br />

Very special thanks go to Dr. Michael Volk. Just like I had my flat mate<br />

vi


Chapter 0: Acknowledgements<br />

in the neighbouring lab at <strong>Swinburne</strong>, I had a friend and former flat mate<br />

at the neighbouring optical table in Hannover. I enjoyed the smoking and<br />

later the non-smoking breaks we had together. Furthermore I want to thank<br />

Dr. Norbert Herschbach and Dr. Peter Spoden for their encouraging words<br />

whenever I felt a bit disheartened, and Sascha Drenkelforth. And <strong>of</strong> course<br />

thanks go to all other group members who I didn’t name here; I wish you all<br />

the best for the future.<br />

I also thank my parents for all they have done and for all their support.<br />

Even though this is now the end <strong>of</strong> my list, the most important person in my<br />

life has not been thanked yet.<br />

Nicole, I love you. I have learnt that I never want to be away from you so<br />

far and for so long.<br />

vii


viii


Declaration<br />

I, Falk Scharnberg, declare that to the best <strong>of</strong> my knowledge, this thesis con-<br />

tains no material which has been submitted to another university for the award<br />

<strong>of</strong> any other degree, previously published or written by another person except<br />

where due reference is made in the text. Where the work is based on joint<br />

research or publications, the relative contributions <strong>of</strong> the respective workers or<br />

authors are disclosed.<br />

ix<br />

Falk Scharnberg,<br />

Hamburg, March 06, 2006


Publications by the Candidate<br />

• A. I. Sidorov, R. J. McLean, F. Scharnberg, D. S. Gough, T. J. Davis, B.<br />

J. Sexton, G. I. Opat and P. Hannaford. Permanent-magnet microstruc-<br />

tures for atom optics. Act. Phys. Pol. B 33, 2137-2155 (2002).<br />

• J. Y. Wang, S. Whitlock, F. Scharnberg, D. S. Gough, A. I. Sidorov,<br />

R. J. McLean and P. Hannaford. Perpendicularly magnetized, grooved<br />

GdTbFeCo microstructures for atom optics. J. Phys. D: Appl. Phys. 38,<br />

4015-4020 (2005). Included as appendix A.3.<br />

• B. V. Hall, S. Whitlock, F. Scharnberg, P. Hannaford and A.I. Sidorov.<br />

Bose-Einstein condensates on a permament magnetic film atom chip. In:<br />

Laser Spectroscopy; Proceedings <strong>of</strong> ICOLS 2005, E. A. Hinds, A. Fer-<br />

guson and E. Riis (Editors), page 275-282 (World Scientific, Singapore,<br />

2005).<br />

• B. V. Hall, S. Whitlock, F. Scharnberg, P. Hannaford and A.I. Sidorov.<br />

A permanent magnetic film atom chip for Bose-Einstein condensation.<br />

J. Phys. B: At. Mol. Opt. Phys. 39, 27-36 (2006). Included as appendix<br />

A.2.<br />

• A. I. Sidorov, B. J. Dalton, S. Whitlock and F. Scharnberg. The asym-<br />

metric double-well potential for single-atom interferometry. Phys. Rev.<br />

A 74, 023612 (1-9) (2006). Included as appendix A.1.<br />

xi


xii


Contents<br />

Abstract iii<br />

Acknowledgements v<br />

Declaration ix<br />

Publications by the Candidate xi<br />

1 Introduction 1<br />

2 Theoretical Background 13<br />

2.1 Atoms and Electromagnetic Fields . . . . . . . . . . . . . . . . . 14<br />

2.1.1 Dressed state model . . . . . . . . . . . . . . . . . . . . 15<br />

2.1.2 Absorption and emission <strong>of</strong> photons . . . . . . . . . . . . 18<br />

2.1.3 Detection <strong>of</strong> atoms . . . . . . . . . . . . . . . . . . . . . 21<br />

2.1.4 Trapping <strong>of</strong> atoms in light fields . . . . . . . . . . . . . . 23<br />

2.1.5 Cooling atoms with light . . . . . . . . . . . . . . . . . . 24<br />

2.2 Atoms and Magnetic Fields . . . . . . . . . . . . . . . . . . . . 28<br />

2.2.1 Magnetic trapping . . . . . . . . . . . . . . . . . . . . . 28<br />

2.2.2 Permanent magnets . . . . . . . . . . . . . . . . . . . . . 34<br />

2.3 Evaporative Cooling . . . . . . . . . . . . . . . . . . . . . . . . 35<br />

2.3.1 A simple model . . . . . . . . . . . . . . . . . . . . . . . 36<br />

xiii


CONTENTS<br />

2.4 Bose-Einstein Condensation . . . . . . . . . . . . . . . . . . . . 39<br />

2.5 The Magneto-Optical Trap . . . . . . . . . . . . . . . . . . . . . 41<br />

2.5.1 Mirror MOT . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

2.6 Atom Interferometry . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

2.6.1 Atom interferometry with symmetric double well poten-<br />

tials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

3 The Asymmetric Double Well 51<br />

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

3.2 Two Mode Approximation and the Bloch Equations . . . . . . . 56<br />

3.2.1 Two mode approximation . . . . . . . . . . . . . . . . . 56<br />

3.2.2 Bloch equations . . . . . . . . . . . . . . . . . . . . . . . 59<br />

3.3 The Results <strong>of</strong> the Model . . . . . . . . . . . . . . . . . . . . . . 65<br />

3.3.1 General results . . . . . . . . . . . . . . . . . . . . . . . 65<br />

3.3.2 Comparison with experimental data . . . . . . . . . . . . 74<br />

3.3.3 CARP: Coherent Adiabatic Readout Process . . . . . . . 76<br />

3.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />

4 The Permanent Magnetic Chip Experiment: Apparatus 81<br />

4.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

4.2 The Laser Systems . . . . . . . . . . . . . . . . . . . . . . . . . 83<br />

4.2.1 Main laser . . . . . . . . . . . . . . . . . . . . . . . . . . 84<br />

4.2.2 Repumping laser . . . . . . . . . . . . . . . . . . . . . . 87<br />

4.2.3 Optical pumping laser . . . . . . . . . . . . . . . . . . . 88<br />

4.2.4 Optical paths on the experiment table . . . . . . . . . . 90<br />

4.3 The Vacuum System . . . . . . . . . . . . . . . . . . . . . . . . 94<br />

4.3.1 Experiment chamber . . . . . . . . . . . . . . . . . . . . 94<br />

4.3.2 Vacuum system . . . . . . . . . . . . . . . . . . . . . . . 95<br />

xiv


CONTENTS<br />

4.3.3 Procedure to reach UHV . . . . . . . . . . . . . . . . . . 96<br />

4.4 The Magnetic Field Coils . . . . . . . . . . . . . . . . . . . . . . 98<br />

4.4.1 Offset coils . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

4.4.2 Quadrupole coils . . . . . . . . . . . . . . . . . . . . . . 100<br />

4.4.3 Bias field coils . . . . . . . . . . . . . . . . . . . . . . . . 100<br />

4.5 The Atom Chip . . . . . . . . . . . . . . . . . . . . . . . . . . . 101<br />

4.5.1 Overall design . . . . . . . . . . . . . . . . . . . . . . . . 101<br />

4.5.2 Current-carrying wires . . . . . . . . . . . . . . . . . . . 103<br />

4.5.3 Magneto-optical film . . . . . . . . . . . . . . . . . . . . 105<br />

5 The Permanent Magnetic Chip Experiment: Results 109<br />

5.1 Overview and Timing Sequence . . . . . . . . . . . . . . . . . . 109<br />

5.2 The Magneto-Optical Traps . . . . . . . . . . . . . . . . . . . . 111<br />

5.3 The Wire Magnetic Trap . . . . . . . . . . . . . . . . . . . . . . 124<br />

5.3.1 Evaporation and BEC . . . . . . . . . . . . . . . . . . . 135<br />

5.4 The Permanent Magnetic Trap . . . . . . . . . . . . . . . . . . 145<br />

6 The All-Optical BEC Experiment: Apparatus 153<br />

6.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153<br />

6.2 The Vacuum System . . . . . . . . . . . . . . . . . . . . . . . . 156<br />

6.3 The Diode Laser Systems for Magneto-Optical Trapping . . . . 157<br />

6.3.1 Main laser . . . . . . . . . . . . . . . . . . . . . . . . . . 158<br />

6.3.2 Repumping laser . . . . . . . . . . . . . . . . . . . . . . 159<br />

6.3.3 Chirp lasers . . . . . . . . . . . . . . . . . . . . . . . . . 160<br />

6.4 The Dipole Trap . . . . . . . . . . . . . . . . . . . . . . . . . . 161<br />

6.5 Detection <strong>of</strong> Atoms . . . . . . . . . . . . . . . . . . . . . . . . . 162<br />

7 The All-Optical BEC Experiment: Results 165<br />

xv


CONTENTS<br />

7.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165<br />

7.2 The Magneto-Optical Trap . . . . . . . . . . . . . . . . . . . . . 166<br />

7.3 Loading the Dipole Trap . . . . . . . . . . . . . . . . . . . . . . 168<br />

7.4 Evaporative Cooling in the Dipole Trap . . . . . . . . . . . . . . 174<br />

8 Summary and Outlook 179<br />

8.1 Summary and Discussion . . . . . . . . . . . . . . . . . . . . . . 179<br />

8.2 Outlook and Future . . . . . . . . . . . . . . . . . . . . . . . . . 181<br />

A Reprints <strong>of</strong> selected Publications by the Candidate 185<br />

A.1 Asymmetric double-well potential for single-atom interferometry 186<br />

A.2 A permanent magnetic film atom chip for Bose-Einstein con-<br />

densation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191<br />

A.3 Perpendicularly magnetized, grooved GdTbFeCo microstructures<br />

for atom optics . . . . . . . . . . . . . . . . . . . . . . . . . . . 196<br />

B Determining the Temperature from One Image 199<br />

C The Atom: 87 Rb 201<br />

D Technical Details <strong>of</strong> the Coils for the Permanent Magnetic<br />

Chip Experiment 203<br />

E List <strong>of</strong> Equipment 205<br />

xvi


Chapter 1<br />

Introduction<br />

One revolution in physics which has had a big influence on everyday life was<br />

the discovery by Planck in 1900 <strong>of</strong> the quantisation <strong>of</strong> action, which led to<br />

quantum mechanics. In slightly more than one human lifetime this has led to<br />

the discovery <strong>of</strong> semi-conductors and their use in computation, and to lasers<br />

and their use in communication, medicine and many other fields. Even more<br />

impressive has been the impact <strong>of</strong> this discovery on physics and the understand-<br />

ing <strong>of</strong> Nature itself. Here, one <strong>of</strong> the most controversially discussed concepts<br />

is the <strong>of</strong>ten appearing duality. One manifestation was labelled “wave-particle-<br />

dualism”, as quantum mechanics allows classical waves (like electromagnetic<br />

waves) to behave like particles [Ein05] and, vice versa, particles to behave like<br />

waves [de 24]. The term is put in inverted commas as it will be later explained<br />

that this “dualism” is not an “exclusive or” in the boolean sense. The wave-<br />

like behaviour usually is connected to the particle’s ability to interfere with<br />

itself, while the particle-like behaviour <strong>of</strong> a wave usually is connected to the<br />

wave’s ability to scatter, to transfer momentum by that scattering and finally<br />

to be localisable.<br />

At first, this duality was explained by and contributed to another duality:<br />

the fundamental impossibility to measure both position and momentum with<br />

1


infinite accuracy [Hei27, Boh27]. Later it was shown that these two dualisms<br />

are not related to each other but both are fundamental in quantum mechan-<br />

ics [Scu91, Eng96]. Indeed the problem to measure wave-like or particle-like<br />

attributes is related closely to the number-phase uncertainty.<br />

This can be easily seen in terms <strong>of</strong> interferometric concepts. Once it is<br />

known which way is taken in the interferometer’s arms (knowledge <strong>of</strong> the num-<br />

bers) we lose the interference (knowledge <strong>of</strong> the phase) [Bru96, D¨98]. These<br />

experiments showed that imperfect knowledge <strong>of</strong> which way leads to a decrease<br />

in the contrast <strong>of</strong> the interference fringes to a correlated degree, and not to a<br />

total destruction <strong>of</strong> the fringes .<br />

Even before this interpretation <strong>of</strong> the wave-particle dualism, other concepts<br />

<strong>of</strong> classical waves were reformulated into quantum mechanics, such as the co-<br />

herent state 1 [Gla63], which corresponds to a monochromatic plane wave and<br />

is the state with a minimum uncertainty in both phase and number. This the-<br />

ory was first formulated for photons, where we have no number conservation.<br />

When dealing with atoms, we find the difficulty that the overall number <strong>of</strong><br />

atoms usually is not only constant but also very well defined. In a single cloud<br />

<strong>of</strong> atoms, we can in principle measure the exact number <strong>of</strong> atoms with zero<br />

uncertainty. Such a state is called a Fock state or number state and has no<br />

defined phase (see for example [Nol94]).<br />

This raises the question: if we can measure the exact number <strong>of</strong> atoms,<br />

how can we still create states with a defined phase for interferometry? How<br />

else can we talk about the wave-like properties <strong>of</strong> atoms but by the ability to<br />

use them in interference experiments?<br />

The answer to this question is given in the field <strong>of</strong> “atom optics”. The name<br />

itself implies atoms, and thus countable particles, with optics, the discipline <strong>of</strong><br />

2005.<br />

1 The quantum theory <strong>of</strong> coherence was awarded half <strong>of</strong> the Nobel prize for physics in<br />

2


Chapter 1: Introduction<br />

physics that totally relies on the wave-like properties <strong>of</strong> its subjects, and indeed<br />

this is the field where the wave character <strong>of</strong> the atoms becomes important. Of<br />

course, when talking <strong>of</strong> wave character, one must be able to parameterise the<br />

atoms by a wavelength. This is done by the so-called ‘de Broglie’ wavelength,<br />

named after the scientist who first postulated the wave-like behaviour [de 24].<br />

This wavelength is the ratio <strong>of</strong> Planck’s action quantum h and the momentum<br />

<strong>of</strong> the particle p, and thus is inversely proportional to the square root <strong>of</strong> the<br />

temperature T <strong>of</strong> a slowly moving particle:<br />

λdB = h<br />

p ∝<br />

�<br />

1<br />

T<br />

This means, the colder an ensemble <strong>of</strong> atoms is, the longer is the wavelength<br />

<strong>of</strong> the particles. Considering Rubidium atoms, at room temperature the wave-<br />

length is 27 pm, while for very low temperatures like 1 µK the corresponding<br />

wavelength is in the optical range <strong>of</strong> 467 nm. Reducing the temperature by<br />

another factor <strong>of</strong> 100 to 10 nK leads to a ten-fold increase in the wavelength<br />

to 4.7 µm, a wavelength that is very much larger than the ‘size’ <strong>of</strong> the atom<br />

in the particle picture.<br />

These temperatures are interesting for a physicist, as in one case the mo-<br />

mentum <strong>of</strong> an atom is comparable to the momentum <strong>of</strong> an interacting photon,<br />

the spectroscopist’s main tool. Then the measurement process is not negligi-<br />

ble anymore. The case <strong>of</strong> even lower temperatures is interesting in itself, as it<br />

seems to be a contradiction - or at least hard to imagine - that a wavelength<br />

can be much longer than the carrier wave’s extension.<br />

Reaching these very low temperatures is not trivial though. Today, tem-<br />

peratures in the region <strong>of</strong> a few µK or less can only be achieved using dilute<br />

gases. With the introduction <strong>of</strong> the magneto-optical trap (MOT) [Raa87] and<br />

the understanding <strong>of</strong> the mechanisms <strong>of</strong> how to cool atoms with the help <strong>of</strong><br />

laser radiation [H¨75, Win75, Dal89], it became possible to create samples <strong>of</strong><br />

3


several million up to several hundred million atoms at these very low tem-<br />

peratures. These break-throughs were rewarded with the award <strong>of</strong> the Nobel<br />

prize for physics to Steven Chu, William Phillips and Claude Cohen-Tannoudji<br />

in 1997. Having such cold atoms in such numbers allowed the creation <strong>of</strong><br />

several types <strong>of</strong> atom interferometers, which the field <strong>of</strong> metrology has bene-<br />

fited from. These interferometers ranged from devices to measure the Earth’s<br />

gravity [Pet99] and rotation [Gus97] to atomic clocks, which in principle are<br />

interferometers only in the time domain [Rus98, Wil02], and to interferome-<br />

ters that were used to measure atomic properties [Eks95] or natural constants<br />

[Gup02]. All <strong>of</strong> these interferometers work with free atoms, either falling or<br />

accelerated from the trap in which they were cooled, and the beamsplitters<br />

needed for interferometry are created by light pulses and the change <strong>of</strong> the<br />

atomic momentum on absorption <strong>of</strong> a photon [Kas91]. With the thermal mo-<br />

tion and kinetic energy being greatly reduced, it became possible to trap and<br />

confine large numbers <strong>of</strong> atoms with rather weak forces, stemming for example<br />

from the interaction between a magnetic field and the magnetic moment <strong>of</strong><br />

an atom or from the interaction <strong>of</strong> the induced atomic electric dipole-moment<br />

with an electro-magnetic field. With these magnetic and optical traps even<br />

more versatile atom interferometer set-ups are possible.<br />

Two main implementations <strong>of</strong> atom interferometers have been proposed: in<br />

the first example, an otherwise stationary trap is split into two, held there and<br />

recombined [Hin01]. The second example works with confined atoms that pass<br />

through wave guides and beam splitters. Only recently it has become possible<br />

to create these time dependent (in the first case) or spatially dependent (in the<br />

second case) potentials. Here either micron-sized optics and lenses are used to<br />

create optical traps which can be split and recombined [Dum02b] or waveguides<br />

are constructed [Dum02a]. For magnetic traps, the proposal [Wei95, Thy99]<br />

and realisation [Rei99, Ott01] <strong>of</strong> the so-called “atom chip” made the tailoring<br />

4


Chapter 1: Introduction<br />

<strong>of</strong> micron-sized potentials possible. Here micron-sized current-carrying wires<br />

on a chip, as known from microelectronics, induce the trapping fields. Both<br />

the time dependent and spatially dependent interferometer types are obtain-<br />

able with magnetic trapping or confinement. Switchable waveguides [M¨01],<br />

beam splitters [Cas00] and the splitting and recombining <strong>of</strong> traps [Est05] have<br />

all been demonstrated. It has to be noted that splitting and merging with<br />

macroscopic traps was also achieved [Tho02, Tie02], but atom chips allow<br />

more precise control and thus are favoured.<br />

Another break-through in the field <strong>of</strong> atom optics, which may prove very<br />

useful for atom interferometry, was the creation <strong>of</strong> Bose-Einstein Condensates<br />

(BEC) from dilute gases <strong>of</strong> alkali atoms [And95, Dav95, Bra95]. The existence<br />

<strong>of</strong> this new state <strong>of</strong> matter was first postulated by S. N. Bose and A. Einstein in<br />

1925 [Ein25], and the experimental pro<strong>of</strong> <strong>of</strong> this phase transition was realised<br />

70 years later, resulting in the award <strong>of</strong> the 2001 Nobel prize in Physics to E.<br />

Cornell, C. Wieman and W. Ketterle. In simple words, the phase transition<br />

from a thermal cloud <strong>of</strong> atoms to the condensate occurs when the de Broglie<br />

wavelength <strong>of</strong> the atoms is larger than the mean distance between them. The<br />

waves then overlap and become coherent. In the end, the waves <strong>of</strong> all atoms<br />

oscillate in phase and a macroscopic coherent matter state is obtained. This<br />

transition was found for dilute gases <strong>of</strong> atoms with temperatures <strong>of</strong> a few hun-<br />

dred nanokelvin, as explained above. Shortly after the experimental realisation<br />

<strong>of</strong> a BEC, theoreticians began investigating its behaviour in double-well po-<br />

tentials, which are traps that are split into two [Mil97, Sme97, Jav97]. These<br />

results were then used to examine the feasibility <strong>of</strong> interferometry with a BEC<br />

[Men01]. An early result was that the splitting <strong>of</strong> the BEC will have a thresh-<br />

old. At this point, it will become virtually impossible for atoms to tunnel from<br />

one well to the other and the condensate “fragments” [Spe99]. This will fix<br />

the atom number in each well, and the coherence between the two wells will be<br />

5


lost—making any meaningful interferometry impossible after this point when<br />

the atoms are localised in the left or right well. This collapse into separate<br />

BECs with undefined relative phase was experimentally seen not only for two<br />

wells [Shi05] but also for an array <strong>of</strong> traps [Gre02a, Gre02b]. These results<br />

seem to indicate that the splitting <strong>of</strong> the BEC has to stay in the so-called<br />

Josephson regime to be useful for interferometry [Jav86, Sme97, Sak02], or<br />

that the whole process <strong>of</strong> splitting, phase evolution and recombining for the<br />

measurement has to be done on short timescales to maintain the coherence<br />

between the wells. The coherence between the wells will only survive as long<br />

as the atom numbers are undefined. One way to achieve this is to split only<br />

so far that some tunnelling is still allowed, so that the atom numbers <strong>of</strong> the<br />

wells fluctuate and are not defined, like in the the above-mentioned Josephson<br />

regime. Another way to keep the coherence is to split the single atoms between<br />

the wells. As long as localisation can be prevented, the atoms in the two wells<br />

then are coherent. However, a BEC consists <strong>of</strong> interacting particles, and these<br />

interactions can be interpreted as measurements <strong>of</strong> the location <strong>of</strong> the atoms.<br />

This limits the coherence time, but there are ways to reduce the interaction<br />

strength by so-called Feshbach resonances [Mar02, Web03, Shi04].<br />

Theoretical work is continuing on this topic, examining the splitting pro-<br />

cess more closely [Bor04] for weakly interacting atoms and ways are being re-<br />

searched on how to increase the sensitivity <strong>of</strong> such an interferometer [Neg04].<br />

Other proposals use the BEC only as a means to create atoms in the ground<br />

state <strong>of</strong> the potential well. Atoms that do not interact with each other, because<br />

they have a scattering length <strong>of</strong> zero, will overcome some <strong>of</strong> the problems <strong>of</strong><br />

the interferometers proposed with interacting BECs. These then in principle<br />

work like multiple single atom interferometers [Dud03], where the single atom<br />

interferometers are the ones proposed for atom chips [Hin01, H¨01c]. Other<br />

6


Chapter 1: Introduction<br />

proposals use the BEC and then use techniques that are beamsplitters in mo-<br />

mentum space rather than in real space [Pou02].<br />

A working interferometer using a BEC and the splitting and recombining<br />

<strong>of</strong> the trap has been experimentally demonstrated [Shi04, Jo07]. In [Shi04]<br />

a limit on the coherence time between the wells was found, after which the<br />

relative phase became undefined. Another similar experiment has shown this<br />

limitation as well, by having random positions <strong>of</strong> the interference fringes from<br />

one experimental run to another [Shi05]. The traps in these two experiments<br />

are comparable for the important parameters; the main difference is that the<br />

splitting in the second experiment takes about 40 times the coherence time<br />

that was determined in the first experiment. How such an increase can affect<br />

a single atom double well device and destroy any interferometric signal is<br />

covered in chapter 3 <strong>of</strong> this thesis. A BEC <strong>of</strong> interacting particles has even<br />

more possible effects that lead to localisation and the loss <strong>of</strong> a defined phase<br />

relation. Nevertheless the chapter presented here (Chapter 3) can already<br />

give an overall idea and lead to a new “gut feeling” <strong>of</strong> suitable timescales<br />

for experiments. In another experiment, a Michelson interferometer using a<br />

magnetic wave guide has also been demonstrated [Wan05b]. Here optical pulses<br />

acted as beam splitters in momentum space, in exactly the same way as the<br />

interferometers with cold atoms in free space which were described above.<br />

This thesis is organised in the following way. Chapter 2 gives information<br />

about the theoretical background that is needed to understand all subsequent<br />

chapters. This background is well known and can be found in textbooks for<br />

graduate students, for example [Met99, Mes90, Mey91, Pet02]. In later sections<br />

<strong>of</strong> the chapter recent developments play a stronger role and the respective<br />

publications are discussed there. The chapter contains no new material.<br />

Chapter 3 deals with a theoretical analysis <strong>of</strong> the single atom interferome-<br />

ter, and early results <strong>of</strong> this work have been published by the author [Sid06].<br />

7


Here a model is presented which in contrast to most publications [Hin01, H¨01c]<br />

incorporates an unavoidable experimental effect: the imperfect symmetry <strong>of</strong><br />

the double well potential. A two mode approximation is developed for this<br />

case. The analogy with all two-level systems allows the use <strong>of</strong> the Bloch equa-<br />

tions. Their validity was checked with the numerical results <strong>of</strong> the full system,<br />

performed by S. Whitlock [Whi04] using the XMDS s<strong>of</strong>tware developed at the<br />

<strong>University</strong> <strong>of</strong> Queensland. The simplicity <strong>of</strong> the Bloch model allows simula-<br />

tions on a much faster scale and application to many related problems. The<br />

presented two mode model is well known [All75, Mey91]. To the author’s<br />

knowledge the application to this kind <strong>of</strong> interferometer is new. The inter-<br />

pretation <strong>of</strong> the model using the Bloch sphere and the solutions <strong>of</strong> the Bloch<br />

equations give further insight into the physics that occurs in such an interfer-<br />

ometer, identifying the main process as Larmor precessions. It is shown that<br />

an asymmetry is crucial and needed for an interferometer to work, and should<br />

be incorporated in any model, as it gives rise to the phase shift that the in-<br />

terferometer measures. The model also shows how experimental imperfections<br />

that translate into asymmetries in the potential can be overcome, clarifying<br />

a common misconception among experimental physicists about adiabatic pro-<br />

cesses and their role in atom interferometry. It also explains where and how<br />

the change from phase difference to measurable number difference is mani-<br />

fested. Although the model strictly is applicable to single atoms only, the<br />

main results have to be kept in mind for BEC-based interferometers. A/Pr<strong>of</strong>.<br />

B. Dalton is acknowledged for correcting a factor in the Bloch equations. Two<br />

main limits <strong>of</strong> the model are identified by comparing the two mode model<br />

with the full numerical analysis. A new way to read out a double well atom<br />

interferometer was inspired by questions <strong>of</strong> Pr<strong>of</strong>. P. Drummond <strong>of</strong> the Uni-<br />

versity <strong>of</strong> Queensland and is based on an idea <strong>of</strong> Pr<strong>of</strong>. A. Sidorov. A related<br />

read out process was later successfully implemented in a BEC-based double<br />

8


Chapter 1: Introduction<br />

well system [Hal07b] on the permanent magnetic chip described in chapters 4<br />

and 5. The model is kept simple, so that different experimental parameters<br />

can be implemented easily. Even though it works with single atoms in double<br />

well, we find a surprisingly close match with the experimental data for a BEC<br />

in a double well [Jo07]. These two experimental publications emphasise the<br />

importance <strong>of</strong> this chapter, although it was mainly intended not as a tool for<br />

prediction, but to give a deeper understanding <strong>of</strong> the actual physical processes<br />

in double well potentials. Further applications to double well problems outside<br />

<strong>of</strong> atom optics are briefly discussed. Although the author is not first author<br />

on a publication that reports the early results [Sid06], the work on which this<br />

publication is based was mainly undertaken by him.<br />

Chapters 4 and 5 contain the description <strong>of</strong> the technical details and the<br />

results <strong>of</strong> the experiment which the author worked on at <strong>Swinburne</strong> <strong>University</strong><br />

<strong>of</strong> <strong>Technology</strong>. Chapter 4 contains a full description <strong>of</strong> the experimental set-up<br />

<strong>of</strong> the permanent magnetic film atom-chip experiment. Here the experimental<br />

apparatus including the laser systems, the vacuum chamber and the novel<br />

hybrid atom chip with both current-carrying wires and a permanent magnetic<br />

film are presented and characterised. This work was done mainly in association<br />

with S. Whitlock and Dr. B. Hall. In the early stages this part <strong>of</strong> the work was<br />

supported by D. Gough and Pr<strong>of</strong>. A. Sidorov. Results have been published<br />

covering the magnetic film and its properties [Wan05a] and experimental data<br />

with some technical details <strong>of</strong> the set-up [Hal06]. The TbGdFeCo magneto-<br />

optical film was fabricated by J. Wang, S. Whitlock and D. Gough. As the<br />

setting up <strong>of</strong> this unique experiment is an integral part <strong>of</strong> this thesis, the set-up<br />

and the individual parts are described in some detail.<br />

After the characterisation <strong>of</strong> the experimental set-up, the analysis <strong>of</strong> the<br />

data taken with the new permanent-magnetic-film atom-chip is presented in<br />

Chapter 5. With the help <strong>of</strong> this chip, atoms were first collected and cooled in a<br />

9


mirror MOT and then transferred into a current-carrying wire magnetic trap,<br />

in which the atomic sample was further cooled by allowing the atoms with<br />

highest kinetic energy to “evaporate” from the trap [Hes86, Mas88]. With<br />

this cooling, the phase transition to a quantum degenerate BEC was achieved.<br />

Then, with the current in the wires turned <strong>of</strong>f, trapping <strong>of</strong> ultracold atoms<br />

in the field <strong>of</strong> the permanent magnetic structure was successfully achieved<br />

and these results are presented at the end <strong>of</strong> this chapter. A BEC was then<br />

successfully realised on a similar permanent magnetic structure [Hal05, Hal06]<br />

while the author was in Hannover working on the all-optical BEC experiment<br />

described in Chapters 6 and 7. The measurements in this chapter were taken<br />

together with S. Whitlock and Dr. B. Hall, and have been published in [Hal06].<br />

Chapters 6 and 7 present the experimental set-up and the results <strong>of</strong> the<br />

experiment at the <strong>University</strong> <strong>of</strong> Hannover, where the author participated in<br />

the experiment for limited periods. A tri-national network grew out <strong>of</strong> this<br />

collaboration. In Chapter 6 the set-up <strong>of</strong> the all-optical BEC experiment is<br />

described. This apparatus was set up and first used by Dr. F. Buchkremer<br />

for the coherence investigations in optical waveguides and traps [Buc01]. The<br />

following experiment implemented miniaturised optical elements like arrays<br />

<strong>of</strong> micron-sized lenses to demonstrate their in-principle use for quantum in-<br />

formation and atom interferometry [Dum03a]. The author first joined the<br />

experiment at that stage. An existing experimental set-up was modified and<br />

improved, and the author was involved in experiments and measurements with<br />

the aim <strong>of</strong> producing an all-optical BEC. This work was done together with Dr.<br />

R. Dumke in the very beginning. Most <strong>of</strong> the changes were implemented after<br />

Dr. Dumke had left and then optimised together with A. Lengwenus [Len04].<br />

While the author was absent, the dipole laser system that is now used to trap<br />

the atoms for evaporation was set up by Dr. T. Müther, J. Nes and A.-L.<br />

Gehrmann. A description <strong>of</strong> the full apparatus can be found in their theses<br />

10


Chapter 1: Introduction<br />

[M¨05, Geh05]. In this thesis the description is kept short, and the interested<br />

reader is referred to the above-mentioned theses for more detailed descriptions.<br />

In Chapter 7 the results from the all-optical BEC experiment are presented.<br />

As a non-resident researcher the author worked with several colleagues over<br />

the time. The collection and pre-cooling <strong>of</strong> the atomic sample in a MOT<br />

is characterised first. This work was done together with A. Lengwenus and<br />

has also been published in his Diploma thesis [Len04]. The results from the<br />

optical trapping, the optimised loading <strong>of</strong> atoms into this trap and the evap-<br />

orative cooling were achieved together with Dr. T. Müther, J. Nes and A.-L.<br />

Gehrmann and published in their respective theses [M¨05, Geh05].<br />

The thesis ends with a summary and discussion and an outlook for possible<br />

future experiments.<br />

11


Chapter 2<br />

Theoretical Background<br />

In the branch <strong>of</strong> atomic physics called “Atom Optics” we deal with cold and<br />

ultra-cold atoms. To exploit the wave like properties <strong>of</strong> atoms [de 24] the<br />

thermal de Broglie wavelength <strong>of</strong> the atoms λdB defined by<br />

λdB = h/p = h/ � 3m · kBT (2.1)<br />

has to be increased by cooling down the atoms while confining them in a trap.<br />

The mechanisms <strong>of</strong> how to cool and how to trap these atoms are thus <strong>of</strong> main<br />

interest. To understand these mechanisms, one has to understand the principal<br />

interactions between the atoms and the fields that are used to trap and cool<br />

the atoms. Using light and magnetic fields for these purposes, we are in the<br />

fortunate position that this theory is well understood.<br />

This chapter will introduce the general theory and emphasize and explain<br />

certain points in more detail, which are needed for a full understanding <strong>of</strong> the<br />

experiments that are the main part <strong>of</strong> this thesis. This chapter starts with<br />

an explanation <strong>of</strong> the interaction between an atom and electromagnetic waves.<br />

Here we need to differ between two main interactions: one that allows the<br />

trapping <strong>of</strong> atoms, and one that allows us to cool and to detect our atomic<br />

sample. The detection <strong>of</strong> atoms will also be addressed here. The next section<br />

13


2.1. Atoms and Electromagnetic Fields<br />

covers the interaction <strong>of</strong> the atoms with constant magnetic fields, which are<br />

another means to trap the atoms. It will then be explained how one can cool<br />

the atoms to quantum degeneracy; here the removal <strong>of</strong> energy from the sample<br />

is coupled to the removal <strong>of</strong> particles. A combination <strong>of</strong> magnetic interaction<br />

and optical forces leads to the magneto-optical trap, which is discussed in the<br />

next section. The experiments discussed in this thesis have slightly different<br />

arrangements for this trap, so both are presented here. The chapter ends with<br />

a section on a main application <strong>of</strong> cold atoms and atom optics: using the wave<br />

like properties <strong>of</strong> cold atoms to perform interferometry.<br />

2.1 Atoms and Electromagnetic Fields<br />

The interaction between atoms and electromagnetic fields is a topic which can<br />

be treated in a nearly arbitrarily detailed fashion [CT97, CT92]. Although,<br />

to understand the basic concepts, some assumptions can be made to simplify<br />

the problem and not lose too much information. If the spectral linewidth <strong>of</strong><br />

the light is small enough so that only two atomic energy levels are coupled<br />

by the light, then other atomic energy levels which do not contribute to the<br />

problem can be ignored, reducing the complexity [Mey91]. Further, one can<br />

ignore the quantised nature <strong>of</strong> the light, assuming classical fields that obey<br />

Maxwells equations: if the light is coherent and <strong>of</strong> sufficient intensity, then<br />

the light field can be well approximated by a classical field [Gla63]. In the<br />

experiments described here, we work with laser light that fulfills both these<br />

conditions. The first simplification will be used to explain how light can create<br />

a conservative potential that can be used to trap atoms. Both simplifications<br />

are needed for the simple model <strong>of</strong> absorption and scattering explained in the<br />

following section. This model in turn is needed to understand the mechanisms<br />

<strong>of</strong> cooling atoms optically [Win79].<br />

14


2.1.1 The dressed state model<br />

Chapter 2: Theoretical Background<br />

To describe the situation <strong>of</strong> an atom in a light field correctly, one needs to<br />

treat the atom, the light field, and their interaction quantum mechanically.<br />

This leads to the so-called dressed states model [Dal85], which will be used<br />

later (section 2.1.4) to explain how atoms can be trapped with light. In the<br />

following, that derivation will be outlined.<br />

If the light is monochromatic, in the sense that the width <strong>of</strong> the frequency<br />

distribution <strong>of</strong> the light is small compared to the energy difference between<br />

atomic levels, the atomic polarisability can be calculated from an ansatz that<br />

treats the atom as a two-level quantum system. The ground state is denoted<br />

by |g〉, the only excited state by |e〉. The Hamiltonian <strong>of</strong> this atomic system<br />

is then<br />

HA = ¯hω0|e〉〈e| (2.2)<br />

where the energy <strong>of</strong> the ground state has been set to zero and the energy<br />

difference between the levels is ¯hω0. If the light has frequency ωL = ω0 + ∆, it<br />

can be described by the Hamiltonian<br />

HL = ¯hωL(a † a + 1<br />

) (2.3)<br />

2<br />

where a † and a are the creation and annihilation operators, and ˆn = a † a is the<br />

number operator with eigenvalue n, the number <strong>of</strong> photons in the field. It is<br />

then said that the light is detuned by ∆ with respect to the transition.<br />

To include the interaction <strong>of</strong> atom and light, we need a third operator<br />

VAL = − � d · � E(�r) (2.4)<br />

Here � d is the operator <strong>of</strong> the induced electric dipole moment <strong>of</strong> the atom and<br />

�E(�r) is the operator <strong>of</strong> the electric field strength at position �r.<br />

15


2.1. Atoms and Electromagnetic Fields<br />

The overall Hamiltonian <strong>of</strong> the interacting atom-light system is thus<br />

H = HA + HL + VAL<br />

= ¯hω0|e〉〈e| + ¯hωL(a † a + 1<br />

2 ) − � d · � E(�r) (2.5)<br />

We now apply the rotating wave approximation, only allowing transitions be-<br />

tween the nearly degenerated states |g, n〉 and |e, n−1〉. There are cases where<br />

this approximation is not justified. This will be addressed in equation (2.13)<br />

later this section. For now we can assume it to hold. The transition matrix<br />

element is then<br />

〈e, n − 1|VAL|g, n〉 = ¯hΩR<br />

2<br />

Here ΩR = d · E/¯h is the Rabi frequency.<br />

(2.6)<br />

The interaction lifts the degeneracy, so that the eigenstates <strong>of</strong> Hamiltonian<br />

(2.5) are linear combinations <strong>of</strong> the eigenstates <strong>of</strong> the uncoupled system |{g, e}, {n, n−<br />

1}〉<br />

|1, n〉 = cos θ|e, n − 1〉 − sin θ|g, n〉<br />

|2, n〉 = sin θ|e, n − 1〉 + cos θ|g, n〉 (2.7)<br />

Here we use the generalised Rabi frequency<br />

to define the angle θ by<br />

Ω =<br />

cos 2θ = − ∆<br />

Ω<br />

�<br />

∆ 2 + Ω 2 R<br />

and sin 2θ = ΩR<br />

Ω<br />

(2.8)<br />

(2.9)<br />

The eigenvalues <strong>of</strong> the energy <strong>of</strong> these dressed states are not degenerate any-<br />

more (see also Fig. 2.1):<br />

E1,n = (n + 1)¯hωL + ¯h<br />

(Ω − ∆)<br />

2<br />

E2,n = (n + 1)¯hωL − ¯h<br />

(Ω + ∆)<br />

2<br />

(2.10)<br />

16


Chapter 2: Theoretical Background<br />

The difference in the energies ±¯hΩ due to the applied light field is called the<br />

ac-Stark shift. In the limit <strong>of</strong> small frequencies ΩR, the Ω in equation (2.8) can<br />

be expanded around ΩR/(4|∆|). This results in energy shifts <strong>of</strong> ¯hΩ 2 R /(4|∆|)<br />

<strong>of</strong> the ground and excited state. For negative detunings ∆ < 0 (red detuned<br />

light), the energy <strong>of</strong> the ground state is lowered by this amount, while the<br />

energy <strong>of</strong> the excited state is raised by the same amount. In this limit we have<br />

θ → 0, and the dressed states become identical to their respective unperturbed<br />

eigenstates.<br />

Ε<br />

e, n<br />

g, n+1<br />

e, n-1<br />

g, n<br />

h∆<br />

h∆<br />

ω<br />

h 0<br />

Figure 2.1: Energy <strong>of</strong> the system atom-light without interaction between<br />

the atomic states and the light field (left side) and including the interaction<br />

(“dressed states”, right side). The notation is explained in the text. The cou-<br />

pling increases the energy gap <strong>of</strong> the doublet states and in case <strong>of</strong> resonance<br />

will cause an “avoided crossing”.<br />

hΩ<br />

hΩ<br />

1, n<br />

2, n<br />

1, n-1<br />

2, n-1<br />

The Rabi frequency <strong>of</strong> a light field is a function <strong>of</strong> its intensity,<br />

Ω 2 R = 1<br />

2 Γ2 · I<br />

17<br />

I0<br />

(2.11)


2.1. Atoms and Electromagnetic Fields<br />

where Γ is the linewidth <strong>of</strong> the atomic transition and I0 = ¯hΓω3 0<br />

12πc 2 the saturation<br />

intensity. We can now calculate the change <strong>of</strong> the energy between the two<br />

levels, as a function <strong>of</strong> intensity and detuning, and get<br />

∆E = U = 3πc2<br />

2ω3 Γ ·<br />

0<br />

I<br />

∆<br />

∝ I<br />

∆<br />

(2.12)<br />

It needs to be remembered that the potential depth scales with U ∝ I/∆,<br />

proportional to the intensity and inversely proportional to the detuning.<br />

In the above, the rotating wave approximation was used when evaluating<br />

the Hamiltonian <strong>of</strong> equation (2.5). This approximation is generally valid if the<br />

difference between the frequency <strong>of</strong> the atomic transition ω0 and <strong>of</strong> a photon ωL<br />

is small. In the correct quantum mechanical treatment, a second term appears<br />

which does not contain the difference but the sum <strong>of</strong> the frequencies in the<br />

denominator. For similar frequencies, especially in the optical regime, the<br />

difference term dominates over the sum term. In the case <strong>of</strong> our experiment,<br />

the transition has a wavelength <strong>of</strong> 780 nm ( 87 Rb D2) while the laser light has a<br />

nearly 300 nm longer wavelength. Here the influence <strong>of</strong> the second term with<br />

the sum is no longer negligible. Instead, we find a corrected energy shift<br />

�<br />

�<br />

Γ Γ<br />

+ · I(�r) (2.13)<br />

∆E(�r) = 3πc2<br />

2ω 3 0<br />

ω0 − ωL<br />

ω0 + ωL<br />

The scattering rate which will be introduced later in equation (2.27) also needs<br />

to be corrected in the same way and the term Γ/∆ be replaced by the sum.<br />

In our case neglecting the counter rotating term leads to an approximation <strong>of</strong><br />

the potential depth that is more than 10% too small, while the scattering rate<br />

in the rotating wave approximation results in a value that is nearly 30% too<br />

small.<br />

2.1.2 Absorption and emission <strong>of</strong> photons<br />

A simple way to describe the absorption and emission <strong>of</strong> photons by atoms is<br />

the density matrix formalism. The quantum mechanical state Ψ <strong>of</strong> the atomic<br />

18


system is expanded into<br />

Ψ = �<br />

i=1,2<br />

Chapter 2: Theoretical Background<br />

ciφi<br />

where the coefficients ci are complex and normalise the state.<br />

(2.14)<br />

A pure state is defined by the density matrix ρ = |Ψ〉〈Ψ|. The time evolution <strong>of</strong><br />

the density matrix for a Hamiltonian H is given by the von Neumann equation:<br />

i¯h dρ<br />

dt<br />

= [H, ρ] (2.15)<br />

As we restrict ourselves to an atom with two levels only, with |φ1〉 = |g〉<br />

the ground state and |φ2〉 = |e〉 the excited state, the density matrix can then<br />

easily be written down<br />

ρ =<br />

⎛<br />

⎝ ρee ρeg<br />

ρge ρgg<br />

⎞<br />

⎠ =<br />

⎛<br />

⎝ |ce| 2 cec ∗ g<br />

cgc ∗ e |cg| 2<br />

⎞<br />

⎠ (2.16)<br />

Normalisation requires |cg| 2 +|ce| 2 = ρgg +ρee = 1. We can now insert equation<br />

(2.16) into equation (2.15). To include spontaneous emission <strong>of</strong> photons by the<br />

atoms, we include a finite lifetime Γ −1 <strong>of</strong> the excited state. The time evolution<br />

<strong>of</strong> the two level system can now be written down as [All75]:<br />

dρgg<br />

dt = +Γρee + i<br />

2 (Ω∗R ˜ρeg − ΩR ˜ρge)<br />

dρee<br />

dt = −Γρee + i<br />

2 (Ω∗R ˜ρge − ΩR ˜ρeg)<br />

d˜ρge<br />

dt<br />

d˜ρeg<br />

dt<br />

� �<br />

Γ<br />

= − + i∆<br />

2<br />

� �<br />

Γ<br />

= − + i∆<br />

2<br />

˜ρge + i<br />

2 Ω∗ R(ρee − ρgg)<br />

˜ρeg + i<br />

2 ΩR(ρgg − ρee) (2.17)<br />

Here ˜ρij = ρije −i∆t , with ∆ being the detuning. The Rabi frequency ΩR is<br />

defined by equation (2.6).<br />

Equations (2.17) are called the optical Bloch equations. Using the inversion<br />

w = ρgg − ρee, the normalisation and the fact that ρeg = ρ ∗ ge, we can simplify<br />

19


2.1. Atoms and Electromagnetic Fields<br />

them to<br />

dw<br />

dt = Γ(1 − w) − i(ΩRρ ∗ eg − Ω ∗ Rρeg)<br />

dρeg<br />

dt<br />

= −<br />

� Γ<br />

2<br />

�<br />

− i∆ ρeg + iΩR<br />

w (2.18)<br />

2<br />

The steady state solutions <strong>of</strong> these equations are retrieved by setting the time<br />

derivatives to zero:<br />

w =<br />

ρeg =<br />

1<br />

1 + s<br />

iΩR<br />

2(Γ/2 − i∆)(1 + s)<br />

Here, we have introduced the saturation parameter s which is given by<br />

s =<br />

|ΩR| 2<br />

2|Γ/2 − i∆| 2 = |ΩR| 2 /2<br />

Γ2 /4 + ∆2 (2.19)<br />

(2.20)<br />

(2.21)<br />

The saturation intensity I0 for a transition <strong>of</strong> wavelength λ into an excited<br />

state <strong>of</strong> lifetime τ = Γ −1 is<br />

I0 = πhc<br />

3λ 3 τ<br />

The saturation parameter s0 for zero detuning, ∆ = 0, is<br />

2 |ΩR|<br />

s0 = 2<br />

Γ2 I<br />

=<br />

Using these definitions equation (2.21) simplifies to<br />

s =<br />

I0<br />

s0<br />

1 + (2∆/Γ) 2<br />

(2.22)<br />

(2.23)<br />

(2.24)<br />

From equation (2.19) we see that with increasing saturation s, the steady<br />

state changes from a high occupation <strong>of</strong> the ground state (w = 1) to an<br />

equal occupation <strong>of</strong> both states w = 0 in the limit <strong>of</strong> infinite saturation. The<br />

occupation probability for the excited state is given by a Lorentzian<br />

ρee = 1<br />

(1 − w) =<br />

2<br />

s<br />

2(1 + s) =<br />

20<br />

s0/2<br />

1 + s0 + (2∆/Γ) 2<br />

(2.25)


Chapter 2: Theoretical Background<br />

In the steady state, the decay and excitation <strong>of</strong> the higher state are equal,<br />

and the decay rate Γ is known. We can now determine the rate <strong>of</strong> scattering<br />

processes Γsc for this case:<br />

Γsc = Γ · ρee = 1 Γs0<br />

2 1 + s0 + (2∆/Γ) 2<br />

(2.26)<br />

The intensity dependence <strong>of</strong> the scattering rate is implicit in the saturation<br />

parameter s0. Substituting the s0 yields<br />

Γsc = 3πc2<br />

2¯hω 3 0<br />

� Γ<br />

∆<br />

� 2<br />

· I ∝ I<br />

∆ 2<br />

(2.27)<br />

We see that the rate <strong>of</strong> photons scattered by the atoms depends linearly on<br />

the intensity <strong>of</strong> the illuminating light and is proportional to the reciprocal <strong>of</strong><br />

the square <strong>of</strong> the detuning.<br />

2.1.3 Detection <strong>of</strong> atoms by fluorescence and absorption<br />

From equation (2.27) we can calculate the scattered light power <strong>of</strong> a single<br />

atom to by multiplying the scattering rate with the energy <strong>of</strong> a single photon<br />

<strong>of</strong> the resonance frequency ω0<br />

PA = ¯hω0 · 3πc2<br />

2¯hω 3 0<br />

� �2 Γ<br />

· I (2.28)<br />

∆<br />

For N atoms, the total scattered power is thus Pt = N · PA. If this is not<br />

spatially resolved but imaged into a photodiode, this photodiode will detect a<br />

power PPD<br />

PPD = Pt · πr2<br />

∝ N (2.29)<br />

4πa2 Here, r denotes the radius <strong>of</strong> the lense used for imaging, and a is the distance<br />

between the atomic cloud and this lense. A single photodiode is sufficient to<br />

measure the atom number.<br />

A better signal to noise ratio in the detection <strong>of</strong> atoms is possible when a low<br />

intensity resonant beam shines though the atomic sample and then is detected.<br />

21


2.1. Atoms and Electromagnetic Fields<br />

It is important that the signal does not contain spontaneously emitted photons.<br />

This is the case when the spatial angle that is gathered by the imaging optics<br />

is much smaller than 4π.We will consider only a single pixel <strong>of</strong> a CCD camera<br />

here, the same argument then also holds for central column density absorption<br />

measurements onto a single photodiode.<br />

We consider an area A inside the atomic cloud, and an initial intensity <strong>of</strong><br />

light IA illuminating this area. Due to the absorption, this intensity is reduced<br />

by ∆I = −PA/A when passing the cloud. Here PA is the scattered power per<br />

atom, see equation (2.28). The measurable is the intensity IA(NA) after NA<br />

absorptions, NA is the number <strong>of</strong> atoms in a column through the cloud with<br />

area A. For a laser frequency <strong>of</strong> ωL and any number <strong>of</strong> atoms N, we receive<br />

the rate equation<br />

dIA(N)<br />

dN<br />

= ¯hωLΓ<br />

2A<br />

IA(N)/I0<br />

1 + IA(N)/I0 + (2∆/Γ) 2<br />

(2.30)<br />

This can be solved for the number <strong>of</strong> atoms as a function <strong>of</strong> the attenuated<br />

intensityIA by substituting N by NA<br />

NA = 2AI0<br />

¯hωLγ<br />

��<br />

1 +<br />

� � �<br />

2 � �<br />

2∆ IA(0)<br />

· ln<br />

+<br />

Γ IA(NA)<br />

IA(0)<br />

�<br />

− IA(NA)<br />

I0<br />

(2.31)<br />

For a weak absorption IA ≪ I0 the last difference term can be neglected. It<br />

has to be noted that the polarisation <strong>of</strong> the absorbed beam influences the<br />

saturation intensity I0. Linear polarised light does not optically pump the<br />

atoms into just one magnetic substate, so that more than one sublevel can<br />

contribute to the transition strength, and the saturation intensity is larger<br />

than for circular polarised light. For a camera, the overall number <strong>of</strong> atoms<br />

in the cloud is then given by summing over all pixels. An extensive overview<br />

on the detection and probing <strong>of</strong> Bose-Einstein condensates can be found in<br />

[Ket99].<br />

22


2.1.4 Trapping <strong>of</strong> atoms in light fields<br />

Chapter 2: Theoretical Background<br />

A spatial varying intensity <strong>of</strong> light I(�r) leads to a trap if the potential has a<br />

local minimum. For red detuned light this is reached by a local maximum <strong>of</strong><br />

the intensity, while for blue detuned light (∆ > 0) the atoms can be trapped in<br />

a local minimum. Figure 2.2 schematically shows how such a trap works by a<br />

spatially dependent shift <strong>of</strong> the energy levels <strong>of</strong> the atoms. In the experiment<br />

described here (chapters 6, 7) we work with red detuned light only. Thus,<br />

trapping occurs in local maxima <strong>of</strong> the intensity. For this, a single focused<br />

beam is already sufficient, with the radial confinement given by the waist <strong>of</strong><br />

the beam and the axial confinement due to the Rayleigh range. This set-up<br />

was used in the first experimental realisation <strong>of</strong> an optical dipole trap [Chu86].<br />

Figure 2.2: A spatially dependent intensity <strong>of</strong> light can be used to trap atoms.<br />

The intensity <strong>of</strong> a red detuned Gaussian beam is causing a trapping potential<br />

by the ac-Stark effect. In high intensity areas the atomic ground state is<br />

shifted to lower energies while in low intensity areas it remains unperturbed<br />

and remains at a relatively higher energy.<br />

Two crossed beams with perpendicular polarisation and same focal char-<br />

acteristics have the advantage that the confinement is much more isotropic.<br />

23


2.1. Atoms and Electromagnetic Fields<br />

The confinement in each axis is given by the waist <strong>of</strong> the beams when their<br />

foci overlap. The potential is twice as deep as that <strong>of</strong> a single beam <strong>of</strong> equal<br />

power, but usually the crossed beam is created from the two beams split <strong>of</strong>f<br />

one single beam and their intensity is only half as big as the intensity <strong>of</strong> the<br />

beam they originate from. Also, one has to note that the particles need a lower<br />

energy than the trap depth to escape the trap. The beams themselves have a<br />

finite intensity outside the crossing region and the atoms can use these arms<br />

as escape routes. This leads to an effective trap depth that is half the absolute<br />

value, or equal to the trap depth that a single uncrossed beam <strong>of</strong> the same<br />

power as one <strong>of</strong> the crossed beams would have.<br />

In the experiments that use dipole traps to evaporate atoms to create a<br />

BEC, crossed beam configurations are used. The experimental set-ups differ in<br />

the lasers used [Bar01], how the confinement is created by two beams [Web03,<br />

Kin05]. A common problem in evaporation with optical traps is that ramping<br />

down the intensity also reduces the gradient <strong>of</strong> the trap: the focal width <strong>of</strong><br />

the light which gives the radial width <strong>of</strong> the trap remains unchanged. A less<br />

deep trap with the same radius is shallower, the gradient and curvature <strong>of</strong> the<br />

trap do not stay constant over the forced evaporation but decrease. For a good<br />

overview on different trapping designs, see [Gri00].<br />

2.1.5 Cooling atoms with light<br />

Doppler cooling<br />

As the photons not only carry an energy E = ¯hω = hc/λ but also a momentum<br />

�p = ¯h � k (2.32)<br />

where � k is the wavevector with absolute value k = | � k| in the propagation<br />

direction <strong>of</strong> the light, it was proposed by [H¨75, Win75] to use light for the<br />

deceleration and cooling <strong>of</strong> atoms.<br />

24


Chapter 2: Theoretical Background<br />

During each scattering event, both energy and momentum have to be con-<br />

served. The force that an atom experiences in a light field is the product <strong>of</strong><br />

the scattering rate and the momentum that is transferred to the atom in each<br />

scattering process.<br />

�F = ¯h Γ<br />

2<br />

s0<br />

1 + s0 + (2∆/Γ) 2 · � k (2.33)<br />

To cool an atom, we need to reduce the thermal motion. This can be<br />

done by the force <strong>of</strong> the spontaneous scattering. Each spatial direction can be<br />

cooled by one pair <strong>of</strong> beams if we use light with a negative detuning ∆ < 0 and<br />

irradiate the atom with a pair <strong>of</strong> counterpropagating beams <strong>of</strong> same intensity<br />

and detuning.<br />

Consider an atom moving with a velocity �v due to the thermal motion. The<br />

Doppler effect will shift the frequency <strong>of</strong> the light <strong>of</strong> the counterpropagating<br />

beam to a different, smaller detuning than the rest frame detuning ∆, so<br />

the effective detuning the atom experiences is ∆ − �v · � k. Under the same<br />

conditions the atom sees a higher detuning from the light propagating in the<br />

same direction. The probability to absorb a photon is different for each beam:<br />

it is higher for the counter propagating beam. After each absorption, a photon<br />

will be emitted by the atom. This is done isotropically and after many cycles<br />

the momentum change due to the emissions averages out. The atom will have<br />

experienced a net momentum transfer that is directed against its propagation<br />

and slowing it in this direction. As this is done in every spatial direction,<br />

the undirected thermal motion is reduced and thus the atom is cooled. This<br />

procedure has been termed “optical molasses”, as in the linear approximation<br />

<strong>of</strong> equation (2.33) the force is linear in the velocity <strong>of</strong> the atom like viscous<br />

damping in mechanics.<br />

Involved with the re-emission <strong>of</strong> the photons is a heating process; so the<br />

lowest possible temperatures that can be reached this way is the equilibrium<br />

between the cooling and the heating process. It can be shown that the lowest<br />

25


2.1. Atoms and Electromagnetic Fields<br />

possible temperature is reached for a detuning <strong>of</strong> ∆ = − Γ<br />

2<br />

[Neu78]. This temperature is called the Doppler-limit<br />

and for 87 Rb is TDoppler = 146 µK.<br />

Sub-doppler cooling<br />

TDoppler = ¯h<br />

kB<br />

Γ<br />

2<br />

or half a linewidth<br />

(2.34)<br />

Soon after the first experimental observation <strong>of</strong> cooling atoms by radiation,<br />

it was observed that the temperature <strong>of</strong> the atoms was actually below the<br />

predicted doppler limit for the temperature [Let88]. This behaviour can be<br />

explained by the fact that the atoms are not pure two level systems [Dal89],<br />

but have magnetic sublevels. The cooling then stems from spatially dependent<br />

stark shifts in an optical standing wave and the optical pumping <strong>of</strong> the atoms<br />

in the substate with less energy.<br />

Assume a light field that is created by two counterpropagating linearly<br />

polarised waves, with perpendicular planes <strong>of</strong> polarisation (lin⊥lin), and an<br />

atom where the ground state has two magentic sublevels mF = ±1/2. In<br />

the basis <strong>of</strong> circular polarisation, the light field creates two standing waves<br />

<strong>of</strong> different polarisation σ +,− that have nodes which are separated by λ/4<br />

from each other. The magnetic substates now are shifted depending on the<br />

substate and the intensity and polarisation <strong>of</strong> the light. Also, σ + -polarised<br />

light drives transitions into the mf = +1/2 state and σ − -polarised light into<br />

the mf = −1/2 state. Now, at positions with high intensity <strong>of</strong> one polarisation,<br />

the level that the atom is pumped into is a strongly negative shifted level. If<br />

the atom moves further, the shifting intensity <strong>of</strong> the standing wave decreases<br />

and the potential energy <strong>of</strong> the state increases. This results in a loss <strong>of</strong> kinetic<br />

energy. A decrease in the intensity <strong>of</strong> one polarisation means an increase in<br />

the intensity <strong>of</strong> the other, so the atom can undergo a new transition and again<br />

26


Chapter 2: Theoretical Background<br />

be pumped into a low-lying energetic state. Figure 2.3 depicts this concept.<br />

It can be seen as the atom is moving uphill in the potential, and close to the<br />

peak, is falling back into a valley where it starts a new climb. Because <strong>of</strong> the<br />

similarity to the fate <strong>of</strong> the greek tragic hero Sisyphos, the cooling mechanism<br />

was dubbed “Sisyphus cooling”.<br />

E<br />

λ/2<br />

m = - 1/2<br />

F<br />

m = + 1/2<br />

F<br />

Figure 2.3: Sub-Doppler cooling: the atom continuously “moves uphill” as<br />

transitions from the energetically higher to lower mF -state are driven.<br />

The ratio <strong>of</strong> this cooling and the Doppler cooling can be shown to be about<br />

2|∆|/Γ [Met99]. For larger detunings ∆ lower temperatures can be reached,<br />

but this cooling mechanism has a smaller capture range and only works with<br />

sufficiently cold, Doppler cooled atoms. The lowest limit here is given by the<br />

momentum <strong>of</strong> the last emitted photon ¯hk<br />

Trec = Erec/kB = (¯hk)2<br />

2mkB<br />

with m being the mass <strong>of</strong> the atom. For 87 Rb this recoil limit is 349 nK.<br />

27<br />

(2.35)


2.2. Atoms and Magnetic Fields<br />

2.2 Atoms and Magnetic Fields<br />

That the interaction between atoms and magnetic fields can expose the atoms<br />

to a force was first shown by Otto Stern 1 and Walther Gerlach in 1922. An<br />

atom with magnetic moment �µ in an inhomogenous field � B will experience a<br />

force<br />

�F = � ∇(�µ · � B) (2.36)<br />

This force can be expressed by a potential U = −�µ · � B and � F = − � ∇U. The<br />

potential finds its extreme values for parallel and antiparallel alignment <strong>of</strong> the<br />

magnetic moment towards the magnetic field and at the position <strong>of</strong> a local<br />

maximum or minimum <strong>of</strong> the field. This can be used for trapping <strong>of</strong> neutral<br />

atoms and was first demonstrated in 1985 [Mig85]. Using magnetic fields to<br />

confine atoms was first proposed by W. Paul [Fri51], although here a hexapole<br />

field was used as lense and not as a trap in 3D.<br />

2.2.1 Magnetic trapping<br />

Unlike the case with optical traps, it is not possible to create local maxima<br />

<strong>of</strong> the field strength with static magnetic fields [Win84]. That means we have<br />

to trap the atoms in a local minimum. For small fields we find anomalous<br />

Zeeman splitting, in which the magnetic moment <strong>of</strong> an atom can be expressed<br />

by µ = −µBgF mF , with µB the Bohr magneton, gF the Landé factor and mF<br />

the magnetic quantum number or substate <strong>of</strong> an atomic state with hyperfine<br />

state F . For a positive Landé factor (gF > 0), this reduces the choice <strong>of</strong><br />

trappable atomic states to those for which the magnetic moment is counter-<br />

aligned with the magnetic field vector. In quantum mechanics, this is achieved<br />

by magnetic substates with positive quantum numbers mF , which then are<br />

1 O. Stern also first showed that whole atoms possess wave-like properties observing<br />

diffraction in a beam <strong>of</strong> He [Kna29]<br />

28


Chapter 2: Theoretical Background<br />

trapped in the local minimum <strong>of</strong> the magnetic field. These are called low or<br />

weak field seeking states.<br />

The simplest way to create a field with a minimum is a quadrupole field,<br />

which can be created by two coils in an anti-Helmholtz configuration. This<br />

field has a zero at its centre and from there a linear increase in the absolute<br />

field strength. Different mF -states thus differ in their energy by an amount<br />

proportional to the magnetic field:<br />

∆E = gF µBB∆mF<br />

(2.37)<br />

The linearity <strong>of</strong> the field <strong>of</strong> this trap is a drawback though. Atoms moving<br />

towards the centre have their magnetic moment counter-aligned with the di-<br />

rection <strong>of</strong> the field vectors. When they pass through the field’s zero crossing,<br />

the atom sees a reversed direction <strong>of</strong> the field and accordingly experiences a<br />

force that expels it from the centre. This has been called a spin flip or a Ma-<br />

jorana spin flip, although it is not the spin that changes, but the direction <strong>of</strong><br />

the magnetic field. In fact, if the spin and thus the magnetic moment is able<br />

to change with the field then the atom will not be lost.<br />

Another simple design creates a quadrupole trap in two dimensions. A<br />

current I is running through an infinitely long wire, and a homogeneous bias<br />

field B0 is applied perpendicular to the wire. There is one point r0 ∝ I/B0<br />

where the field <strong>of</strong> the wire Bw ∝ I/r is just canceled by the bias field. This<br />

field can be well approximated by a quadrupole field. If we now add a second<br />

bias field along the wire, the overall field never crosses zero, although it has a<br />

field minimum at the same position as before. Here the atomic spin can now<br />

adiabatically follow the changing field direction and thus Majorana spin flips<br />

are suppressed. Such a configuration is generally called a I<strong>of</strong>fe-Pritchard (IP)<br />

configuration [Pri83].<br />

29


2.2. Atoms and Magnetic Fields<br />

Miniaturised traps: Atom chips<br />

A single wire and bias field producing a magnetic quadrupole field was used<br />

in 1933 to study spin flips in an atomic beam [Fri33]. Experiments with free<br />

standing wires and magnetic bias fields were successful in deflecting, guid-<br />

ing and trapping atoms [Row96, For98b, For00]. It was proposed to use<br />

micr<strong>of</strong>abricated current-carrying wires [Wei95], much like the conductors in<br />

micro-electronics, to create the trapping magnetic potential. This device<br />

was called an atom chip. Soon after the proposal the trapping <strong>of</strong> atoms<br />

[Rei99, Fol00, For00] and the use <strong>of</strong> such a chip as a magnetic mirror for atoms<br />

[Lau99b] were experimentally demonstrated. The use <strong>of</strong> atom chips proved to<br />

be a breakthrough technology in evaporative cooling and Bose-Einstein con-<br />

densation [Ott01, H¨01a]. Traps created by these chips in general have a much<br />

higher gradient than “macroscopic” traps. This increases the collision rate<br />

and reduces the time needed equilibrating the atomic sample and in this way<br />

reduces the time needed for the whole evaporation process. A shorter time<br />

relaxes the conditions on the lifetime <strong>of</strong> the trap and the surrounding vacuum.<br />

