15.02.2014 Views

Evanescent-Wave Mirrors for Cold Atoms - Faculteit der ...

Evanescent-Wave Mirrors for Cold Atoms - Faculteit der ...

Evanescent-Wave Mirrors for Cold Atoms - Faculteit der ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Evanescent</strong>-<strong>Wave</strong> <strong>Mirrors</strong><br />

<strong>for</strong> <strong>Cold</strong> <strong>Atoms</strong><br />

Dirk Voigt


This thesis studies elastic and inelastic mirrors <strong>for</strong> atoms,<br />

based on the repelling optical potential from an evanescent<br />

wave in the vicinity of a surface. Light scattering by<br />

cold (10 K) rubidium atoms bouncing at normal incidence<br />

on such mirrors was investigated experimentally.<br />

It was observed as sideward radiation pressure exerted<br />

on elastically bouncing atoms. Inelastic bounces as a<br />

consequence of optical hyperfine pumping were also<br />

observed.<br />

The inelastic process could be used as a<br />

loading mechanism <strong>for</strong> low-dimensional optical traps,<br />

possibly leading to a low-dimensional quantum gas. New<br />

prospects to reduce light scattering of trapped atoms,<br />

using dark states in circularly-polarised evanescentwaves,<br />

are discussed.<br />

ISBN 90-6464-438-1


<strong>Evanescent</strong>-<strong>Wave</strong> <strong>Mirrors</strong><br />

<strong>for</strong><br />

<strong>Cold</strong> <strong>Atoms</strong>


<strong>Evanescent</strong>-<strong>Wave</strong> <strong>Mirrors</strong><br />

<strong>for</strong><br />

<strong>Cold</strong> <strong>Atoms</strong><br />

ACADEMISCH PROEFSCHRIFT<br />

ter verkrijging van de graad van doctor<br />

aan de Universiteit van Amsterdam,<br />

op gezag van de Rector Magnificus<br />

prof. dr. J.J.M. Franse<br />

ten overstaan van een door het college voor promoties ingestelde<br />

commissie, in het openbaar te verdedigen in de Aula <strong>der</strong> Universiteit<br />

op maandag 18 december 2000 te 15:00 uur<br />

door<br />

Dirk Voigt<br />

geboren te Tübingen


Promotor:<br />

Co-promotor:<br />

Commissie:<br />

prof. dr. H.B. van Linden van den Heuvell<br />

dr. R.J.C. Spreeuw<br />

dr. T.W. Hijmans<br />

prof. dr. W. Hogervorst<br />

prof. dr. J.A. Schouten<br />

prof. dr. G.V. Shlyapnikov<br />

prof. dr. P. van <strong>der</strong> Straten<br />

prof. dr. J.T.M. Walraven<br />

<strong>Faculteit</strong> <strong>der</strong> Natuurwetenschappen, Wiskunde en In<strong>for</strong>matica<br />

The work described in this thesis was part of the research program of the<br />

“Stichting voor Fundamenteel On<strong>der</strong>zoek <strong>der</strong> Materie” (FOM),<br />

which is financially supported by the<br />

“Ne<strong>der</strong>landse Organisatie voor Wetenschappelijk On<strong>der</strong>zoek” (NWO),<br />

and was carried out at the<br />

Van <strong>der</strong> Waals-Zeeman Instituut<br />

Universiteit van Amsterdam<br />

Valckenierstraat 65<br />

1018 XE Amsterdam<br />

A limited number of copies of this thesis is available at this address.<br />

ISBN 90-6464-438-1


für meine Eltern<br />

Sieglinde und Edgar


Contents<br />

1 General introduction 9<br />

1 .1 Atom optics and laser cooling . . . . . . . . . . . . . . . . . . . . . . 9<br />

1 .2 Quantum gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 0<br />

1<br />

1<br />

2 A low-dimensional quantum gas by means of dark states 13<br />

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

2.2 <strong>Evanescent</strong>-wave mirrors . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

2.3 Generic trap loading scheme . . . . . . . . . . . . . . . . . . . . . . . 8<br />

2.4 Loading a low dimensional trap . . . . . . . . . . . . . . . . . . . . . 20<br />

2.5 Photon scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

2.6 Circularly-polarised evanescent waves . . . . . . . . . . . . . . . . . . 26<br />

2.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

3 Experimental setup 31<br />

3.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

3.2 Atomic species — rubidium . . . . . . . . . . . . . . . . . . . . . . . 32<br />

3.3 Ultra-high vacuum system . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

3.4 Optical access to the UHV system . . . . . . . . . . . . . . . . . . . . 37<br />

3.5 Semiconductor lasers <strong>for</strong> cooling and trapping . . . . . . . . . . . . . 43<br />

3.6 Real-time experimental control . . . . . . . . . . . . . . . . . . . . . 55<br />

3.7 The magneto-optical trap . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

4A high-power tapered semiconductor amplifier system 69<br />

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

4.2 Amplifier setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

4.3 Unseeded operation of the amplifier . . . . . . . . . . . . . . . . . . . 74<br />

4.4 Amplification of a seed beam . . . . . . . . . . . . . . . . . . . . . . . 75<br />

4.5 Spatial and spectral filtering using an optical fibre . . . . . . . . . . . 77<br />

4.6 Variations of individual gain elements . . . . . . . . . . . . . . . . . . 79<br />

4.7 Far off-resonance dipole potentials with spectral background . . . . . 80<br />

4.8 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

5 The evanescent-wave atom mirror 83<br />

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84<br />

5.2 Fraction of bouncing atoms . . . . . . . . . . . . . . . . . . . . . . . 84<br />

5.3 Time-of-flight detection of bouncing atoms . . . . . . . . . . . . . . . 88<br />

5.4 Investigation of bouncing atoms . . . . . . . . . . . . . . . . . . . . . 91<br />

5.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96<br />

7


8 CONTENTS<br />

1<br />

1<br />

6 Radiation pressure exerted by evanescent waves 97<br />

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

6.2 Photon scattering by bouncing atoms . . . . . . . . . . . . . . . . . . 98<br />

6.3 Observation of bouncing atoms . . . . . . . . . . . . . . . . . . . . . 99<br />

6.4 The observation of radiation pressure . . . . . . . . . . . . . . . . . . 02<br />

6.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 08<br />

1 1<br />

1 1 1<br />

1 1<br />

1 1<br />

1 1<br />

7 Inelastic evanescent-wave mirrors 109<br />

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0<br />

7.2 Principle of inelastic evanescent-wave mirrors . . . . . . . . . . . . . .<br />

7.3 Configuration of the inelastic mirror . . . . . . . . . . . . . . . . . . . 2<br />

7.4 Observation of inelastically bouncing atoms . . . . . . . . . . . . . . 4<br />

7.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

A Appendix 119<br />

A.1Useful atom-optical numbers <strong>for</strong> 87 Rb . . . . . . . . . . . . . . . . . . 1 1 9<br />

A.2 Fresnel coefficients <strong>for</strong> evanescent waves . . . . . . . . . . . . . . . . . 1 20<br />

A.3 Light <strong>for</strong>ces and scattering rate . . . . . . . . . . . . . . . . . . . . . 1 21<br />

A.4 Analysis of absorption images . . . . . . . . . . . . . . . . . . . . . . 1 26<br />

References 127<br />

Summary 137<br />

Samenvatting / Zusammenfassung 139<br />

Nawoord 149<br />

Curriculum Vitae / Publications 151


1<br />

General introduction<br />

1.1 Atom optics and laser cooling<br />

The motion of neutral atoms can be manipulated by laser light. Similar to optical<br />

elements <strong>for</strong> light, “atom optical” mirrors, lenses, gratings or beamsplitters can be<br />

realised <strong>for</strong> atomic matter waves [1].<br />

Light <strong>for</strong>ces on atoms are commonly classified as “dipole <strong>for</strong>ces” and “spontaneous<br />

<strong>for</strong>ces” [2]. The dipole <strong>for</strong>ce results from the interaction of the electromagnetic<br />

field with the induced electric dipole of an atom and can be described by an optical<br />

potential, also called “light-shift” potential. This corresponds to the refractive index<br />

of a dielectric medium in optics. The spontaneous <strong>for</strong>ce is based on the absorption<br />

of photons from a preferential direction, such that the related photon recoils represent<br />

the <strong>for</strong>ce. Because of the involved spontaneous emission processes this <strong>for</strong>ce is<br />

dissipative.<br />

Of particular interest are evanescent waves, which can <strong>for</strong>m a mirror <strong>for</strong> atoms<br />

by the induced repelling optical potential. Such a mirror was proposed by Cook and<br />

Hill [3] and was first demonstrated with an atomic beam at grazing incidence by<br />

Balykin et al. [4], and with cold atoms at normal incidence by Kasevich et al. [5].<br />

An evanescent wave occurs when light un<strong>der</strong>goes total internal reflection, e.g. at a<br />

glass surface in vacuum [6,7]. On the dark side, the electric field decays exponentially<br />

away from the surface. The property that makes such a wave particularly useful<br />

<strong>for</strong> atom optics, as compared to freely propagating beams, is the characteristic<br />

decay length which can be less than the optical wavelength. <strong>Atoms</strong> are quickly<br />

(∼ µs) reflected by a steep potential. Of course, compared to the typical refractive<br />

index step <strong>for</strong>, say, a (partially) light transmitting window, an evanescent-wave<br />

potential is relatively smooth: the atomic De Broglie wavelength ranges usually<br />

between 10 − 100 nm.<br />

This thesis investigates the light scattering by cold (10 µK) rubidium atoms<br />

( 87 Rb) in the optical potential of an evanescent-wave mirror. The atoms bounce at<br />

normal incidence either elastically or inelastically from such a mirror. In case of an<br />

elastic mirror, we observed radiation pressure that is exerted on the atoms by the<br />

evanescent wave. Since the evanescent wave propagates only along the glass surface,<br />

also the radiation pressure is expected to be directed parallel to the surface (see<br />

chapter 6). The inelastic mirror is the consequence of optical hyperfine pumping<br />

9


10 General introduction<br />

of bouncing atoms (see chapter 7). Such a dissipative atom-optical element has<br />

no analogy in optics.<br />

A striking consequence of dissipation is that it allows to cool atoms with laser<br />

light. This was proposed in 1975 by Hänsch and Schawlow [8], and by Wineland<br />

and Dehmelt [9]. When cooling atoms to a few µK, they are also sufficiently slow<br />

(a few cm/s) to be bound in optical traps [10]. In the past years vast progress in<br />

laser cooling and trapping has taken place, see the reviews [11,12] or the 1997 Nobel<br />

prize lectures of Chu, Cohen-Tannoudji and Phillips [13–15].<br />

Due to dissipation, the inelastic evanescent-wave mirrors investigated in this<br />

thesis are closely related to reflection cooling of atoms by evanescent waves, which<br />

was demonstrated by Laryushin et al. [16] and used by Ovchinnikov et al. [17] to<br />

cool atoms in a gravito-optical surface trap.<br />

Various fields in fundamental and applied physics take benefit of the achievements<br />

in atom optics and laser cooling. Examples are atom-interferometric measurements<br />

of the gravitational and finestructure constant, spectroscopy of atoms<br />

and molecules, cavity quantum electrodynamics (CQED), and ultracold atomic collisions<br />

[18–26]. <strong>Cold</strong> atoms are also used to model more complex systems, such as<br />

Bloch states in solids [27, 28]. Prospected technical applications are atomic clocks,<br />

e.g. <strong>for</strong> navigational systems, atom lithography on the nm-scale, or quantum computational<br />

devices [29–32].<br />

1.2 Quantum gases<br />

The mentioned applications of cold atoms commonly employ incoherent, thermal<br />

sources of atoms. This might be an atomic beam evading from an oven and collimated<br />

by diaphragms, or a cloud of cold atoms from a magneto-optical trap<br />

(MOT) [33]. Lasers [34] have triggered and improved many optical techniques since<br />

the first demonstration by Maiman [35] in 1960. Similarly, an “atom laser” as a<br />

bright source of coherent matter waves may have a consi<strong>der</strong>able impact on developments<br />

in atom optics.<br />

An atom laser requires the creation of “quantum gas” or, a “quantum degenerate”<br />

gas. Quantum degeneracy requires that an ensemble of atoms is sufficiently<br />

dense and cold, such that the atomic wavefunctions overlap, i.e. the atomic separation<br />

is less than the De Broglie wavelength Λ ∝ √ 1/T ,whereT is the temperature.<br />

The “phase-space” is the combined space of position and momentum coordinates.<br />

The phase-space density of a degenerate gas is thus Φ = nΛ 3 1,wheren is the<br />

spatial density. Equivalently, Φ describes the occupation number of a volume h 3<br />

per available quantum state in the phase-space of a system [36], where h is Planck’s<br />

constant. For bosonic atoms (with integer spin quantum number), degeneracy leads<br />

to the <strong>for</strong>mation of a Bose-Einstein condensate (BEC), in which a single quantum<br />

state is macroscopically occupied by indistinguishable atoms.<br />

BEC was predicted already in 1924 [36,37]. It is un<strong>der</strong>stood as a consequence of<br />

the bosonic quantum statistics that is employed instead of the classical Boltzmann<br />

statistics in the case of a degenerate gas. Ef<strong>for</strong>ts to achieve BEC in dilute, weakly


1.2 Quantum gases 11<br />

BEC outcoupling<br />

Atom Laser<br />

optically driven<br />

LASER<br />

thermal atoms<br />

gain medium<br />

stimulated emission of bosons into the same lasing mode<br />

coherent matter wave<br />

coherent light field<br />

stimulated collisions<br />

stimulated optical transition<br />

evaporation, thermalisation<br />

collisions not required<br />

critical temperature<br />

pumping threshold<br />

3D trap 2D or 1D trap optical cavity<br />

2D <strong>for</strong> hydrogen<br />

due to reabsorption<br />

thermal equilibrium<br />

open, driven non-equilibrium system<br />

ground state occupied excited state possible higher or<strong>der</strong> cavity modes<br />

demonstrated [49–52] proposed [58–60, 84, 86] commonly used [34, 35]<br />

Table 1.1: Atom laser and optical laser.<br />

interacting gases first focused on atomic hydrogen, see e.g. Walraven in Ref. [38].<br />

Accompanied by the development of laser (pre)cooling techniques, BEC was however<br />

first observed with alkali atoms in 1995 by An<strong>der</strong>son et al., Daviset al., andBradley<br />

et al. [39–41]. In 1998, Fried et al. [42] achieved BEC in hydrogen, and Safonov et<br />

al. [43] realised a two-dimensional degenerate hydrogen gas. Recently, DeMarco<br />

and Jin [44] observed also a degenerate fermionic gas. For reviews on BEC, see<br />

Refs. [45–48].<br />

By outcoupling atoms from a BEC, atom lasers have been demonstrated by<br />

Mewes et al., An<strong>der</strong>son et al., Blochet al., and Hagley et al. [49–52]. However, the<br />

only method to achieve BEC so far, is evaporative cooling [53–55] of thermal atoms<br />

down to the critical temperature of condensation into the ground state of a magnetic<br />

trap. Compared to laser cooling, evaporation is a slow process lasting tens of seconds,<br />

since it relies on thermalising collisions. The mentioned atom lasers deplete the<br />

condensate. Their output is there<strong>for</strong>e pulsed, or at most quasicontinuous. As a<br />

remedy it was recently proposed to feed a condensate continuously with atoms [56],<br />

or to cool an atomic beam by evaporation [57].<br />

It remains a challenging goal to achieve quantum degeneracy by purely optical<br />

means, possibly leading to cw atom lasers as open, driven systems out of thermal<br />

equilibrium. Various schemes have been proposed by Wiseman and Collett, Spreeuw<br />

et al., andOlshaniiet al. [58–60]. Commonly, they employ a close analogy to the<br />

optical laser, see Table 1.1. The coherent matter wave takes the role of the lasing<br />

field, and the cavity is realised by an optical trap. The dissipative process involved<br />

is a spontaneous Raman transition, that pumps thermal atoms in a single step from


12 General introduction<br />

a reservoir into the trap. At the onset of degeneracy, i.e. the lasing threshold, the<br />

feeding of the lasing mode becomes a stimulated process, due to the bosonic nature<br />

of the atoms. A related type of matter-wave amplification was demonstrated by<br />

Inouye et al. [61] using samples of condensed atoms from a BEC.<br />

An optical trap loading scheme is independent from collisional properties and<br />

may also work <strong>for</strong> species with an unsuitably small s-wave scattering length <strong>for</strong><br />

evaporative cooling, see e.g. Refs. [62, 63].<br />

Optical cooling schemes suffer from heating by reabsorbed photons [64–67] and<br />

from light-assisted collisional losses [68, 69]. The phase-space density in a MOT<br />

with optical molasses cooling [70,71] is there<strong>for</strong>e limited to Φ 10 −4 [69,72–74]. A<br />

breakthrough was achieved with Raman sideband cooling [75–78] of atoms that were<br />

tightly confined at the sites of an optical lattice [79, 80]. Han et al. [78] reported<br />

Φ ≈ 1/30 using this technique. An even larger density of Φ ≈ 0.1was achieved by<br />

Ido et al. [81] with Doppler-cooled strontium atoms using a spin-<strong>for</strong>bidden optical<br />

transition. This extremely narrow transition reduces the photon reabsorption.<br />

To overcome the limitations of cooling in a trap, the optical atom laser schemes<br />

[58–60] attempt to use a single dissipative trap loading process to bridge the gap in<br />

phase space density between a precooled sample and a degenerate sample. Several<br />

specific schemes were proposed [82–87], which employ a low-dimensional trapping<br />

geometry, such that photons can escape into a large solid angle without reabsorption.<br />

In particular, it was proposed [82–84,87] to employ optical pumping of atoms at their<br />

turning point on an evanescent-wave mirror to load a low-dimensional trap in the<br />

vicinity of a surface.<br />

Low-dimensional quantum gases are also of conceptual interest <strong>for</strong> the un<strong>der</strong>standing<br />

of phase transition phenomena [43, 45, 88, 89] such as, e.g., the predicted<br />

Kosterlitz-Thouless transition in two dimensions [90]. Furthermore a source of cold<br />

atoms in the vicinity of a surface may be a valuable tool in the emerging field of<br />

guiding atoms along surfaces (“integrated atom optics”) [91–94].<br />

Gauck et al. [86] demonstrated a first realisation of evanescent-wave trap loading<br />

using metastable argon atoms. This species suffers, however, from Penning ionisation<br />

losses [95,96]. Chapter 2 of this thesis describes our own proposal, extending the<br />

trapping scheme to be applicable also with alkali atoms and, more specifically, with<br />

87 Rb. “Dark states” in circularly-polarised evanescent-waves are proposed to reduce<br />

the scattering rate of trapped atoms by several or<strong>der</strong>s in magnitude. The highly spatially<br />

selective pumping by an evanescent wave can be matched to a tightly confining<br />

trap [83]. Un<strong>der</strong>standing and control of photon scattering by bouncing atoms, is<br />

there<strong>for</strong>e necessary to optimise the efficiency and reduce losses of these trap loading<br />

schemes.<br />

In chapter 5, basic properties of our evanescent-wave mirror are characterised,<br />

such as the effective mirror surface. Also the influence of the Van <strong>der</strong> Waals attraction<br />

between atoms and glass surface is discussed. The chapters 6 and 7 describe<br />

different aspects of photon scattering by evanescent waves.