Recently chips with slightly larger structures showed the same benefits in fast<br />

and ‘easy’ condensation [Sch03, Val04].<br />

The field <strong>of</strong> an infinitely long wire with current I at a distance �r with an<br />

external bias field � B0 is<br />

�B = µ0<br />

�I × �r<br />

2π r2 + � B0 (2.38)<br />

We choose the component <strong>of</strong> the external field that is perpendicular to the<br />

wire and reduce the problem to this dimension, where we put the origin <strong>of</strong> the<br />

axis into the wire. Then we have<br />

B = µ0 I<br />

− B⊥0<br />

(2.39)<br />

2π r<br />

This field has a zero at the position r0 = µ0<br />

2π<br />

30<br />

I<br />

B⊥0<br />

, and for the gradient <strong>of</strong> this


magnetic field at the trap position we find<br />

�<br />

∂B �<br />

�<br />

∂r � = −<br />

r=r0<br />

µ0 B<br />

2π<br />

2 ⊥0<br />

I<br />

In the Gaussian cgs system this is simply<br />

�<br />

∂B �<br />

� =<br />

∂r<br />

1 B<br />

2<br />

2 ⊥0<br />

I [cgs]<br />

� r=r0, [cgs]<br />

Chapter 2: Theoretical Background<br />

(2.40)<br />

(2.41)<br />

Of course, an infinite wire is not realisable and does not allow confinement<br />

in the direction <strong>of</strong> the wire. Bent wires can create the <strong>of</strong>fset field � B0 and<br />

create the confining field. Here two major designs have been considered. In<br />

the first design the wire is bent into a U-shaped form, where the connecting<br />

parts are bent away from the ‘infinite’ part in the same direction. The overall<br />

field <strong>of</strong> this wire alone is a quadrupole field in all three dimensions, the field<br />

components <strong>of</strong> the side wires cancelling themselves at the trap centre. This<br />

trap’s centre is not directly above the wire. It is shifted slightly away from the<br />

wires in the dimension <strong>of</strong> the bent wires, see Fig. 2.4.<br />

A more symmetric arrangement is created by bending the wire in a Z-<br />

shape. This creates a harmonic trap, as the fields from the side wires add<br />

constructively at the trap’s centre. The field <strong>of</strong> this trap for a current I, from<br />

the centre <strong>of</strong> the ‘infinite’ central wire bar with length wx, perpendicular to<br />

the chip’s plane can be calculated to:<br />

⎛<br />

z ·<br />

⎜<br />

�B(z) = I · ⎜<br />

⎝<br />

2wy<br />

(w 2 x/4+z 2 ) √ w 2 x/4+z 2 +w 2 y<br />

− 1<br />

z · wx √<br />

w2 x /4+z2 0<br />

⎞<br />

⎟<br />

⎠ + � B0<br />

(2.42)<br />

In this thesis the Gaussian cgs system is used when covering magnetic fields,<br />

as for applications in atom optics it yields convenient numbers. The field is<br />

expressed in Cartesian coordinates, where the x-direction is chosen along the<br />

wire and the z-axis is perpendicular to the chip’s plane. The terms wx and wy<br />

31


2.2. Atoms and Magnetic Fields<br />

give the extensions <strong>of</strong> the wire in the labelled direction, where it is assumed<br />

that the bent parts <strong>of</strong> the wire have the same length wy. The finite width <strong>of</strong><br />

the wire is still neglected here; the correction terms can be found in [Rei02].<br />

A wire configuration in an “H”-shape now allows one to produce a U-shape<br />

or a Z-shape, depending on the arms the current passes through [Fol00]. The<br />

different different trapping potentials for a single wire, a U-shaped and a Z-<br />

shaped wire with their respective bias fields are depicted in Fig. 2.4, taken<br />

from an article with an extensive overview on atom chips [Fol02].<br />

Figure 2.4: Magnetic fields and currents (top) and corresponding potentials<br />

(bottom) <strong>of</strong> different possible wire configurations for trapping atoms: (a) a<br />

single wire with a quadrupole field, (b) a U-shaped wire with homogenous bias<br />

field, (c) a Z-shaped wre with homogenous bais field. Taken from [Fol02]<br />

Using typical parameters (I = 30 A, B0 = 60 G, wy = 4 · wx = 10 mm),<br />

we expect our trap to be about 0.8 mm away from the wire with trapping<br />

frequencies <strong>of</strong> ν ≈ 300 Hz (see Figure 2.5).<br />

It is possible to use more than one wire to create trapping potentials. If<br />

more than one parallel wire is used, then it is possible to create configurations<br />

where the magnetic potential has more than one minimum (a double well<br />

potential) [Hin01, Est05], waveguides for atoms [Thy99], or traps where the<br />

32


absolute magnetic field / G<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

0.5 0.6 0.7 0.8 0.9 1 1.1 1.2<br />

distance from wire / mm<br />

Chapter 2: Theoretical Background<br />

Figure 2.5: Magnetic field <strong>of</strong> a Z-shaped wire plus bias field with parameters<br />

close to the experimental parameters. The field is given as a function from the<br />

distance from the wire perpendicular to the plane defined by the Z-structure<br />

and calculated by equation (2.42).<br />

wires create their own bias field [Ott01]. Using bent wires can also create IP-<br />

type traps without external fields [Wei95] or even complicated structures like<br />

conveyor belts for atoms [H¨01b], waveguides [Lea02] or beamsplitters [Cas00,<br />

M¨01]. A new generation <strong>of</strong> chips has structures to capture and evaporate the<br />

atoms, and structures which create the field for the experiment [Zim04]. When<br />

these structures create a periodic potential with or without perturbations,<br />

they allow us to investigate problems <strong>of</strong> solid-state physics with a mesoscopic<br />

number <strong>of</strong> atoms.<br />

In general one can state that atom chips simplify the creation <strong>of</strong> a BEC. The<br />

high gradients allow very efficient evaporation, relaxing the requirements on<br />

the vacuum by reducing the evaporation and rethermalisation times. Current-<br />

carrying wires allow complicated structures and fields, such as atom guides,<br />

beam splitters and switches, which are all promising candidates to become<br />

parts <strong>of</strong> an atom interferometer on a chip [Shi05, Wan05b]. Several overview<br />

articles on the topic have been published [Rei02, Fol02, For03].<br />

33


2.2. Atoms and Magnetic Fields<br />

2.2.2 Permanent magnets<br />

Instead <strong>of</strong> using current-carrying wires on a chip to induce magnetic fields,<br />

it is possible to use permanent magnetic structures instead. Among the first<br />

applications <strong>of</strong> permanent magnetic structures in atom optics was a periodic<br />

array <strong>of</strong> magnets [Roa95, Sid96, Lau99a], which has a field strength that decays<br />

exponentially with distance from the surface. This decaying field can then be<br />

used as a mirror, similar to evanescent light [Bal87]. The materials used in<br />

mirror experiments were magneto-optical ferrimagnetic films like TbGdFeCo,<br />

alloys like CoCr [Sid02b] and even computer hard drives [Lev03b], audio tape<br />

[Roa95] or video tape [Ros00].<br />

When using periodic arrays like the above, the magnetic fields <strong>of</strong> the indi-<br />

vidual elements are summed. We now consider a single magnet <strong>of</strong> thickness<br />

h uniformly magnetised in perpendicular direction, which means it has its<br />

magnetisation axis along its height and perpendicular to its plane. Then, the<br />

magnetic field � Bfilm produced by this magnet will be the same as the field <strong>of</strong><br />

a current Ieff that propagates along the edges <strong>of</strong> the magnet. The strength <strong>of</strong><br />

this effective current depends on the magnetisation MR and the thickness h<br />

<strong>of</strong> the element, Ieff = hMR. The magnitude <strong>of</strong> the field and the field gradient<br />

near the edge <strong>of</strong> the magnetic element is then like the field <strong>of</strong> a single wire (see<br />

equation (2.39) )<br />

Bfilm = µ0 hMR<br />

2π z<br />

∂<br />

∂z Bfilm = − µ0<br />

2π<br />

hMR<br />

z 2<br />

(2.43)<br />

where z is the distance from the magnetic element and much larger than the<br />

thickness h. As the edge <strong>of</strong> a magnet can be seen as a current through a wire,<br />

it was proposed to implement atom traps by replacing the current-carrying<br />

wires by permanent magnetic structures [Dav99a, Dav99b, Dav99c, Sid02b].<br />

Here it was proposed to use parallel slabs <strong>of</strong> magnets, so that a single slab<br />

34


Chapter 2: Theoretical Background<br />

would be equivalent to a double-wire structure, and two slabs would create the<br />

field <strong>of</strong> four wires. These waveguides and even 2D magneto-optical trapping<br />

was demonstrated [Ven02]. Even though these magnetic fields are permanent,<br />

it would still be possible to load them from the outside by using bias fields<br />

to compensate the barriers that otherwise confine the atoms once they are<br />

trapped [Sid02b]. Careful design <strong>of</strong> perpendicular slabs can lead to a 2D<br />

array <strong>of</strong> traps for quantum computation [Gha06], similar to dipole traps using<br />

microstructured optics [Bir01, Dum03b].<br />

In our experiment, we use the edge <strong>of</strong> a ferrimagnetic magneto-optical film<br />

(TbGdFeCo) sandwich structure [Wan05a] as the equivalent <strong>of</strong> a single straight<br />

wire. Other experiments use a layered structure <strong>of</strong> a different magneto-optical<br />

film (Pt-Co), with a corner cut out <strong>of</strong> their slab (this is equivalent to a Z-<br />

shaped wire) [Eri04], or a comparatively thick layer <strong>of</strong> Fe-Pt film [Xin04].<br />

In the latter experiment the permanent magnetic material has the form <strong>of</strong> a<br />

capital F, for which the resulting field creates a self-biased I<strong>of</strong>fe-Pritchard type<br />

trap [Bar05]. Recently, the first Bose-Einstein condensate was formed in a trap<br />

from a permanent magnetic chip [Sin05].<br />

2.3 Evaporative Cooling<br />

Optical cooling methods are limited by the lowest temperatures that can be<br />

achieved. High density samples can show strong absorption, leading to an<br />

anisotropic cooling, and have a higher chance <strong>of</strong> reabsorbing photons emitted<br />

by cooled atoms, which results in a higher heating rate. For Bose-Einstein<br />

condensation, it is important to reach a high phase space density: that is<br />

very low temperatures (a narrow distribution in momentum space) and high<br />

densities (a narrow distribution in space). Both <strong>of</strong> these factors are directly<br />

limited when cooling with radiation.<br />

35


2.3. Evaporative Cooling<br />

The light field <strong>of</strong> a laser is a highly ordered system and can be considered<br />

as a coolant into which the sample is immersed; the emission <strong>of</strong> photons into<br />

different modes adds entropy and transfers energy from the atoms into the<br />

light field. Instead <strong>of</strong> having the sample immersed in a cooling medium, a way<br />

to further cool the sample is to let parts <strong>of</strong> the sample carry away an amount<br />

<strong>of</strong> entropy and energy that exceeds the mean value <strong>of</strong> the sample. This <strong>of</strong><br />

course reduces the number <strong>of</strong> particles in the cloud. For this the trap depth is<br />

lowered, so that particles with high energies can escape the trap. Because it<br />

can be compared to hot water, where the highest energy molecules evaporate<br />

and leave a cooler sample, this method has been called evaporative cooling<br />

[Hes86, Mas88].<br />

2.3.1 A simple model<br />

A simple model <strong>of</strong> evaporative cooling is presented now, following the treat-<br />

ments <strong>of</strong> [Met99] and [Arn99]. This model allows us to introduce some quan-<br />

titative measures to characterise the cooling process. It highly simplifies the<br />

actual processes: ergodic behaviour <strong>of</strong> the atoms is assumed, as is classical<br />

thermodynamics without quantum influences. Only s-wave scattering is taken<br />

into account and only with a constant cross section for elastic scattering. Fur-<br />

ther, full evaporation is assumed and the rethermalisation is much faster than<br />

the cooling rate. For the experiment here the model applies sufficiently well,<br />

but one must be aware <strong>of</strong> its limitations.<br />

We assume our trapping potential to be <strong>of</strong> the form<br />

U(�r) = �<br />

i=1,2,3<br />

ci<br />

�<br />

�<br />

�<br />

�<br />

xi<br />

ai<br />

�<br />

�<br />

�<br />

�<br />

si<br />

(2.44)<br />

where ai is the characteristic length <strong>of</strong> the trap in the xi direction, and si gives<br />

the power scaling for each axis. It can be shown that the volume V <strong>of</strong> the<br />

36


Chapter 2: Theoretical Background<br />

atomic cloud is then proportional to a power <strong>of</strong> the temperature T<br />

V ∝ T ξ<br />

ξ = � 1<br />

si<br />

(2.45)<br />

(2.46)<br />

and the dependency on the trap geometry is reduced to a single parameter ξ<br />

[Met99, Arn99]. For a linear trap in 3 dimensions, si = 1, we get ξ = 3; for a<br />

harmonic trap in 3D with si = 2 it is ξ = 3/2.<br />

We now assume N atoms <strong>of</strong> temperature T in this trap which we assume<br />

to be infinitely high initially. We reduce the trap depth to a value <strong>of</strong> ηkBT ,<br />

so η is a truncation parameter. This will let some atoms escape the trap, we<br />

will label the ratio <strong>of</strong> the remaining atoms N ′ and the initial number N with<br />

ν ≡ N ′ /N. Now we can define a measure for the temperature decrease from<br />

T to T ′<br />

γ ≡ log(T ′ /T )<br />

log(N ′ /N) = log(T ′ /T )<br />

log ν<br />

(2.47)<br />

From this we can deduce power laws for the thermodynamic quantities.<br />

The primed values are the ones after the reduction <strong>of</strong> the trap height. They<br />

are:<br />

T ′ = T · ν γ<br />

N ′ = N · ν<br />

V ′ = V · ν γξ<br />

(2.48)<br />

Additional important quantities when evaporating atoms from a trap are the<br />

atom density n = N/V <strong>of</strong> the remaining cloud and the phase space density<br />

(PSD) ρ which scales with the atom density and the cube <strong>of</strong> the thermal de<br />

Broglie wavelength:<br />

ρ = nλ 3 dB<br />

37<br />

(2.49)


2.3. Evaporative Cooling<br />

The change <strong>of</strong> these two quantities can also be expressed in terms <strong>of</strong> ν, γ and<br />

ξ:<br />

n ′ = n · ν 1−γξ<br />

ρ ′ = ρ · ν 1−γ(ξ+3/2)<br />

(2.50)<br />

One can now show [Met99, Arn99] from the density <strong>of</strong> states in an external<br />

potential that the fraction <strong>of</strong> atoms remaining in the trap after decreasing the<br />

trap depth to ηkBT becomes<br />

ν =<br />

N ′<br />

N<br />

= γ(ξ + 3/2, η)<br />

Γ(ξ + 3/2)<br />

(2.51)<br />

where γ(ξ + 3/2, η) is the lower incomplete Gamma-function, with Γ(ξ + 3/2)<br />

being the corresponding complete Gamma-function.<br />

The dynamics <strong>of</strong> the evaporation process are covered in [Pet02]. The key<br />

results are that the decay times for evaporation τev and other losses that do<br />

not change the temperature <strong>of</strong> the atoms, τloss, need to be comparable for the<br />

optimal choice <strong>of</strong> the threshold energy at which atoms can leave the trap. It<br />

is also shown that for evaporation trap potentials with larger values <strong>of</strong> the<br />

parameter ξ are preferable, which means linear traps are better than harmonic<br />

ones. Most important in experiments is that the time needed to rethermalise<br />

the atoms, and thus the elastic scattering time τel, decreases as the evaporation<br />

proceeds. The scattering rate scales as the atomic density times the thermal<br />

velocity ∝ T 1/2 . To achieve runaway evaporation one needs<br />

d<br />

d log N log τel = γ ·<br />

τloss<br />

τev + τloss<br />

giving a stringent requirement for successful evaporation.<br />

· (ξ − 1<br />

) − 1 > 0 (2.52)<br />

2<br />

Of course, this model can not be applied directly to optical traps, as there<br />

it is not only the trap height, but also the trap’s curvature and gradient that<br />

change during the evaporation ramping. Thus, the characteristic length in<br />

38


Chapter 2: Theoretical Background<br />

equation (2.44) is not constant. With the failure <strong>of</strong> this approximation the<br />

model as presented here becomes invalid. The problem <strong>of</strong> evaporating in an<br />

optical trap was treated in [O’H01], where a constant η and one case <strong>of</strong> ramping<br />

down the potential was assumed. Still, one can use the parameters as defined<br />

here for the evaporative cooling in optical traps to get a rough idea <strong>of</strong> the<br />

processes.<br />

2.4 Bose-Einstein Condensation<br />

The main reason why such an effort is made to reach ultra-cold temperatures<br />

is the condensation <strong>of</strong> atoms into a Bose-Einstein condensate (BEC) [Ein25,<br />

And95, Dav95, Bra95]. In this state, all atoms occupy the same quantum state<br />

<strong>of</strong> the trap, the ground state. One then has a macroscopic and coherent matter<br />

wave, a tool which promises to be <strong>of</strong> great use in atom interferometry.<br />

Bosonic particles do not follow the Maxwell-Boltzmann distribution. In-<br />

stead quantum effects lead to the Bose-Einstein distribution, so that for a<br />

sample at temperature T N(E) particles will be found at an energy E:<br />

N(E) =<br />

1<br />

e (E−µ)/kBT − 1<br />

(2.53)<br />

The chemical potential µ vanishes for photons, leading to Planck’s equation<br />

for black body radiation. This was the historical problem which S. Bose solved<br />

and A. Einstein extended for quantized energies. A new state <strong>of</strong> matter was<br />

predicted by this distribution. For very low temperatures, a phase transition<br />

will occur. Then the lowest possible state, the ground state, would gain a<br />

sudden and large increase in its population. The threshold condition for this<br />

phase transition is related to the phase space density, using the same notation<br />

as above:<br />

ρ = nλ 3 dB ≥ 2.612 (2.54)<br />

39


2.4. Bose-Einstein Condensation<br />

This is the critical value for a three dimensional box. In a 3D harmonic po-<br />

tential the threshold is less than half that value [Pet02, Arn99], although older<br />

sources claim it is independent <strong>of</strong> the actual trap geometry [Bag87]. Once the<br />

critical PSD is reached, the other atoms will condense into that region. A more<br />

pictorial description <strong>of</strong> this condition is that the wavefunctions <strong>of</strong> the atoms<br />

must overlap each other, so that they would then constructively add up to one<br />

large coherent wave function.<br />

As these atoms are interacting with each other, the description <strong>of</strong> the whole<br />

system is not trivial. Fortunately, in many cases one can use standard approxi-<br />

mation methods. Using the Hartree-Fock approximation allows one to describe<br />

the interaction <strong>of</strong> a single particle with many others by estimating a mean field,<br />

which averages the influences <strong>of</strong> all other N particles into one additional non-<br />

linear potential term. When this additional non-linear term is added to the<br />

Schrödinger equation, one gets the Gross-Pitaevskii equation [Gro61, Pit61]<br />

for the wavefunction Φ <strong>of</strong> the condensate:<br />

�<br />

− ¯h2<br />

�<br />

2<br />

∆ + V + NVNL|Φ| · Φ = i¯h<br />

2m ∂<br />

Φ (2.55)<br />

∂t<br />

The term N · VNL · |Φ| 2 describing the interaction between the atoms is depen-<br />

dent on the scattering <strong>of</strong> the atoms: it is assumed that only s-wave scatter-<br />

ing takes place between the atoms with scattering length a0, so that VNL =<br />

4π¯h 2 a0/m. For a BEC this assumption in general is valid. The scattering<br />

length <strong>of</strong> Rubidium can be taken as constant over the range <strong>of</strong> parameters<br />

that are used in the experiments here. It is positive, leading to a repulsive in-<br />

teraction between the atoms. Magnetic fields and optical fields can be used to<br />

tune the scattering length using resonances between atomic states and molec-<br />

ular states, so-called Feshbach resonances. By tuning the respective field, so<br />

that the scattering length vanishes, it is possible to create a coherent mat-<br />

ter wave <strong>of</strong> non-interacting particles. This simplifies the description, as the<br />

40


Chapter 2: Theoretical Background<br />

non-linear term vanishes and the whole sample can be described by the single<br />

particle Schrödinger equation. Especially for atom interferometry in double<br />

well potentials this can be advantageous (see section 2.6.1 and chapter 3).<br />

Unfortunately for the isotope <strong>of</strong> 87 Rb the Feshbach resonances are not easily<br />

accessible [Mar02, The04].<br />

For an extensive description <strong>of</strong> this model with a detailed assessment <strong>of</strong><br />

when the assumptions and approximations made are valid, see for example<br />

[Pet02, Arn99]. An overview suited for experimental physicists covering the<br />

route towards Bose-Einstein condensation, detection <strong>of</strong> condensates, early re-<br />

sults and an extensive list <strong>of</strong> literature can be found in [Ket99].<br />

2.5 The Magneto-Optical Trap<br />

Both the optical and the magnetic traps described in the earlier sections have<br />

one common drawback: their depths allow us to trap only atoms that are<br />

already sufficiently cold. If one tried to fill these traps from an atomic reservoir<br />

at room temperature, the number <strong>of</strong> captured atoms would be negligible. The<br />

method <strong>of</strong> choice to overcome this is to trap and pre-cool the atoms in a<br />

magneto-optical trap (MOT) [Raa87], a combination <strong>of</strong> a quadrupole magnetic<br />

field and a three dimensional optical molasses with circular polarisation. The<br />

absolute magnetic field increases linearly from its zero point in each direction,<br />

| � B| ∝ |�x|. If we consider a single transition |J = 0〉 → |J ′ = 1〉, we will<br />

find a Zeeman splitting <strong>of</strong> the magnetic sublevels <strong>of</strong> ∆E = gJµBBmJ ′. This<br />

Zeeman splitting due to the magnetic fields provides a natural quantisation<br />

axis radially along the � B field, with mJ ′ being a “good” quantum number for<br />

our purposes. This quantisation only depends on the magnetic field, and due to<br />

its radial symmetry there are no angular dependencies. Electronic transitions<br />

with mJ ′ = 0 are called π transitions, transitions with mJ ′ = ±1 are called<br />

41


2.5. The Magneto-Optical Trap<br />

σ ± transitions. We will consider circularly polarised light. This polarisation<br />

is independent <strong>of</strong> the magnetic field. It is always defined in relation to the<br />

direction <strong>of</strong> propagation. One can define this by the projection <strong>of</strong> the photon’s<br />

spin onto its direction <strong>of</strong> propagation. This is called the helicity and possible<br />

values are ±¯h. Another common way to define the polarisation is to describe<br />

the evolution <strong>of</strong> the electrical field vector along its propagation: with the light<br />

coming towards the observer, it is called right-handed circular (RHC) when<br />

the vector rotates clockwise and left-handed circular (LHC) when the vector<br />

rotates anti-clockwise. Positive helicity is the same as LHC [Bra86].<br />

We now illuminate the atoms with counterpropagating light that is po-<br />

larised to drive transitions from the ground state to the excited state with<br />

mJ ′ = −1, called σ− light. Of course, the light that is σ − on one side <strong>of</strong><br />

the trap centre will change to σ + once it has passed the centre point: the<br />

handedness <strong>of</strong> the light does not change, but the direction <strong>of</strong> the magnetic<br />

field reverses. So, at each point we find not only the cooling and trapping σ −<br />

light, but also σ + light. Fortunately, the Zeeman effect shifts the energy <strong>of</strong> the<br />

mJ ′ = +1 state upwards, so that the detuning is further increased and absorp-<br />

tion <strong>of</strong> this light is not probable. This scheme is depicted in Figure 2.6. We<br />

can thus say that the incoming light has to have the same circular polarisation,<br />

which translates to one beam pair (passing through the magnetic field coils)<br />

with one handedness, while the other two pairs have the opposite handedness.<br />

Up to the point where the sum <strong>of</strong> the Zeeman shift and the thermal Doppler<br />

shift <strong>of</strong> the atoms are less than the red detuning <strong>of</strong> the light this results in a<br />

force that drives the atoms back to the centre. When the detuning is large<br />

compared to these shifts, the force can be written as a sum <strong>of</strong> a spatially and<br />

a velocity dependent part:<br />

�F = −β�v − κ�r (2.56)<br />

This force leads to a damped harmonic oscillation <strong>of</strong> the atoms with a spring<br />

42


negative<br />

helicity,<br />

RHC<br />

LHC<br />

I coils<br />

positive<br />

helicity,<br />

LHC<br />

left handed circular<br />

LHC<br />

right<br />

handed<br />

circular<br />

Chapter 2: Theoretical Background<br />

E<br />

hνLaser<br />

m =+1<br />

J<br />

m =0<br />

J<br />

m =0<br />

J<br />

J=1<br />

m =-1<br />

J<br />

Figure 2.6: Schematic view on a magneto-optical trap with helicities <strong>of</strong> the light<br />

for a chosen direction <strong>of</strong> the current in the quadrupole coils (left). Influence<br />

<strong>of</strong> the magnetic field on the energy <strong>of</strong> the atomic sublevels, increasing the<br />

probability to absorb a photon from the σ − polarised beam (right).<br />

constant κ. For the standard MOT situation the motion is overdamped. For a<br />

MOT with Rb, one finds temperatures that are below the Doppler limit. The<br />

cooling mechanism is not the Sisyphus cooling described above, but due to<br />

non adiabatic passage through polarisation gradients, also called σ + σ − -cooling<br />

[Dal89].<br />

2.5.1 The mirror MOT<br />

In the standard design <strong>of</strong> a MOT the optical molasses is created by three<br />

pairs <strong>of</strong> counterpropagating beams. The whole design is symmetric around the<br />

centre point, and thus to a plane through the centre point. This symmetry can<br />

be used to further simplify the set-up. If a mirror is placed such that it stands<br />

at an angle <strong>of</strong> 45 degrees to the coils that create the quadrupole field, then only<br />

43<br />

σ +<br />

σ −<br />

J=0<br />

r


2.5. The Magneto-Optical Trap<br />

two pairs <strong>of</strong> light beams are needed [Rei99]. The left side <strong>of</strong> Figure 2.7 shows<br />

a sketch <strong>of</strong> this. Here one pair <strong>of</strong> beams passes through without reflection and<br />

LHC<br />

I coils<br />

positive<br />

helicity,<br />

left handed<br />

circular<br />

Mirror<br />

LHC<br />

negative<br />

helicity,<br />

right handed<br />

circular<br />

LHC<br />

σ −<br />

σ +<br />

LHC<br />

Mirror<br />

Figure 2.7: Schematic view on a mirror MOT with helicities <strong>of</strong> the light (left).<br />

Sketch <strong>of</strong> the change <strong>of</strong> polarisation and handedness <strong>of</strong> the light close to the<br />

surface (right)<br />

this does not differ from the description above: as stated, the light changes its<br />

polarisation from σ − to σ + upon passing through the field’s zero point while<br />

it keeps its handedness. The other pair is made up <strong>of</strong> two beams that are<br />

reflected into each other. One <strong>of</strong> these beams initially points with the field<br />

lines <strong>of</strong> the magnetic field; the other beam has a � k vector that points against<br />

the field. They are chosen with different handedness so that with the different<br />

relative alignment to the magnetic field their circular polarisation is thus the<br />

same. Upon reflection, only the handedness <strong>of</strong> the light changes. The relative<br />

alignment <strong>of</strong> � k vector and magnetic field changes once before the reflection, and<br />

once after the reflection: the light leaving the trap has the opposite circular<br />

polarisation. Figure 2.7, right, explains these changes graphically.<br />

σ −<br />

RHC<br />

σ +<br />

Instead <strong>of</strong> using external coils to create the quadrupole field, it is possible<br />

to use current-carrying structures on the mirror holder (see section 2.2.1).<br />

A U-shaped wire then creates the magnetic field [Rei99, Fol00]. This MOT<br />

44<br />

RHC


Chapter 2: Theoretical Background<br />

typically has a higher field gradient, and is closer to the surface <strong>of</strong> the chip.<br />

The terms U-wire MOT and compressed MOT will be used synonymously.<br />

The possibility <strong>of</strong> creating first a mirror-MOT with external coils for a large<br />

trapping volume and high atom numbers, which is then changed to a U-wire<br />

MOT, has some important benefits. The U-wire MOT allows easy alignment<br />

<strong>of</strong> the atomic cloud with the chip regarding the relative position to the wire<br />

and the matching <strong>of</strong> the cloud size to the potential that the Z-wire will create<br />

for magnetic trapping. Also, the switching times <strong>of</strong> the quadrupole coils is<br />

replaced by the switching time <strong>of</strong> the bent wire, which is significantly shorter.<br />

Furthermore, the fabrication <strong>of</strong> these chips uses standard technologies, making<br />

these chips rather easily accessible [Lev03a]. These reasons are partly why<br />

atom chips and mirror MOTs have become so successful and part <strong>of</strong> many<br />

atom optics experiments [Sch03, Wil04, Val04].<br />

2.6 Atom Interferometry<br />

Interferometry with atoms is an important branch <strong>of</strong> today’s physics, and<br />

one <strong>of</strong> the two most promising future main applications 2 <strong>of</strong> cold atoms and<br />

atom optics, just like interferometry is an important part <strong>of</strong> optics. The pos-<br />

sibility to confine atoms in free space and to cool them to low temperatures<br />

opens the door towards interferometry with heavy particles. Here we will con-<br />

centrate on interferometry with trapped atoms, though many schemes exist<br />

which use cold atoms that are not confined during the interferometric process<br />

[Gus97, Pet99, Stu03]. This is closely related to the question <strong>of</strong> using internal<br />

or external [H¨01b] degrees <strong>of</strong> freedom for the interference. As a rule <strong>of</strong> thumb,<br />

the proposals using internal degrees use unconfined atoms, and usually the<br />

2 The other possible application is quantum information processing. How closely these<br />

two are related can be seen for example in [Dum03a], [M¨05] and [Vol05].<br />

45


2.6. Atom Interferometry<br />

beamsplitting process is done by momentum transfer due to the process <strong>of</strong><br />

manipulating these degrees <strong>of</strong> freedom [Kas91]. The exception to this rule<br />

are proposals that use state selective potentials. The use <strong>of</strong> external degrees<br />

<strong>of</strong> freedom for interferometry is usually connected to using trapped or guided<br />

atoms, where the these degrees then are quantised. The condensation <strong>of</strong> atoms<br />

into a BEC and with the trapping <strong>of</strong> atoms in microscopic potentials further in-<br />

creased the interest in atom interferometry. The production <strong>of</strong> cold atoms and<br />

BEC in microtraps on atom chips [Fol02, H¨01a, Ott01] and in micro-optical<br />

systems [Dum02a, Dum02b] has stimulated a great interest towards novel im-<br />

plementations <strong>of</strong> atom interferometers [Hin01, H¨01b, And02] that are based<br />

on the use <strong>of</strong> double well potentials. Now, a large number <strong>of</strong> coherent atoms, a<br />

macroscopic wavefunction, can be manipulated with very precise control. Only<br />

recently chip-based atom interferometers [Sch05b, Est05, Wan05b, Shi05] and<br />

an interferometer using an optical double well trap [Shi04] were demonstrated.<br />

These interferometers used cold atom clouds and BECs. A good overview <strong>of</strong><br />

the whole field <strong>of</strong> interferometry with atoms is given in the book <strong>of</strong> [Ber96].<br />

2.6.1 Atom interferometry with symmetric double well<br />

potentials<br />

The double well potential in quantum mechanics has for a long time been ex-<br />

amined by theoretical physicists. The first studies were in the field <strong>of</strong> solid<br />

state physics [Set81, Wei87] or inspired by effects that that were known from<br />

this field like Josephson oscillations [Jav86, Jac96]. With the advent <strong>of</strong> the<br />

Bose-Einstein condensate in atomic physics [And95, Dav95, Bra95], interfer-<br />

ence <strong>of</strong> a single BEC distributed over both wells and between BECs located<br />

in a well each was studied [Cas97, And97]. These experiments showed that<br />

the phase between two independent condensates was unpredictable, which was<br />

46


Chapter 2: Theoretical Background<br />

explained by the fact that the BECs in this case are Fock states with a fixed<br />

number <strong>of</strong> atoms and as such their relative phase is undefined [Jav96, R¨97].<br />

A phase between two BECs is always measurable, but it will be different and<br />

unpredictable from shot to shot.<br />

Double well atom interferometers are suited to be realised on atom chips<br />

[Jo07]. Precise control over the splitting and merging processes becomes possi-<br />

ble on a sub-micron scale when micr<strong>of</strong>abricated structures are used. Splitting<br />

[Cas00] and merging [H¨01c] <strong>of</strong> cold atomic clouds was demonstrated soon after<br />

the first atom chips became available, and interference <strong>of</strong> BECs split and re-<br />

combined in double wells [Shi04, Shi05] or in a chip based waveguide [Wan05b]<br />

has been demonstrated. The use <strong>of</strong> a BEC in interferometry can enhance the<br />

precision <strong>of</strong> measurements by a factor <strong>of</strong> √ N[Kas02]. At the moment, phase<br />

diffusion [Jav97] due to localisation <strong>of</strong> atoms is a major limitation [Shi04].<br />

Double well interferometry with single atoms will allow longer measurement<br />

time and larger splitting distances. In these regards single atom interferometry<br />

has an advantage. On the other hand, by the statistical nature <strong>of</strong> quantum<br />

mechanics, a single measurement <strong>of</strong> a single atom will not yield information<br />

about the phase. A way to overcome this is to repeat the experiment, or run<br />

identical experiments at the same time. Even more promising is a BEC <strong>of</strong><br />

non-interacting particles. Then the phase diffusion due to mean field effects<br />

can be neglected and we look at N single atoms in a single trap.<br />

Two schemes have been proposed to split a potential into a double well us-<br />

ing the technology <strong>of</strong> atom chips [Hin01, H¨01c]. These schemes use a three step<br />

approach. The starting point is an atom in the ground state <strong>of</strong> a symmetric<br />

potential. In the first step this potential is split adiabatically into a symmetric<br />

double well. Then a phase difference between the parts <strong>of</strong> the atom in each<br />

well is then induced. The double well is then adiabatically recombined into the<br />

original potential, and the population <strong>of</strong> the excited state measures the effect<br />

47


2.6. Atom Interferometry<br />

<strong>of</strong> the spatially-asymmetric potential. This can be viewed as a Mach-Zehnder<br />

interferometer where the two separated wells give rise to different optical paths.<br />

The three main steps are shown schematically in Figure 2.8.<br />

Figure 2.8: Schematic view on the interferometer process. An atom is prepared<br />

in the ground state <strong>of</strong> a potential (left). This potential is then split, so that for<br />

a symmetric the potential the ground state <strong>of</strong> that split potential populates<br />

both wells. Then a phase shift between the wells is introduced (centre). Upon<br />

merging the two wells again, this leads to a population in the first excited state<br />

<strong>of</strong> the merged potential. Solid line: potential, dashed: ground state, dotted:<br />

first excited state wavefunction<br />

The concept <strong>of</strong> adiabaticity needs to be addressed further. We start with<br />

the atom in the lowest energy eigenstate |φ0〉 <strong>of</strong> the single well. We can sup-<br />

press transitions to excited states by choosing the time scale for splitting and<br />

recombination stages much longer than the inverse energy gap between the<br />

relevant states. The energy gap between the ground state energy E0 and E2,<br />

the energy <strong>of</strong> the second excited state, always exceeds the gap between the<br />

ground state energy and the first excited state energy E1. We can adiabat-<br />

ically isolate the two lowest energy eigenstates (|φ0〉 and |φ1〉) from higher<br />

excited states (|φ2〉, |φ3〉, etc) by a suitable choice <strong>of</strong> the time scales:<br />

h · (E1 − E0) −1 ≫ Tadiabatic isolation ≫ h · (E2 − E0) −1<br />

(2.57)<br />

In the symmetric case, the states |φ2n〉 and |φ2n+1〉 become degenerate for large<br />

splittings, then the left hand side <strong>of</strong> inequality (2.57) becomes infinite and can<br />

48


Chapter 2: Theoretical Background<br />

be neglected [H¨01c]. In this case, one <strong>of</strong>ten changes from the energy basis to<br />

a localised basis, where<br />

|L〉 = 1<br />

√ 2 (|φ0〉 + |φ1〉)<br />

|R〉 = 1<br />

√ 2 (|φ0〉 − |φ1〉) (2.58)<br />

The basis vectors |L, R〉 indicate whether the atom is localised in the left or<br />

the right well. This can be seen from the central figure <strong>of</strong> in Figure 2.8: the<br />

energetic states |φ0,1〉 differ by a phase factor <strong>of</strong> −1 in their respective right<br />

well population. In the sum <strong>of</strong> these states, only the left well population is<br />

non-vanishing. In subtracting the states, this phase factor is canceled and only<br />

the right well is populated.<br />

This allows us to apply a two-mode approximation, using the two lowest<br />

energy eigenstates <strong>of</strong> the describing Hamiltonian. The theory <strong>of</strong> the inter-<br />

ferometer is then simpler and allows us to introduce a the two level Bloch<br />

equations. These can be visualised by the motion <strong>of</strong> the Bloch vector on the<br />

Bloch sphere and simplifies the interpretation <strong>of</strong> the physics. This is done in<br />

the following chapter, where the need for a symmetric potential is dropped<br />

and the more general case <strong>of</strong> an asymmetric double well potential is discussed.<br />

49


2.6. Atom Interferometry<br />

50


Chapter 3<br />

A single particle in an<br />

asymmetric double well<br />

potential<br />

This chapter addresses one problem that atom interferometric experiments<br />

have to deal with. The problem addressed here is an imperfection that is<br />

unavoidable in experimental set-ups and actually needed for an interferometer:<br />

the imbalance in the potentials <strong>of</strong> the two arms <strong>of</strong> a single particle, double<br />

well interferometer. Because this imbalance is needed to give rise to the phase<br />

difference, it will be explained how it has to be introduced and how it can<br />

destroy the interferometer if this is not done properly. Although only Mach-<br />

Zehnder-like interferometers will be specifically addressed, like in section 2.6.1,<br />

the results <strong>of</strong> this chapter can easily be generalized to other situations.<br />

A known approach is to express the single atom in a double well system by<br />

the localised wavefunctions <strong>of</strong> the left and right well as in equation (2.58). This<br />

requires the truncation to the lowest two states <strong>of</strong> the symmetric double well,<br />

which is a well known method when treating double well problems [Gri98].<br />

51


Localised wavefunctions were also used to examine a time dependent symmet-<br />

ric double well [H¨01c] and time independent asymmetric double wells [Hu00].<br />

The problem <strong>of</strong> non-adiabatic transitions in two level systems was first raised<br />

in 1927 by F. Hund [Hun27] and treated in 1932 individually by L.D. Landau,<br />

C. Zener and E.G.C. Stückelberg [Lan32, Zen32, Stu32]. Their work is now<br />

known as the Landau-Zener-Stückelberg (LZS) theory and is concerned with<br />

exact solutions for special forms <strong>of</strong> potential curves. For a so-called Landau-<br />

Zener (LZ) type <strong>of</strong> curve crossing, the slopes <strong>of</strong> the potential curves have the<br />

same sign. In terms <strong>of</strong> the double well potential, the LZ model has a linearly<br />

varying asymmetry and a constant underlying symmetric potential. For an<br />

extensive overview <strong>of</strong> the topic and several exact analytical solutions <strong>of</strong> special<br />

potentials, see [Nak02].<br />

In our case <strong>of</strong> the asymmetric double well potential, we also deal with a<br />

LZ-type <strong>of</strong> curve crossing, except that the potentials do not cross but asymp-<br />

totically become degenerate. More importantly the use <strong>of</strong> the LZS theory is<br />

impossible in our case, as neither <strong>of</strong> the two important parameters is constant;<br />

both are time varying. In addition, neither <strong>of</strong> them can be neglected over the<br />

full range <strong>of</strong> our interest. Indeed, we are most interested in the range where<br />

both parameters are <strong>of</strong> the same order <strong>of</strong> magnitude. So, instead <strong>of</strong> <strong>of</strong>fer-<br />

ing exact or approximation solutions for special cases [Bam81, Rob82], here<br />

the emphasis is on the interpretation and understanding <strong>of</strong> the physics. That<br />

this understanding is needed can be seen by the the examples in the field <strong>of</strong><br />

atom interferometry where the influence <strong>of</strong> asymmetries in experimental set-<br />

ups seems to be underestimated or ignored, see for example footnote [24] <strong>of</strong><br />

[Shi05], where a difference between the trap frequencies <strong>of</strong> the wells <strong>of</strong> 12%<br />

was mentioned but the effect <strong>of</strong> that asymmetry on the coherence between the<br />

separated BECs was neglected.<br />

The behaviour <strong>of</strong> a BEC in a symmetric double well system has been the<br />

52


Chapter 3: The Asymmetric Double Well<br />

subject <strong>of</strong> theoretical research [Jav97, Mil97, Sti02b, Tho02, Sak02], includ-<br />

ing analytical investigations using a two mode approximation [Jav99, Cor99,<br />

Spe99, Men01, Sti02a, Mah02]. The influence <strong>of</strong> non-adiabaticity <strong>of</strong> the split-<br />

ting <strong>of</strong> a BEC has also been examined [Bor04, Mel03] during the time the<br />

model presented here has been developed. Mellish et al. use a two mode ap-<br />

proach and “dressed state” basis states like in solid state physics to describe<br />

the loading <strong>of</strong> a BEC into an optical lattice [Mel03]. Smerzi et al. have exam-<br />

ined a BEC in an asymmetric double well, using two coupled Gross-Pitaevskii<br />

equations in the far split regime only, without overlap between the wavefunc-<br />

tions <strong>of</strong> the different wells [Sme97]. During the time the model here has been<br />

developed, Sakellari et al. have studied Josephson effects <strong>of</strong> BECs in double<br />

wells by introducing a time dependent linear asymmetry and using a two mode<br />

approach [Sak04].<br />

Single particles in double well systems have also been the subject <strong>of</strong> theo-<br />

retical research. The use <strong>of</strong> degenerate isomers <strong>of</strong> molecules as double well sys-<br />

tems for detection <strong>of</strong> very small energy differences has been proposed [Har78],<br />

and later the theoretical justification for the sensitivity <strong>of</strong> double well systems<br />

has been studied [Sim85]. The multi-state dynamics <strong>of</strong> magnetically trapped<br />

atoms with a double well formed by the avoided crossing <strong>of</strong> atomic spin states<br />

and RF radiation have been studied [Vit97], based on previous work with<br />

time independent fields, by Cook and Shore [Coo79]. Double wells and the<br />

influence <strong>of</strong> pulsed coupling have been examined from a LZS point <strong>of</strong> view<br />

[Bam81, Rob82]. Tunnelling processes, their resonances and how a modula-<br />

tion <strong>of</strong> the potential changes these resonances has been studied [Gro91]. A<br />

good overview on two mode models and double well systems with an emphasis<br />

on these tunnelling processes is presented in [Gri98].<br />

This chapter begins with the introduction <strong>of</strong> the model system and Hamil-<br />

tonian and how the imbalance is introduced. This system is kept in a most<br />

53


3.1. Introduction<br />

general form, so that it is easily applicable to other similar situations. The sec-<br />

ond section deals with the main approximation made to simplify the model:<br />

the system is restricted to the lowest two levels leading to the well known<br />

Bloch equations. These equations allow easy interpretation <strong>of</strong> the different<br />

parameters and their roles in the interferometric process. Some obvious lim-<br />

itations <strong>of</strong> the model will be discussed here as well. This is then followed by<br />

the numerical results <strong>of</strong> this model and compared with the numerical solutions<br />

<strong>of</strong> the full Hamiltonian and experimental data <strong>of</strong> a BEC based interferometer<br />

[Jo07]. This part <strong>of</strong> the chapter includes a novel way to read out an atom in-<br />

terferometer, similar to that which has been used for a BEC based double well<br />

experiment in our group [Hal07b]. The last section <strong>of</strong> this chapter summarizes<br />

the insights <strong>of</strong> the model and its limitations. A short list <strong>of</strong> similar physical<br />

problems where this model can be used is presented. The results <strong>of</strong> the first<br />

sections <strong>of</strong> the chapter have been published [Sid06]; some later results were<br />

obtained after the contents <strong>of</strong> the publication had been decided and for that<br />

reason only have not been included in the paper. The author withdrew from<br />

his position as first author on that publication after agreement could not be<br />

reached with some <strong>of</strong> the co-authors.<br />

3.1 Introduction<br />

We investigate the dynamics <strong>of</strong> a single particle due to a time dependent, one<br />

dimensional Hamiltonian ˆ H(t) [Sid06]. Any Hamiltonian can be expanded into<br />

a sum <strong>of</strong> one part with even symmetry and one part with odd symmetry. The<br />

Hamiltonian thus can be written as the sum <strong>of</strong> its symmetric part ˆ H0(t) and an<br />

antisymmetric part ˆ Vas. For the symmetric potential ˆ V0(x, t) a rather general<br />

54


form is chosen [Pul03]:<br />

ˆH = ˆ H0 + ˆ Vas<br />

∂2 Chapter 3: The Asymmetric Double Well<br />

ˆH0 = − 1<br />

2 ∂x2 + ˆ V0(x, t) (3.1)<br />

� �<br />

ˆV0(x, t) = 1 + β(t) − x2<br />

� �<br />

2<br />

1/2<br />

(3.2)<br />

2<br />

ˆVas = Vasˆx (3.3)<br />

The Hamiltonian is normalised by dividing by ¯hω0, and the variables are<br />

rescaled by<br />

x = x′<br />

a0<br />

t = ω0 · t ′<br />

V =<br />

a0 =<br />

V ′<br />

¯hω0<br />

�<br />

¯h<br />

mω0<br />

(3.4)<br />

where the primed quantities are the variables <strong>of</strong> the unscaled Hamiltonian.<br />

Dimensionless harmonic oscillator units are used. The normalised asymmetric<br />

part <strong>of</strong> the potential ˆ Vas is linear in the spatial variable x and constant in time.<br />

The eigenvectors <strong>of</strong> ˆ H(t) will be labelled |φi〉 with eigenvalues Ei. The<br />

eigenvectors <strong>of</strong> ˆ H0(t) are |σi〉 and their eigenvalues are Eis, where for both<br />

cases i = 0, 1, 2 . . .. All eigenvalues and eigenvectors are time dependent. The<br />

eigenvectors <strong>of</strong> the symmetric Hamiltonian have defined symmetry. A notation<br />

common in the field <strong>of</strong> solid states physics [Hu00] calls the symmetric ground<br />

state |S〉 = |σ0〉 and the first excited and thus lowest anti-symmetric state<br />

|AS〉 = |σ1〉.<br />

The shape <strong>of</strong> the potential is characterised by a time dependent splitting<br />

parameter β(t). For β = 0 it is a single well that is nearly quartic. The<br />

splitting is realised by increasing this parameter to a large value. For β ≫ 0<br />

55


3.2. Two Mode Approximation and the Bloch Equations<br />

the double well potential appears with two nearly harmonic wells. The distance<br />

between the minima is 2 √ 2β. In the following β(t) will change linearly or will<br />

be constant. We are allowed to choose our eigenvectors |φi〉, |σi〉 to be real, as<br />

a one dimensional problem is considered.<br />

3.2 Two Mode Approximation and the Bloch<br />

Equations<br />

3.2.1 The two mode approximation<br />

The two mode approximation is now applied. The eigenstates <strong>of</strong> the total<br />

Hamiltonian ˆ H(t) do not possess a defined symmetry. We will use a standard<br />

matrix mechanics approach [Mes90] in the basis <strong>of</strong> localised atoms, the so-<br />

called left, right basis |L, R〉 (L-R basis). The basis vectors are defined as<br />

|L〉 = 1<br />

√ 2 (|σ0〉 + |σ1〉)<br />

|R〉 = 1<br />

√ 2 (|σ0〉 − |σ1〉) (3.5)<br />

For the split potential, β ≫ 0, the states |L, R〉 describe atoms that are lo-<br />

calised in the left or right well. For the unsplit potential the states |L, R〉<br />

correspond to higher probability amplitudes at the sides <strong>of</strong> the single well. In<br />

the left, right basis, the Hamiltonian takes the form<br />

ˆHL,R = 1<br />

⎡<br />

⎛<br />

⎣(E0 + E1) · ˆ1 −<br />

2<br />

56<br />

⎝ Vasym<br />

∆0<br />

∆0 −Vasym<br />

⎞⎤<br />

⎠⎦<br />

(3.6)


Chapter 3: The Asymmetric Double Well<br />

Here ˆ1 is the unity matrix, the columns and rows are ordered L-R and the<br />

following definitions are used:<br />

∆0 = E1s − E0s (3.7)<br />

Vasym = 〈R| ˆ Vas|R〉 − 〈L| ˆ Vas|L〉 (3.8)<br />

= −〈σ0| ˆ Vas|σ1〉 − 〈σ1| ˆ Vas|σ0〉<br />

All <strong>of</strong> these quantities are real. The energy gap between the eigenstates <strong>of</strong> the<br />

symmetric Hamiltonian |σ0,1〉 is given by ∆0. In solid state physics this term<br />

sometimes is denoted by ∆SAS as the energy difference between |S〉 = |σ0〉 and<br />

|AS〉 = |σ1〉. The quantity Vasym is a measure <strong>of</strong> the asymmetry between the<br />

wells and would be zero for a symmetric Hamiltonian. These two quantities<br />

are related to the energy gap <strong>of</strong> the total Hamiltonian ˆ H(t), which we call ∆<br />

following a notation used in [Hu00], by:<br />

∆ 2 = ∆ 2 0 + V 2<br />

asym = (E1 − E0) 2<br />

(3.9)<br />

The dependence <strong>of</strong> ∆, Vasym and ∆0 on the splitting parameter β(t) is shown<br />

in Fig. 3.1. The energies were calculated numerically for the symmetric and<br />

asymmetric potential using a Numerov recursion method, and the difference<br />

was calculated from equation (3.9). All quantities are expressed in the natural<br />

oscillator units <strong>of</strong> equations (3.4), so that they can easily be rescaled with<br />

the actual experimental parameters. The antisymmetric contribution roughly<br />

follows a square root behaviour, but deviates for low splitting values. The<br />

square root can be explained thus. The definition <strong>of</strong> the symmetric splitting<br />

potential, equation (3.2), establishes a square root relation between β (linear)<br />

and x (squared). The asymmetry is defined by equation (3.8) and is linear in<br />

the distance between the wells. This means it has a square root dependence<br />

on β. The overlap between the states |L, R〉 is suspected to be the reason<br />

for the deviation from the square root. The symmetric part shows a decay<br />

57


3.2. Two Mode Approximation and the Bloch Equations<br />

proportional to e−(β−δ)2, like the falling slope <strong>of</strong> a Gaussian displaced by δ.<br />

A square root with a quartic root correction and a Gaussian had been fitted<br />

to the points that were directly determined from the stationary Schrödinger<br />

equation, and were used in the later calculations <strong>of</strong> the model.<br />

Energy in hν<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 1 2 3 4 5 6 7<br />

splitting parameter β<br />

Figure 3.1: Energy difference ∆ between ground and first excited state (solid<br />

line) for ˆ Vas = 0.02x, as a function <strong>of</strong> the splitting parameter β. Dotted line:<br />

symmetric fraction ∆0; dashed: asymmetric fraction Vasym.<br />

The energy eigenvectors can be expressed in terms <strong>of</strong> the L,R basis, and<br />

are given by<br />

|φ0〉 = 1<br />

√ 2 ( √ 1 + V |L〉 + √ 1 − V |R〉)<br />

|φ1〉 = 1<br />

√ 2 ( √ 1 − V |L〉 − √ 1 + V |R〉) (3.10)<br />

where the asymmetry is now represented by the scaled quantity<br />

V = Vasym<br />

∆<br />

(3.11)<br />

The degree <strong>of</strong> asymmetry is described by V , which is related to the angle θ<br />

used for similar situations in the field <strong>of</strong> solid state physics, so that V=tan θ or<br />

58


Chapter 3: The Asymmetric Double Well<br />

V=cot θ. This θ also appears in equations (2.7) and (2.9) <strong>of</strong> the dressed states<br />

model in section 2.1.1. It has to be pointed out that for a nearly symmetric<br />

(unsplit) trap, we still have<br />

|L〉 = 1<br />

√ 2 (|φ0〉 + |φ1〉)<br />

|R〉 = 1<br />

√ 2 (|φ0〉 − |φ1〉) (3.12)<br />

but for a far split trap, when V ≈ 1, the left and right well states can be<br />

expressed by<br />

|L〉 = |φ0〉<br />

|R〉 = −|φ1〉 (3.13)<br />

When looking at the eigenfunctions |φ0,1〉 for different splittings β, we can<br />

see why a two mode approximation should be sufficient to describe the inter-<br />

ferometer. Figure 3.2 shows these eigenfunctions for three values <strong>of</strong> β and a<br />

fixed asymmetry. We see how over a small increase in the splitting distance<br />

the wave functions change from more or less symmetric functions to fully lo-<br />

calised eigenfunctions. This has been called the “flea on the elephant” effect<br />

[Sim85], when a very small disturbance (for Rb in a trap <strong>of</strong> frequency 1 kHz,<br />

the asymmetry only needs to be about 1/36 <strong>of</strong> gravity to reach our model value<br />

<strong>of</strong> 0.02·x) has a huge effect on the system [Har78]. This is the main reason why<br />

one should not consider symmetric potentials as described in section 2.6.1.<br />

3.2.2 The Bloch equations<br />

When dealing with the dynamics <strong>of</strong> a two level system, like in section 2.1, one<br />

can always describe the system by the Bloch equations, as all two level prob-<br />

lems have been shown to be equivalent [Fey57] and can be directly mapped<br />

onto the special unitary group SU(2). Another way to represent such a prob-<br />

lem is by rotating vectors on a unit sphere, the so-called Bloch sphere. This<br />

59


3.2. Two Mode Approximation and the Bloch Equations<br />

ψ<br />

1<br />

0.5<br />

0<br />

−0.5<br />

(a) β=1<br />

−1<br />

−6 −4 −2 0<br />

x<br />

2 4 6<br />

20<br />

15<br />

10<br />

5<br />

0<br />

V(x)<br />

ψ<br />

1<br />

0.5<br />

0<br />

−0.5<br />

(b) β=2.5<br />

−1<br />

−6 −4 −2 0<br />

x<br />

2 4 6<br />

20<br />

15<br />

10<br />

5<br />

0<br />

V(x)<br />

ψ<br />

1<br />

0.5<br />

0<br />

−0.5<br />

(c) β=5<br />

−1<br />

−6 −4 −2 0<br />

x<br />

2 4 6<br />

Figure 3.2: Instantaneous energy eigenfunctions <strong>of</strong> the ground state (dashed-<br />

dotted line) and the first excited state (dotted line) for different splittings β =<br />

(a) 1, (b) 2.5,(c) 5. The solid line is the asymmetric potential. The asymmetry<br />

is ˆ Vas = 0.02ˆx<br />

allows a graphical representation <strong>of</strong> the different states and the parameter V ,<br />

which were introduced in the previous paragraphs. A projection onto two di-<br />

mensions <strong>of</strong> this sphere is shown in Fig. 3.3. As our physical system is one<br />

dimensional and we can choose our eigenfunctions to be real, this projection is<br />

a good means to visualise the problem and understand the underlying physics.<br />

In [All75, Mey91] extensive treatments <strong>of</strong> the two level problem are given,<br />

including the derivation <strong>of</strong> the Bloch equations. It was pointed out by [Dal04]<br />

that these derivations [All75] rely on time independent basis vectors. In the<br />

case <strong>of</strong> time dependent basis vectors these derivations are valid if and only<br />

if there exists a set <strong>of</strong> basis vectors that are real and with opposite symme-<br />

try. Fortunately these two restrictions are trivially fulfilled in any two level<br />

problem: any Hamiltonian can be expanded as a sum <strong>of</strong> even and odd func-<br />

tions, and as basis vectors the eigenvectors <strong>of</strong> the even part <strong>of</strong> the Hamiltonian<br />

are used. The eigenfunctions <strong>of</strong> this symmetric Hamiltonian always have de-<br />

fined symmetry. All two level systems are one dimensional or have degenerate<br />

energy levels. Degeneracy happens through dimensions in the Hamiltonian<br />

being symmetric. The one dimensionality <strong>of</strong> the problem allows us to choose<br />

the phase <strong>of</strong> these eigenvectors as zero so that the eigenvectors are real. Thus,<br />

60<br />

20<br />

15<br />

10<br />

5<br />

0<br />

V(x)


R<br />

V<br />

σ 0<br />

Ψ<br />

Chapter 3: The Asymmetric Double Well<br />

φ 1<br />

Figure 3.3: Projection <strong>of</strong> the Bloch sphere, with energy eigenstates |φ0,1〉, and<br />

different basis states |L, R〉, |σ0,1〉. The situation is an incomplete split <strong>of</strong> the<br />

traps, V ≈ 1/2. The state |Ψ〉 is closer to |φ0〉 than to |σ0〉, so the splitting<br />

is nearly adiabatic. The projection onto the |L, R〉 axis shows how the atomic<br />

wavefunction is overbalanced to the left well.<br />

it is valid to use the derivation as presented in [All75]. We also find the first<br />

restriction to our approximation. It will not hold when the ground or first<br />

excited state <strong>of</strong> the asymmetric potential is a linear combination with a signif-<br />

icant contribution from the second or higher excited states <strong>of</strong> the symmetric<br />

basis. This can happen for example when the asymmetry and splitting are too<br />

large, so that the asymmetric first excited state becomes localised in the same<br />

well as the ground state.<br />

φ 0<br />

We can now apply the standard treatment <strong>of</strong> the two level problem to our<br />

system: any state vector can be written as<br />

σ1<br />

|Ψ〉 = cL|L〉 + cR|R〉 (3.14)<br />

61<br />

L


3.2. Two Mode Approximation and the Bloch Equations<br />

where the complex amplitudes fulfil |cL| 2 + |cR| 2 = 1. The state vector |Ψ〉<br />

fulfils the normalised time-dependent Schrödinger equation<br />

where the Hamiltonian is given by equation (3.6).<br />

i ∂<br />

∂t |Ψ(t)〉 = ˆ H(t)|Ψ(t)〉 (3.15)<br />

We identify the components <strong>of</strong> the Bloch vector −→ Y as<br />

Yx = cLc ∗ R + c ∗ LcR = Re(cLc ∗ R)<br />

Yy = i · (c ∗ LcR − cLc ∗ R) = Im(cLc ∗ R)<br />

Yz = |cR| 2 − |cL| 2 ,<br />

The Bloch equations can be obtained from the Schrödinger equation (3.15)<br />

when the state vector is substituted using equation (3.14). The Bloch equa-<br />

tions 1 are then given by<br />

˙<br />

Yx= −Yy · Vasym<br />

˙<br />

Yy= Yx · Vasym +Yz · ∆0<br />

˙<br />

Yz= −Yy · ∆0<br />

(3.16)<br />

and can be solved numerically as rate equations using the Runge-Kutta algo-<br />

rithm. These equations are analogous to those for the case <strong>of</strong> a two level atom<br />

driven by a single mode laser field. The symmetric transition frequency ∆0<br />

is equivalent to the Rabi frequency, while the asymmetry frequency Vasym is<br />

equivalent to the detuning. The component Yz is a measure <strong>of</strong> the imbalance<br />

<strong>of</strong> the atomic population between the wells (Yz = 0 means balanced popula-<br />

tion), and in the far split case it is a measure <strong>of</strong> the excitation amplitude <strong>of</strong><br />

the first vibrational state. The component Yx contains information about the<br />

symmetry <strong>of</strong> the atomic wave function: for Yx = −1 the state is symmetric,<br />

while the antisymmetric state is given by Yx = 1. It is thus a measure <strong>of</strong> the<br />

excitation <strong>of</strong> the vibrational state in the unsplit trap.<br />

1 I thank B. Dalton for correcting a factor in these equations [Dal04].<br />

62


Chapter 3: The Asymmetric Double Well<br />

The Bloch equations can be expressed as a vector product<br />

d −→ Y<br />

dt = −→ Ω × −→ Y , (3.17)<br />

where the torque vector is −→ Ω = (−∆0, 0, Vasym). This form <strong>of</strong> the vector with<br />

‘Rabi frequency’ ∆0 and ‘detuning’ Vasym also explains the definition <strong>of</strong> the<br />

‘generalised Rabi frequency’ ∆ in equation (3.9). In the case <strong>of</strong> atoms in RF<br />

or laser fields, the Rabi frequency is determined by the external driving field.<br />

This driving field usually is approximated by the rotating wave approximation<br />

(RWA). In our case, no such approximation is needed; the Rabi frequency is<br />

fully determined by the symmetric energy gap. It is noted that no relaxation<br />

terms appear in the Hamiltonian and therefore no decay to the ground state<br />

or other states is taken into account.<br />

A high potential barrier suppresses tunnelling between the wells (∆0 → 0).<br />

In the symmetric case (Vasym = 0) the left and right well states become de-<br />

generate. This is equivalent to zero detuning or resonance. Thus, in an inter-<br />

ferometric process, we do not observe Rabi oscillations (like the Stückelberg<br />

oscillations seen in the group <strong>of</strong> R. Grimm [Chi05a] on a system <strong>of</strong> ultracold Cs<br />

molecules [Chi05b]), but instead we observe Larmor precessions. The principal<br />

difference between these two is shown in the schematic diagram <strong>of</strong> Fig. 3.4.<br />

Here the spectral flow <strong>of</strong> the levels used for interferometry is depicted. The<br />

thick lines show the evolution <strong>of</strong> the coupled levels, or dressed states, during<br />

the beam splitting (or trap splitting in our case), the free evolution or phase<br />

accumulation, and the final beam recombining (in our case the recombining <strong>of</strong><br />

the trap). In the Rabi case (Fig. 3.4 left) we see the uncoupled states (dotted<br />

lines) cross so that the previously higher excited state becomes the state with<br />

lower energy. The coupling between the states leads to an avoided crossing.<br />

The Larmor case (Fig. 3.4 right) is different; here we see the uncoupled states<br />

degenerate, and only the coupling lifts this degeneracy, leading to a split <strong>of</strong><br />

63


3.2. Two Mode Approximation and the Bloch Equations<br />

Energy<br />

Time<br />

Figure 3.4: Schematic view <strong>of</strong> the spectral flow <strong>of</strong> the coupled (thick line)<br />

and uncoupled (dotted line) energy levels for an interferometer using Rabi<br />

oscillations (left) and Larmor precession (right), as functions <strong>of</strong> time. The plots<br />

show the full interferometer process including splitting, holding and merging.<br />

The double well interferometer is <strong>of</strong> the Larmor kind with coupled levels.<br />

the two states <strong>of</strong> the interferometer. In a sense, the Larmor interferometer is<br />

a special case <strong>of</strong> the Rabi interferometer: it just stops at the crossing point <strong>of</strong><br />

the uncoupled levels and stays there during the phase evolution.<br />

Energy<br />

We are interested in the excitation probability after the interferometric<br />

process. Thus, when we assume an arbitrary state |Ψ〉 = cL|L〉 + cR|R〉, we<br />

want to calculate |〈φ1||Ψ〉| 2 . We find:<br />

|〈φ1|||Ψ〉| 2 = |〈φ1|(cL|L〉 + cR|R〉)| 2<br />

= 1/2((|cL| 2 + |cR| 2 ) + V (|cR| 2 − |cL| 2 )<br />

− √ 1 − V 2 (cLc ∗ R + c ∗ LcR)<br />

= 1<br />

2 (1 + Yz · V − Yx · √ 1 − V 2 ) (3.18)<br />

The result depends on the parameter V . It is thus easy to probe two <strong>of</strong> the<br />

three Bloch components by measuring in a merged trap (V ≈ 0) or in a split<br />

trap (V ≈ 1).<br />

In the recombined trap, the influence <strong>of</strong> the asymmetry is small, ∆ ≈ ∆0<br />

64<br />

Time


Chapter 3: The Asymmetric Double Well<br />

and V ≪ 1. In the far split trap the asymmetry dominates: ∆ ≈ Vasym and<br />

V ≈ 1. Equation (3.18) then reduces to<br />

|〈φ1|||Ψ〉| 2 = 1 − Yx for 2<br />

|〈φ1|||Ψ〉| 2 = 1 + Yz for 2<br />

the unsplit case<br />

the split case<br />

(3.19)<br />

Especially the second equation is <strong>of</strong> interest. The consequences <strong>of</strong> that equation<br />

are discussed in section 3.3.3.<br />

3.3 The Results <strong>of</strong> the Model<br />

3.3.1 General results: the interferometer and the im-<br />

pact <strong>of</strong> adiabaticity<br />

In this section the results <strong>of</strong> the rate equation (3.16) will be presented. These<br />

are then compared with the results <strong>of</strong> the numerical integration <strong>of</strong> the Schrödinger<br />

equation without the two mode approximation, which were calculated in our<br />

group by S. Whitlock [Whi04] using the eXtensible Multi-Dimensional Sim-<br />

ulator (XMDS) s<strong>of</strong>tware developed at the <strong>University</strong> <strong>of</strong> Queensland 2 . This<br />

comparison will show another limitation <strong>of</strong> the approximation and a good<br />

agreement apart from this, with a considerably shorter calculation time. The<br />

effects will be addressed and the physical meaning will be explained. Unless<br />

otherwise noted, the asymmetry will be ˆ Vas = 0.02ˆx and any ramps <strong>of</strong> the<br />

splitting β will be linear.<br />

The rate equations can be used to examine the important parameters. For<br />

the symmetric case, it is important not to excite the atom into the second ex-<br />

cited state. Thus, experimentalists might be tempted to extend the splitting<br />

time when they encounter problems in balancing the atomic population in the<br />

traps. Figure 3.5 shows the behaviour that is to be expected in this case. Here<br />

2 For more information see: http://www.xmds.org<br />

65


3.3. The Results <strong>of</strong> the Model<br />

V; Y z ; |〈 φ 1 (x) | Ψ 〉| 2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

−0.2<br />

−0.4<br />

−0.6<br />

−0.8<br />

−1<br />

0 20 40 60 80 100 120 140 160 180 200<br />

Figure 3.5: Dynamic evolution <strong>of</strong> the parameters V (dashed), the probability<br />

for excitation <strong>of</strong> the first excited state |〈φ1||Ψ〉| 2 (dotted) and the Bloch vector<br />

component Yz (solid line) for a splitting time <strong>of</strong> Ts = 200 and a final splitting<br />

β = 3.5. This figure shows an adiabatic splitting, with nearly no excitation<br />

<strong>of</strong> the atom in the potential well. After the split, the atom is localised in the<br />

energetically lower well.<br />

the splitting time was Ts = 200 for a final splitting <strong>of</strong> β = 3.5. The plot shows<br />

the evolution <strong>of</strong> the parameters V (dashed line), the excitation <strong>of</strong> the first<br />

excited state (dotted line) and the parameter Yz (solid line), which contains<br />

information about the balance between the wells during the splitting. The ex-<br />

citation shows us, that for this splitting time and parameter values, the overall<br />

evolution is adiabatic. Of course, this leads to the population accumulating in<br />

the lower well (Yz = −1 corresponds to all <strong>of</strong> the atom in the lower well). With<br />

the atom localised in one well, this cannot be used as an interferometer, as we<br />

have intrinsic which-way information [Eng96]. Thus, when doing experiments<br />

with double wells one has to take care that the splitting time is adiabatic for<br />

the transition |φ0〉 → |φ2〉 but non adiabatic for |φ0〉 → |φ1〉.<br />

T s<br />

If we keep this mind, we see how the interferometer works. This is shown in<br />

66


Chapter 3: The Asymmetric Double Well<br />

Fig. 3.6 for the same splitting, holding and recombining times Ts,h,r = 20 and<br />

a final splitting <strong>of</strong> β = 12.5. Here the figure shows the component Yx (solid<br />

V; Y x ; |〈 φ 1 (x) | Ψ 〉| 2<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

0 10 20 30 40 50 60<br />

total time T<br />

Figure 3.6: Dynamic evolution <strong>of</strong> the parameters V (dashed), the probability<br />

for excitation <strong>of</strong> the first excited state |〈φ1||Ψ〉| 2 (dotted) and the Bloch vector<br />

component Yx (solid line) for equal times <strong>of</strong> Ts,h,r = 20 and a final β = 12.5.<br />

This figure shows a full interferometer process with splitting, phase evolution<br />

and merging. During the non adiabatic split the atom is excited into a balanced<br />

superposition <strong>of</strong> the ground and first excited state. In the holding stage, the<br />

excitation <strong>of</strong> the atom does not change in the potential well, although the<br />

phase between the population in the left and right well oscillates. This is the<br />

well known Larmor precession. When the trap is recombined non adiabatically<br />

this phase difference gives rise to a population difference in the energy levels.<br />

line), the measure <strong>of</strong> the symmetry and thus <strong>of</strong> the inversion <strong>of</strong> the energetic<br />

states in the unsplit trap. As the Bloch vector is normalised, | � Y | 2 = 1, our<br />

knowledge <strong>of</strong> Yz ≈ 0 3 allows us to directly relate Yy to Yx. As the splitting is<br />

much larger than before, the critical time for splitting and merging is less than<br />

3 In the limit <strong>of</strong> small times, non adiabatic splitting is the same as the sudden approxi-<br />

mation [Sch65] which keeps the shape and thus the symmetry <strong>of</strong> the wavefunction.<br />

67


3.3. The Results <strong>of</strong> the Model<br />

the actual time <strong>of</strong> the ramps. During a time T <strong>of</strong> about 5, the parameter V<br />

goes from zero to one (and in the same amount <strong>of</strong> time it goes back to zero<br />

later in the merging process). After this time we see that the component Yx<br />

precesses around the revolution axis. The precession frequency is constant for<br />

the time between 20 and 40; before and after that the precession frequency<br />

changes slightly as the trap is still being split/merged and thus the driving<br />

potential difference between the wells changes. A good interferometer works<br />

when both paths are equally occupied [D¨98]; so in our case we need a balanced<br />

distribution between the left and right well. We know from equations (3.13)<br />

that this means we need a state corresponding to |Ψ〉 ≈ 1<br />

√ 2 (|φ0〉 + |φ1〉). From<br />

Equation (3.18) we also know that the excitation does not change during the<br />

phase evolution stage. A change in excitation in the split trap would result<br />

in a change <strong>of</strong> the population difference between the wells, which is unwanted<br />

and, due to the high tunnelling barrier, also impossible. The plot, Fig. 3.6,<br />

shows this behaviour clearly: the dotted line depicts the excitation probability.<br />

It stays constant at about 1<br />

2<br />

when the parameter V changes.<br />

during the phase accumulation, and changes only<br />

During the interferometric process, a phase difference between the localised<br />

wavefunctions is accumulated. This illustrates the fact that we look at a<br />

precession (Larmor) and not an oscillation (Rabi). Only during the splitting<br />

and merging is this phase difference translated into an excitation.<br />

These results were compared with the results from the above mentioned<br />

full numerical integration <strong>of</strong> the Schrödinger equation <strong>of</strong> S. Whitlock [Whi04].<br />

Figure 3.7 shows the excitation probability after a complete interferometer cy-<br />

cle, from left to right with different splitting and recombination times Ts,r =<br />

5, 20, 200, as a function <strong>of</strong> the holding time, with a final β = 12.5. This value<br />

for β was chosen as it translates into a splitting <strong>of</strong> the two wells <strong>of</strong> ten times<br />

the ground state sizea0. The points are the results <strong>of</strong> the full numerical sim-<br />

68


|〈 φ 1 (x) | Ψ 〉| 2<br />

1<br />

0.5<br />

0<br />

0 10 20 30 T 40<br />

h<br />

a)<br />

|〈 φ 1 (x) | Ψ 〉| 2<br />

1<br />

0.5<br />

Chapter 3: The Asymmetric Double Well<br />

0<br />

0 10 20 30 T 40<br />

h<br />

b)<br />

|〈 φ 1 (x) | Ψ 〉| 2<br />

1<br />

0.5<br />

0<br />

0 10 20 30 T 40<br />

h<br />

Figure 3.7: Excitation probability for different splitting and recombining times<br />

Ts,r = (a) 5, (b) 20, (c) 200 as a function <strong>of</strong> holding time Th. Solid line: Bloch<br />

model results; points: results from full numerical simulation [Whi04]. (a): A<br />

too short a splitting/recombining time leads to non adiabatic excitation to<br />

higher states, causing a deviation between the two-mode-approximation and<br />

the full analysis. (b): for longer times the agreement is very good. (c): A<br />

too long a splitting/recombining time leads to adiabatic splitting and merging<br />

with reduced fringe size. The different lines here are from different fits to the<br />

asymmetric component Vasym (see text). Errors in this fit lead to phase errors<br />

that accumulate during a prolonged splitting and merging.<br />

ulation; the lines are the results <strong>of</strong> the two mode approximation. For short<br />

splitting/recombining times Ts,r = 5, we see a deviation. While the Bloch<br />

model reaches unity values in the excitation probability, this is not the case<br />

for the numeric calculations which incorporate more levels. This is due to non<br />

adiabatic transitions to higher energy levels which the Bloch model ignores.<br />

Indeed for even smaller times the signal from the full numeric calculation shows<br />

an obvious modulation on the signal with a higher frequency, and the results<br />

<strong>of</strong> the Bloch model act as an envelope. For splitting/recombining times <strong>of</strong><br />

20, both models show remarkable agreement. In Fig. 3.7 (c), the signal <strong>of</strong><br />

the excitation reaches only values <strong>of</strong> |〈φ1||Ψ〉| 2 ≈ 1/2. This can be attributed<br />

to the onset <strong>of</strong> adiabaticity in the transition |φ0〉 → |φ1〉. This results in an<br />

unbalanced distribution <strong>of</strong> the atom between the wells, which in turn leads to<br />

69<br />

c)


3.3. The Results <strong>of</strong> the Model<br />

a reduction in the final signal. It can be seen as intrinsic which-way informa-<br />

tion when the atom preferentially populates one well, and this destroys the<br />

interference [Eng96]. The dashed and dotted line in Figure 3.7 (c) use differ-<br />

ent fits for the antisymmetric part: one is a square root without the quartic<br />

root correction, and the other is a natural logarithm. It is notable that the<br />

model is robust with regard to the maximum signal, but is fragile regarding<br />

the phase <strong>of</strong> the signal. This error is accumulated during the splitting and<br />

merging phases, when 0 < V < 1, as a result <strong>of</strong> the fact that Vasym �= 0 in a<br />

merged trap because <strong>of</strong> the overlap between |L〉 and |R〉. Any error will scale<br />

with the splitting/recombining time and thus it appears dominant when these<br />

times become longer, and a small relative error can add up to a large absolute<br />

error.<br />

As the quality <strong>of</strong> the interferometric signal will depend on how symmetric<br />

the atomic population is distributed between the wells, we examine the popu-<br />

lation <strong>of</strong> both wells after the splitting process only. We define a filling factor<br />

F as a measure <strong>of</strong> the symmetry, requiring F = 0 if the atom is fully localised<br />

in one well and F = 1 if it is evenly distributed. With |Ψ〉 = a · |φ0〉 + b · |φ1〉,<br />

we define F as<br />

F = 2 · |a| · |b| (3.20)<br />

This fulfils our requirements. The decay <strong>of</strong> the filling factor is shown in Fig.<br />

3.8, for a splitting <strong>of</strong> β = 12.5 and different asymmetries. Split times range<br />

from Ts= 0.5 to Ts = 1000. The results <strong>of</strong> the Bloch model (dotted line) and the<br />

full numerics (solid line) are in good agreement for times Ts > 20. For smaller<br />

times, the two mode approximation <strong>of</strong> the Bloch model fails, and excitations<br />

into higher modes take place. In the case <strong>of</strong> the high asymmetry, Fig. 3.8<br />

(d), the two mode approximation reaches its limit: for higher asymmetries,<br />

the adiabatic isolation to two levels begins to fail (see equation (2.57)). Then<br />

both lowest states <strong>of</strong> the system appear in the same well especially for small<br />

70


Chapter 3: The Asymmetric Double Well<br />

values <strong>of</strong> the splitting parameter β(t). This fact <strong>of</strong> adiabatic following is not<br />

to be confused with the decay into the ground state caused by a relaxation<br />

term in the Hamiltonian. Most <strong>of</strong> the results up to this point have also been<br />

published in [Sid06]; the following results are exclusive to this thesis.<br />

F<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

10 0<br />

10 1<br />

Figure 3.8: Decay <strong>of</strong> the filling factor F as a function <strong>of</strong> splitting time; Bloch<br />

model: dotted line; full numerical simulation: solid line [Whi04]. Asymmetries<br />

ˆVas = 0.01ˆx (a), 0.02ˆx (b), 0.05ˆx (c), 0.1ˆx (d). The failure <strong>of</strong> adiabatic isolation<br />

to two levels is apparent for short times. Part (d) shows how large asymmetries<br />

also limit the Bloch model.<br />

The behaviour <strong>of</strong> the filling factor F translates into a reduced fringe ampli-<br />

tude. Using only the Bloch model, the maximum fringe size, max(|〈φ1|||Ψ〉| 2 ),<br />

was calculated as a function <strong>of</strong> the same splitting and recombining time, for<br />

different asymmetries, for a maximum splitting <strong>of</strong> β = 12.5. The results for<br />

some asymmetries are shown in Fig. 3.9. The solid lines are the results <strong>of</strong> the<br />

Bloch model; the dashed lines are fits <strong>of</strong> exponential functions <strong>of</strong> the type<br />

d)<br />

10 2<br />

c)<br />

max(|〈φ1|||Ψ〉| 2 ) = C · e κTs,r (3.21)<br />

where κ is a negative decay rate, and the factor C can be used to determine<br />

71<br />

a)<br />

b)<br />

T s<br />

10 3


3.3. The Results <strong>of</strong> the Model<br />

max( |〈 Ψ| φ 1 〉| 2 )<br />

10 1<br />

10 0<br />

10 −1<br />

10 −2<br />

10 −3<br />

10 −4<br />

10<br />

0 100 200 300 400 500 600 700 800 900 1000<br />

−5<br />

Figure 3.9: Decay <strong>of</strong> the interferometric signal (on a logarithmic scale) versus<br />

splitting/recombining time for different asymmetries: ˆ Vas = 0.01ˆx (a), 0.02ˆx<br />

(b), 0.1ˆx (c). Solid line: Bloch model results; dashed line: fitted simple expo-<br />

nentials. Part (c) is at the edge <strong>of</strong> the applicability <strong>of</strong> the model.<br />

the time where the fit crosses the unit signal line. This time can be used as<br />

a characteristic timescale, which should not be exceeded in order to retain<br />

good visibility <strong>of</strong> the fringes and thus a working interferometer. As we can<br />

see from Figure 3.9, both the decay rate and the time for unity maximum<br />

signal are functions <strong>of</strong> the asymmetry Vas, where ˆ Vas = Vasˆx. If we plot these<br />

dependences, we obtain Figure 3.10.<br />

T s,r<br />

On the left <strong>of</strong> Figure 3.10, we see the time where the exponential fit crosses<br />

the unity signal threshold as a function <strong>of</strong> the asymmetry. The circles are the<br />

results from the Bloch model, and the line is a fitted function 4 . The results<br />

for Vas = 0 are included, which corresponds to no decay (κ = 0) and an<br />

infinitely large unity signal threshold time. We see that the dependence <strong>of</strong><br />

the threshold time on the applied asymmetry Vas is inverse to a high degree.<br />

We relate the imperfections to the problem <strong>of</strong> setting the interval <strong>of</strong> the data<br />

4 Time for maximum unity signal = (−0.006 + 0.382 · Vas) −1<br />

72<br />

(c)<br />

(b)<br />

(a)


unity maximum signal<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12<br />

V as<br />

Chapter 3: The Asymmetric Double Well<br />

0<br />

−0.01<br />

−0.02<br />

−0.03<br />

−0.04<br />

−0.05<br />

−0.06<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2<br />

Figure 3.10: Time for maximum unity signal (left) and decay rate κ (right) <strong>of</strong><br />

the signal as functions <strong>of</strong> the applied asymmetry Vas, with fits as explained in<br />

the text. The model seems limited to asymmetries that do not exceed 10% <strong>of</strong><br />

the trapping frequency during the splitting and merging. The triangle in the<br />

right figure is the result for an asymmetry that exceeds this limiting value.<br />

to which the exponential function is fitted. Especially for large asymmetries,<br />

when the model begins to break down, the signal shows some modulation (see<br />

Fig. 3.9 (c)) which makes a decision difficult on which points to include in the<br />

fitting process. Even though this was decided “by eye”, the data shows good<br />

agreement with the reciprocal behaviour.<br />

decay rate κ<br />

On the right <strong>of</strong> Fig. 3.10 we see the decay rate κ as a function <strong>of</strong> the<br />

asymmetry. The circles and the triangle are results from the fit to the Bloch<br />

model. The line is a linear fit 5 , enforcing the results for the symmetric trap<br />

but ignoring the two points with Vas ≥ 0.1. Here we see the breakdown <strong>of</strong> the<br />

model due to large asymmetries more clearly. The model fails as the ground<br />

and first excited state both occupy the same well for a non negligible time <strong>of</strong><br />

the splitting process. Thus, the deviation from the linear behaviour can be<br />

5 κ = −0.403 · Vas<br />

73<br />

V as


3.3. The Results <strong>of</strong> the Model<br />

used to check an experimental set-up (assuming it has a variable asymmetry)<br />

and determine the maximum asymmetry allowed during the splitting process.<br />

3.3.2 Addendum: Comparison with experimental data<br />

During the period amendments were being made to this thesis, the results<br />

<strong>of</strong> an atom interferometer using a BEC in an asymmetric double well were<br />

published in the group <strong>of</strong> W. Ketterle [Jo07]. The experiment described in<br />

that publication differs from the situation in our model only by the fact that<br />

the experiment used a BEC while our model is based on a single atom. One<br />

<strong>of</strong> the key results <strong>of</strong> the publication is an observed decay in the fringe contrast<br />

with increasing merging time. The authors state that the “dependence [...]<br />

on the recombination time allows [...] to speculate” about the causes <strong>of</strong> this<br />

decay. “Furthermore, it is not clear during what fraction <strong>of</strong> the ramp time <strong>of</strong><br />

the [...] recombination time the effective merging [...] occurs. [...] Another<br />

open question is what the rate <strong>of</strong> phase evolution is at the moment <strong>of</strong> the<br />

merger.” [Jo07].<br />

Even though the situations <strong>of</strong> the experiment (uses a BEC) and the model<br />

(which uses a single atom) are not identical using either a BEC or a single<br />

atom, the experimental parameters given in [Jo07] can be changed into the<br />

model’s harmonic oscillator units by use <strong>of</strong> Equation 3.4. The separation <strong>of</strong><br />

the wells <strong>of</strong> d ∼ 6 µm corresponds to a splitting parameter <strong>of</strong> β = 39.05. The<br />

experiment showed an interference that oscillated with a frequency <strong>of</strong> 500 Hz,<br />

corresponding to an asymmetry between the wells <strong>of</strong> Vasym = 0.0283 when fully<br />

split. The results <strong>of</strong> the experiment and the model can now be compared. This<br />

is shown in Fig. 3.11. The results <strong>of</strong> the model have to be seen as preliminary:<br />

the experiment’s far splitting <strong>of</strong> the wells leads to a large potential difference<br />

between the wells compared to the case for which the model was used before.<br />

74


enormalised interference fringe contrast<br />

1<br />

0.95<br />

0.9<br />

0.85<br />

0.8<br />

0.75<br />

0.7<br />

0.65<br />

Chapter 3: The Asymmetric Double Well<br />

0.6<br />

0 10 20 30 40 50 60 70<br />

merging time in ms<br />

Figure 3.11: Comparison <strong>of</strong> the experimental data <strong>of</strong> [Jo07] with the predic-<br />

tions <strong>of</strong> our model. The experimental data is shown by the squares. The<br />

results <strong>of</strong> the model (crosses interpolated by the thick line) are taken from<br />

three individual simulations for a splitting time <strong>of</strong> 75 ms with different tem-<br />

poral resolution. The thin line shows the results <strong>of</strong> one <strong>of</strong> these simulations.<br />

The reason for the oscillation is given in the text.<br />

Thus, the temporal resolution in the simulations was chosen to be too low,<br />

and a stroboscopic effect caused by the difference <strong>of</strong> the atomic interference<br />

frequency and the sampling frequency <strong>of</strong> the calculation appears in the calcula-<br />

tions. An attempt was made to overcome this effect by increasing the sampling<br />

frequency. Unfortunately, even after three runs with increasing time needed<br />

for the calculations, the beating in the calculations still appeared. The figure<br />

shows the envelope <strong>of</strong> the results <strong>of</strong> the three calculations with different tem-<br />

poral resolution (crosses and thick line). The result <strong>of</strong> one <strong>of</strong> these simulations<br />

(thin line in the figure) shows the stroboscopic effect. The step-like structures<br />

at times <strong>of</strong> ∼ 25 ms and ∼ 45 ms are clearly caused by the still insufficient<br />

temporal resolution. The agreement between the experimental results and the<br />

model is remarkable, especially since the experiment uses a BEC and the model<br />

75


3.3. The Results <strong>of</strong> the Model<br />

uses a single atom. The remaining small deviations may be explained by this<br />

difference. Because <strong>of</strong> this good agreement, the model presented here can be<br />

used to answer the above questions raised in the publication [Jo07].<br />

3.3.3 CARP: a Coherent Adiabatic Readout Process for<br />

double well atom interferometers<br />

The results presented above <strong>of</strong>fer an intriguing new possibility to read out<br />

an atom interferometer based on a double well. While for the interferometric<br />

process an adiabatic splitting <strong>of</strong> the asymmetric well is unwanted as it reduces<br />

the signal, it provides a simple means to read out the energy state <strong>of</strong> the particle<br />

in the trap. Equation (3.19) for the split trap shows that the occupation<br />

probability for the first excited state is only dependent on the Bloch vector<br />

component Yz. This was already identified as a measure <strong>of</strong> the imbalance <strong>of</strong><br />

the atomic population between the wells.<br />

Thus, if we add an additional phase <strong>of</strong> adiabatic splitting to the interfero-<br />

meter which so far consisted <strong>of</strong> a non adiabatic splitting, a phase evolution<br />

and a non adiabatic recombination phase, we can read out the interferometer<br />

by simply measuring the population <strong>of</strong> one <strong>of</strong> the wells. For this phase it is<br />

advisable to apply the highest possible asymmetry that does not lead to the<br />

two lowest states being located in the same well in order to reduce the required<br />

time for adiabaticity.<br />

If we plot the energy spectrum <strong>of</strong> the two lowest, uncoupled states (as in<br />

Fig. 3.4), we obtain a graph that resembles a fish for the overall process. This<br />

is shown schematically in Fig. 3.12, where the time is progressing along the<br />

horizontal axis and the energy <strong>of</strong> the states is in the vertical axis.<br />

A similar read out process that relies on the relative atom numbers in each<br />

well after adiabatic splitting <strong>of</strong> the wells has recently been used in our group to<br />

76


Energy<br />

Chapter 3: The Asymmetric Double Well<br />

Interferometer Read−out<br />

Figure 3.12: Schematic spectral flow, like in Fig. 3.4, showing the two lowest,<br />

uncoupled states for an interferometer with a simple read-out possibility: the<br />

CARP. The times for the actual interferometer and the read-out process are<br />

not to scale.<br />

measure the double well asymmetry <strong>of</strong> a trapped BEC in a permanent magnet<br />

film trap [Hal07b]. Here an experimental cycle time <strong>of</strong> 30 s and a single shot<br />

sensitivity <strong>of</strong> 16 Hz or δg<br />

g = 2 · 10−4 was achieved. The main source <strong>of</strong> noise<br />

has been identified as the shot-to-shot number fluctuations in the condensate.<br />

3.4 Summary<br />

In this chapter it was shown how the two level Bloch equations can be applied<br />

to an asymmetric double well system. This was done with an emphasis on the<br />

application as a single atom interferometer, and the restrictions that an asym-<br />

metry imposes on the splitting and merging time scales were shown. The main<br />

advantages <strong>of</strong> this model are that the physics becomes easily understandable:<br />

any double well interferometer works by Larmor precession. Only a certain<br />

asymmetry can be applied, so that during the processes the ground and first<br />

77<br />

Time


3.4. Summary<br />

excited states do not significantly occupy the same well. This is one limit to<br />

the model, but it also promotes the use <strong>of</strong> traps with high trapping frequencies<br />

(miniaturized magnetic or optical traps) as the increased steepness allows one<br />

to deal with higher asymmetries. The only disadvantage <strong>of</strong> this model is that<br />

higher modes are neglected. The model fails for processes that non adiabati-<br />

cally couple these higher levels to the lowest two. For longer times, the results<br />

<strong>of</strong> the model agree with the results <strong>of</strong> the full numerical integration.<br />

The most important result <strong>of</strong> this model is the fact that, for a real double<br />

well interferometer, an optimum time for splitting and merging the trap exists.<br />

A too short a time will lead to non adiabatic transitions from the ground state<br />

to the second excited state or higher states [H¨01c]. A too long a time will be<br />

adiabatic in the transition between the ground state and the first excited state<br />

and will inevitably lead to a localisation <strong>of</strong> the atom. This adiabaticity can be<br />

useful in the read out <strong>of</strong> the interferometer, as it translates the population <strong>of</strong><br />

the energy states into a spatially distinguishable population, as pointed out in<br />

section 3.3.3.<br />

To extract information from a single atom interferometer, it is necessary to<br />

have a statistical sample <strong>of</strong> identical experiments. This can be done either by<br />

parallel experiments or by repeating the experiment under identical conditions,<br />

or by a combination <strong>of</strong> both. It is possible to create arrays <strong>of</strong> several tens <strong>of</strong><br />

atom traps [Dum02b]. Let us assume that we want at least 100 measurements,<br />

which means we need to run the overall experiment 10 times. If we further<br />

assume we have double well potentials with a trap frequency <strong>of</strong> ω0<br />

2π<br />

= 1 kHz,<br />

we have a characteristic time <strong>of</strong> Tchar ≈ 0.16 ms. From Figure 3.9 we see that<br />

a linear and not-optimised splitting (and recombination) should take at least<br />

10 · Tchar. The holding phase then is variable, while for the proposed read-out<br />

we are limited by the maximum <strong>of</strong> the asymmetry that can be applied. Figure<br />

3.9 proposes times <strong>of</strong> the order <strong>of</strong> magnitude 100·Tchar for a well localised atom.<br />

78


Chapter 3: The Asymmetric Double Well<br />

So, as a rough estimation, one can say that one run <strong>of</strong> the interferometer should<br />

take about T = 150 · Tchar or T ≈ 24 ms for a 1 kHz trap. This corresponds to<br />

a repetition rate <strong>of</strong> about 40 Hz ignoring preparation times. These preparation<br />

times will be the dominant factor in the usefulness <strong>of</strong> the double well potential<br />

as an interferometer to measure spatially or temporally varying external fields<br />

[Hal07b]. To test fundamental quantum mechanics or in the application to<br />

quantum computing the external conditions are either under our control or not<br />

<strong>of</strong> interest as a measurable. Here the demand for a fast repetition <strong>of</strong> identical<br />

interferometers is not as strong as if used as a measuring device for external<br />

fields. The application <strong>of</strong> the double well single atom interferometer for the<br />

purposes <strong>of</strong> quantum computing or tests <strong>of</strong> quantum mechanics is thus more<br />

likely. For atom interferometers to measure imbalances between the wells, it<br />

is more likely that trapped BECs are used (see [Hal07b, Jo07] and references<br />

therein).<br />

The model here was kept simple. It can easily be applied to a plethora<br />

<strong>of</strong> problems, and its simplicity allows easy adaption to these problems. For<br />

example, a proposed atomic double well interferometer on a chip can lead to<br />

different trap frequencies when split [Hin01]. These different trap frequencies<br />

can be translated into an asymmetry Vasym and the model here can be used.<br />

Another direct application <strong>of</strong> this model is possible in double well problems<br />

in solid state physics, like [Det04, Hu00, Hol04], and to macroscopic double<br />

wells in atom optics [Tho02]. Some quantum computing proposals also rely<br />

on double well potentials, both on the solid state side [Hol04] and the atom<br />

optics side [Cal00, Mom03, Dum02b, Buc02]. A similar model was used to de-<br />

scribe STIRAP-like processes and other properties <strong>of</strong> three level systems in a<br />

triple well system [Eck04]. The model presented here neglects any interactions<br />

between atoms. Interactions enhance the localising effect <strong>of</strong> the asymmetry,<br />

as they would add a non-linear term to the Schrödinger equation which is de-<br />

79


3.4. Summary<br />

pendent on the atomic wavefunction or atomic density in the traps, similar<br />

to the Gross-Pitaevskii equation (2.55). To include such a term in the model<br />

would require a complex model, as the two mode approximation is question-<br />

able for a N atom system where the first excited state is N-fold degenerate.<br />

Even though these interactions were neglected, the read out process presented<br />

inspired a way to read out asymmetry measurements <strong>of</strong> a double well with a<br />

BEC [Hal07b] with high accuracy.<br />

The only things that have to be changed to the specific problem are the<br />

temporal flows <strong>of</strong> the matrix elements ∆0 and Vasym, according to the spectral<br />

flow <strong>of</strong> the energy states and the ramps that are used to split or merge the<br />

traps. These are the parameters that take the roles <strong>of</strong> the Rabi frequency and<br />

the detuning in the more common form <strong>of</strong> the universal two level problem.<br />

This makes it easy to compare our model to experimental results <strong>of</strong> the decay<br />

<strong>of</strong> fringe contrast in an atom interferometer that utilizes BECs [Jo07]. Even<br />

with the obvious difference between a single atom and a BEC the comparison<br />

shows good agreement <strong>of</strong> the model.<br />

80


Chapter 4<br />

The Permanent Magnetic Chip<br />

Experiment: Apparatus<br />

4.1 Overview<br />

This chapter describes the permanent magnetic chip experiment that is located<br />

at <strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong>, Australia. The major goal was to<br />

achieve Bose-Einstein Condensation <strong>of</strong> 87 Rb atoms held in a magnetic trapping<br />

field, where this field was produced by a permanent magnetic microstructure<br />

on an atom chip. To achieve this, atoms were collected and optically precooled<br />

by the standard methods <strong>of</strong> laser cooling and trapping. For magnetic trapping<br />

and evaporation <strong>of</strong> the atomic ensemble to quantum degeneracy an ultra-high<br />

vacuum (UHV) system was set up, containing the atom chip that was produced<br />

in house. The chip was designed to trap 87 Rb atoms magnetically with a<br />

potential that permits fast and efficient evaporative cooling towards quantum<br />

degeneracy. This experiment was newly set up as a part <strong>of</strong> this PhD project<br />

and it will be described and explained in detail in the following sections.<br />

Permanent magnetic structures have a number <strong>of</strong> potential advantages over<br />

81


4.1. Overview<br />

current-carrying wires, but they also have a handicap: their fields are perma-<br />

nent, and only in some situations is it possible to fully cancel the effect <strong>of</strong> these<br />

fields when needed. This makes some standard procedures like temperature<br />

measurements by TOF more difficult or even impossible. On the other hand<br />

permanent magnets allow high effective currents, without any ohmic heating<br />

and without any chance <strong>of</strong> breaking the wire. This is true especially for small<br />

structures, which would be realisable only by very thin wires. Permanent<br />

magnets exhibit no current noise and reduced thermally induced current fluc-<br />

tuations in the conducting materials, which can cause spin-flip losses [Jon03].<br />

This can lead to extraordinary low heating rates <strong>of</strong> the atoms in the trap,<br />

which for wire-based traps can not be easily realised. With the magnetic film<br />

trap presented here, heating rates as low as 3 nK/s were observed, while the<br />

current-carrying wire <strong>of</strong> the set-up presented here had a heating rate <strong>of</strong> 270<br />

nK/s [Hal07b, Hal07a]. Another remarkable advantage <strong>of</strong> permanent magnetic<br />

structures is the ability to create true closed loops <strong>of</strong> the effective currents,<br />

whereas for a current-carrying wire, there always has to be a connection to<br />

the power supply. This allows the creation <strong>of</strong> potentials with a magnetic film<br />

that are not possible with current-carrying wires. Here permanent magnetic<br />

structures are advantageous.<br />

Section 4.2 contains a description <strong>of</strong> the optical system, starting with the<br />

laser systems that are used to cool, trap, manipulate and image the atoms.<br />

This is followed by a description <strong>of</strong> the UHV apparatus and the procedure<br />

to reach a sufficiently high vacuum. The third and fourth parts describe the<br />

sources <strong>of</strong> the magnetic fields needed to trap the atoms. The third section<br />

contains information on the magnetic field coils required for the complete op-<br />

eration <strong>of</strong> the experiment. This is followed by the section describing the central<br />

piece <strong>of</strong> the experiment, the permanent magnetic atom chip. Here the design<br />

<strong>of</strong> the chip is introduced. This also covers the current-carrying wire structure<br />

82


Chapter 4: The Permanent Magnetic Chip Experiment: Apparatus<br />

and the permanent magnetic structure. A standard PC with two digital and<br />

analogue I/O cards is used to control the experiment. Via connector boards [E<br />

21] the AOMs (RF frequency and switching), the quadrupole and <strong>of</strong>fset coils<br />

(switched by MosFets), the Rb-dispensers (switched by MosFets), the shutters<br />

and the CCD camera can all be addressed. This is done using the LabView<br />

s<strong>of</strong>tware package and a home-coded program.<br />

4.2 The Laser Systems<br />

This experiment uses three laser systems to provide light <strong>of</strong> the appropriate<br />

frequencies, powers, linewidths, polarisations and beam shapes, necessary to<br />

trap, cool, and engineer the atomic states and to image the ensemble. Each<br />

system is based on diode laser technology with a wavelength <strong>of</strong> 780 nm and<br />

resonant with the D2 line <strong>of</strong> 87 Rb. These laser diodes are readily available<br />

since they are mass produced for CD players and recorders. The laser systems<br />

are labelled according to their use in the experiment. They are:<br />

• The main laser: a laser system based on a commercial tapered amplifier<br />

system. It is used to produce the light needed for trapping and cooling<br />

the atoms in the MOT and the light which is used for the absorption<br />

imaging, resonant with the |F = 2〉 → |F ′ = 3〉 transition <strong>of</strong> the D2 line<br />

(see Figure C.1 in the appendix C).<br />

• The repumping laser: a commercial diode laser system. As the main<br />

laser also excites some atoms into the |F ′ = 2〉 state, from where they<br />

can decay into the |F = 1〉 ground state, the repumping light is tuned<br />

to the |F = 1〉 → |F ′ = 2〉 transition. The repumping laser recycles any<br />

population in the unwanted atomic substate |F = 1〉 back to the cooling<br />

cycle.<br />

83


4.2. The Laser Systems<br />

• The optical pumping laser: a home built, grating stabilized laser<br />

system, similar to the repumping laser. Its light is used to optically<br />

pump the atoms into the state |F = 2, mF = 2〉 before the transfer from<br />

the MOT to the magnetic trap.<br />

All laser systems are located on a separate optical table, and their light is<br />

brought to the experiment table by optical fibres. To reduce the amount <strong>of</strong><br />

stray light that could interfere with the experiment, the lasers are shielded by<br />

an opaque PVC box.<br />

4.2.1 The main laser<br />

The main light source is a commercial tapered amplifier [E 1]. Here a low<br />

power extended cavity laser diode, which is frequency stabilised by controling<br />

temperature, current and the grating, is used to seed a high power tapered<br />

amplifier. This produces a single tranverse mode, 500 mW output beam which<br />

can be frequency tuned by 10 GHz across the D2 spectrum <strong>of</strong> Rb without<br />

mode hops.<br />

Frequency stabilisation follows the dither-free polarisation spectroscopy<br />

scheme presented in [Pea02, Pet03], which has the big advantage <strong>of</strong> having<br />

only one zero crossing in the error signal. This makes jumps to other frequen-<br />

cies impossible, and is in this way superior to standard frequency-modulated<br />

peak-lock techniques. On the other hand, it is not a peak-locking technique<br />

and thus relies on high stability <strong>of</strong> laser powers and the used electronic devices<br />

to avoid frequency drifts <strong>of</strong> the locking point.<br />

The set-up is depicted in Fig. 4.1. Part <strong>of</strong> the light <strong>of</strong> the diode laser in Lit-<br />

trow arrangement (the master laser) is split <strong>of</strong>f after the optical diode. This<br />

part double-passes an acousto-optical modulator (AOM) [E 2], with a centre<br />

84


xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

AOM<br />

λ/2<br />

Chapter 4: The Permanent Magnetic Chip Experiment: Apparatus<br />

400<br />

λ/4<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

PBS<br />

400<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

λ/2<br />

λ/4<br />

Rb cell<br />

TA 100<br />

PD<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxxx<br />

xxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

PD<br />

PBS<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

λ/2<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

400<br />

AOM<br />

Figure 4.1: The main laser system: tapered amplifier, spectroscopy <strong>of</strong> the<br />

master laser and path <strong>of</strong> the amplified light. In this and all following sketches<br />

lenses are depicted by double arrows. Other labels are explained in the text.<br />

frequency <strong>of</strong> 80 MHz and a bandwidth <strong>of</strong> 30 MHz. The AOM 1 is used to<br />

detune the light for the different tasks and is driven with frequencies ν1 ≈ 63<br />

MHz for cooling or ν1 ≈ 55 MHz for absorption imaging. A polarising beam<br />

splitter cube (PBS) is used to deflect the detuned light, that had its polari-<br />

sation rotated by 90 o by double-passing a quarter-wave plate (λ/4), into the<br />

spectroscopy branch. There it is split again: most <strong>of</strong> the light passes another<br />

λ/4-plate and irradiates the 87 Rb atoms in a glass cell (Rb cell) as circular po-<br />

larised pumping light. This pump beam induces a birefringence in the atomic<br />

vapour. It is interrogated by the linearly polarised remaining amount <strong>of</strong> light,<br />

which first passes a half-wave plate (λ/2) and then the Rb cell. The probe<br />

400<br />

1 The AOM output frequency fAOM was calibrated to fAOM = (525.87 · U + 1279.97) 1/2<br />

MHz as a function <strong>of</strong> applied voltage U.<br />

85


4.2. The Laser Systems<br />

light is then split at another PBS (stray light is blocked by a pinhole in front<br />

<strong>of</strong> the PBS) and collected by two photodiodes (PD).<br />

The difference <strong>of</strong> the two signals is taken electronically and fed into a PID<br />

regulator [E 3], which in turn drives the piezo-electric translators (PZT) <strong>of</strong> the<br />

grating and stabilizes the master laser frequency to the |F = 2〉 → |F ′ = 3〉<br />

transition. The internal feed forward from the PZT-driver to the driver <strong>of</strong><br />

the laser diode current is set to zero. Instead, for a fast correction, the error<br />

signal is also fed into the modulation input <strong>of</strong> the current driver. The resulting<br />

reduction in the noise is shown by the error signals and the power spectra <strong>of</strong><br />

these signals in Fig. 4.2. The modulation <strong>of</strong> the current as the fast branch <strong>of</strong><br />

the frequency stabilisation shows a significant effect in the acoustic frequencies<br />

up to 1.5 kHz.<br />

error signal / V<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

−0.01<br />

−0.02<br />

−0.03<br />

−0.04<br />

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04<br />

time / s<br />

(a)<br />

noise Amplitude, linear a.u.<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

0 200 400 600 800 1000 1200 1400 1600 1800 2000<br />

frequency / Hz<br />

Figure 4.2: Closed loop error signals (a) and power spectra (b) <strong>of</strong> the main laser<br />

without (dashed line) and with (solid line) current as fast correcting element.<br />

To reduce drifts in the signal, precautions have been taken. After the de-<br />

flection into the spectroscopy part, the light instantly passes a second, perpen-<br />

dicular PBS. This ensures that we have well linearly polarised light. Second,<br />

lenses are placed in front <strong>of</strong> the PDs. With their focussing they correct for<br />

inevitable slight misalignments <strong>of</strong> the lenses in front <strong>of</strong> and after the AOM,<br />

which in turn lead to a frequency dependent position <strong>of</strong> the beam.<br />

86<br />

(b)


Chapter 4: The Permanent Magnetic Chip Experiment: Apparatus<br />

The master laser is used to seed a tapered amplifier. The amplified light<br />

passes another AOM 2 [E 4] which runs at a frequency <strong>of</strong> ν2 ≈ 110 MHz. This<br />

AOM has its centre frequency at 110 MHz and a bandwidth <strong>of</strong> 50 MHz. The<br />

first order <strong>of</strong> the light is coupled into an optical fibre [E 5] [E 6], while the zeroth<br />

order is lost. The optical fibre cleans the mode <strong>of</strong> the light and transfers it to<br />

the experiment table. As the coupling efficiency into the fibre is very sensitive<br />

to the position and angle <strong>of</strong> the incoming light, the AOM is used as a fast<br />

switch to turn the light on and <strong>of</strong>f. The two AOMs are set up so that they<br />

work against each other in their frequency shifts. The light that is used in the<br />

experiment thus has an overall detuning <strong>of</strong> 2ν1 − ν2 against the locking point,<br />

which is close to but not exactly equal to the |F = 2〉 → |F ′ = 3〉 frequency, as<br />

the polarisation locking scheme here does not lock to the peak <strong>of</strong> the absorption<br />

like it does with the conventional modulated lock-in techniques. The overall<br />

spectral width <strong>of</strong> the frequency stabilized laser light has been estimated to<br />

≤ 1.6 MHz using a scanning etalon.<br />

4.2.2 The repumping laser<br />

The light source for closing the |F = 2〉 → |F ′ = 3〉 cooling cycle is produced<br />

by a commercial diode laser system [E 7], which is a grating stabilised laser<br />

in Littrow configuration. The output wavelength is stabilised by a Doppler-<br />

free saturation spectroscopy using a peak-lock lock-in technique (see Fig. 4.3).<br />

After passing the optical diode, part <strong>of</strong> the light is split into the spectroscopy<br />

arm. Here another beamsplitter reflects a small amount <strong>of</strong> the light through<br />

a 87 Rb-cell on a photodiode. The light transmitted by this beamsplitter is<br />

counterpropagating this probe beam in the cell as the saturation light. The<br />

wavelength <strong>of</strong> the light is modulated via the PZT <strong>of</strong> the grating, and the<br />

2 This AOM was calibrated to fAOM = (997.09 · U + 2332.31) 1/2 MHz.<br />

87


4.2. The Laser Systems<br />

photodiode’s signal feeds the lock-in electronics [E 8]. The resulting error signal<br />

is used to stabilise the laser directly to the |F = 1〉 → |F ′ = 2〉 transition.<br />

TUI<br />

DL 100<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

PD<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

Rb cell<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

λ/2<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

Figure 4.3: The repumping laser: extended cavity diode laser system with<br />

spectroscopy<br />

Most <strong>of</strong> the light <strong>of</strong> the diode laser is not used for the spectroscopy and is<br />

coupled into an optical fibre, [E 5], and transported to the experiment table.<br />

As there is no AOM in this set-up there is no possibility to detune or quickly<br />

switch the light. The switching is done by mechanical shutters in other parts<br />

<strong>of</strong> the set-up (see section 4.2.4).<br />

4.2.3 The optical pumping laser<br />

A home built diode laser is the light source for the optical pumping. This<br />

laser is an extended cavity diode laser in the Littrow design, similar to the<br />

repumping laser above. It is current and temperature stabilised [E 9]. Its<br />

frequency stabilisation is a polarisation spectroscopy scheme like the one used<br />

for the main laser in section 4.2.1. The set-up is depicted in Fig. 4.4.<br />

The spectroscopy works in the same way as the spectroscopy <strong>of</strong> the main<br />

laser, with a circularly polarised pump beam and a linearly polarised probe<br />

88


Chapter 4: The Permanent Magnetic Chip Experiment: Apparatus<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

PBS<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx xxxxxxxxxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

λ/2<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

PBS<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

λ/2<br />

Rb cell<br />

AOM<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

Figure 4.4: The optical pumping laser: extended cavity diode laser, spec-<br />

200<br />

troscopy and an AOM to detune and switch the light.<br />

beam which is then split onto two photodiodes. The difference <strong>of</strong> the signals<br />

<strong>of</strong> the PDs is the error signal which is fed into a home built PID regulator.<br />

The resulting output signal is fed to the PZT <strong>of</strong> the diode laser to stabilise<br />

the laser to the frequency <strong>of</strong> the |F = 2〉 → |F ′ = 3〉 transition. Unlike the<br />

PI-controling electronics for the tapered amplifier, we have no fast branch in<br />

the correction signal working with the current <strong>of</strong> the laser diode.<br />

The light for the experiment is deflected by a PBS. It then double-passes an<br />

AOM, which is used to detune the light to 50 MHz above the |F = 2〉 → |F ′ =<br />

2〉 transition. Switching the light is carried out by turning the AOM <strong>of</strong>f while<br />

still applying the frequency, so that the undiffracted light is dumped. This<br />

allows us to switch much faster than the irradiation time <strong>of</strong> the atoms required<br />

to optically pump the atoms, which is <strong>of</strong> the order <strong>of</strong> a few 100 µs. After<br />

89<br />

PD<br />

PD<br />

λ/4<br />

200<br />

λ/4<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx


4.2. The Laser Systems<br />

double-passing a λ/4-plate, the light passes the PBS without deflection and<br />

is coupled into an optical fibre [E 5] and transferred to the experiment. This<br />

set-up allows us to tune the frequency <strong>of</strong> the light to the |F = 2〉 → |F ′ = 2〉<br />

transition frequency and with sufficient fast switching times.<br />

4.2.4 The optical paths on the experiment table<br />

After the light has been transferred to the table with the vacuum chamber,<br />

some beams <strong>of</strong> the different sources have to be overlapped and some have to<br />

be split. For a working MOT, the light <strong>of</strong> the main laser and the repumping<br />

laser have to be overlapped. Some <strong>of</strong> the light <strong>of</strong> the main laser has to be split<br />

<strong>of</strong>f before the addition <strong>of</strong> the repumping light for the absorption imaging. How<br />

this is achieved is depicted in Fig. 4.5. Some beam shaping occurs in this part,<br />

and there are mechanical shutters to fully block the light when unwanted. The<br />

whole set-up is enclosed in a PVC box to reduce the amount <strong>of</strong> stray light that<br />

could disturb the experiment.<br />

The light for the MOT is prepared at a PBS, where the light <strong>of</strong> the main<br />

laser (cooler) and the light <strong>of</strong> the repumping laser (repumper) are added after<br />

both have passed a λ/2-plate. This light is then expanded by a factor <strong>of</strong> 10<br />

in a telescope. A shutter [E 10] is placed in the focus <strong>of</strong> this telescope to stop<br />

the light when trapping magnetically. This light proceeds towards the MOT.<br />

The light for the absorption imaging is split <strong>of</strong>f the cooler light before it is<br />

combined with the repumper light. A rotatable PBS allows control over the<br />

intensity <strong>of</strong> the absorption light for a good signal to noise ratio in the image <strong>of</strong><br />

the atoms. A λ/2-plate and a second PBS are used to redefine the polarisation<br />

axis. The half-wave plate is not necessarily needed, but adds robustness to the<br />

system by always allowing us to work in a regime <strong>of</strong> Malus’ law which is<br />

insensitive to slight polarisation changes and mismatches. The beam is then<br />

90


Chapter 4: The Permanent Magnetic Chip Experiment: Apparatus<br />

1000<br />

absorption<br />

light<br />

50<br />

PBS<br />

λ/2<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

shutter<br />

rotatable<br />

PBS<br />

opt. pumping<br />

10<br />

PBS<br />

10<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

λ/2<br />

opt. pumper<br />

shutter<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

λ/2<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

cooler<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

20<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

repumper<br />

MOT light<br />

PBS<br />

500<br />

shutter<br />

50<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx10<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

Figure 4.5: The optical paths to overlap the different beams. The set-up is on<br />

the same optical table as the vacuum chamber and is enclosed in a PVC box.<br />

expanded by a factor <strong>of</strong> 20. This is done to achieve a flat intensity distribution<br />

<strong>of</strong> the light at the location <strong>of</strong> the atoms. A shutter [E 10] is placed in the focus<br />

<strong>of</strong> this telescope and blocks the light until the actual pictures are taken. The<br />

shutter starts to open at 1.7 ms and finishes 2.5 ms after the opening signal<br />

arrives. It begins to close 1.5 ms and is fully closed 2.3 ms after the signal. The<br />

times were determined using a fast photodiode [E 11] directly after the shutter,<br />

with the beam appropriately weakened. Both opening and closing signal reach<br />

the 3 · σ-line, which is calculated from the datapoints <strong>of</strong> the last 0.4 ms, after<br />

2.5 ms. The results for the shutter <strong>of</strong> the MOT light are consistent with this.<br />

91


4.2. The Laser Systems<br />

To the light <strong>of</strong> the optical pumping laser we add a portion <strong>of</strong> the repumping<br />

light by a PBS. This light is then expanded in a telescope, before it continues<br />

its way to the chamber.<br />

The final manipulation <strong>of</strong> the MOT light occurs outside the PVC enclosure.<br />

It is shown in Fig. 4.6, together with the optical paths <strong>of</strong> the absorption and<br />

the imaging light. These beams are sent to the sides <strong>of</strong> the chamber to an<br />

elevator stage. The MOT light is split into 4 beams <strong>of</strong> equal power (about 25<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

elevator<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

MOT<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

100 50 λ/4<br />

100<br />

50<br />

λ/4<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxxx<br />

xxxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

100<br />

50<br />

λ/4<br />

PBS λ/2<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

100<br />

50<br />

λ/4<br />

PBS<br />

λ/2<br />

PBS<br />

λ/2<br />

absorption<br />

light opt. pumping<br />

MOT light<br />

Figure 4.6: The optical paths close to the chamber. This part <strong>of</strong> the set-up is<br />

following the part <strong>of</strong> Fig. 4.5, but outside the PVC enclosute.<br />

mW and diameter 3.5 cm) using a combination <strong>of</strong> λ/2-plates and a PBS. Each<br />

beam then passes a λ/4-plate to be circularly polarised and is expanded by a<br />

92


Chapter 4: The Permanent Magnetic Chip Experiment: Apparatus<br />

telescope with f1 = −50 mm, f2 = 100 mm. The beams are the two pairs <strong>of</strong><br />

counterpropagating light needed for a surface MOT.<br />

Both the optical pumping light and the light for the absorption imaging<br />

are elevated to the position <strong>of</strong> the trap. Their polarisations are defined by a<br />

PBS with following λ/4-plates (see Fig. 4.7). The optical pumping beam is<br />

slightly tilted towards the axis given by the absorption beam and the magnetic<br />

fields, and retroreflected on itself, while keeping its handedness after double-<br />

passing a second λ/4-plate. The absorption imaging light passes through the<br />

atomic cloud and is imaged on a CCD camera [E 12]. The imaging optics <strong>of</strong><br />

the camera consists <strong>of</strong> an achromatic doublet with f = 100 mm and allows a<br />

resolution <strong>of</strong> 9.5 × 9.5 µm 2 per pixel. The camera itself has a pixel size <strong>of</strong> 9 × 9<br />

µm 2 .<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxx<br />

xxxxxxxxxx<br />

absorption<br />

light<br />

PBS<br />

PD<br />

λ/4<br />

λ/4<br />

PBS<br />

opt. pumping<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

20<br />

atomic<br />

cloud<br />

λ/4<br />

100<br />

xxxx<br />

xxxx<br />

xxxx<br />

xxxx<br />

xxxx<br />

xxxx<br />

xxxx<br />

xxxx<br />

xxxx<br />

xxxx<br />

xxxx<br />

CCD<br />

camera<br />

Figure 4.7: The optical pumping and the absorption light at the chamber. The<br />

light is brought to height by an ‘elevator’ (see top <strong>of</strong> Fig. 4.6).<br />

The fluorescence <strong>of</strong> the atoms, when trapped in the MOT, is focused onto<br />

a photodiode. A confocal lens (f = 20 mm) is placed at a distance <strong>of</strong> twice its<br />

focal length from both the atoms and the photodiode. An aperture in front <strong>of</strong><br />

the photodiode filters out unwanted scattered light. Infrared sensitive cameras<br />

93


4.3. The Vacuum System<br />

are mounted so that the atoms can be directly monitored from the bottom and<br />

at an angle from the side, for quick online position control.<br />

4.3 The Vacuum System<br />

To magnetically trap atoms and to have trap lifetimes that allow evaporative<br />

cooling towards a BEC, collisions with untrapped and hot background atoms<br />

have to be avoided. This is done by trapping the atoms in an Ultra High<br />

Vacuum (UHV) with a pressure <strong>of</strong> the order 10 −10 − 10 −11 Torr. To achieve<br />

such a vacuum, care has to be taken in the design <strong>of</strong> the experiment chamber<br />

and the pumping down <strong>of</strong> the vacuum. The chamber has to be baked out for<br />

several days while pumping down. In the following the chamber itself will be<br />

introduced. After that the pumps and their arrangement will be described.<br />

This section finishes with the procedure for baking and pumping to reach the<br />

desired UHV.<br />

4.3.1 The experiment chamber<br />

The chamber is shown in Fig. 4.8. It was designed by A. Sidorov [Sid02a], and<br />

is made <strong>of</strong> electropolished steel (304SS), which has a magnetic permeability<br />

close to the vacuum permeability, µ304SS < 1.05 × µ0. The overall dimensions<br />

<strong>of</strong> the chamber are 9.6 × 9.6 × 4.9 cubic inches, the diagonal extension is 11”,<br />

and the inner diameter is 6”.<br />

The sides allow attachment <strong>of</strong> 8” (outer diameter) vacuum components<br />

at the tapped holes. We have attached windows with anti-reflection (AR)<br />

coating on the outside. Two <strong>of</strong> the MOT beams are passed through these<br />

windows. In the horizontal and vertical plane, there are 4.5” flanges with<br />

tapped holes. AR coated viewports are mounted on the horizontal flanges, for<br />

the optical pumping and the absorption imaging. Through the top opening,<br />

94


9.6"<br />

Chapter 4: The Permanent Magnetic Chip Experiment: Apparatus<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx xxxxxxxx<br />

xxxxxxxx xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

4.5"<br />

11"<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx<br />

xxxxxxxx xxxxxxx<br />

xxxxxxxx xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx xxxxxxx<br />

xxxxxxx xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

2.75"<br />

4.9"<br />

x<br />

x<br />

x xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxxx x<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxxx x<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxxx x<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxxx x<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxxx x<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxxx x<br />

xxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxxx<br />

xxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx x<br />

xxxxxxxxxxxxxxx<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x 6" x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx x<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxxxx<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxxx<br />

xxx<br />

xxxxxxxxxxxxxx<br />

xxx<br />

Figure 4.8: The experiment chamber, front and side view. Dimensions are<br />

given in the imperial system.<br />

the chip and its holder are inserted and fixed. The bottom flange leads to<br />

the vacuum pumps. The four 2.75” flanges with clear holes at 45 o have the<br />

following attachments. The two lower flanges hold AR coated viewports which<br />

are used for the second pair <strong>of</strong> MOT beams. One <strong>of</strong> the upper flanges holds<br />

a cold cathode gauge [E 13] for monitoring the pressure. At the other upper<br />

flange an electrical feedthrough (ceramaseal 8962-06-CF, 12 pin, copper, 55A)<br />

is mounted which provides electrical connections to the Rb dispensers and the<br />

chip wires.<br />

4.3.2 Vacuum pumps and further vacuum system<br />

The chamber is mounted on top <strong>of</strong> the vacuum system that includes the vac-<br />

uum pumps. The tubes <strong>of</strong> the vacuum system are commercially available tubes<br />

95<br />

8"


4.3. The Vacuum System<br />

produced from stainless steel. The main axis <strong>of</strong> this system is parallel to the<br />

short axis <strong>of</strong> the chamber (see Fig. 4.9).<br />

Ti: sublimation<br />

pump<br />

experiment<br />

chamber<br />

xxxxxxx<br />

x<br />

xxxxxxx<br />

x<br />

xxxxxxx<br />

x<br />

xxxxxxx<br />

xxxx x<br />

xxxxxxx<br />

xxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxx<br />

xxxxxx<br />

xxxx<br />

xxxxxx xx<br />

xxxxxx xx<br />

xxxxxx xx<br />

xxxxxx xx<br />

xxxxxxx xx<br />

xxxxxxx xx<br />

xxxxxxx xx<br />

xxxxxxx<br />

xxxxx xx<br />

xxxxxxx<br />

xxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxx<br />

xxxxxxx<br />

xxxxx<br />

xxxxxxx xx<br />

xxxxxxx xx<br />

xxxxxxx xx<br />

xxxxxxx xx<br />

valve<br />

Ion pump<br />

Turbo<br />

pump<br />

Figure 4.9: The vacuum system with its pumps. The experiment chamber is<br />

mounted on top <strong>of</strong> this set-up. The diaphragm pump (see text) is omitted.<br />

During the pumping down stage, the valve to the turbo pump [E 14] is<br />

opened. This pump has a pumping capacity <strong>of</strong> 50 l/s. A diaphragm pump<br />

[E 15] with ≈ 0.9 l/s is connected to it to produce a sufficiently low pressure<br />

for the turbo pump to operate. Once the pumping down is completed and<br />

the required pressure has been reached, the valve is closed and the pressure is<br />

maintained by a 75 l/s ion pump [E 16]. A Titanium sublimation pump [E 17]<br />

with four filaments can be used to temporarily further reduce the pressure.<br />

4.3.3 Procedure to reach UHV<br />

To reach the ultra low pressure multiple steps were needed. Before any bak-<br />

ing, the chamber was flushed with Argon. The windows and viewports were<br />

replaced by metal blanks, so that a higher temperature could be applied: after<br />

about a week <strong>of</strong> baking at 300 degrees, a pressure <strong>of</strong> p < 7.5 × 10 −12 Torr<br />

was reached (the pressure was lower than the minimum pressure that could<br />

be measured by the cold cathode gauge). During that week, the chip was<br />

96


Chapter 4: The Permanent Magnetic Chip Experiment: Apparatus<br />

allowed to outgas in a second small metal chamber at 10 −9 Torr. After this<br />

initial cleaning, the viewports were put back and the chip was placed inside<br />

the chamber. The vacuum was then pumped down and baked out according<br />

to Table 4.1.<br />

Time 0 35 60.5; 67 80 96<br />

(h) → 35 → 62; → → →<br />

35 59 65.5 80 96 103<br />

Pressure → → max → → →<br />

(Torr) 3× 4× 1.7× 2.2× 8.3× 1.7<br />

10 −7 10 −6 10 −6 10 −9 10 −11 10 −11<br />

Temperature 20 20 140 100<br />

( o C) → → →<br />

special (a) (b)<br />

140 100 20<br />

Table 4.1: The pumping down process. Special: (a) all 4 Ti:siblimation pump<br />

filaments were fired, the Rb-dispensers were outgased, baking was prepared,<br />

(b) at 60.5 h the turbo pump heating was turned on; at 62 h the Ti:sublimation<br />

pump filaments were fired; at 65.5 h turbo pump heating was turned <strong>of</strong>f.<br />

After an overall time <strong>of</strong> more than 100 hours, a final pressure <strong>of</strong> < 2×10 −11<br />

Torr was achieved. The development <strong>of</strong> temperature and pressure are also<br />

shown in Fig. 4.10 starting after the first 35 hours <strong>of</strong> pumping without heating.<br />

The firing <strong>of</strong> the sublimation pump’s filaments is clearly visible by the peak at<br />

t = 1630 min.<br />

Before the experiment reached BEC, we had to open the vacuum once more.<br />

Due to unknown reasons the vacuum had degenerated by roughly one order <strong>of</strong><br />

97


4.4. The Magnetic Field Coils<br />

temperature / o C<br />

200<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0 1000 2000 3000 4000 5000 6000 7000<br />

time / min<br />

10−11<br />

Figure 4.10: The pressure (solid line) and the temperature (dotted line) during<br />

the pump down. The zero point here is set after the system was pumped out<br />

for 35 hours.<br />

magnitude. After the renewed baking and pumping down, the vacuum values<br />

as stated here were reached again.<br />

4.4 The Magnetic Field Coils<br />

To trap the atoms, we use magnetic fields, which couple to the magnetic mo-<br />

ment <strong>of</strong> the atoms. These fields are produced by different parts <strong>of</strong> the set-up.<br />

To compensate for the Earth’s magnetic field and other stray fields that may<br />

perturb the atoms, three pairs <strong>of</strong> coils surround the chamber so that for each<br />

main spatial axis the compensation can be done independently. For the initial<br />

loading <strong>of</strong> atoms in the mirror MOT, we have a set <strong>of</strong> large quadrupole coils.<br />

The atoms are then transferred to a surface MOT whose field is created by a<br />

U-shaped wire on the chip and a pair <strong>of</strong> bias coils. These are also used for<br />

98<br />

10 −5<br />

10 −6<br />

10 −7<br />

10 −8<br />

10 −9<br />

10 −10<br />

pressure / Torr


Chapter 4: The Permanent Magnetic Chip Experiment: Apparatus<br />

magnetic trapping. The chip also carries a permanent magnetic structure that<br />

is used to trap the atoms.<br />

We will now describe the different sets <strong>of</strong> coils, each <strong>of</strong> which is labelled<br />

according to its activity. The arrangement <strong>of</strong> all pairs <strong>of</strong> coils is depicted in<br />

Fig. 4.11.<br />

240 mm<br />

160 mm<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxx xxxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

300 mm<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxx<br />

xxxx<br />

xxxx<br />

xxxx<br />

xxxx<br />

xxxx<br />

xxxx<br />

210 mm<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxxx xxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

Figure 4.11: External coil arrangement around the chamber. Coils with di-<br />

ameter 160 mm are quadrupole coils, coils with diameter <strong>of</strong> 400 mm are bias<br />

coils. The other coils are <strong>of</strong>fset coils.<br />

4.4.1 The <strong>of</strong>fset coils<br />

To compensate for unwanted magnetic fields, pairs <strong>of</strong> square coils have been<br />

aligned along the three orthogonal axes around the chamber. Individual dc cur-<br />

rents run through these coils continuously, reducing the influence <strong>of</strong> the Earth’s<br />

magnetic field. This allows us to realize uniform ballistic expansion during<br />

a molasses stage and to achieve a cold atomic ensemble with sub Doppler<br />

99<br />

400 mm


4.4. The Magnetic Field Coils<br />

temperatures, reducing the temperature by a factor <strong>of</strong> ≈ 5 compared to the<br />

temperature <strong>of</strong> the atoms in the MOT.<br />

To these compensation coils, a second set <strong>of</strong> windings has been added<br />

which can be addressed independently <strong>of</strong> the compensation coils, to give us<br />

more flexibility. These fields can be switched on and <strong>of</strong>f and are used during<br />

the loading <strong>of</strong> the surface MOT and to reduce the magnetic field at the bottom<br />

<strong>of</strong> the magnetic trap, thus increasing the trap frequency. The dimensions and<br />

characteristics <strong>of</strong> these pairs <strong>of</strong> coils are given in table D.1 <strong>of</strong> appendix D.<br />

4.4.2 The quadrupole coils<br />

To create a surface MOT, the symmetry axis <strong>of</strong> a magnetic quadrupole field<br />

has to to be aligned at 45 o to the reflecting surface [Rei99]. We achieve this<br />

by a set <strong>of</strong> external coils in an anti-Helmholtz configuration. The dimensions<br />

<strong>of</strong> this pair <strong>of</strong> coils is given in table D.2 in appendix D. These coils generate<br />

a spatially large field, which is roughly matched with the beam size <strong>of</strong> the<br />

cooling beams, so that the atoms see the MOT forces over a large volume. This<br />

increases the capture volume and effectively the capture velocity <strong>of</strong> the MOT.<br />

We water-cool the coils, so that they can dissipate the power without heating<br />

themselves. The coils are driven with a current <strong>of</strong> 10 A, giving a gradient <strong>of</strong><br />

10 G/cm. Due to the high inductance, the switching time is around 10 ms.<br />

As this MOT is needed to collect a large number <strong>of</strong> atoms initially only, which<br />

are then transferred to a MOT created by the chip, the long switching time is<br />

not a limitation to us.<br />

4.4.3 The bias field coils<br />

For trapping with a single wire, a homogeneous <strong>of</strong>fset field has to be applied<br />

to cancel out the wire’s field. The coils have been wound from a copper wire<br />

100


Chapter 4: The Permanent Magnetic Chip Experiment: Apparatus<br />

with a polyesterimide coating. The coils are wound 6 by 11 turns (width by<br />

height), with a coil former <strong>of</strong> PVC. The maximum current is limited by the<br />

melting point <strong>of</strong> the PVC. At 30 A with a 50:50 duty cycle, the coils reach a<br />

temperature <strong>of</strong> 60 o C and create a field <strong>of</strong> 85 G. The coils are driven by two<br />

power supplies (18 V, 20A) in parallel. The dimensions <strong>of</strong> this piar <strong>of</strong> coils in<br />

Helmhotlz configuration is given in table D.3 in appendix D.<br />

The time needed to switch <strong>of</strong>f the coils is sub ms, although there is a 1.5<br />

ms delay between the signal and the onset <strong>of</strong> the decay <strong>of</strong> the field. If this is<br />

taken into account, the fast switching allows the imaging <strong>of</strong> ballistic expansion<br />

and time <strong>of</strong> flight measurements. The longer time needed to switch on the<br />

coil (≈ 20 ms) is not a problem, as the magnetic field here needs to increase<br />

gradually when changing from the mirror MOT to the compressed MOT and<br />

later to the magnetic trap (see chapter 5).<br />

4.5 The Atom Chip<br />

This section describes the main part <strong>of</strong> the experimental set-up. The atom<br />

chip used in this experiment is novel in its combination <strong>of</strong> current-carrying and<br />

permanent magnetic structures. It allows us to trap atoms with magnetic fields<br />

produced by a current-carrying wire or by a permanent magnetic strip. This<br />

section starts with a presentation <strong>of</strong> the overall arrangement <strong>of</strong> the different<br />

parts. Then these parts are described in detail, starting with the wire structure.<br />

The section ends with the description <strong>of</strong> the permanent magnetic film and its<br />

properties.<br />

4.5.1 Overall design<br />

The atom chip and how its single parts are assembled is shown on a photo-<br />

graph and schematically in Fig. 4.12. A copper block acts as the base piece<br />

101


4.5. The Atom Chip<br />

and as a heat sink. On one side <strong>of</strong> the block the chip is mounted, and the<br />

other side is connected to a solid copper feedthrough with 19 mm diameter<br />

(Ceramaseal, 800 A rating) and thus to the top <strong>of</strong> the vacuum chamber. This<br />

feedthrough acts as a thermal conductor to the atmosphere and can be used as<br />

a possible cold finger when cooled with liquid Nitrogen. Our atom sources are<br />

Rb dispensers, as described and used in [Rap01, For98a]. They are mounted<br />

on ceramic blocks (Macor) at the sides <strong>of</strong> the copper block. To insulate the<br />

current-carrying structure from the copper block, a ceramic slide is glued be-<br />

neath it. The reflective surface needed for a surface MOT is created by the<br />

gold-coated magnetic film deposited on a glass slide on one half; the other<br />

half is made up by a gold-coated glass slide. The adjacent glass slides have a<br />

thickness <strong>of</strong> 300 µm each, with the long edges polished. The thickness <strong>of</strong> the<br />

gold coating is 170 nm. The electrical connections use bare copper wires (Ø =<br />

1.6 mm), BeCu barrel connectors, and a power feedthrough (Ceramaseal, 55<br />

A rating, 12 pin). A publication describing the chip is available [Hal06].<br />

Magnetic film<br />

and coated glass<br />

Atom chip<br />

Insulation<br />

Copper block<br />

with dispensers<br />

Figure 4.12: Photograph (left) and schematic view (right) <strong>of</strong> the overall chip<br />

arrangement.<br />

102


Chapter 4: The Permanent Magnetic Chip Experiment: Apparatus<br />

4.5.2 The current-carrying wires<br />

We now describe the current-carrying wire structure. The procedure for pro-<br />

ducing the chip was developed by [Val04] A 0.5 mm silver foil (99.99 % purity)<br />

was glued onto a Shapal-M ceramic base-plate (thickness 2 mm) using epoxy<br />

Epotek H77. The structure itself consists <strong>of</strong> an “H”-like structure and two<br />

end-wires on each side <strong>of</strong> the “H” (see Fig. 4.13).<br />

22 mm<br />

4 mm<br />

6mm<br />

Figure 4.13: Schematic view <strong>of</strong> the current-carrying wires on the chip. All<br />

distances are from centres <strong>of</strong> the wires. White: remaining foil; grey: foil cut<br />

away, black: holes drilled through the foil and the Macor. The ‘H’-structure<br />