2<br />

A low-dimensional quantum gas<br />

by means of dark states in an<br />

inelastic evanescent-wave mirror<br />

An experimental scheme to create a low-dimensional gas of cold<br />

atoms is discussed, based on inelastic bouncing of cold atoms on<br />

an evanescent-wave mirror. Close to the turning point on the mirror,<br />

atoms are transferred into an optical dipole trap. This scheme<br />

can compress the phase-space density and may ultimately yield an<br />

optically-driven “atom laser”. An important issue is the suppression<br />

of photon scattering due to “cross-talk” between the mirror<br />

potential and the trapping potential. It is proposed that <strong>for</strong> alkalimetal<br />

atoms the photon scattering rate can be suppressed by several<br />

or<strong>der</strong>s of magnitude if the atoms are decoupled from the evanescent<br />

wave. It is discussed how such dark states can be achieved by<br />

making use of circularly-polarised evanescent waves.<br />

This chapter is based on the publication<br />

R.J.C. Spreeuw, D. Voigt, B.T. Wolschrijn, and H.B. van Linden van den Heuvell,<br />

Phys. Rev. A 61, 053604 (2000).<br />

13


14A low-dimensional quantum gas by means of dark states<br />

2.1 Introduction<br />

The only route to quantum degeneracy in a dilute atomic gas which has been experimentally<br />

successful so far [39–44] is evaporative cooling [53–55]. Other routes<br />

to quantum degeneracy, in particular all-optical methods, have been elusive until<br />

now. Nevertheless it is interesting as well as important to keep exploring alternative<br />

methods which do not rely on atomic collisions. Such systems may be held<br />

away from thermal equilibrium and may there<strong>for</strong>e constitute a closer matter-wave<br />

analogy to the optical laser, as compared to “atom lasers” based on Bose-Einstein<br />

condensation [49–52]. In addition, the physics will be quite different because a different<br />

physical, viz. optical, interaction would be used to populate the macroscopic<br />

quantum state: the amplification of a coherent matter wave while emitting photons<br />

(cf. Ref. [61]).<br />

Several proposals <strong>for</strong> an optically-driven atom laser have previously been published.<br />

They have in common that a macroscopic quantum state is populated using<br />

an optical Raman transition [58–60]. Note that also atom laser schemes were proposed<br />

[97–99] which make use of binary atomic collisions, i.e. evaporative cooling.<br />

For an overview of the various proposed schemes, see e.g. Ref. [48].<br />

One problem that has been anticipated from the beginning, is heating and trap<br />

loss caused by reabsorption of the emitted photons [64–66, 68]. There<strong>for</strong>e later<br />

proposals [82–85] and current experiments [17, 86] have aimed at a reduced dimensionality,<br />

based on optical pumping close to a surface. At the same time, there is<br />

also increasing interest in the low-dimensional equivalents of Bose-Einstein condensation<br />

in cold gases [43]. A trap close to a surface is also very interesting from the<br />

viewpoint of cavity QED [25]. The proximity of a dielectric surface can change the<br />

radiative properties of an atom [100], and <strong>for</strong> circularly-polarised evanescent waves<br />

it has been predicted that the radiation pressure (see Chap. 6) is not parallel to the<br />

Poynting vector [101].<br />

In this chapter, it is argued that an evanescent-wave mirror is particularly promising<br />

<strong>for</strong> loading a low-dimensional trap close to a surface. Previous work [59, 60] is<br />

extended so that it can be applied to the alkali-metal atoms. Since these are favourite<br />

atoms <strong>for</strong> laser cooling, the application to alkali-metal atoms will make these kinds<br />

of experiments more easily accessible. In comparison to previous experiments with<br />

metastable noble gas atoms [86], the alkali metals have the advantage that they do<br />

not suffer from Penning ionisation [95,96]. Furthermore, several alkali-metal species<br />

have been cooled to the Bose-Einstein condensation, which makes them good candidates<br />

to create low-dimensional quantum degeneracy also. The extension to the<br />

alkali metals is nontrivial because the splitting between the hyperfine ground states<br />

is not large enough to address them separately with far detuned lasers. The resulting<br />

“cross-talk” would lead to large photon scattering rates in the trap, as is<br />

explained below. It is proposed to use circularly-polarised evanescent waves and to<br />

trap alkali-metal atoms in “dark states”. This allows the detuning to be increased<br />

and the photon scattering rate to be reduced by several or<strong>der</strong>s of magnitude.


2.2 <strong>Evanescent</strong>-wave mirrors 15<br />

2.2 <strong>Evanescent</strong>-wave mirrors<br />

As an introductory excursion, this section describes the phenomenon of evanescent<br />

waves and, in particular, the amplitude and polarisation properties of such optical<br />

waves. Cook and Hill [3] proposed to use an evanescent wave as a mirror <strong>for</strong> slow<br />

neutral atoms, based on the “dipole <strong>for</strong>ce”. <strong>Evanescent</strong>-wave mirrors have since<br />

become an important tool in atom optics [1]. They have been demonstrated <strong>for</strong><br />

atomic beams at grazing incidence [4] and <strong>for</strong> cold atoms at normal incidence [5].<br />

2.2.1 <strong>Evanescent</strong> waves<br />

An evanescent wave appears whenever an electromagnetic wave un<strong>der</strong>goes total<br />

internal reflection (TIR) at a dielectric interface [6,7]. If we consi<strong>der</strong> such an interface<br />

between two dielectrics, a light wave incident on the interface is usually partly<br />

reflected. In “internal” reflection the light is reflecting off the medium with lower<br />

refractive index, see Fig. 2.1(a). In our experiments this is a glass surface in vacuum.<br />

When the angle of incidence, θ i , relative to the surface normal exceeds the critical<br />

angle, θ c =arcsin(n −1 ), the reflection coefficient is unity, i.e. all light is reflected.<br />

For example, the BK 7 glass prism used in our experiments has a refractive index<br />

n =1.511 <strong>for</strong> the rubidium lines at 780 nm and 795 nm wavelength. The critical<br />

angle is thus 41.44 ◦ , or 0.7232 rad. Although no light propagates into the vacuum,<br />

TIR gives rise to an electric field in the vacuum close to the glass surface. This<br />

“evanescent wave” decays exponentially with the distance from the surface on a<br />

length scale of the or<strong>der</strong> of the reduced optical wavelength λ 0 /2π =1/k 0 ,where<br />

k 0 = ω/c is the vacuum wave number.<br />

The evanescent wave can be un<strong>der</strong>stood from Maxwell’s equations with momentum<br />

conservation along the surface. We consi<strong>der</strong> a monochromatic wave,<br />

E(r,t) = 1 2ˆɛ E exp[i(k · r − ω Lt)] + c.c. , (2.1)<br />

with wave vector k and frequency ω L . The complex polarisation vector is denoted<br />

as ˆɛ and the field amplitude is E. The z-direction is taken as surface normal and<br />

k x is assumed to be the wave-vector component parallel to the surface (k y =0).<br />

Maxwell’s equations, expressed as wave equation <strong>for</strong> the electric field, require on the<br />

vacuum side of the surface:<br />

∇ 2 E = 1 ∂ 2 E<br />

=⇒ k 2<br />

c 2 ∂t 2 x + k2 z = k2 0 . (2.2)<br />

Translational invariance of the surface implies momentum conservation in the<br />

x-direction, that is conservation of k x :<br />

k x = k 0 n sin θ i . (2.3)<br />

In TIR, due to k x >k 0 , the normal wave-vector component k z is complex imaginary:<br />

k z = iκ , κ= k 0<br />

√<br />

n2 sin 2 θ i − 1. (2.4)


16 A low-dimensional quantum gas by means of dark states<br />

The electric field thus decays exponentially away from the surface, E ∝ exp(−κz).<br />

Fig. 2.1(b) shows the decay length ξ(θ i )=1/κ(θ i ) as a function of the angle θ i .<br />

In Fig. 2.1(a), the two fundamental linear polarisation vectors of the incident<br />

wave are assigned as s i and p i . In the s, or TE mode, the electric field vector is<br />

directed in the y-direction, perpendicular to the xz-plane of incidence. In the p,<br />

or TM mode, the electric field vector is in the plane of incidence. The polarisation<br />

vectors of the evanescent wave are assigned as s t and p t . From the amplitudes of the<br />

incident field, E s,i and E p,i , the corresponding amplitudes of the evanescent wave, E s,t<br />

and E p,t , are calculated using the same expressions <strong>for</strong> the Fresnel coefficients as <strong>for</strong> a<br />

propagating wave that would be transmitted trough a dielectric interface. However,<br />

the “transmission” angle θ t is complex in TIR. The Fresnel transmission coefficients<br />

t j <strong>for</strong> the two polarisation modes, j = {s, p} = {TE, TM}, and the polarisation<br />

vectors are listed in the Appendix A.2.<br />

The intensity of the incident (and reflected) beam, which propagates inside<br />

the glass substrate, is expressed as I j,i = (1/2)nε 0 c |E j,i | 2 . An effective “intensity”<br />

can also be defined at the glass surface (z = 0) <strong>for</strong> the evanescent wave,<br />

I j,t =(1/2)ε 0 c |E j,t | 2 . The transmittance, i.e. the intensity ratios T j = I j,t /I j,i ,are<br />

T p =(1/n)t ∗ p t p (p ∗ t · p t)andT s =(1/n)t ∗ s t s,<strong>for</strong>p and s polarisation, respectively.<br />

Note that p ∗ t · p t > 1<strong>for</strong> the p polarisation vector, whereas the corresponding expression<br />

<strong>for</strong> the s polarisation, s ∗ t · s t = 1, drops out.<br />

In our experiments, we use an uncoated right-angle prism. Since the evanescentwave<br />

angle of incidence is close to the critical angle, the laser beam with intensity I L<br />

is almost normally incident on the hypotenuse of the prism. The transmittance into<br />

the prism, T L = I j,i /I L , is here independent of the polarisation and approximated<br />

<strong>for</strong> normal incidence by T L ≈ 4n/(n +1) 2 =0.96. The evanescent-wave intensity is<br />

thus enhanced above the laser intensity I L by a factor T j = I j,t /I L = T L T j :<br />

4n cos 2 θ i (2n 2 sin 2 θ i − 1)<br />

T TM = T L<br />

cos 2 θ i + n 2 (n 2 sin 2 θ i − 1) , (2.5)<br />

T TE = T L<br />

4n cos 2 θ i<br />

n 2 − 1 . (2.6)<br />

2.2.2 The evanescent-wave as a mirror <strong>for</strong> atoms<br />

The evanescent-wave dipole, or “light-shift” potential <strong>for</strong> a two-level atom at a<br />

distance z above the surface can be written as (see Appendix A.3 and Ref. [2]):<br />

U dip (z) = U 0 exp(−2κz) , (2.7)<br />

U 0 = 1 2 s 0 δ. (2.8)<br />

The maximum potential at the prism surface, U 0 , is written here in the limit of large<br />

laser detuning, |δ| ≫Γ, and low saturation, s 0 ≪ 1. The saturation parameter is<br />

approximated as<br />

s 0<br />

≃<br />

( Γ<br />

2δ<br />

) 2<br />

T j I L<br />

I 0<br />

. (2.9)


2.2 <strong>Evanescent</strong>-wave mirrors 17<br />

(a)<br />

(b)<br />

3<br />

degrees<br />

0.6 1.1 1.7 2.3<br />

z<br />

y<br />

s<br />

i<br />

x<br />

p<br />

EW<br />

i<br />

<br />

i<br />

p<br />

t<br />

s<br />

t<br />

n<br />

decay length (m)<br />

2<br />

1<br />

0<br />

0 10 20 30 40 50<br />

angle i<br />

- c<br />

(mrad)<br />

Figure 2.1: (a) An evanescent wave (EW) occurs in total internal reflection at a dielectric<br />

interface: refractive index n, angle of incidence θ i >θ c . Incident polarisation vectors s i<br />

and p i . The s polarisation is unchanged in the evanescent wave (s t = s i ), whereas the<br />

p polarisation is elliptical in the xz-plane. (b) <strong>Evanescent</strong>-wave decay length ξ(θ i ),as<br />

calculated with a wavelength λ 0 = 780 nm.<br />

For the D2 line of rubidium, I 0 =1.67 mW/cm 2 is the saturation intensity and<br />

Γ=2π × 6 MHz is the natural transition linewidth. From the Eqs. (2.5) and (2.6)<br />

it follows that the dipole potential induced by a TM-polarised beam always exceeds<br />

that of a TE-polarised beam of equal intensity. Close to the critical angle, θ i ≈ θ c ,<br />

the ratio in optical potential is T TM /T TE ≈ n 2 ≈ 2.28. In our experiments the angle<br />

varies between 0 − 25 mrad from the critical angle, such that T TM ranges between<br />

5.4 − 6.0 andT TE ranges between 2.5 − 2.65.<br />

The detuning of the laser frequency ω L with respect to the atomic transition<br />

frequency ω 0 , is defined as δ = ω L − ω 0 . Thus, a detuning above the resonance<br />

(δ >0, “blue” detuning) yields an exponential potential barrier <strong>for</strong> incoming atoms.<br />

A classical turning point of the motion exists if the barrier height exceeds the kinetic<br />

energy of the atom with incident momentum p i and mass M. This defines the<br />

required threshold potential, U th = p 2 i /2M, <strong>for</strong> atoms being reflected by the mirror.<br />

A plot of such a potential is shown with realistic experimental parameters in Fig 5.2.<br />

For a purely optical potential, the barrier height is U 0 . In reality, the potential<br />

is also influenced by gravity and the Van <strong>der</strong> Waals interaction [102, 103]:<br />

U = U dip + U grav + U VdW , (2.10)<br />

U dip (z) = U 0 exp(−2κz) ∼ 10exp(−2κz) Γ , (2.11)<br />

U grav (z) = Mgz = 4.5 10 −5 k 0 z Γ , (2.12)<br />

( ) 3 ( ) 3<br />

U VdW (z) = − 3(n2 − 1) 1<br />

1<br />

Γ = −0.073 Γ . (2.13)<br />

16(n 2 +1) k 0 z<br />

k 0 z


18 A low-dimensional quantum gas by means of dark states<br />

The gravitational potential can be neglected on the length scale of the evanescentwave<br />

decay length. In contrast, the Van <strong>der</strong> Waals interaction significantly lowers<br />

the potential maximum close to the prism surface. Thus, in combination with the<br />

Gaussian transverse intensity profile of the evanescent wave, the Van <strong>der</strong> Waals<br />

interaction decreases the effective mirror surface on which atoms can bounce. This<br />

effect was experimentally investigated previously by Landragin et al. [103] and is<br />

discussed also in Chap. 5.<br />

2.3 Generic trap loading scheme<br />

2.3.1 An optical trap loaded by a spontaneous Raman<br />

transition<br />

In the following, the generic idea of loading an optical atom trap by an optical<br />

(Raman) transition is briefly reviewed. The original proposal described in Ref. [59]<br />

is based on a Λ-type configuration of three atomic levels, which are indicated here<br />

by |t〉, |b〉 and |e〉, as shown in Fig. 2.2. The levels |t〉 and |b〉 <strong>for</strong> “trapping” and<br />

“bouncing” state, respectively, are electronic ground (or metastable) states, |e〉 is an<br />

electronically excited state. An optical trap is created <strong>for</strong> atoms in level |t〉 using the<br />

optical dipole potential induced by a far off-resonance laser (see e.g. Refs. [10,104]).<br />

Level |b〉 serves as a reservoir of cold atoms, prepared by laser cooling. The cold<br />

atoms are transferred from the reservoir into the trap by a spontaneous Raman<br />

transition |b〉 →|e〉 →|t〉.<br />

The goal is to load a large number of atoms into a single bound state |t, ν〉 of the<br />

trapping potential, where ν is the vibrational quantum number. If the atoms are<br />

bosons, the transition probability into state |t, ν〉 should be enhanced by a factor 1+<br />

N ν ,whereN ν is the occupation of the final state |t, ν〉. If the rate at which atoms are<br />

pumped from |b〉 to |t〉 exceeds a threshold value, the buildup of atoms in |t, ν〉 should<br />

rapidly increase. The Raman filling process can thus be stimulated by the matter<br />

wave in the trapped final state, leading to matter-wave amplification [61]. The<br />

associated threshold is reached when, <strong>for</strong> some bound state |t, ν〉, the unenhanced<br />

filling rate exceeds the unavoidable loss rate. The threshold can be lowered either<br />

by decreasing the loss rate or by increasing the overlap of wave functions (“Franck-<br />

Condon factor”).<br />

Ideally, the energy separation between states |t〉 and |b〉 should be so large that<br />

they can be addressed separately by different lasers. Examples are alkali-metal atoms<br />

or metastable noble gas atoms. The loading scheme has been applied successfully<br />

to load metastable argon atoms into a far off-resonance lattice [105] and into a<br />

quasi-two-dimensional planar matter waveguide [86]. The two metastable states of<br />

Ar ∗ are separated by 42 THz. This chapter focuses on 87 Rb atoms, which are used<br />

in our experiments. Here the separation between the two hyperfine ground states<br />

F g = {1, 2} is only δ GHF =6.8 GHz (see Figs. 2.2(b) and 3.7). This requires a<br />

modification of the scheme as is discussed below.