9 mm<br />

allows the wire to be used in ‘U’- or ‘Z’-configuration.<br />

The structure was cut into the foil using a PCB milling machine. This<br />

is a cost effective solution which allows rapid prototyping and the in-house<br />

production <strong>of</strong> the chip. The thick parts, leading to the connectors, have a<br />

width <strong>of</strong> 4.7 mm for the H-wire and 4.5 mm for the end wires. The thin<br />

wire parts have widths <strong>of</strong> 1 mm. The length <strong>of</strong> the bridge part <strong>of</strong> the “H”<br />

is thus 4 mm long. The structure was cut into the foil with a cutting depth<br />

<strong>of</strong> approximately 0.5 mm, in 6 consecutive cuts <strong>of</strong> about 90 µm each. The<br />

width <strong>of</strong> the insulating grooves is 0.5 mm. The holes on two sides <strong>of</strong> the chip<br />

103<br />

32 mm


4.5. The Atom Chip<br />

are 2.0 mm and 1.7 mm in diameter and used to connect the wires to the<br />

feedthrough and to hold the chip to the copper block. This connection was<br />

later identified to be the point <strong>of</strong> highest resistance in the system, and led<br />

to substantial heating at the connections. This led to outgassing <strong>of</strong> the chip,<br />

and in turn to a strongly reduced lifetime <strong>of</strong> the magnetic trap, one <strong>of</strong> the<br />

main reasons why BEC was not reached in the initial runs. To reduce this<br />

effect each connection was bridged by an additional thin sheet <strong>of</strong> metal from<br />

the chip to the wire. These new, additional connections led to a decrease in<br />

the overall resistance <strong>of</strong> chip and connectors from about 0.5 Ω to 0.32 Ω. The<br />

wire resistance <strong>of</strong> the “Z”-wire was measured to be 4.6 mΩ while the chip<br />

was changed, highlighting the need to measure the resistances before the set-<br />

up is assembled and placed into the vacuum. Before the modification <strong>of</strong> the<br />

connectors, a maximum current <strong>of</strong> 25 A could be run through the wire without<br />

adverse effect on the lifetime. After the modifications a current <strong>of</strong> 31.1 A was<br />

used without decreasing the lifetime <strong>of</strong> the magnetic trap. After the machining<br />

and before the chip was mounted on the copper block, the silver was polished<br />

using 600 grade sandpaper.<br />

Having an “H”-shaped structure allows versatile operation. An “U”-shaped<br />

wire (creating a quadrupole trap) and a “Z”-shape wire (for a harmonic trap)<br />

are both possible. In the “U”-shaped regime, trap gradients <strong>of</strong> up to 40 G/cm<br />

are possible for magneto-optical trapping limited only by the breakdown <strong>of</strong><br />

polarisation gradient cooling at high field gradients.<br />

The end wires, with a separation <strong>of</strong> 9 mm, are not only used to provide a<br />

confining field when trapping with the edge <strong>of</strong> the magnetic film, but also serve<br />

as our RF antennas for the evaporative cooling. A frequency synthesizer [E<br />

19] is connected to the wires to create the RF frequency. A sensing resistance<br />

<strong>of</strong> 5 Ω is used to measure the resonance spectrum <strong>of</strong> the antenna loop. The<br />

resonance shows a Lorentzian curve, centred at 11.8 MHz with a FWHM <strong>of</strong><br />

104


Chapter 4: The Permanent Magnetic Chip Experiment: Apparatus<br />

2.9 MHz. This knowledge allows better control over the power <strong>of</strong> the radiation<br />

during a ramping <strong>of</strong> the frequency as done in the process <strong>of</strong> evaporative cooling.<br />

4.5.3 The magneto-optical film<br />

The main component <strong>of</strong> the permanent magnetic strip is a ferrimagnetic magneto-<br />

optical (MO) film. The structure is made from a TbGdFeCo / Cr sandwich<br />

structure [Wan05a], on a glass slide, coated with gold for reflectivity. The<br />

TbGdFeCo material has its magnetisation axis perpendicular to the film plane.<br />

The structure is schematically shown in Fig. 4.14. The film was produced at<br />

<strong>Swinburne</strong> <strong>University</strong>, with in-house technology that was developed for this<br />

task. It was prepared using a commercial thin film deposition system [E 20]<br />

with three magnetron sputtering guns and electron beam evaporation. The<br />

sandwich consists <strong>of</strong> six alternating layers <strong>of</strong> the magneto-optical material<br />

(Tb6Gd10Fe80Co4, ≈ 170 nm thickness) and Chromium (≈ 140 nm thickness).<br />

This is coated by a thin (10 nm) Cr layer and a gold layer <strong>of</strong> 170 nm thickness<br />

for reflectivity.<br />

A hysteresis curve <strong>of</strong> the film was taken using the Magneto-Optical Kerr<br />

Effect (MOKE) and shows a nearly square-like behaviour (Fig. 4.15). The co-<br />

ercivity is around Hc = 4 kOe. The remanent magnetic field was determined<br />

to be Brem = 2.8 kG, which is equivalent to an effective current <strong>of</strong> Ieff = 200<br />

mA. The film was also analysed by SQUID measurements [Whi05, Hal06]. The<br />

coercivity here was determined to be 2.7 kOe and the remanent magnetisation<br />

to be 3.3 kG, in reasonable agreement with the MOKE results. The Curie<br />

temperature <strong>of</strong> a similar film was measured by Imation to be TC ≈ 300 o C. A<br />

similar film was produced together with samples for a magnetic mirror exper-<br />

iment [Wan05a]. Here a single-layer TbGdFeCo/Cr film was examined for the<br />

influence <strong>of</strong> the Cr underlayer on the grain size <strong>of</strong> the MO film using atomic<br />

105


4.5. The Atom Chip<br />

Cr<br />

Au<br />

Glass<br />

MO<br />

film<br />

Figure 4.14: Schematic view <strong>of</strong> the<br />

magnetic film, with alternating layers<br />

<strong>of</strong> Cr and magneto-optical film on a<br />

glass slide, with gold coating.<br />

photodiode signal / V<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

−10<br />

−8<br />

− H c<br />

−6<br />

−4<br />

−2<br />

0<br />

∝ B rem<br />

applied field / kOe<br />

Figure 4.15: Hysteresis curve <strong>of</strong> the<br />

film by MOKE, used to determine the<br />

magnetic polarisation Brem and the<br />

coercivity Hc.<br />

force microscopy (AFM). The MO film on Cr shows a smaller grain size (about<br />

40 nm) and smoother rounder grains than a Cr or MO film alone (grain size<br />

<strong>of</strong> 50 nm). It was found that without the Cr mid-layers, the coercivity was<br />

reduced to less than 1.5 kOe. In addition, the remanent magnetisation and co-<br />

ercivity deteriorated for film thicknesses larger than 200 nm. Using magnetic<br />

force microscopy (MFM) the domain structure <strong>of</strong> the magnetised sample was<br />

examined, showing excellent magnetic homogeneity (see Fig. 4.16). No do-<br />

main structures are visible in the magnetised sample, limited by the resolution<br />

<strong>of</strong> the MFM <strong>of</strong> 100 nm.<br />

106<br />

2<br />

4<br />

6<br />

8<br />

10


Chapter 4: The Permanent Magnetic Chip Experiment: Apparatus<br />

Figure 4.16: Magnetic force microscope (MFM) images <strong>of</strong> the Gd10Tb6Fe80Co4<br />

magnetic film surface, with 1 µm/scale. Left (a) the unmagnetised sample<br />

showing magnetic domains <strong>of</strong> micron size. Right (b) the uniformly magnetised<br />

sample showing no visible magnetic structure. Taken from [Hal06].<br />

107


4.5. The Atom Chip<br />

108


Chapter 5<br />

The Permanent Magnetic Chip<br />

Experiment: Results<br />

5.1 Overview and Timing Sequence<br />

In this chapter the results <strong>of</strong> the permanent magnetic chip experiment located<br />

at <strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong> are presented. In this overview the<br />

steps taken to reach degeneracy <strong>of</strong> the atomic sample are outlined, with the<br />

most important parameters and results. In the following sections, the impor-<br />

tant steps are explained and further results are presented there.<br />

Each experiment starts with a phase <strong>of</strong> loading atoms in the mirror magneto-<br />

optical trap. For this, the Rb dispensers are pulsed for 12 s at 6.5 A. To allow<br />

the vacuum to recover, this is followed by a 25 s waiting phase. During the<br />

loading the cooling light is detuned by 18.5 MHz, and in the waiting phase it is<br />

detuned by 15 MHz. With a gradient <strong>of</strong> about 10 G/cm from the quadrupole<br />

coils, we trap about 5 · 10 8 atoms at about 90 µK. The atoms are then trans-<br />

ferred to the compressed MOT (CMOT) by linearly ramping up the current in<br />

the U-wire to 8 A and applying a bias field <strong>of</strong> about 7 G perpendicular to the<br />

109


5.1. Overview and Timing Sequence<br />

wire in the plane <strong>of</strong> the chip. This ramp takes 20 ms, and at the same time<br />

the current in the quadrupole coils is reduced to zero. This is performed at a<br />

detuning <strong>of</strong> 15 MHz. As the gradient increases to about 33 G/cm, the atomic<br />

cloud heats up to a temperature <strong>of</strong> 140 µK, while it is compressed to less than<br />

one quarter <strong>of</strong> its previous size. To further cool the atoms, the field gradient <strong>of</strong><br />

the CMOT is decreased to 11.2 G/cm, while the light is turned <strong>of</strong>f and detuned<br />

to 56 MHz in a time <strong>of</strong> 2 ms. The position <strong>of</strong> the cloud is held constant. The<br />

sample <strong>of</strong> atoms is then optically cooled to 40 µK by illuminating with the<br />

highly detuned light for 2 ms. This is followed by a stage <strong>of</strong> optical pumping<br />

in which the magnetic substates <strong>of</strong> the atoms are changed to the mF = 2 state.<br />

For this, the CMOT is turned <strong>of</strong>f, while the bias field is increased to 6 G over<br />

2.4 ms. The optical pumping is then performed with a 200 µs pulse <strong>of</strong> about<br />

0.2 mW/cm 2 , with a frequency <strong>of</strong> about 50 MHz below the F = 2 → F ′ = 2<br />

transition. A 500 µs pulse <strong>of</strong> repumping light <strong>of</strong> low power is also applied here.<br />

The atoms are now prepared to be loaded into the magnetic trap created<br />

by the wire. The bias field is increased to 19.5 G in 1 ms, and a current <strong>of</strong> 21.5<br />

A is passed through the Z-shaped wire. Mode matching is achieved by using<br />

the bias coils. We collect 8 · 10 7 atoms at 50 µK in this initial magnetic trap.<br />

In a 100 ms linear ramp, this trap is adiabatically compressed to a gradient<br />

<strong>of</strong> about 510 G/cm by further increasing the currents in the wire (to 31.1 A)<br />

and bias coils (to a field <strong>of</strong> 56.3 G). This moves the trap closer to the chip<br />

surface and heats the atoms to 160 µK. Further cooling is performed using<br />

the mechanism <strong>of</strong> forced evaporative cooling by applying a RF-current to the<br />

end-wires. The frequency is then ramped down from 20 MHz to 0.78 MHz in<br />

10 seconds. At the same time the trap is further compressed by reducing the<br />

current in the Z-wire and keeping the bias field constant. At the end <strong>of</strong> the<br />

evaporation, the wire current is thus 24.9 A, leaving us with a gradient <strong>of</strong> 635<br />

G/cm if we ignore the effect <strong>of</strong> the permanent magnetic film. At this stage<br />

110


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

the transition to a Bose-Einstein condensate has taken place and we have up<br />

to 10 5 atoms in the condensate.<br />

As the atom cloud is close to the magnetic film, the effect <strong>of</strong> its field can<br />

not be ignored, and for imaging the atoms with the time-<strong>of</strong>-flight technique,<br />

we need to adiabatically move the atoms away from the chip. This is done by<br />

a logarithmic ramp <strong>of</strong> 50 ms decreasing the bias field to 37.5 G, which moves<br />

the atoms about 1 mm away from the surface. Imaging is performed by taking<br />

4 images, each with a 100 µs pulse <strong>of</strong> resonant light. The first picture is taken<br />

<strong>of</strong> the MOT before the transfer. The second is taken without illumination<br />

during the RF-sweep as a correction for dark counts <strong>of</strong> the camera. The third<br />

picture is the main picture. This is divided by the fourth image, which is a<br />

background picture without atoms.<br />

This chapter starts with the characterisation <strong>of</strong> the initial atomic ensem-<br />

ble in the mirror magneto-optical trap, followed by a characterisation <strong>of</strong> the<br />

compressed MOT. The next part covers the initial magnetic trapping and the<br />

adiabatic compression <strong>of</strong> that trap. The subsequent section deals with the<br />

cooling by RF-induced evaporation and the reaching <strong>of</strong> quantum degeneracy.<br />

The chapter closes with some results on the trapping <strong>of</strong> atoms in the field <strong>of</strong><br />

the permanent magnetic film.<br />

5.2 The Magneto-Optical Traps<br />

The starting point <strong>of</strong> all experiments is collecting a large number <strong>of</strong> atoms at<br />

reasonably low temperatures. This is done by trapping atoms in a mirror MOT<br />

(see section 2.5.1). We use quadrupole coils (see section 4.4.2) to create a MOT<br />

with a large trapping volume, which enables us to trap and cool a very large<br />

number <strong>of</strong> atoms. From this point, the current in the coils is ramped down.<br />

Ramping up currents in the wire structure on the chip (see section 4.5.2) and<br />

111


5.2. The Magneto-Optical Traps<br />

in the Helmholtz coils that are placed around the chamber (the ‘bias coils’,<br />

see section 4.4.3) creates a new magnetic field that replaces the initial field.<br />

Increasing the current in the bias coils moves the trap centre towards the chip’s<br />

surface and increases the gradient <strong>of</strong> the magnetic field, creating a compressed<br />

MOT (CMOT). The new gradients are chosen so that the atomic cloud can<br />

be loaded into the magnetic trap with minimal losses. To further reduce these<br />

losses, a stage <strong>of</strong> molasses-like cooling in a less steep trap is applied. A further<br />

important step for optimal loading is to optically pump the atoms into the<br />

mF = 2 magnetic substate.<br />

In this section data from the mirror MOT is presented first, which can be<br />

used to calibrate some laser parameters <strong>of</strong> the system. This is followed by the<br />

results for the compressed MOT including further cooling. The section ends<br />

with a description <strong>of</strong> the optical pumping as a preparation for the magnetic<br />

trapping.<br />

The mirror MOT<br />

Early measurements on the atomic ensemble in the MOT were performed using<br />

column density absorption. For the column density absorption, a weak laser<br />

beam was scanned in frequency over the |F = 2〉 → |F ′ = 3〉 resonance <strong>of</strong> the<br />

D2-line. Its intensity was measured after passing through the atomic cloud as<br />

a function <strong>of</strong> the frequency. The intensity had to be weak so that this directed<br />

beam did not destroy the MOT. The result <strong>of</strong> this measurement is shown in<br />

Fig. 5.1. Fitting a Lorentzian to the logarithm <strong>of</strong> the normalised absorption<br />

curve gave a maximum absorption <strong>of</strong> 96% and a full width at half maximum<br />

(FWHM) <strong>of</strong> 9.1 MHz. The Lorentzian curve was not perfectly centred at zero<br />

detuning, marked by the right dotted line. At a detuning <strong>of</strong> δf ≈ −16 MHz, a<br />

small yet steep dispersive-like structure is visible. The left dotted line marks<br />

the structure and intercepts the frequency scale. The structure appears at the<br />

112


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

logarithm <strong>of</strong> normalised Intensity<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

−2<br />

−2.5<br />

−3<br />

−50 −40 −30 −20 −10 0 10 20 30 40 50<br />

Detuning / MHz<br />

Figure 5.1: Column density measurement <strong>of</strong> the MOT by absorption; thin line:<br />

a Lorentzian fitted to the logarithm <strong>of</strong> the absorption signal, dotted lines are<br />

explained in the text.<br />

detuning <strong>of</strong> the trapping light fields from the atomic resonance and has its<br />

source in a Λ-system where the probe and the trapping lasers couple the atom<br />

states [Tab91]. We thus work with a trapping laser detuned by δf = 15.1 MHz,<br />

the distance between the two dotted lines.<br />

To evaluate the quality <strong>of</strong> the magneto-optical trap we determined the<br />

linewidth <strong>of</strong> the |F = 2〉 → |F ′ = 3〉 transition <strong>of</strong> the D2-line absorption<br />

spectrum by absorption imaging: our goal is to collect a large number <strong>of</strong><br />

atoms at a low temperature at a high density in the trap. The frequency <strong>of</strong><br />

the absorption light was tuned over the atomic resonance. Each picture was<br />

taken with the MOT turned <strong>of</strong>f after a loading time <strong>of</strong> 30 s and corrected for<br />

the background. The data are depicted in Fig. 5.2. It shows the number <strong>of</strong><br />

absorbing atoms versus the detuning <strong>of</strong> the imaging light, and a Lorentzian was<br />

fitted to the data. The Lorentzian has a FWHM <strong>of</strong> 8.7 MHz. This is slightly<br />

closer to the natural linewidth <strong>of</strong> 6.1 MHz than the result <strong>of</strong> the column density<br />

measurement. Here the MOT was turned <strong>of</strong>f before each image was taken. As<br />

the pr<strong>of</strong>ile is Lorentzian, only the natural linewidth <strong>of</strong> the transition, collisional<br />

113


5.2. The Magneto-Optical Traps<br />

Number <strong>of</strong><br />

absorbing atoms<br />

x 108<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

−30 −20 −10 0 10 20 30<br />

Detuning / MHz<br />

Figure 5.2: Integrated absorption spectrum <strong>of</strong> the MOT, plotted as number<br />

<strong>of</strong> absorbing atoms versus detuning <strong>of</strong> the imaging light. Solid line: fitted<br />

Lorentzian.<br />

broadening and the linewidth <strong>of</strong> the absorption laser contribute significantly<br />

to this result. Doppler broadening has a Gaussian form and would have led to<br />

a Voigt pr<strong>of</strong>ile in the absorption spectrum. In the convolution <strong>of</strong> Lorentzian<br />

curves, the widths <strong>of</strong> the pr<strong>of</strong>iles add to the resulting width. The pr<strong>of</strong>ile being<br />

Lorentzian means that we deal with cold atoms. Allowing for the natural<br />

linewidth only we can give an upper bound for the emission linewidth <strong>of</strong> the<br />

main laser: ∆νMOPA < 2.6 MHz. A scanning etalon was used to measure<br />

the spectral linewidth <strong>of</strong> the laser directly. The result here was ∆νMOPA ≤ 1.6<br />

MHz. The difference probably stems from collisional broadening, and indicates<br />

that we have a high atomic density in the MOT.<br />

We can use the above measurement to measure the overall number <strong>of</strong> atoms<br />

if we correct for the shape <strong>of</strong> the absorption probability. The above yields a<br />

number <strong>of</strong> atoms <strong>of</strong> N = (2.2 ± 0.8) · 10 9 atoms, using equation (2.31) summed<br />

over all camera pixels and taking the average over the atom numbers calculated<br />

from the data for different detunings. We qualitatively find that our MOT<br />

fulfills the requirements <strong>of</strong> collecting a large number <strong>of</strong> cold atoms at high<br />

114


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

density. Typically, the trap is at a distance <strong>of</strong> 4.6 mm from the chip surface<br />

and the atomic cloud <strong>of</strong> 5 · 10 8 atoms has a 1/e size <strong>of</strong> about 1.1 mm. The left<br />

side <strong>of</strong> Fig. 5.3 shows the absorption image after correction for the background<br />

and taking the logarithm <strong>of</strong> the intensity. The radial pr<strong>of</strong>ile <strong>of</strong> this cloud is<br />

depicted in Fig. 5.3, right, with a Gaussian curve fitted to it.<br />

log absorption signal<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 500 1000 1500 2000 2500 3000<br />

radial distance / µm<br />

Figure 5.3: Picture (left) and radial pr<strong>of</strong>ile (right) <strong>of</strong> the atoms in the mirror<br />

MOT.<br />

The final particle number in a MOT is determined by the ratio <strong>of</strong> its loading<br />

and loss rates. These rates can be determined when the fluorescence signal <strong>of</strong><br />

the MOT is collected by a photodiode as a function <strong>of</strong> the time after turning on<br />

the MOT and after turning <strong>of</strong>f the source <strong>of</strong> atoms. A graph <strong>of</strong> the photodiode<br />

signal when the MOT is loaded with atoms is shown in Fig. 5.4 (a). A graph<br />

<strong>of</strong> the decaying fluorescence signal <strong>of</strong> the MOT which is not being replenished<br />

with atoms is shown in Fig. 5.4 (b). The photodiode was calibrated with<br />

3.2 · 10 7 atoms/V. As the maximum voltage is slightly less than 15 V, the<br />

photodiode saturates at about 4.8 · 10 8 atoms in the trap and does not allow<br />

the detection <strong>of</strong> higher atom numbers.<br />

The dynamics <strong>of</strong> the atom number are fully characterised by the loading<br />

115


5.2. The Magneto-Optical Traps<br />

numer <strong>of</strong> atoms<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

x 108<br />

5<br />

(a)<br />

0<br />

0 10 20 30 40 50 60 70 80 90<br />

time /s<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

x 108<br />

5<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Figure 5.4: Time dependence <strong>of</strong> the fluorescence signal <strong>of</strong> the mirror MOT,<br />

number <strong>of</strong> atoms<br />

(b)<br />

time /s<br />

(a) after being turned on, (b) after turning <strong>of</strong>f the atom source.<br />

rate RL and the loss rate γ; the atom number N(t) follows a rate equation:<br />

Integration gives<br />

dN(t)<br />

dt = RL − γ · N(t)<br />

N(t) = RL<br />

γ · (1 − e−γt ) (5.1)<br />

In the case <strong>of</strong> having turned <strong>of</strong>f the atomic source, this simplifies to<br />

with the solution<br />

dN(t)<br />

dt<br />

= −γ · N(t)<br />

N(t) = N(0) · e −γt . (5.2)<br />

Using these solutions functions have been fitted to the data points. The<br />

loading rate is given here as RL = 1.82 · 10 7 s −1 . The loss rate from the<br />

loading data is γ1 = 0.037 s −1 , and from the decaying signal is γ2 = 0.011 s −1 .<br />

These correspond to lifetimes <strong>of</strong> τ1 = 27 s and τ2 = 91 s, respectively. As we<br />

use dispensers for loading the trap, switching <strong>of</strong>f the source does not instantly<br />

affect the MOT. The dispensers give rise to a high partial pressure <strong>of</strong> Rb atoms<br />

in the vacuum. As long as these Rb atoms are still in the vicinity <strong>of</strong> the trap,<br />

116


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

the MOT can use them as a background gas and replenish itself from there.<br />

Thus, we have to expect the measured loss rate for the turning <strong>of</strong>f to be lower<br />

than for the turning on, or start the measurement after the vacuum definitely<br />

has recovered.<br />

To measure the temperature <strong>of</strong> an atomic ensemble we use the “time-<strong>of</strong>-<br />

flight” (TOF) or ballistic expansion technique. Here the trapping potential<br />

and the cooling lasers are turned <strong>of</strong>f, releasing the atomic cloud which starts<br />

to fall under gravity. It will also expand along all three spatial directions.<br />

Using the identity between thermal and kinetic energy along one axis, we get<br />

Ekin = Eth<br />

1<br />

2 mv2 = 1<br />

2 kBT<br />

T = m<br />

v 2 . (5.3)<br />

Here, the mean velocity <strong>of</strong> the ensemble has to be taken. At each time, the<br />

atomic cloud will show a variance. This statistical term is also called the<br />

second order central moment <strong>of</strong> a distribution and is a measure <strong>of</strong> the spread<br />

<strong>of</strong> the distribution around the mean value. This value can either be taken<br />

explicitly from the image taken or after fitting a Gaussian curve to the data,<br />

where the variance σ appears in the functional term:<br />

kB<br />

x2<br />

−<br />

fGauss ∝ e 2σ2 Using the fit to a Gaussian distribution will lead to smaller values than calcu-<br />

lating the second order moment directly, as the fitting procedure will smooth<br />

out noise in the wings <strong>of</strong> the distribution which in the other case will add to<br />

the moment.<br />

As the cloud in the trap before the release at t = 0 has a variance σ0, the<br />

image at a time ti will have a variance <strong>of</strong> (under the assumption that the cloud<br />

117


5.2. The Magneto-Optical Traps<br />

at all times has a Gaussian pr<strong>of</strong>ile):<br />

σ 2 i = σ 2 v(ti) + σ 2 0<br />

(5.4)<br />

If we assume the dynamic part to behave like σv(ti) = v · ti, we can substitute<br />

after squaring and get v2 = σ2 i −σ2 0<br />

t2 . If the initial size is unknown, the mean<br />

i<br />

velocity can be determined by taking two images at different times, so that<br />

the temperature is<br />

T = m<br />

As we work with 87 Rb only, we can substitute and get<br />

kB<br />

σ2 1 − σ2 2<br />

t2 1 − t2 . (5.5)<br />

2<br />

T = 10.45 nK σ2 1 − σ2 2<br />

t2 1 − t2 , (5.6)<br />

2<br />

where the variances σi are measured in µm and the times ti in ms. The size<br />

<strong>of</strong> the atomic cloud in the trap can be determined from the intercept <strong>of</strong> the<br />

graph.<br />

The final temperature <strong>of</strong> the atomic ensemble is determined by the detun-<br />

ing <strong>of</strong> the trapping light. We have examined the temperature <strong>of</strong> the atoms as<br />

a function <strong>of</strong> the detuning using TOF measurements. For each detuning the<br />

temperature was determined by linear regression. The variance was calculated<br />

directly from the data. The results are shown in Fig. 5.5. The maximum pos-<br />

sible detuning here was ∆ = −36.5 MHz. This led to a minimum temperature<br />

<strong>of</strong> 27 µK which is considerably smaller than the Doppler-limited temperature.<br />

With the recoil limit for 87 Rb, Trec = 359 nK, we had a temperature <strong>of</strong> about<br />

75 · Trec. The set-up was later changed to allow a larger detuning.<br />

The compressed MOT (CMOT)<br />

The atoms that were collected in the mirror MOT are then transferred into a<br />

MOT where the magnetic field is created by the wire on the chip. The current<br />

runs through the wire in a U-configuration creating a quadrupole field when a<br />

118


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

Temperature / µ K<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

−40 −35 −30 −25 −20 −15 −10 −5 0<br />

Detuning / MHz<br />

Figure 5.5: Temperature <strong>of</strong> the atoms in the mirror MOT as a function <strong>of</strong> the<br />

detuning <strong>of</strong> the trapping light. Each temperature and the error are the results<br />

<strong>of</strong> a linear regression for time-<strong>of</strong>-flight measurements.<br />

uniform bias field is added. The trap is moved closer to the surface while at<br />

the same time the gradient is increased. This is done as a preparation for the<br />

magnetic trapping, trying to keep a large number <strong>of</strong> atoms at a low temperature<br />

in a potential that is deformed towards the potential <strong>of</strong> the magnetic trap.<br />

Here it is described how both low temperatures and a large atom number in<br />

the magnetically trappable substate F = 2, mF = 2 can be realised.<br />

We observed a loss <strong>of</strong> atoms when moving our trap. This loss appeared<br />

when our trap centre moved to about 1 mm from the surface <strong>of</strong> the atom chip.<br />

As in this phase both gradient and position are changed, we examined the loss<br />

mechanism more closely.<br />

The loss <strong>of</strong> atoms from the compressed MOT was measured for different<br />

settings. In one setting, the distance <strong>of</strong> the atomic cloud from the atom chip<br />

surface was kept constant (d = 1.29 mm), while the gradient <strong>of</strong> the magnetic<br />

field was varied ( ∂B<br />

∂z<br />

= 39 G/cm and ∂B<br />

∂z<br />

the other setting, the gradient was fixed at ∂B<br />

∂z<br />

= 77 G/cm)(see Fig. 5.6 (a)). In<br />

= 77 G/cm and measurements<br />

were taken at distances <strong>of</strong> d = 0.78 mm and d = 1.29 mm (shown in Fig. 5.6<br />

119


5.2. The Magneto-Optical Traps<br />

(b)). For the same distance with different gradients we see different loss rates.<br />

A higher gradient causes a higher loss rate. The loss seems to be independent<br />

<strong>of</strong> the distance, as long as the cloud does not touch the surface.<br />

Remaining fraction<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

(a)<br />

10 1<br />

10 2<br />

Hold time / ms<br />

10 3<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

(b)<br />

10 1<br />

Hold time / ms<br />

Figure 5.6: Examination <strong>of</strong> the main loss mechanism when the MOT was<br />

compressed and moved towards the chip: remaining fraction <strong>of</strong> atoms in the<br />

CMOT as a function <strong>of</strong> the hold time at: (a) same distance from the surface<br />

<strong>of</strong> 1.29 mm and different magnetic field gradients <strong>of</strong> 39 G/cm (crosses) and<br />

77 G/cm (dots); (b) same magnetic field gradient <strong>of</strong> 77 G/cm and different<br />

distances from the surface <strong>of</strong> 0.78 mm (crosses) and 1.29 mm (dots). The lines<br />

are included to guide the eye.<br />

Remaining fraction<br />

The influence <strong>of</strong> the permanent magnetic film also becomes noticable here.<br />

We compared the value <strong>of</strong> the applied bias field calculated from the applied<br />

current and the known dimensions <strong>of</strong> the coils with the value <strong>of</strong> the field from<br />

the distance between the trap and the wire rw. When using the simple equation<br />

for a wire with current Iw and perpendicular bias field Bb (see equation (2.39)):<br />

Bb = µ0 Iw<br />

2π rw<br />

10 2<br />

(5.7)<br />

the values deviate at small distances from the wire. When correcting for the<br />

magnetic field <strong>of</strong> the film, it has to be kept in mind that the wires and film are<br />

120


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

separated by a small distance δR, so that the distance between the magnetic<br />

film and the trap is R = rw − δR. For a film with a magnetisation equivalent<br />

to a current IF we thus get:<br />

Bb = µ0 Iw<br />

2π rw<br />

+ µ0 IF<br />

2π R<br />

(5.8)<br />

We now have the option to fit this function while keeping one parameter fixed:<br />

if we assume to know the distance in our experiment (δR = 0.55 mm, the<br />

sum <strong>of</strong> the thickness <strong>of</strong> the glass slide with the film and half the thickness <strong>of</strong><br />

the silver foil) we obtain much better agreement between the two values <strong>of</strong><br />

the magnetic fields by using an equivalent current <strong>of</strong> IF = 0.44 A, which is<br />

in rough agreement with the value for the effective current obtained from the<br />

SQUID measurement [Hal06]. If we assume to know the effective current <strong>of</strong><br />

the film IF = 0.2 A from the SQUID measurement, then the data yields a<br />

distance <strong>of</strong> δR = 0.46 mm between the magnetic film and the current-carrying<br />

wire. Figure 5.7 shows a plot <strong>of</strong> the values <strong>of</strong> the bias magnetic field calculated<br />

from the coil dimensions, from equation (5.7) and from equation (5.8), for a<br />

wire current Iw = 6 A. The solid line is the result <strong>of</strong> either fitted function<br />

including the influence <strong>of</strong> the film. The dotted line is the result <strong>of</strong> equation<br />

(5.7) for a distance between film and wire <strong>of</strong> δR = 0.46 mm, the dash-dotted<br />

line is the result <strong>of</strong> the same equation for δR = 0.55 mm. It is clear that for<br />

distances smaller than 1 mm the influence <strong>of</strong> the magnetic film is no longer<br />

negligible, but it is not clear which <strong>of</strong> the two cases is the actual case.<br />

Typically, the compressed MOT is run with a current <strong>of</strong> Iw = 8 A in the<br />

wire. The bias field is created by operating the bias coils (section 4.4.3) with<br />

a field <strong>of</strong> Bb,1 = 18.7 G and a set <strong>of</strong> the smaller compensation coils in the x-<br />

direction (section 4.4.1) with a field <strong>of</strong> Bb,2 = 11.2 G in the opposite direction.<br />

This leads to an overall bias field <strong>of</strong> Bb ≈ 7.5 G. Two counteracting fields are<br />

used in this way to reduce the switching time, which in our set-up is longer<br />

121


5.2. The Magneto-Optical Traps<br />

Applied bias field / G<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 0.5 1 1.5 2 2.5 3<br />

Distance from surface / mm<br />

Figure 5.7: Examining the influence <strong>of</strong> the magneto-optical film: the applied<br />

bias field Bb is plotted against the distance between the trap and the surface<br />

R. Circles are the calculated bias fields from the current and the dimensions <strong>of</strong><br />

the coils against the measured trap position for this current. The dash-dotted<br />

and dotted line result from equation (5.7) for a wire trap without influence<br />

<strong>of</strong> the magnetic film with Iw = 6 A and different distances between film and<br />

wire. The solid line is calculated from equation (5.8) with the influence <strong>of</strong> the<br />

film, IF = 0.44 A.<br />

for building up a field. This longer time is avoided by the faster turn-<strong>of</strong>f <strong>of</strong> the<br />

countering field, which effectively results in a fast increase. Taking an effective<br />

current <strong>of</strong> 0.44 A for the film, the trap emerges at R = 1.6 mm from the surface<br />

with a gradient <strong>of</strong> ∂B<br />

∂z<br />

losses in the transfer to a negligible amount.<br />

= 33 G/cm. This gradient is low enough to reduce the<br />

For efficient loading into the magnetic trap we need both cold atoms and<br />

atoms close to the chip surface. At distances as close to the surface as ours,<br />

the proximity <strong>of</strong> the permanent magnetic film makes the usual mechanism <strong>of</strong><br />

sub-Doppler cooling impossible. Thus a compromise had to be made: cooling<br />

with far detuned light at a lower gradient. For this, the field <strong>of</strong> the bias coils<br />

is reduced to Bb,1 = 13.7 G over 1 ms; thus we reach an overall bias field <strong>of</strong><br />

122


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

Bb = 2.5 G. At the same time the current in the U-wire is reduced to Iw = 2.8<br />

A. This keeps the trap at the same position, but quickly lowers the gradient to<br />

about one third: ∂B<br />

∂z<br />

= 11.2 G/cm. During this switching, the light is turned<br />

<strong>of</strong>f and changed to a much higher detuning <strong>of</strong> 56 MHz. We examined the final<br />

temperature <strong>of</strong> the atoms in the cloud as a function <strong>of</strong> both the final detuning<br />

and the time the cooling light was applied. The results are shown in Fig. 5.8.<br />

A short pulse <strong>of</strong> 2 ms is sufficient to cool the atoms, and as expected a larger<br />

Temperature / µ K<br />

150<br />

100<br />

50<br />

0<br />

0<br />

−10<br />

−20<br />

−30<br />

detuning / MHz<br />

−40<br />

−50<br />

−60<br />

12<br />

10<br />

8<br />

6<br />

4<br />

cooling time / ms<br />

Figure 5.8: Sub-Doppler cooling the atoms in a weak magnetic field: the<br />

temperature <strong>of</strong> the trapped atoms against the detuning and the duration <strong>of</strong><br />

the cooling light. The lowest temperature is achieved with a cooling pulse <strong>of</strong><br />

≥ 2 ms with maximum detuning. The distance <strong>of</strong> the trap from the surface is<br />

1.6 mm.<br />

detuning leads to lower final temperatures. We thus apply a pulse <strong>of</strong> 2.5 ms<br />

with light detuned by 56 MHz and reach temperatures as low as 40 µK for a<br />

distance from the the surface <strong>of</strong> 1.6 mm.<br />

To increase the population in the magnetically trappable atomic substate<br />

F = 2, mF = 2, we turn <strong>of</strong>f the light and the current in the wire, while<br />

increasing the bias field to Bb = 6 G over 2.4 ms. The bias field now acts<br />

123<br />

2<br />

0


5.3. The Wire Magnetic Trap<br />

as a quantisation axis for the atomic spin and lifts the degeneracy <strong>of</strong> the<br />

substates by the Zeeman effect. Using σ + -polarised light will move the atomic<br />

population towards the state with the highest magnetic quantum number mF<br />

by conservation <strong>of</strong> angular momentum. As any heating or a change in position<br />

<strong>of</strong> the atoms is undesirable, the pumping light is retroreflected so that the net<br />

heating is kept minimal. Further, only a low light power <strong>of</strong> 2 mW is applied<br />

for a time <strong>of</strong> 200 µs. It was found that the number <strong>of</strong> atoms trapped by<br />

the magnetic field could be increased by a factor <strong>of</strong> about 5 by locking the<br />

light to the frequency <strong>of</strong> the F = 2 → F ′ = 2 transition detuned by 50 MHz<br />

to the blue. This increase shows that before the pumping we have the five<br />

magnetic substates very evenly populated and that our pumping works with<br />

nearly 100% efficiency. This high efficiency is unexpected, as the detuning <strong>of</strong><br />

50 MHz corresponds to a Zeeman splitting between adjacent mF -states when<br />

the magnetic field is about 71.5 G. In our case, we have a field <strong>of</strong> 6 G only,<br />

and thus would expect the best pumping to occur at a detuning <strong>of</strong> roughly 4.5<br />

MHz. The discrepancy remains unexplained.<br />

5.3 The Wire Magnetic Trap<br />

The transfer trap<br />

Once the atoms have been transferred to the magnetically trappable state, we<br />

have to create a magnetic potential that captures these atoms at their position.<br />

From there the atoms can be further manipulated towards Bose degeneracy.<br />

To trap the atoms magnetically, the bias field is ramped to Bb = 19.5<br />

G in 1 ms, while the current in the wire is switched to Iw = 21.5 A in the<br />

Z-configuration. This leads to a magnetic trap at the same distance <strong>of</strong> 1.6<br />

mm from the surface with a gradient ∂B<br />

∂z<br />

= 88.4 G/cm. Strictly speaking, as<br />

we use the Z-wire configuration for magnetic trapping, the gradient is not a<br />

124


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

good parameter, as the trap is harmonic at the bottom, with a gradient <strong>of</strong><br />

zero but a finite curvature. In non-central regions <strong>of</strong> the trap, the harmonic<br />

approximation fails (see Figure 2.5). But as all trapping parameters are given<br />

by the current through the wire and the bias field, one can use any combination<br />

<strong>of</strong> two parameters that only depend on Iw and Bb. The reasons why the bias<br />

field and the gradient are chosen as parameters are explained below. Additional<br />

bias fields <strong>of</strong> 3 G along the wire direction and perpendicular to the surface are<br />

applied to match the field <strong>of</strong> the U-wire with the field <strong>of</strong> the Z-wire and to<br />

further suppress Majorana spin flips. This trap holds ≈ 8 · 10 7 atoms at a<br />

temperature <strong>of</strong> 50 µK.<br />

In the Z-magnetic trap, the number <strong>of</strong> atoms slowly decays due to differ-<br />

ent mechanisms. To measure these decays, a simple procedure is used: after<br />

different times <strong>of</strong> magnetic trapping, the trap is changed back to a CMOT. If<br />

needed, the trap is first moved back towards the CMOT position by ramping<br />

the wire current and the bias field accordingly. Now, the fluorescence <strong>of</strong> the<br />

recaptured atoms in the CMOT can be detected. This provides an easy means<br />

to measure the lifetime <strong>of</strong> the magnetic trap.<br />

To find the optimum trapping parameters, different settings were examined.<br />

The trap depth and the compression are the important parameters. The trap<br />

depth and its position r0 are determined by the bias field Bb and the current<br />

through the wire Iw over Bb ∼ Iw/r0, where Bb is a direct measure <strong>of</strong> the<br />

trap depth. The compression is determined by the gradient which scales like<br />

∂B/∂r ∼ I/r 2 0 ∼ B 2 b /Iw. We want the trap to be as deep as possible to collect<br />

all atoms, but we are limited by the maximum current through the wire and<br />

the fact that our wire is separated from the chip surface, limiting the choice <strong>of</strong><br />

r0. Close to the surface we need a higher compression to prevent atoms from<br />

touching the surface and being removed from the trap. On the other hand, too<br />

high a gradient increases the chance <strong>of</strong> atoms not being able to adiabatically<br />

125


5.3. The Wire Magnetic Trap<br />

follow the field and leads to Majorana ‘spin flip’ losses. We thus need to find<br />

a compromise. For this we first used two different initial numbers <strong>of</strong> atoms in<br />

the same magnetic trap (Iw = 21.7 A, Bb = 39 G). The result is shown in Fig.<br />

5.9 (a) on a logarithmic scale. We also varied the trap parameters. The larger<br />

trap above was compared to a very shallow trap with a bias field <strong>of</strong> 13.7 G<br />

and a current <strong>of</strong> 7 A and a steep deep trap with 54 G and 31.1 A. In terms <strong>of</strong><br />

the gradient, we thus compare 350 G/cm with 134 G/cm and 467 G/cm. The<br />

results are shown in Fig. 5.9 (b) on a logarithmic plot.<br />

relative atom number<br />

1<br />

0.5<br />

0.3<br />

0.2<br />

0 5 10 15 20 25 30<br />

time / s<br />

(a)<br />

1<br />

0.5<br />

0.3<br />

0.1<br />

0.04<br />

0 5 10 15 20 25 30 35 40<br />

Figure 5.9: Examining the lifetime <strong>of</strong> different magnetic traps: decay <strong>of</strong> the<br />

number <strong>of</strong> trapped atoms on a logarithmic scale as a function <strong>of</strong> time. (a)<br />

same trap parameters, different starting populations <strong>of</strong> 7.4·10 7 atoms (circles),<br />

4.3 · 10 7 atoms (crosses). (b) different trap parameters: normal (o), shallow<br />

relative atom number<br />

time / s<br />

(x), steep (+). The lines are fits <strong>of</strong> equation (5.9) to the data.<br />

If we take into account density dependent effects, the loss rate changes from<br />

dN/dt = −α · N to dN/dt = −α · N − β · N 2 . The loss coefficient α describes<br />

constant losses like in radioactive decay, so the loss rate is proportional to<br />

the number <strong>of</strong> atoms and each atom has the same probability to be lost.<br />

Density dependent losses are described by β. The proportionality to N 2 is an<br />

approximation for a large number <strong>of</strong> atoms N. Strictly speaking, a density<br />

126<br />

(b)


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

dependent loss is proportional to the number <strong>of</strong> atoms N and to the number<br />

<strong>of</strong> remaining atoms N − 1 which can interact with the first atom. For large<br />

numbers, the approximation N · (N − 1) ≈ N 2 is valid. This rate equation can<br />

be solved analytically [Bro91], and still leads to a single exponential form with<br />

the single particle loss rate α in the power <strong>of</strong> the exponential and the starting<br />

particle number N0:<br />

N(t) =<br />

αN0<br />

(α + βN0) · e αt − βN0<br />

(5.9)<br />

If we fit this function to the data for the different traps, we find the following:<br />

first <strong>of</strong> all, the data <strong>of</strong> the first 0.5 seconds doesn’t fit to the model. During<br />

this time, the cloud <strong>of</strong> atoms touches the surface <strong>of</strong> the chip and atoms are<br />

lost at a high rate, which drops to zero once the cloud loses contact with the<br />

chip. This effect is not covered by the above model, and thus the data for<br />

times t ≤ 0.5 s was ignored. Second, from Fig. 5.9 (a) we can see that the<br />

atom number is not the decisive parameter. The single loss parameter α is<br />

negligible in both cases. If we look at the product <strong>of</strong> the density dependent<br />

loss parameter and the starting number <strong>of</strong> atoms, βN0, we have a parameter<br />

that is comparable to α. The value βN0 ≈ 0.1 s −1 is similar for both starting<br />

atom numbers. If we now look at Fig. 5.9 (b), we can see that the steeper and<br />

the shallower traps lose atoms at a higher rate than the intermediate trap. In<br />

the steep trap, the two body loss rate is slightly higher. This has to be related<br />

to the higher collision rate and thus the larger number <strong>of</strong> inelastic collisions<br />

which cause loss. In this trap, one particle losses can’t be neglected anymore<br />

and have to be taken into account with α ≈ 0.02 s −1 . This loss has to be<br />

attributed to ‘spin flips’ when the atoms are unable to adiabatically follow the<br />

changing magnetic field at the trap’s minimum. In the shallow trap the one<br />

particle loss parameter is slightly higher than in the other traps with α ≈ 0.05<br />

127


5.3. The Wire Magnetic Trap<br />

s −1 , while the two body losses βN0 ≈ 1 s −1 exceed the values <strong>of</strong> the other<br />

traps by an order <strong>of</strong> magnitude. Here the atoms are freely evaporating.<br />

In summary, we infer that the main losses are caused by three mechanisms:<br />

initially, a large number <strong>of</strong> atoms is lost by contact with the atom chip and<br />

later, collisions between atoms and the atoms’ inability to adiabatically follow<br />

the magnetic potential. In the shallow trap, the atoms can escape easier than<br />

in the other two traps and any single collision can lead to the escape <strong>of</strong> atoms.<br />

In the steep trap, the main loss mechanisms are ‘spin flips’ directly related to<br />

the high curvature <strong>of</strong> the trapping potential and a slightly increased number<br />

<strong>of</strong> inelastic collisions compared to the intermediate trap.<br />

The compressed magnetic trap<br />

Once we have trapped the atoms magnetically, this trap is compressed slowly<br />

to improve the scattering characteristics <strong>of</strong> the atoms. We deepen the trap<br />

and increase the gradient and by so doing we increase the density and the<br />

temperature <strong>of</strong> the atomic sample. This is done ‘adiabatically’ in the sense<br />

that the phase space density ρ does not change. In our set-up, we increase the<br />

current to Iw = 31.1 A and the bias field to Bb = 56.25 G in a time <strong>of</strong> 100 ms.<br />

As mentioned above, the set-up <strong>of</strong> the wire not being coincident with the<br />

surface is a limiting factor. We examined how far the trap can be compressed<br />

by measuring the recaptured atom number for variable bias fields at different<br />

currents. The results are plotted in Fig. 5.10, with the atom number as a<br />

function <strong>of</strong> the gradient. The magnetic film was taken into account, as 0.55<br />

mm above the wire with an effective current <strong>of</strong> 0.44 A. The current-carrying<br />

wire was taken as infinitely long with infinitesimal width. The dependence on<br />

the trap depth given by the bias field is obvious: for the same gradient a higher<br />

current generates a deeper trap, as it needs a higher bias field. A deeper trap<br />

is capable <strong>of</strong> trapping more atoms. The influence <strong>of</strong> the gradient shows the<br />

128


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

number <strong>of</strong> atoms / 10 7<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

200 400 600 800 1000 1200 1400 1600<br />

Gradient / Gcm −1<br />

Figure 5.10: Compressing the magnetic trap: atom number versus magnetic<br />

field gradient, for wire currents <strong>of</strong> 31.1 A (crosses), 20.5 A (circles). The<br />

optimum gradient is about 850 G/cm.<br />

best trapping for values <strong>of</strong> around 850 G/cm. This optimum value is nearly<br />

independent <strong>of</strong> the trap depth.<br />

This Z-trap being <strong>of</strong> the IP-type has no zero crossing <strong>of</strong> the field. This<br />

difference <strong>of</strong> the field’s minimum from zero can be measured by changing the<br />

strength <strong>of</strong> the bias field along the wire direction and measuring the losses.<br />

Once this field is scanned to a value that compensates the intrinsic <strong>of</strong>fset field<br />

<strong>of</strong> the Z-wire, a sudden decrease in the number <strong>of</strong> atoms indicates the losses<br />

due to the spin flips which can occur in the vicinity <strong>of</strong> the field’s zero-crossing.<br />

The data presented in Fig. 5.11 was taken with a wire current <strong>of</strong> 31.1 A<br />

and a perpendicular bias field <strong>of</strong> 61.6 G, leading to a nominal gradient <strong>of</strong> 611<br />

G/cm ignoring the film. The drop in the atom number at a field <strong>of</strong> about 3<br />

G indicates that here the magnetic fields cancel each other. The wire alone<br />

creates a bias field <strong>of</strong> 3 G which is counteracted by the field <strong>of</strong> the external<br />

bias coils.<br />

The increase up to 3 G is due to the slight changes in trapping depth and<br />

gradient during the scan. The magnetic field changes the energy <strong>of</strong> the state<br />

129


5.3. The Wire Magnetic Trap<br />

number <strong>of</strong> atoms / 10 7<br />

3<br />

2.9<br />

2.8<br />

2.7<br />

2.6<br />

2.5<br />

2.4<br />

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5<br />

bias field along wire / G<br />

Figure 5.11: Atoms are lost from the trap as the magnetic field in the direction<br />

<strong>of</strong> the wire cancelled out. The drop in the atom number at a bias field between<br />

3 G and 3.5 G (arrow) is caused by Majorana ‘spin flips’, indicating the value<br />

<strong>of</strong> the magnetic field component along the wire, induced by the Z-wire itself.<br />

The dotted lines are to guide the eye.<br />

|F, mF 〉 by ∆E due to the Zeeman effect (see equation (2.37)). The splitting<br />

between the adjacent mF -levels at a field <strong>of</strong> 3 G corresponds to a frequency<br />

f = ∆E/h = 2.1 MHz. In our case, in the compressed stage <strong>of</strong> the magnetic<br />

trap, we apply a bias field <strong>of</strong> 1.6 G along the wire. The resulting effective<br />

field <strong>of</strong> 1.4 G leads to an energy difference <strong>of</strong> 1.96 MHz between the trapped<br />

|F = 2, mF = 2〉 and the untrapped |F = 2, mF = 0〉 state. This <strong>of</strong>fset has to<br />

be taken into account in the next paragraph.<br />

Radi<strong>of</strong>requency radiation can be used to cool the atoms by evaporation. It<br />

can also be used to examine the ensemble by a spectroscopic means. Ramping<br />

down the frequency to different end values fi leads to different numbers <strong>of</strong><br />

atoms remaining in the trap Ni. Atoms with energies higher than Ei = hfi<br />

will be ejected from the trap and not be detected in a recapture fluorescence<br />

measurement. Figure 5.12 shows how this number decreases when the fre-<br />

quency <strong>of</strong> the radiation is linearly ramped down in 5 s from 25 MHz to 3.5<br />

130


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

MHz. The cut-<strong>of</strong>f energy is given in thermal units E ∝ kBT with correspond-<br />

ing temperatures in the low mK regime to allow the relation to the following<br />

calculations. The data was taken from two scans that differed slightly in the<br />

fraction <strong>of</strong> remaining atoms<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4<br />

cut−<strong>of</strong>f energy in thermal units k B ⋅mK<br />

Figure 5.12: Atoms remaining in the trap as a function <strong>of</strong> the cut-<strong>of</strong>f energy<br />

<strong>of</strong> the RF radiation. This energy is equivalent to the trap depth and is given<br />

in thermal units.<br />

rate <strong>of</strong> the ramp. The trap is created by the same magnetic fields and wire<br />

current as in the previous paragraphs. The data is corrected for an RF-<strong>of</strong>fset<br />

<strong>of</strong> 2 MHz.<br />

If we now take the differences <strong>of</strong> the atom number ∆N and the energy for<br />

adjacent frequency values, ∆E = h(fi+1 − fi), we obtain an estimate <strong>of</strong> the<br />

number <strong>of</strong> atoms per unit energy interval as a function <strong>of</strong> their energy ñ(E) :<br />

∆N<br />

∆E<br />

≈ dN<br />

dE<br />

= ñ(E) (5.10)<br />

This can be plotted against the energy, where the lower frequency <strong>of</strong> the dif-<br />

ference equation gives the value for the abscissa. Figure 5.13 shows such a<br />

plot, using the data shown in Fig. 5.12. To guide the eye, two additional lines<br />

are shown in the plot. The dotted line is the classical Maxwell-Boltzmann<br />

131


5.3. The Wire Magnetic Trap<br />

distribution for an ideal gas <strong>of</strong> N atoms at temperature T = 300 µK [Alo92]:<br />

dN<br />

dE =<br />

2πN<br />

(πkBT ) 3/2<br />

√ −E<br />

k Ee B T (5.11)<br />

The solid line is for a Bose-Einstein distribution with temperature T = 120 µK.<br />

With gi being the density <strong>of</strong> states, the distribution here is given by [Bec85]:<br />

ñ(Ei) =<br />

gi<br />

e (µ−Ei)/(kBT ) − 1<br />

(5.12)<br />

Here the quantisation <strong>of</strong> the energy becomes important. Depending on the trap<br />

geometry, gi states can be degenerate at the same energy Ei. The chemical<br />

potential µ is the energy that is needed to add an atom to the ensemble while<br />

keeping the volume and temperature constant. For a harmonic trap, the degree<br />

<strong>of</strong> degeneracy depends on the number <strong>of</strong> dimensions. In a one dimensional trap,<br />

there is no degeneracy. In a 2D trap, the energy Ei is i+1-fold degenerate, and<br />

in the 3D case it is gi = 1/2(i + 1)(i + 2). For the plot an isotropic, harmonic<br />

3D trap with a trapping frequency <strong>of</strong> 200 Hz was chosen. The first 100,000<br />

levels were taken into consideration, using every 10th for the calculation. The<br />

chemical potential µ <strong>of</strong> the Bose-Einstein distribution and the atom number <strong>of</strong><br />

the Maxwell-Boltzmann distribution were chosen to roughly fit the amplitude.<br />

As the occupation was not calculated for every single level the physical meaning<br />

<strong>of</strong> the chemical potential µ was lost anyway. The temperatures were chosen<br />

to roughly fit the form <strong>of</strong> the data. We can see that the Maxwell-Boltzmann<br />

distribution does not fit the actual distribution <strong>of</strong> the atoms in the trap. Using<br />

only atom number and temperature, it can not be modified to fit both the<br />

rising edge and the decaying tail. The Maxwell-Boltzmann distribution has<br />

to be modified to take the confining potential into account, which is properly<br />

done in the evaporation model <strong>of</strong> [Met99].<br />

We can take the data points shown in Fig. 5.12 as a function <strong>of</strong> RF-<br />

frequency νRF against atomic loss. We rescale and normalise the RF frequency<br />

132


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

atom number per energy interval<br />

0.1<br />

0.09<br />

0.08<br />

0.07<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

0 0.5 1 1.5<br />

energy in thermal units k B ⋅mK<br />

Figure 5.13: Distribution <strong>of</strong> the number <strong>of</strong> atoms in a compressed magnetic<br />

trap per energy interval. The points are derived from the results shown in<br />

Fig. 5.12. To guide the eye, lines have been added: the dotted line shows<br />

the Maxwell-Boltzmann distribution <strong>of</strong> an ideal gas, the solid line shows the<br />

Bose-Einstein distribution for a 3D isotropic harmonic trap. The energy is<br />

expressed in thermal units.<br />

into units <strong>of</strong> an unknown temperature. This gives us the atom loss as a function<br />

<strong>of</strong> the truncation parameter η (used in the evaporation model, see section<br />

2.3.1). This parameter is defined by the energy E = η · kBT at which the trap<br />

is cut <strong>of</strong>f. Thus, we get<br />

η = E<br />

kBT<br />

h νRF − ν<strong>of</strong>fset<br />

=<br />

kB T<br />

(5.13)<br />

where ν<strong>of</strong>fset is the RF-<strong>of</strong>fset frequency which is defined by the parallel bias<br />

field. We can now use the relation <strong>of</strong> remaining atoms N ′ /N and the truncation<br />

parameter <strong>of</strong> equation (2.51)<br />

N ′<br />

N<br />

= γ(ξ + 3/2, η)<br />

Γ(ξ + 3/2)<br />

(5.14)<br />

where γ(ξ + 3/2, η) is the lower incomplete Gamma-function as a fitting func-<br />

tion. Here the data was fitted with the <strong>of</strong>fset frequency and the temperature<br />

133


5.3. The Wire Magnetic Trap<br />

<strong>of</strong> the ensemble as free parameters, for the case <strong>of</strong> an harmonic (ξ = 3/2) and<br />

a quadrupole (ξ = 3) trap. This is shown in Fig. 5.14. For the harmonic trap,<br />

η<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

fraction <strong>of</strong> remaining atoms<br />

Figure 5.14: Truncation parameter η as a function <strong>of</strong> the atoms remaining in<br />

the trap after reducing the trap depth. The results <strong>of</strong> Fig. 5.12 have been<br />

used to calculate the parameter for the case <strong>of</strong> a harmonic trap (circles) and<br />

a linear trap (crosses). Solid lines: fitted functions for the respective trap (see<br />

equation (5.14)).<br />

we obtain an <strong>of</strong>fset between the trap bottom and the field zero <strong>of</strong> 2.36 MHz and<br />

a temperature <strong>of</strong> T = 142 µK. In the case <strong>of</strong> the linear (quadrupole) trap, the<br />

<strong>of</strong>fset frequency is much smaller with 790 kHz and the temperature is slightly<br />

smaller with T = 110 µK. When we allow the trap geometry parameter ξ as a<br />

free parameter in the fit as well, then the results become inconsistent with all<br />

a priori knowledge we have <strong>of</strong> the trap: with fitted results <strong>of</strong> ξ = 3.8 we would<br />

have a less than linear rising trap. The fit then results in an <strong>of</strong>fset frequency<br />

<strong>of</strong> 1.5 Hz where we would have had to see massive losses due to Majorana<br />

spin flips. When we compare the confidence levels <strong>of</strong> the three fits, all <strong>of</strong> them<br />

reach values <strong>of</strong> r 2 > 0.995. Now, there are arguments for assuming a linear<br />

trap and arguments that would propose a harmonic trap. The geometry with<br />

a Z-shaped wire leads to a trapping potential which is harmonic at the bot-<br />

134


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

tom, for low energies, while it is well approximated by a linear trap for higher<br />

frequencies (ignoring the influence <strong>of</strong> the magnetic film) (see also Fig. 5.16).<br />

A strong argument for the harmonicity <strong>of</strong> the trap is the <strong>of</strong>fset <strong>of</strong> the trap <strong>of</strong><br />

2.36 MHz, which is in good agreement with the previously determined <strong>of</strong>fset<br />

from the above measurements on the bias field along the wire (Fig. 5.11).<br />

Time-<strong>of</strong>-flight measurements <strong>of</strong> the atom cloud’s expansion have resulted in a<br />

temperature <strong>of</strong> about 150 µK, close to the temperature we deduce from the fit<br />

for a harmonic potential.<br />

We see that we can not clearly specify the power-law <strong>of</strong> the trap. For<br />

either a purely linear or for a purely harmonic trap the results are not fully<br />

satisfying. This is closely related to the geometry <strong>of</strong> the wire which leads to two<br />

regimes, one scaling with the square <strong>of</strong> the distance from the wire, the other<br />

scaling linearly. From the above, we can conclude that our atomic cloud after<br />

compression is in the middle region, with a tendency towards the harmonic<br />

trapping regime.<br />

After this compression stage, we typically work with 8 · 10 7 atoms at a<br />

temperature <strong>of</strong> 160 µK. The trap is 560 µm from the chip surface, with a<br />

gradient <strong>of</strong> ∼ 510 G/cm (ignoring the influence <strong>of</strong> the film). A bias field in the<br />

direction <strong>of</strong> the wire <strong>of</strong> 1.6 G is applied, and the trap bottom for the atoms in<br />

the mF = 2 state is about 2 MHz above the level for the mF = 0 state or zero<br />

magnetic-field energy.<br />

5.3.1 Evaporation and BEC<br />

Once the trap has been compressed, we begin the cooling by evaporation. This<br />

is done by applying a so-called “RF-knife”. By passing a radi<strong>of</strong>requency cur-<br />

rent along the end wires <strong>of</strong> the chip, we create radiation <strong>of</strong> the same frequency<br />

at the trap. This causes atoms which have an energy that matches the energy<br />

135


5.3. The Wire Magnetic Trap<br />

corresponding to the applied frequency to undergo spin flip transitions into<br />

magnetically untrapped states. In this way we can remove in a controlled way<br />

atoms that have an energy higher than a threshold energy, leaving only low<br />

energy atoms in the trap. Because <strong>of</strong> this, the RF-aided evaporation is a tech-<br />

nique <strong>of</strong> forced evaporative cooling. The underlying mechanisms are explained<br />

in section 2.3. This part is divided into two: one part covers the evaporation<br />

process and the related important properties <strong>of</strong> the cold atoms, and the other<br />

part shows the results once the BEC was reached.<br />

Evaporation<br />

Here results are presented that were obtained before the chip’s connectors were<br />

modified, and with a vacuum an order <strong>of</strong> magnitude worse than is presented in<br />

section 4.3. This vacuum problem was caused by the heating <strong>of</strong> the connectors<br />

and wires. These will be compared with recent results [Whi05] <strong>of</strong> the BEC. The<br />

results showing the BEC were taken with a substitute main laser, a commercial<br />

high power laser diode [E 33]. The stabilisation set-up remained unchanged.<br />

The most important parameter that characterises the atomic sample is the<br />

phase space density ρ = n0 · λ 3 dB . Here the peak density n0 is given by the<br />

the number <strong>of</strong> atoms N divided the volume which can be approximated by<br />

the measured size <strong>of</strong> the atomic cloud, V = 1<br />

√ 3 σaxial · σ<br />

2π 2 radial , assuming a<br />

3D gaussian distribution. The de Broglie wavelength can be calculated from<br />

equation (2.1) once the temperature <strong>of</strong> the atomic cloud is known.<br />

In the following, the data from the BEC includes TOF measurements on the<br />

temperature for expansion times <strong>of</strong> 5, 10, and 15 ms. The cloud sizes σi were<br />

determined by the intercept <strong>of</strong> the linear regression and by in-trap absorption<br />

images. The absolute atom number N is the mean value taken from three<br />

ballistic expansions. The data taken earlier provided no direct determination<br />

136


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

<strong>of</strong> the temperature but a means to achieve this is presented in Appendix B.<br />

Here the atomic spread σ was taken from in-trap absorption measurements.<br />

Different ramps were used when ramping down the RF frequency. They<br />

are shown in Fig. 5.15. The markers on the lines depict the times and the<br />

value <strong>of</strong> the frequency at which data was taken for the following comparison.<br />

The solid line marked with circles is the actual ramp that is now used to reach<br />

Bose-Einstein condensation. This is a logarithmic ramp from a frequency <strong>of</strong><br />

20 MHz to 0.78 MHz in 10 seconds. The trap parameters also differed for the<br />

RF frequency / MHz<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 2 4 6 8 10 12<br />

time / s<br />

Figure 5.15: The different logarithmic ramps <strong>of</strong> the radio frequency. The ramp<br />

that leads to BEC is marked by circles.<br />

measurements. The successful trap was run at a wire current <strong>of</strong> Iw = 31.1 A,<br />

with bias fields <strong>of</strong> By = 1.6 G along and Bx = 54 G perpendicular to the wire.<br />

Figure 5.16 shows the calculated magnetic field as a function <strong>of</strong> the distance<br />

from the chip surface. Here the Z-shape and the magnetic film were taken<br />

into account, and the finite width <strong>of</strong> the wires was neglected. The dotted<br />

line shows a quadratic fit to the trap’s base. It leads to a trap frequency <strong>of</strong><br />

ν = 217 Hz. The magnetic field <strong>of</strong>fset <strong>of</strong> this plot is much larger than is seen in<br />

the experiment. This has to be explained by the calculation which takes into<br />

account the Z-shape <strong>of</strong> the wire and the bias field along the wire, but not any<br />

137


5.3. The Wire Magnetic Trap<br />

other sources <strong>of</strong> fields in this direction, which can counteract these fields. The<br />

magnetic field / G<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

distance from surface / mm<br />

Figure 5.16: Calculated magnetic field <strong>of</strong> the trap that leads to quantum<br />

degeneracy, plotted against the distance from the chip (solid line). Harmonic<br />

approximation (dotted line) leads to a trap frequency <strong>of</strong> 217 Hz.<br />

unsuccessful first attempt to reach BEC presented here used a wire current <strong>of</strong><br />

Iw = 19 A and a perpendicular bias field <strong>of</strong> Bx = 25 G. This leads to a trap<br />

frequency in the harmonic approximation <strong>of</strong> ν = 100 Hz. Going back to Fig.<br />

5.15 its RF ramp is shown by a solid line marked with crosses.<br />

The evolution <strong>of</strong> the temperature during the evaporation process is shown<br />

in Fig. 5.17 (a). The marking is the same as for the RF ramps in Fig. 5.15:<br />

the trap that leads to a BEC is marked by circles. The older trap which did<br />

not reach degeneracy is marked by crosses. Solid lines interpolate between the<br />

points. The influence <strong>of</strong> the compression on the temperature can be seen on<br />

this graph: the higher compressed trap has an initial temperature <strong>of</strong> nearly one<br />

order <strong>of</strong> magnitude higher than the less steep trap. Both ramps lead to final<br />

temperatures <strong>of</strong> the order <strong>of</strong> Tfinal ≈ 300 nK, close to the recoil temperature<br />

<strong>of</strong> 87 Rb.<br />

Knowing the temperatures, we can calculate the de Broglie wavelengths<br />

using equation (2.1). These are shown in Fig. 5.17 (b), where the marking<br />

138


temperature / µ K<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

(a)<br />

Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

10<br />

0 5 10 15 20 25<br />

−1<br />

cut <strong>of</strong>f frequency / MHz<br />

atomic de Broglie wavelength / m<br />

10 −6<br />

10 −7<br />

(b)<br />

10<br />

0 2 4 6 8 10 12 14 16 18 20<br />

−8<br />

cut <strong>of</strong>f frequency / MHz<br />

Figure 5.17: (a) Temperature <strong>of</strong> the atomic cloud, using the approximation<br />

introduced in the text. (b) De Broglie wavelength <strong>of</strong> the atoms. Both graphs<br />

are on a logarithmic scale as a function <strong>of</strong> the end value <strong>of</strong> the radio-frequency.<br />

Circles: the trap that leads to BEC; +: the older trap. The lines are interpo-<br />

lations between the points.<br />

is the same as above. In the shallow, older trap we measure a slightly longer<br />

wavelength than in the trap that leads to BEC. They are <strong>of</strong> the same order<br />

<strong>of</strong> magnitude and with λdB ≈ 300 nm slightly shorter than the de Broglie<br />

wavelength <strong>of</strong> the recoil temperature.<br />

An experimental challenge, and how it can be overcome, becomes apparent<br />

when the atomic density n0 is plotted against the cut-<strong>of</strong>f frequency. Here for<br />

the newer set <strong>of</strong> data the density was calculated once using the atomic cloud<br />

size obtained directly from the in-trap imaging and once using the extrapo-<br />

lated value for the size that comes from the linear regression which is used<br />

to determine the temperature by ballistic expansion. The result <strong>of</strong> this dual<br />

approach, together with the result <strong>of</strong> the older data, is shown in Fig. 5.18 (a),<br />

using the same markings to depict the different measurements.<br />

When we compare the different results, we see that the one that relies<br />

on the in-trap imaging (circles) shows a breakdown in the density for end-<br />

139


5.3. The Wire Magnetic Trap<br />

atomic peak density in m −3<br />

10 18<br />

10 17<br />

10 16<br />

0 2 4 6 8 10 12 14 16 18 20<br />

cut <strong>of</strong>f frequency / MHz<br />

(a)<br />

phase space density<br />

10 −1<br />

10 −2<br />

10 −3<br />

10 −4<br />

10 −5<br />

10 −6<br />

10 −7<br />

0 2 4 6 8 10 12 14 16 18 20<br />

cut <strong>of</strong>f frequency / MHz<br />

Figure 5.18: (a) Density <strong>of</strong> the atomic cloud, (b) Phase space density <strong>of</strong> the<br />

atomic cloud, on logarithmic scales as a function <strong>of</strong> the end value <strong>of</strong> the radio-<br />

frequency. Circles: the trap that leads to BEC using the spread <strong>of</strong> the in-trap<br />

imaging; squares: the same trap using the spread calculated from the ballistic<br />

expansion; +: the older trap. The lines are interpolations between the points.<br />

frequencies smaller than 2 MHz. On the other hand, the results that use<br />

the extrapolation from the ballistic expansion (squares) are much smaller for<br />

higher frequencies. We can relate this to the problem <strong>of</strong> resolution, when the<br />

trap reaches a size comparable to the resolution <strong>of</strong> the system. In our case,<br />

for the points with the smallest frequency, we have a trap size <strong>of</strong> less than<br />

10 pixels on the camera. Here possible errors in the focussing and imaging<br />

will lead to a large overestimation <strong>of</strong> the atom cloud’s size, which in turn will<br />

lead to too large a volume and thus to a systematically too small density.<br />

It is thus better to rely on the ballistic expansion to achieve values for the<br />

atomic spread. The problem <strong>of</strong> the breakdown <strong>of</strong> the density can also be<br />

seen on the graph for the shallow, older trap. The threshold <strong>of</strong> the breaking<br />

down is at a slightly smaller frequency. As this trap is less confining, this is<br />

consistent with the above explanation. Unfortunately it is impossible to use<br />

the different approach for this set <strong>of</strong> data, as here only in-trap images were<br />

taken. The influence will be discussed further when the phase space densities<br />

140<br />

(b)


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

are compared. Apart from systematical errors in the imaging the small size<br />

<strong>of</strong> the trap also leads to large statistical errors as can be seen in the error<br />

bars. Here an absolute minimum uncertainty <strong>of</strong> one pixel to either side was<br />

assumed together with a five percent relative uncertainty. The absolute error<br />

dominates for small cut <strong>of</strong>f frequencies in the measurements that rely on the<br />

in-trap imaging, and is absent in the measurement that uses a linear regression<br />

for determination <strong>of</strong> the trap size. It is favourable to use a linear regression<br />

rather than the in-trap imaging technique for small trap sizes.<br />

The most important figure <strong>of</strong> merit <strong>of</strong> any experiment that tries to reach<br />

quantum degeneracy is the phase space density. Bose-Einstein condensation<br />

requires a value <strong>of</strong> the phase space density <strong>of</strong> the order <strong>of</strong> one. The low<br />

temperatures <strong>of</strong> the atoms have a major effect on this density, as they increase<br />

the de Broglie wavelength which appears as a cube in the definition in equation<br />

(2.49). The results <strong>of</strong> our experiments are shown in Fig. 5.18 (b), with the<br />

same markings as above. We see that the trap that finally leads to BEC and<br />

the older trap that was unsuccessful show a very similar behaviour and a large<br />

increase in the PSD towards the end <strong>of</strong> the evaporation. The lowest frequency<br />

<strong>of</strong> the RF radiation in Fig. 5.18 (b) is 850 kHz, which corresponds to a value<br />

<strong>of</strong> ρ ≈ 0.03. The phase transition was observed at a frequency <strong>of</strong> 763 kHz. It<br />

is difficult to extrapolate the PSD for that frequency from Fig. 5.18 (b), but<br />

it is clearly very much greater than 0.03. The trap that did not lead to BEC<br />

reached a maximum value <strong>of</strong> ρ ≈ 0.01 for a lowest frequency <strong>of</strong> about 1.1 MHz.<br />

Fig. 5.18 (b) also shows the main reason why the older trap did not reach<br />

quantum degeneracy. When we ignore the problems <strong>of</strong> determining the tem-<br />

perature and the subsequent uncertainties, it appears as if the lowest cut-<strong>of</strong>f<br />

frequency <strong>of</strong> the RF knife was chosen to be too high. However, this is not the<br />

only reason. It is also important that the trap was shallower by a factor <strong>of</strong> two.<br />

This led to a density that was lower by roughly the same factor (Fig. 5.18, (a)).<br />

141


5.3. The Wire Magnetic Trap<br />

All figures here show the resulting problem with the shallow trap: the forced<br />

evaporation started with RF frequencies <strong>of</strong> about 8 MHz. For higher frequen-<br />

cies neither the temperature nor the phase space density change appreciably.<br />

A lower atomic density also reduces the elastic collision rate in the trap. This<br />

increases the time needed for rethermalisation and can be a further reason why<br />

quantum degeneracy was not reached. The currents running through the chip<br />

wires in the unsuccessful attempt were low enough not to heat the chip and<br />

reduce the lifetime <strong>of</strong> the trap by outgassing. The problem <strong>of</strong> losing atoms to<br />

the surface initially in the magnetic trap is a further reason for the failure <strong>of</strong><br />

these attempts.<br />

Bose-Einstein Condensation<br />

Figure 5.19 shows absorption images after ramping the cut-<strong>of</strong>f frequency from<br />

20 MHz down to 0.78 MHz in a logarithmic sweep over 10 seconds. The phase<br />

transition to quantum degeneracy is clear. Here the wire current was also<br />

ramped from 31.1 A initially down to 24.9 A. This leads to a further increase<br />

in the trapping frequency and thus in the collision rates. A bias field <strong>of</strong> 56.25<br />

G was applied perpendicular to the wire. A second bias field along the wire<br />

had a strength <strong>of</strong> 1.6 G. We can see the phase transition starting at a cut-<strong>of</strong>f<br />

frequency <strong>of</strong> 763 kHz. The trap was measured to be 350 µm from the surface;<br />

the calculated value ignoring the magnetic field <strong>of</strong> the film is 280 µm. The<br />

trap’s gradient, also ignoring the magnetic field <strong>of</strong> the film, is 635 G/cm. In<br />

the images shown here, the atom number goes from 9 · 10 4 atoms at 771 kHz<br />

down to 4 · 10 4 atoms at 755 kHz. In the intermediate step we find 7.6 · 10 4<br />

atoms.<br />

The pr<strong>of</strong>iles <strong>of</strong> the atomic clouds for each end frequency are shown in<br />

Figs. 5.20 and 5.21. Here the signal was summed over one axis and plotted<br />

against the other. The plots show the data points and fitted functions. For<br />

142


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

771 kHz 763 kHz 755 kHz<br />

Figure 5.19: The phase transition from a thermal cloud to a BEC can be seen<br />

by absorption images <strong>of</strong> the atomic cloud for different final radi<strong>of</strong>requencies<br />

(771 kHz, 763 kHz, 755 kHz). Each image shows an area <strong>of</strong> 1.26 × 1.17 mm 2 .<br />

integrated atomic density axis in a.u.<br />

0.16<br />

0.15<br />

0.14<br />

0.13<br />

0.12<br />

0.11<br />

0.1<br />

0.09<br />

(a)<br />

0.08<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

vertical position in mm<br />

integrated atomic density axis in a.u.<br />

0.16<br />

0.15<br />

0.14<br />

0.13<br />

0.12<br />

0.11<br />

0.1<br />

0.09<br />

(b)<br />

0.08<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

vertical position in mm<br />

integrated atomic density axis in a.u.<br />

0.16<br />

0.15<br />

0.14<br />

0.13<br />

0.12<br />

0.11<br />

0.1<br />

0.09<br />

(c)<br />

0.08<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

vertical position in mm<br />

Figure 5.20: Vertical pr<strong>of</strong>ile <strong>of</strong> the atom cloud for end frequencies <strong>of</strong> (a) 771,<br />

(b) 763 and (c) 755 kHz. Solid lines are fitted Gaussians for 771 and 763<br />

kHz (thermal clouds). For 755 kHz the sum <strong>of</strong> a Gaussian and a quadratic<br />

Thomas-Fermi pr<strong>of</strong>ile (BEC) was fitted. The dotted line is the Gaussian ther-<br />

mal contribution.<br />

the thermal clouds, the data was fitted to a Gaussian distribution, while the<br />

pr<strong>of</strong>iles <strong>of</strong> the condensed atoms were fitted to the quadratic function <strong>of</strong> the<br />

Thomas-Fermi pr<strong>of</strong>ile. For Fig. 5.21 (b) the phase transition is clearly visible.<br />

The underlying Gaussian <strong>of</strong> the thermal cloud is shown by the dotted line in<br />

that plot. Figure 5.21 (c) shows deviations and oscillations in the wings <strong>of</strong> the<br />

Gaussian. These are also clearly visible in Fig. 5.19, 755 kHz. The structure<br />

in the wings indicates that we are looking at a surrounding thermal cloud here,<br />

and see a super-imposed pattern <strong>of</strong> the light being diffracted by the BEC.<br />

143


5.3. The Wire Magnetic Trap<br />

integrated atomic density axis in a.u.<br />

0.18 (a)<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0 0.2 0.4 0.6 0.8 1<br />

horizontal position in mm<br />

integrated atomic density axis in a.u.<br />

0.18 (b)<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0 0.2 0.4 0.6 0.8 1<br />

horizontal position in mm<br />

integrated atomic density axis in a.u.<br />

0.18 (c)<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0 0.2 0.4 0.6 0.8 1<br />

horizontal position in mm<br />

Figure 5.21: Horizontal pr<strong>of</strong>ile <strong>of</strong> the atom cloud for end frequencies <strong>of</strong> (a) 771,<br />

(b) 763 and (c) 755 kHz. The solid line is a fitted Gaussian for 771 kHz. For<br />

763 and 755 kHz the sum <strong>of</strong> a Gaussian and a quadratic Thomas-Fermi pr<strong>of</strong>ile<br />

(BEC) were fitted. The dotted lines are the Gaussian thermal contribution.<br />

There is an inconsistency in the trap bottom, which has been measured to<br />

be about 0.5 G or a frequency <strong>of</strong> 350 kHz. Above in section 5.3 we found that<br />

for a trap with a wire current <strong>of</strong> 31.1 A, a perpendicular bias field <strong>of</strong> 61.6 G<br />

and a parallel bias field <strong>of</strong> 1.6 G, the trap has an <strong>of</strong>fset field <strong>of</strong> 3 G, while a<br />

calculation would lead to an <strong>of</strong>fset field <strong>of</strong> 21.2 G, so that our calculation is<br />

too high by 18 G absolute. When calculating the fields with the parameters<br />

used for the BEC, we see that the ramping down <strong>of</strong> the current as in the<br />

above reduces the <strong>of</strong>fset field to 16 G, which is still too high by about 15.5 G.<br />

As well, the calculation has its minimum at 200 µm from the trap, less than<br />

the measured 350 µm. Part <strong>of</strong> the discrepancy between measurement and<br />

calculation has to be related to the difficulties <strong>of</strong> adjusting the bias field to be<br />

perpendicular to the wire and adjusting the edge <strong>of</strong> the magnetic film to be<br />

parallel to the wire. Any angle between them leads to additional parallel fields.<br />

If we incorporate this in the calculations, we see that the largest influence is<br />

due to possible misalignments between the wire and the bias field, outweighing<br />

the influence <strong>of</strong> the film by a factor <strong>of</strong> about 5. In the linear regime <strong>of</strong> small<br />

angles, it contributes to roughly 1 G per degree away from the perpendicular.<br />

If we allow an angular error between the chip and the bias field <strong>of</strong> 5 degrees,<br />

144


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

we can reduce the difference between theory and measurement by 6 G. An<br />

angular misplacement <strong>of</strong> 10 degrees would account for nearly all the observed<br />

differences, but appears rather large and should be noticeable “by eye” in the<br />

experimental set-up. It was later seen that the substrate with the magnetic<br />

film moved by about 50 µm due to unevenly cured epoxy [Hal06]. Thus it is<br />

not clear where the discrepancy originates, but it seems to be rather constant<br />

in absolute values, and explainable with geometric displacements.<br />

5.4 The Permanent Magnetic Trap<br />

A film-based MOT<br />

As a very early result <strong>of</strong> the implementation <strong>of</strong> the magnetic film, we were able<br />

to show that the film is sufficient to create one <strong>of</strong> the basic magnetic fields for<br />

a magneto-optical trap, similar to [Ven02]. This was done purely qualitatively,<br />

and has to be seen as an “in principle” demonstration. Figure 5.22 shows<br />

the fluorescence from this MOT, where the trapping field is created solely by<br />

the film, a homogeneous bias field and the confining field from the end wires,<br />

running anti-parallel currents to create a quadrupole potential. This picture is<br />

an integration over 5 shots, each corrected for the background. On the top the<br />

reflection from the chip is visible. As the camera resolution was not calibrated,<br />

no measure <strong>of</strong> the atom number or temperature was performed. This result was<br />

very promising. Not only were we able to trap atoms with the film, the trap<br />

also was strong enough to be detected with fluorescence detection, even though<br />

for signal-to-noise reasons it was integrated over several shots. The reason for<br />

the low atom number lies in the small volume <strong>of</strong> the quadrupole potential, and<br />

thus the velocity range for capture is small. The small number <strong>of</strong> atoms in the<br />

film MOT was insufficient for further experiments, and highlights the need for<br />

a large MOT from where the atoms are then transfered to the magnetic trap.<br />

145


5.4. The Permanent Magnetic Trap<br />

Figure 5.22: Fluorescence <strong>of</strong> atoms in a MOT created by the permanent mag-<br />

netic film, integrated over 5 images which were corrected for the background.<br />

To align the external coils for the large MOT and the current direction in the<br />

wire magnetic trap, one needs to know the direction <strong>of</strong> the effective current<br />

in the film. This knowledge can be deduced from the polarisation <strong>of</strong> the film-<br />

MOT beams once the film MOT appears, and the direction <strong>of</strong> the current in<br />

the wire for magnetic trapping can be matched to the direction <strong>of</strong> the effective<br />

current <strong>of</strong> the magnetic film.<br />

The film-based magnetic trap<br />

The first experiments that were performed with this trap involved transfering<br />

atoms from the “Z”-wire magnetic trap to the permanent magnetic film trap.<br />

The trapping potential there is created by the magnetic fields <strong>of</strong> the film, a<br />

homogenous bias field perpendicular to it and the field <strong>of</strong> the two end wires<br />

with currents running through them in a parallel direction. It is not possible<br />

to transfer a large number <strong>of</strong> atoms to the magnetic film trap directly without<br />

an intermediate RF evaporation cooling stage in the wire. The reason for this<br />

is the constant effective current <strong>of</strong> the film. While with the two degrees <strong>of</strong><br />

freedom <strong>of</strong> a variable wire current and bias field it is possible to change the<br />

146


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

B,I<br />

B',<br />

d<br />

B depth<br />

I wire<br />

time<br />

trap distance<br />

from chip<br />

B': gradient<br />

Figure 5.23: Schematical view <strong>of</strong> the ramps for wire current and bias field<br />

to transfer the atoms from the wire trap to the film trap. Additionally, the<br />

development <strong>of</strong> the trap position from the surface and its gradient are sketched.<br />

trap depth and gradient independently, with the fixed effective current and a<br />

variable bias field these two properties are connected. For mode matching, it<br />

is possible to choose the bias field such that the trap based on the magnetic<br />

film has the same gradient as the one based on the wire (for 21 A this is<br />

approximately 100 G/cm), but the depth <strong>of</strong> the film based trap is reduced.<br />

Without intermediate cooling, the atoms in the wire trap would be too hot to<br />

stay in the film trap and would spill over, making the capture <strong>of</strong> a large atom<br />

number impossible. To avoid this, a stage <strong>of</strong> RF cooling down to about 50 µK<br />

was applied, with a ramp <strong>of</strong> 10 seconds and an end frequency <strong>of</strong> about 2 MHz.<br />

To then transfer the atoms to the magnetic film trap, both the wire current<br />

and the bias field were ramped down, as shown schematically in Fig. 5.23.<br />

This led to a trap <strong>of</strong> 8 · 10 6 atoms roughly 220 µm below the surface. By<br />

turning <strong>of</strong>f the current in the end wires, a ballistic expansion along the wire<br />

can be detected. Some results are shown in Fig. 5.24.<br />

The density <strong>of</strong> the atomic sample was summed up perpendicular to the wire<br />

to result in a density pr<strong>of</strong>ile. Unfortunately, the wire axis does not completely<br />

147


5.4. The Permanent Magnetic Trap<br />

0 ms<br />

8 ms<br />

16 ms<br />

Figure 5.24: Ballistic expansion <strong>of</strong> the atoms along the magnetic film. For<br />

this only the current in the confining end wires was turned <strong>of</strong>f. The left side<br />

shows absorption pictures. The right side shows the atomic density pr<strong>of</strong>iles<br />

in arbitrary units versus position, including a Gaussian fit. Expansion times,<br />

from top to bottom are T = 0 ms, 8 ms, 16 ms. The size <strong>of</strong> each image is<br />

5.9 × 0.4 mm 2 . The pr<strong>of</strong>iles are scaled with one mm/tick.<br />

coincide with the camera axis. To compensate for this, a linear function was<br />

added to the Gaussian function in the fitting process. From the Gaussian fits<br />

we can deduce the spread σx and calculate an axial temperature along the wire<br />

axis from this data using equation (5.6). The results are shown in Fig. 5.25.<br />

Measuring the temperature by standard ballistic expansion is not possible in<br />

this set-up, as the film’s inhomogenous magnetic field will push the atoms away<br />

and not allow a measurement <strong>of</strong> the thermal spread <strong>of</strong> the atomic cloud.<br />

When we omit the first two data points in Fig. 5.25, this procedure results<br />

in a temperature <strong>of</strong> Tx = 54 µK. It is clear that the linear behaviour fails<br />

for expansion times t < 8 ms. A high density effect may be the reason: the<br />

atoms are first confined tightly in two dimensions and loosely in the third. The<br />

barrier in the third direction is removed. The sudden energy gain for t < 8 ms<br />

can be related to atoms that relax from the high energy states in the remaining<br />

confined directions by collisions and convert their potential energy from one<br />

axis to kinetic energy along another axis. Only when the density has been<br />

sufficiently reduced the expected linear behaviour can be seen and thus we can<br />

148<br />

0.5<br />

0<br />

0.5<br />

0<br />

0.5<br />

0


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

2 2<br />

σ / mm<br />

x<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0 50 100 150 200 250 300 350 400<br />

T 2 / ms 2<br />

Figure 5.25: Squared atomic cloud size σ 2 x as a function <strong>of</strong> the squared ex-<br />

pansion time T 2 , taken from the results and Gaussian fits as presented in Fig.<br />

5.24. The solid line allows one to extract the temperature after relaxation<br />

processes have ended and results in Tx = 54 µK.<br />

indeed approximate the overall temperature by this one measurement. These<br />

results were obtained before the BEC was reached and demonstrate that even<br />

with the early experimental imperfections we were still able to cool the atoms<br />

enough to transfer them to the magnetic film trap.<br />

With the atoms trapped by the field <strong>of</strong> the magnetic film, we are able<br />

to further examine the properties and characterise the film. We see the first<br />

evidence <strong>of</strong> the so-called fragmentation <strong>of</strong> the atomic cloud [Lea03, For02,<br />

Kra02, Jon04]. To check for this, the difference between the actual pr<strong>of</strong>ile and<br />

the fitted function was taken for all data points. It is depicted in Fig. 5.26,<br />

where the differences are plotted for all six expansion times from 0 to 20 ms.<br />

The differences between the actual pr<strong>of</strong>ile and the fit show a common pr<strong>of</strong>ile.<br />

This can be explained by the finite roughness or inhomogenous magnetisation<br />

<strong>of</strong> the film, with roughness leading to an effective current that is corrugated<br />

by perpendicular currents [Est04, Sch05a, Wan04]. These lead to changes in<br />

the magnetic field, which affect the density <strong>of</strong> the atoms. Positions with a<br />

149


5.4. The Permanent Magnetic Trap<br />

(density pr<strong>of</strong>ile − Gaussian fit )/ a.u.<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

−0.02<br />

−0.04<br />

−0.06<br />

−0.08<br />

0 1 2 3 4 5 6<br />

position / mm<br />

Figure 5.26: Differences between actual pr<strong>of</strong>iles <strong>of</strong> the atomic cloud after bal-<br />

listic expansion and the fitted Gaussian pr<strong>of</strong>iles (see Fig. 5.24) as a function<br />

<strong>of</strong> position along the wire. The plot overlaps the differences <strong>of</strong> 6 expansion<br />

times from 0 to 20 ms.<br />

higher overall field will be less populated than positions with lower magnetic<br />

field. Further investigations have later been carried out and the source <strong>of</strong> the<br />

fragmentation has been identified as inhomogeneities in the magnetisation <strong>of</strong><br />

the film [Whi07].<br />

During the writing <strong>of</strong> this thesis, evaporation and Bose-Einstein conden-<br />

sation was achieved while trapping the atoms in the field <strong>of</strong> the permanent<br />

magnetic film. This work has now been published [Hal05, Hal06]. As a result,<br />

further characterisation <strong>of</strong> the properties <strong>of</strong> the film was possible. We use eqns.<br />

(2.43) and add a bias field with components parallel, By, and perpendicular,<br />

Bx, to the edge <strong>of</strong> the magnetic film. The trap will now appear at a distance<br />

from the film z0 where Bx = Bfilm, while By lifts the trap bottom from zero<br />

field. The result is a 3D harmonic trap with a radial frequency <strong>of</strong><br />

ωradial = µ0<br />

2π<br />

hM<br />

z2 �<br />

µBgF mF<br />

0 mBy<br />

(5.15)<br />

where h and M are the thickness and magnetisation <strong>of</strong> the film, µB is the Bohr<br />

150


Chapter 5: The Permanent Magnetic Chip Experiment: Results<br />

magneton, gF the Landé factor, mF the magnetic quantum number and m the<br />

atomic mass.<br />

Increasing the homogeneous bias field by about 5% for a time <strong>of</strong> 2 to 5 ms<br />

excited harmonic oscillations <strong>of</strong> the BEC’s radial centre <strong>of</strong> mass within the film<br />

trap. The position <strong>of</strong> the atomic cloud was then measured after 10 ms <strong>of</strong> free<br />

expansion over five periods <strong>of</strong> the oscillation, allowing a determination <strong>of</strong> the<br />

trap frequency with an accuracy <strong>of</strong> about 0.1%. Additionally, the trap bottom<br />

was measured using RF outcoupling with an accuracy <strong>of</strong> better than 1%. These<br />

sets <strong>of</strong> data were taken as functions <strong>of</strong> the distance from the magnetic film and<br />

allow a calculation <strong>of</strong> both the magnetic field strength and the gradient. The<br />

results are shown in Fig. 5.27. The figure also shows the expected results<br />

calculated from eq. (5.15) when using an effective current <strong>of</strong> hM = 0.2 A.<br />

Figure 5.27: The magnetic field strength (a) and gradient (b) as functions <strong>of</strong><br />

the distance from the chip surface. Data points (circles) and predictions <strong>of</strong><br />

a simple model (dotted line) show good agreement. Experimental errors are<br />

dominated by the image resolution. Taken from [Hal06].<br />

151


5.4. The Permanent Magnetic Trap<br />

152


Chapter 6<br />

The All-Optical Bose-Einstein<br />

Condensate Experiment:<br />

Apparatus<br />

6.1 Overview<br />

This chapter describes the all-optical Bose-Einstein Condensate experiment at<br />

the Institute for Quantum Optics, <strong>University</strong> <strong>of</strong> Hannover, Germany. Its goal<br />

was to achieve Bose-Einstein condensation <strong>of</strong> 87 Rb atoms held in an optical<br />

dipole trap, without any magnetic fields. Atoms were collected from a decel-<br />

erated atomic beam into a magneto-optical trap, and after pre-cooling loaded<br />

into the dipole trap. Evaporation was forced by ramping down the amplitude<br />

<strong>of</strong> the trapping light. The apparatus used for this experiment was first set up<br />

to examine the coherence <strong>of</strong> 85 Rb atoms in miniaturised trapping and guiding<br />

potentials [Buc01]. Following this project, the apparatus was used to analyze<br />

microstructured optical elements for use in atom optics and quantum informa-<br />

tion processing [Dum03a]. The success <strong>of</strong> this work stimulated the interest in<br />

153


6.1. Overview<br />

using a degenerate atomic ensemble in these micro-optical potentials for atom<br />

interferometry. To achieve this, the experimental set-up had to be changed to<br />

a more suitable isotope ( 87 Rb) and higher atom numbers in the initial MOT<br />

and optical trap. The optical trap was changed, using a laser that with a larger<br />

detuning, which means fewer scattered photons and less heating for the atoms<br />

in the trap. The vacuum system was changed so that the pressure could be<br />

further reduced. These changes are described here. The original set-up as de-<br />

scribed in [Buc01, Dum03a] will be described very briefly only where needed.<br />

The changes to the 87 Rb isotope and for higher atom numbers in the MOT<br />

can also be found in [Len04], the diploma thesis <strong>of</strong> A. Lengwenus, who assisted<br />

the author and took over the experiment after the author’s first stay, and in<br />

[M¨05, Geh05], which contain information about the improved optical trapping<br />

and evaporation.<br />

While the standard approach to reach Bose-Einstein condensation is to trap<br />

the atoms in the minimum <strong>of</strong> a magnetic potential, such an approach does not<br />

allow one to trap all possible atomic states. Magnetic fields cannot be designed<br />

to have a local maximum, so states that have a negative magnetic quantum<br />

number mF , so called strong field seeking states, cannot be trapped. Precision<br />

experiments for metrology with atoms in the magnetically neutral states, mF =<br />

0, are not possible with magnetic traps. This also means that optical traps<br />

allow one to collect atoms in different magnetic states at the same time, and<br />

to create a BEC consisting <strong>of</strong> several components. A further advantage <strong>of</strong><br />

optical dipole traps over magnetic traps is that they can be used to study<br />

the influence <strong>of</strong> homogeneous magnetic fields, which are needed to implement<br />

Feshbach resonances [Chi05b]. For applications in atom interferometry based<br />

on microstructured potentials, a technical advantage <strong>of</strong> optical dipole traps is<br />

the use <strong>of</strong> micro-optical arrays that are positioned outside the vacuum chamber<br />

154


Chapter 6: The All-Optical BEC Experiment: Apparatus<br />

and have their foci imaged into the chamber. This allows a quick change <strong>of</strong> the<br />

experimental set-up without the need to break the vacuum [Dum02a, Dum02b].<br />

The first all-optical BEC was realised in 2001 [Bar01]. It used crossed<br />

beams <strong>of</strong> a CO2 laser to produce a condensate <strong>of</strong> 87 Rb atoms. The same iso-<br />

tope was used in an experiment that reached BEC in a single beam <strong>of</strong> a CO2<br />

laser [Cen03]. A trap where the rubidium atoms were pre-cooled in an opti-<br />

cal lattice and then condensed in a compressible trap has been demonstrated,<br />

using a Nd:YAG laser [Kin05]. A BEC <strong>of</strong> Cs was created within a so-called<br />

“dimple” trap [Web03]. Here the focussed 1064 nm light from a YAG laser<br />

was added to the crossed beams <strong>of</strong> a CO2 laser. A two dimensional surface<br />

trap using blue detuned light was also successful in condensing Cs [Ryc04].<br />

These experiments demonstrate the importance <strong>of</strong> optical dipole traps, as Cs<br />

atoms in the magnetically trappable state could not be condensed due to high<br />

losses caused by a high rate <strong>of</strong> inelastic two-body collisions [Kok98]. A further<br />

element to be trapped optically and then cooled down to BEC was ytterbium.<br />

The trap here was created by crossed beams <strong>of</strong> frequency doubled light (523<br />

nm) from a YAG laser [Tak03]. All <strong>of</strong> these experiments are technically very<br />

demanding: they either work with light <strong>of</strong> a very long wavelength, needing spe-<br />

cial optics and windows, with frequency-doubled light, or with an intermediate<br />

cooling stage in an optical lattice.<br />

Our goal is to set up an apparatus that allows the creation <strong>of</strong> an all-optical<br />

BEC by a much simpler means, using standard optics and no intermediate<br />

stages.<br />

This chapter begins with the vacuum system which provides the environ-<br />

ment to cool and trap atoms in an optical dipole trap. This is followed by a<br />

description <strong>of</strong> the laser systems used to load atoms into and to trap and cool<br />

the atoms in a magneto-optical trap. The next section covers the laser system<br />

155


6.2. The Vacuum System<br />

used for trapping by the dipole force <strong>of</strong> the light. The chapter ends with a<br />

short description <strong>of</strong> the atom detection.<br />

6.2 The Vacuum System<br />

The vacuum system has been described in detail in [Buc01]. The changed set-<br />

up is depicted in Fig. 6.1. To achieve a low background pressure in the space<br />

experiment<br />

chamber<br />

Ion pump<br />

Ti: sublimation<br />

pump & cold<br />

finger (LN )<br />

2<br />

valve &<br />

diff. pumping<br />

stage<br />

Turbo<br />

pump<br />

oven<br />

chamber<br />

Figure 6.1: Schematic view <strong>of</strong> the vacuum system. The atomic beam starts<br />

from the oven on the right and passes through the differential pumping stage<br />

into the experiment chamber. The sublimation pump and cold finger are ad-<br />

ditions to the set-up described in [Buc01].<br />

<strong>of</strong> the atom traps, the system consists <strong>of</strong> two chambers. One chamber holds<br />

the source <strong>of</strong> the Rubidium atoms. This oven chamber is continually pumped<br />

by a turbo pump [E 22], leading to a typical pressure <strong>of</strong> 2.5 × 10 −7 Torr. The<br />

oven heats the Rubidium atoms which create an atomic beam. Between the<br />

oven and the experiment chamber a differential pumping stage with an inner<br />

156<br />

oven


Chapter 6: The All-Optical BEC Experiment: Apparatus<br />

diameter <strong>of</strong> 3 mm and a length <strong>of</strong> 15 cm is placed and the atomic beam is<br />

directed through it. The experiment chamber is connected to an ion pump<br />

[E 23] and to a Titanium sublimation pump [E 24]. Close to the sublimation<br />

pump a cold finger for liquid nitrogen has been installed. These last two devices<br />

were not part <strong>of</strong> the original set-up <strong>of</strong> [Buc01] and allow a further temporary<br />

reduction in the pressure if needed. The design <strong>of</strong> two chambers connected by<br />

a differential pumping stage allows the ion pump to maintain a pressure <strong>of</strong> less<br />

than 1 × 10 −10 Torr in the experiment chamber.<br />

As previous experiments on this set-up only used light close to 780 nm,<br />

four windows were anti-reflection coated for the new used wavelength <strong>of</strong> 1030<br />

nm.<br />

6.3 The Diode Laser Systems for Magneto-<br />

Optical Trapping<br />

As both the experiments at <strong>Swinburne</strong> and Hannover work with the same<br />

isotope, the laser systems employed in the experiments only differ in minor<br />

details, mostly in the techniques to detune and stabilise the frequency <strong>of</strong> the<br />

emitted light. The work <strong>of</strong> [M¨05, Len04] contain details about the laser sys-<br />

tems. Each system is based on diode laser technology with a wavelength <strong>of</strong> 780<br />

nm and labelled according to their use in the experiment. Unless noted, their<br />

roles are equivalent to the systems described in chapter 4.2. Additionally there<br />

is a set <strong>of</strong> two lasers which are used to decelerate the atomic beam so that the<br />

velocity <strong>of</strong> the atoms is slow enough to be captured by the MOT [Ert85]. In<br />

turn, they are labelled “Chirp lasers”.<br />

157


6.3. The Diode Laser Systems for Magneto-Optical Trapping<br />

6.3.1 The main laser<br />

As in the experiment described in chapter 4 the main light source for this<br />

experiment is a commercial tapered amplifier, type TA100 [E 25]. It shows<br />

similar spectral characteristics, but is different in the specifics <strong>of</strong> the stabilisa-<br />

tion and detuning schemes. The set-up up to the fibre coupling is depicted in<br />

Fig. 6.2. The frequency <strong>of</strong> this laser is stabilised using saturated absorption<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

200<br />

shutter<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

PBS<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

PBS<br />

ν=227 MHz<br />

AOM<br />

shutter<br />

AOM ν=233 MHz<br />

λ/4<br />

λ/4<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

λ/2<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxxx<br />

xxxxx xxxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

xxxxxxx<br />

PD<br />

100<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

TA 100<br />

λ/4<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

xxxxx<br />

Rb cell<br />

AOM<br />

ν=121 MHz<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

Figure 6.2: The main laser system: tapered amplifier, spectroscopy <strong>of</strong> the<br />

master laser and path <strong>of</strong> the amplified light.<br />

spectroscopy with a frequency modulated pump beam in conjunction with a<br />

commercial lock-in amplifier [E 26] and a home-built PID regulator. The fre-<br />

quency is modulated by an AOM [E 27]; the small modulation amplitude is<br />

added to a fixed frequency <strong>of</strong> ν1 = 121 MHz. As only the light <strong>of</strong> the saturation<br />

158<br />

λ/2<br />

PBS


Chapter 6: The All-Optical BEC Experiment: Apparatus<br />

beam is detuned, but not the detected light, the frequencies <strong>of</strong> the two beams<br />

now differ by 2 · ν1. The spectroscopy does not address atoms <strong>of</strong> velocity class<br />

v = 0, but rather those that have a Doppler detuning <strong>of</strong> ν1. Overall, the out-<br />

put light <strong>of</strong> the master laser is detuned by a frequency −ν1 with regard to the<br />

lock point <strong>of</strong> the spectroscopy. This lock point is the Lamb dip <strong>of</strong> the 87 Rb D2<br />

line which is created by the crossover resonance |F = 2〉 → |F ′ = 1, 2〉; here<br />

a peak-lock technique is used. As this line is less energetic than the cooling<br />

transition |F = 2〉 → |F ′ = 3〉 by 346 MHz, the overall detuning <strong>of</strong> the laser<br />

light frequency to the cooling transition is 467 MHz.<br />

This light is used to seed the tapered amplifier crystal <strong>of</strong> the TA100, which<br />

is current and temperature controlled [E 28]. The amplified light is split at<br />

a polarising beam splitter: one arm is used for the absorption imaging, the<br />

other as the cooling light <strong>of</strong> the MOT. Each arm passes an AOM twice for fast<br />

switching and detuning to the required frequencies, and a mechanical shutter<br />

[E 10] to fully block the light from the experiment before the light <strong>of</strong> each arm<br />

is coupled into a fibre to be transferred to the experimental chamber. The<br />

light for cooling and trapping the atoms is detuned by the AOM [E 29], so<br />

that the frequency <strong>of</strong> the light is 13 MHz below the resonance. This leads to<br />

a final power <strong>of</strong> ≈ 50 mW available at the position <strong>of</strong> the MOT. The other<br />

arm double-passes an AOM [E 30], so that this light is resonant with the<br />

|F = 2〉 → |F ′ = 3〉 transition with a maximum power <strong>of</strong> slightly less than 200<br />

µW available for imaging.<br />

6.3.2 The repumping laser<br />

To close the |F = 2〉 → |F ′ = 3〉 cooling cycle a home-built diode laser system<br />

is used. This system is described fully in [Buc01]. It consists <strong>of</strong> a grating<br />

stabilised diode laser in Littrow arrangement, which is temperature and current<br />

159


6.3. The Diode Laser Systems for Magneto-Optical Trapping<br />

stabilised by home-built electronics. The frequency is controlled by saturated<br />

absorption spectroscopy similar to that described in the section above, locking<br />

directly onto the |F = 1〉 → |F ′ = 2〉 transition. Only the saturation beam<br />

double-passes an AOM which is driven at ν = 72 MHz, leading to an overall<br />

detuning <strong>of</strong> the same magnitude. This light can be used to injection-lock a<br />

second laser diode for higher power. One result <strong>of</strong> the experiments was that<br />

for the loading <strong>of</strong> the dipole trap a minimal amount <strong>of</strong> this repumping light<br />

is beneficial (see chapter 7). For some experiments a grey filter was used to<br />

reduce this intensity; for others the light <strong>of</strong> the master laser has been used<br />

directly without any amplification. A second AOM, also running at ν = 72<br />

MHz, is used to tune the frequency <strong>of</strong> the light back to resonance and to<br />

quickly switch the light. The light <strong>of</strong> the zeroth order is fed into a scanning<br />

etalon to monitor the quality <strong>of</strong> the injection lock. A mechanical shutter in<br />

the path can be used to completely block the repumping light.<br />

6.3.3 The chirp lasers<br />

The atoms which are finally trapped in the MOT originate from an oven at<br />

several hundred Kelvin. The resulting beam <strong>of</strong> hot thermal atoms has to be<br />

decelerated so that the slowed atoms can be captured by the MOT forces.<br />

In this set-up this is done by chirp cooling the beam [Ert85]. This needs a<br />

cooling and a repumping laser. The light sources here are home-built, grating<br />

stabilised diode lasers, whose frequencies are stabilised by saturated absorption<br />

spectroscopy on the required wavelengths. The cooling laser consists <strong>of</strong> a<br />

master-slave system, where the light <strong>of</strong> the frequency stabilised master laser is<br />

used to injection-lock a single laser diode. The repumping light comes from a<br />

single stabilised diode laser. The chirp which detunes the lasers to keep them<br />

in resonance with the slowed atoms is created by a ramp voltage which is<br />

applied to the PZT <strong>of</strong> both the cooling master laser and the repumping laser.<br />

160


Chapter 6: The All-Optical BEC Experiment: Apparatus<br />

This chirp signal is then repeatedly applied to the lasers for a quasi-continuous<br />

loading <strong>of</strong> the MOT.<br />

6.4 The Dipole Trap<br />

The aim <strong>of</strong> this experiment was to reach degeneracy <strong>of</strong> the bosonic atomic<br />

ensemble by trapping in a purely optical trap without the help <strong>of</strong> any magnetic<br />

fields. This requires light <strong>of</strong> sufficient power and detuning from the transition<br />

to create a trap deep enough to trap many atoms and so far-detuned that<br />

the heating rate <strong>of</strong> the atoms due to scattering <strong>of</strong> photons from the beam is<br />

sufficiently small. While a Ti:Sa laser allows the creation <strong>of</strong> deep traps, the<br />

small detuning leads to a high scattering rate and allows temperatures just<br />

below 1 µK. CO2 lasers are available with high powers and wavelengths <strong>of</strong><br />

10.6 µm, and have been used for evaporation and creation <strong>of</strong> BECs [Bar01,<br />

Cen03]. The long wavelength requires the use <strong>of</strong> special materials for the<br />

optical elements, such as windows made <strong>of</strong> ZnSe. We use a Yb:YAG disk-laser<br />

[E 31] with an output wavelength <strong>of</strong> 1030 nm as the light source. Its emission<br />

wavelength is sufficiently far away for low scattering rates, the output power<br />

is high enough for a sufficiently deep trap, and standard optical components<br />

can be used. This laser was also chosen because its output is single mode and<br />

single frequency. Inside its folded cavity, an etalon and a Lyot-filter are used as<br />

frequency selective elements. These features are beneficial for the experiments<br />

as the trap created by the light has a high stability which is an important factor<br />

in reducing heating effects. The laser has a maximum output power <strong>of</strong> 50 W<br />

(without frequency selective elements) and 25 W (in single mode operation).<br />

To trap atoms, the focus <strong>of</strong> the laser beam has to be imaged into the atomic<br />

cloud. The set-up <strong>of</strong> the optics to achieve this is shown in Fig. 6.3. The<br />

emitted light first passes a combination <strong>of</strong> a half-wave plate and a polarising<br />

161


6.5. Detection <strong>of</strong> Atoms<br />

beam splitter for a rough intensity control. A telescope then reduces the beam<br />

diameter by a factor <strong>of</strong> two. An AOM is used for fine control <strong>of</strong> the light<br />

intensity, so that the power in the beam and thus the trap depth can be<br />

ramped down in a controlled way. A lens <strong>of</strong> f = 200 mm focusses the beam.<br />

Before this focus is imaged in the experiment chamber, the beam is split into<br />

two by a polarising beam splitter. The lenses <strong>of</strong> f = 250 and f = 300 mm<br />

each create an image <strong>of</strong> the focus. The two beams are aligned so that they<br />

cross in their foci in the atomic cloud. The perpendicular polarisation <strong>of</strong> the<br />

beams reduces interference effects in the crossed trap. The light in one <strong>of</strong> the<br />

arms is then used to control and monitor the laser. This light is split and sent<br />

to a photodiode and to a scanning etalon. The photodiode feeds its signal to a<br />

PI-regulator, which in turn uses the AOM to stabilise the intensity, increasing<br />

the long term stability <strong>of</strong> the intensity. The signal <strong>of</strong> the scanning etalon shows<br />

whether the laser is emitting a single mode or several modes <strong>of</strong> light.<br />

6.5 Detection <strong>of</strong> Atoms<br />

At the top <strong>of</strong> the chamber, a CCD camera [E 32] is mounted. This camera<br />

can be used for detecting the atoms either by their fluorescence or by their<br />

absorption. In this experiment fluorescence detection was used only in the<br />

early stages and for the MOT. All images <strong>of</strong> atoms trapped in the dipole trap<br />

were taken using absorption imaging. The camera used in this experiment<br />

has a pixel size <strong>of</strong> 9 × 9 µm 2 . The magnification <strong>of</strong> the imaging optics has<br />

been determined to 1.06 ± 0.1, so that each pixel gathers light from an area <strong>of</strong><br />

(8.5 ± 0.8 µm) 2 .<br />

162


PD<br />

Chapter 6: The All-Optical BEC Experiment: Apparatus<br />

scanning etalon<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

VersaDisk<br />

disk laser<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

experiment<br />

chamber<br />

λ/2<br />

200<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

xxx<br />

250<br />

λ/2 PBS<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxx<br />

PD<br />

250 300<br />

PBS<br />

60 -30<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

AOM<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

xxxxxxxxxxxx<br />

Figure 6.3: Set-up <strong>of</strong> the dipole trap. The focus <strong>of</strong> a f = 200 mm lens is<br />

imaged into the experiment chamber by a f = 250 mm and a f = 300 mm<br />

lens. The atomic beam enters the chamber from the right.<br />

163


6.5. Detection <strong>of</strong> Atoms<br />

164


Chapter 7<br />

The All-Optical Bose-Einstein<br />

Condensate Experiment:<br />

Results<br />

7.1 Overview<br />

In this chapter the results from the all-optical BEC experiment located at<br />

the Institut für Quantenoptik in Hannover, Germany, are presented. As the<br />

author worked on the experiment as a non-resident researcher, these results also<br />

appear in colleagues’ theses [M¨05, Geh05, Len04]. Apart from two exceptions<br />

only original work for which the author actively contributed to the experiment<br />

is included in this thesis. The exceptions are marked as citations in the text<br />

and figures.<br />

The typical experiment starts with the trapping <strong>of</strong> atoms in a magneto-<br />

optical trap. This trap is loaded from a chirp cooled atomic beam, and it is<br />

possible to load up to 10 9 87 Rb atoms at a temperature <strong>of</strong> as low as 35 µK. The<br />

optical trap is created by two crossed beams <strong>of</strong> 8 W <strong>of</strong> power at a wavelength<br />

165


7.2. The Magneto-Optical Trap<br />

<strong>of</strong> 1030 nm and is loaded directly from the MOT, with both traps active during<br />

the loading. After 30 ms loading time, we find a maximum <strong>of</strong> 1.5·10 5 atoms in<br />

the dipole trap. Evaporation takes place during three linear ramps, in which<br />

the trap depth is decreased by two orders <strong>of</strong> magnitude. Typical values before<br />

the evaporation are 6 · 10 4 atoms at a phase space density <strong>of</strong> about 2 · 10 −4 .<br />

After the evaporation ramp, we are left with ≈ 10 3 atoms and a phase space<br />

density <strong>of</strong> about 0.2.<br />

In the following sections, the results achieved so far will be described, start-<br />

ing with the characterisation <strong>of</strong> the magneto-optical trap. This is then followed<br />

by the description <strong>of</strong> the optical trap. Here the first section deals with the<br />

loading and the different steps to optimise the loading process. The chapter<br />

ends with the results from the evaporative cooling implemented so far. Un-<br />

fortunately quantum degeneracy was not reached at the time this thesis was<br />

written.<br />

7.2 The Magneto-Optical Trap<br />

The magneto-optical trap in the Hannover experiment is a six-beam MOT.<br />

Three beams are separated before the experiment chamber and each is retrore-<br />

flected onto itself after passing the chamber and a quarter-wave plate. The<br />

atom source in this experiment is a chirp cooled atom-beam, where the actual<br />

Rb-oven is located in a second chamber and the beam is guided through a<br />

differential pumping stage. The quadrupole field is created by anti-Helmholtz<br />

configured coils inside the vacuum chamber. The results presented here were<br />

achieved during the author’s first stay in Hannover. The data was taken to-<br />

gether with A. Lengwenus and our results also are part <strong>of</strong> his thesis [Len04],<br />

where they are discussed in greater detail.<br />

A part <strong>of</strong> the fluorescence from the captured atoms is captured by a photo-<br />

166


Chapter 7: The All-Optical BEC Experiment: Results<br />

diode, allowing an online measurement <strong>of</strong> the number <strong>of</strong> atoms. This was<br />

used to measure the loading and loss characteristics <strong>of</strong> this trap. The loading<br />

process is shown in Figure 7.1, (a). Use <strong>of</strong> equation (5.1) allows us to calculate<br />

number <strong>of</strong> atoms<br />

x 108<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0 5 10 15 20 25 30 35 40 45<br />

time / s<br />

number <strong>of</strong> atoms<br />

x 108<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0 5 10 15 20 25 30 35 40 45<br />

time / s<br />

Figure 7.1: Loading and decay curve <strong>of</strong> the Hannover MOT. (a) Loading:<br />

fitted to the data is equation (5.1), with the loading rate RL = 6.4 · 10 7 atoms<br />

s −1 , and the loss γ = 0.08 s −1 ; (b) Decay: equation (5.2) has been fitted to<br />

the data, with a loss rate <strong>of</strong> γ = 0.06 s −1 .<br />

the loading rate RL = 6.4 · 10 7 atoms per second from the fit <strong>of</strong> the equation<br />

to the data. The loss rate from the trap is fitted to γ = 0.08 s −1 . The highest<br />

number <strong>of</strong> atoms that were trapped at this stage was more than 1.3·10 9 atoms.<br />

The loading process can be turned <strong>of</strong>f by shutting <strong>of</strong>f the chirp lasers, so<br />

the atoms in the beams are too fast to be captured. The atom number in the<br />

trap then begins to decay according to equation (5.2). This is shown in Figure<br />

7.1, (b). The loss coefficient here has been determined to be γ = 0.06 s −1 .<br />

The lifetime <strong>of</strong> the MOT is the inverse <strong>of</strong> the loss parameter, τ = 14.3 s for<br />

the mean <strong>of</strong> both values <strong>of</strong> γ. Compared to the MOT in Melbourne, this set-up<br />

has a faster loading rate but a shorter lifetime. The first can be explained by<br />

the more efficient loading from an atomic beam. The shorter lifetime probably<br />

has its origin in the fact that these measurements were taken with a detuning<br />

167


7.3. Loading the Dipole Trap<br />

from the atomic resonance <strong>of</strong> about 6 MHz only. The atoms in the trap thus<br />

are not cooled as much. The detuning was then changed to a higher value <strong>of</strong><br />

about 15 MHz when it came to load the atoms into the dipole trap.<br />

Before the loading into the optical trapping potential, the atoms are cooled<br />

by a phase <strong>of</strong> optical molasses. The fluorescence <strong>of</strong> the atoms was used in time<br />

<strong>of</strong> flight measurements to determine the temperature <strong>of</strong> the atomic cloud.<br />

Using equation (5.6), the temperature reached after this stage was measured<br />

as T = 34 µK.<br />

7.3 Loading the Dipole Trap<br />

In this section the loading <strong>of</strong> atoms from the MOT into the optical dipole<br />

potential is described. This work was done during the author’s second and<br />

third stay in Hannover. The results have thus also been published in the<br />

theses <strong>of</strong> the author’s colleagues at that time [Geh05, M¨05] in greater detail<br />

and among their individually taken results.<br />

To achieve quantum degeneracy by evaporative cooling in optical traps,<br />

a high initial number <strong>of</strong> trapped atoms is even more important than it is<br />

for the experiments with magnetic traps. This is because the trap widens<br />

during the evaporation ramp, and thus the time for rethermalisation increases<br />

with shallower traps. Indeed, the mean time between two elastic collisions<br />

scales with the inverse <strong>of</strong> the product <strong>of</strong> the trap frequencies <strong>of</strong> each direction<br />

[Geh05]. This need for a high initial atom number led to the optimisation <strong>of</strong><br />

several parameters. These results are presented here, where all data was taken<br />

by absorption measurements.<br />

For an isotropic confinement <strong>of</strong> the atoms, the trapping potential was cre-<br />

ated by two crossed, red detuned laser beams. Unless noted, a power <strong>of</strong> 8W<br />

per beam <strong>of</strong> wavelength 1030 nm was focussed down to a size <strong>of</strong> 60 µm. This<br />

168


Chapter 7: The All-Optical BEC Experiment: Results<br />

corresponds to a trap depth <strong>of</strong> about U = kB ·300 µK or less than 1/10th <strong>of</strong> the<br />

depth <strong>of</strong> the magnetic trap in section 5. This trap was loaded directly from the<br />

magneto-optical trap. This is shown in Figure 7.2, where the trap formed by a<br />

single beam is loaded from the MOT. The absorption light’s frequency here is<br />

Figure 7.2: Absorption image <strong>of</strong> a dipole trap and the MOT during the loading<br />

stage for a single beam optical trap. The detuning <strong>of</strong> the absorption light is 4<br />

MHz.<br />

shifted by 4 MHz from the atomic resonance. The atoms in the MOT and the<br />

dipole trapped atoms show their peak absorption at different laser frequencies.<br />

This difference <strong>of</strong> 6 MHz is due to the ac-Stark shift that the energy levels <strong>of</strong><br />

the atoms in the optical trap experience which is absent for the atoms in the<br />

MOT.<br />

The lifetime in the crossed trap was measured for both hyperfine states <strong>of</strong><br />

the ground state, F = 1 and F = 2. The rate equation for the atom number<br />

N takes into account constant losses α and losses that are linearly density<br />

dependent, like two-body-collisions, β:<br />

dN<br />

dt<br />

= −αN − βN 2<br />

(7.1)<br />

Solving this equation for N yields an exponential function in α for an initial<br />

atom number <strong>of</strong> N0 (see equation (5.9)). This formula was fitted to the data<br />

169


7.3. Loading the Dipole Trap<br />

number <strong>of</strong> atoms<br />

x 104<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000<br />

time / ms<br />

Figure 7.3: Lifetime measurements <strong>of</strong> the dipole traps for both ground states<br />

F = 1 and F = 2.<br />

<strong>of</strong> the lifetime measurements, and the results are shown in Figure 7.3. For<br />

both ground states, the constant loss parameter α was negligible. The density<br />

dependent loss was measured to be βF =1 = 4.6·10 −6 s −1 for the lower hyperfine<br />

state and βF =2 = 3.8 · 10 −5 s −1 for the upper state. The parameter for the<br />

state with higher energy is thus about one order <strong>of</strong> magnitude larger and these<br />

atoms have a reduced lifetime in the trap. The reason for this is that the atoms<br />

in the F = 1 state cannot undergo hyperfine-state changing collisions, while<br />

the atoms in F = 2 can. These hyperfine-state changing collisions and photo<br />

association <strong>of</strong> atoms to molecules are the main loss mechanisms in the dipole<br />

trap. Scaled with the initial atom number, the density dependent losses <strong>of</strong> the<br />

F = 1 state are comparable with the losses in the magnetic trap <strong>of</strong> section 5.3,<br />

while N0βF =2 ≈ 2.5 s −1 is about one order <strong>of</strong> magnitude larger.<br />

During the loading <strong>of</strong> the trap, which is under the influence <strong>of</strong> the near-<br />

resonant light <strong>of</strong> the MOT lasers, light assisted collisions dominate the losses.<br />

Here we again find collisions that change the hyperfine state <strong>of</strong> the atom, but<br />

now losses due to radiative escape are by far the most dominating [Kup00]. In<br />

this case, one <strong>of</strong> the two colliding atoms has to be in the upper ground state,<br />

170<br />

F = 1<br />

F = 2


Chapter 7: The All-Optical BEC Experiment: Results<br />

F = 2, as the detuning <strong>of</strong> the light for the MOT is too large to couple from<br />

the lower state F = 1 to a molecular state during a collision. Thus, the rate<br />

equation here will have a loss term β ∝ NF =2 · (NF =1 + NF =2). As it is the<br />

role <strong>of</strong> the repumping laser <strong>of</strong> the MOT to close the cooling cycle by driving<br />

the transition F = 1 → F ′ = 2 → F = 2 one can already see that during the<br />

loading <strong>of</strong> atoms into the dipole trap a reduced power in the repumping light<br />

is advisable as it minimizes β by minimizing NF =2. The detailed calculation<br />

can be found in [Kup00].<br />

The dynamics <strong>of</strong> the loading can be deduced from the loss equation by<br />

adding a term to equation (7.1) to cover the transfer <strong>of</strong> atoms from the MOT.<br />

dN<br />

dt = −αN − βN 2 + R0 · e −γt<br />

(7.2)<br />

Here R0 is the initial loading rate from the maximum sized MOT. The MOT<br />

size will decrease though as the settings <strong>of</strong> the cooling and repumping lasers for<br />

an optimum transfer from the MOT to the dipole trap are different from the<br />

settings for a standard MOT. One has to keep in mind that the parameters<br />

α and β here are under the influence <strong>of</strong> the near-resonant light, and hence<br />

β = 3.55 · 10 −4 s −1 is about two orders <strong>of</strong> magnitude larger than without the<br />

light [M¨05]. The parameter γ = 10 s −1<br />

can be determined by measuring the<br />

lifetime <strong>of</strong> the MOT with the settings used for the transfer.<br />

The number <strong>of</strong> atoms in the dipole trap was measured as a function <strong>of</strong><br />

loading time without the help <strong>of</strong> the author [M¨05]. The results are shown in<br />

Figure 7.4. The loading parameter R0 is given by the initial gradient <strong>of</strong> the<br />

loading curve and was measured to be R0 = 9.2·10 6 s −1 . The maximum number<br />

<strong>of</strong> atoms here was Nmax = 1.4 · 10 5 atoms. This is largely consistent with<br />

an approximated equilibrium value for N from equation (7.2). The density<br />

independent losses α are negligible, and if one assumes the MOT to be constant<br />

during the loading (γ ≡ 0) then we can approximate the maximum number <strong>of</strong><br />

171


7.3. Loading the Dipole Trap<br />

number <strong>of</strong> atoms<br />

x 104<br />

15<br />

10<br />

5<br />

0<br />

0 100 200 300 400 500 600 700 800 900 1000<br />

time / s<br />

Figure 7.4: Loading curve <strong>of</strong> the dipole trap from the MOT: number <strong>of</strong> opti-<br />

cally trapped atoms as a function <strong>of</strong> time. The markers are results from the<br />

experiment, the solid line is the numerical fit <strong>of</strong> equation (7.2), taken from<br />

[M¨05].<br />

atoms Nmax,approx from equation (7.2) to<br />

Nmax,approx = � R0/β ′ = 1.6 · 10 5<br />

(7.3)<br />

Also, from the graph, the optimum loading time can be deduced to be about<br />

30 ms for our experimental setting.<br />

The loading and loss parameters depend on the settings <strong>of</strong> the MOT lasers.<br />

A higher detuning <strong>of</strong> the cooling laser leads to a cooler atomic cloud to load<br />

from, increasing the number <strong>of</strong> atoms in the dipole trap. On the other hand,<br />

if the detuning is chosen to be too high, then the atom number in the MOT<br />

itself decreases and thus the loading characteristics <strong>of</strong> the optical trap decrease.<br />

Figure 7.5 (a) shows the number <strong>of</strong> optically trapped atoms in a trap depth <strong>of</strong><br />

about 300 µK (a crossed trap with a waist <strong>of</strong> w0 = 45 µm and 3.5 W per beam)<br />

as a function <strong>of</strong> the detuning <strong>of</strong> the MOT light. The effect <strong>of</strong> the detuning was<br />

examined without the help <strong>of</strong> the author [Geh05].<br />

The optimum detuning can be determined to be −50 MHz from the atomic<br />

172


Chapter 7: The All-Optical BEC Experiment: Results<br />

resonance. Increasing the light’s power in the arms <strong>of</strong> the trap will cause the<br />

distribution to flatten up to the point where no influence <strong>of</strong> the detuning on the<br />

atom number can be seen anymore for high power [Geh05]. The reason why<br />

the optimum detuning here differs by about a factor <strong>of</strong> three from experiments<br />

that use CO2 lasers to create the trap [Cen03] (the detuning there is about<br />

−150 MHz) is that the energy levels are affected differently by the radiation’s<br />

ac-Stark shift. In the case <strong>of</strong> the quasi-static trap with a radiation wavelength<br />

<strong>of</strong> 10 µm all energy eigenvalues are shifted to lower energies. In the case <strong>of</strong> the<br />

Rb D2 line the differential Stark shift there can detune the transition by about<br />

60 MHz towards the red [Cen04]. In our case the upper energy level is shifted<br />

further up while the lower state is even lowered. Here the transition energy<br />

is increased, and the transition is blue detuned. If we take into account the<br />

nominal detuning <strong>of</strong> the light towards the unperturbed atomic energy levels<br />

and the shift <strong>of</strong> the energy levels, then the optimum detuning for CO2 traps<br />

is about −90 MHz while it is between −70 and −80 MHz in our case. The<br />

difference then is comparable.<br />

The intensity <strong>of</strong> the MOT’s repumping light will directly affect the loading<br />

by pumping atoms into a hyperfine state that is trapped with a shorter lifetime.<br />

The number <strong>of</strong> atoms in the trap will thus increase for a lower intensity <strong>of</strong> this<br />

laser light, although it has to be kept sufficiently high to keep the magneto-<br />

optical trap working. Figure 7.5 (b) shows the number <strong>of</strong> atoms in the dipole<br />

trap for different powers <strong>of</strong> repumping light. Here the frequency <strong>of</strong> the light<br />

is fixed to the F = 1 → F ′ = 2 transition frequency <strong>of</strong> the atoms. One<br />

can clearly see an optimum power at about 50 µW. This corresponds to an<br />

intensity <strong>of</strong> about 30 µW/cm 2 and a factor <strong>of</strong> 10 less than the power that is<br />

usually applied to the MOT.<br />

173


7.4. Evaporative Cooling in the Dipole Trap<br />

number <strong>of</strong> atoms<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

x 104<br />

4<br />

0<br />

−80 −70 −60 −50 −40 −30 −20<br />

detuning / MHz<br />

x 104<br />

6.5<br />

6<br />

5.5<br />

5<br />

4.5<br />

4<br />

10 0<br />

3.5<br />

10 1<br />

power / µW<br />

Figure 7.5: Number <strong>of</strong> atoms in the optical trap: (a) as a function <strong>of</strong> the<br />

detuning <strong>of</strong> the MOT cooling laser, taken from [Geh05], (b) as a function <strong>of</strong><br />

the power in the MOT’s repumping light on a logarithmic scale.<br />

7.4 Evaporative Cooling in the Dipole Trap<br />

Evaporative cooling is possible in an optical trap, although it works differently<br />

from the case <strong>of</strong> a magnetic trap. In the magnetic case, the trap remains<br />

unchanged and the high energy atoms are pumped into untrapped mF -states<br />

by RF-radiation. In optical traps, one has the possibility to reduce the trap<br />

depth directly by reducing the intensity <strong>of</strong> the trapping light. As the beam<br />

waist is constant regardless <strong>of</strong> the intensity in the beam, this results in a<br />

change <strong>of</strong> the trapping gradient and curvature. This shallower trap increases<br />

the rethermalisation times needed in the evaporation process. This fact has<br />

to be taken into account when the form <strong>of</strong> the evaporation ramp is chosen.<br />

Figure 7.6 shows how the the intensity in the trapping beams is decreased over<br />

a ramp <strong>of</strong> four seconds in four steps. The first step is a 100 ms long phase<br />

<strong>of</strong> free evaporation at the maximum power <strong>of</strong> 17 W trapping light (or a trap<br />

depth <strong>of</strong> 390 µK). This is followed by three linear ramps to a final value <strong>of</strong> 110<br />

mW or a trap depth <strong>of</strong> 2 µK in the lowest case. For laser powers <strong>of</strong> 70 mW or<br />

less, it is expected that the optical potential is insufficient to overcome gravity<br />

174<br />

number <strong>of</strong> atoms<br />

10 2<br />

10 3


trapping light power / W<br />

18<br />

16<br />

14<br />

12<br />

10<br />

Chapter 7: The All-Optical BEC Experiment: Results<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 0.5 1 1.5 2 2.5 3 3.5 4<br />

time / s<br />

Figure 7.6: The power in the trapping beams as a function <strong>of</strong> time: the evap-<br />

oration ramp.<br />

and trap atoms. The ramps were optimized by varying the ramp length for a<br />

chosen final value <strong>of</strong> that ramp. The figure <strong>of</strong> merit was the atom number <strong>of</strong><br />

the atomic cloud. This number first increased with a longer ramp, as such a<br />

ramp reduced the spilling [Geh05]. After an optimum ramping time, the atom<br />

number decreased again due to the finite lifetime <strong>of</strong> the trap.<br />

In our case, the best results were achieved with a focus <strong>of</strong> about 40 µm.<br />

In this case, 6 · 10 4 atoms are loaded from the MOT into the optical trap.<br />

This results in an initial phase space density <strong>of</strong> 3 · 2 · 10 −4 ; the factor <strong>of</strong> three<br />

stems from the distribution over all three Zeeman sublevels. Assuming these<br />

sublevels are populated evenly, the starting phase space density is reduced to<br />

2 · 10 −4 .<br />

The final laser power has been varied from 210 mW to 110 mW. At the<br />

end <strong>of</strong> the ramp, absorption images have been taken to determine the atom<br />

number and the temperature <strong>of</strong> the atomic cloud. To increase the signal to<br />

noise ratio the images <strong>of</strong> three identical experiment runs have been summed<br />

up before further evaluation. These sum images are shown in Figure 7.7.<br />

The temperatures were determined using the time <strong>of</strong> flight technique. The<br />

175


7.4. Evaporative Cooling in the Dipole Trap<br />

213 mW 183 mW 170 mW<br />

153 mW 143 mW 113 mW<br />

Figure 7.7: Absorption images <strong>of</strong> the atomic cloud after the evaporation ramp<br />

for different final values <strong>of</strong> the trapping power. Each image shown is the sum<br />

over three absorption images after 10ms TOF.<br />

results are depicted in Figure 7.8. Although a rather sudden decrease in the<br />

temperature can be seen for trapping powers less than 170 mW, this is not<br />

a signature <strong>of</strong> the phase transition to a BEC as it is not accompanied by a<br />

sufficient increase in the atomic density: the density for the 210 mW trap is 6 ·<br />

10 12 cm −3 and increases to 8.5·10 12 cm −3 for the 150 mW trap. The phase space<br />

density in all images is <strong>of</strong> the order <strong>of</strong> 0.2 (assuming equal distribution over all<br />

mF sublevels). Compared to the initial value <strong>of</strong> the PSD before evaporation<br />

(2 · 10 −4 ) this is an increase by a factor <strong>of</strong> 1000, which is twice what would<br />

be expected by the scaling laws for optical traps [O’H01]. Unfortunately, this<br />

176


temperature / nK<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

Chapter 7: The All-Optical BEC Experiment: Results<br />

0<br />

100 120 140 160 180 200 220<br />

final power / mW<br />

Figure 7.8: Temperatures from the TOF measurements <strong>of</strong> the images in Fig.<br />

7.7<br />

factor can be explained by experimental errors, as this factor <strong>of</strong> two in phase<br />

space density is equivalent to an error <strong>of</strong> 25% in the temperature measurement.<br />

As we look at only a few thousand atoms in the clouds and as is obvious from<br />

Fig. 7.8, this error is <strong>of</strong> the order <strong>of</strong> the experimental error. To this we can<br />

add the error in the density measurement, as the focal size also can only be<br />

measured with a 30% accuracy. Still, the evaporation works and only one<br />

order <strong>of</strong> magnitude in phase space density has to be gained to reach quantum<br />

degeneracy. This should be achievable by improving the parameters during<br />

the loading and an optimised focal and thus trap size. Employing a so-called<br />