2.3 Generic trap loading scheme 19<br />

(a)<br />

(b)<br />

e<br />

e<br />

F e=0,1,2,3<br />

pump<br />

FORT<br />

EW<br />

FORT<br />

b<br />

t<br />

2<br />

1<br />

=0<br />

t<br />

b<br />

<br />

GHF<br />

F g=2<br />

F g=1<br />

Figure 2.2: Three-level optical trap loading scheme. (a) Internal atomic states |b〉, |e〉,<br />

and |t〉. <strong>Atoms</strong> are accumulated by means of a spontaneous Raman transition from the<br />

unbound state |b〉 into the bound levels of a far off-resonance trapping potential (FORT),<br />

operating on atoms in the state |t〉. Bosonic enhancement eventually channels all atoms<br />

into the same level ν. (b) 87 Rb hyperfine states, evanescent-wave laser (EW), see text.<br />

2.3.2 The problem of photon reabsorption<br />

It has early been recognised that the photon emitted during the Raman process can<br />

be reabsorbed and thus remove another atom from the trap. This will obviously<br />

counteract the gain process and may even ren<strong>der</strong> the threshold unreachable [60].<br />

This conclusion may be mitigated in certain situations, such as in highly anisotropic<br />

traps [64], in small traps with a size of the or<strong>der</strong> of the optical wavelength [65], and<br />

in the so-called “festina lente” regime [66].<br />

The approach discussed here, is to aim <strong>for</strong> a low-dimensional geometry, with<br />

at least one strongly confining direction z, so that the Lamb-Dicke parameter is<br />

k 0 z ω = √ ω R /ω ≪ 1in that direction [76, 77]. Here k 0 is the optical wave vector,<br />

z ω = √ /2Mω is the rms width of the ground state of the trap with frequency ω <strong>for</strong><br />

an atomic mass M, andω R = k 2 /2M is the recoil frequency. A low-dimensional<br />

geometry should reduce the reabsorption problem because the emitted photon has a<br />

large solid angle available to escape without encountering trapped atoms. Furthermore,<br />

we expect to compress the phase-space density by loading the low-dimensional<br />

optical trap by an evanescent-wave mirror, using optical pumping.


20 A low-dimensional quantum gas by means of dark states<br />

2.4Loading a low dimensional trap<br />

2.4.1 Inelastic evanescent-wave mirror<br />

In the following the specific way is discussed in which the generic scheme from above<br />

may be realised in an experiment. Our implementation is based on an evanescentwave<br />

mirror, using explicitly the level scheme of 87 Rb atoms. The role of the states<br />

|t〉 and |b〉 is played by the two hyperfine sublevels of the ground state 5s 2 S 1/2<br />

(F g =1, 2), which are separated by δ GHF =6.8 GHz. We take the lower level,<br />

F g = 1as the “bouncing state” |b〉 and the upper level, F g = 2, as the “trapping<br />

state” |t〉, as illustrated in Fig. 2.2(b).<br />

The consi<strong>der</strong>ed configuration of laser beams is sketched in Fig. 2.3(a). An evanescent<br />

wave is generated by total internal reflection of a “bouncer” beam inside a<br />

prism. This bouncer is blue with respect to a transition starting from the F g =1<br />

ground state, with a detuning δ 1 . A second laser beam, the “trapper” beam, is<br />

incident on the prism surface from the vacuum side and is partially reflected from<br />

the surface. The reflected wave interferes with the incident wave to produce a set<br />

of planar fringes, parallel to the prism surface. Note that even with 4 % reflectivity<br />

of an uncoated glass surface (refractive index n =1.5), the fringe visibility will<br />

be V =(I max − I min )/(I max + I min )=0.38, where I max and I min are the intensity<br />

maxima and minima, respectively. There<strong>for</strong>e a specific reflection coating may not<br />

be necessary. Note that a possible coating must not inhibit the application of the<br />

bouncer beam.<br />

The trapper beam can be either red or blue detuned, the <strong>for</strong>mer having the<br />

advantage that it automatically provides also transverse confinement. In Fig. 2.3(b)<br />

the situation <strong>for</strong> blue detuning is sketched, confining the atoms vertically in the<br />

intensity minima, but allowing them to move freely in the transverse direction. We<br />

assume that the loss rate due to moving out of the beam is slow compared to other<br />

loss rates, such as that due to photon scattering. Alternatively, one can obtain<br />

transverse confinement by using multiple trapper beams from different directions,<br />

which interfere to yield a lattice potential. Similarly, one can create an optical lattice<br />

using multiple bouncer beams (see e.g. Fig. 2.6). Also an additional hollow beam<br />

may provide transverse confinement, as reported in Ref. [17].<br />

<strong>Cold</strong> atoms, in the bouncing state F g = 1, are dropped onto the prism and are<br />

slowed down by the repulsive light-shift potential induced by the bouncer beam [see<br />

Fig. 2.3(b)]. If the potential is strong enough, the atoms turn around be<strong>for</strong>e they<br />

hit the prism and bounce back up. Thus, an evanescent-wave mirror, or “atomic<br />

trampoline” is <strong>for</strong>med.<br />

We are here interested in interrupting the bouncing atoms halfway during the<br />

bounce, near the classical turning point. The interruption can occur when the atom<br />

scatters an evanescent-wave photon and makes a Raman transition to the other<br />

hyperfine ground state, F g = 2. This Raman transition yields a sudden change<br />

of the optical potential, because <strong>for</strong> an atom in F g = 2 the detuning is larger by<br />

approximately the ground state hyperfine splitting δ GHF . This mechanism has been<br />

used <strong>for</strong> evanescent-wave reflection cooling [16, 17, 106].


2.4Loading a low dimensional trap 21<br />

(a)<br />

trapper<br />

(FORT)<br />

87<br />

Rb<br />

(b)<br />

Uz ()/ h<br />

1 MHz<br />

<br />

pumping<br />

trapping<br />

( F =2)<br />

g<br />

z<br />

bouncing<br />

( F =1)<br />

g<br />

bouncer<br />

(EW)<br />

10 MHz<br />

<br />

<br />

<br />

z<br />

Figure 2.3: Trap loading using an inelastic evanescent-wave mirror. (a) Geometry of<br />

laser beams, incident on a vacuum-dielectric interface. (b) Corresponding potential curves<br />

<strong>for</strong> “bouncing” and “trapping” state <strong>for</strong> 87 Rb (F g =1, 2). <strong>Cold</strong> atoms fall towards the<br />

surface, where they are slowed down by the repulsive potential due to the evanescent<br />

“bouncing” field. Near the turning point atoms un<strong>der</strong>go a spontaneous Raman transition<br />

and become trapped in the optical potential of a standing “trapping” wave. The ripple<br />

on the evanescent wave represents cross-talk from the standing wave (see text). The tick<br />

mark at one-half the optical wavelength, λ 0 /2, indicates the typical length scale. The axis<br />

break indicates the hyperfine splitting, δ GHF =6.8 GHz.<br />

In our case, we tailor the potentials so that the bouncer potential dominates <strong>for</strong><br />

F g = 1and the trapper <strong>for</strong> F g = 2. The atom is thus slowed down by the bouncer<br />

and then transferred into the trapping potential.<br />

As long as the probability <strong>for</strong> un<strong>der</strong>going a Raman transition during the bounce<br />

is not too large (P 1−e −2 , see below), the transition will take place predominantly<br />

near the turning point, <strong>for</strong> two reasons. First, the atoms spend a relatively long time<br />

near the turning point. Secondly, the intensity of the optical pump (the evanescent<br />

wave) is highest in the turning point. The probability that the atoms end up in<br />

the lowest bound state of the trapping potential has been estimated to be on the<br />

or<strong>der</strong> of 10 − 20 %, albeit <strong>for</strong> somewhat different geometries [82, 83]. The resulting<br />

compression of a three-dimensional cloud into two dimensions is in fact dissipative<br />

and can there<strong>for</strong>e increase the phase-space density.


22 A low-dimensional quantum gas by means of dark states<br />

2.4.2 Phase-space compression<br />

In the following, the result of a classical trajectory simulation is discussed, starting<br />

from the dimensionless phase-space distribution Φ 0 (z, v) <strong>for</strong> the vertical motion of<br />

a single atom cooled in optical molasses, shown in Fig. 2.4(a). The vertical velocity<br />

component is denoted as v here, and the subscript z in v z is dropped throughout<br />

this chapter. The phase-space density has been made dimensionless by dividing it<br />

by the phase-space density of quantum states. The latter is given by M/h [quantum<br />

states per unit area in the (z, v) space], where h is Planck’s constant. The distribution<br />

Φ 0 (z, v) can be interpreted as the probability that the atom is in an arbitrary<br />

quantum state localised around (z, v).<br />

The atom, described by the classical distribution Φ 0 (z, v), is assumed to enter<br />

the evanescent wave at a velocity v i = p i /M, determined by its velocity in the<br />

molasses v 0 and the height z 0 from which it falls. Inside the evanescent wave the<br />

atom moves as a point particle along a phase-space trajectory (z(t),v(t)), governed<br />

by the evanescent-wave potential U dip (z) from Eq. (2.7).<br />

Similar to the optical potential, the photon scattering rate Γ ′ of a two-level atom<br />

in steady-state and at low saturation is proportional to the saturation parameter s 0 ,<br />

and can be expressed using U dip (z):<br />

Γ ′ (z) = Γ ′ 0 exp(−2κz) = Γ δ U dip(z) , (2.14)<br />

Finally, the Raman transition rate is given by<br />

Γ ′ 0 = 1 2 s 0Γ . (2.15)<br />

R(z) = R 0 exp(−2κz) , (2.16)<br />

R 0 = q Γ ′ 0 , (2.17)<br />

where q is the branching ratio, i.e. 1 − q is the probability that photon scattering<br />

leads to a Raman transition. The Raman rate gives the local probability per unit<br />

time that the trajectory is interrupted.<br />

The moving atom in the evanescent-wave perceives a time-dependent saturation<br />

parameter, s(t) =s 0 exp(−2κz(t)). Assuming that the excited state population<br />

follows adiabatically, we can integrate the scattering rate along the trajectory to<br />

obtain the number of scattered photons,<br />

∫<br />

N scat = Γ ′ (t)dt = Γ ∫ +pi<br />

( )<br />

Udip<br />

dp . (2.18)<br />

δ −∂ z U<br />

If neglecting the Van <strong>der</strong> Waals contribution and gravity in Eq. (2.10), that is <strong>for</strong> a<br />

purely optical potential U∝exp(−2κz), this leads to an analytical solution:<br />

N scat = Γ δ<br />

−p i<br />

p i<br />

κ . (2.19)


2.4Loading a low dimensional trap 23<br />

Figure 2.4: Phase-space compression due to inelastic bouncing on an evanescent-wave<br />

mirror, based on a classical trajectory simulation. (a) Initial one-dimensional phase-space<br />

distribution of a single atom. (b) Distribution of phase-space coordinates where the<br />

bounce was interrupted due to a spontaneous Raman transition. Note that the spatial<br />

scale changes from cm to µm and that the peak phase-space density along the line v p =0<br />

increases by a factor ∼ 10 3 .<br />

Because of the stochastic nature of the spontaneous Raman transition, we obtain<br />

a probability distribution, Φ p (z p ,v p ), over pumping coordinates where the trajectory<br />

through the phase space is interrupted due to the transition. Not surprisingly, we<br />

see in Fig. 2.4(b) that this distribution has the shape of a (decreasing) “mountain<br />

ridge” following the phase-space trajectory.<br />

Our goal is to load the pumped atoms into a bound state of a trap near the<br />

surface. There<strong>for</strong>e the number of interest is the peak value of Φ p (z p , 0), which<br />

occurs <strong>for</strong> a value of z near the turning point. Fig. 2.4(b) shows that the peak value<br />

of Φ p (z p , 0) is about 1000 times higher than the initial peak value of Φ 0 (z, 0) in<br />

optical molasses [73, 74], see Fig. 2.4(a). The peak value of 0.11 can be interpreted<br />

as the trapping probability in the ground state of the trap that collects the atoms.<br />

This value is quite comparable to previous calculations by different methods using<br />

quantum Monte Carlo simulations [82, 83].<br />

The position of the turning point should be adjusted to coincide with the centre<br />

of the trap, <strong>for</strong> example by adjusting U 0 or κ. Using the Raman transition rate,<br />

we can define a survival probability Q(v) <strong>for</strong> a bouncing atom, with velocities v =<br />

−v i ...+ v i along the trajectory, dQ/dt = −RQ. UsingV (v) =−dQ/dv, this leads<br />

to a distribution V (v) in pumping velocities:<br />

(<br />

Q(v) = exp − R )<br />

0M(v + v i )<br />

, (2.20)<br />

U 0 2κ<br />

V (v) = R 0M<br />

U 0 2κ<br />

Q(v) . (2.21)<br />

The trapping probability, i.e. V (0) can be maximised by changing the value of κ<br />

and the ratio U 0 /R 0 , in such a way that U 0 /R 0 = Mv i /2κ. This corresponds to a


24A low-dimensional quantum gas by means of dark states<br />

situation where the probability <strong>for</strong> reaching the turning point without being optically<br />

pumped is Q(0) = 1/e, orQ(v i )=1/e 2 <strong>for</strong> completing the bounce. If the pumping<br />

rate is very high, too many atoms are pumped be<strong>for</strong>e they reach the turning point.<br />

If the pumping rate is very low, too many atoms bounce without being pumped<br />

at all. If the optical pumping is done by the same laser that induces the bouncing<br />

potential, we have U 0 /R 0 = δ/Γq, so that we obtain an optimum value <strong>for</strong> the<br />

detuning:<br />

δ = q p i<br />

2κ Γ . (2.22)<br />

Experimentally it may be advantageous to use separate lasers <strong>for</strong> the mirror potential<br />

and <strong>for</strong> pumping so that this restriction on the detuning does not apply.<br />

Obviously, one should be somewhat careful in assigning quantitative meaning to<br />

the result of this classical simulation. In particular it has to be verified that the<br />

distribution Φ p (z p , 0) is broad on the characteristic length scale of the (quantum<br />

mechanical) atomic wavefunction near the turning point. The latter is determined<br />

by the slope of the bouncing potential near the turning point. With the approximation<br />

of a constant slope near the turning point, the corresponding Schrödinger<br />

equation is solved by an Airy function with a characteristic width of the first lobe of<br />

∼ κ −1 (κ/p i ) 2/3 . For the same parameters as used in Fig. 2.4(b) this characteristic<br />

width is ∼ 22 nm, indeed smaller than the width of Φ p (z p , 0), which is ∼ 50 nm.<br />

2.5 Photon scattering<br />

2.5.1 Metastable atoms versus alkali-metal atoms<br />

The level scheme used in the proposal of Ref. [59] was inspired by metastable noble<br />

gas or alkaline earth atoms. In those cases two (meta)stable states can usually be<br />

found with a large energy separation. This makes it relatively straight<strong>for</strong>ward to<br />

separate the bouncing and trapping processes, as demonstrated experimentally <strong>for</strong><br />

Ar ∗ in Ref. [86]. Note however, that Penning ionisation of the metastable species<br />

constitutes a severe loss mechanism and has to be taken into account in the regime<br />

of large atomic density [95, 96]. In our scheme those ideas are extended, applying<br />

them to the typical level scheme of the alkali metals. In this case the separation<br />

between two stable states is limited to the ground state hyperfine splitting.<br />

There<strong>for</strong>e, the issue of photon scattering by atoms that have been transferred<br />

into the trap, is addressed in the following. More specifically, our main concern is<br />

scattering of bouncer light. Since, by Eq. (2.14), the rate of scattering light from the<br />

trapping laser is related to the optical potential, Γ ′ /U dip ∝ 1/δ, it can in principle<br />

be made negligibly small by choosing a large enough detuning. This can be done<br />

because the trapping potential can be much shallower than the bouncing potential<br />

and there<strong>for</strong>e need not be F -state specific. (In fact, it will be F -state specific <strong>for</strong><br />

the dark states discussed below.) For example, if the atoms are dropped from 6 mm<br />

above the prism, their incident kinetic energy is E i /k B =0.6 mK, corresponding


2.5 Photon scattering 25<br />

to a required bouncing potential of U th /h = 12 MHz. For the trapping potential,<br />

on the other hand, a depth of less than 50 µK (1MHz) should be sufficient, since<br />

most of the external energy of the atom has been used <strong>for</strong> climbing the bouncing<br />

potential. For the bouncing state F g = 1, the trapping potential then appears as a<br />

small ripple superimposed on the bouncing potential.<br />

The scattering of bouncer light is more difficult to avoid. Ideally, the interaction<br />

of the atoms with the bouncer should vanish completely as soon as they are transferred<br />

into the F g = 2 state. In reality, the bouncer connects both ground states,<br />

F g =1andF g = 2 to the excited state through a dipole-allowed transition. We can<br />

approach the ideal situation by a proper choice of the bouncer detuning. For the<br />

simplified three-level scheme of Fig. 2.2(b), a limitation is imposed by the ground<br />

state hyperfine splitting δ GHF . A good distinction between the F g =1andF g =2<br />

states is only obtained if the bouncer detuning is small, δ 1 ≪ δ GHF . However, a very<br />

small detuning is undesirable because it leads to an increased photon scattering rate<br />

and thus heating during the bounce and also in the trapped final state.<br />

Typical experimental settings are p i ≃ 60 k 0 <strong>for</strong> the momentum of a rubidium<br />

atom falling from a height of about 6 mm, and κ ≃ 0.15 k 0 <strong>for</strong> an angle of incidence<br />

θ i = θ c + 10 mrad. If we operate in the regime qN scat ≃ 2 (i.e. until the turning<br />

point we have qN scat ≃ 1)andsetq =0.5, this requires a detuning δ 1 ≃ 100 Γ ≃<br />

2π × 0.6 GHz. After the atom has been transferred into the trapping potential <strong>for</strong><br />

F g = 2, the detuning of the bouncer will be δ 2 = δ 1 +δ GHF ≃ 2π ×7.4 GHz≃ 1200 Γ.<br />

The trapped atoms will then scatter bouncer light at an unacceptably high rate of<br />

typically 5 × 10 3 s −1 .<br />

2.5.2 Dark states<br />

The limitation imposed by the hyperfine splitting, δ 1 ≪ δ GHF , can be overcome by<br />

making use of dark states, see e.g. Refs. [107, 108]. This requires a more detailed<br />

look at the Zeeman sublevels of the hyperfine ground states. We consi<strong>der</strong> the state<br />

|F g = m g = 2〉 and tune the bouncer laser to the D1resonance line (795 nm,<br />

5s 2 S 1/2 → 5p 2 S 1/2 ), see Fig. 2.5(a). If this light is σ + -polarised, the selection rules<br />

require an excited state |F e = m e =3〉, which is not available in the 5p 2 S 1/2 manifold<br />

and so |F g = m g =2〉 is a dark state with respect to the entire D1line.<br />

The state selectivity of the interaction with bouncer light no longer depends on<br />

the detuning, but rather on a selection rule. There<strong>for</strong>e the bouncer detuning can<br />

be chosen large compared to δ GHF . The new limitation on the detuning is the fine<br />

structure splitting of the D-lines, 7.2 THz (or 15 nm) <strong>for</strong> rubidium. This reduces<br />

the photon scattering rate by 3 or<strong>der</strong>s of magnitude. Note that heavier alkalimetal<br />

atoms are more favourable in this respect because of the larger fine structure<br />

splitting. The price to be paid is the restriction to two specific Zeeman sublevels<br />

|F g = ±m g =2〉 and the need <strong>for</strong> a circularly-polarised evanescent wave.