“dark” MOT allows one to increase the density <strong>of</strong> atoms before loading them<br />

into the dipole trap [Ket93, Tow96]. Changing the experimental set-up to<br />

create such a MOT was considered but postponed until other enhancements<br />

are implemented.<br />

177


7.4. Evaporative Cooling in the Dipole Trap<br />

178


Chapter 8<br />

Summary and Outlook<br />

8.1 Summary and Discussion<br />

In this thesis, two atom optics experiments and their results have been pre-<br />

sented. The far goals <strong>of</strong> these experiments are to set up atom interferometers<br />

in microstructured potentials. Near goals were to achieve quantum degeneracy<br />

<strong>of</strong> the trapped bosonic atoms. Unfortunately, the phase transition to BEC was<br />

reached in only one <strong>of</strong> the experiments. Both experiments have in common<br />

that they work with the same atomic species <strong>of</strong> 87 Rb, an isotope which has<br />

become a standard in atom optics. Yet, they use different interactions to trap<br />

the atoms.<br />

The first experiment is the permanent magnetic chip experiment at Swin-<br />

burne <strong>University</strong> <strong>of</strong> <strong>Technology</strong>. Here the trapping potential was created by<br />

magnetic fields. In this experiment a new hybrid atom chip was used, in which<br />

current-carrying structures and a magneto-optical film were combined. This<br />

film was produced in the group, and consists <strong>of</strong> multiple layers, where layers<br />

<strong>of</strong> the ferrimagnetic alloy TbGdFeCo alternate with layers <strong>of</strong> Chromium. It<br />

allows the trapping <strong>of</strong> atoms in a micron sized potential and their cooling down<br />

179


8.1. Summary and Discussion<br />

to Bose-Einstein condensation. Not only is this chip new, the whole experi-<br />

mental apparatus was first set up as part <strong>of</strong> this thesis. In chapter 4 the whole<br />

apparatus, starting with the laser systems, including the vacuum system, and<br />

ending with the above-mentioned chip, is described in detail.<br />

The results <strong>of</strong> this experiment, from the initial collection <strong>of</strong> atoms in a<br />

MOT up to quantum degeneracy <strong>of</strong> the atoms, can be found in chapter 5. In<br />

the MOT, we trap 5·10 8 atoms at about 90 µK initially and cool to 40 µK after<br />

compression to match the atomic cloud with the magnetic trapping potential.<br />

After optically pumping all atoms into a magnetically trappable substate, we<br />

collect 8 · 10 7 atoms at 50 µK in an initial Z-wire magnetic trap. This trap<br />

is compressed, and then the atoms are cooled by forced evaporation using RF<br />

radiation. This leads us to a BEC with up to 10 5 atoms. We were also able<br />

to trap atoms without the current-carrying wires and with only the magneto-<br />

optical film. Chapter 4 ends with a presentation <strong>of</strong> the early results <strong>of</strong> these<br />

traps. The work on this experiment, together with the theoretical work on<br />

the double well atom interferometer, finally led to a double well system in the<br />

potential <strong>of</strong> the magnetic film using a BEC capable <strong>of</strong> measuring asymmetries<br />

between the wells [Hal07b].<br />

In the second experiment the trap’s potential was provided by spatially<br />

intensity-dependent light, a dipole trap. As part <strong>of</strong> this thesis the existing<br />

experimental set-up in Hannover was improved (see chapter 6). The number<br />

<strong>of</strong> trapped atoms was improved for each step in the experiment to 10 9 atoms<br />

during the MOT phase and 1.5 · 10 5 atoms in the dipole trap. A much better<br />

transfer seems unrealistic, as the dipole trap’s radius is <strong>of</strong> the order <strong>of</strong> 10<br />

µm and thus several orders <strong>of</strong> magnitude smaller than the MOT from where<br />

the atoms are loaded. Cooling by evaporation <strong>of</strong> atoms from the trap was<br />

demonstrated. With this the temperature <strong>of</strong> the atomic cloud could be reduced<br />

from about 30 µK in the MOT to below 200 nK. Unfortunately, the atom<br />

180


Chapter 8: Summary and Outlook<br />

number could not be sufficiently increased to demonstrate beyond doubt that<br />

Bose-Einstein condensation in this set-up was achieved, although an increase<br />

in the phase space density by a factor <strong>of</strong> 1000 was shown. With this set-up,<br />

only a further factor <strong>of</strong> 10 in the phase space density is required to reach<br />

quantum degeneracy. The complete results <strong>of</strong> the dipole trap experiment can<br />

be found in chapter 7.<br />

In both experiments presented here atoms are trapped in micron-sized po-<br />

tentials. With both kinds <strong>of</strong> trapping the splitting and merging <strong>of</strong> traps have<br />

already been demonstrated [Cas00, H¨01c, Dum02b]. A beamsplitter and a<br />

recombiner are the main components <strong>of</strong> optical interferometry; so with the<br />

splitting and merging <strong>of</strong> the traps the two main requirements for double well<br />

atom interferometers are available. Some theoretical models for such an inter-<br />

ferometer exist, but these in general work with an overly optimistic assumption.<br />

In chapter 3, this point is addressed. It is shown that for a single atom in a<br />

double well, a two mode approximation and the Bloch equations are helpful to<br />

understand the physics <strong>of</strong> a realistic double well interferometer. The process <strong>of</strong><br />

phase accumulation has been identified as Larmor precession. The two mode<br />

Bloch model itself is well known. The results from this model are thus not<br />

surprising, though they are counter-intuitive if one has worked with the old<br />

and too optimistic models. The model presented here can easily be adapted to<br />

further problems, for example, in some quantum computation schemes or in<br />

the field <strong>of</strong> solid state physics. In addition, an easy to implement and robust<br />

way to read out a double-well atom interferometer is presented.<br />

8.2 Outlook and Future<br />

The use <strong>of</strong> both magnetic and optical double well potentials for interferometry<br />

with Bose-Einstein condensates has already be shown in principle [Shi04, Shi05,<br />

181


8.2. Outlook and Future<br />

Jo07]. Here one <strong>of</strong> the demonstrations using the magnetic field <strong>of</strong> an atom<br />

chip [Jo07] examined increasing losses <strong>of</strong> the interference signal as a function<br />

<strong>of</strong> the merging time, as has been treated theoretically for single atoms in this<br />

thesis in part 3.3.2. The experimental data was compared with the results<br />

<strong>of</strong> the model and showed remarkable agreement. This publication shows how<br />

important a deeper understanding <strong>of</strong> the underlying physics is, as the authors<br />

only speculate about the reasons for the reducing signal. To understand the<br />

physics <strong>of</strong> a single atom in a double well is an important and crucial step to<br />

understand the behaviour <strong>of</strong> a BEC in such a system. The other magnetic<br />

trap experiment suffered from a non-repeatable phase relation between the<br />

BECs in the two wells [Shi05]. As the double well <strong>of</strong> optical dipole traps has<br />

comparable parameters [Shi04], the main difference seems to be the splitting<br />

time, which is a factor <strong>of</strong> 40 higher in the magnetic trap. The failure <strong>of</strong> being<br />

able to reproduce a phase relation between the wells can be attributed to the<br />

localisation <strong>of</strong> the atoms. This number squeezing leads to a phase diffusion.<br />

This magnetic version was implemented on an atom chip with two wires. Just<br />

as the edge <strong>of</strong> a permanent magnet is equivalent to a single current-carrying<br />

wire, a system <strong>of</strong> two current-carrying wires can be replaced by the two edges <strong>of</strong><br />

a slab <strong>of</strong> permanent magnetic material. The use <strong>of</strong> two slabs, equivalent to four<br />

wires, allows a field with a guiding potential between the slabs without the need<br />

for any external bias fields. One possible future application <strong>of</strong> our magneto-<br />

optical film is thus in structures like single and double slabs. These can then be<br />

used as waveguides. A double well created by a permanent magnetic film has<br />

been demonstrated [Hal07b]. Although not an atom interferometric process,<br />

the experiment allows the measurement <strong>of</strong> asymmetries between the wells. An<br />

accuracy <strong>of</strong> δg<br />

g = 10−4 for gravity fields has been inferred, and a read out<br />

process that was based on adiabatic splitting similar to the one presented in<br />

section 3.3.3 has been implemented. A more complicated structure <strong>of</strong> crossed<br />

182


Chapter 8: Summary and Outlook<br />

slabs has been proposed to create a 2D array <strong>of</strong> magnetic traps [Gha06]. This<br />

has possible applications mainly in the field <strong>of</strong> quantum information.<br />

Permanent magnets have many advantages over current-carrying wires.<br />

The most important is the absence <strong>of</strong> thermal noise that is always found in the<br />

current <strong>of</strong> a wire and leads to fluctuations in the magnetic field. These fluctu-<br />

ations can cause decoherence and are unwanted for atom interferometry. Also,<br />

it is possible to construct a vacuum chamber that does not need any electrical<br />

connections. The permanent magnet can act as a wire, and the desorption <strong>of</strong><br />

atoms from the chamber surface by light can act as the atom source [And01].<br />

This could create a device that should be robust enough to be taken into the<br />

field. An atom interferometer using these techniques could well be taken to<br />

measure gravitational gradients outside the laboratory.<br />

All-optical Bose-Einstein condensates have already been created [Bar01,<br />

Web03, Cen03, Kin05]. The most important up to date double well interfer-<br />

ometer with a BEC works with a BEC that is trapped optically, although here<br />

the BEC was created in a magnetic trap [Shi04]. When optical traps are used<br />

to manipulate the atoms for interferometry or quantum computation, the cre-<br />

ation <strong>of</strong> a BEC in situ in these traps would be a considerable simplification<br />

<strong>of</strong> the experiment. Once a BEC is trapped optically, micro-optical devices<br />

<strong>of</strong>fer a plethora <strong>of</strong> possibilities. The BEC could be loaded into waveguides<br />

[Dum02a, Kre04], as has been demonstrated already with cold atoms, to cre-<br />

ate a spatial atom interferometer. Loading the BEC into an array <strong>of</strong> microtraps<br />

leads to even more options. Each <strong>of</strong> these traps could be split and merged, so<br />

that an array <strong>of</strong> atom interferometers could be created. A possible scheme has<br />

been proposed in [Dud03]. Here a single measurement could include several<br />

tens <strong>of</strong> interferometers with one experimental run, and thus reduce the number<br />

<strong>of</strong> measurements needed to reduce statistical errors. Another possible applica-<br />

tion <strong>of</strong> a BEC loaded into an array <strong>of</strong> traps is in quantum information. Using a<br />

183


8.2. Outlook and Future<br />

BEC can be beneficial when the loading <strong>of</strong> the array can be achieved in a way<br />

similar to the transition between a BEC and a Mott insulator in an optical<br />

lattice [Gre02a]. This would lead to a fixed and well defined number <strong>of</strong> atoms<br />

in each trap and most importantly there would be the same number <strong>of</strong> atoms<br />

in each trap. Proposals exist for using an array <strong>of</strong> optical traps for quantum<br />

computation, including the localisation <strong>of</strong> a single atom in a double well as the<br />

qubit [Mom03]. Also, many proposals for atoms in optical lattices suffer from<br />

the drawback that the traps are not addressable. An array <strong>of</strong> microtraps can<br />

overcome this problem.<br />

184


Appendix A<br />

Reprints <strong>of</strong> selected<br />

Publications by the Candidate<br />

• A. I. Sidorov, R. J. McLean, F. Scharnberg, D. S. Gough, T. J. Davis, B.<br />

J. Sexton, G. I. Opat and P. Hannaford. Permanent-magnet microstructures<br />

for atom optics. Act. Phys. Pol. B 33, 2137-2155 (2002).<br />

• J. Y. Wang, S. Whitlock, F. Scharnberg, D. S. Gough, A. I. Sidorov,<br />

R. J. McLean and P. Hannaford. Perpendicularly magnetized, grooved<br />

GdTbFeCo microstructures for atom optics. J. Phys. D: Appl. Phys. 38,<br />

4015-4020 (2005). Included as appendix A.3.<br />

• B. V. Hall, S. Whitlock, F. Scharnberg, P. Hannaford and A.I. Sidorov.<br />

Bose-Einstein condensates on a permament magnetic film atom chip. In:<br />

Laser Spectroscopy; Proceedings <strong>of</strong> ICOLS 2005, E. A. Hinds, A. Ferguson<br />

and E. Riis (Editors), page 275-282 (World Scientific, Singapore,<br />

2005).<br />

• B. V. Hall, S. Whitlock, F. Scharnberg, P. Hannaford and A.I. Sidorov.<br />

A permanent magnetic film atom chip for Bose-Einstein condensation.<br />

J. Phys. B: At. Mol. Opt. Phys. 39, 27-36 (2006). Included as appendix<br />

A.2.<br />

• A. I. Sidorov, B. J. Dalton, S. Whitlock and F. Scharnberg. The asymmetric<br />

double-well potential for single-atom interferometry. Phys. Rev.<br />

A 74, 023612 (1-9) (2006). Included as appendix A.1.<br />

185


A.1. Asymmetric double-well potential for single-atom interferometry<br />

A.1 Asymmetric double-well potential for singleatom<br />

interferometry<br />

SIDOROV et al. PHYSICAL REVIEW A 74, 023612 2006<br />

PHYSICAL REVIEW A 74, 023612 2006<br />

harmon ic well app ears with asepa ration bet wee n minima <strong>of</strong><br />

2 2 . For zero and for large x the symmet ric pot ential<br />

app roximate s tha t for aqua ntu m harm onic oscillato r with<br />

freque ncy 0 and mass m. Key results in the pape r would<br />

still app ly if oth er suitable forms for the symmet ric pot ential<br />

are used .<br />

We den ote the eigenvectors <strong>of</strong>H ˆ as i and the ir ene rgy<br />

eigenvalues asEi, where i= 0,1,2,..., and Ei+1 Ei. The cor-<br />

Sec. III tha t provide s an adeq uat e descripti on <strong>of</strong> the dynam -<br />

ics <strong>of</strong> the splitting, holding, and recombi nation process es in<br />

the pres ence <strong>of</strong> an asymmet ric component , and is the n used<br />

Sec. IV in describing the interferometric process. The validity<br />

<strong>of</strong> the two-mo de appr oach is explored in Sec. V<br />

throug h the comparison <strong>of</strong> predictions <strong>of</strong> the Bloch vecto r<br />

mode l with the results <strong>of</strong> direct nume rical simulations <strong>of</strong> the<br />

time-depend ent multimod e Schrödinger equation .Adiscussion<br />

<strong>of</strong> results follows in Sec. VI and include saschem e to<br />

mea sure the popul ation <strong>of</strong> the excited state . Theoretical<br />

details are dealt with in the Append ix.<br />

Asymmetric double-well potential for single-atom interferometry<br />

A. I. Sidorov, * B. J. Dalton, S. M. Whitlock, and F. Scharnbe rg †<br />

ARC Centre <strong>of</strong> Excellence for Quantu m-Atom Optics<br />

and Centre for Atom Optics and Ultrafast Spect roscopy, Swinburn e <strong>University</strong> <strong>of</strong> <strong>Technology</strong> , Melbou rne, Victoria 3122, Australia<br />

Received 20 March 2006; published 18 August 2006<br />

II. THEORETICALFRAME<br />

In gene ral, the evolut ion <strong>of</strong> asingle ato m in an interferomete<br />

r must be describe d using athree-dimensio nal quan tum<br />

treatment . Howeve r, for asystem <strong>of</strong> cylindrical symmetryas<br />

is prese nt in typical ato m chip experime ntsit is possible to<br />

ignore excitations associate d with the two tightl y confined<br />

dime nsions, as long as the dynamic s through out the proces s<br />

is restricted to the dimension <strong>of</strong> weak confineme nt longitudinal<br />

splitting . In this system it is pos sible to redu ce the<br />

quantu m trea tmen t to tha t <strong>of</strong> aone-dime nsiona l problem .<br />

We consider the one- dime nsiona l evolution <strong>of</strong> asingleato<br />

m system due to atime-depend ent Hamiltonian Hˆ t that<br />

can be written as the sum <strong>of</strong> asymmetric HamiltonianH ˆ<br />

0 t<br />

and an asymmetric potentialV ˆ as xˆ<br />

Hˆ t = Hˆ 0 t + Vas ˆ xˆ<br />

= pˆ 2<br />

2 + V0 ˆ xˆ,t + Vˆ as xˆ , 1<br />

ˆ<br />

xˆ<br />

V0 xˆ,t = 1 + t − 2 2 1/2<br />

, 2<br />

2<br />

where aspecific form for the symmetric potentialV ˆ<br />

0 is chosen<br />

26 . The Hamiltonian and oth er physical quantit ies have<br />

bee n writte n in dime nsionles s quantu m harmo nic oscillator<br />

units associated with atomi c massm and angu lar frequency<br />

0. With the original quan tities denote d by primes we have<br />

xˆ = xˆ<br />

, pˆ =<br />

a0 a respo ndin g qua ntities for the symmet ric compo nen t <strong>of</strong> the<br />

Hamiltonian, Hˆ 0,<br />

will be den ote d Si and ESi. Both sets <strong>of</strong><br />

eigenvecto rs are orthonor mal, and all ene rgies and eigenvectors<br />

are time depe nde nt.Hˆ 0 is symmet ric and the ground<br />

state S0 is symmet ric and deno ted asS , while the first<br />

excited stat e S1 is ant isymmet ric and den ote d asAS . Their<br />

ene rgies are den ote dE S and EAS. The one -dimensional nat ure<br />

<strong>of</strong> the system allows real eigenfunctions i x , Si x to be<br />

chosen. In this case the geo met ric pha se27 is zero.<br />

We can illustrate the gen eral beh avior <strong>of</strong> the lowest few<br />

ene rgy eigenvalues E0,E1,E2,E3, ... , <strong>of</strong> the tota l Hamilton<br />

ian H<br />

0<br />

pˆ ,<br />

t = 0t , a0 = . 3<br />

m 0<br />

The dimensio nless Hamiltonians, potential s, and ene rgies are<br />

obtaine d by dividing the original quantitie s by 0. Vas ˆ will<br />

be taken as alinear function <strong>of</strong>xˆ.<br />

The symmetric potenti al depend s on a time-de pen den t<br />

splitting paramete r , whose chan ge from zero to alarge<br />

value and back to zero aga in convenientl y describ es the splitting,<br />

holding, and recom binati on processe s with periods Ts, Th, and Tr, respec tively. For zero the symmetric pot ential<br />

involves asingle quar tic well. When it is large adoub le<br />

ˆ as the splitting parame ter is increased from zero to<br />

a large value and back. At the begi nning and the end <strong>of</strong> the<br />

process whe n 0 the ene rgy eigenvalues are well sep arated<br />

. Here the effect <strong>of</strong> asymmetryV ˆ<br />

as is small and the<br />

eigenvalues and eigenvecto rs resemb le tho se for the symmet -<br />

ric Hamiltonian. When the splitting paramet er increase s and<br />

the trap ping pot ential change s to adou ble well, pairs <strong>of</strong> eigen<br />

values E0 and E1, E2 and E3, etc. beco me very close. At<br />

this stag e the qua ntu m system is very sensit ive to the presence<br />

<strong>of</strong> Vˆ as which breaks the symmet ry, allows tran sitions<br />

bet wee n 0 and 1 to occur, and causes the eigenvecto rs<br />

0 and 1 as well as 2 and 3 to be localized in the<br />

individual wells in this far split regime.<br />

Initially, the ato m is prepare d in the lowest ene rgy eigenstate<br />

0 <strong>of</strong> the single well. Transitions to excited state s are<br />

supp ressed if the time scale for splitting and recombinatio n is<br />

much longer tha n the inverse freque ncy gap bet wee n the relevant<br />

state s. The ene rgy gap bet wee nE 0 and E2 is always<br />

larger tha n the gap bet wee nE 0 and E1, and by choo sing<br />

app ropriate time scales we can isolate the two lowest ene rgy<br />

eigensta tes 0 and 1 from highe r excited state s 2 ,<br />

3 , etc. , but still allow for tran sitions bet wee n the two<br />

lowest ene rgy eigensta tes to occur. As aconsequenc e the<br />

dyna mics <strong>of</strong> the DWAI can be treate d unde r the two-mode<br />

app roximation, in which only the two lowest ene rgy eigenstate<br />

s <strong>of</strong> the tota l HamiltonianHˆ and the symmet ric Hamilton<br />

ian Hˆ 0 nee d to be considered . In this case the first excited<br />

stat e probab ility a mea surable qua ntitycan vary from zero<br />

to one .Aproposal for mea suring the excited stat e popu lation<br />

is out lined in Sec. VI.<br />

Using the two-mod e app roximation expression s for the<br />

lowest two ene rgy eigenvalues E0,E1 and eigenvecto rs<br />

0 , 1 for the Hamiltonian Hˆ will be obt ained .Astan -<br />

dard mat rix mech anics app roach will be used , but instea d <strong>of</strong><br />

using the symmet ric pot ential ene rgy eigenvecto rs S , AS<br />

as basis vectors, we use the left, rightL-R basis vectors L ,<br />

R , which are defined by<br />

023612-2<br />

We conside r the evolutio n <strong>of</strong> asingle-ato m wave function in atime-depen den t dou ble-well interferomete r in<br />

the presence <strong>of</strong> aspatially asymmet ric pot ential. We examine acase where asingle trapp ing pot ent ial is split<br />

into an asymmet ric dou ble well and the n recomb ined again. The interferomete r involves ameas urem ent <strong>of</strong> the<br />

first excited stat e pop ulation as asensitive mea sure <strong>of</strong> the asymm etric pot ent ial. Based on atwo-mode approximation<br />

aBloch vecto r mode l provides asimple and satisfactory des cription <strong>of</strong> the dynami cal evolution.<br />

We discuss the roles <strong>of</strong> adiabaticity and asymmet ry in the doubl e-well interferomete r. The Bloch mode l allows<br />

us to accoun t for the effects <strong>of</strong> asymmet ry on the excited stat e pop ulation throug hou t the interferometric<br />

process and to choos e the app ropriate splitting, holding, and recombin ation periods in order to maximize the<br />

outpu t signal. We also compare the out come s <strong>of</strong> the Bloch vector mode l with the results <strong>of</strong> num erical<br />

simulation s <strong>of</strong> the multistat e time-dep enden t Schrödinger equatio n.<br />

DOI: 10.1103/PhysRevA.74.023612 PACS num ber s : 39.20. q, 03.75.Dg, 03.75.Be<br />

trap ping pot ential. An ato m is initially prepared in the<br />

grou nd stat e <strong>of</strong> asingle symmet ric trap ping pot ential, which<br />

is the n split into asymmet ric doub le well. Aspatially asymmet<br />

ric pot ent ial is the n app lied and anona diabatic evolution<br />

leads to tran sitions bet wee n grou nd and excited state s. The<br />

asymmetry is the n switched <strong>of</strong>f and the doubl e well is recomb<br />

ined into the original pot ential. The popul ation <strong>of</strong> the<br />

excited stat e mea sures the effect <strong>of</strong> the asymmetric pot ential.<br />

The DWAI can be considered as a qua ntu m-stat e Mach-<br />

Zehn der interferomet er where the evolution <strong>of</strong> the qua ntu m<br />

stat e via the two sepa rate d wells is analo gou s to the propagat<br />

ion <strong>of</strong> an opt ical field via two pat hways.<br />

However, thes e treatm ent s ignore the effect <strong>of</strong> asymmet ry<br />

during the splitting and merging stag es. In reality asymmetry<br />

is always present and could be the result <strong>of</strong> imperfect horizontal<br />

splitting introducing agravity-based asymmetry, extern<br />

al spatially variable magn etic and electric fields or different<br />

left and right trap freque ncies. We show tha t the presence<br />

<strong>of</strong> small asymmet ries has dram atic consequ ences on the interferom<br />

etric process. We have produce dasimple mode l in<br />

term s <strong>of</strong> aBloch vector evolution tha t ena bles us to consider<br />

a splitting-ho lding-merging sequ ence involving a doub le<br />

well and takes into accoun t the presence <strong>of</strong> asymmetry at all<br />

stag es. The two key paramet ers are the ene rgy gap bet wee n<br />

the lowest two state s <strong>of</strong> the symmet ric comp one nt <strong>of</strong> the<br />

trap ping pot ential and an asymmetry parame ter, which is related<br />

to mat rix elem ent s <strong>of</strong> the asymmet ric compo nen t <strong>of</strong> the<br />

pot ential. Nona diabat ic evolution only occurs during the<br />

splitting and recombining stag es whe n the torq ue vector<br />

change s much more rapidly compa red to the Larmor precession<br />

<strong>of</strong> the Bloch vector. It is impo rtan t tha t the torq ue vector<br />

rema ins constan t during the holding stag e, and tha t this period<br />

is long compa red to the splitting and recombin ation<br />

times. In this case the final excited stat e popu lation is asinusoidal<br />

function <strong>of</strong> the holding time, with aperiod det ermined<br />

via the asymmetry paramet er.<br />

In this pap er we consider the dyna mics <strong>of</strong> asingle ato m in<br />

an asymmetric DWAI, with the basic theo retical framew ork<br />

being out lined in Sec. II. Using atwo-mod e app roximation<br />

we develop aBloch vector mode l for time-depen den t DWAI<br />

I. INTRODUCTION<br />

The evolution <strong>of</strong> aquantu m system in adou ble-well potenti<br />

al has bee n the sub ject <strong>of</strong> num erou s theore tical studi es.<br />

These include trea tment s <strong>of</strong> Joseph son-like oscillations1,2 ,<br />

dynami c splitting 3 , and interference 4 <strong>of</strong> Bose-Einstein<br />

condensate s BECs . The interpreta tion <strong>of</strong> thes e effects is<br />

base d on the app roach5 , in which interference patte rns are<br />

seen to evolve as aresult <strong>of</strong> successive boso n mea sureme nts<br />

which do not ident ify the originating cond ensate . The product<br />

ion <strong>of</strong> cold atom s and BEC in microtraps on ato m chips<br />

6–8 and in micro-opti cal system s 9 has stimulate dagrea t<br />

interes t towa rds novel implementa tions <strong>of</strong> ato m interferomete<br />

rs 10–15 tha t are base d on the use <strong>of</strong> dou ble-well potentials.<br />

Double-well ato m interferome ters DWAI <strong>of</strong> bot h the<br />

single-atom and the BEC varieties are well suited to implemen<br />

tat ion on an ato m chip. Here micr<strong>of</strong>abricate d structures<br />

allow us aprecise control on asubmi cron scale over the<br />

splitting and merging processes . The key processe s <strong>of</strong> splitting<br />

16–18 and merging 19 <strong>of</strong> cold ato mic clouds and<br />

even interfe rence <strong>of</strong> aBEC after splitting in adoubl e well<br />

20–22 have bee n already demo nstrat ed. Althoug h the<br />

impleme ntati on <strong>of</strong> a DWAI using a BEC can lead to a<br />

N-fold enhance men t in precision mea sureme nts23 , pha se<br />

diffusion 24 associate d with mea n field effects is <strong>of</strong> concern<br />

20 . Double-well interfe rometry with asingle ato m can<br />

allow us alonge r mea surem ent time and in this regar d has a<br />

potentia l advanta ge. An on-chip single-atom interferome ter<br />

can be integra ted with the source <strong>of</strong> ato ms in a ground<br />

state —the Bose-Einstein condens ate —and be used for<br />

sensitive mea surem ent s <strong>of</strong> gravitationa l fields. DWAI may<br />

also be app lied to mea sure collisiona l phas e shifts induc ed<br />

by the ato m-ato m interaction, which is useful for qua ntum<br />

compu tatio n processe s25 .<br />

Two proposed scheme s <strong>of</strong> asingle-atom DWAI involve<br />

time-depend ent transverse10 and axial 11 splittings <strong>of</strong> a<br />

186<br />

*Email addres s: asidorov@swin.edu.au<br />

†<br />

Also at IQO, <strong>University</strong> <strong>of</strong> Hannov er.<br />

1050-2947/2006/742 /023612 9 023612-1<br />

©2006 The American Physical Society


SIDOROV et al. PHYSICAL REVIEW A 74, 023612 2006<br />

ASYMMETRIC DOUBLE-WELL POTENTIAL FOR ? PHYSICAL REVIEW A 74, 023612 2006<br />

Chapter A: Reprints <strong>of</strong> selected Publications by the Candidate<br />

t = C L L + CR R . 12<br />

1<br />

E0 = 2 E 1<br />

S + EAS − 2 ,<br />

20<br />

1<br />

0.8<br />

Our interferomet er will be described in term s <strong>of</strong> the Bloch<br />

vector and its dynamics determ ined via Bloch equat ions. We<br />

now introduce Pauli spin ope rato rs and the Bloch vector.<br />

Time-depen dent Pauli spin ope rato rs ˆ<br />

a a=x,y,z are defined<br />

in the Schrödinger picture<br />

ˆ<br />

x = R L + L R ,<br />

0.7<br />

1<br />

E1 = 2 E 1<br />

S + EAS + 2 , 8<br />

15<br />

0.5<br />

0.6<br />

as<br />

where the quantit y gives the ene rgy gap for the total<br />

Hamiltonian Hˆ and is defined by<br />

V<br />

0.5<br />

,<br />

)<br />

0<br />

x<br />

R L − L R ,<br />

ˆ<br />

1<br />

y =<br />

i<br />

= 0 2 2<br />

+ Vas = E1 − E0. 9<br />

In term s <strong>of</strong> the laser-driven two-level ato m analogy, would<br />

be ana logou s to the gene ralized Rabi frequency.<br />

The ortho norma l ene rgy eigenve ctors for the tota l Hamilton<br />

ian Hˆ are given by<br />

(<br />

10<br />

0<br />

ψ<br />

0.4<br />

V<br />

,<br />

0.3<br />

5<br />

-0.5<br />

0.2<br />

(a)<br />

0.1<br />

ˆ<br />

z = R R − L L . 13<br />

From its mat rix represen tat ion in theL-R basis 5 , the<br />

dimensionless Hamiltonian ope rato rHˆ in the Schrödinger<br />

picture can be expressed in term s <strong>of</strong> the Pauli spin ope rato rs<br />

as<br />

0<br />

-6 -4 -2 0 2 4 6<br />

-1<br />

0 2 4 6 8 10 12<br />

splitting parameter<br />

0<br />

R ,<br />

1 − V<br />

2<br />

L +<br />

1 + V<br />

2<br />

0 =<br />

x<br />

20<br />

1<br />

R , 10<br />

1 + V<br />

2<br />

L −<br />

1 − V<br />

2<br />

1 =<br />

FIG. 1. Energy difference between ground and first excited<br />

states solid line for Vˆ as=<br />

0.02xˆ as afunction <strong>of</strong> the splitting paramete<br />

r . Dotted line—energy difference 0 for symmetric Hamiltonian,<br />

dashe d line—asymmetry quantityV as.<br />

15<br />

0.5<br />

Hˆ = 1<br />

2 01ˆ + ˆ<br />

x x + ˆ<br />

y y + ˆ<br />

z z , 14<br />

where the effect <strong>of</strong> asymmet ry is now represente d by the<br />

variable<br />

)<br />

whe re<br />

x<br />

(<br />

S + AS ,<br />

V<br />

10<br />

0<br />

ψ<br />

L = 1<br />

2<br />

0 = E S + EAS ,<br />

V = V as . 11<br />

x = − 0, y = 0, z = Vas. 15<br />

It is conve nient to introduce aso-called torque vector, defined<br />

as = x, y, z .<br />

The Bloch vector is defined to have comp one nts which<br />

are the expe ctation values <strong>of</strong> the Pauli spin ope ratoˆrsa in<br />

the qua ntu m state . These compon ent s will be deno ted as<br />

a. Hence in the Schrödinger picture<br />

a = t a<br />

ˆ t t a = x,y,z . 16<br />

The Bloch vector is defined as = x, y, z . The Bloch<br />

comp one nts are bilinear functions <strong>of</strong> the amp litudes CL and<br />

CR. The evolution <strong>of</strong> the DWAI system is now described by a<br />

set <strong>of</strong> real variables x, y, z and each <strong>of</strong> thes e variables has<br />

a certain physical mea ning. The compon ent z is amea sure<br />

<strong>of</strong> the imbalance <strong>of</strong> the atom ic popu lation <strong>of</strong> the localized<br />

state s L , R . The compon ent x is amea sure <strong>of</strong> the atom ic<br />

pop ulation imbalance betwee n the delocalized stateS s ,<br />

AS , as can be seen if the qua ntu m stat e is expan ded in the<br />

symmet ric basis. For x= + 1 all the popul ation is in the symmet<br />

ric state S , for x= −1 it is all in the antisymme tric state<br />

AS . It is thu samea sure <strong>of</strong> the excitation <strong>of</strong> the first excited<br />

stat e in the unsplit trap regime.<br />

Equa tions for the compon ent s <strong>of</strong> the Bloch vector can be<br />

obtained from Heisenberg equati ons for the Pauli spin ope rators.<br />

The derivation must take into accoun t the present situation<br />

where the Pauli spin ope rato rs are explicitly time depen<br />

dent since the basis vectorsL , R chang e with time.<br />

This differs from the stan dard situation <strong>of</strong> time indepe ndent<br />

basis vectors 28,29 . However, the add itional term in the<br />

Heisenberg equati ons can be shown to cont ribute zero to the<br />

Bloch equ ations due to the two eigenfunctions in the sym-<br />

On substituting for L , R the eigenve ctors for the total<br />

Hamiltonian can be related to thos e for the symmetric<br />

Hamiltonian Hˆ 0.<br />

At the beginn ing and end <strong>of</strong> the interferom -<br />

ete r process we find tha t the asymmet ry parame ter Vas is<br />

small comp ared to the symmetric ene rgy gap0. For V 1<br />

the eigenve ctors 0 , 1 become similar to S and AS ,<br />

respe ctively. For Vas 0 V 1 , the eigenve ctors 0 ,<br />

1 become equa l to L , R , respe ctively, the localized<br />

eigenve ctors for the separa te wells.<br />

The behavior <strong>of</strong> the quantit ies , Vas, and 0 as the splitting<br />

paramet er is chan ged is shown in Fig.1 for the case<br />

where the asymmetric potenti alV as x varies linearly with<br />

the coordinatex, spe cifically withˆ Vas = 0.02xˆ. The symmetric<br />

ene rgy gap 0 become s close to zero for 4 and the n the<br />

actua l ene rgy gap is appr oximately given byV as. The energy<br />

eigenfun ctions 0 x and 1 x for different splitting<br />

paramete rs are depicted in Fig.2, again with Vˆ as=<br />

0.02xˆ.<br />

The beha vior <strong>of</strong> the potenti alV x =V0 x +Vas x is also<br />

shown . For small Fig. 2 a the potentia l is asingle well<br />

and the eige nfunctions are app roximately symmetric and antisymmet<br />

ric. For larger Fig. 2 c the potent ial is adouble<br />

well, which still appea rs to be symmetric. However, even<br />

with asmall asymmetry in the potenti al the eigenfun ctions<br />

are no longer symme tric and antisymme tric, but instea d are<br />

each localized in separa te wells. This sensitivity <strong>of</strong> the<br />

eigenfun ctions to asmall asymmetry is critical to the perfo rman<br />

ce <strong>of</strong> the presen t interfe romete r.<br />

5<br />

-0.5<br />

187<br />

(b)<br />

0<br />

-6 -4 -2 0 2 4 6<br />

-1<br />

x<br />

20<br />

1<br />

R = 1<br />

S − AS . 4<br />

2<br />

The stat es L , R are orthonorm al and for large correspon<br />

d to an atom localized in the left or right well, respectively.<br />

However, even for a single well theL-R basis is still<br />

applicable. The matrix for the HamiltonianHˆ in the L-R<br />

basis is given by<br />

15<br />

0.5<br />

, 5<br />

− Vas − 0<br />

− 0 + Vas<br />

1<br />

+<br />

2<br />

1 0<br />

0 1<br />

Hˆ L−R = 1<br />

2 ES + EAS )<br />

x<br />

(<br />

10<br />

0<br />

ψ<br />

where the order <strong>of</strong> the columns and rows isL,R and we<br />

define the convenient real quantities<br />

V<br />

5<br />

-0.5<br />

0 = E AS − ES, 6<br />

(c)<br />

Vas = R Vas ˆ R − Lˆ Vas L<br />

0<br />

-1<br />

-5 0 5<br />

x<br />

FIG. 2. Stat ionary eigenfunctions <strong>of</strong> the ground statedasheddotted<br />

line and the first excited state dotted line for different<br />

values <strong>of</strong> the splitting paramete r 1 a , 2.5 b , and 5 c . The<br />

potentialV x is shown as the solid line.<br />

II. BLOCH VECTOR MODEL<br />

We can expres sagen eral time-depen dent normalized stat e<br />

vector as aquantu m sup erposition <strong>of</strong> the state sL and<br />

R<br />

Hamiltonian transition frequency 0 is analogous to the<br />

Rabi frequency, while the quantityV as is analogous to<br />

the detuning.<br />

The energy eigenvalues for the total HamiltonianH ˆ are<br />

obtained from the dete rminental equation as the eigenvalues<br />

<strong>of</strong> the matrix Hˆ L−R , and are given by<br />

=− S Vˆ as AS + AS Vas ˆ S . 7<br />

The derivation <strong>of</strong> the Hamiltonian matrix uses the<br />

symmetry prop erties <strong>of</strong> S , AS , and the reality <strong>of</strong> the<br />

eigenfunctions. The total energy for the symmetric Hamiltonian<br />

is given by ES+EAS, and the energy gap is given<br />

by 0. The quantity Vas describes the asymmetry <strong>of</strong> the<br />

system, and would be zero if the Hamiltonian was symmetric.<br />

The second equation relatesV as to <strong>of</strong>f-diagon al elements<br />

<strong>of</strong> the asymmetric contribution to the Hamiltonian, indicating<br />

its role in causing transitions between the eigenstate s<br />

S , AS <strong>of</strong> the symmetric Hamiltonian. The Hamiltonian<br />

matrix 5 is analogous to tha t for a two-level atom interacting<br />

with amon ochromatic light field28 . The symmetric<br />

023612-4<br />

023612-3


A.1. Asymmetric double-well potential for single-atom interferometry<br />

SIDOROV et al. PHYSICAL REVIEW A 74, 023612 2006<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

- 0.2<br />

- 0.4<br />

- 0.6<br />

(a)<br />

- 0.8<br />

-1<br />

0 10 20 30 40 50 60<br />

time<br />

1<br />

0.8<br />

ASYMMETRIC DOUBLE-WELL POTENTIAL FOR ? PHYSICAL REVIEW A 74, 023612 2006<br />

σ x , σ y , σ z<br />

0.6<br />

0.4<br />

(b)<br />

0.2<br />

β / 12.<br />

5,<br />

V,<br />

P<br />

1<br />

0 10 20 30 40 50 60<br />

time<br />

0<br />

FIG. 4. a Time evolution <strong>of</strong> the Bloch vector componen ts x<br />

solid line , y dashed line, and z dotted line for Ts=Th=Tr = 20 and max= 12.5; b time evolution <strong>of</strong> the first excited state<br />

popu lation P1 dotted line, the asymmetry paramete rV dashed<br />

line , and the splitting paramete r / m ax solid line .<br />

the n chang ed linearly to zero during the recombination<br />

period. The dynamical beha vior <strong>of</strong> the Bloch vector<br />

comp onent s is shown in Fig.4 a along with the time<br />

depend ence <strong>of</strong> the asymmetry paramete rV=V as/ , the<br />

aligned with the x axis. For small values <strong>of</strong> the splitting<br />

paramete r the abso lute value <strong>of</strong> the torqu e vector is mainly<br />

dete rmined by the symmetric ene rgy gap0 Fig. 1 and its<br />

direction remains along the −x direction. The pos ition <strong>of</strong> the<br />

Bloch vector remains mostly along the +x direction Fig.<br />

3 b during early stag es <strong>of</strong> the splitting process. When the<br />

splitting paramete r is increased furthe r the decreasing ene rgy<br />

gap 0 becomes comparable with and later much smaller<br />

tha n the asymmetry quantityV as. As a result the torque vector<br />

rotate s in ax-z plane until it is aligned along thez direction<br />

0,0,V as . It is impor tan t to make this chan ge<br />

nonadia batically so tha t the Bloch vector does not follow the<br />

torqu e rotation. If the Bloch vector were to follow the<br />

change s <strong>of</strong> the torqu e vector adiabatically the atom will always<br />

stay in the ground stat e and no interference would be<br />

observed.<br />

During the holding stag e the torque vector is constan t and<br />

the Bloch vector precesses around the torqu e vector with a<br />

constan t ang ular velocityV as, and hen ce both thex and y<br />

compo nent s oscillate with aperiod 2/V as Fig. 3 c . In an<br />

ideal doub le-well interferomete r the splitting and recomb ination<br />

stage s are short and the value <strong>of</strong> thex compo nen t does<br />

not chang e much during the se stage s, so thatx T which<br />

defines the final excited stat e popu lationis basically given<br />

by its value at the end <strong>of</strong> the holding period. The simple<br />

beha vior during the holding period indicate s tha t the excited<br />

stat e popu lation would have aperiod 2/V as considered as a<br />

function <strong>of</strong> holding time. A similar description in terms <strong>of</strong><br />

the evolution <strong>of</strong> a Bloch vector also app lies to the scheme<br />

described in Ref. 11 , though the dynamical beha vior <strong>of</strong> the<br />

Bloch vector is different .<br />

The beha vior <strong>of</strong> the interferomete r may also be described<br />

in terms <strong>of</strong> time-dependen t state sL , R , which during the<br />

holding period represent atoms localized in the left and right<br />

wells. The interferomete r process involves the transition<br />

S 0 ? AS T , which involves two pathw ays S 0<br />

? L T/ 2 ? AS T and S 0 ? R T/ 2 ? AS T , involving<br />

two poss ible localized intermediate stat es associated<br />

with the left or right wells. The overall transition amp litude<br />

is the sum <strong>of</strong> amplitudes for the two pathw ays, and depend -<br />

z<br />

(a)<br />

metric basis bein g real and having opposit e symmetrysee<br />

Appendi x . The Bloch equ ation s are given by dt d/dt<br />

Ω<br />

d t x = − Vas y,<br />

d t y = Vas x + 0 z,<br />

d t z = − 0 y, 17<br />

y<br />

σ<br />

and can be solved nume rically using the Runge -Kutt a<br />

algorithm . In vecto r nota tion the Bloch equ ation s are<br />

x<br />

d t = . 18<br />

This form <strong>of</strong> the Bloch equation s is adirect conseq uence<br />

<strong>of</strong> the equiva lence <strong>of</strong> the two-mod e dou ble-well interferom -<br />

ete r to aspin 1<br />

2 system . The Bloch vecto r precesse s at the<br />

Larmo r frequenc y arou nd the torq ue vecto r, which in detail<br />

is<br />

z<br />

(b)<br />

= − 0,0,Vas . 19<br />

If ther e is no asymmetr y, thex compo nen t <strong>of</strong> the Bloch<br />

vecto r rema ins unch ange d, while its compo nen t in the y-z<br />

plane just rotate s abou t thex axis Fig. 3 .<br />

Ω<br />

188<br />

y<br />

σ<br />

x<br />

splitting paramete r and the excited stat e popu lationP 1<br />

Fig. 4 b . The paramete rs used areV ˆ<br />

as=<br />

0.02xˆ, max= 12.5,<br />

and Ts= 20, Th= 20, Tr= 20 in dimensionless harmonic<br />

oscillator units. Here we observe comp lex oscillatory<br />

beha vior for the x and y comp onent s <strong>of</strong> the Bloch vector<br />

which occurs during the splitting and merging stag es. During<br />

the holding stag e the y exhibit simp le periodic variations<br />

ing on the relative phase betwee n the se amplitudes eithe r<br />

with frequencyV as= 0.2. At the same time thez comp onent<br />

constructive or destructive interference may occur. For maxi- develops a small neg ative value during splitting and<br />

mum contrast it is desirable tha t the magnitudes <strong>of</strong> the two increases the abso lute value even furthe r during merging.<br />

partial amp litudes be equal, so tha t during the holding period The x comp onen t reaches a neg ative value <strong>of</strong> −0.9 at the<br />

the popu lations <strong>of</strong> the left and right well stat es should be end <strong>of</strong> the process. This corresponds to an excited state<br />

abou t the same. After optimal splitting thez compo nen t <strong>of</strong> popu lation <strong>of</strong> 0.95 and represent s a case <strong>of</strong> constructive<br />

the Bloch vector z should be kept close to zero during the interference.<br />

holding period, however a phase difference between the lo- By monitoring the beha vior <strong>of</strong>P 1 during the interferometcalized<br />

state s accumu lates. Only at the end <strong>of</strong> the recomb iric process we can see whe n nonad iabatic evolution occurs.<br />

nation stag e this phase is translated into the popu lation <strong>of</strong> the The pop ulationP1 change s from 0to 0.47 Fig. 4 b at the<br />

excited state .<br />

beginning <strong>of</strong> the splitting process and does not reach the<br />

optimal value<br />

V. RESULTS OF NUMERI CALSIMULATIONS<br />

We studied the evolution <strong>of</strong> a Bloch vector and the<br />

popu lation <strong>of</strong> the excited stat e by solving Eqs.17 numerically.<br />

The splitting paramete r is chan ged linearly from zero<br />

up to a maximum max during the splitting period. It is then<br />

held constan t at max during the holding period, and<br />

1<br />

2 as a result <strong>of</strong> the nonzeroz comp onen t <strong>of</strong> the<br />

Bloch vector. The variable P1 does not chang e during the<br />

adiabatic precession <strong>of</strong> the Bloch vector around the torque<br />

vector during the rest <strong>of</strong> the splitting, holding, and the begin -<br />

ning <strong>of</strong> recomb ining stag es. It again exhibits drastic change s<br />

in a short period during the remerging whe n the torque vector<br />

rotate s rapidly and the Larmor frequency is relatively<br />

small.<br />

z Ω<br />

(c)<br />

IV. MODELOFA SINGLE-ATOM<br />

DOUBLE-WELATOM INTERFEROMETER<br />

In the single-ato m interferome ter unde r considerat ion the<br />

ato m is always locate d in atrappin g pot ential, which changes<br />

from a single well to adoubl e well—which, in gene ral, is<br />

slightly asymmetric—and back again to the original single<br />

well. The interferom ete r is used to mea sure the effects <strong>of</strong> this<br />

asymmetr y, the cause <strong>of</strong> which may be <strong>of</strong> mea surable interest<br />

e.g., as in agravity gradio mete r. The ato m is initially in the<br />

groun d state 0 0 <strong>of</strong> the original unsplit pote ntial, and as<br />

Vas is the n small compared to 0, 0 0 is the n appr oximatel<br />

y the same asS 0 . The popul ation <strong>of</strong> the excited stat e<br />

at the end <strong>of</strong> the recombinati on process is the mea surable<br />

interferome ter out put . The probabil ityP 1 <strong>of</strong> findin g the atom<br />

in the uppe r ene rgy eigenstat e at any time is given by<br />

y<br />

σ<br />

x<br />

2 P1 = 1 , 20<br />

and this will rema in zero unless an asymmetr y is present<br />

toget her with suitabl y short splitting and recom bining stages<br />

for the interferom ete r process—so tha t transition s occur<br />

bet ween 0 and 1 due to the presenc e <strong>of</strong>Vas ˆ .<br />

We find that<br />

· . 21<br />

1<br />

+<br />

2<br />

1<br />

P1 =<br />

2 1 + 2 1<br />

zV − x 1 − V =<br />

2<br />

FIG. 3. Evolution <strong>of</strong> the Bloch vector and the torque vector<br />

at differen t mome nts <strong>of</strong> the splitting stag e:a at the beg inning<br />

when 0 Vas and − 0,0,0 , b when 0=Vas, and c ,<br />

when Vas 0 and 0,0,Vas .<br />

recombining stag es. At the start <strong>of</strong> the process the Bloch and<br />

torq ue vectors are antiparallelFig. 3 a and app roximately<br />

At the beg inning and end <strong>of</strong> the interferomete r proce ss<br />

V 1 and henc e the probabil ityP 1 only dep end s on thex<br />

compo nen t <strong>of</strong> the Bloch vecto r. The probabilityP 1 T thus<br />

depend s on how this componen t has change d from its initial<br />

value <strong>of</strong> 1. We can, there fore, describ e the dynamical behav -<br />

ior <strong>of</strong> the single-ato m interferome ter in term s <strong>of</strong> the evolution<br />

<strong>of</strong> the Bloch vector during the splitting, holding, and<br />

023612-6<br />

023612-5


SIDOROV et al. PHYSICAL REVIEW A 74, 023612 2006<br />

ASYMMETRIC DOUBLE-WELL POTENTIAL FOR ? PHYSICAL REVIEW A 74, 023612 2006<br />

Chapter A: Reprints <strong>of</strong> selected Publications by the Candidate<br />

The equat ion <strong>of</strong> motion for the Bloch vector compone nts<br />

can be derived using the Heisen be rg picture via<br />

ˆ H<br />

a 0 , A1<br />

d<br />

dt<br />

tion betwee n out come s <strong>of</strong> two mod els. For large values <strong>of</strong><br />

the asymmetr y frequencyV as it is impossible to adiabat ically<br />

isolate two lower state s and the transition s to highe r stat es<br />

have to be taken into account.<br />

0.9<br />

1<br />

(a)<br />

1<br />

(a)<br />

d<br />

a = 0<br />

dt<br />

0.8<br />

0.7<br />

0.5<br />

whe re the Heisen be rg equati on <strong>of</strong> mot ion for the Pauli spin<br />

operat or in dimensionless units is<br />

VI. DISCUSION AND CONCLUSIONS<br />

(c)<br />

0.6<br />

0.5<br />

F<br />

population P1<br />

d<br />

ˆ H<br />

a = − i ˆ H<br />

a , Hˆ H + ˆ<br />

a<br />

dt<br />

t<br />

H<br />

. A2<br />

The derivation involves the use <strong>of</strong> the commu tation rules<br />

for the Pauli spin operat ors. For the first term , we have after<br />

substi tuting forHˆ from Eq. 14<br />

b a<br />

ˆ H<br />

, ˆ<br />

b<br />

0 a<br />

ˆ ,1ˆ +<br />

b=x,y,z<br />

− i<br />

2<br />

− i a<br />

ˆ H , Hˆ H =<br />

= ˆ H a. A3<br />

We have app lied aBloch vecto r mode l to des cribe the<br />

quantum -stat e interference <strong>of</strong> asingle-ato m wave function in<br />

a time-variable asymmetric double -well pote ntial. The probability<br />

<strong>of</strong> findin g the ato m in the first excited stat e is closely<br />

associate d with the mag nitude <strong>of</strong> the spatially dependen t externa<br />

l pote ntial and could be used as amea sure <strong>of</strong> the app lied<br />

asymmetr y. Transition s betwe en groun d and excited state s<br />

occur during the splitting and recom binati on stages . Larmo r<br />

precess ion <strong>of</strong> the Bloch vecto r during the holdin g stag e is<br />

induc ed by the asymmet ry, will effect an interfe rome tric<br />

phase , and dete rmine the final value <strong>of</strong> the excited state<br />

population. The evolut ion <strong>of</strong> the Bloch vecto r during the<br />

splitting and merging stag es is also importan t bec ause it will<br />

(b)<br />

0.4<br />

0.3<br />

0 10 20 30 40<br />

Th<br />

0<br />

(d)<br />

0.2<br />

0.1<br />

0<br />

10 3<br />

T s<br />

10 2<br />

10 1<br />

10 0<br />

(b)<br />

1<br />

FIG. 6. Depen den ce <strong>of</strong> the filling factorF on the duration <strong>of</strong> the<br />

splitting stag eTs for max= 12.5 and various values <strong>of</strong> asymmet ry<br />

Vˆ as=<br />

0.01xˆ a , 0.02xˆ b , 0.05xˆ c , and 0.1xˆ d .<br />

0.5<br />

population P1<br />

Hence the cont ribution from the first term in Eq.A2 is<br />

given by<br />

ˆ H<br />

a 0 1 = a. A4<br />

0 d<br />

dt<br />

For the contribution from the second term in Eq.A2 , we<br />

may first write<br />

are time independ ent<br />

a a<br />

KAB A B , where the KAB<br />

ˆ<br />

a=<br />

A,B=L,R<br />

coefficients tha t can be read from Eqs.13 , and then<br />

H<br />

.<br />

A B + A<br />

a<br />

KAB = A,B=L,R<br />

t B<br />

t<br />

ˆ<br />

a<br />

H<br />

t<br />

A5<br />

Using Eq. 12 for the stat e vector and reverting to the<br />

Schrödinger picture we have<br />

0 t a<br />

ˆ H a<br />

0 2 = t KAB t A B + A<br />

A,B=L,R<br />

S<br />

t B t<br />

affect the mea surable probabilityP 1. We have also shown<br />

tha t special requ irement s app ly to the durat ion <strong>of</strong> the splitting<br />

and merging stage s in orde r to avoid excitation <strong>of</strong> high er<br />

mode s for short times and partial adiabatic following if the<br />

splitting is too long. Both thes e effects lead to adecre ase <strong>of</strong><br />

the mea sured signal. Interest ingly enoug h the y do not affect<br />

the contras t <strong>of</strong> the interfe rence fringes if the first excited state<br />

is not initially populate d.<br />

Adiaba tic evolut ion <strong>of</strong> the Bloch vecto r can <strong>of</strong>fer anew<br />

way to mea sure the first excited stat e population after the<br />

doubl e-well interferome tric process. We have already men -<br />

tione d tha t in the far-split regime the excited stat e wave function<br />

doe s not overlap with the groun d stat e wave function<br />

and will predomina ntly occupy the high er ene rgy wellFig.<br />

2 c . If at the end <strong>of</strong> the nonadia batic splitting, phase evolution,<br />

and non adiabatic recom binati on process we also add<br />

an addit ional stag e <strong>of</strong> adiabatic splitting in aknown asymmetrical<br />

potenti al, the n the wavefunctions <strong>of</strong> the two state s<br />

will be spatially separated. For recording the out putP 1 we<br />

now simply mea sure the population <strong>of</strong> the high er ene rgy<br />

well. To shorte n the adiabatic evolution time we nee d to<br />

app ly the highest available asymmetryFig. 6 .<br />

signal and aredu ced maximum popu lation <strong>of</strong> the first excited<br />

state .<br />

For splitting and merging timesTs=Tr= 20 Fig. 5 b<br />

bot h mode ls show excellent agreemen t indicating asimple<br />

sinusoidal variation <strong>of</strong> the first excited stat e popu lation with<br />

holding time. This simple beh avior is also obs erved for long<br />

splitting time Fig. 5 c , but with significantly redu ced amplitud<br />

e <strong>of</strong> the oscillations. The redu ced fringing is att ributed<br />

to the onset <strong>of</strong> adiab atic following <strong>of</strong> the Bloch vector during<br />

splitting and recombinatio n which is shown by the presence<br />

<strong>of</strong> anon zero z comp one nt Fig. 4 a . We not ed tha t our<br />

num erical solution <strong>of</strong> the Bloch equa tion is robust with regard<br />

to the variations <strong>of</strong> the signal but is fragile rega rding the<br />

phas e. The error was accumulat ed during the splitting stag e<br />

as aresult <strong>of</strong>Vas 0 in amerged trap and will scale with the<br />

splitting time.<br />

In the asymmetric doub le-well pot ent ial the ground state<br />

eigenfunction will predom inantly occupy the lower well, and<br />

the excited stat e eigenfunction will be localized in the upp er<br />

well Fig. 2 c . In the slow splitting regime the onset <strong>of</strong> the<br />

adiab atic evolution will lead to the unbalan ced distribution<br />

<strong>of</strong> the ato mic wave function bet wee n the wells, which in turn<br />

leads to aredu ction in the mea sured signal. In app lication to<br />

interferomet ry it can be seen as intrinsic which-way information<br />

whe n the ato m predom inantly follows one pat h after the<br />

0 10 20 30 T 40<br />

h<br />

0<br />

1<br />

(c)<br />

0.5<br />

population P1<br />

189<br />

0 10 20 30 T 40<br />

h<br />

0<br />

FIG. 5. Dependenc e <strong>of</strong> the first excited stat e popu lationP 1 at the<br />

end <strong>of</strong> the interferometri c process on the duratio n <strong>of</strong> the holdin g<br />

stage Th for various dura tions <strong>of</strong> the splitting and recombining<br />

stage sTs=Tr= 5 a , 20 b , and 200 c . Results <strong>of</strong> the Bloch vector<br />

mod el are represente d by solid lines and out come s <strong>of</strong> full num erical<br />

simulation s are presented by circles.<br />

a *<br />

= KAB CDCB<br />

D t A<br />

A,B,D=L,R<br />

*<br />

+ CACD t B D . A6<br />

To evaluate this result, a consideration <strong>of</strong> the four qua ntities<br />

A t B , where A,B=L,R is requ ired, not ing also<br />

tha t t B A = A t B * . These four qua ntities can be<br />

expressed in term s <strong>of</strong> related mat rix elem ent s in the symmet -<br />

ric basis Si t Sj , where i, j=0,1 . Note S0 S and<br />

S1 AS .<br />

Using the normalization and reality property, we first<br />

show tha t the diagona l term sSi t Si are zero. For the<br />

<strong>of</strong>f-diagona l term s Si t Sj , the se are zero becau se Si<br />

and t Sj have opp osite symmet ry. From the se considerations<br />

we find tha t all mat rix elemen tsA t B , where<br />

splitting. In gen eral,<br />

ACKNOWLEDGMENTS<br />

= a 0 + b 1 + i , 22<br />

It is tempti ng to limit the evolution <strong>of</strong> the Bloch vecto r<br />

and the relevan t phas e accum ulation during the splitting<br />

and merging stage s by making thes e stage s shorte r. However,<br />

this can lead to excitation s <strong>of</strong> highe r excited stat es. We<br />

have compared out come s <strong>of</strong> the Bloch vecto r mode l with<br />

the results <strong>of</strong> the numeri cal solution <strong>of</strong> a multist ate<br />

Schrödinge r equatio n using theXMDS code 30 . The<br />

behavi or <strong>of</strong> the excited stat e popula tionP 1 T at the end<br />

We than k T.D. Kieu for fruitful discussions, and T.<br />

Vaug han and P. Drummon d for the introdu ction to the<br />

XMDS code . This work has bee n supp orte d by the ARC<br />

Centr e <strong>of</strong> Excellence for Quant um-Atom Optics.<br />

whe re i is alinear combination <strong>of</strong> all oth er excited state s.<br />

We define afilling factor<br />

F = 2ab, 23<br />

APPENDIX: DERIVATION OF BLOCH EQUATIONS<br />

The state vectors at timet and at time 0 are related via the<br />

unitary evolution operat orU t ˆ as t =U t ˆ 0 . Operator<br />

s in the Heisenbe rg and Schrödinge r pictures are related<br />

viaÛ as ˆ H † = Û ˆ S Û . The expectatio n values <strong>of</strong><br />

operator s in the two pictures are related as ˆ<br />

which will des cribe the balan ce <strong>of</strong> ground and excited state s<br />

popu lations. The dep end ence <strong>of</strong> the filling factor on splitting<br />

time is shown in Fig.6 for asplitting <strong>of</strong> = 12.5 and different<br />

asymmetries. The results <strong>of</strong> the Bloch mod eldot ted line<br />

and the multistate num erical simulationssolid line show<br />

goo d agreemen t for splitting timesT s 20. For shorter splitting<br />

stag es the two-mod e app roximation fails and excitations<br />

into higher mod es take place. In the case <strong>of</strong> the high asym-<br />

= t ˆ S t = 0 ˆ H 0 .<br />

<strong>of</strong> the interfe romet er proces s as afunction <strong>of</strong> the holding<br />

period Th is shown in Fig. 5 for the parameters Vas ˆ = 0.02xˆ,<br />

max= 12.5, Ts=Tr= 5 Fig. 5 a , Ts=Tr= 20 Fig. 5 b ,<br />

and Ts=Tr= 200 Fig. 5 c . In all cases the sinusoidal beh avior<br />

<strong>of</strong> the excited stat e pop ulation as a function <strong>of</strong> the<br />

holding period can be seen. Situations rangin g from<br />

compl ete destructi ve interfe rence to perfect constru ctive<br />

interference are bot h present . For short splitting timesFig.<br />

5 a we observe d a discrepanc y bet wee n the two mod els.<br />

Multistate nume rical simulations indicate the presenc e <strong>of</strong><br />

popul ate d high er ene rgy state s which the Bloch vecto r mod el<br />

ignores. The full num erical calculations show an irregular<br />

high freque ncy modulat ion <strong>of</strong> the fundame nta l frequenc y met ry Vˆ as=<br />

0.1xˆ Fig. 6 d we obs erve asignificant devia-<br />

023612-7<br />

023612-8


A.1. Asymmetric double-well potential for single-atom interferometry<br />

ASYMMETRIC DOUBLE-WELL POTENTIAL FOR ? PHYSICAL REVIEW A 74, 023612 2006<br />

Combining bot h contributions we find tha t<br />

d<br />

a = a. A8<br />

dt<br />

Thus the Bloch equ ations can be expressed in vector form as<br />

in Eq. 18 and in det ail as in Eqs. 17 .<br />

A,B=L,R , are zero. Hence the secon d contributi on to the<br />

equatio n <strong>of</strong> motio n for the Bloch vector componen t is zero<br />

ˆ H<br />

a 0 2 = 0. A7<br />

0 d<br />

dt<br />

17 D. Müller, E. A. Cornell, M. Prevedel li, P. D. D. Schwindt, A.<br />

Zozulya, and D. Z. Anderson, Opt. Lett. 25, 1382 2000 .<br />

18 J. Estève, T. Schum m, J.-B. Trebb ia, I. Bouchoule, A. Aspect,<br />

and C. I. Westbrook, Eur. Phys. J. D 35, 141 2005 .<br />

19 W. Hänsel, J. Reichel, P. Hommelh<strong>of</strong>f, and T. W. Hänsch,<br />

Phys. Rev. Lett. 86, 608 2001 .<br />

20 Y. Shin, M. Saba , T. A. Pasquini, W. Ketterl e, D. E. Pritchard,<br />

and A. E. Leanha rdt, Phys. Rev. Lett. 92, 050405 2004 .<br />

21 T. Schum m, S. H<strong>of</strong>ferber th, L. M. Andersson, S. Wildermuth ,<br />

S. Groth, I. Bar-Joseph, J. Schmiedmayer, and P. Krüger, Nat.<br />

Phys. 1, 57 2005 .<br />

22 Y. Shin, C. Sanne r, G.-B. Jo, T. A. Pasquin i, M. Saba , W.<br />

Kett erle, D. E. Pritchard, M. Veng alattore, and M. Prent iss,<br />

Phys. Rev. A 72, 021604 R 2005 .<br />

23 M. A. Kasevich, Science 298, 1363 2002 .<br />

24 J. Javanainen and M. Wilkens, Phys. Rev. Lett. 78, 4675<br />

1997 .<br />

25 T. Calarco, E. A. Hinds, D. Jaksch, J. Schmiedmayer, J. I.<br />

Cirac, and P. Zoller, Phys. Rev. A 61, 022304 2000 .<br />

26 D. Pulido, Master thesis , Worcester Politechnic Institute, 2003<br />

unp ublished.<br />

27 M. V. Berry, Proc. R. Soc. London , Ser. A 392, 45 1984 .<br />

28 L. Allen and J. H. Eberly, Optical Resonanc e and Two-Level<br />

Atom s Wiley, New York, 1975 .<br />

29 C. Cohen-Tann oud ji, B. Diu, and F. Laloë, Quantu m Mechanics,<br />

Volume I Wiley, New York, 1977 .<br />

30 XMDS code is acode gen erat or tha t integra tes equatio ns. It is<br />

develope d at the <strong>University</strong> <strong>of</strong> Quee nsland, Brisbane , Australia,<br />

www.xmds.org.<br />

1 J. Javana inen, Phys. Rev. Lett. 57, 3164 1986 .<br />

2 M. W. Jack, M. J. Collett, and D. F. Walls, Phys. Rev. A 54,<br />

R4625 1996 .<br />

3 C. Menot ti, J. R. Anglin, J. I. Cirac, and P. Zoller, Phys. Rev. A<br />

63, 023601 2001 .<br />

4 Y. Castin and J. Dalibard, Phys. Rev. A 55, 4330 1997 .<br />

5 J. Javana inen and S. M. Yoo, Phys. Rev. Lett. 76, 161 1996 .<br />

6 R. Folman, P. Krüger, J. Schmiedmayer , J. Denschl ag, and C.<br />

Henkel, Adv. At., Mol., Opt. Phys. 48, 263 2002 .<br />

7 W. Hänsel, P. Hommelho ff, T. W. Hänsch, and J. Reichel, Nature<br />

London 413, 498 2001 .<br />

8 H. Ott, J. Fortagh , G. Schlotterbec k, A. Grossmann , and C.<br />

Zimmermann , Phys. Rev. Lett. 87, 230401 2001 .<br />

9 R. Dumke, M. Volk, T. Müther , F. B. J. Buchkremer , G. Birkl,<br />

and W. Ertmer, Phys. Rev. Lett. 89, 097903 2002 .<br />

10 E. A. Hinds, C. J. Vale, and M. G. Boshier, Phys. Rev. Lett.<br />

86, 1462 2001 .<br />

11 W. Hänsel, J. Reichel, P. Hommelh<strong>of</strong>f, and T. W. Hänsch,<br />

Phys. Rev. A 64, 063607 2001 .<br />

12 E. Andersson, T. Calarco, R. Folman, M. Andersson , B.<br />

Hessmo, and J. Schmiedmayer, Phys. Rev. Lett. 88, 100401<br />

2002 .<br />

13 J. A. Stickney and A. A. Zozulya, Phys. Rev. A 66, 053601<br />

2002 .<br />

14 A. Negrett i and C. Henkel, J. Phys. B 37, 385 2004 .<br />

15 H. Kreut zmann , U. V. Poulsen, M. Lewenstein , R. Dumke, W.<br />

Ertmer, G. Birkl, and A. Sanp era, Phys. Rev. Lett. 92, 163201<br />

2004 .<br />

16 D. Cassett ari, B. Hessmo, R. Folman, T. Maier, and J.<br />

Schmiedmayer , Phys. Rev. Lett. 85, 5483 2000 .<br />

190<br />

023612-9


Chapter A: Reprints <strong>of</strong> selected Publications by the Candidate<br />

A.2 A permanent magnetic film atom chip for<br />

Bose-Einstein condensation<br />

INSTITUTE OF PHYSICS PUBLISHING JOURNAL OF PHYSICS B: A TOMIC, MOLECULAR AND OPTICAL PHYSICS<br />

28 BVHall et al<br />

J. Phys. B: At. Mol. Opt. Phys. 39 (2006) 27–36 doi:10.1088/0953- 4075/39/ 1/004<br />

z 0B s<br />

a<br />

i<br />

b<br />

A permanent magnetic film atom chip for<br />

Bose–Einstein condensation<br />

z<br />

y<br />

x<br />

BVHall, S Whitlock, FScharnberg 1 ,PHannaford and A Sidorov<br />

ARC Centre <strong>of</strong> Excellence for Quantum-Atom Optics and Centre for Atom Optics and Ultrafast<br />

Spectroscopy, <strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong>, Hawtho rn, Victoria 3122, Australia<br />

Figure 1. A simple mode l des cribes the magn etic field <strong>of</strong> asemi-infinite, perpendic ularly<br />

magneti zed thin film in comb ination with a uniform bias mag netic field.<br />

E-mail: brhall@groupwise.swin.edu .au<br />

material s <strong>of</strong>fer the possibility <strong>of</strong> ultra-stable mag netic pote ntials due to their intrinsically low<br />

magneti c field noise. Moreover, permanen t magn etic films are thin and relatively high in<br />

resistanc e comp ared to current -carrying wires, properties which strongly supp ress the rmal<br />

magneti c field noise [12]. Permanen t mag net ic mate rials with in-plane mag netization have<br />

recentl y bee n used to demo nstra te trapp ing <strong>of</strong> cold atom13] s [ and to prod uce BEC on a<br />

magneti c videot ape 14, [ 15]. Here we empl oy a novel mag netic mat erial with perpen dicular<br />

anisotro py, develope d spec ifically for appl ications with ultraco ld atoms . Perpen dicularly<br />

magnetize d mat erials allow arbitrary 2D patter ns to be writte n in the plane <strong>of</strong> the film and<br />

provide magnet ic field configurati ons analogou s to thos e produced by planar micr<strong>of</strong>abricated<br />

wires [16, 17].<br />

In this pap er, we repor t the realization <strong>of</strong>apermanen t mag netic film /machined conductor<br />

87 ato m chip which has bee n used to produ ce aRb<br />

BEC. In section 2, we present a simple<br />

mode l for a thin film <strong>of</strong> perpen dicularly magnetize d mat erial which results in straigh tforward<br />

equat ions for the mag netic field nea r the edg e <strong>of</strong> the film. We the n desc ribe the principle<br />

<strong>of</strong> trapp ing ultracold atom s in the potentia l formed by the permanen t mag netic film (the<br />

film trap ). TbGdFeCo material s are the n introd uced in section3 with adesc ription <strong>of</strong> the<br />

deposit ion process. We mea sured the bulk prop erties <strong>of</strong> the mag netic film using bot h a<br />

superco nductin g quan tum interference device (SQUID) and a mag netic force microscope<br />

(MFM). Section 4 describe s the constructio n <strong>of</strong> the ultra-hig h vacuu m (UHV) compa tible<br />

ato m chip. This include sacurrent-carr ying struct ure to provide time-dep enden t control <strong>of</strong><br />

surface-ba sed pot entials and is used to form a conduc tor-base d mag netic trap (the wire trap).<br />

The app aratus and experime nta l procedure s used for making a BEC indep enden tly with<br />

the film trap or the wire trap are describ ed in section5. In section 6, we appl y the BEC<br />

as a novel ultraco ld ato m mag net omet er by meas uring the spatia l decay <strong>of</strong> the mag netic<br />

field from the film. High precision trap frequen cy meas uremen ts in conjunct ion with radio<br />

frequen cy outp ut coupl ing also allow the direct det erminat ion <strong>of</strong> the associated mag netic field<br />

gradient . In conclusion, we specul ate on future directions for our permanen t mag netic film<br />

ato m chip.<br />

Received 19 July 2005, in final form 25 July 2005<br />

Published 5 Decem be r 2005<br />

Online at stacks.iop.org/JPhysB/39/ 27<br />

Abstract<br />

We present a hybrid atom chip which combine s a permanen t mag netic film with<br />

a micromachined current-carrying structure used to realize a Bose–Einstein<br />

conden sate (BEC). A novel TbGdFeCo material with large perpendic ular<br />

magne tization has bee n tailored to allow small scale, stable mag netic pot entials<br />

for ultracold atoms. We are able to produce 87Rb BECinamag netic trap based<br />

on either the permanen t mag netic film or the current-carrying structure. Using<br />

the conde nsate as a mag netic field probe we perform cold atom mag netometry<br />

to pr<strong>of</strong>ile both the field magni tude and gradient as a function <strong>of</strong>distance from the<br />

magne tic film surface. Finally, we discuss future directions for our permanen t<br />

magne tic film atom chip.<br />

191<br />

2. The simple model <strong>of</strong> a permanent magnetic film<br />

1. Introduction<br />

A recent technological advance in the area <strong>of</strong> quan tum degen erate gases has bee n the<br />

developmen t <strong>of</strong> the ‘atom chip’. These devices exploit tightly confining , mag netic pot entials,<br />

created by low power current-carrying wires to simplify the production <strong>of</strong> Bose–Einstein<br />

conden sates (BEC) [1, 2]. In addition, they provide the freedom to realize intricate mag netic<br />

pot entials with features <strong>of</strong> the size comparable to the atomic de Broglie waveleng th. Atom<br />

chips have bee n used to realize atomic waveguides and transpo rt devices for BEC[ 3, 4]. These<br />

tools allow controllable manipulation <strong>of</strong> ultracold neu tral atoms, with pot ential app lications<br />

in quan tum information processing 5, [ 6] and atom interferometry 7, [ 8].<br />

For current-carrying wire-based atom chips, technical limitations are imposed by current<br />

noise and spatial fluctuations in the current den sity leadin g to increased heating rates and<br />

fragmen tation <strong>of</strong> cold clouds 9, [ 10]. In addition, nea r-field therma l noise in condu ctors<br />

is respo nsible for afundamen tal atom loss mechanism 11]. [ Atom chips incorporating<br />

perman ent magne tic materials are expected to overcome many <strong>of</strong> these difficulties. These<br />

Consider a semi-infinite rectangula r magne t with the mag netization M and thickness h<br />

(figure 1). The magne t lies in thexy plane with one edg e aligned along they axis. The<br />

magne t is uniformly magnet ized in the z+ direction and the distance from the film is large<br />

with respec t to the film thickness(z h). The mag netic field and the field gradient directly<br />

1 Present addr ess: Institut für Quantenopt ik, Universität Hann over, 30167 Hann over, Germany.<br />

0953-4075/06/010027 +10$30.00 © 2006 IOP Publishing Ltd Printed in the UK 27


A.2. A permanent magnetic film atom chip for Bose-Einstein condensation<br />

30 BVHall et al<br />

A perm ane nt magnet ic film ato m chip for Bose–Einstein condens ation 29<br />

above the edg e can be written as<br />

0.3<br />

0.2<br />

hM<br />

. (1)<br />

z2 and Bfilm =− µ0<br />

2?<br />

hM<br />

z<br />

Bfilm = µ0<br />

2?<br />

( T)<br />

0.1<br />

M<br />

0<br />

−0.1<br />

magnetization, μ<br />

0<br />

−0.2<br />

−0.3<br />

−1.2 −0.8 −0.4 0 0.4 0.8 1.2<br />

applied field, μ H (T)<br />

0<br />

Figure 2. Ahysteresis loop derived from SQUID mag neto met ry <strong>of</strong> a multilayer Tb 6Gd10Fe80Co4<br />

magneti c film. The film mag netizat ion isµ0M S ? µ0M R = 0.28 T and the coercivity is<br />

µ0H C = 0.32 T.<br />

These expression s are analogo us to thos e derived using Biot–Savart’s law for the mag netic<br />

field above an infinitely long and thin current -carrying wire. The similarity betw een a<br />

permanen t mag netic film and a current -carrying wire can be explained using asimple mod el.<br />

An unmag net ized film is comprise d <strong>of</strong> many small mag netic doma ins <strong>of</strong> rando m orientat ion.<br />

The mag netic field prod uced by each dom ain is equivalent to tha t from an imaginary surface<br />

current flowing along the doma in border s, perp endi cular to the mag netization vector 18]. [<br />

For a uniformly mag net ized film with perpend icular anisot ropy all dom ains are aligned in the<br />

same direction (out <strong>of</strong> plane) and within the bulk the mag netic fields <strong>of</strong> neighb ouring dom ains<br />

cancel. A net effective current exists abo ut the perimet er <strong>of</strong> the film with amag nitude given<br />

by the prod uct <strong>of</strong> the magneti zation and the film thickness (Ieff = hM).<br />

The appl ication <strong>of</strong> auniform bias field(B bias)in the −x direction produces a radially<br />

symmetric two-dimensional quadrupole mag netic field above the film edg e at the heigh z0, t<br />

where the magnit udes <strong>of</strong>Bbias andBfilm are equal. To realizeathree-dimensional(3D)mag netic<br />

trap for weak-field seeking atom s a non -uniform axial fieldBy is provided by two parallel<br />

current s located beneat h and perpendic ular to the waveguide. Additionally, By supp resses<br />

spin-flip losses by prevent ing the tota l mag net ic field at the trap botto m from going to zero.<br />

This results in a3D harmoni c film trap at adistanc ez0 from the surface with the radial<br />

frequency given by<br />

goo d magneti c prop erties and increase the mag netic field strengt h nea r the surface we have<br />

implem ente d a multilayer depo sition which produc es high quality TbGdFeCo mag netic films<br />

with atot al thickness app roach ing 1µm. A glass slide subs trate was cleaned in an ultraso nic<br />

bat h using a nitric acid solution the n carefully rinsed before being mounte d in the dep osition<br />

chambe r. The base pressure was less tha n 5× 10− 8 Torr prior to introducing the argon buffer<br />

gas (? 4 mTorr). The sub strate was the n hea ted to 100 ? C and a bon ding layer <strong>of</strong> chromium<br />

(120 nm) was spu ttere d on the surface. This was followed by the depo sition <strong>of</strong> six bi-layers<br />

<strong>of</strong> TbGdFeCo (150 nm) and Cr (120 nm) films.<br />

The mag netic proper ties <strong>of</strong> the multilayer TbGdFeCo/Cr film were characterized by a<br />

SQUID magnetomete r (figure2). The hysteres is loop indicate s a remanen t mag netization<br />

<strong>of</strong> 0.28 T for a tot al magne t thickness <strong>of</strong> 900 nm hM ( = 0.20± 0.01 A). The comp lete<br />

magnetizat ion <strong>of</strong> the film can be achieved by applying a field ? <strong>of</strong>0.8<br />

T, while the film<br />

magnetizat ion is robust in the presence <strong>of</strong> externa l fields below ? 0.1 T. The surface features<br />

<strong>of</strong> the films have also bee n examined byahigh-resolution atom ic force microscope ope rating<br />

in the magnet ic force mod e (figure3). An unmag netized sampl e shows micron-sized features<br />

consisten t with domai n stripes , while an example <strong>of</strong> a uniformly mag net ized sample exhibits<br />

excellent magnet ic homo geneit y.<br />

192<br />

, (2)<br />

µBgFm F<br />

mBy<br />

hM<br />

z0 2<br />

2?f radial = µ0<br />

2?<br />

whereµB is the Bohr mag neton ,gF is the Landé factor,m F is the mag netic quant um number<br />

and m is the atomi c mass. The ability to prod uce high quality, thick mag netic films with large<br />

magneti zation is necessary to prod uce tightly confining mag net ic traps for ultracold ato ms.<br />

3. Tb6Gd10Fe 80Co4 magneto-o ptical films and their properties<br />

4. The atom chip design<br />

This device represen ts the first ato m chip base d on a perpen dicularly mag net ized permanen t<br />

magneti c film for trap ping ultracold ato ms. It has bee n designe d for the production and<br />

manipulation <strong>of</strong> a BEC nea r the surface <strong>of</strong> the mag netic mat erial. Althoug h thes e films are<br />

well suited for making tight and stable trapp ing pote ntials up to a few 100 µm from the<br />

surface, the small volume <strong>of</strong> the film trap is not suitab le for efficient loading directly from a<br />

magneto -optica l trap (MOT). To circumvent this difficulty a current -carrying wire structure<br />

locate d beneat h the magneti c film provides an add itiona l trapp ing field. The comb ination <strong>of</strong><br />

bot h the magneti c film and the wire structur e represen ts the hybrid ato m chip des ign shown<br />

schematicall y in figure4.<br />

The des ire for large capa city information storage devices has enco urage d an extens ive<br />

investmen t towards develop ing novel mag netic comp ositions. These are primarily opti mized<br />

to achieve small scale, recordabl e pat ternin g <strong>of</strong> mag netic media . While it is poss ible to<br />

benefi t from this experience , app lications with cold ato ms have several add itional yet very<br />

specific requiremen ts. Firstly, a high Curie tem perature(TC)will prevent dem agnet ization<br />

during the bake-out procedure, a necessary step in achieving UHV conditions. Secondly,<br />

a high coercivity(H C)will prevent the loss <strong>of</strong> magne tization whe n app lying large external<br />

magneti c fields. Finally, the remanen t magneti zation(M R)and the saturation mag netization<br />

(M S) should be large and nearly equivalent, an indication <strong>of</strong> goo d mag netic homo gen eity.<br />

These conditi ons are satisfied by Tb6Gd10Fe80Co4 mag neto -opti cal films which have a high<br />

Curie temperat ure (TC ? 300 ? C), perpendic ular anisotropy and a square hysteresis loop.<br />

TbGdFeCo films were prod uced using athin film depo sition system (Kurt J Lesker<br />

CMS-18) equipped with magnetro n sputterin g and elect ron bea m evapo ration sources 19]. [<br />

A compo site target with a nominal atomi c comp osition <strong>of</strong> 6Gd10Fe80Co4 Tb and ahigh purity<br />

chromium target are used in the prod uction <strong>of</strong> the mag netic films. A systema tic study <strong>of</strong> the<br />

influence <strong>of</strong> process parame ters over the prop erties <strong>of</strong> the film indicate d tha t det eriorat ion <strong>of</strong><br />

the mag netic anisotro py occurs for the film thickness <strong>of</strong> above 250 nm. In order to mainta in


32 BVHall et al<br />

A perm ane nt magnet ic film ato m chip for Bose–Einstein condens ation 31<br />

Chapter A: Reprints <strong>of</strong> selected Publications by the Candidate<br />

The second layer <strong>of</strong>the hybrid chip is a wire structure which was produced using the micromachine<br />

d silver foiltechniqu e develope d by Valet al[20]. A500 µm thick silver foil(99.99%<br />

purity) was fixed with epo xy (Epot ek H77) to a 2 mm thick Shapal -M machineab le ceramic<br />

base-plate . A comp ute r cont rolled Quick Circuit 5000 PCB mill was used to cut 500µm<br />

wide insulating grooves in the foil. Each wire has a width <strong>of</strong> 1 mm which is broadene d to<br />

6 mm far from the trapp ing region to facilitate goo d elect rical connections . After cutting, the<br />

insulatin g channe ls were filled with add itional epo xy to increase the structural integrity and<br />

therma l cond uctivity. The wire structur e including elect rical connection s hasatot al resistan ce<br />

<strong>of</strong> 4.6 m . A continuou s current <strong>of</strong> 30 A can be app lied with an associated tem peratur e rise<br />

<strong>of</strong> less tha n 40 ? C and negligible increase in vacuum pressure.<br />

In convent ionalato m chips U- or Z-shap e wires are used for creating qua drupole and I<strong>of</strong>fe–<br />

Pritchard(IP)magneti c field geom etries to realize mirror MOTs and mag netic microtraps 21]. [<br />

In the prese nt ato m chip, high current s are used to formatight trap relatively far from the wire,<br />

thereb y avoiding unwante d collisions with the surface <strong>of</strong> the slide. Conse que ntly, the use <strong>of</strong><br />

broa d conduc tors proh ibit the use <strong>of</strong> sep arate U- and Z-shape wires. This is circumvent ed<br />

with a plana r H-shap e structure , designe d to allow bot h U- and Z-shap e current path s with<br />

goo d spatia l overlap <strong>of</strong> the associated traps . Axial confinemen t for the film trap is provided<br />

by additiona l parallel cond uctor s separate d by 9.5 mm and located eithe r side <strong>of</strong> the H-shape<br />

structure . The top surface <strong>of</strong> the machine d silver foil was later polished flat to supp ort the<br />

glass slides.<br />

During assembl y, the polished edg e <strong>of</strong> the TbGdFeCo film coat ed slide is aligned to the<br />

midd le <strong>of</strong> the H-shape structure and set with epo xy. The second gold-coa ted slide is epo xied<br />

adjacen t to the magnet ic film slide to comp lete the reflective chip surface. Two rubidium<br />

dispe nsers are mounte d on two ceramic blocks (Macor) which are recesse d below the chip<br />

surface. The two glass slides, mach ined silver foil and ceramic base -plate are the n fixed<br />

to acopp er hea t sink. The comp leted chip is clamped to a19 mm diamete r solid copper<br />

feed throug h (Cerama seal, 800 A rating) and mou nte d in the vacuu m cham be r. Electrical<br />

conn ection s are mad e using a1.6 mm diamete r bare coppe r wire and BeCu barrel connectors<br />

in conjunction with a 12 pin power feed throug h (Ceramaseal , 55 Arating ). A cold cathod e<br />

gaug e indicate d a pressur e below× 110−<br />

11 Torr after baking at 140 ? Cfor 4 days, highlight ing<br />

the UHVcomp atibility <strong>of</strong> all mat erials.<br />

−5<br />

(a) (b)<br />

−4<br />

−3<br />

−2<br />

−1<br />

0<br />

1<br />

position ( μ m)<br />

2<br />

3<br />

4<br />

5<br />

Figure 3. The mag netic force microscope (MFM) image <strong>of</strong> a Tb6Gd10Fe80Co4 mag net ic film<br />

surface. (a) The unm agne tized sampl e shows the dom ain structu re with micron-sized features.<br />

(b) The uniformly mag netize d sampl e is free <strong>of</strong> any visible mag net ic structur e.<br />

193<br />

5. Bose–Einstein condensation on a permanent magnetic film<br />

The reflective surface <strong>of</strong> the ato m chip is used to form amirror MOT and accomm oda tes<br />

30 mm diamet er laser beam s provided byahigh-p ower diod e laser (Topti ca DLX110) locked<br />

to the D2 (F = 2 ? 3)cooling transition <strong>of</strong>87Rb. The trappin g light is det une d 18 MHz<br />

below resona nce and has an inten sity <strong>of</strong> 4 mW cm − 2 in each bea m. Arepum ping laser locked<br />

to the D2 (F = 1 ? 2)trans ition is comb ined with the trapp ing light with an intens ity <strong>of</strong><br />

0.5 mW cm− 2 pe r beam . Two wate r-cooled coils mount ed outs ide the vacuu m cham ber<br />

provide a quadrup ole magneti c field with gradient 0.1 Tm − 1 centered 4.6 mm below the chip<br />

surface. To load the mirror MOT a current <strong>of</strong> 6.5Ais pulsed for 9.5 s throug h one resistively<br />

heate d Rb dispe nser, allowing the collection <strong>of</strong> 2× 108 atoms . The atom s are held for a further<br />

15 s while the UHVpressur e recovers, read y for trans fer to the chip-base d pot entia ls.<br />

Transfer begins by simultaneousl y ramping a current thro ugh the U-shap e circuit<br />

(IU = 0 ? 8 A), increasing the uniform fieldBbias and turning <strong>of</strong>f the external qua drupole<br />

magneti c field over 50 ms. This moves the atom s withou t loss, into a U-wire MOT locate d at<br />

1.6 mm from the surface and increase s the radial gradient to 0.4 Tm − 1 . While this compression<br />

increases the spat ial overlap with theIP potential , it also hea ts the cloud. To counte ract this,<br />

Figure 4. Schematic view <strong>of</strong> the hybrid ato m chip. Inset: TbGdFeCo/Cr multilayer film and Au<br />

overlayer. From the top down, glass slide coat ed with magn etic film, machined silver foil H-wire<br />

and end wires, Shapa l-M base -plate and Cu hea t sink. Missing from the schema tic is the second<br />

glass slide and two rubidium dispe nsers.<br />

The top layer <strong>of</strong> the ato m chip consists <strong>of</strong> two adjacen t 300 µm thick glass slides which<br />

are sturdy eno ugh to prevent warping . The long edg es <strong>of</strong> the glass slides were polished with<br />

aluminium oxide grit prior to depo sition to remove visible chips. Amultilayer TbGdFeCo /Cr<br />

film was depo sited on one slide using the proced ure out lined in section 3. Both slides were<br />

the n coat ed withagold overlayer (170 nm) and toge the r formalarge reflective surface × (40<br />

46 mm2 ). This allows the collection <strong>of</strong> a large num be r <strong>of</strong> atom s into a mirror MOT within a<br />

single-cham be r UHVsystem . The glass slide coat ed with the TbGdFeCo/Cr multilayer film<br />

was the n mag netized inauniform field <strong>of</strong> 1 T pend ing assembl y.


A.2. A permanent magnetic film atom chip for Bose-Einstein condensation<br />

34 BVHall et al<br />

A perm ane nt magnet ic film ato m chip for Bose–Einstein condens ation 33<br />

1.5<br />

ty<br />

nsi<br />

1<br />

e<br />

d<br />

0.5<br />

optical<br />

0<br />

−0.6 −0.3 0 0.3 0.6<br />

−0.6 − 0.3 0 0.3 0.6<br />

position (mm)<br />

−0.6 −0.3 0 0.3 0.6<br />

Figure 5. Typical absor ptio n images and opti cal den sity pr<strong>of</strong>iles <strong>of</strong> a ballistically expand ed atom<br />

cloud. Each image is asingle realization <strong>of</strong> the experiment where evapora tion is performe d in<br />

the perma nen t magnet ic film pot ential. After truncating the evaporat ion ramp , atom s are held for<br />

150 ms and ballistically expande d for 30 ms befo re imaging. (a) RF final = 804 kHz—the rmal<br />

cloud, (b) RFfinal = 788 kHz—part ially condens ed cloud, and (c) RFfinal = 760 kHz—an almost<br />

pure condens ate.<br />

cold atom s have bee n empl oyed to chara cterize the mag netic field produced by the film inside<br />

the vacuum cham be r. This allowsadirect comp arison with the simple mod el desc ribed earlier.<br />

A magnet ically trapp ed cloud <strong>of</strong> cold ato ms or a BEC beh aves as an ultra-sensitive probe to<br />

the local mag net ic field. A measur e <strong>of</strong> the trap position as a functionBbias <strong>of</strong> det ermines<br />

Bfilm(z), while an indepen den t measur e <strong>of</strong> the trap frequen cy is used to det ermine Bfilm (z).<br />

Once the BEC is confin ed by the film trap it is poss ible to pr<strong>of</strong>ile the mag netic field<br />

depen den ce nea r the surface. The pot ent ial minimum is located at the poin t whe re the uniform<br />

magneti c field is equal in mag nitud e to and cancels the field from the film (B bias = − Bfilm).<br />

The uniform magneti c field can be increased (decreased ) to move the trap minimum closer to<br />

(furthe r from) the film surface. The BECfollows the pot ent ial minimum and the measur ement<br />

<strong>of</strong> the cloud posit ion with respec t to the film surface det ermines Bfilm(z). The stren gth <strong>of</strong>Bbias<br />

is calibrated within the vacuu m cham be r using untr app ed atom s (far from the film) and a short<br />

RF pulse resona nt with the Zeema n splitting. The pixel size in the imaging plane is calibrated<br />

against the gravitationa l acceleration <strong>of</strong> freely falling atom s and agree s with the calibration<br />

given by imaging areference rule external to the app aratus. Unfortunatel y thoug h, the glass<br />

substrat e coate d with mag net ic mat erial has recessed appr oximat elyµm 50 beh ind the second<br />

blank glass slide as aconse quenc e <strong>of</strong> unevenly cured epo xy. The exact position <strong>of</strong> the film<br />

surface (in relation to the image) is theref ore unknown and present s an uncert ainty Bfilm(z). in<br />

For this reaso nasecon d techni que has bee n app lied to provide more information abo ut the<br />

magneti c field from the film.<br />

Harmonic oscillation s with the small amplit ude and frequenc ies up to 10 kHz can be<br />

measure d accura tely over many period s with a BECdue to low dam ping rate s and small spatial<br />

extent . In this case, trap frequen cies are measure d by exciting radial centre <strong>of</strong> the mass motion<br />

within the film trap and have bee n measure d to bet ter tha n1Hz ? 0.1% ( accuracy). These<br />

excitation s were obser ved by rapidly increasing the uniform mag netic field by app roximately<br />

5% befo re returnin g to the original position within 2 to5ms. The cloud position was measur ed<br />

after 10 ms <strong>of</strong> free expansion and dat a have bee n taken over five periods <strong>of</strong> oscillation.<br />

the radial gradient is reduc ed rapidly to 0.11 T m − 1 with the trap light <strong>of</strong>f to minimize any<br />

force on the atoms . Polarization gradien t cooling is applie d for 2 ms with 56 MHz red-detu ned<br />

trap light to reduc e the tem peratur e from 140 µKto 40 µK. Both the MOT light and IU are<br />

the n turned <strong>of</strong>f leaving the cold ato ms inauniform mag netic field.<br />

Next a 200µs opt ical pum ping pulse is appl ied to maximize the num ber <strong>of</strong> atom s in the<br />

|F = 2,m F = +2 weak-field seeking stat e read y for mag netic trapp ing. A current (IZ)<strong>of</strong><br />

21.5 Ais switched on throug h the Z-shap e circuit whileBbias is increase d to 1.3 mTto form an<br />

IP wire trap at the same posit ion. Atot al <strong>of</strong> 4× 107 atom s are held withabackground -limited<br />

lifetime greate r tha n 60 s. Adiabatic compres sion <strong>of</strong> this trap is performed by ramping IZ up<br />

to 31 A andBbias up to 4.0 mT over 100 ms. Furth er comp ression results in loss <strong>of</strong> ato ms to<br />

the surface. The compres sed mag netic trap is 560µm from the film surface where the radial<br />

and axial trap frequencies are 2?× 530 Hz and 2?× 18 Hz, resp ectively. The elastic collision<br />

rate in this trap (?el ? 50 s− 1 ) is high enoug h to begin evapo rative cooling.<br />

Forced evapo rative cooling to the BEC trans ition begins in the wire trap and is the n<br />

tran sferred to the film trap during a single logarithmic radio frequen cy (RF) ramp . The first<br />

8.85 s <strong>of</strong> this ramp is apreliminary cooling stag e in the wire trap down to atem peratur e <strong>of</strong><br />

? 5 µK. As the cloud is cooled the trap is compress ed furthe r to improve the evapo ration<br />

efficiency by loweringIZ to 25 A, moving the trap to 350µm from the surface and increasing<br />

the radial trap frequen cy to? 2? × 660 Hz. The RF amp litude is the n reduc ed to zero for<br />

150 ms while the ato ms are trans ferred closer to the chip surface and finally to the film trap.<br />

In this trapIZ is zero and axial confin eme nt on the mag netic film edg e is provided by the two<br />

end wires, each with a current <strong>of</strong> 6 A. The trap botto m is tune d using an add itiona l mag netic<br />

field para llel to the film edg e to minimize any discon tinuity in the RF evapo ration trajectory.<br />

The radial and axial trap frequencie s are 2? × 700 Hz and 2? × 8 Hz, resp ectively. The<br />

RF amplit ude is the n increase d again and evapo ration continues for 1sto the BEC pha se<br />

trans ition.<br />

Before imaging, the mag netic film trap is adiabatically moved 0.17 mm from the surface<br />

to avoid excessive field grad ients from the film. The cloud is the n release d by switching<br />

<strong>of</strong>f Bbias and the ato ms fall unde r gravity with minor acceleration from the perman ent field<br />

gradient . Resonan t opt ical absorptio n is used to image the atom s with a100 µs ? + light<br />

pulse parallel to the gold surface and tun ed to the2 D(F<br />

= 2 ? 3) transition. A CCD<br />

camer a records the absorptio n image <strong>of</strong> the cloud using an achroma tic dou blet telescope with<br />

a resolution <strong>of</strong> 5µm/pixel. Using the above proced ure a new conden sate <strong>of</strong>× 1105<br />

atoms<br />

is creat ed every 50 s. Figure 5 shows absorption image s and opt ical dens ity pr<strong>of</strong>iles after<br />

30 ms <strong>of</strong> ballistic expans ion. The forced RF evapo ration is truncat ed at 804 kHz, 788 kHz<br />

and 760 kHz revealing a therma l cloud, part ially conden sed cloud and nearly pure conden sate ,<br />

respe ctively.<br />

It is also possible to form a cond ensat e trapp ed solely by the wire trap. Here asingle,<br />

uninterrupt ed, 10 s RF ramp results in a BECwith ato m num ber compa rable to tha t realized in<br />

the film trap . This provides a unique possibility for stud ying the prop erties <strong>of</strong> aBEC in both<br />

permanen t mag net ic and current -carrying trap ping environment s. In add ition, the formation<br />

<strong>of</strong> a BEC inde penden t <strong>of</strong> the top layer will allow new mag netic structures or mat erials to be<br />

replaced with ease . The wire trap can also be used to transp ort a BEC to regions on the chip<br />

where the mag net ic field topolog y may be different from thos e nea r the substrat e edg e.<br />

194<br />

6. Magnetic field characterization<br />

The mag netic prop erties <strong>of</strong> the TbGdFeCo film were mea sured prior to mountin g on the ato m<br />

chip using a combination <strong>of</strong> SQUID and magneti c force microscopy. In situ techniques using


A perm ane nt magnet ic film ato m chip for Bose–Einstein condens ation 35<br />

0.8<br />

(a)<br />

−3<br />

T )<br />

Chapter A: Reprints <strong>of</strong> selected Publications by the Candidate<br />

0.6<br />

0<br />

( 1<br />

0.4<br />

th<br />

g<br />

n<br />

0.2<br />

stre<br />

36 BVHall et al<br />

ld<br />

fie<br />

0<br />

Acknowledgments<br />

(b)<br />

12<br />

We would like to than k J Wang and D Gough for carrying out the mag net ic film dep osition.<br />

This projec t is supp orte d by the ARC Centre <strong>of</strong> Excellence for Quantum –Atom Optics and a<br />

<strong>Swinburne</strong> <strong>University</strong> Strategic Initiative fund .<br />

8<br />

4<br />

g r a d i e n t ( T T/<br />

/ m )<br />

References<br />

f i e l d<br />

0<br />

0 50 100 150 200 250<br />

distance from surface ( μm)<br />

[1] Hänsel W, Homme lh<strong>of</strong>f P, Hänsch TWand JReichel J 2001 Nature 413 498–501<br />

[2] Ott H, Fort ágh J, Schlotterbe ck G, Grossmann A and Zimmerman n C 2001Phys. Rev. Lett. 23 230401<br />

[3] Lean hardt A, Chikkatu r A, Kielpinski D, Shin Y, Gustavson T, Ketterl e W and Pritchard D 2002 Phys. Rev.<br />

Lett. 89 040401<br />

[4] Müller D, Anderson D Z, GrowRJ, Schwindt P DDand Cornell E A 1999 Phys. Rev. Lett. 83 5194<br />

[5] Calarco T, HindsEA, Jaksch D, Schmiedm ayer J, Cirac J I and Zoller P 1999 Phys. Rev. A 61 022304<br />

[6] Wang Y-J, Anderson D Z, Bright V M, Cornell E A, Diot Q, Kishimoto T, Prentiss M, Saravana nRA,<br />

Segal S Rand Wu S 2005 Phys. Rev. Lett. 94 090405<br />

[7] Shin Y, Sanne r C, Jo G-B, Pasquini TA, Saba M, Kett erle Wand Pritchard D E2005 Phys. Rev. Lett. 95 170402<br />

[8] Schumm T, H<strong>of</strong>ferbe rth S, Andersson L M, Wildermut h S, Groth S, Bar-Josep h I, Schmiedm ayer J and<br />

Kr üge r P 2005Nature Phys. 1 57–62<br />

[9] Fort ágh J, Ott H, Kraft S, G ünthe rAand Zimmerma nn C 2002Phys. Rev. A 66 041604<br />

[10] Estève J, Aussibal C, Schumm T, Figl C, Mailly D, Bouchoule I, Westbro ok C I and AspectA2004 Phys. Rev.<br />

A 70 043629<br />

[11] Jone s MPA, Vale C J, Sahagun D, Hall BVand Hinds E A 2003 Phys. Rev. Lett. 91 080401<br />

[12] Scheel S, Rekdal P K, Knight P Land HindsEA2005 Preprint quan t-ph/0501149<br />

[13] Barba I, Gerritsma R, Xing YT, GoedkoopJBand SpreeuwRJC2005 Eur. Phys. J. D 00055-3<br />

[14] Sinclair CDJ, Curtis E A, Llorent e Garcia I, RetterJA, Hall BV, Eriksson S, SauerBEand Hinds EA2005<br />

Phys. Rev. A 72 031603<br />

[15] Sinclair CDJ, Curtis E A, Llorent e Garcia I, RetterJA, Hall BV, Eriksson S, SauerBEand Hinds EA2005<br />

Eur. Phys. J. D 35 105–110<br />

[16] Eriksson S, Ramirez-Martinez F, Curtis EA, Sauer BE, Nutt er P W, Hill EWand Hinds EA2004 App l. Phys.<br />

B 79 811<br />

[17] Jaakkola A, Shevchenkon A, Lindfors K, Hautakorpi M, Il’yashen ko E, JohansenTHand Kaivola M 2005 Eur.<br />

Phys. J. D 00176-7<br />

[18] Jackson JD1999 Classical Electrodynamics 3rd edn (New York: Wiley) chapte r 5<br />

[19] WangJY, Whitlock S, Scharnberg F, Gough D S, Sidorov A I, McLean RJ and Hannaford P 2005 J. Phys. D:<br />

App l. Phys. 38 4015–20<br />

[20] Vale C J, Upcr<strong>of</strong>t B, Davis M J, Heckenb erg N R and Rubinsztein Dunlop H 2004 J. Phys. B: At. Mol. Opt.<br />

Phys. 37 2959–67<br />

[21] Reichel J, H änsel W and Hänsch TD1999 Phys. Rev. Lett. 83 3398<br />

[22] Wang D, Lukin M and Demler E 2004 Phys. Rev. Lett. 92 076802<br />

Figure 6. Measureme nts <strong>of</strong> the mag netic field streng th (a) and field gradient (b) as a function <strong>of</strong><br />

distanc e from the surface. The dat a (open circles) agree well with pred ictions (dotte d line) <strong>of</strong> the<br />

simple mode l (see equat ion (1)). Experiment al errors are mostly det ermined by image resolut ion<br />

and a small unce rtainty in the pixel size calibration.<br />

195<br />

In add ition, the trap bot tom was mea sured using RF out coupling with an accuracy bet ter than<br />

10 mG(? 1%). The measure men t <strong>of</strong> trap frequen cy in comb ination with the trap bott(By) om<br />

unambigu ously determi nes the local mag netic field gradient (see equa tion 2). This comb ined<br />

with the trap posit ion meas uremen ts have bee n used to provide the mag netic field and the<br />

magneti c field gradien t as afunction <strong>of</strong> heigh t above the surface (figure 6). These dat a are<br />

consisten t withaprediction base d on the simple mode l whe re the film thickness -mag netization<br />

produc t is given by the prior SQUID measure men t hM ( = 0.20 A).<br />

7. Discussion and conclusion<br />

We have dem onstrate d a hybrid ato m chip tha t exploits perpen dicularly mag netized film or<br />

current -carrying wires for the prod uction <strong>of</strong>aBEC. We have develop ed a multilayer mag netic<br />

film structur e (TbGdFeCo/Cr) tha t provides large mag netization and thickness, important<br />

for realizing tight and flexible mag netic microtrap s. We have used the BEC as asens itive<br />

prob e to directly mea sure the local magneti c field and gradient associated with the mag netic<br />

film. These measure ment s justify the use <strong>of</strong> the simple mod el for perpen dicularly mag net ized<br />

magneti c microstructu res.<br />

At prese nt we are extendin g the techniqu e <strong>of</strong> cold ato m mag neto metr y to the measur ement<br />

<strong>of</strong> the spatia l dependenc e <strong>of</strong> the mag netic field along the film edg e. Spatially dep endent<br />

magneti c field variations have bee n obser ved above micr<strong>of</strong>abricate d wire-base d ato m chips<br />

and have bee n att ribute d to spatia l deviations along the wire edg 10, e 22]. [ Similar phe nom ena<br />

obser ved in permane nt mag netic structures may be caused by sub strat e roug hness , dep osition<br />

irregularity or ultimate ly doma in reversal. Future stud ies are aimed at the interac tion betw een a<br />

BEC and mag netic thin films. Acompa rison <strong>of</strong> the dec ohe rence rate s <strong>of</strong> conden sate s confined<br />

in eithe r the film or wire-base d microtraps may reveal intriguing pos sibilities for coherent<br />

manipulation <strong>of</strong> cold ato ms in microstructu red perman ent mag netic pot entia ls.


A.3. Perpendicularly magnetized, grooved GdTbFeCo microstructures for<br />

atom optics<br />

A.3 Perpendicularly magnetized, grooved GdTbFeCo<br />

microstructures for atom optics<br />

JYWang et al<br />

2. Principles <strong>of</strong> magnetic atom optics with periodic<br />

structures<br />

A one-dime nsional periodic array <strong>of</strong> mag net s <strong>of</strong> alterna ting<br />

polarity or aone- dime nsiona l periodi cally grooved magn etic<br />

structur e produce samagneti c field pat tern tha t is well suited<br />

to ato m optic s [2]. For such an array in the xz plane with<br />

periodicity in thex direction , the mag nitude <strong>of</strong> the mag net ic<br />

field depen ds on heig hty above the surface as given by 3, [ 4]:<br />

|B(x,y)|= B0e −ky [(1− e −kb )<br />

+ 1<br />

3 (1− e− 3kb )e − 2ky cos 2kx +..], (1)<br />

wherek− 1 = a/2? is the deca y leng th,a is the perio d <strong>of</strong> the<br />

array,bis the thickness <strong>of</strong>the mag net s andB0 isacharacteristic<br />

magneti c field tha t is define d by the magneti zation M 0 <strong>of</strong><br />

the mat erial. For an array <strong>of</strong> magn ets <strong>of</strong> alternat ing polarity<br />

B0 = 8M 0 (Gaussian units) and for agrooved structure<br />

B0 = 4M 0. The factors (1− e−nkb ) accoun t for the finite<br />

thickness <strong>of</strong> the magneti c material . For heig htsy a/4?<br />

above the surface equation 1) ( reduce s to<br />

|B(x,y)|= B0(1− e −kb )e −ky it may be turne d into an ato mic mat ter wave diffraction<br />

grating by app lying asmall bias magn etic field normal to the<br />

microstruct ure surface to produce aspatial diffraction grating<br />

[2,5] or by app lying an oscillating orthogona l mag netic field to<br />

createatempora l diffraction grating for ato ms 6]. [ In addition,<br />

it is possible to gen erate magn etic microtraps and waveguides<br />

for low magn etic field-seeking ato ms by app lying app ropriate<br />

dc bias fields to produce aseries <strong>of</strong> mag net ic field minima<br />

above the array surface [7]. A mag netic tube for tran sporting<br />

ato ms may be formed from acylindrically shape d periodic<br />

magn etic structu re producin garadially varying magn etic field<br />

tha t guides ato ms along the axis <strong>of</strong> the cylinder 8, [ 9].<br />

The first ato mic mirror to retro -reflect cold ato ms was<br />

based on aud iotape ont o which asinusoidal mag netic pat tern<br />

<strong>of</strong> period 9.5µm had been recorded 10]. [ Subsequen tly, sine<br />

waves with periods <strong>of</strong>around 15µm were recorded ont o floppy<br />

disks [11, 12] and videota pe [13]. These magn etic recording<br />

medi a magn etize in-plane, which limits the smallest period<br />

pat tern tha t can be recorded and makes the recording <strong>of</strong>pat terns<br />

<strong>of</strong> arbitrary shape difficult.<br />

Magn etic mirrors have also been constructe d based on<br />

, (2) periodic arrays <strong>of</strong> perma nen t NdFeB [14] and SmCo [15]<br />

magn ets <strong>of</strong> alternating polarity. Althoug h such magn ets<br />

so tha t the magn itude <strong>of</strong> the magneti c field decay s<br />

can produce large fields, they cann ot be used to produce<br />

exponen tially with heig hty. The x andy component s <strong>of</strong> the<br />

structu res with micron-scale periodicities. In our magn etic<br />

magneti c field both vary sinusoidally in thex direction , with<br />

ato m opt ics programme, we have previously att emp ted to<br />

a pha se differenc e <strong>of</strong>?/2, so tha t they combine to produc e<br />

construct periodic grooved microstructure s <strong>of</strong> ferroma gne tic<br />

flat magn etic equipote ntials. When slowly moving atom s in<br />

nickel, cobalt and alnico [16,17]. These microstructu res were<br />

posit ive or low field-seeking mag netic state s mgF ( > 0, whe re<br />

mad e by electron beam lithog raphy followed by sputt ering<br />

m is the mag netic quant um num ber <strong>of</strong> the stat e and gF the<br />

and electrop lating processes tha t resulted in the ent ire grooved<br />

Land e g-factor) app roach the surface <strong>of</strong> such an array, they<br />

structu re being mad e <strong>of</strong> the ferromag netic mat erial. In such<br />

are repelle d by the increasing mag netic field strengt h and the<br />

a structure , the se magn etic med ia also magn etize in-plane<br />

array behaves as an atomi c mirror (figure1). The origin <strong>of</strong><br />

and have astrong preference for mag netizing parallel to the<br />

the repuls ive force is the mag netic dipole interaction tha t has<br />

direct ion <strong>of</strong> the grooves. However, for mag netic ato m opt ics,<br />

potentialU int(x,y,z) =− µ·B(x,y,z) produci ng the gradien t<br />

in order to produce the app ropriate magn etic field distribution<br />

force Fgrad = ? (µ·B)= − mgFµB?B(x,y,z) , where µB is<br />

above the grooved surface <strong>of</strong>materia ltha t magn etizes in-plane,<br />

the Bohr magn eton.<br />

it is nece ssary to magn etize the structu res at right ang les to the<br />

Aperiodic magneti c array in the form <strong>of</strong>amicrostruct ure<br />

groove direct ion. The structu res failed to produce satisfactory<br />

is also the basis <strong>of</strong> othe r ato m opti cs devices. For examp le,<br />

results because <strong>of</strong> the difficulty in mag net izing the micronscale<br />

protrusions between the grooves in this way. Magn etic<br />

force microscope (MFM) images revealed dom ain structu re in<br />

the protrusions tha t indicated incomp lete magn etization, and a<br />

magn etic field tha t fell <strong>of</strong>f withadeca y constant characteristic<br />

<strong>of</strong> the mag netic dom ain size rather tha n the periodicity <strong>of</strong> the<br />

structu re [17].<br />

Mate rials with a perpen dicularmagn etic anisotro py do<br />

produce the required magn etic field distribution whe n mag -<br />

netized along the easy axis <strong>of</strong> mag netization. Microstructure<br />

s compr ising Co0.8Cr0.2 films on anon -magne tic grooved<br />

substrate and magn etized perpen dicular to the array surface<br />

have been successfully used as mag netic mirrors 18,19]. [ For<br />

ato m opt ics appl ications, however, the mag net ic properties <strong>of</strong><br />

Co0.8Cr0.2 films are inferior to tho se <strong>of</strong> GdTbFeCo magn eto -<br />

opt ical films. In part icular, the shape <strong>of</strong> the hyste resis loop<br />

indicate s tha t the rema nen t mag netization is only about one<br />

quarte r <strong>of</strong> the satura tion magn etization and tha t the mag netic<br />

dom ains are not comp letely orient ed, giving rise to magn etic<br />

inhomo gen eities.<br />

Magn eto -optical thin films, such as ferrimagnetic<br />

Figure 1. Reflection <strong>of</strong> ato ms by aperiodic ally grooved structure (Gd,Tb)FeCo and (Dy,Tb)FeCo are widely used in mag-<br />

coate d with magn etic film with perpen dicular magn etic anisot ropy. netic recording and device app lications due to the ir high<br />

INSTITUTE OF PHYSICS PUBLISHING JOURNAL OF PHYSICS D: A PPLIED PHYSICS<br />

J. Phys. D: Appl. Phys. 38 (2005) 4015–4020 doi:10.1088/00 22-3727/38/ 22/003<br />

Perpendicularly magnetized, grooved<br />

GdTbFeCo microstructures for atom<br />

optics<br />

JYWang,SWhitlock, FScharnberg ,DSGough ,AISidorov,<br />

RJMcLean andPHannaford<br />

Centr e for Atom Optics and Ultrafast Spectro scopy and ARC Cent re <strong>of</strong> Excellence for<br />

Quantum- Atom Optics, <strong>Swinburne</strong> <strong>University</strong> <strong>of</strong> <strong>Technology</strong>, Hawthorn, 3122, Australia<br />

Received 15 March 2005, in final form 26 Augu st 2005<br />

Published 7Novemb er 2005<br />

Online at stacks.iop.org/JPhysD/38/4015<br />

Abstract<br />

Periodically grooved, micron-scale structure s incorporating perpend icularly<br />

magnetiz ed Gd10Tb 6Fe80Co4 magneto -optical films have bee n fabricated<br />

and characterized . Such structure s produceamagn etic field having flat<br />

equi pote ntials and whose magnitu de decays expone ntially with distanc e<br />

above the surface, making them att ractive for manipulating ultracold ato ms<br />

in ato m optics. The GdTbFeCo films have bee n deposited on aCr<br />

under layer on asilicon (100) wafer and on agrooved silicon microstructure<br />

using DC magnetro n sputt ering. The films are found to have excellent<br />

magneti c properties for mag net ic ato m optics app lications, including high<br />

remanen t magneti zation, high coercivity and excellent homog eneity.<br />

196<br />

(Some figure s in this article are in colour only in the electro nic version)<br />

1. Introduction<br />

practicalato m optics-based devices willincorporate permanent<br />

magn ets.<br />

Atom optics involves the manipulation <strong>of</strong> atom s, particularly Acommon requireme nt in magn etic ato m optical devices<br />

cold atom s, in an ana logo us wayto the waylight is manipulate dis<br />

that the mag netic struct ure be periodic and have feature s<br />

in optics. High qua lity eleme nts are nee ded in atom on the scale <strong>of</strong> amicron. This is nece ssary to produc e a<br />

optics for many purposes includin g the reflection, diffraction, ‘hard’ magn etic mirror where the magn etic field gradien t is<br />

beam splitting, trapp ing, storag e and guiding <strong>of</strong> slowly movinglarge<br />

eno ugh so that the ato m interacts with it over ashort<br />

atom s and atomi c mat ter waves. The force commo nly used todistance<br />

, and to use the periodic struct ure as the basis for a<br />

manipulate the atom s derives from the interaction betw een the diffraction grating for ato ms, requiring the period <strong>of</strong>the grating<br />

induced electric dipole mome nt <strong>of</strong> the ato m and the electricto<br />

be comparable to the de Broglie wavelengt h <strong>of</strong> the ato m<br />

field gradien t associated with a laser light field (see, for matte r waves for reasonab le diffraction ang les and inten sities.<br />

example [1]), but there are many advant age s in exploiting theDifficulties<br />

in micromachining and magn etizing materia ls on<br />

force that results from the interaction betwee n the magnet ic this scale can be avoide d by the use <strong>of</strong>mag net ic films deposited<br />

dipole mom ent <strong>of</strong> the ato m and the magn etic field gradie nt on non- magn etic microstruct ures. In this paper we discuss<br />

nea r, e.g. a current carrying wire or apermanen t mag net icthe<br />

app lication <strong>of</strong> perpend icularly mag netized GdTbFeCo<br />

structure [2]. These advanta ges include eliminating the nee d magn eto -optical films widely used in the recording indu stry<br />

for alaser to generat e the inhomogene ous light field, and theto<br />

the fabrication <strong>of</strong> periodic microstruct ures for magn etic<br />

fact that the atom s can remain in the groun d stat e so that the ato m optics. In the following section we out line the principles<br />

coheren ce-destroying process <strong>of</strong> spontaneo us emission doe s<strong>of</strong><br />

mag netic ato m optics with periodic struct ures and the use<br />

not occur (see, for example [3]). Furthermore, perman ent <strong>of</strong> grooved microstruct ures for producing asuitable magn etic<br />

magnet s have advant age s over the use <strong>of</strong> current carryingfield<br />

gradien t, while in section3 we describe the production<br />

wires to produc e the magneti c field gradien t that includeand<br />

characterization <strong>of</strong> GdTbFeCo films and GdTbFeCo<br />

eliminatin g the problems <strong>of</strong> heating, curren t instabilities andfilm-ba<br />

sed microstruct ures with magn etic properties that are<br />

short and open circuits. It is likely, therefore , that many attrac tive for ato m optics.<br />

4016<br />

0022-3727/05/224015+06$30 .00 © 2005 IOP Publishing Ltd Printed in the UK 4015


JYWang et al<br />

Perpen dicularly magn etized, grooved GdTbFeCo microstructure s<br />

(a)<br />

Chapter A: Reprints <strong>of</strong> selected Publications by the Candidate<br />

(a)<br />

(b)<br />

(b)<br />

Figure 3. MFM micrograph s showing the domain structur e <strong>of</strong> a<br />

Figure 2. AFM micrograph s <strong>of</strong> the surface morpho logy <strong>of</strong> 150 nm thick Gd10Tb6Fe80Co 4 film prepar ed on a140 nm thick Cr<br />

(a) a 150 nm thick Gd10Tb6Fe80Co 4 film prepared on a140 nm thick underlayer on an Si wafer (a) unma gne tized and b) ( mag netized.<br />

Cr underlayer on an Si wafer and (b) a 140 nm thick single-layer Cr<br />

film deposite d on an Si wafer.<br />

AFM and MFM imaging. Figure 5 shows typical AFM<br />

top ograph y and MFM phas e images <strong>of</strong> the 3µm period<br />

SQUID magn etome ter mea surement s <strong>of</strong> the magn eticgrooved<br />

microstructu re obtained by scanning a) ( the surface<br />

propertie s <strong>of</strong> the Gd10Tb6Fe80Co4 films with Cr underlayers for the AFM image and (b) 100 nm above the top surface <strong>of</strong><br />

deposite d on an Si wafer were also carried out , up to the magn etized grooved microstructu re for the MFM image.<br />

a maximum applied field <strong>of</strong> 10 kOe. Figure 4 shows The AFM scan shows uniform periodic signals and indicates<br />

a hysteresis loop mea sured at room tem perat ure in the directiontha<br />

t the side walls are reasonably perpen dicular with some<br />

perpe ndicular to the film. It has ashap e close to rectan gularroun<br />

ding due to instrumen tal effects <strong>of</strong> the AFM. A statistical<br />

and indicates the film has an intrinsic coercivity <strong>of</strong> about analysis indicates asurface roug hnes s <strong>of</strong> about 2 nm (rms).<br />

− 3<br />

2.7 kOe and a rema nen t magn etization <strong>of</strong> about 265 emu cm The MFM images indicate excellent magn etic homog eneity,<br />

(3.3 kG). The GdTbFeCo films deposite d directly on Si were with no evidence <strong>of</strong> the domain structu re <strong>of</strong> figure 3(a). The<br />

found to have inferior magneti c prope rties to thos e depo sitedvariation<br />

<strong>of</strong> they (vertical) compo nen t <strong>of</strong> the mag net ic field (to<br />

on the Cr underlayer, particularly in term s <strong>of</strong> coercivity which the MFM is sensitive) with distancex is appro ximately<br />

(< 1.5 kOe).<br />

sinusoidal (figure5(b)), even ataheight <strong>of</strong> only 100 nm (about<br />

The surface topolog y and magn etic characteristics <strong>of</strong> the 0.2× a/2?) above the top surface. The observed sinusoidal<br />

magneti zed grooved microstruct ures with periodicities <strong>of</strong> 1.5depen<br />

dence at distances very close to the surface is att ribute d<br />

and 3µm coate d with a150 nm thick GdTbFeCo film on to some roun ding <strong>of</strong> the top edges <strong>of</strong> the groove walls during<br />

a 140 nm thick Cr underlayer were also investigated with the fabrication process, which significantly decrease s the<br />

perpe ndicular magn etic anisotro py, high saturation magne -tempe<br />

ratur e <strong>of</strong> around 300 ˚C [26]. The effect <strong>of</strong> process<br />

tization and large coercivity (see, for example 20]). [ We paramet ers, such as argon gas pres sure, subst rate temp erat ure,<br />

have fabricated periodically grooved microstruct ures based DC power and dep osition time, on mag netic properties <strong>of</strong><br />

on Gd10Tb6Fe80Co4 films with perpendi cular anisotro py and the films was investigated , and the dep osition conditions for<br />

investigate d the ir properti es. Optimizing the magn etic prop erprepa<br />

ring the bes t GdTbFeCo films within the capa bilities<br />

ties <strong>of</strong> thes e films is crucial to successful device developmen t <strong>of</strong> the dep osition system were established using astatistical<br />

and this deman ds careful prepar ation <strong>of</strong> the magn eto-opt ical design <strong>of</strong> experiment meth odo logy. The opt imal dep osition<br />

films and charac terization <strong>of</strong> the ir prop erties. This pap er paramet ers were found to be an argon pressure <strong>of</strong> 4 mTorr,<br />

repor ts investigation into thes e issues, and the results obtai nedsubst<br />

rate tempe ratur e <strong>of</strong> 100 ˚C , dc discharge power <strong>of</strong> 150 W<br />

indicate tha t the structure s are highly suitabl e for ato m opticsand<br />

a dep osition time <strong>of</strong> 50 min. After opt imizing the<br />

applica tions.<br />

process paramet ers, a 140 nm thick chromium underl ayer was<br />

Anothe r mag neto -opt ical materia l with perpe ndiculardep<br />

osited ont o an Si (100) wafer and Si grating structure s<br />

anisotr opy tha t is a promising candid ate for app lications(periodicitiesa=<br />

1.5 and 3µm and groove dep th <strong>of</strong> 0.5µm)<br />

in ato m optic s is CoPt. CoPt films usually compri se atadep osition rate <strong>of</strong>10 nm min<br />

multiple alterna ting layers <strong>of</strong> Co and Pt. Such astruct ure<br />

was recently mag netized with aview to produc ing atom ic<br />

microtrap s above its surface by ruling pattern s <strong>of</strong> opposite<br />

magneti c polarity in the film usingamag net o-optica lrecording<br />

techniq ue [21]. We have used asimilar magneto -opt ical<br />

recordin g techniq ue with thicker TbFeCo magn eto-opti cal<br />

films [17, 18, 22] but the qua lity <strong>of</strong> the recorded pat tern s was<br />

limited by dema gnetizat ion <strong>of</strong> the film adjacen t to the writing<br />

laser during the recordin g process . Both CoPt and GdTbFeCo<br />

films appea r to have excellent magn etic characteristic s for<br />

ato m opti cs, altho ugh one poss ible limitation <strong>of</strong> CoPt may<br />

be the relatively small thicknesse s (<strong>of</strong> the orde r <strong>of</strong> 50 nm) tha t<br />

appea r to be necess ary if the perpe ndicular anisotro py is to be<br />

maintained.<br />

Many ato m opti cs experimen ts now involve miniatu rizing<br />

and integrating ato m optica l elem ent s on the surface <strong>of</strong> an<br />

‘atom chip’(see, for example [23]) and magn eto-opt ical films<br />

will be useful in this tech nolog y as well. The use <strong>of</strong> permane nt<br />

magneti c mat erials in ato m chips <strong>of</strong>fers potent ial advant age s<br />

in overcoming the instab ility <strong>of</strong> the mag netic potentia l due to<br />

curren t fluctuat ions, as well as the problem s men tioned earlier<br />

<strong>of</strong> heati ng from current s and short and ope n circuits. In our<br />

group , cold Rb atom s have recently bee n successfully trapped<br />

in amag net ic trap generat ed by amagn eto-opt ical film above<br />

the surface <strong>of</strong> an ato m chip 24]. [ In this case, no grooved<br />

structur e is involved, but the magneto -opt ical film is similar to<br />

tha t used for the magn etic mirror microstructur e discussed in<br />

this pape r.<br />

3. Film preparation and characterization<br />

The GdTbFeCo films were prepare d using a thin film<br />

depositi on system (Kurt J Lesker CMS-18) capabl e <strong>of</strong><br />

magnetr on sputt ering and electro n bea m evapor ation.<br />

A comp osite target with anominal atomi c composi tion <strong>of</strong><br />

Gd10Tb6Fe80Co4 and a chromium target were magnetr on<br />

spu ttered in the system . The magn etic prop erties <strong>of</strong>GdTbFeCo<br />

film vary considerabl y with composi tion (see, for example<br />

[25]). Briefly, the Curie temp eratur e increases with the Co/Fe<br />

ratio, and the mag net ization and coercivity vary not only<br />

with the Tb/Gd ratio but also with the amou nt <strong>of</strong> Co; the<br />

degree <strong>of</strong> perpen dicular anisotr opy depe nds on the Tb/Gd<br />

ratio and the rare earth to transition met al ratio. Film<br />

prepar ation parameter s also influence the mag netic prop erties.<br />

The composi tion selecte d was expected to produc e films<br />

with high rema nen t magn etization and coercivity and aCurie<br />

− 1 , followed bya150 nm thick<br />

GdTbFeCo film at adep osition rate <strong>of</strong> 3 nm min − 1 . The base<br />

pres sure <strong>of</strong> the chambe r was less tha n 5× 10− 8 Torr prior to<br />

introdu cing the argon gas and the target to subs trat e distance<br />

was 0.2 m. Finally, a 20 nm thickY2O3 film was dep osited ont o<br />

the GdTbFeCo film as aprotect ive layer using electron beam<br />

evapora tion. The distanc e bet wee n the evapo ration source and<br />

the subst rate was 0.7 m.<br />

Analysis <strong>of</strong> the film compo sition by indu ctively coupled<br />

plasma spe ctroscopy gives an ato mic compo sition <strong>of</strong><br />

Gd9.6Tb6Fe80Co4.4, which is close to the nomina l compo sition<br />

<strong>of</strong> the target . It is well known tha t having aCr underlayer<br />

can pos itively influen ce the surface morpho logy and improve<br />

the mag netic properties <strong>of</strong> magn etic films. To give some<br />

insight into the value <strong>of</strong> prepa ring the magn eto -opt ical film<br />

on a Cr und erlayer, a single-layer 140 nm thick Cr film<br />

and asingle-layer 150 nm thick GdTbFeCo film were also<br />

prepa red. All thre e films were prepare d on Si wafers rath er<br />

tha n grooved structu res to facilitate the characterization. The<br />

surface feature s <strong>of</strong> the films were examined immediately after<br />

the sampl es were removed from the chambe r by an ato mic<br />

force microscope (AFM) ope rating in high resolution, semicontact<br />

mode .<br />

Figure 2 shows AFM micrographs <strong>of</strong> the surface<br />

morpho logy <strong>of</strong> the GdTbFeCo film on aCr underlaye r and<br />

the single-layer Cr film. The two films exhibit similar surface<br />

morpho logies. Both are den se and the ir surfaces are smooth .<br />

The grain shape on bot h surfaces is found to be round , with<br />

an average grain size <strong>of</strong> appro ximately 40 nm. The single<br />

Cr layer possesses aslight ly larger grain size and is alittle<br />

roug her tha n the GdTbFeCo film on aCr under layer. By<br />

contrast, GdTbFeCo films dep osited direct ly ont o Si wafers<br />

have roug her surfaces and larger grain sizes (50 nm). The<br />

smaller grain size <strong>of</strong> the GdTbFeCo films whe n dep osited on a<br />

Cr under layer may result from the enh anced surface roug hness<br />

<strong>of</strong>the underlaye r,suggesting tha t it aids the fabrication <strong>of</strong>den se<br />

GdTbFeCo films with smaller grain size and smooth er surface.<br />

Figure 3 shows MFMmicrographs <strong>of</strong> the domai n structu re<br />

<strong>of</strong> the same film as in figure2 mad e by scanning ataheig ht <strong>of</strong><br />

100 nm above the film surface. In the unma gne tized state a), ( a<br />

labyrinth <strong>of</strong> doma in pat terns with smooth surface contour s can<br />

be clearly obser ved. These are typical <strong>of</strong> GdTbFeCo magn eto -<br />

opt ical thin films with large perpen dicular anisotro py. In the<br />

magn etized state b), ( the re is no domai n structure visible,<br />

indicating tha t the film has excellent magn etic homog enei ty<br />

down to the lower limit <strong>of</strong> resolution <strong>of</strong> the MFM (about<br />

100 nm).<br />

4017<br />

197<br />

4018


A.3. Perpendicularly magnetized, grooved GdTbFeCo microstructures for<br />

atom optics<br />

[3] Hind s E A and Hughes IG 1999 J. Phys. D:Appl. Phys. 32<br />

R119<br />

[4] Sidorov A I, Lau D C, Opat G I, McLean RJ, Rowlands W J<br />

and Hanna ford P 1998Laser Phys. 8 642<br />

[5] Davis T J 2001 Eur. Phys. J. D 14 111<br />

[6] Opat G I, Nic Chormaic S, Cant wel l BP and Richmond J A<br />

1999 J. Opt. B:Quantum Semiclass. Opt.1 415<br />

[7] Sinclai r C D J, Rette r J A, Curtis E A, Hall BV, Llorente<br />

Garcia I, Eriksson S, Saue r BEand Hind s E A 2005<br />

Preprint physics/0502073<br />

[8] Myatt C J, Newbury N R, Guist RW, Luitzenhiser S and<br />

Wieman C E 1996 Opt. Lett 21 290<br />

[9] Richmond J A, Cant well BP, Nic Chormaic S, Lau D C,<br />

Akulshi n A M and Opat G I 2002 Phys. Rev. A 65 33422<br />

JYWang et al<br />

in thezdirection. If an add itional field <strong>of</strong> 1 gauss is app lied in<br />

thex direction to sup press spin-flips, the n the trap would have<br />

a radial trapp ing frequency <strong>of</strong> 5.4 kHz for ground stateF = 2,<br />

m = 2 87Rb atoms.<br />

Perpen dicularly magn etized, grooved GdTbFeCo microstructure s<br />

2<br />

c)<br />

1.5<br />

u/<br />

c<br />

4. Conclusions<br />

m<br />

( e<br />

[10] Roach T M, Abele H, Boshier M G, Grossman H H, Zetie KP<br />

and Hinds E A 1995 Phys. Rev. Lett. 75 629<br />

[11] Hughe sIG, Barton P A, Roach T M, Boshie r M G and<br />

Hinds E A 1997 J. Phys. B:At. Mol. Opt. Phys. 30 647<br />

[12] Hughe sIG, Barton P A, Roac h T M and Hinds E A 1997<br />

J. Phys. B:At. Mol. Opt. Phys. 30 2119<br />

[13] Saba C V, Barton P A, Boshier M G, Hughe sIG,<br />

Rosenbusch P, Sauer BEand Hind s E A 1999 Phys. Rev.<br />

Lett. 82 468<br />

[14] Sidorov A I, McLean RJ, Rowlands W J, Lau D C,<br />

Murphy J E, Walkiewicz M, Opa t G I and Hanna ford P<br />

1996 Quantum Semiclass. Opt.8 713<br />

[15] Meschede D, Bloch I, Goepfert A, Haubrich D, Kreis M,<br />

Lison F, Schutze Rand Wynan ds R1997 Atom Optics<br />

Proc. SPIE 2995 191<br />

[16] Sidorov A I, Lau D C, Opat G I, McLean RJ, Rowlands W J<br />

and Hanna ford P 1997Proc. 13th Int. Conf. on Laser<br />

Spe ctroscopy (Hangzhou, China, 1997)(Singapore: World<br />

Scientific) p 252<br />

[17] Lau D C, McLean RJ, Sidorov A I, Gough D S, Koperski J,<br />

Rowlands W J, Sexton B A, Opa t G I and Hanna ford P<br />

1999 J. Opt. B:Quantum Semiclass. Opt 1 371<br />

[18] Sidorov A I, McLean RJ, Sexton BA, Gough D S, Davis T J,<br />

Akulshin A M, Opa t G I and Hanna ford P 2001C. R. Acad.<br />

Sci. Ser. IV 2 565<br />

[19] Sidorov A I, McLean RJ, Scharnberg F, Gough D S,<br />

Davis T J, Sexton B A, Opat G I and Hann aford P 2002<br />

Acta Phys. Pol. B 33 2137<br />

[20] Tsunashima S 2001 J. Phys. D:Appl. Phys 34 R87<br />

[21] Eriksson S, Ramirez-Martinez F, Curtis E A, Sauer BE,<br />

Nutter P W, Hil l E W and Hind s E A 2004 Appl. Phys. B<br />

79 811<br />

[22] Gough D S, McLean R J, Sidorov A I, La u D C, KoperskiJ,<br />

Rowlands W J, Sexton B A, Hannaford P and Opat G I<br />

1999 Proc. 14th Int. Conf. on Laser Spe ctroscopy<br />

(Innsbruck, Austria) (Singapore: World Scientific) p 380<br />

[23] Folman R, Kruger P, Schmiedmayer J, DenschlagJand<br />

Henkel C 2002 Adv. At. Mol. Opt. Phys. 48 263<br />

[24] Hall BV, Whitlock S, Scharnberg F, WangJY, Dalton BJ,<br />

McLean RJ, Kieu T D, Hanna ford P and Sidoro v A I 2004<br />

XIX Int. Conf. on Atomic Physics (Rio de Janeiro, Brazil,<br />

2004) Book <strong>of</strong> Abstracts, p 87<br />

[25] Kryde r M H 1993 Annu. Rev. Mater. Sci 23 411<br />

[26] Challene r W A 1997 Private communication<br />

We have fabricate d magn etic microstructures tha t show<br />

considerable promise as atom optical devices by depositing<br />

Gd10Tb 6Fe80Co4 magn eto-optical film with perpendicular<br />

magn etic anisotropy on grooved silicon microstructures with<br />

periodicities <strong>of</strong> 1.5 and 3µm. When the magnet o-optical<br />

mate rial was deposited onto a Cr underlayer the magne tic<br />

characteristics <strong>of</strong> the films were found to be significant ly<br />

improved, along with their surface topolo gy and density.<br />

A single layer <strong>of</strong> the magn eto-optical film deposited on aCr<br />

underlayer was measured at room temperature in the direction<br />

perpendicular to the film to have an intrinsic coercivity<br />

− 3<br />

<strong>of</strong> 2.7 kOe and a remanen t magn etization <strong>of</strong> 265 emu cm<br />