26 A low-dimensional quantum gas by means of dark states<br />

2.6 Circularly-polarised evanescent waves<br />

In this section, two methods <strong>for</strong> the generation of evanescent waves with circular<br />

polarisation are described, using either a single bouncer beam or a combination of<br />

two. The resulting photon scattering rates are also calculated.<br />

2.6.1 Single beam<br />

A circularly-polarised evanescent wave can be obtained using a single incident laser<br />

beam if it has the proper elliptical polarisation, i.e. the proper superposition of TE<br />

and TM polarisation. The TE-mode yields an evanescent electric field parallel to<br />

the surface and perpendicular to the plane of incidence. The evanescent field of the<br />

TM-mode is elliptically polarised in the plane of incidence, with the long axis of the<br />

ellipse along the surface normal. This was shown in Fig. 2.1(a).<br />

It is straight<strong>for</strong>ward to calculate the input polarisation that yields circular polarisation<br />

in the evanescent wave. We find that the required ellipticity of the input<br />

polarisation is the inverse of the refractive index, 1/n. Here the ellipticity is defined<br />

as the ratio of the minor and major axes of the ellipse traced out by the electric field<br />

vector. The orientation φ of the ellipse is defined as the angle of its major axis with<br />

respect to the normal of the xz-plane of incidence, see Fig. 2.5(b). The required<br />

orientation depends on the angle of incidence:<br />

√<br />

n2 sin 2 θ i − 1<br />

tan φ = −<br />

. (2.23)<br />

cos θ i<br />

Close to the critical angle this is φ ≈ 0, and the ellipse has its major axis perpendicular<br />

to the plane of incidence.<br />

Following this prescription, the resulting evanescent wave will be circularly polarised,<br />

with the plane of polarisation perpendicular to the surface. However, the<br />

plane of polarisation is not perpendicular to the in-plane component, k x ,ofthe<br />

k-vector. Here the evanescent wave differs from a propagating wave, which has its<br />

plane of polarisation always perpendicular to the k-vector (and Poynting vector).<br />

For the evanescent wave the plane of circular polarisation is also perpendicular to<br />

the Poynting vector. However, the Poynting vector is not parallel to the in-plane<br />

k-vector, but tilted sideways by an angle ±χ <strong>for</strong> σ ± polarisation. It is given by<br />

tan χ = √ n 2 sin 2 θ i − 1=κ λ 0<br />

2π . (2.24)<br />

Close to the critical angle, χ ≈ 0, and the plane of polarisation becomes perpendicular<br />

to the in-plane wave vector, as it is <strong>for</strong> propagating waves.<br />

We can estimate the photon scattering rate of an atom in the dark state<br />

|F g = m g =2〉, residing in the circularly polarised evanescent wave of the bouncer<br />

beam. Ideally, this scattering rate is only due to off-resonance excitation to the<br />

5p 2 P 3/2 manifold (D2 line, 780 nm). Choosing the bouncer detuning at 100 GHz<br />

(with respect to the D1line) yields a scattering rate of Γ ′ D2 =3.5 s−1 . In practice<br />

there will also be scattering due to polarisation impurity. For example, assuming<br />

this impurity to be 10 −3 , we obtain a scattering rate of Γ ′ D1,σ<br />

=10.6 s −1 .<br />


2.6 Circularly-polarised evanescent waves 27<br />

(a)<br />

(b)<br />

z<br />

-2<br />

-1 0 +1 +2<br />

-1 0 +1<br />

F e<br />

2<br />

1<br />

x<br />

S<br />

<br />

+<br />

EW<br />

y<br />

glass surface<br />

-2<br />

-1 0 +1 +2<br />

F g<br />

2<br />

EW<br />

<br />

i 0<br />

<br />

m = g<br />

-1 0 +1<br />

1<br />

Figure 2.5: Dark state in a single-beam of a σ + -circularly-polarised evanescent wave.<br />

(a) Dark state |F g = m g =+2〉 in σ + -polarised light, tuned above the D1 line of 87 Rb.<br />

(b) Glass surface in the xy-plane, evanescent-wave (EW) angle of incidence θ i . Elliptical<br />

incident polarisation, rotated by the angle φ with respect to the normal to the xz-plane<br />

of incidence. The thin dashed line indicating 90 ◦ − φ is normal to the EW beam and in<br />

the plane of incidence. The Poynting vector S is in the xy-plane, rotated by the angle +χ<br />

out of the x-direction.<br />

2.6.2 Two crossing TE waves<br />

Alternatively, evanescent waves of circular polarisation can be produced using two<br />

(or more) bouncer beams. Two TE polarised evanescent waves, crossed at 90 ◦ ,<br />

will produce a polarisation gradient as sketched in Fig. 2.6(a). Lines of circular<br />

polarisation are now produced with the plane of polarisation parallel to the surface.<br />

Lines of opposite circular polarisations alternate, with a distance of approximately<br />

λ 0 /2 √ 2 between neighbouring σ + and σ − lines.<br />

This configuration offers interesting opportunities. The light field can be decomposed<br />

into two interleaved standing wave patterns, <strong>for</strong> σ + and σ − polarisation,<br />

respectively. An atom in the state |F g = m g =2〉 is dark with respect to the σ +<br />

standing wave only. However it does interact with the σ − standing wave and there<strong>for</strong>e<br />

can be trapped in its nodes. The bouncer light will thus play a double role.<br />

First it slows the atoms on their way down to the surface. Then, after the atoms<br />

have been optically pumped, the bouncer light will transversely confine the atoms.<br />

The situation be<strong>for</strong>e and after pumping is shown in Fig. 2.6(b,c). We thus expect<br />

a 1D lattice of atomic quantum wires with alternating spin states, very much like<br />

a surface version of previously demonstrated optical lattices [109,110]. The vertical<br />

confinement in the z-direction can still be achieved by an additional trapping field,


28 A low-dimensional quantum gas by means of dark states<br />

(a)<br />

r<br />

(b)<br />

(c)<br />

U<br />

15<br />

10<br />

5<br />

0<br />

U<br />

15<br />

10<br />

5<br />

0<br />

/ h<br />

0.5<br />

r/ 0<br />

/ h<br />

0.5<br />

r/ 0<br />

1<br />

1<br />

1, 0<br />

1.5 0<br />

0.5<br />

2, 2<br />

1.5 0<br />

0.5<br />

1<br />

z/ 0<br />

2<br />

1.5<br />

2<br />

1.5<br />

1<br />

z/ 0<br />

Figure 2.6: (a) Generating circularly-polarised evanescent waves by crossing two TEpolarised<br />

waves at a right angle (looking down at the prism surface). The polarisations<br />

and in-plane wave vector components yield a fringe pattern of alternating lines of opposite<br />

circular polarisation. The total intensity is constant across the pattern, since the two TE<br />

polarisations are orthogonal. The optical potential is shown in (b) <strong>for</strong> an atom incident<br />

in state |F g ,m g 〉 = |1, 0〉, andin(c) <strong>for</strong> a pumped atom in the (locally) dark state |2, 2〉.<br />

The potentials are plotted vs. the height z and the transverse r-direction in the xy-plane<br />

(orthogonally crossing the fringe pattern).<br />

see also Fig. 2.3. The transverse lattice structure may allow postcooling of atoms in<br />

the trap by Sisyphus cooling [111, 112] or Raman sideband cooling [75–78].<br />

It is not strictly necessary to cross the evanescent waves at a right angle, but it<br />

has the advantage that the total intensity is constant across the polarisation pattern.<br />

The same could also be achieved by using counter-propagating evanescent waves with<br />

orthogonal polarisations. For any other angle, the intensity varies spatially so that<br />

the atoms bounce on a corrugated optical potential. However, even with a uni<strong>for</strong>m<br />

intensity, most atoms will experience a corrugated potential, as shown in Fig. 2.7.<br />

The potential depends on the local polarisation and on the atom’s magnetic sublevel<br />

through the Clebsch-Gordan coefficients. Only <strong>for</strong> the state |F g =1,m g =0〉 is<br />

the dipole potential independent of the polarisation. One could of course prepare<br />

the falling atoms in |F g =1,m g =0〉 using optical pumping. The local circular<br />

polarisation σ ± will tend to pump the atom into the local dark state |F g = ±m g =2〉.<br />

However the optical pumping transition then has a branching ratio of only 1/6<br />

(using a dedicated resonant pumping beam). By contrast, <strong>for</strong> an atom starting in<br />

|F g =1,m g =1〉, the branching ratio is 1/2. There<strong>for</strong>e starting in |F g =1,m g =0〉<br />

is conceptually simple, but probably not optimal.


2.6 Circularly-polarised evanescent waves 29<br />

(a)<br />

U/ h<br />

(b)<br />

U/ h<br />

15<br />

10<br />

5<br />

15<br />

10<br />

1, m<br />

2, m<br />

g<br />

0.5 1 1.5 2<br />

g<br />

r/ 0<br />

(c)<br />

U/ h U/ h<br />

(d)<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0.2<br />

0<br />

1, 0<br />

2, 2<br />

0 1 2 3<br />

2, 2<br />

z/ 0<br />

5<br />

-0.2<br />

0.5 1 1.5 2<br />

r/ 0<br />

0 1 2 3<br />

z/ 0<br />

Figure 2.7: Potentials <strong>for</strong> crossed TE-polarised evanescent waves. (a) Bouncing potential<br />

<strong>for</strong> the sublevels of an atom in |1,m g =0, ±1〉. The mirror is smooth only <strong>for</strong> m g =0.<br />

(b) Transverse evanescent-wave trapping potential <strong>for</strong> an atom in |2,m g = −2 ...2〉.<br />

Dark states (zero light shift) occur only <strong>for</strong> m g = ±2, at alternating transverse locations.<br />

(c,d) Vertical confinement by a standing wave potential, calculated <strong>for</strong> a transverse<br />

r-coordinate of a dark state |2, 2〉 in a σ − -node. (c) shows the modulation of the bouncing<br />

potential by the trapping laser. (d) zooms in to the trapping potential. The dark state<br />

is decoupled from the bouncing laser on the D1 line, it does however perceive a weakly<br />

attractive potential due to coupling on the D2 line, <strong>for</strong> which the evanescent wave is very<br />

far red detuned. This effect is also visible in (b) by a slightly negative light shift.<br />

A disadvantage of creating circularly-polarised evanescent waves on a lattice<br />

is that an additional source of photon scattering appears. We approximate the<br />

transverse potential near the minimum as a harmonic oscillator. Choosing again<br />

the bouncer detuning at 100 GHz, the harmonic oscillator frequency will be about<br />

ω/2π = 480 kHz. An atom in state |2, 2〉, i.e. the ground state of the harmonic<br />

oscillator associated with the σ − node, has a Gaussian wavefunction with wings extending<br />

into the region with σ − light. The resulting scattering rate can be estimated<br />

as:<br />

Γ ′ HO ≈ ω Γ<br />

4(δ 1 + δ GHF ) ≈ 52 s−1 . (2.25)<br />

Here the bouncer detuning was again chosen at 100 GHz. The scattering rate can<br />

be further suppressed to Γ ′ HO ≈ 18s−1 by raising the bouncer detuning to 300 GHz.<br />

For an even larger detuning the off-resonance scattering by the D2 line, Γ ′ D2 , starts<br />

to dominate.


30 A low-dimensional quantum gas by means of dark states<br />

2.6.3 Feasibility<br />

One should point out that the examples to produce circularly polarised evanescent<br />

waves are not meant to be exhaustive. Several other methods can be devised, some<br />

being more experimentally challenging than others.<br />

For the single-beam method the incident beam must be prepared with the correct<br />

ellipticity as well as the correct orientation. It will probably be difficult to measure<br />

the polarisation of the evanescent wave directly. One should there<strong>for</strong>e prepare the<br />

incident polarisation using well-calibrated optical retar<strong>der</strong>s and using calculated<br />

initial settings. For example, as an experimental method, the use of reversibility<br />

theorems <strong>for</strong> the polarisation of plane waves was discussed in Ref. [113]. The fine<br />

tuning could then be done, e.g., by optimising the lifetime of the trapped atoms in<br />

the dark state.<br />

For the two-beam method of Fig. 2.6(a) we have assumed <strong>for</strong> simplicity that<br />

the two interfering evanescent waves have the same decay length and the same<br />

amplitude. Equal decay lengths <strong>for</strong> the two waves can be en<strong>for</strong>ced by making use<br />

of a dielectric waveguide [114]. Alternatively, one may deliberately give the two<br />

beams a slightly unequal decay length and, at the same time, give the wave with<br />

the shorter decay length a larger amplitude. In this case there will always be one<br />

particular height above the surface where the two beams have equal amplitude, as<br />

required <strong>for</strong> superposing to circular polarisation. This procedure would make the<br />

circular polarisation somewhat self-adjusting. The height where circular polarisation<br />

occurs is tunable by changing the relative intensity of the two beams.<br />

Obviously, the final word on the feasibility can only be given experimentally. In<br />

our ongoing experiments, a variation on Fig. 2.6(a) is pursued, including the just<br />

mentioned self-adjusting properties.<br />

2.7 Conclusion<br />

It was discussed that inelastic bouncing on an evanescent-wave mirror is a promising<br />

method <strong>for</strong> achieving high phase-space density in low-dimensional optical traps. The<br />

phase space compression is achieved by means of a spontaneous Raman transition,<br />

which is highly spatially selective <strong>for</strong> atoms near the turning point of the evanescentwave<br />

mirror potential.<br />

Previous work based on the level schemes of metastable noble gas atoms was<br />

extended <strong>for</strong> application to alkali-metal atoms. This requires suppression of the high<br />

photon scattering rate, resulting from the relatively small ground state hyperfine<br />

splitting of the alkali-metal atoms. It was shown how the photon scattering rate<br />

can be reduced by several or<strong>der</strong>s of magnitude, by trapping the atoms in dark<br />

states. This requires the use of circularly-polarised evanescent waves, which can<br />

be generated by several methods. If built up from multiple beams, the evanescent<br />

field may play a double role, generating a bouncing potential as well as a trapping<br />

potential. This could lead to an array of quantum wires <strong>for</strong> atoms.


3<br />

Experimental setup<br />

A table-top ultra-high vacuum rubidium vapour cell has been built.<br />

Optical access to the vacuum system is achieved by use of a rectangular<br />

glass cell. Two techniques of vacuum sealing of such glass cells<br />

using either a knife-edged metal gasket or epoxy glue are discussed.<br />

In the vapour cell a magneto-optical trap is operated. With additional<br />

optical molasses, cooling provides samples of ≈ 10 7 atoms at<br />

temperatures of ≈ 10 µK. Frequency-stabilised diode lasers serve<br />

as trapping and cooling light sources. Their output is amplified by<br />

injection-locked single-mode diode lasers or, <strong>for</strong> high-power applications,<br />

tapered semiconductor gain elements. Real-time control of<br />

the experiment is achieved by a personal computer with an additional<br />

digital signal processor. <strong>Cold</strong> atoms are detected by imaging<br />

with a triggered digital frame-transfer CCD camera system.<br />

31


32 Experimental setup<br />

3.1 Overview<br />

An optical trapping scheme <strong>for</strong> atoms has to be realised in ultra-high vacuum (UHV)<br />

to avoid atom loss due to collisions with room temperature gas. Since the first<br />

demonstration, the magneto-optical trap (MOT) [33] has become a standard tool in<br />

atomic physics. Usually, a MOT provides a cloud of cold atoms after a single loading<br />

cycle. Alternatively, a slow continuous atomic beam is extracted [115, 116]. In our<br />

experiments, a MOT with subsequent polarisation gradient cooling (PGC) [70] was<br />

used to prepare a cloud of atoms a few mm above an evanescent-wave atom mirror.<br />

We have chosen frequency-stabilised diode lasers to provide the various light frequencies<br />

required <strong>for</strong> the MOT, PGC, optical pumping, dipole trapping, and probing of<br />

atoms. Such devices are a low-cost and less maintenance demanding alternative to<br />

Ti:Sapphire laser systems. Their compactness permits to assemble a larger number<br />

of laser sources together with an UHV setup on a single table. The lasers consist<br />

of external grating diode lasers [117, 118], the output of which is amplified by<br />

injection-locked diode lasers. The laser stabilisation schemes are based on frequencymodulation<br />

spectroscopy [119] and “Zeeman polarisation spectroscopy” [120–122].<br />

The complexity of the experiments demands real-time computer control of experimental<br />

parameters. A digital signal processor is in charge of this task. Images<br />

of atoms bouncing on evanescent-wave mirrors were acquired with a digital CCD<br />

camera system.<br />

In this chapter, the relevant properties of rubidium are discussed and the design<br />

of the table-top UHV rubidium vapour cell is described, in which our experiments<br />

were per<strong>for</strong>med. A separate section is dedicated to the delicate issue of connecting<br />

and sealing glass cuvettes and window substrates to standard CF 40 and CF 16<br />

Conflat TM steel knife-edge flanges. Also an overview of the used laser systems and<br />

the controlling computer hardware is given. In the last section, the MOT is described.<br />

The temperature of atom clouds, achieved by PGC, was determined by a<br />

time-of-flight method using falling atom clouds. The characterisation of one particular<br />

device, a tapered semiconductor amplifier, is given in Chap. 4.<br />

3.2 Atomic species — rubidium<br />

The experimental choice of an atomic species depends on physical properties including,<br />

(i) appropriate optical transition frequencies and the availability of laser<br />

sources operating on these frequencies, (ii) the atomic collisional properties and,<br />