(3.3 kG). The periodically grooved microstructures coate d<br />

with GdTbFeCo films exhibit reasonably perpendicular sidewalls<br />

and uniform periodic modu lation in AFM and MFM<br />

scans. Optimum magne tic characteristics were found for<br />

multilayer structures <strong>of</strong> GdTbFeCo alternating with Cr. The<br />

amplitude <strong>of</strong> the componen t <strong>of</strong> magneti c field in the direction<br />

perpendicular to the grooved microstructure surface was found<br />

to decrease exponen tially with heig ht above the surface with<br />

a decay constan t consisten t with the the oretical value given<br />

by the period <strong>of</strong> the microstructure. Such perpendicularly<br />

magn etized grooved microstructures coate d with GdTbFeCo<br />

films should be well suited for atom optical app lications.<br />

1<br />

l y ( relative)<br />

on<br />

n B<br />

0.5<br />

netizati<br />

g<br />

a<br />

M<br />

0<br />

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4<br />

Height above surface ( µm)<br />

Figure 6. Plot <strong>of</strong> lnBy versus heigh t above the surface for the<br />

a= 1.5µm perpendic ularly mag netize d grooved microstructu re.<br />

The slope gives adec ay lengt h <strong>of</strong> 0.24µm.<br />

Figure 4. Hysteresis loop measur ed by SQUID mag net ome try at<br />

room temp erat ure in the perpendicula r directio n for a150 nm thick<br />

Gd10Tb6Fe80Co4 film deposite d ona140 nm thick Cr underlaye r on<br />

an Si wafer.<br />

Acknowledgments<br />

This work is funded by a Systemic Infrastructure Initiative<br />

(SII) grant from the Department <strong>of</strong> Education, Science and<br />

Training (DEST), the ARCCent re <strong>of</strong> Excellence for Quant um-<br />

Atom Optics and a <strong>Swinburne</strong> <strong>University</strong> Strategic Initiative<br />

grant . We tha nk Dr T Hicks and Mr D Robinson for their<br />

help with the SQUIDana lysis at Monash <strong>University</strong>, Australia,<br />

and W A Challene r and J Sexton <strong>of</strong> Imation Corp., USA for<br />

supp lying the early GdTbFeCo samples.<br />

References<br />

[1] Metcalf H J and van der Strate n P 1999Laser Cooling and<br />

Trapping (Heidelberg: Springe r)<br />

[2] Opat G I, Wark S J and Cimmino A 1992 Appl. Phys. B<br />

54 396<br />

supe rior magn etic characteristics to films tha t had asingle<br />

450 nm thick layer <strong>of</strong> GdTbFeCo. Withou t the Cr midlayers<br />

present the remane nt mag netization and coercivity were<br />

found to deter iorate at thicknesses larger tha n abo ut 200 nm,<br />

probab ly due to the magn etic anisotro py direct ion bein g less<br />

well defined ; but by using the multilayer structure the overall<br />

thickness <strong>of</strong> the film could be built up beyond the thickness <strong>of</strong><br />

1µm while still preserving goo d magn etic properties.<br />

A strong indicator <strong>of</strong> how the grooved microstructures<br />

work as mag netic mirrors is the dep end ence <strong>of</strong> the magn etic<br />

field on heig ht above the surface <strong>of</strong> the structu re. The<br />

dep end ence on heighty above the 1.5µm period grooved<br />

microstruct ure <strong>of</strong> they-compo nen t <strong>of</strong> the magn etic field was<br />

test ed by making aseries <strong>of</strong> MFM scans with the magn etic<br />

tip at different height s ranging from 100 to 1500 nm above the<br />

top <strong>of</strong> the microstructure . From the MFM dat a it is possible<br />

to plot the dep end ence <strong>of</strong>By on heighty (figure 6). The<br />

values <strong>of</strong>By in figure 6, which are relative values dete rmined<br />

from the amp litude <strong>of</strong> the MFM phas e signal, indicate tha t the<br />

amp litude <strong>of</strong>By decre ases expon ent ially with heig ht above<br />

the surface. The mag nitud e <strong>of</strong> the magn etic field is expect ed<br />

to deca y with the same dep end ence. The slope yields a<br />

deca y lengt h <strong>of</strong> (0.24 ± 0.02)µm, which is in agreemen t<br />

with a/2? = 0.24µm given by the ory [2], with most <strong>of</strong> the<br />

unce rtainty arising from the calibration <strong>of</strong> the vertical position<br />

<strong>of</strong> the MFM tip.<br />

The MFMmea sureme nts do not give absolute values <strong>of</strong>the<br />

magn etic field above the surface, but we can use the value <strong>of</strong>the<br />

magn etization from the SQUIDmea sureme nts to estimate how<br />

Figure 5. Micrograph s <strong>of</strong>aperpe ndicularly mag netize d grooved we expec t the ato m opt ical elemen ts based on the multilayer<br />

microstruct ure fabricate d with a150 nm thick Gd 10Tb6Fe80Co4 film GdTbFeCo film to perform. From equation (2), the magn itude<br />

ona140 nm thick Cr under layer onasilicon gratin g structure with a <strong>of</strong> the magn etic field at the surface <strong>of</strong> the structu re, using<br />

perio d <strong>of</strong> 3µm. (a) AFM scan and (b) MFM scan. In ( b) the<br />

a= 1.5µm,b= 0.45µm andM = 265 emu cm<br />

grooves are represente d by the light regions and the inset shows a<br />

cross-sectio n <strong>of</strong> the signal along the indicate d horizonta l line.<br />

cont ribution <strong>of</strong> higher orde r spatial harm onics in the mag netic<br />

potentia l [19].<br />

The films depo sited ont o the grating struct ure for cold<br />

ato m experiments were fabricated with amultilayer structu re<br />

comprisin g thre e 140 nm thick layers <strong>of</strong> Cr alternat ing with<br />

thre e 150 nm thick layers <strong>of</strong> Gd10Tb6Fe80Co4. This gave<br />

− 3 , is abo ut<br />

900 gau ss. Given tha t a87Rb ato m in theF = 2, m = 2<br />

ground state dropp ed fromaheig ht <strong>of</strong> 10 mm will be reflected<br />

by afield <strong>of</strong> 16 gau ss, the reflection should occur abo ut one<br />

micron above the surface, well above any surface effects and<br />

effects from the higher order term s in equat ion 1). (<br />

If the film is used on the surface <strong>of</strong> an ato m chip, the n,<br />

for example, a two-dime nsiona l quadrup ole magn etic trap is<br />

produce d at aheight <strong>of</strong> 25µm above an edg e <strong>of</strong> the film in<br />

the x direct ion by appl ying abias magn etic field <strong>of</strong> 10 gau ss<br />

4019<br />

198<br />

4020


Appendix B<br />

Determining the Temperature<br />

from One Image<br />

In the case in which the trap can be assumed to be harmonic, we have a relation<br />

between the temperature T <strong>of</strong> the atoms and the trap frequency νT = ωT<br />

2π from<br />

equating the energies:<br />

1<br />

2 kBT = 1<br />

2 m(ωt · q) 2<br />

⇒ T = m<br />

(2π · νT · σ) 2<br />

kB<br />

(B.1)<br />

where the spatial variable q can be substituted by the measurable spread <strong>of</strong><br />

the atom cloud in the trap σ. This can be determined directly by an in-situ<br />

absorption image <strong>of</strong> the atoms or can be calculated from the intercept <strong>of</strong> the<br />

linear regression <strong>of</strong> the TOF images which also yield the temperature.<br />

We already know that the harmonic approximation will not hold, especially<br />

at the beginning <strong>of</strong> the evaporation process. We thus can not use equation<br />

(B.1) to calculate the temperature from the size <strong>of</strong> the atomic cloud. On the<br />

other hand, we have one set <strong>of</strong> TOF measurements <strong>of</strong> the temperature and<br />

<strong>of</strong> the size <strong>of</strong> the atomic cloud in the trap for different RF frequencies for a<br />

trap <strong>of</strong> a nominal frequency <strong>of</strong> ν1 = 217 Hz. If we plot the temperature <strong>of</strong><br />

the atomic cloud for the axial and the radial direction against the respective<br />

spatial widths, as in Fig. B.1, we can use a commercial fitting s<strong>of</strong>tware 1 to<br />

interpolate between the points. As a fitting function, we chose the function<br />

with the least number <strong>of</strong> free parameters that describes the behaviour to the<br />

best degree in both directions. It showed that this was a function <strong>of</strong> the type<br />

1 TableCurve 2D v5.01<br />

T = (a + b ·<br />

199<br />

ln σi<br />

σ2 )<br />

i<br />

−1<br />

(B.2)


temperature in µK<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

atomic spread σ in mm<br />

Figure B.1: To determine the temperature <strong>of</strong> an atomic cloud from a single<br />

image, the temperature was plotted as a function <strong>of</strong> atomic cloud size for axial<br />

(solid line, circles) and radial (dotted line, crosses) direction. Marked are the<br />

data points. The fitted functions are explained in the text.<br />

The fitted functions are also shown in the plot.<br />

These functions can now be used to approximate the temperature for the<br />

unsuccessful try. As we used a different trap with ν2 = 100 Hz, we can not<br />

simply use the values for the atom cloud’s spread <strong>of</strong> those experiments in<br />

the formula. Instead, the cloud size σi has to be corrected for the changing<br />

steepness and curvature. This in a first approximation can be done if we<br />

substitute the spread by ˜σi = ν2<br />

ν1 · σi. Here we assume that the trap frequencies<br />

scale the same in both radial and axial direction. The approximation was<br />

chosen so that it will agree with the behaviour in a harmonic trap, where the<br />

temperature T ∝ (ν · σ) 2 is proportional to the square <strong>of</strong> the product <strong>of</strong> the<br />

frequency and the spread. The different number <strong>of</strong> atoms in the trap for the<br />

different tries was ignored here, as the maximum height <strong>of</strong> a Gaussian should<br />

not influence the standard deviation or width <strong>of</strong> that curve.<br />

200


Appendix C<br />

The Atom: 87 Rb<br />

The isotope used in both experiments that are described in this thesis is 87 Rb.<br />

Although the nucleons and electrons are fermionic, having an odd number <strong>of</strong><br />

nucleons and electrons adds to an overall even number. Their half-integral<br />

spins couple to an integer spin, and the overall atom is bosonic. Together with<br />

the collision properties and the easily obtained light sources to manipulate<br />

the atoms, this is one <strong>of</strong> the reasons why 87 Rb has found a widespread use:<br />

this isotope is probably the most common element to be found in atom optics<br />

and Bose-Einstein condensation experiments. Although 87 Rb is unstable and<br />

radioactive, the lifetime is more than 10 10 years and it can be considered as<br />

stable for our purposes.<br />

Rubidium, element 37, is an alkali-metal. The electronic configuration in<br />

the ground state is [Kr]5s; the ground state is thus 5 2 S1/2. The strongest<br />

transition is the D line which is fine split into the transitions to the two states<br />

5 2 P1/2 and 5 2 P3/2, with wavelengths <strong>of</strong> 794 nm (D1 line) and 780 nm (D2<br />

line). We here consider the D2 line only (see Figure C.1). The states show<br />

a hyperfine splitting, with F = 1, 2 for the ground and F ′ = 0, 1, 2, 3 for the<br />

excited state. These levels are themselves degenerate (2F + 1) times. When<br />

examining absorption spectra, we find two Doppler broadened signals from the<br />

transitions F = 1 → F ′ and F = 2 → F ′ with the different F ′ -states washed<br />

out as their spectral difference is more than a factor <strong>of</strong> 10 smaller than the<br />

separation <strong>of</strong> the ground states. Doppler free spectroscopy allows us to resolve<br />

these transitions and their crossover peaks. Extensive information about these<br />

transitions can be found in [Ste03].<br />

201


Figure C.1: The energy levels <strong>of</strong> the 87 Rb D2 line, taken from [Ste03].<br />

202


Appendix D<br />

Technical Details <strong>of</strong> the Coils<br />

for the Permanent Magnetic<br />

Chip Experiment<br />

The <strong>of</strong>fset coils<br />

Coils side length separation windings slope<br />

(mm) (mm) (G/A)<br />

Bx 240 335 10 0.235<br />

2nd set 20 0.47<br />

By 210 115 9 0.681<br />

2nd set 20 1.515<br />

Bz 300 55 7 0.49<br />

2nd set 20 1.404<br />

Table D.1: Dimensions and characteristics <strong>of</strong> the compensating <strong>of</strong>fset coils.<br />

The quadrupole coils<br />

Coils centre separation windings width depth gradient<br />

diameter<br />

(mm) (mm) (mm) (mm) (G/(Acm))<br />

Quad 160 280 400 16 35 ≈ 1<br />

Table D.2: Dimensions and characteristics <strong>of</strong> the quadrupole coils.<br />

203


The bias field coils<br />

Coils inner diameter separation windings width depth slope<br />

(mm) (mm) (mm) (mm) (G/A)<br />

Bias 400 400 66 24 27.5 2.82<br />

Table D.3: Dimensions and characteristics <strong>of</strong> the bias field coils for trapping<br />

with a single wire.<br />

trigger / V; current / A<br />

5.5<br />

5<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

(a)<br />

0 5 10 15 20 25 30 35<br />

time / ms<br />

trigger / V; current / A<br />

5.5<br />

5<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

(b)<br />

0 0.5 1 1.5 2 2.5 3 3.5<br />

time / ms<br />

Figure D.1: Switching times <strong>of</strong> the coils: (a) switch on, (b) switch <strong>of</strong>f. Dotted/thin<br />

line: trigger pulse; solid/thick line: current through the coils.<br />

204


Appendix E<br />

List <strong>of</strong> Equipment<br />

[E 1]: tapered amplifier: TA 100, electronics: SC 100, DC 100, DCC<br />

100, DTC 100; TOptica<br />

[E 2]: acousto-optical modulator: 1205C-2; Isomet<br />

[E 3]: PID Regulator: PID 100 ; TOptica<br />

[E 4]: acousto-optical modulator: 1206C; Isomet<br />

[E 5]: polarisation preserving fibre: PMJ-A3A-3AF-800-5/125-3-5-1;<br />

Oz Optics<br />

[E 6]: coupler and collimation: HPUC-2,A3A-780-P-11AS-11 and HPUCO-<br />

23AF-800-P-6.2AS; Oz Optics<br />

[E 7]: diode laser: DL 100, electronics: SC 100, DCC 100, DTC 100;<br />

TUI Optics (now: TOptica)<br />

[E 8]: lock-in regulator: LIR 100; TUI Optics (now TOptica)<br />

[E 9]: current and temperature stabilisation: LDC500 and TED200;<br />

Thorlabs<br />

[E 10]: shutter and driver: LS612 and VMM-D1; UniBlitz<br />

[E 11]: fast photodiode: Si Nanosecond Photodetector (1621); Newfocus<br />

[E 12]: CCD camera: Micromax 1024CCD; Princeton Instruments<br />

[E 13]: cold cathode gauge: TPG 300; Pfeiffer Vacuum<br />

[E 14]: turbo pump: TMU 065 DN 63 CF-F, 1P (50 l/s); Balzers Pfeiffer<br />

205


[E 15]: diaphragm pump: MD4T (3.3 m 3 /h); Vacuubrand, distributor:<br />

Balzers Pfeiffer<br />

[E 16]: ion pump: Varian Triode (75 l/s); Varian<br />

[E 17]: Ti: sublimation pump and control unit: TSP 2140412 and 224-<br />

0550; Physical Electronics<br />

[E 18]: PCB mill: Quick Circuit 5000<br />

[E 19]: RF synthesizer: SRS 200, Stanford<br />

[E 20]: thin film deposition system: CMS-18; Kurt J. Lasker<br />

[E 21]: digital and analogue I/O board: PCI-6733, BNC 2110; National<br />

Instruments<br />

[E 22]: turbo pump: TMU 260 (210 l/s); Pfeiffer Vacuum<br />

[E 23]: ion pump: Varian Noble Diode; Varian<br />

[E 24]: Ti: sublimation pump: SS-400/275; Thermionics Laboratory<br />

Inc.<br />

[E 25]: tapered amplifier: TA 100; TOptica<br />

[E 26]: lock-in amplifier: Type 401A; Brookdeal<br />

[E 27]: acousto-optical modulator: 3110-120; Crystal <strong>Technology</strong><br />

[E 28]: current, temperature controller: DCC 100, DTC 100; TOptica<br />

[E 29]: acousto-optical modulator: 3200-124; Crystal <strong>Technology</strong><br />

[E 30]: acousto-optical modulator: 3200-121; Crystal <strong>Technology</strong><br />

[E 31]: disk laser: VersaDisk; ELS Elektronik Laser System GmbH<br />

[E 32]: CCD camera: chip type: KAF0400; Photometrics Sensys<br />

[E 33]: high power laser diode: DLX 100; TOptica<br />

206


Bibliography<br />

[All75] L. Allen and J. Eberly, Optical Resonance and Two-level atoms<br />

(Dover, New York, 1975).<br />

[Alo92] M. Alonso and E. Finn, Physics (Addison-Wesley, Wokingham,<br />

1992), first ed.<br />

[And95] M. Anderson, J. Ensher, M. Mathews, C. Wieman and E. Cornell,<br />

Observation <strong>of</strong> Bose-Einstein condenstaion in a dilute atomic vapor.<br />

Science 269, 198 (1995).<br />

[And97] M. Andrews, C. Townsend, H.-J. Meisner, D. Durfee, D. Kurn and<br />

W. Ketterle, Observation <strong>of</strong> interference between two Bose condensates.<br />

Science 275, 637 (1997).<br />

[And01] B. Anderson and M. Kasevich, Loading a vapor-cell MOT using<br />

light-induced atom desorption. Phys. Rev. A 63, 023404 (2001).<br />

[And02] E. Andersson, T. Calarco, R. Folman, M. Andersson, B. Hessmo<br />

and J. Schmiedmayer, Multimode interferometer for guided matter<br />

waves. Phys. Rev. Lett. 88, 100401 (2002).<br />

[Arn99] A. Arnold, Preparation and manipulation <strong>of</strong> a 87 Rb Bose-Einstein<br />

condensate. Ph.D. thesis, <strong>University</strong> <strong>of</strong> Sussex (1999).<br />

[Bag87] V. Bagnato, D. Pritchard and D. Kleppner, Bose-Einstein condensation<br />

in an external potential. Phys. Rev. A 35, 4354 (1987).<br />

[Bal87] V. Balykin, V. Letokhov, Y. Ovchinnikov and A. Sidorov, Reflection<br />

<strong>of</strong> an atomic beam from a gradient <strong>of</strong> an optical field. JETP Lett.<br />

45, 353 (1987).<br />

[Bam81] A. Bambini and P. Berman, Analytic solutions to the two-state problem<br />

for a class <strong>of</strong> coupling potentials. Phys. Rev. A 23, 2496 (1981).<br />

207


BIBLIOGRAPHY<br />

[Bar01] M. Barrett, J. Sauer and M. Chapman, All-optical formation <strong>of</strong><br />

an atomic Bose-Einstein condensate. Phys. Rev. Lett. 87, 010404<br />

(2001).<br />

[Bar05] I. Barb, R. Gerritsma, Y. Xing, J. Goedkoop and R. Spreeuw, Creating<br />

I<strong>of</strong>fe-Pritchard micro-traps from permanent magnetic film with<br />

in-plane magnetization. Eur. J. Phys. D 35, 75 (2005).<br />

[Bec85] R. Becker, Theorie der Wärme (Springer Verlag, Berlin, 1985),<br />

third ed.<br />

[Ber96] P. Berman, Atom Interferometry (Academic Press, New York,<br />

1996).<br />

[Bir01] G. Birkl, F. Buchkremer, R. Dumke and W. Ertmer, Atom optics<br />

with micr<strong>of</strong>abricated optical elements. Opt. Comm. 191, 67 (2001).<br />

[Boh27] N. Bohr, Das Quantenpostulat und die neuere Entwicklung der<br />

Atomistik. Naturwissenschaften 16, 245 (1927).<br />

[Bor04] D. Bortolotti and J. Bohn, Wave mechanics <strong>of</strong> a two-wire atomic<br />

beam splitter. Phys. Rev. A 69, 033607 (2004).<br />

[Bra86] B. Bransden and C. Joachain, Physics <strong>of</strong> atoms and molecules<br />

(Longman Scientific, London, 1986).<br />

[Bra95] C. Bradley, C. Sacket and R. Hulet, Evidence <strong>of</strong> Bose-Einstein condensation<br />

in an atomic gas with attractive interactions. Phys. Rev.<br />

Lett. 75, 1687 (1995).<br />

[Bro91] I. Bronstein and K. Semendjajew, Taschenbuch der Mathematik<br />

(B.G. Teubner Verlag, Nauka, Stuttgart, Leipzig, Moscow, 1991),<br />

25th ed.<br />

[Bru96] M. Brune, E. Hagley, J. Dreyer, X. Maître, A. Maali, C. Wunderlich,<br />

J. Raimond and S. Haroche, Observing the progressive decoherence<br />

<strong>of</strong> the “meter” in a quantum measurement. Phys. Rev. Lett. 77,<br />

4887 (1996).<br />

[Buc01] F. Buchkremer, Kohärenz in miniaturisierten Speicher- und Leiterstrukturen<br />

für neutrale Atome. Ph.D. thesis, Universität Hannover<br />

(2001).<br />

[Buc02] F. Buchkremer, R. Dumke, M. Volk, T. Müther, G. Birkl and<br />

W. Ertmer, Quantum information processing with micr<strong>of</strong>abricated<br />

optical elements. Laser Physics 12, 736 (2002).<br />

208


BIBLIOGRAPHY<br />

[Cal00] T. Calarco, E. Hinds, D. Jaksch, J. Schmiedmayer, J. Cirac and<br />

P. Zoller, Quantum gates with neutral atoms: controlling collisional<br />

interactions in time-dependent traps. Phys. Rev. A 61, 022304<br />

(2000).<br />

[Cas97] Y. Castin and J. Dalibard, Relative phase <strong>of</strong> two Bose-Einstein<br />

condensates. Phys. Rev. A 55, 4330 (1997).<br />

[Cas00] D. Cassettari, B. Hessmo, R. Folman, T. Maier and J. Schmiedmayer,<br />

A beam splitter for guided atoms on an atom chip. Phys.<br />

Rev. Lett. 85, 5483 (2000).<br />

[Cen03] G. Cennini, G. Ritt, C. Geckeler and M. Weitz, Bose-Einstein condensation<br />

in a CO2-laser optical dipole trap. Appl. Phys. B 77, 773<br />

(2003).<br />

[Cen04] G. Cennini, Field insensitive Bose-Einstein condensates and an alloptical<br />

atom laser. Ph.D. thesis, Universität Tübingen (2004).<br />

[Chi05a] C. Chin, T. Kraemer, M. Mark, J. Herbig,<br />

P. Waldburger, H.-C. Naegerl and R. Grimm<br />

(2005), http://www.quantum.univie.ac.at/QUPON/<br />

TalksOnline/Rudi Grimm.<strong>pdf</strong>.<br />

[Chi05b] C. Chin, T. Kraemer, M. Mark, J. Herbig, P. Waldburger, H.-C.<br />

Naegerl and R. Grimm, Observation <strong>of</strong> Feshbach-like resonances in<br />

collisions between ultracold molecules. Phys. Rev. Lett. 94, 123201<br />

(2005).<br />

[Chu86] S. Chu, J. Bjorkholm, A. Ashkin and A. Cable, Experimental observation<br />

<strong>of</strong> optically trapped atoms. Phys. Rev. Lett. 57, 314 (1986).<br />

[Coo79] R. Cook and B. Shore, Coherent dynamics <strong>of</strong> N-level atoms and<br />

molecules. III. An analytically soluble periodic case. Phys. Rev. A<br />

20, 539 (1979).<br />

[Cor99] J. Corney, G. Milburn and W. Zhang, Weak force detection using a<br />

double Bose-Einstein condensate. Phys. Rev. A 59, 4630 (1999).<br />

[CT92] C. Cohen-Tannoudji, J. Dupont-Roc and G. Grynberg, Atom-<br />

Photon Interactions (John Wiley and Sons, Inc., New York, 1992).<br />

[CT97] C. Cohen-Tannoudji, J. Dupont-Roc and G. Grynberg, Photons and<br />

Atoms (John Wiley and Sons, Inc., New York, 1997).<br />

209


BIBLIOGRAPHY<br />

[D¨98] S. Dürr, T. Nonn and G. Rempe, Fringe visibility and which-way<br />

information in an atom interferometer. Phys. Rev. Lett. 26, 5705<br />

(1998).<br />

[Dal85] J. Dalibard and C. Cohen-Tannoudji, Dressed-atom approach for<br />

atomic motion in laser light: the dipole force revisited. J. Opt. Soc.<br />

Am. B 2, 1707 (1985).<br />

[Dal89] J. Dalibard and C. Cohen-Tannoudji, Laser cooling below the<br />

Doppler limit by polarization gradients: Simple theoretical models.<br />

J. Opt. Soc. Am. B 6, 2023 (1989).<br />

[Dal04] B. Dalton (2004), private communication.<br />

[Dav95] K. Davis, M.-O. Mewes, M. Andrews, N. van Druten, D. Durfee,<br />

D. Kurn and W. Ketterle, Bose-Einstein condensation in a gas <strong>of</strong><br />

sodium atoms. Phys. Rev. Lett. 75, 3969 (1995).<br />

[Dav99a] T. Davis, Atomic de Broglie waveguides and integrated atom-optics<br />

using permanent magnets. J. Opt. B: Quant. Semiclass. Opt. 1, 408<br />

(1999).<br />

[Dav99b] T. Davis, The properties <strong>of</strong> two-dimensional magnetic multipole<br />

traps from ideal planar magnets. Tech. rep., CSIRO (1999), internal<br />

Report.<br />

[Dav99c] T. Davis, Thin-film magnetic devices for monolithic integrated atom<br />

optics. Tech. rep., CSIRO (1999), internal Report.<br />

[de 24] L.-V. de Broglie, Recherches sur la théorie des quanta. Ph.D. thesis,<br />

Faculty <strong>of</strong> Sciences, Paris <strong>University</strong> (1924).<br />

[Det04] A. Dethlefsen, Elektronische Raman-Spektrospokie an ein- und<br />

doppellagigen Quantendrähten. Ph.D. thesis, Universität Hamburg<br />

(2004).<br />

[Dud03] A. Dudarev, R. Diener, B. Wu, M. Raizen and Q. Niu, Entanglement<br />

generation and multiparticle interferometry with neutral<br />

atoms. Phys. Rev. Lett. 91, 010402 (2003).<br />

[Dum02a] R. Dumke, T. Müther, M. Volk, W. Ertmer and G. Birkl,<br />

Interferometer-type structures for guided atoms. Phys. Rev. Lett.<br />

89, 220402 (2002).<br />

210


BIBLIOGRAPHY<br />

[Dum02b] R. Dumke, M. Volk, T. Müther, F. Buchkremer, G. Birkl and<br />

W. Ertmer, Micro-optical realization <strong>of</strong> arrays <strong>of</strong> selectively addressable<br />

dipole traps: a scalable configuration for quantum computation<br />

with atomic qubits. Phys. Rev. Lett. 89, 097903 (2002).<br />

[Dum03a] R. Dumke, Atomoptik und Quanteninformationsverarbeitung mit<br />

mikrostrukturierten optischen Elementen. Ph.D. thesis, Universität<br />

Hannover (2003).<br />

[Dum03b] R. Dumke, M. Volk, T. Müther, F. Buchkremer, W. Ertmer and<br />

G. Birkl, Quantum Information Processing with Atoms in Optical<br />

Micro-Structures (G. Leuchs and T. Beth, Weinheim, 2003).<br />

[Eck04] K. Eckert, M. Lewenstein, R. Corbalán, G. Birkl, W. Ertmer and<br />

J. Mompart, Three-level atom optics via the tunneling interaction.<br />

Phys. Rev. A 70, 023606 (2004).<br />

[Ein05] A. Einstein, Über einen die Erzeugung und Verwandlung des Lichtes<br />

betreffenden heuristischen Gesichtspunkt. Annalen d. Physik 17,<br />

132 (1905).<br />

[Ein25] A. Einstein, Zur Quantentheorie des idealen Gases. Sitzungsber.<br />

Preuss. Akad. Wiss., Phys.-math. Klasse 23, 3 (1925).<br />

[Eks95] C. Ekstrom, J. Schmiedmayer, M. Chapman, T. Hammond and<br />

D. Pritchard, Measurement <strong>of</strong> the electric polarizability <strong>of</strong> sodium<br />

with an atom interferometer. Phys. Rev. A 51, 3883 (1995).<br />

[Eng96] B.-G. Englert, Fringe visibility and which-way information: An inequality.<br />

Phys. Rev. Lett. 77, 2154 (1996).<br />

[Eri04] S. Eriksson, F. Ramirez-Martinez, E. Curtis, B. Sauer, P. Nutter,<br />

E. Hill and E. Hinds, Micron sized atom traps made from magnetooptical<br />

thin films. Appl. Phys. B 79, 811 (2004).<br />

[Ert85] W. Ertmer, R. Blatt, J. Hall and M. Zhu, Laser manipulation <strong>of</strong><br />

atomic beam velocities: demonstration <strong>of</strong> stopped atoms and velocity<br />

reversal. Phys. Rev. Lett. 54, 996 (1985).<br />

[Est04] J. Estève, C. Aussibal, T. Schumm, C. Figl, D. Mailly, I. Bouchoule,<br />

C. Westbrook and A. Aspect, The role <strong>of</strong> wire imperfections in<br />

micro magnetic traps for atoms. Phys. Rev. A 70, 043629 (2004).<br />

[Est05] J. Estève, T. Schumm, J. Trebbia, I. Bouchoule, A. Aspect and<br />

C. Westbrook, Realising a stable magnetic double-well potential on<br />

an atom chip. Eur. J. Phys. D 35, 141 (2005).<br />

211


BIBLIOGRAPHY<br />

[Fey57] R. Feynman, F. Vernon and R. Hellwarth, Geometrical representation<br />

<strong>of</strong> the Schrödinger equation for solving the maser problem. J.<br />

Appl. Phys. 28, 49 (1957).<br />

[Fol00] R. Folman, P. Krüger, D. Cassettari, B. Hessmo, T. Maier and<br />

J. Schmiedmayer, Controlling cold atoms using nan<strong>of</strong>abricated surfaces:<br />

Atom chips. Phys. Rev. Lett. 84, 4749 (2000).<br />

[Fol02] R. Folman, P. Krüger, J. Schmiedmayer, J. Denschlag and<br />

C. Henkel, Microscopic atom optics: from wires to an atom chip.<br />

Adv. in Atom, Mol. and Opt. Phys. 48, 263 (2002).<br />

[For98a] J. Fortagh, A. Grossmann, T. Hänsch and C. Zimmermann, Fast<br />

loading <strong>of</strong> a magneto-optical trap from a pulsed thermal source. J.<br />

Appl. Phys. 84, 6499 (1998).<br />

[For98b] J. Fortágh, A. Grossmann, C. Zimmermann and T. Hänsch, Miniaturized<br />

wire trap for neutral atoms. Phys. Rev. Lett. 81, 5310<br />

(1998).<br />

[For00] J. Fortágh, H. Ott, A. Grossmann and C. Zimmermann, Miniaturized<br />

magnetic guide for neutral atoms. Appl. Phys. B 70, 701<br />

(2000).<br />

[For02] J. Fortágh, H. Ott, S. Kraft, A. Günther and C. Zimmermann,<br />

Surface effects in magnetic microtraps. Phys. Rev. A 66, 041604<br />

(2002).<br />

[For03] J. Fortágh, H. Ott, S. Kraft, A. Günther and C. Zimmermann,<br />

Bose-Einstein condensates in magnetic waveguides. Appl. Phys. B<br />

76, 157 (2003).<br />

[Fri33] R. Frisch and E. Segrè, Über die Einstellung der Richtungsquantelung.<br />

II. Z. Phys. 80, 610 (1933).<br />

[Fri51] H. Friedberg and W. Paul, Optische Abbildung mit neutralen<br />

Atomen. Naturwissenschaften 38, 159 (1951).<br />

[Geh05] A.-L. Gehrmann, Evaporative Kühlung in optischen Dipolfallen.<br />

Master’s thesis, Universität Hannover (2005).<br />

[Gha06] S. Ghanbari, T. Kieu, A. Sidorov and P. Hannaford, Permanent<br />

magnetic lattices for ultracold atoms and quantum degenerate gases.<br />

J. Phys. B: At. Mol. Opt. Phys. 39, 1 (2006).<br />

212


BIBLIOGRAPHY<br />

[Gla63] R. Glauber, The quantum theory <strong>of</strong> optical coherence. Phys. Rev.<br />

130, 2529 (1963).<br />

[Gre02a] M. Greiner, O. Mandel, T. Esslinger and T. Hänsch, Quantum phase<br />

transition from a superfluid to a Mott insulator in a gas <strong>of</strong> ultracold<br />

atoms. Nature 415, 39 (2002).<br />

[Gre02b] M. Greiner, O. Mandel, T. Hänsch and I. Bloch, Collapse and revival<br />

<strong>of</strong> the matter wave <strong>of</strong> a Bose-Einstein condensate. Nature 419,<br />

51 (2002).<br />

[Gri98] M. Grifoni and P. Hänggi, Driven quantum tunneling. Phys. Reports<br />

304, 229 (1998).<br />

[Gri00] R. Grimm, M. Weidemüller and Y. Ovchinnikov, Optical dipole<br />

traps for neutral atoms. Adv. At. Mol. Opt. Phys. 42, 95 (2000).<br />

[Gro61] E. Gross, Structure <strong>of</strong> quantized vortex. Nuovo Cimento 20, 454<br />

(1961).<br />

[Gro91] F. Grossmann, T. Dittrich, P. Jung and P. Hänggi, Coherent destruction<br />

<strong>of</strong> tunneling. Phys. Rev. Lett. 67, 516 (1991).<br />

[Gup02] S. Gupta, K. Dieckmann, Z. Hadzibabic and D. Pritchard, Contrast<br />

interferometry using Bose-Einstein condensates to measure<br />

h/m and α. Phys. Rev. Lett. 89, 140401 (2002).<br />

[Gus97] T. Gustavson, P. Bouyer and M. Kasevich, Precision rotation measurements<br />

with an atom interferometer gyroscope. Phys. Rev. Lett.<br />

78, 2046 (1997).<br />

[H¨75] T. Hänsch and A. Schawlow, Cooling <strong>of</strong> gases by laser radiation.<br />

Opt. Comm. 13, 68 (1975).<br />

[H¨01a] W. Hänsel, P. Hommelh<strong>of</strong>f, T. Hänsch and J. Reichel, Bose-Einstein<br />

condensation on a microelectronic chip. Nature 413, 498 (2001).<br />

[H¨01b] W. Hänsel, J. Reichel, P. Hommelh<strong>of</strong>f and T. Hänsch, Magnetic conveyor<br />

belt for transporting and merging trapped atom clouds. Phys.<br />

Rev. Lett. 86, 608 (2001).<br />

[H¨01c] W. Hänsel, J. Reichel, P. Hommelh<strong>of</strong>f and T. Hänsch, Trapped-atom<br />

interferometer in a magnetic microtrap. Phys. Rev. A 64, 063607<br />

(2001).<br />

213


BIBLIOGRAPHY<br />

[Hal05] B. Hall, S. Whitlock, F. Scharnberg, P. Hannaford and A. Sidorov,<br />

Bose-Einstein condensates on a permament magnetic film atom<br />

chip. In E. Hinds, A. Ferguson and E. Riis, (Eds.) Laser Spectroscopy,<br />

Proceedings <strong>of</strong> ICOLS (World Scientific, Singapore, 2005),<br />

275.<br />

[Hal06] B. Hall, S. Whitlock, F. Scharnberg, P. Hannaford and A. Sidorov,<br />

A permanent magnetic film atom chip for Bose-Einstein condensation.<br />

J. Phys. B: At. Mol. Opt. Phys. 39, 27 (2006).<br />

[Hal07a] B. Hall (2007), private communication.<br />

[Hal07b] B. Hall, S. Whitlock, R. Anderson, P. Hannaford and A. Sidorov,<br />

Condensate splitting in an asymmetric double well for atom chip<br />

based sensors. Phys. Rev. Lett. 98, 030402 (2007).<br />

[Har78] R. Harris and L. Stodolsky, Quantum beats in optical activity and<br />

weak interactions. Phys. Lett. B 78, 313 (1978).<br />

[Hei27] W. Heisenberg, Über den anschaulichen Inhalt der quantentheoretischen<br />

Kinematik und Mechanik. Z. Phys. 43, 172 (1927).<br />

[Hes86] H. Hess, Evaporative cooling <strong>of</strong> magnetically trapped and compressed<br />

spin-polarized hydrogen. Phys. Rev. B 34, 3476 (1986).<br />

[Hin01] E. Hinds, C. Vale and M. Boshier, Two wire waveguide and interferometer<br />

for cold atoms. Phys. Rev. Lett. 86, 1462 (2001).<br />

[Hol04] L. Hollenberg, A. Dzurak, C. Wellard, A. Hamilton, D. Reilly,<br />

G. Milburn and R. Clark, Charge-based quantum computing using<br />

single donors in semiconductors. Phys. Rev. B 69, 113301 (2004).<br />

[Hu00] C.-M. Hu and D. Heitmann, Bilayer quantum transistor. Appl.<br />

Phys. Lett. 77, 1475 (2000).<br />

[Hun27] F. Hund, Zur Deutung der Molekelspektren. III. Z. Phys. A 43<br />

(1927).<br />

[Jac96] M. Jack, M. Collett and D. Walls, Coherent quantum tunneling<br />

between two Bose-Einstein condensates. Phys. Rev. A 54, R4625<br />

(1996).<br />

[Jav86] J. Javanainen, Oscillatory exchange <strong>of</strong> atoms between traps containing<br />

Bose condensates. Phys. Rev. Lett. 57, 3164 (1986).<br />

214


BIBLIOGRAPHY<br />

[Jav96] J. Javanainen and S. Yoo, Quantum phase <strong>of</strong> a Bose-Einstein condensate<br />

with an arbitrary number <strong>of</strong> atoms. Phys. Rev. Lett. 76,<br />

161 (1996).<br />

[Jav97] J. Javanainen and M. Wilkens, Phase and phase diffusion <strong>of</strong> a split<br />

Bose-Einstein condensate. Phys. Rev. Lett. 78, 4675 (1997).<br />

[Jav99] J. Javanainen and M. Ivanov, Splitting a trap containing a Bose-<br />

Einstein condensate: Atom number fluctuations. Phys. Rev. A 60,<br />

2351 (1999).<br />

[Jo07] G.-B. Jo, J.-H. Choi, C. Christensen, T. Pasquini, Y.-R. Lee,<br />

W. Ketterle and D. Pritchard, Phase sensitive recombination <strong>of</strong><br />

two Bose-Einstein condensates on an atom chip. Phys. Rev. Lett.<br />

98 (2007).<br />

[Jon03] M. Jones, C. Vale, D. Sahagun, B. Hall and E. Hinds, Spin coupling<br />

between cold atoms and the thermal fluctuations <strong>of</strong> a metal surface.<br />

Phys. Rev. Lett. 91, 080401 (2003).<br />

[Jon04] M. Jones, C. Vale, D. Sahagun, B. Hall, C. Eberlein, B. Sauer,<br />

K. Furusawa, D. Richardson and E. Hinds, Cold atoms probe the<br />

magnetic field near a wire. J. Phys. B: At. Mol. Opt. Phys. 37, L15<br />

(2004).<br />

[Kas91] M. Kasevich and S. Chu, Atomic interferometry using stimulated<br />

Raman transitions. Phys. Rev. Lett. 67, 181 (1991).<br />

[Kas02] M. Kasevich, Coherence with atoms. Science 298, 1363 (2002).<br />

[Ket93] W. Ketterle, K. Davis, M. J<strong>of</strong>fe, A. Martin and D. Pritchard, High<br />

densities <strong>of</strong> cold atoms in a dark spontaneous-force trap. Phys. Rev.<br />

Lett. 70, 2253 (1993).<br />

[Ket99] W. Ketterle, D. Durfee and D. Stamper-Kurn, Making, probing and<br />

understanding Bose-Einstein condensates (Int. School <strong>of</strong> Physics<br />

“Enrico Fermi”, Course CXL, 1999), cond-mat/9904034.<br />

[Kin05] T. Kinoshita, T. Wenger and D. Weiss, All-optical Bose-Einstein<br />

condensation using a compressible crossed dipole trap. Phys. Rev.<br />

A 71, 011602 (2005).<br />

[Kna29] F. Knaur and O. Stern, Über die Reflexion von Molekularstrahlen.<br />

Z. Phys. 53, 779 (1929).<br />

215


BIBLIOGRAPHY<br />

[Kok98] S. Kokkelmans, B. Verhaar and K. Gibble, Prospects for Bose-<br />

Einstein condensation in Cesium. Phys. Rev. Lett. 81, 951 (1998).<br />

[Kra02] S. Kraft, A. Günther, H. Ott, D. Wharam, C. Zimmermann and<br />

J. Fortagh, Anomalous longitudinal magnetic field near the surface<br />

<strong>of</strong> copper conductors. J. Phys. B: At. Mol. Opt. Phys. 35, L469<br />

(2002).<br />

[Kre04] H. Kreutzmann, U. Poulsen, M. Lewenstein, R. Dumke, W. Ertmer,<br />

G. Birkl and A. Sanpera, Coherence properties <strong>of</strong> guided-atom<br />

interferometers. Phys. Rev. Lett. 92, 163201 (2004).<br />

[Kup00] S. Kuppens, K. Corwin, K. Miller, T. Chupp and C. Wieman, Loading<br />

an optical dipole trap. Phys. Rev. A 62, 013406 (2000).<br />

[Lan32] L. Landau, Zur Theorie der Energieübertragung. II. Phys. Z. Sowjetunion<br />

1, 89 (1932).<br />

[Lau99a] D. Lau, R. McLean, A. Sidorov, D. Gough, J. Koperski, W. Rowlands,<br />

B. Sexton, G. Opat and P. Hannaford, Magnetic atom optical<br />

elements for laser-cooled atoms. J. Korean Phys. Soc. 35, 127<br />

(1999).<br />

[Lau99b] D. Lau, A. Sidorov, G. Opat, R. McLean, W. Rowlands and P. Hannaford,<br />

Reflection <strong>of</strong> cold atoms from an array <strong>of</strong> current-carrying<br />

wires. Eur. J. Phys. D 5, 193 (1999).<br />

[Lea02] A. Leanhardt, A. Chikkatur, D. Kielpinski, Y. Shin, T. Gustavson,<br />

W. Ketterle and D. Pritchard, Propagation <strong>of</strong> Bose-Einstein condensates<br />

in a magnetic waveguide. Phys. Rev. Lett. 89, 040401<br />

(2002).<br />

[Lea03] A. Leanhardt, Y. Shin, A. Chikkatur, D. Kielpinski, W. Ketterle<br />

and D. Pritchard, Bose-Einstein condensates near a micr<strong>of</strong>abricated<br />

surface. Phys. Rev. Lett. 90, 100404 (2003).<br />

[Len04] A. Lengwenus, Effiziente Präparation von Atomen zur Erzeugung<br />

von Bose-Einstein-Kondensaten in optischen Dipolpotentialen.<br />

Master’s thesis, Universität Hannover (2004).<br />

[Let88] P. Lett, R. Watts, C. Westbrook, W. Phillips, P. Gould and H. Metcalf,<br />

Observation <strong>of</strong> atoms laser-cooled below the Doppler limit.<br />

Phys. Rev. Lett. 61, 169 (1988).<br />

[Lev03a] B. Lev, Fabrication <strong>of</strong> micro-magnetic traps for cold neutral atoms.<br />

Published in Quantum Information and Computation 3, 450 (2003).<br />

216


BIBLIOGRAPHY<br />

[Lev03b] B. Lev, Y. Lassailly, C. Lee, A. Scherer and H. Mabuchi, An atom<br />

mirror etched from a hard drive. Appl. Phys. Lett. 83, 395 (2003).<br />

[M¨01] D. Müller, E. Cornell, M. Prevedelli, P. Schwindt, Y.-J. Wang and<br />

D. Anderson, Magnetic switch for integrated atom optics. Phys. Rev.<br />

A 63, 041602 (2001).<br />

[M¨05] T. Müther, Evaporative Kühlung in optischen Dipolpotentialen.<br />

Ph.D. thesis, Universität Hannover (2005).<br />

[Mah02] K. Mahmud, J. Kutz and W. Reinhardt, Bose-Einstein condensates<br />

in a one-dimensional double square well: Analytical solutions <strong>of</strong> the<br />

nonlinear Schrödinger equation. Phys. Rev. A 66, 063607 (2002).<br />

[Mar02] A. Marte, T. Volz, J. Schuster, S. Dürr, G. Rempe, E. van Kampen<br />

and B. Verhaar, Feshbach resonances in Rubidium 87: Precision<br />

measurement and analysis. Phys. Rev. Lett. 89, 283202 (2002).<br />

[Mas88] N. Masuhara, J. Doyle, J. Sandberg, D. Kleppner, T. Greytak,<br />

H. Hess and G. Kochanski, Evaporative cooling <strong>of</strong> spin-polarized<br />

atomic hydrogen. Phys. Rev. Lett. 61, 935 (1988).<br />

[Mel03] A. Mellish, G. Duffy, G. McKenzie, R. Geursen and A. Wilson,<br />

Nonadiabatic loading <strong>of</strong> a Bose-Einstein condensate into the ground<br />

state <strong>of</strong> an optical lattice. Phys. Rev. A 68, 051601 (2003).<br />

[Men01] C. Menotti, J. Anglin, J. Cirac and P. Zoller, Dynamic splitting <strong>of</strong><br />

a Bose-Einstein condensate. Phys. Rev. A 63, 023601 (2001).<br />

[Mes90] A. Messiah, Quantenmechanik 2 (Walter de Gruyter, Berlin, New<br />

York, 1990), third ed.<br />

[Met99] H. Metcalf and P. van der Straten, Laser cooling and trapping<br />

(Springer Verlag, New York, 1999), second ed.<br />

[Mey91] P. Meystre and M. Sargent III, Elements <strong>of</strong> Quantum Optics<br />

(Springer Verlag, Berlin, Heidelberg, 1991), second ed.<br />

[Mig85] A. Migdall, J. Prodan, W. Phillips, T. Bergeman and H. Metcalf,<br />

First observation <strong>of</strong> magnetically trapped neutral atoms. Phys. Rev.<br />

Lett. 54, 2596 (1985).<br />

[Mil97] G. Milburn, J. Corney, E. Wright and D. Walls, Quantum dynamics<br />

<strong>of</strong> an atomic Bose-Eeinstein condensate in a double-well potential.<br />

Phys. Rev. A 55, 4318 (1997).<br />

217


BIBLIOGRAPHY<br />

[Mom03] J. Mompart, K. Eckert, W. Ertmer, G. Birkl and M. Lewenstein,<br />

Quantum computing with spatially delocalized qubits. Phys. Rev.<br />

Lett. 90, 147901 (2003).<br />

[Nak02] H. Nakamura, Nonadiabatic Transition: Concepts, Basic Theories<br />

and Applications (World Scientific, 2002).<br />

[Neg04] A. Negretti and C. Henkel, Enhanced phase sensitivity and soliton<br />

formation in an integrated BEC interferometer. J. Phys. B: At. Mol.<br />

Opt. Phys. 37, L1 (2004).<br />

[Neu78] W. Neuhauser, M. Hohenstatt, P. Toschek and H. Dehmelt, Opticalsideband<br />

cooling <strong>of</strong> visible atom cloud confined in parabolic well.<br />

Phys. Rev. Lett. 41, 233 (1978).<br />

[Nol94] W. Nolting, Grundkurs: Theoretische Physik 5, Quantenmechanik<br />

2 (Verlag Zimmermann- Neufang, Ulmen, 1994), second ed.<br />

[O’H01] K. O’Hara, M. Gehm, S. Granade and J. Thomas, Scaling laws for<br />

evaporative cooling in time-dependent optical traps. Phys. Rev. A<br />

64, 051403 (2001).<br />

[Ott01] H. Ott, J. Fortágh, G. Schlotterbeck, A. Grossmann and C. Zimmermann,<br />

Bose-Einstein condensation in a surface microtrap. Phys.<br />

Rev. Lett. 87, 230401 (2001).<br />

[Pea02] C. Pearman, C. Adams, S. Cox, P. Griffin, D. Smith and I. Hughes,<br />

Polarization spectroscopy <strong>of</strong> a closed atomic transition: applications<br />

to laser frequency locking. J. Phys. B: At. Mol. Opt. Phys. 35, 5141<br />

(2002).<br />

[Pet99] Y. Peters, K. Chung and S. Chu, Measurement <strong>of</strong> gravitational acceleration<br />

by dropping atoms. Nature 400, 849 (1999).<br />

[Pet02] C. Pethick and H. Smith, Bose-Einstein Condensation in Dilute<br />

Gases (Cambridge <strong>University</strong> Press, Cambridge, 2002), first ed.<br />

[Pet03] T. Petelski, M. Fattori, G. Lamporesi, J. Stuhler and G. Tino,<br />

Doppler-free spectroscopy using magnetically induced dichroism <strong>of</strong><br />

atomic vapor: a new scheme for laser frequency locking. Eur. J.<br />

Phys. D 22, 279 (2003).<br />

[Pit61] L. Pitaevskii, Vortex lines in an imperfect Bose gas. Sov. Phys.<br />

JETP 13, 451 (1961).<br />

218


BIBLIOGRAPHY<br />

[Pou02] U. Poulsen and K. Mølmer, A quantum beam splitter for atoms.<br />

Phys. Rev. A 65, 033613 (2002).<br />

[Pri83] D. Pritchard, Cooling neutral atoms in a magnetic trap for precision<br />

spectroscopy. Phys. Rev. Lett. 51, 1336 (1983).<br />

[Pul03] D. Pulido, Instability in a cold atom interferometer. Master’s thesis,<br />

Worcester Polytechnic Institute (2003).<br />

[R¨97] A. Röhrl, M. Naraschewski, A. Schenzle and H. Wallis, Transition<br />

from phase locking to the interference <strong>of</strong> independent Bose condensates:<br />

theory vs experiment. Phys. Rev. Lett. 78, 4143 (1997).<br />

[Raa87] E. Raab, M. Prentiss, A. Cable, S. Chu and D. Pritchard, Trapping<br />

<strong>of</strong> neutral sodium atoms with radiation pressure. Phys. Rev. Lett.<br />

59, 2631 (1987).<br />

[Rap01] U. Rapol, A. Wasan and V. Natarajan, Loading <strong>of</strong> a Rb magnetooptic<br />

trap from a getter source. Phys. Rev. A 64, 023402 (2001).<br />

[Rei99] J. Reichel, W. Hänsel and T. Hänsch, Atomic micromanipulation<br />

with magnetic surface traps. Phys. Rev. Lett. 83, 3398 (1999).<br />

[Rei02] J. Reichel, Microchip traps and Bose-Einstein condensation. Appl.<br />

Phys. B 75, 469 (2002).<br />

[Roa95] T. Roach, H. Abele, M. Boshier, H. Grossman, K. Zetie and<br />

E. Hinds, Realization <strong>of</strong> a magnetic mirror for cold atoms. Phys.<br />

Rev. Lett. 75, 629 (1995).<br />

[Rob82] R. Robiscoe, Perturbative solution to the time-dependent two-level<br />

problem and the validity <strong>of</strong> the Rosen-Zener conjecture. Phys. Rev.<br />

A 27, 1365 (1982).<br />

[Ros00] P. Rosenbusch, B. Hall, I. Hughes, C. Saba and E. Hinds, Manipulation<br />

<strong>of</strong> cold atoms using a corrugated magnetic reflector. Phys.<br />

Rev. A 61, 031404 (2000).<br />

[Row96] W. Rowlands, D. Lau, G. Opat, R. McLean and P. Hannaford, Magnetostatic<br />

state-selective deflection <strong>of</strong> a beam <strong>of</strong> laser-cooled atoms.<br />

Opt. Comm. 126, 55 (1996).<br />

[Rus98] F. Ruschewitz, J. Peng, H. Hinterthür, N. Schaffrath, K. Sengstock<br />

and W. Ertmer, Sub-kilohertz optical spectroscopy with a time domain<br />

atom interferometer. Phys. Rev. Lett. 80, 3173 (1998).<br />

219


BIBLIOGRAPHY<br />

[Ryc04] D. Rychtarik, B. Engeser, H.-C. Nägerl and R. Grimm, Twodimensional<br />

Bose-Einstein condensation in an optical surface trap.<br />

Phys. Rev. Lett. 92, 173003 (2004).<br />

[Sak02] E. Sakellari, M. Leadbeater, N. Kylstra and C. Adams, Josephson<br />

spectroscopy <strong>of</strong> a dilute Bose-Einstein condensate in a double-well<br />

potential. Phys. Rev. A 66, 033612 (2002).<br />

[Sak04] E. Sakellari, N. Proukakis, M. Leadbeater and C. Adams, Josephson<br />

tunneling <strong>of</strong> a phase-imprinted Bose-Einstein condensate in a timedependent<br />

double-well potential. New J. Physics 6, 42 (2004).<br />

[Sch65] L. Schiff, Quantum Mechanics (McGraw-Hill, Novaro, Mexico,<br />

1965), second ed.<br />

[Sch03] S. Schneider, A. Kasper, C. vom Hagen, M. Bartenstein, B. Engeser,<br />

T. Schumm, I. Bar-Joseph, R. Folman, L. Feenstra and J. Schmiedmayer,<br />

Bose-Einstein condensation in a simple microtrap. Phys.<br />

Rev. A 67, 023612 (2003).<br />

[Sch05a] T. Schumm, J. Estève, C. Figl, J.-B. Trebbia, C. Aussibal,<br />

H. Nguyen, D. Mailly, I. Bouchoule, C. Westbrook and A. Aspect,<br />

Atom chips in the real world: the effects <strong>of</strong> wire corrugation. Eur.<br />

J. Phys. D 32, 171 (2005).<br />

[Sch05b] T. Schumm, S. H<strong>of</strong>ferberth, L. Andersson, S. Wildermuth, S. Groth,<br />

I. Bar-Joseph, J. Schmiedmayer and P. Krüger, Matter-wave interferometry<br />

in a double well on an atom chip. Nature Physics 1, 57<br />

(2005).<br />

[Scu91] M. Scully, B.-G. Englert and H. Walther, Quantum optical tests <strong>of</strong><br />

complementarity. Nature 351, 111 (1991).<br />

[Set81] J. Sethna, Phonon coupling in tunneling systems at zero temperature:<br />

An instanton approach. Phys. Rev. B 24, 698 (1981).<br />

[Shi04] Y. Shin, Q. Saba, T. Pasquini, W. Ketterle, D. Pritchard and<br />

A. Leanhardt, Atom interferometry with Bose-Einstein condensates<br />

in a double-well potential. Phys. Rev. Lett. 92, 050405 (2004).<br />

[Shi05] Y. Shin, C. Sanner, G.-B. Jo, T. Pasquini, M. Saba, W. Ketterle,<br />

D. Pritchard, M. Vengalattore and M. Prentiss, Interference <strong>of</strong><br />

Bose-Einstein condensates split with an atom chip. Phys. Rev. A<br />

72, 021604 (2005).<br />

220


BIBLIOGRAPHY<br />

[Sid96] A. Sidorov, R. McLean, W. Rowlands, D. Lau, J. Murphy,<br />

M. Walkiewicz, G. Opat and P. Hannaford, Specular reflection <strong>of</strong><br />

cold caesium atoms from a magnetostatic mirror. Quantum Semiclass.<br />

Opt. 8, 713 (1996).<br />

[Sid02a] A. Sidorov (2002), private communication.<br />

[Sid02b] A. Sidorov, R. McLean, F. Scharnberg, D. Gough, T. Davis, B. Sexton,<br />

G. Opat and P. Hannaford, Permanent-magnet microstructures<br />

for atom optics. Act. Phys. Pol. B 33, 2137 (2002).<br />

[Sid06] A. Sidorov, B. Dalton, S. Whitlock and F. Scharnberg, The asymmetric<br />

double-well potential for single-atom interferometry. Phys.<br />

Rev. A 74, 023612 (2006).<br />

[Sim85] B. Simon, Semiclassical analysis <strong>of</strong> low lying eigenvalues. iv. The<br />

flea on the elephant. J. Funct. Anal. 63, 123 (1985).<br />

[Sin05] C. Sinclair, E. Curtis, J. Retter, B. Hall, I. Llaurente Garica,<br />

S. Eriksson, B. Sauer and E. Hinds, Preparation <strong>of</strong> a Bose-Einstein<br />

condensate on a permanent-magnet atom chip. In Conference on<br />

Atoms and Molecules Near Surfaces 19 (2005), Conference series,<br />

74–77.<br />

[Sme97] A. Smerzi, S. Fantoni, S. Giovanazzi and S. Shenoy, Quantum coherent<br />

atomic tunneling between two trapped Bose-Einstein condensates.<br />

Phys. Rev. Lett. 79, 4950 (1997).<br />

[Spe99] R. Spekkens and J. Sipe, Spatial fragmentation <strong>of</strong> a Bose-Einstein<br />

condensate in a double well potential. Phys. Rev. A 59, 3868 (1999).<br />

[Ste03] D. Steck, Rubidium 87 D Line Data (2003),<br />

http://steck.us/alkalidata.<br />

[Sti02a] J. Stickney and A. Zozulya, Expansion <strong>of</strong> a Bose-Einstein condensate<br />

from a microtrap into a waveguide. Phys. Rev. A 65, 053612<br />

(2002).<br />

[Sti02b] J. Stickney and A. Zozulya, Wave-function recombination instability<br />

in cold-atom interferometers. Phys. Rev. A 66, 053601 (2002).<br />

[Stu32] E. Stueckelberg, Theory <strong>of</strong> inelastic collisions between atoms. Helv.<br />

Phys. Acta 5, 369 (1932).<br />

[Stu03] J. Stuhler, M. Fattori, T. Petelski and G. Tino, MAGIA - using<br />

atom interferometry to determine the Newtonian gravitational constant.<br />

J. Opt. B: Quant. Semiclass. Opt. 5, 75 (2003).<br />

221


BIBLIOGRAPHY<br />

[Tab91] J. Tabosa, G. Chen, Z. Hu, R. Lee and H. Kimble, Nonlinear spectroscopy<br />

<strong>of</strong> cold atoms in a spontaneous-force optical trap. Phys.<br />

Rev. Lett. 66, 3245 (1991).<br />

[Tak03] Y. Takaso, K. Maki, K. Komori, T. Takano, K. H. andM. Kumakura,<br />

T. Yabuzaki and Y. Takahashi, Spin-singlet Bose-Einstein<br />

condensation <strong>of</strong> two-electron atoms. Phys. Rev. Lett. 91, 040404<br />

(2003).<br />

[The04] M. Theis, G. Thalhammer, K. Winkler, M. Hellwig, G. Ruff,<br />

R. Grimm and J. Hecker Denschlag, Tuning the scattering length<br />

with an optically induced Feshbach resonance. Phys. Rev. Lett. 93,<br />

123001 (2004).<br />

[Tho02] N. Thomas, C. Foot and A. Wilson, A double-well magnetic trap<br />

for Bose-Einstein condensates. Phys. Rev. A 65, 063406 (2002).<br />

[Thy99] J. Thywissen, M. Olshanii, G. Zabow, M. Drndić, K. Johnson,<br />

R. Westervelt and M. Prentiss, micr<strong>of</strong>abricated magnetic waveguides<br />

for neutral atoms. Eur. J. Phys. D 7, 361 (1999).<br />

[Tie02] T. Tiecke, M. Kemmann, C. Buggle, I. Shvarchuck, W. von Klitzing<br />

and J. Walraven. J. Opt. B: Quant. Semiclass. Opt. 5, S119 (2002).<br />

[Tow96] C. Townsend, N. Edwards, K. Zetie, C. Cooper, J. Rink and<br />

C. Foot, High-density trapping <strong>of</strong> cesium atoms in a dark magnetooptical<br />

trap. Phys. Rev. A 53, 1702 (1996).<br />

[Val04] C. Vale, B. Upcr<strong>of</strong>t, M. Davis, N. Heckenberg and H. Rubinzstein-<br />

Dunlop, Foil based atom chip for Bose-Einstein condensates. J.<br />

Phys. B: At. Mol. Opt. Phys. 37, 2959 (2004).<br />

[Ven02] M. Vengalattore, W. Rooijakkers and M. Prentiss, Ferromagnetic<br />

atom guide with in situ loading. Phys. Rev. A 66, 053403 (2002).<br />

[Vit97] N. Vitanov and K.-A. Suominen, Time-dependent control <strong>of</strong> ultracold<br />

atoms in magnetic traps. Phys. Rev. A 56, R4377 (1997).<br />

[Vol05] M. Volk, Kalte Atome für die Quanteninformationsverarbeitung.<br />

Ph.D. thesis, Universität Hannover (2005).<br />

[Wan04] D.-W. Wang, M. Lukin and E. Demler, Disordered Bose-Einstein<br />

condensates in quasi one-dimensional magnetic microtraps. Phys.<br />

Rev. Lett. 92, 076802 (2004).<br />

222


BIBLIOGRAPHY<br />

[Wan05a] J. Wang, S. Whitlock, F. Scharnberg, D. Gough, A. Sidorov,<br />

R. McLean and P. Hannaford, Perpendicularly magnetized, grooved<br />

GdTbFeCo microstructures for atom optics. J. Phys. D: Appl. Phys.<br />

38, 4015 (2005).<br />

[Wan05b] Y.-J. Wang, D. Anderson, V. Bright, E. Cornell, Q. Diot, T. Kishimoto,<br />

M. Prentiss, R. Saravan, S. Segal and S. Wu, Atom Michelson<br />

interferometer on a chip using a Bose-Einstein condensate. Phys.<br />

Rev. Lett. 94, 090405 (2005).<br />

[Web03] T. Weber, J. Herbig, M. Mark, H.-C. Nägerl and R. Grimm, Bose-<br />

Einstein condensation <strong>of</strong> cesium. Science 299, 232 (2003).<br />

[Wei87] U. Weiss, H. Grabert, P. Hänggi and P. Riseborough, Incoherent<br />

tunneling in a double well. Phys. Rev. B 35, 9535 (1987).<br />

[Wei95] J. Weinstein and K. Libbrecht, Microscopic magnetic traps for neutral<br />

atoms. Phys. Rev. A 52, 4004 (1995).<br />

[Whi04] S. Whitlock (2004), private communication.<br />

[Whi05] S. Whitlock (2005), private communication.<br />

[Whi07] S. Whitlock, B. Hall, T. Roach, R. Anderson, M. Volk, P. Hannaford<br />

and A. Sidorov, Effect <strong>of</strong> magnetization inhomogenity on magnetic<br />

microtraps for atoms. Phys. Rev. A 75, 043602 (2007).<br />

[Wil02] G. Wilpers, T. Binnewies, C. Degenhardt, U. Sterr, J. Helmcke and<br />

F. Riehle, Optical clock with ultracold neutral atoms. Phys. Rev.<br />

Lett. 89, 230801 (2002).<br />

[Wil04] S. Wildermuth, P. Krüger, C. Becker, M. Brajdic, A. Kasper, R. Folman<br />

and J. Schmiedmayer, Optimized U-MOT for experiments with<br />

ultracold atoms near surfaces. Phys. Rev. A 69, 030901 (2004).<br />

[Win75] D. Wineland and H. Dehmelt, Proposed 10 14 δν/ν laser fluorescence<br />

spectroscopy on Tl + mono-ion oscillator. Bull. Am. Phys. Soc. 20,<br />

637 (1975).<br />

[Win79] D. Wineland and W. Itano, Laser cooling <strong>of</strong> atoms. Phys. Rev. A<br />

20, 1521 (1979).<br />

[Win84] W. Wing, On neutral particle trapping in quasistatic electromagnetic<br />

fields. Prog. Quant. Elect. 8, 181 (1984).<br />

223


BIBLIOGRAPHY<br />

[Xin04] Y. Xing, A. Eljaouhari, I. Barb, R. Gerritsma, R. Spreeuw and<br />

J. Goedkoop, Hard magnetic FePt films for atom chips. Phys. Stat.<br />

Sol. (c) 1, 3702 (2004).<br />

[Zen32] C. Zener, Non-adiabatic crossing <strong>of</strong> energy levels. Proc. R. Soc.<br />

London, Ser. A 137, 696 (1932).<br />

[Zim04] C. Zimmermann, S. Kraft, A. Günther, H. Ott and J. Fortágh,<br />

Two experiments with Bose-Einstein condensates in magnetic micro<br />

traps. J. Phys. IV France 116, 275 (2004).<br />

224

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!