(iii) the ease of handling in an UHV system.<br />

(i) Optical transitions.— Optical cooling techniques require well separated<br />

optical transitions of sufficiently narrow natural linewidth, among which cycling<br />

(“closed”) transitions. Methods like PGC or velocity-selective coherent population<br />

trapping (VSCPT) [107] in “dark states” rely on optical pumping between magnetic<br />

sublevels or hyperfine states. The hyperfine-split D1and D2 fine structure lines of<br />

alkali-metal atoms and the optical transitions of metastable noble gas atoms allow<br />

the use of dye lasers and Ti:Sapphire lasers with, wavelengths from the visible to


3.3 Ultra-high vacuum system 33<br />

the near-infrared spectrum. With LNA lasers at 1083 nm also metastable helium<br />

became usable [123]. For an overview on common elements <strong>for</strong> laser cooling, see<br />

e.g. Ref. [12].<br />

Well established in atomic physics are meanwhile stabilised diode laser systems,<br />

if providing sufficient optical output power together with spectral and spatial beam<br />

quality. The availability of the laser diodes is generally determined by commercial<br />

applications, e.g. <strong>for</strong> CD disk drives (785 nm wavelength), Nd:YAG laser pumping<br />

sources (808 nm), DVD drives (650 − 670 nm), or magnetometers <strong>for</strong> navigational<br />

systems using helium (1083 nm). Particularly, low-cost high-power laser diodes in<br />

the near infrared make rubidium an attractive choice, due to optical resonances at<br />

780 nm and 795 nm wavelength.<br />

(ii) Collisional properties.— In high density applications of cold atoms the<br />

s-wave scattering length is an important parameter, e.g. <strong>for</strong> evaporative cooling<br />

and <strong>for</strong> the properties of a Bose-Einstein condensate. For an overview of scattering<br />

lengths <strong>for</strong> various atomic species, see Ref. [63]. More specifically 87 Rb, due to<br />

its suitable positive scattering length (a ≈ 109 a 0 ), may be the most promising<br />

candidate to reach quantum degeneracy in a purely optical scheme, as envisaged in<br />

Chap. 2. There<strong>for</strong>e we use this isotope in our experiments.<br />

(iii) Handling in UHV.— A reliable and compact technical solution to handle<br />

rubidium atoms is a table-top UHV vapour cell setup, in which a MOT can be quickly<br />

and directly loaded from the room temperature vapour [124–126]. Rubidium can<br />

be used at convenient temperatures. For example, the (saturated) rubidium vapour<br />

pressure at room temperature is between 10 −7 − 10 −6 mbar and the melting point is<br />

38.5 ◦ C [127]. The natural abundance of the 87 Rb isotope is 27.9 %, next to 72.1%<br />

of 85 Rb. Some useful numbers <strong>for</strong> experiments with rubidium are listed in the<br />

Appendix A.1. The hyperfine level structure of the D1 (795 nm) and D2 (780 nm)<br />

line is shown in Fig. 3.7.<br />

3.3 Ultra-high vacuum system<br />

3.3.1 Requirements on a rubidium vapour cell<br />

The used vapour cell is an UHV system which maintains a partial rubidium pressure<br />

of typically 10 −8 mbar and a background gas pressure of 10 −9 mbar. In<br />

an optical trap, light scattering will dominate the loss of trapped atoms, rather<br />

than background gas collisions. An experiment involving bouncing atoms from an<br />

evanescent-wave mirror typically lasts less than 100 ms, whereas the mean collision<br />

time of cold atoms with room temperature atoms from the vapour is ∼ 350 ms<br />

(mean free path ∼ 100 m). Experiments of longer duration, such as evaporative<br />

cooling of atoms towards BEC, require significantly better vacuum ( 10 −11 mbar).<br />

In these cases differentially pumped “double-MOT” systems [128] or bright beams<br />

of slow rubidium atoms [115,116] are employed to load a MOT in good vacuum with<br />

a sufficient number of atoms.


34Experimental setup<br />

A vapour cell can be economically realised as a small stand-alone system using<br />

mostly commercial components. The very low rubidium consumption reduces maintenance<br />

tasks. For example, using 10 mg of rubidium in a reservoir, the operating<br />

time is limited by constructional changes on the system rather than by rubidium<br />

depletion. Due to vibrations, the use of turbo-molecular pumps located on a table<br />

together with stabilised lasers is undesirable. There<strong>for</strong>e we employ ion pumps,<br />

though these pumps require care <strong>for</strong> shielding or compensating their stray magnetic<br />

fields. Usually it is sufficient to place an ion pump far enough away ( 0.5 m)from<br />

the experimental region. However, this is at the cost of pumping speed and vacuum<br />

pressure.<br />

In addition to vacuum specifications, also the optical properties of a vapour cell<br />

have to be consi<strong>der</strong>ed. Experiments on evanescent-wave atom mirrors as discussed<br />

in this thesis, require a prism as the only optical component mounted inside the<br />

vacuum system. Nevertheless, optical access from various directions is needed to<br />

apply the numerous laser beams. The windows should be of laser optical quality and,<br />

if possible, antireflection coated. In the present system, an uncoated rectangular<br />

glass cell is used. Beside good optical access, a particular feature of such a cell is<br />

that magnetic field coils can closely approach the region of interest and hence can<br />

be small sized and of low power consumption. Since a glass cell is nonmagnetic,<br />

experiments are not perturbed by eddy currents caused by switching field coils.<br />

3.3.2 Vapour cell setup<br />

The vacuum system, shown in Fig. 3.1, consists of, (i) a lower UHV chamber, pumped<br />

bya15l/s ion pump and, (ii) an upper differentially pumped vapour cell connected<br />

to a glass cuvette and to a rubidium reservoir. The vapour cell is (optionally)<br />

pumped by a 8 l/s ion pump. The typical background pressure achieved in this<br />

system is ≈ 10 −9 mbar, after gentle bakeout up to 114 ◦ C. The epoxy-glued glass<br />

cell used so far, did not allow warmer baking.<br />

(i) UHV section.— The components of the UHV system are grouped in the<br />

horizontal plane at a 5-way CF 40 cross. The system is clamped to the optical<br />

table by aluminium mounts that can be water cooled, in or<strong>der</strong> to protect the laser<br />

table during bakeout. An all-metal sealed valve (Granville-Phillips, gold-seal type<br />

204) leads via bellows to a roughing turbo-molecular pump. When the valve is<br />

closed, the system is self-sustaining with an ion pump of 15 l/s (N 2 ) pumping speed<br />

(Varian, VacIon Plus 20 StarCell with ferrite magnets). The achieved pressure can<br />

be monitored by an ionisation gauge in a range between 10 −12 − 10 −3 mbar (Varian,<br />

type UHV-24p). A pressure below 10 −9 mbar might be possible by extending the<br />

system with a titanium sublimation pump or non-evaporative getter materials.<br />

The UHV section is separated from the upper vapour cell section by a blank<br />

CF 40 copper gasket with a hole of 1.5 mm diameter. Differential pumping reduces<br />

the pumping speed in or<strong>der</strong> to maintain the rubidium pressure in the vapour cell<br />

during experiments.


3.3 Ultra-high vacuum system 35<br />

IP1<br />

2<br />

Rb<br />

IP2<br />

IG<br />

VP<br />

1<br />

roughing<br />

3<br />

prism<br />

1.5 mm diaphragm<br />

(differential pumping)<br />

vapour cell<br />

-8<br />

10 mbar<br />

IP<br />

15 l/s<br />

2<br />

Rb<br />

-9<br />

UHV 10 mbar<br />

3<br />

IP<br />

8 l/s<br />

Figure 3.1: Vacuum system. Lower UHV section: ion pump (IP1), ion gauge (IG)<br />

and valve ( ⊗ 1) to roughing pump. Upper vapour cell section: rubidium reservoir (Rb,<br />

unmounted when photograph was taken) with valve ( ⊗ 2), and small ion pump (IP2,<br />

out of sight). A bypass valve ( ⊗ 3) connects the vapour cell and the UHV section.<br />

A diaphragm between the sections enables differential pumping. The closeup shows the<br />

10 × 10 × 4 mm 3 right-angle BK 7 glass prism used <strong>for</strong> the evanescent-wave mirror (Melles<br />

Griot, high precision prism, no. 01 PRB 009, cut to a width of 4 mm).


36 Experimental setup<br />

For pumping down from atmospheric pressure and during bakeout, an all-metal<br />

CF 16 valve (Vacuum Generators, type ZCR 20 R) is opened in a bypass from the<br />

5-way cross to the vapour cell. The strong magnets of the ion pump are approximately<br />

35 cm away from the prism. The ion pump manual specifies a stray magnetic<br />

field of 1.5 Gatadistance 15 cm from the pump.<br />

(ii) Vapour cell section.— A hexagonal section with six CF 16 ports is mounted<br />

on top of the UHV section. It interconnects the cuvette, the rubidium reservoir and<br />

the pumping bypass. In addition, an in-line pair of custom-made optical viewports<br />

is mounted, that provides optical access <strong>for</strong>, e.g., time-of-flight diagnostics of falling<br />

atoms (if no prism is mounted). The horizontal tube of the bypass leads 40 cm away<br />

toasmall8l/s ion pump (Varian, VacIon with AlNiCo magnet). If necessary <strong>for</strong><br />

stray field minimization and if the pump is not in use, the magnet can be removed.<br />

Together with the bypass, this pump assists in stabilising the rubidium vapour<br />

pressure or to reduce background gas pressure, respectively.<br />

The rubidium reservoir is connected to the vapour cell by a short spacer tube<br />

and an all-metal valve. It consists of a flexible tube with a short intermediate bellows<br />

section. Be<strong>for</strong>e evacuating the system, a small cylindrical quartz ampule was<br />

inserted, containing a few milligram rubidium. When the final roughing pressure<br />

≈ 10 −6 mbar was established after bakeout, the ampule was broken by bending the<br />

bellows. When the pressure settled again, the system was sealed off from the roughing<br />

line and further pumped down by the ion pump. Commercial standard rubidium<br />

ampules can be used in the setup (Aldrich Chemical, 2 g, no. 38,599-9). However,<br />

a few milligram suffices to keep the system operable <strong>for</strong> years. For constructional<br />

changes, the reservoir can stay evacuated <strong>for</strong> a short time, avoiding a replacement<br />

of the ampule. Hence it is more economical (and more safely) to distil only a small<br />

amount of rubidium into custom reservoir ampules.<br />

In case the ampule breaks too neatly, it might be necessary to keep the bellows<br />

bent to increase the rubidium diffusion out of the ampule. The reservoir is wrapped<br />

with a heating cord. When preparing experiments, the reservoir is gently baked with<br />

open valve until the desired vapour pressure in the cuvette is reached. (The saturated<br />

rubidium vapour pressure is, e.g., ≈ 10 −5 mbar at 60 ◦ C.) It can be monitored<br />

by observing the fluorescence from a laser beam tuned to an optical resonance of<br />

rubidium. After cooling down the reservoir, the valve is kept open and adjusted to<br />

maintain a constant vapour pressure. Due to the differential pumping, the vapour<br />

pressure decays with a time constant of ∼ 30 min if the reservoir is closed. Be<strong>for</strong>e<br />

inserting the differential pumping hole this was less than 5 min. It is difficult to<br />

estimate the rubidium diffusion and pumping speed <strong>for</strong> two reasons: First, the<br />

system has many bends and apertures. Second, rubidium is strongly sticking to<br />

surfaces. Since the surface-to-volume ratio is large, there is a delay of several hours<br />

in vapour pressure build-up when charging the system <strong>for</strong> the first time. One has<br />

to avoid saturating the entire system and, particularly, the cuvette by a rubidium<br />

droplet that slowly “creeps” through the system.


3.4Optical access to the UHV system 37<br />

It is worth mentioning two alternative techniques of charging a vapour cell with<br />

rubidium, taking less constructional and machining ef<strong>for</strong>ts: (i) commercial singleuse<br />

quench-seal copper tubes as containment <strong>for</strong> the rubidium ampule and, (ii) a<br />

saturated dispenser compound that releases rubidium when heated by an electrical<br />

current (SAES Getters, type Rb/NF/3.4/12 FT10+10, 2.6 mg yield). The latter has<br />

the advantage that it offers cw and pulsed operation with short time constants ∼min,<br />

and may charge the vacuum system only locally with rubidium. A disadvantage are<br />

the electrical UHV feedthroughs and the limited rubidium load.<br />

3.4Optical access to the UHV system<br />

In cooling and trapping experiments, laser-beam wavefronts must not be distorted<br />

by the UHV viewports. Also (stress-induced) birefringence of the viewports is undesirable,<br />

since it might perturb polarisation sensitive applications such as polarisation<br />

gradient cooling or “dark state” trapping. Furthermore, the vacuum sealing has to<br />

withstand common bakeout temperatures above 200 ◦ C. Commercial viewports are<br />

usually costly and have clear apertures that are significantly smaller than the Conflat<br />

flange counterpart. In or<strong>der</strong> to achieve optical quality access from many directions,<br />

we have chosen a rectangular cuvette.<br />

The first choice material was fused silica (“quartz glass”). It is available as laseroptical<br />

plate elements, that are welded in a baking process using glass weld pow<strong>der</strong><br />

at the connecting faces. The surface flatness is preserved locally in this process.<br />

A disadvantage is that no inside antireflection (AR) coating can be applied. A<br />

pre-applied coating would be destroyed during welding, and the elongated geometry<br />

makes the application of an UHV compatible coating after welding impossible.<br />

Such cuvettes are usually supplied with a “graded-seal” transition, with which<br />

the mismatch in the thermal expansion of the fused silica cuvette and an Invar steel<br />

flange is compensated. The graded seal consists of a succession of tubular segments<br />

that change gradually in composition from fused silica to Pyrex glass. Un<strong>for</strong>tunately,<br />

the minimal length of the graded seal is > 10 cm, which may degrade the UHV in<br />

the cuvette. For our vacuum system, we have there<strong>for</strong>e extended an earlier reported<br />

window sealing technique [129] to seal a cuvette directly to a CF 40 flange.<br />

In the following, our application of this technique to CF 16 viewports is discussed,<br />

and the CF 40 scheme is refined using spring-loaded knife-edge seals, that reduce<br />

stress on the glass substrate. Finally, a less complex preliminary solution based<br />

on an epoxy-glued “Optical Glass” cuvette is presented. Note that this cuvette,<br />

despite of modest bakeout temperatures, allowed <strong>for</strong> sufficient UHV to per<strong>for</strong>m all<br />

experiments reported in this thesis (see also the photographs in Fig 3.1).


38 Experimental setup<br />

8mm<br />

compression ring<br />

OFHC gasket<br />

34 mm<br />

16 mm<br />

18 mm<br />

o<br />

60<br />

0.4 mm<br />

CF16 flange<br />

Cu knife<br />

Al foil<br />

BK 7 (dia. 22 mm x 6 mm)<br />

Figure 3.2: All-metal sealed optical CF 16 viewport.<br />

3.4.1 All-metal sealed optical quality UHV viewports<br />

The viewport concept in Ref. [129] makes use of a standard 50 mm diameter laser<br />

window that is sealed to a CF 40 knife-edge flange. The substrate is pressed on a<br />

knife edge milled onto the outer surface of a common OFHC (copper) gasket ring.<br />

Un<strong>der</strong> compression, the de<strong>for</strong>ming copper knife edge seals the window with a leak<br />

tightness of 10 −12 mbar l/s helium leakage, comparable with usual CF connections.<br />

We implemented this technique, to realise viewports also <strong>for</strong> CF 16 flanges.<br />

The maximisation of the clear aperture required custom-sized windows of 22 mm<br />

diameter and 6 mm thickness (Melles Griot, BK7,AR/AR HEBBAR coating).<br />

Fig. 3.2 shows a cross section and a frontal view of the viewport construction.<br />

A blank CF 16 flange was milled as a compression ring to clamp the window onto<br />

the copper seal. The ring has a circular overlap of 2 mm width with the window.<br />

Between them we use as a cushion a stack of 10 − 15 punched rings of aluminium<br />

foil. The clear aperture of the viewport is limited by the flange bore diameter and<br />

by the inner diameter of the gasket, both ≈ 16 mm. The knife edge in the copper<br />

gasket is also shown in the figure. Tightening of the six (lubricated) bolts was done<br />

with a torque wrench uni<strong>for</strong>mly and in small steps. A torque of 4.9 Nm provided<br />

good sealing without damage, whereas <strong>for</strong> 5.5 Nm we observed cracks in the AR<br />

coatings of the window. Also the copper gasket was significantly de<strong>for</strong>med due to a<br />

slight mismatch in the knife edge diameters. Compared to the CF 16 steel knife of<br />

18.5 mm diameter, we used initially a slightly smaller diameter of 18.0 mm <strong>for</strong> the<br />

copper knife, in or<strong>der</strong> to keep more space to the edge of the small glass substrate.<br />

Later we also used copper knifes of 18.5 mm. The final sealing torque on the bolts<br />

was 3.9 Nm, similar to that in Ref. [129]. The copper knife was then compressed to<br />

a flat ring of 0.5 mm width, by a total loading <strong>for</strong>ce of ≈ 150 kN from the 6 bolts,<br />

or ≈ 26 kN/cm along the knife.<br />

So far, three of our four windows withstood several bakeout cycles up to 200 ◦ C<br />

(max. 250 ◦ C). One window broke while being unmounted. Another window showed<br />

a slight edge damage by the compression ring but did not leak.


3.4Optical access to the UHV system 39<br />

3.4.2 All-metal sealed fused silica cell<br />

The knife-edged Conflat seal.— In a first attempt, we adapted the knifeedged<br />

CF 40 gasket of the viewports also to a cuvette. Fused silica cuvettes were<br />

manufactured by Optiglass (England) and supplied by Starna Analytical Accessoires<br />

(Austria). They were made from 4 mm thick plates with a square outside width<br />

of 30 mm and lengths of 100 mm and 150 mm. The material is Spectrosil B from<br />

Thermal Syndicate. The cuvettes were molded each on a 15 mm thick ring disk<br />

substrate of 50 mm outer and 22 mm inner diameter. Particularly the lower (sealing)<br />

disk surface was polished.<br />

Thus, the disk resembled a 50 mm dia. (CF 40) fused silica window. Here, the<br />

knife edge milled onto the copper gasket had a diameter of 42 mm, like a CF 40<br />

steel knife. For a window, we achieved a good seal with a torque of 5 Nm on each of<br />

the 6 flange bolts, corresponding with 173 kN total load (13 kN/cm). Nevertheless,<br />

the disk-mounted cuvette did not withstand the compression clamp. Be<strong>for</strong>e vacuum<br />

sealing was achieved, the disk cracked at the corners of the cuvette, when the torque<br />

at the bolts was increased to about 3.4 Nm.<br />

There<strong>for</strong>e it seemed necessary to realise a seal using consi<strong>der</strong>ably less loading<br />

<strong>for</strong>ce on the disk than with an OFHC gasket. A possibility might be a softer gasket,<br />

e.g. made from nickel. However, nickel is ferromagnetic and there<strong>for</strong>e undesirable<br />

close to the experimental region.<br />

The Helicoflex ∆ spring-loaded seal.— With the single-side knife-edged gasket,<br />

most of the compression was needed to de<strong>for</strong>m the bulk gasket material by<br />

the CF steel knife. There<strong>for</strong>e, a double knife-edged gasket between two flat surfaces<br />

promised a stress reduction. A commercial solution is the Helicoflex ∆ gasket<br />

(Le Carbone-Lorraine, type HNV 200 ∆ (DN 25), spring Nimonic 90, lining aluminium/Inconel<br />

600) [130]. The gasket consists of a toroidal lining made from<br />

the sealing material. It has tiny knife edges milled on the top and bottom circumference.<br />

Inside, as an elastic core, the torus contains a helical spring. This<br />

spring provides a homogenous compression all around the sealing circumference and<br />

avoids (torsion) stress on the sealed UHV components. A cross section of the Helicoflex<br />

∆ gasket is shown in Fig. 3.3 (not to scale). The helium leakage is specified<br />

as < 10 −10 mbar l/s [130].<br />

In the present setup, we have chosen an aluminium lining. Apart from the low<br />

costs, it offers advantageous properties in sealing our particular glass substrates.<br />

Aluminium is a ductile material: the knife edge is consumed un<strong>der</strong> compression and<br />

requires less loading <strong>for</strong>ce compared to a nonductile material. Among other ductile<br />

materials like silver or copper, aluminium gaskets require less loading <strong>for</strong>ce, whereas<br />

the specified final compression is even larger. Note the distinction made here between<br />

“(linear) loading <strong>for</strong>ce” and ”compression”. The <strong>for</strong>mer is <strong>der</strong>ived from applied<br />

torques when tightening flange bolts, the latter describes the visible geometrical<br />

de<strong>for</strong>mation of the toroidal gasket and the knife edges. A large compression of the<br />

gasket promises accurate control during the sealing procedure. Aluminium gaskets<br />

require a Vickers hardness of the sealing surfaces of 65 only, in contrast to minimal


40 Experimental setup<br />

Quartz cuvette<br />

compression nut<br />

sliding ring<br />

cushion ring<br />

compression ring<br />

(stationary)<br />

side view<br />

Helicoflex <br />

30 mm Al lining<br />

spring<br />

top view<br />

polished surface<br />

lath milled surface<br />

fine thread<br />

CF 40 knife edge<br />

hexagonal section<br />

Figure 3.3: All-metal UHV sealed fused silica cell.<br />

100 − 120 <strong>for</strong> silver and copper. A disadvantage may be the maximum bakeout<br />

temperature of 280 ◦ C, which is, however, still above the attempted rating <strong>for</strong> our<br />

cuvettes. For detailed requirements on machining and finish of the sealed surfaces,<br />

see Ref. [130]. As an alternative to the Helicoflex gasket, also a metal wire seal<br />

might offer a solution. However, common wires from gold need significantly more<br />

loading <strong>for</strong>ce and, the softer indium does not allow larger bakeout temperatures<br />

than the epoxy-glued connection, discussed below.<br />

Fig. 3.3(a) shows a cross section and a top view of the sealed cuvette. This<br />

corresponds to the most recent construction. In or<strong>der</strong> to provide a flat sealing<br />

surface, our workshop machined a CF 40 adapter to be connected to the hexagonal<br />

vapour cell section. This adapter accepts the Helicoflex gasket. The base of the<br />

cuvette assembly consists of a 50 mm diameter quartz disk of 20 mm thickness. In<br />

addition, between this disk and the rectangular cell, there is an intermediate quartz<br />

ring of 10 mm thickness, the outer diameter of which matches the 30 mm cross<br />

section of the cell. This assembly was clamped onto the Helicoflex by a combination<br />

of a stationary compression ring and a single compression nut, with a fine thread of<br />

72 mm diameter and a pitch of 1.2 mm/turn. A thin ring of annealed aluminium<br />

<strong>for</strong>ms a cushion between the compression ring and the quartz disk. A similar cushion,<br />

lubricated with MoS paste, serves as a sliding ring between compression ring and nut.<br />

The 10 mm spacer ring was inserted to avoid focused stress at the cell corners.<br />

Standard DN 25 Helicoflex ∆ gaskets were used. The inner and outer diameters<br />

are 30.4 mm and 40.2 mm, respectively. The torus cross section is 4.8 mm, with<br />

a nominal linear sealing load of 245 N/cm and a nominal compression of 0.9 mm.<br />

The load required to achieve sealing should thus be 100 times smaller than with the<br />

CF-sealed windows which were described in the previous section. The diameter of


3.4Optical access to the UHV system 41<br />

the knife edges is 35.2 mm. The larger inner diameter of the compression ring causes<br />

torsion stress in the quartz disk. Note that the choice of the standard DN 25 gasket<br />

was motivated earlier by modifying an existing CF 40 flange rather than machining<br />

a custom adapter to the vapour cell section.<br />

The final construction was tested while mounted directly on top of the inlet of<br />

the turbo-molecular pump. Sealing was achieved with a compression of 0.49 mm<br />

(55 % of the nominal one). The pressure reading from a Penning detector indicated<br />

4.0 × 10 −8 mbar be<strong>for</strong>e bakeout. After a bakeout cycle up to 230 ◦ C, the pressure<br />

settled at 1.8 × 10 −8 mbar, the same as when operating the terminated pump alone.<br />

Our “consumption” of numerous gaskets reflects their quality. Un<strong>der</strong> visual inspection<br />

both used and freshly unpacked gaskets occasionally showed tiny scratches or<br />

material faults in the aluminium knife-edges. In these cases sealing of the cuvette<br />

was not achieved within the nominal compression. It is strongly recommended to<br />

consume as many gaskets as necessary until sealing is achieved within the nominal<br />

compression. In fact, a glance at the catalogues [130] indicates that these gaskets<br />

are originally designed to seal reactor vessels rather than atomic vapour cells.<br />

The actual two-disk cuvette was motivated by an unsuccessful attempt to use the<br />

single-disk construction with a Helicoflex ∆ gasket. The 15 mm thick single disk was<br />

clamped on the gasket by 6 bolts of a modified CF 40 flange, similar to the viewport<br />

construction. After sealing was achieved, the system was gently baked at 70 ◦ C.<br />

Having cooled down slowly and after some hours of settling at room temperature, a<br />

crack at one of the cuvette’s corners occurred, similar to the crack with the previously<br />

used CF gasket. The reason was probably that the compression clamp finally made<br />

a consi<strong>der</strong>able wedge with the flange counterpart. The bolts had been tightened<br />

evenly by observing the applied torque. This suggests that either the spring load<br />

of the gasket was not constant along the circumference, or the torque readings<br />

of the wrench were not reliable due to variable bolt friction. This is the reason<br />

why we finally used a single screw terminal to control the compression rather than<br />

controlling the linear load by individual bolts. Nevertheless, the latter concept has<br />

been successfully realised by Dieckmann et al. [131], including bakeout above 250 ◦ C.


42 Experimental setup<br />

side view<br />

top view<br />

Glass cuvette<br />

42 mm<br />

TorrSeal<br />

(epoxy glue)<br />

stainless steel<br />

(304)<br />

weld<br />

CF 40<br />

knife edge<br />

Figure 3.4: UHV sealing of an epoxy-glued glass cell.<br />

3.4.3 Epoxy-glued glass cell<br />

We started experiments with an improvised glass cell, which we glued to a stainless<br />

steel rectangular plat<strong>for</strong>m, see Fig. 3.4. The low-vapour pressure epoxy resin<br />

was TorrSeal (Varian) which allows pressures down to 10 −9 mbar and bakeout<br />

temperatures up to 120 ◦ C. The glass cuvette is a standard “Large Cell” from<br />

HELLMA (Germany) and made from “Optical Glass” (B 270-Superwite crone glass,<br />

from DESAG). The outside dimensions are 130 × 42 × 42 mm 3 with a wall thickness<br />

of 4 mm.<br />

The figure shows the rectangular stainless steel plat<strong>for</strong>m, that was welded to<br />

a tubular CF 40 flange. The epoxy resin <strong>for</strong>ms a seam of triangular cross section<br />

along the bottom face of the cuvette. Thus, direct and polluting contact of the resin<br />

with the vacuum is kept small. The triangular steel edge is supported by a thin<br />

rim of steel. This proved to be necessary in or<strong>der</strong> to allow the seal to relax from<br />

stress after bakeout. The resin seemed to soften at bakeout temperatures and to<br />

relax stress that has been induced by the thermal expansion mismatch of steel and<br />

glass. When cooling down after bakeout the resin hardens too quickly. In a first<br />

construction without any significant elasticity, this resin property caused the glass<br />

to break at several locations at the epoxy seam, some hours after cooling down two<br />

room temperature.<br />

The ultimate pressure of 10 −9 mbar was reached after bakeout of the system,<br />

during which we kept the resin temperature below 115 ◦ C. In fact, this pressure<br />

was permissible <strong>for</strong> the experiments reported in this thesis. Thus, the glued cell<br />

proofed to be a low-cost, reliable concept, maybe even simpler and more robust<br />

than a cuvette that is assembled from loose glass plates as reported in Ref. [132].<br />

Of course, it was not possible to apply an optical AR coating to the inner surfaces,<br />

neither did we apply any coating at the outside.


3.5 Semiconductor lasers <strong>for</strong> cooling and trapping 43<br />

3.5 Semiconductor lasers <strong>for</strong> cooling and trapping<br />

3.5.1 Requirements<br />

There are three essential specifications of a laser system <strong>for</strong> atom-optical experiments,<br />

(i) frequency stability, (ii) optical output power and, (iii) beam quality,<br />

which are briefly discussed here, followed by a detailed description of our frequency<br />

stabilised diode lasers and injection-locked diode lasers.<br />

(i) Frequency stability.— The laser linewidth must be smaller than the atomic<br />

transition linewidth. The frequency should also be stable on this scale. Using rubidium<br />

(Γ/2π =6.0 MHz), this requirement is usually fulfilled with common laser<br />

sources. Frequency drift stability within 1MHz is achieved by “locking” the laser<br />

to an atomic resonance using feedback from a reference spectroscopy signal [133].<br />

Magneto-optical trapping and polarisation gradient cooling typically demand a laser<br />

detuning from the atomic resonance of a few times the transition linewidth Γ. Continuous<br />

and fast detuning control is achieved by frequency shifting acousto-optical<br />

modulators (AOM). Passive drift stability of the laser source is desirable when large<br />

detuning (δ ≫ Γ) is necessary and suitable references <strong>for</strong> locking are not available,<br />

e.g., when working with far off-resonance dipole potentials. In some applications also<br />

the spectral background has to be consi<strong>der</strong>ed. In Chap. 4, the amplified spontaneous<br />

emission background (ASE) of a tapered amplifier system (TA) is discussed.<br />

(ii) Output power.— In applications with near-resonance light, laser intensities<br />

of a few times the saturation intensity are usually sufficient, e.g. I 0 =1.67 mW/cm 2<br />

<strong>for</strong> the rubidium D2 line. A laser output of 15 mW allows operating a rubidium MOT<br />

with beam waists 5 mm. Much more power is usually needed to realise optical<br />

dipole potentials. The trapping scheme envisaged in Chap. 2 requires intensities<br />

∼ 10 6 I 0 . In this case, laser power constitutes the limiting factor to the spatial<br />

extension of the trapping potential.<br />

(iii) Beam quality.— Most applications demand good beam quality and a well<br />

defined polarisation. For this reason single-mode optical fibres are used as spatial<br />

filters. In case of the TA system, such a fibre also provides spectral filtering of ASE<br />

background in the amplifier output.<br />

3.5.2 Compact external grating diode lasers<br />

A single-transverse-mode laser diode emits a diffraction-limited, elliptical beam.<br />

The emission linewidth of such a laser is typically several tens of MHz. The emitted<br />

centre frequency is determined by both the internal cavity <strong>for</strong>med by the reflective<br />

waveguide facets and the spectral gain profile. Between “mode hops”, it can be<br />

continuously tuned by means of operating temperature and injection current.<br />

The most common technique to narrow the linewidth of a diode laser to below<br />

100 kHz is optical feedback by the first diffraction or<strong>der</strong> from a grating, see Fig. 3.5<br />

(“external grating diode laser”, EGDL). The grating establishes an external cavity,<br />

while the specular reflection is coupled out. Simultaneous control of the grating angle


44 Experimental setup<br />

EGDL<br />

HW<br />

AP<br />

OI<br />

G<br />

l<br />

A<br />

<br />

OC<br />

LD<br />

Figure 3.5: External grating diode laser in Littrow configuration. A rotation ∆α of the<br />

grating (G) around an axis (A) causes a simultaneous displacement ∆l; laserdiode(LD),<br />

output collimator (OC), half-wave plate (HW), anamorphic prism pair (AP), and optical<br />

isolator (OI).<br />

and distance, and of the diode current allows tuning and locking of the laser to a<br />

reference frequency [117, 118]. In atomic, physics this method provides a standard<br />

laser tool, reaching from the mid-infrared [134] to the red [135, 136] and, recently,<br />

to the blue [137] range of the spectrum.<br />

Spectroscopic applications usually demand a wide tuning range, free of modehops<br />

and covering various atomic or molecular resonances. A common realisation of<br />

an EGDL is the Littrow configuration (see e.g. Ref. [133]). The rotation axis of the<br />

grating is chosen such, that a change in feedback frequency ∆ω, due to a rotation<br />

∆α, is matched with the grating displacement ∆l, or(∂ω/∂α)∆α =(∂ω/∂l)∆l.<br />

Simultaneous modulation of the laser diode current provides continuous tuning over<br />

tens of GHz.<br />

Efficient coupling of the laser output to a single-mode optical fibre requires the<br />

elliptical beam profile to be circularised. This is done immediately after the EGDL<br />

by a pair of anamorphic prisms. Since the Brewster effect assists in reducing reflection<br />

losses, the laser polarisation is first rotated to horizontal by means of a half-wave<br />

plate. The resulting circular beam typically has a waist of 0.5 mm (1/e 2 intensity<br />

radius). An EGDL demands good optical isolation against backreflections. In most<br />

situations an isolation of 30 dB is sufficient. However, if the EGDL is used as “master”<br />

oscillator to seed an amplifier or an injection-locked “slave” laser [138, 139],<br />

60 dB isolation may be required to prevent direct feedback to the master laser from<br />

the mode-matched slave output, see below.<br />

An external cavity makes the system susceptible to vibrations and thermal drift.<br />

Various designs use compact realisations of the Littrow type to improve laser stability<br />

[135,140]. Other concepts put emphasis also on economical usage of commercial<br />

opto-mechanical components [141, 142]. In the experiments reported in this thesis,<br />

our interest was in locking lasers to a single atomic resonance rather than a wide continuous<br />

tuning range. Hence, a very compact EGDL design was chosen, based on a<br />

stimulating idea from Poul Jessen [142]: the grating angle is preset manually with a


3.5 Semiconductor lasers <strong>for</strong> cooling and trapping 45<br />

horizontal tilt<br />

fine gear<br />

heating<br />

resistor<br />

clamp<br />

(lens tube)<br />

collimation<br />

lens tube<br />

laser beam<br />

horizontal<br />

flexure<br />

1cm<br />

vertical flexure<br />

horizontal tilt<br />

coarse<br />

vertical tilt<br />

(coarse)<br />

grating<br />

PZT<br />

Two-stage flexure gear<br />

d<br />

travel 0.5 mm<br />

Grating mount<br />

grating<br />

glue<br />

action 70<br />

m<br />

d’<br />

d’’<br />

d = 2.6 d’ = 7.0 d’’<br />

low voltage<br />

PZT stack<br />

mounting<br />

ring<br />

Figure 3.6: Compact external grating diode laser.<br />

resolution ∼ 100 MHz and kept fixed in experiments. Fine adjustment is per<strong>for</strong>med<br />

by tuning the grating distance from the laser facet with a piezo actuator (PZT).<br />

The construction of the laser head is shown in Fig. 3.6. A laser diode (TO-5<br />

window package, 9 mm dia.) is mounted in a collimation lens tube (ThorLabs,<br />

type LT 230 B, f =4.5 mm, N.A.=0.55). The beam can thus conveniently be collimated<br />

be<strong>for</strong>e the lens tube is mounted in the laser head. A gold coated holographic<br />

grating with 1800 lines/mm provides feedback to the laser diode (Carl Zeiss Jena,<br />

no. 263232-9451-325, 10 × 10 × 6mm 3 ). The feedback angle <strong>for</strong> 780 nm and 795 nm<br />

wavelength is α =44.6 ◦ and 45.7 ◦ , respectively. In a similar construction also gratings<br />

with 1200 lines/mm were used (Zeiss, no. 263232-9052-825). The feedback angle<br />

was there ≈ 19 ◦ . However, these gratings provided significantly less output power.


46 Experimental setup<br />

Both collimation package and grating are integrated in a single, milled block<br />

from copper-bronze. The rotational degrees of freedom <strong>for</strong> the grating are provided<br />

by a flexure construction. Adjustment of optical feedback (coarse vertical tilt) is<br />

achieved with a small screw using an Allan key. The feedback wavelength (horizontal<br />

tilt α) is coarsely set with a screw also at the flexure mount, whereas fine<br />

adjustment is achieved using an additional double-stage flexure gear. Not shown in<br />

the drawing is an AD 590 temperature sensor, attached to the base of the laser head,<br />

close to a Peltier thermo-electric cooler (TEC). The small volume of the laser head<br />

allows relatively fast temperature control and provides good thermal drift stability<br />

to the external cavity. In particular, the flexure grating mount has better thermal<br />

conductance than a comparable spring-loaded construction using ball-bearings. The<br />

choice of copper-bronze (7 % Sn) is a compromise between thermal and elastic properties<br />

[140]. The typical passive stability of this system is ∼ 100 MHz per hour and<br />

limited by both thermal drift and drift of the PZT.<br />

The laser head is mounted via the TEC on a brass heatsink that can be water<br />

cooled and also includes a compartment <strong>for</strong> laser current modulation circuitry.<br />

Vibrations from the optical table are damped by a polymer sheet un<strong>der</strong>neath the<br />

heatsink (Edmund Scientific, Sorbothane) The laser head is shielded from surrounding<br />

airflow and from electro-magnetic noise by a metallised cap. This cap allows<br />

access to the fine adjustment gear.<br />

The flexure gear.— In the drawing of the double-stage gear, the concentric<br />

pairs of large and small circles indicate the motion of parts from the gear, or the<br />

translations d ↦−→ d ′ and d ′ ↦−→ d ′′ , respectively. The total translation is 7:1. A<br />

single turn of the screw (d =0.5 mm) results in a travel of d ′′ =70µm. The length<br />

change of the external cavity is ∆l ≈ 0.3 d ′′ ,or20µm per turn of the screw. The<br />

frequency change of the cavity is ≈ 20 GHz/µm. (With a cavity length of ≈ 20 mm,<br />

the free spectral range is FSR≈ 7.5 GHz.) This resolution is sufficient to allow<br />

smooth manual presetting of the laser frequency. Essential is, that the user touches<br />

an actuator fixed to the bulk of the laser head rather than to the vibrationally<br />

sensitive grating hol<strong>der</strong>.<br />

The PZT stack actuator.— The grating is directly glued to a polymer-molded<br />

low-voltage PZT stack actuator (Piezomechanik, bare actuator type PSt 150/7/7,<br />

travel 7 µm <strong>for</strong> 150 V). The length or frequency tuning of the cavity is given by<br />

≈ 30 nm/V, or ≈ 0.6 MHz/mV, respectively. The only moving mass is that of the<br />

grating. This allows a faster response when tuning the cavity length as compared<br />

with a rotational grating mount. A prestressed PZT device in a tubular steel case<br />

has also been tested (PSt 150/5/7 VS10), in first instance promising a more linear<br />

response. Un<strong>for</strong>tunately, the weight of the horizontally mounted grating bent the<br />

PZT. This resulted in friction with the steel case, causing an unreliable actuator<br />

response [Pickelmann, Piezomechanik, private communication].


3.5 Semiconductor lasers <strong>for</strong> cooling and trapping 47<br />

85 Lasers<br />

Rb (I=5/2)<br />

87<br />

Rb (I=3/2)<br />

5p 2 P3/2<br />

( =26.2 ns)<br />

121<br />

63<br />

29<br />

5p 2 P1/2 362<br />

( =27.7 ns)<br />

F<br />

3<br />

2<br />

1<br />

5s 2 3<br />

S1/2 3036 MHz<br />

2<br />

4<br />

3<br />

2<br />

D2 (780.2 nm)<br />

D1 (795.0 nm)<br />

F<br />

3<br />

267<br />

2<br />

157<br />

1<br />

72<br />

0<br />

2<br />

812<br />

1<br />

2<br />

6835 MHz<br />

1<br />

2<br />

3<br />

4<br />

5 6 7<br />

1<br />

Figure 3.7: Hyperfine structure of rubidium [143, 144].<br />

(i) Near resonance cooling, probing and hyperfine pumping:<br />

beam line F g → F e detuning fibre output<br />

(1) D2 2 → 3 0 − (±8) Γ 0 − 500 µW probing on cycling transition<br />

(2) D2 2 → 2 0 − (±8) Γ 0 − 500 µW hyperfine depumping to F g =1<br />

(3) D2 2 → 3 (−10) − 0Γ 20 mW MOT, molasses cooling<br />

(4) D1 1 → 2 resonant 10 mW hyperfine repumping to F g =2<br />

(ii) Far off-resonance dipole potentials:<br />

(5) D1 1 → 2 ±2nm 120 mW inelastic EW mirror,<br />

dark state trapping<br />

(6) D2 2 → 3 ±2nm 200 mW elastic EW mirror,<br />

atom guiding, trapping<br />

(7) D2 1 → 2 ±2nm 200 mW inelastic EW mirror<br />

Table 3.1: Laser frequencies used in experiments with 87 Rb.


48 Experimental setup<br />

3.5.3 The laser park <strong>for</strong> atom-optical experiments<br />

An experiment with rubidium requires various frequency-stabilised laser sources<br />

tuned to specific optical resonances of the consi<strong>der</strong>ed isotope, here 87 Rb. Fig. 3.7<br />

shows the hyperfine energy levels of the rubidium D1and D2 line. The optical<br />

transitions, labelled (1)−(7), indicate the corresponding laser frequencies in our<br />

setup of diode laser systems. The specific usage of the lasers is listed in Table 3.1.<br />

An overview of this setup is given in Fig. 3.8. It is specified essentially by two<br />

groups of lasers, (i) EGDL’s tuned close to a rubidium resonance, some of which are<br />

amplified by an “injection-locked” laser diode, and (ii) high-power tapered amplifiers<br />

used far off-resonance (see Chap. 4).<br />

(i) Cooling, probing and optical hyperfine pumping.— The laser frequencies<br />

of the beams (1) − (3) are <strong>der</strong>ived from an EGDL that is stabilised by feedback<br />

from a frequency modulation (FM) spectroscopy on rubidium. The EGDL serves<br />

as a master oscillator <strong>for</strong> an injection-locked slave diode laser (ILDL), providing<br />

the trapping and cooling light of beam (3). The laser diodes used <strong>for</strong> the EGDL’s<br />

were 60 mW single-spatial and -frequency mode laser diodes with a specified wavelength<br />

close to the D1or D2 line (Hitachi, HL 7851G98, selected 781− 785 nm;<br />

Mitsubishi, ML 64114R, selected 788 − 793 nm). Recently also an 80 mW device<br />

became available (Sanyo, DL-7140-001, specified 785 nm).<br />

The optical output power of our EGDL’s ranges between 5 − 30 mW after the<br />

isolator, depending on the used laser diode. A small fraction (∼ 0.5 mW) is split<br />

off to be used <strong>for</strong> spectroscopy. The lasing frequency mode of both master and<br />

slave can be permanently monitored by an optical spectrum analyser. The seeding<br />

beam from the master laser was inserted into the slave’s beam path by means of the<br />

accessible output polariser of a 30 dB optical isolator (Gsänger, single-stage type<br />

FR 780). Perturbing feedback from the slave to the master is thus prevented by this<br />

isolator, in addition to the 60 dB isolator directly after the EGDL. Mode matching<br />

of master and slave was achieved using identical beam collimation and circularising<br />

optics, see Fig. 3.5. If optimally aligned, a seeding input of 100 µW wassufficientto<br />

provide a stable locking range over > 10 Γ, as required to load a MOT and per<strong>for</strong>m<br />

molasses cooling.<br />

Fast and continuous frequency control of the beams (1) − (3) was achieved using<br />

acousto-optic modulators (AOM) in double-pass, see below. For the beams (1) and<br />

(2), the AOM also serves as a power modulator and a shutter.<br />

Figure 3.8: The system of stabilised lasers and amplifiers (previous page). Injectionlocked<br />

diode laser (ILDL), tapered amplifier (TA), half-wave plates (HW), optical spectrum<br />

analysers (SA), grating spectrometer (GS) and wavelength meter (WM). The frequency<br />

modulation (FM) and Zeeman polarisation (ZS) spectroscopy schemes are indicates<br />

symbolically (Rb). Frequency shifting AOM’s in double-pass are explained in detail below:<br />

lens (L), quarter-wave plate (QW), diaphragm (D), and mirror (M). The laser frequency<br />

ω L is shifted by twice the RF frequency of the acoustic wave, ω RF .


3.5 Semiconductor lasers <strong>for</strong> cooling and trapping 49<br />

(i)<br />

Near-resonance cooling, probing, pumping<br />

EGDL<br />

isolator<br />

60 dB<br />

HW<br />

HW<br />

A<br />

M O<br />

shutter<br />

(1) probe<br />

FM<br />

feedback<br />

SA<br />

spectroscopy<br />

Rb<br />

A<br />

M O<br />

(2) depump<br />

ILDL<br />

injection lock<br />

30 dB<br />

AOM<br />

EOM<br />

(3)<br />

trap/cool<br />

MOT (D2)<br />

30 dB<br />

EGDL<br />

FM<br />

Rb<br />

optical fibre<br />

(4) repump (D1)<br />

(ii) Far off-resonance dipole potentials<br />

ZS<br />

EGDL<br />

60 dB 60 dB<br />

Rb<br />

A<br />

M O<br />

TA<br />

GS<br />

WM<br />

(5), (6), (7)<br />

bounce/trap<br />

(D1,D2)<br />

AOM frequency control:<br />

A<br />

M O<br />

M<br />

D<br />

QW<br />

f<br />

L<br />

f<br />

<br />

RF<br />

<br />

L<br />

<br />

L<br />

+2<br />

RF


50 Experimental setup<br />

The EGDL of beam (4) provides hyperfine repumping light to transfer atoms<br />

from F g =1toF g = 2, mainly <strong>for</strong> operating the MOT and cooling but also <strong>for</strong><br />

specific probing techniques, see Chap. 7.<br />

All beams were coupled to single-mode optical fibres, as indicated <strong>for</strong> beam<br />

(4). Coupling efficiencies were achieved, ranging between 70 − 85 % <strong>for</strong> circularised<br />

beams of single-mode laser diodes, using compact fibre coupling ports, designed <strong>for</strong><br />

beam input diameters between 0.9−1.8 mm (OFR, type PAF-X-5-780). These ports<br />

were used with standard fibre patchcords with angle-polished fibre connectors (type<br />

FC/APC). This avoids etalon effects from reflections at the fibre facets. The fibres<br />

were not polarisation conserving. However, twisting the fibres in loops and fixing<br />

them to the optical table provided arbitrary polarisation control of the output beam.<br />

(ii) Far off-resonance dipole potentials.— High power output up to 200 mW<br />

from a single-mode fibre is achieved in two systems using tapered semiconductor<br />

amplifiers (TA). Only one scheme is sketched in Fig. 3.8. One system provides beam<br />

(5), the other provides the beams (6) and (7). The gain elements are each seeded by<br />

a well isolated EGDL. A detailed characterisation of these systems is given in the<br />

next chapter.<br />

For near-resonance applications, the EGDL can be frequency stabilised by a Zeeman<br />

polarisation spectroscopy (ZS). A tunable frequency offset between ±500 MHz<br />

from the referenced atomic resonance is achieved using an AOM. For larger detunings,<br />

the EGDL remains unlocked and the frequency can be monitored by means<br />

of an optical spectrum analyser, a wavelength meter (Coherent, <strong>Wave</strong>Mate), or a<br />

grating spectrometer (Ocean Optics, PC 2000).<br />

Laser frequency tuning by acousto-optical modulators.— Laser frequencies<br />

are shifted using acousto-optic modulators in double-pass, also shown in Fig. 3.8.<br />

A lens (L) of focal length f (between 10 − 20 cm) and a mirror (M) <strong>for</strong>m a folded<br />

telescope. After the first passage of the AOM, the Bragg-deflected beam is retroreflected<br />

and collimated again be<strong>for</strong>e passing the AOM a second time. All light<br />

but the selected diffraction or<strong>der</strong> is blocked by a diaphragm (D). The polarisation<br />

of the retro-reflected beam has been turned by 90 ◦ , passing twice a quarter-wave<br />

plate. The light is coupled out by a polarising beam splitter cube. Using Bragg<br />

deflection in the first diffraction or<strong>der</strong>, the light un<strong>der</strong>goes a net frequency shift<br />

of twice the RF modulation frequency ω RF , with no net deflection or frequency<br />

dependent displacement. This property is particularly important when shifting the<br />

frequency of the master laser in an injection-locking scheme or when coupling light<br />

to an optical fibre. Both cases require excellent directional beam stability.<br />

A typical double-pass efficiency is 50 % in first diffraction or<strong>der</strong>. Higher or<strong>der</strong>s are<br />

not practical due to their low efficiency. Our modulators have PbMoO 4 crystals and<br />

accept random polarisation [A.A. Opto-Electronique, type AA.SP.200/B100/A0.5-<br />

ir (ω RF = 200 ± 50 MHz) and AA.MP.25-IR (110±30 MHz); Isomet, type 1205 C<br />

(80±15 MHz)]. It is recommended to use linear polarisations only, since birefringence<br />

of the crystal together with varying RF load may lead to severe thermal drift<br />

in the diffracted beam power.


3.5 Semiconductor lasers <strong>for</strong> cooling and trapping 51<br />

In the laser setup of Fig. 3.8(i), AOM’s are used to <strong>der</strong>ive the required frequencies<br />

from the master laser that is locked to the (bf) cross-over spectroscopy signal of the<br />

F g =2−→ F e = {1, 3} transitions, see Fig. 3.9. The cross-over is centred between<br />

these transitions. Hence, a blue shift of 212 MHz realises the resonant probe (1) on<br />

the cycling transition F g =2−→ F e = 3, using an AOM with 110 MHz specified<br />

centre frequency. By a red shift of 133 MHz, the depumping beam (2) on the open<br />

transition F g =2−→ F e = 2 is realised. The injection-locked slave is supplied with<br />

a shifted seed beam, thus saving power from the slave <strong>for</strong> the experiment. Using<br />

an AOM with 80 MHz centre frequency, beam (4) is thus tuned 0 − 10 Γ to the red<br />

of the F g =2−→ F e = 3 transition. The collimation of the (astigmatic) master<br />

laser beam was optimised <strong>for</strong> mode-matching the slave laser. This resulted in poor<br />

beam quality and thus poor AOM efficiencies in beam (1) and (2). In a later stage, a<br />

second injection-locked slave laser supplied these beams with more power and better<br />

beam quality.<br />

When using an AOM as a switch or power modulator, “leakage” into the selected<br />

diffraction or<strong>der</strong> reduces the extinction to typically 1: 1000. There<strong>for</strong>e we use also<br />

mechanical shutters. Power modulation with an extinction of 1: 200 is obtained <strong>for</strong><br />

beam (3) by an electro-optical modulator. [Gsänger, type LM 0202 5W IR, aperture<br />

3 × 3mm 2 . We use also a version of the LM 0202 5W IR with 5 × 5mm 2 aperture.]<br />

3.5.4Laser frequency stabilisation<br />

The lasers were locked to rubidium resonances using Doppler-free saturation spectroscopy<br />

[133]. Rubidium is commonly used as a saturated vapour in spectroscopy<br />

cells at room temperature. Most spectroscopy schemes provide absorptive signals<br />

(“dips”), resolving the natural transition linewidth Γ. It is necessary to <strong>der</strong>ive a<br />

dispersive signal with a zero-crossing as feedback to the laser. Three common techniques<br />

are:<br />

• Frequency modulation (FM) of the laser creates RF sidebands [119]. A dispersive<br />

signal is obtained by mixing the spectroscopy signal with the local RF<br />

oscillator and adjusting the phase.<br />

• Zeeman spectroscopy employs nondegenerate magnetic sublevels. “Dispersion”<br />

signals are electronically generated from oppositely frequency-shifted<br />

absorptive signals of orthogonal polarisations [120–122].<br />

• Polarisation spectroscopy [133,145] probes the dispersion of the atomic species<br />

rather than the absorption, and a feedback signal is directly obtained. This can<br />

also be used in passive schemes, relying on purely optical feedback [146, 147].<br />

FM spectroscopy has an intrinsically large bandwidth providing fast feedback to<br />

the laser, with good distinction between neighbouring optical transitions. However,<br />

this technique is relatively complex due to RF electronics. More important, the FM<br />

sidebands imprinted onto the laser output may perturb the laser application. (This<br />

could be avoided by using an EOM to modulate only the light used <strong>for</strong> spectroscopy.)


52 Experimental setup<br />

We employ the FM technique there<strong>for</strong>e as a robust locking scheme <strong>for</strong> less sensitive<br />

tasks, such as the MOT, optical pumping or probing atoms. Zeeman spectroscopy<br />

is less demanding in electronics and optics equipment. We use it, with the far offresonance<br />

lasers, were the mo<strong>der</strong>ate accuracy of the artificially dispersive locking<br />

signal is not an issue. In the following, a brief description of both methods is given.<br />

Frequency modulation spectroscopy.— The FM scheme is shown in Fig. 3.9.<br />

The optical part is based on Doppler-free saturation spectroscopy: Light is split<br />

off from the output of an EGDL and sent in a first pass through a spectroscopy<br />

cell with rubidium vapour. If resonant within the Doppler-broadened absorption<br />

profiles, rubidium optical transitions are saturated. When the laser scans across<br />

a resonance, the retro-reflected beam probes these transitions. The Doppler effect<br />

cancels out on a resonance and an absorptive signal with a width ∼ Γ is recorded.<br />

Optimal retro-reflection, i.e. Doppler cancelling is achieved using a quarter-wave<br />

plate and a polarising beam splitter cube.<br />

A dispersive signal is achieved by modulation of the diode laser current, I, with<br />

a local oscillator radio frequency, here ω RF = 40 MHz. This results in frequency<br />

sidebands, ω L ± ω RF , next to the laser carrier, ω L , shown in the inset of the figure.<br />

Both sidebands beat with the carrier. If no spectral atomic feature is covered by<br />

any of these laser frequencies, the net beating cancels out, due to the opposite phase<br />

of the sidebands. The photodetector then receives no signal ∝ ω RF . However, if<br />

one of the frequencies probes a resonance, the beating is out of balance and the<br />

photodetector detects an RF signal. (The detector is supplied with a RF bandpass<br />

filter.) By amplification of this signal, adjusting the phase, and mixing with a local<br />

oscillator, the dispersive (low frequency) signal <strong>for</strong> the laser lock is obtained and fed<br />

back to both, laser current and grating actuator. The grating feedback tackles down<br />

slow drifts of the laser using a longer integration time constant than the intrinsically<br />

fast current feedback.<br />

Two exemplary FM spectra are plotted in the figure. For the (f) transition also<br />

the sidebands are resolved. Typical <strong>for</strong> this type of Doppler-free spectroscopy is<br />

the occurrence of so-called “cross-over” resonances, given the Doppler broadened<br />

absorption profiles of several resonances overlap. This is the case with room temperature<br />

rubidium vapour. The cross-over resonance of two transitions is located at<br />

the average transition frequency. (The spectra in the figure show that the overlap<br />

of the Doppler profiles is larger <strong>for</strong> the D2 line.) In our laser setup we locked the<br />

EGDL <strong>for</strong> the beams (1)−(3) to the (bf) cross-over.


3.5 Semiconductor lasers <strong>for</strong> cooling and trapping 53<br />

OI<br />

HW<br />

<br />

EGDL<br />

RF<br />

0<br />

<br />

L<br />

<br />

FM<br />

I<br />

PZT<br />

QW<br />

Rb<br />

PD<br />

RF<br />

LO<br />

DC<br />

RF<br />

<br />

D1 line<br />

(795.0 nm)<br />

b) Fg<br />

=1 Fe<br />

=1<br />

d) 1 2<br />

D2 line<br />

(780.2 nm)<br />

b) Fg=2 Fe=1<br />

d) 2 2<br />

f) 2 3<br />

b<br />

d<br />

arb. units<br />

(bd)<br />

-1.0 -0.5 0.0 0.5 1.0<br />

relative frequency (GHz)<br />

arb. units<br />

b<br />

(bd)<br />

d<br />

(bf)<br />

(df)<br />

0.0 0.2 0.4 0.6 0.8<br />

relative frequency (GHz)<br />

f<br />

Figure 3.9: Laser stabilisation by FM spectroscopy. Scheme: Half-wave plate (HW),<br />

double pass through a rubidium cell (Rb), outcoupling using a quarter-wave plate (QW)<br />

and a polarising cube, detection with a photodiode (PD). Inductive frequency modulation<br />

of the laser current I using a RF oscillator. The photodiode signal is phase shifted (φ)<br />

and mixed ( ⊗ ) with the local oscillator (LO). The resulting dispersive DC signal is fed<br />

back to laser current and grating actuator (PZT). Inset: Laser carrier frequency ω L and<br />

FM sidebands ω L ± ω RF , spectral feature of natural linewidth Γ at atomic resonance ω 0 .<br />

Graphs: FM signals of 87 Rb (see Ref [148]), labelling as in Ref. [143]. For comparison:<br />

absorption signals from a DC photodiode (thin curves).


54Experimental setup<br />

(a)<br />

QW<br />

PB<br />

(b)<br />

Rb -<br />

-<br />

+ +<br />

- +<br />

-<br />

B<br />

o<br />

+1<br />

(c)<br />

0<br />

-1<br />

F =1<br />

- - +<br />

e<br />

<br />

- +<br />

- + <br />

+1<br />

F g =1<br />

<br />

m g=0<br />

-1<br />

<br />

Figure 3.10: Laser stabilisation by Zeeman spectroscopy. (a) Doppler-free saturation<br />

spectroscopy: rubidium cell (Rb) with axial magnetic field (B), quarter-wave plate (QW),<br />

polarising beam splitter cube, and photodetectors <strong>for</strong> σ ± polarisation. (b) Exemplary<br />

Zeeman shift of magnetic sublevels: the resonance is blue (red) shifted <strong>for</strong> σ + (σ − )polarised<br />

light. (c) Zeeman-shifted resonances of both circular polarisations. The dispersive<br />

signal obtained by subtraction.<br />

Zeeman spectroscopy.— The scheme <strong>for</strong> this method is shown in Fig. 3.10 (see<br />

also Ref [148]). It uses the decomposition of linearly polarised light into circular<br />

polarisations, σ ± . In Doppler-free saturation spectroscopy, rubidium vapour is made<br />

birefringent by using the Zeeman shift of magnetic sublevels in an axial, homogeneous<br />

magnetic field. There<strong>for</strong>e, the orthogonal circular polarisations encounter<br />

oppositely shifted resonances. Both polarisations are detected independently using<br />

a combination of a quarter-wave plate and a polarising cube as an analyser <strong>for</strong> the<br />

circular polarisation basis. A dispersive laser-lock signal is obtained electronically.<br />

This method requires an optical transition scheme with different g-factors, i.e. different<br />

Zeeman shifts in the ground and excited state sublevels. For example, <strong>for</strong> the<br />

D1-line F g =1−→ F e = 1transition this are g g =9/4 andg e =3/2, respectively.


3.6 Real-time experimental control 55<br />

3.6 Real-time experimental control<br />

An atom-optical experiment constitutes a series of processes in quick succession,<br />

demanding real-time application of analogue and digital control signals. Data acquisition<br />

(DAQ) also requires precise triggering with µs-resolution. A typical experimental<br />

sequence consists of loading a MOT, cooling atoms in optical molasses, releasing<br />

them <strong>for</strong> bouncing on an evanescent-wave mirror and, finally, imaging them<br />

with a CCD camera. Laser beams must be switched on time scales of typically<br />

0.1ms. We employ a common personal computer (PC), that operates LabVIEW<br />

un<strong>der</strong> Windows NT, to do both real-time control and data acquisition. This provides<br />

a flexible system with various software-controlled input and output channels. It can<br />

be configured <strong>for</strong> arbitrary time sequences. These tasks are per<strong>for</strong>med by several<br />

hardware extension cards. In particular, a self-sustaining digital signal processor<br />

(DSP) per<strong>for</strong>ms the real-time control of digital output (trigger) channels, thus circumventing<br />

perturbing interrupts of the PC processor. Table 3.2 gives an overview<br />

on the various hardware components.<br />

PC plat<strong>for</strong>m and DAQ hardware.— The system is based on a PC with Pentium<br />

II processor. An analogue output board (AT-AO-10) controls the modulation<br />

of AOM and EOM drivers. A general purpose DAQ board (MIO-16E-4) provides<br />

analogue inputs, which are used, e.g., to acquire photodetector signals <strong>for</strong> time-offlight<br />

measurements. This board has additionally two wave<strong>for</strong>m output channels<br />

and two general purpose counters. There<strong>for</strong>e, it is also used as a versatile function<br />

and pulse generator. Two more slots of the PC are occupied by the DSP (DIO-128)<br />

and by the interface (ST-138) of a digital camera system.<br />

The fluorescence of trapped atoms is permanently monitored by several analogue<br />

video cameras. These cheap surveillance cameras (Conrad Electronic) havenonearinfrared<br />

blocking filters and are thus sensitive to the rubidium fluorescence. The<br />

video signals can be recorded by a framegrabber (FlashBus), e.g., <strong>for</strong> beam profiling<br />

tasks or assisting laser beam alignment in the UHV vapour cell. A grating spectrometer<br />

(PC 2000) allows monitoring laser wavelengths. The framegrabber and the<br />

spectrometer are operated by a second PC.<br />

Digital signal processor and LabVIEW user interface.— The main task of<br />

the DSP is to provide precise timing during the experiments. In a screen interface,<br />

the user fills in a time schedule of the experimental events. This record contains<br />

the possibly altered status of digital and analogue output ports <strong>for</strong> a given event,<br />

including the time of the event with 1 µs resolution <strong>for</strong> the DSP timer. Using a<br />

LabVIEW driver, the DSP loads the time table and the digital output record into<br />

its on-board memory. The analogue output record is buffered in the PC’s memory<br />

and handed over on request to the FIFO-buffered analogue output board by the<br />

DMA (“direct memory access”) controller. All input and output channels are experimentally<br />

accessible through a front-end connector panel. One digital output of<br />

the DSP supplies a hardware event-update trigger to the analogue output board.<br />

Other digital outputs provide modulation signals <strong>for</strong> AOM/EOM drivers, magnetic


56 Experimental setup<br />

field coil current supplies and mechanical shutter drivers. CCD image capture and<br />

input of photodetector signals are triggered similarly. The DSP works independently<br />

from the PC. Thus, other LabVIEW routines can be used on the PC to acquire measurement<br />

data. The LabVIEW interface <strong>for</strong> the experiments discussed in this thesis<br />

was programmed in a simple and effective way, mainly using exemplary routines<br />

from the DSP driver library. Meanwhile, we use a commercially available program<br />

that has been developed by H. Alberda (AMOLF Institute, Amsterdam) and is very<br />

convenient in use, including also DAQ functions.<br />

CCD digital camera imaging system.— An imaging system <strong>for</strong> (cold) atoms<br />

must have an accurate image capture trigger and a well defined exposure time.<br />

Mechanical shutters are usually too slow. There<strong>for</strong>e we use a frame-transfer system<br />

(Princeton Instruments). Half of the CCD array is covered by a mask. After an<br />

exposure, the image is shifted within 1.6 ms un<strong>der</strong> the mask to be shielded against<br />

further illumination and is read out. After shifting, the CCD is ready <strong>for</strong> another<br />

image capture. If masking a larger area of the sensor (1024 × 512 pixels in total),<br />

an even faster sequence of more than two image frames, although smaller in size,<br />

can be captured (cf. Refs. [149, 150]). An advantage of the frame transfer <strong>for</strong> our<br />

experiments is, that the sensitive area can be kept “clean” (unexposed) by means of<br />

continuous line shifting, until ∼ 2 ms be<strong>for</strong>e an image capture. This is particularly<br />

important, if an image is taken only a few ms after a strong (saturating) illumination<br />

source, e.g. an evanescent-wave, has been switched off.<br />

Other important CCD specification are the spectral sensitivity, the pixel size, the<br />

pixel filling ratio, the pixel well depth (electron capacity), and the noise properties<br />

(dark current). These specifications are discussed in Ref. [151]. We may expect a<br />

strong background illumination while imaging optically trapped atoms. There<strong>for</strong>e,<br />

the well depth of the pixels must be sufficiently deep ( 10 5 electron charges) to<br />

avoid saturation, and the resolution of signal digitisation should be at least 10 bits.<br />

It is also this expected background illumination why we don’t use an intensified<br />

imaging system.<br />

There exist also “interline transfer” systems, which have read-out registers between<br />

adjacent pixel lines. This allows <strong>for</strong> even faster cleaning, shuttering and read<br />

out. However, the pixel filling ratio and the well depth of these CCD’s are low.<br />

We use our CCD system with either a commercial 50 mm camera objective to<br />

capture fluorescence images of bouncing atoms (see Chap. 6) or with a relay telescope<br />

to do absorption imaging (see Chap. 7).


3.6 Real-time experimental control 57<br />

Host system:<br />

Personal computer<br />

Pentium II, 300 MHz, 256 MB RAM;<br />

Windows NT 4.0<br />

Programming LabVIEW 5.1,<br />

user interface by H. Alberda, AMOLF Institute, Amsterdam<br />

Experimental control:<br />

Real-time control, Viewpoint Software Solutions, DIO-128 (PCI-bus),<br />

digital output Dynamic Digital I/O System,<br />

64 inputs/64 outputs (128 inputs),<br />

timer 32 bit, resolution 1 µs;<br />

LabVIEW driver library<br />

Analogue output National Instruments, AT-AO-10 (ISA-bus),<br />

10 channels, resolution 12 bit, max. sampling 300 kS/s<br />

DAQ, imaging:<br />

Analogue input,<br />

counters,<br />

wave<strong>for</strong>ms<br />

Digital imaging<br />

Video cameras<br />

Framegrabbing<br />

National Instruments, MIO-16E-4 (PCI-bus),<br />

16 single-ended (8 differential) analogue input channels,<br />

resolution 12 bit, max. sampling 300 kS/s,<br />

2 general purpose counters (24 bit),<br />

2 wave<strong>for</strong>m analogue output channels (12 bit)<br />

Princeton Instruments/Roper Scientific,<br />

TE/CCD-512 EFT frame transfer digital camera system,<br />

sensor EEV 37, grade 1, 512 × 512 pixel,<br />

pixel size 15 × 15 µm, 100 % pixel filling,<br />

shift time 1.6 ms/frame,<br />

dark current 11 e − /pixel s (@ − 40 ◦ C, fan cooled),<br />

NIR AR-coated vacuum window, no window on CCD;<br />

controller ST-138 (PCI-bus), A/D converter 12 bit (@ 1 MHz);<br />

WinView 32 imaging software, LabVIEW driver library<br />

Conrad Electronic, b&w miniature camera module,<br />

no. 19-27-75, tele-lens no. 11-65-32;<br />

EHDPhysikalische Technik, KAM 08, b&w 1/3” CCD sensor<br />

Integral Technologies, FlashBus BV Lite (PCI-bus)<br />

Laser diagnostics:<br />

Spectrometer Ocean Optics, PC 2000,<br />

miniature PC-card fibre optic grating spectrometer (ISA-bus),<br />

grating no. 6, range 650 − 850 nm, entrance slit 10 µm,<br />

resolution 0.4 nm, accuracy 0.1 nm<br />

Table 3.2: (Computer) hardware <strong>for</strong> experimental control and data acquisition.


58 Experimental setup<br />

3.7 The magneto-optical trap<br />

3.7.1 Trapping principle and molasses cooling<br />

Since the first demonstration [33], the magneto-optical trap and optical molasses<br />

cooling have been topic of numerous experimental and theoretical investigations.<br />

For detailed in<strong>for</strong>mation see, e.g., Refs. [11, 12, 70, 71]. Particular work on vapourcell<br />

configurations has been reported in Refs. [124–126, 132].<br />

The MOT is based on the spontaneous light <strong>for</strong>ce [2,152,153], i.e. the transfer of<br />

photon recoil momenta, k 0 , to atoms, where k 0 =2π/λ 0 is the vacuum wave vector<br />

of near-resonance laser light. The configuration of the light field is chosen such that<br />

this momentum transfer occurs with a preferential direction and in a succession of<br />

absorption and spontaneous emission cycles. A central trapping <strong>for</strong>ce is established<br />

in combination with a (velocity dependent) friction <strong>for</strong>ce that cools the atoms to<br />

a temperature close to the “Doppler limit”, T D = Γ/2k B . For rubidium, with a<br />

transition linewidth Γ/2π =6.1MHz, this limit is T D =146µK.<br />

Sub-Doppler cooling schemes, such as polarisation gradient cooling in optical<br />

molasses, provide temperatures close the “recoil limit”, T R =(k 0 ) 2 /Mk B . With<br />

the rubidium mass M =87amu,thisisT R = 361nK. The corresponding photon<br />

recoil velocity is v rec =5.88 mm/s. (Some additional, useful numbers <strong>for</strong> rubidium<br />

are listed in the Appendix A.1).<br />

Fig. 3.11(a) shows the principle of the MOT in a one-dimensional scheme. For<br />

convenience, a J g =0−→ J e = 1transition scheme is consi<strong>der</strong>ed. In a magnetic<br />

field of constant gradient, B(z) =bz, the excited state sublevels, m e = {0, ±1},<br />

are Zeeman-shifted in a position dependent way by ω Z (z, m e )=m e g e (µ B /) bz.For<br />

simplicity, we assume a Landé factorg e = 1. The counter-propagating laser beams<br />

are red detuned, δ = ω L − ω 0 < 0. The circular polarisations are assigned with<br />

respect to the z-axis, which is the atomic quantisation axis. The σ + -beam that<br />

travels to the right carries + angular momentum, the σ − -beam that travels to the<br />

left carries −. However, when defining the polarisation state with respect to the<br />

propagation direction, and following the notation of Ref. [7], both beams are “leftcircularly”<br />

polarised (L): An observer facing the source sees the electric field vector<br />

in a counter-clockwise rotation. The Zeeman shift brings the red detuned light in<br />

resonance with the appropriate transition to a sublevel m e , such that the atom is<br />

pushed towards z = 0. This establishes a central trapping <strong>for</strong>ce.<br />

The cooling effect in the MOT configuration is based on the Doppler shift of<br />

the atomic resonance, ω D (v) =−k 0 v. Due to the red laser detuning, moving atoms<br />

are shifted into resonance with the counter-propagating laser (ω D > 0). Hence, a<br />

moving atom preferentially absorbs decelerating photons. This mechanism is called<br />

“Doppler cooling”. Note that it is effective also in the presence of the magnetic<br />

field gradient, thus enabling a combined spontaneous trapping and cooling <strong>for</strong>ce F.<br />

This <strong>for</strong>ce can be written in terms of the scattering rate of an atom in the beams<br />

travelling to the right (F + ) and to the left (F − ), as shown in Fig. 3.11(a):<br />

F(z, v) = F + (z, v)+F − (z, v) , (3.1)


3.7 The magneto-optical trap 59<br />

(a)<br />

<br />

J =1<br />

m =+1<br />

e<br />

e<br />

(b)<br />

R<br />

R<br />

B3<br />

m =+1<br />

e<br />

<br />

L<br />

+ -<br />

B=<br />

b z<br />

z<br />

m =0<br />

m<br />

e=-1<br />

m =-1<br />

( L )<br />

( L)<br />

e<br />

e<br />

J =0<br />

g<br />

L<br />

-I<br />

R<br />

x<br />

70 o<br />

z<br />

y<br />

4cm<br />

L<br />

I<br />

R<br />

B2<br />

B1<br />

Figure 3.11: The magneto-optical trap. (a) Simplified 1D two-level scheme with a red<br />

detuned laser, δ = ω L − ω 0 < 0. Trapping: The sublevels are Zeeman-shifted in a<br />

magnetic field gradient. An atom at a position z, preferentially absorbs resonant light<br />

that pushes the atom towards z =0. Doppler cooling: An atom preferentially absorbs<br />

counter-propagating light that is Doppler-shifted into resonance. (b) 3D configuration of<br />

three counter-propagating beam pairs, B1, B2, and B3 in the vapour cell. The mutual<br />

angle of B2 and B3 in the yz-plane is 70 ◦ . The opposing currents, ±I, in the coils generate<br />

a quadrupole field gradient. The polarisation notation of right (R) andleft(L) circular<br />

light visualises the symmetries of the configuration.<br />

F ± (z, v) = ±k 0<br />

Γ<br />

2<br />

I L<br />

1<br />

I 0 4δ± 2 (z, v)<br />

+1+ I . (3.2)<br />

L<br />

Γ 2 I 0<br />

Here, I 0 is the saturation intensity <strong>for</strong> the atomic species, and δ ± (z, v) is the effective<br />

detuning including Zeeman and Doppler shift:<br />

δ ± (z, v) = δ ∓ µ B<br />

bz ± k 0v. (3.3)<br />

Usually, a detuning of |δ| ∼ Γ is applied and the magnetic field gradient is<br />

b ∼ 10 G/cm.<br />

In or<strong>der</strong> to achieve temperatures close to the recoil temperature, we apply PGC<br />

in “σ + σ − ” optical molasses. Comprehensive descriptions of the cooling process can<br />

be found in [11,12,70]. This technique is particularly useful here: when the magnetic<br />

field is switched off, the MOT laser configuration just results in the required “σ + σ − ”<br />

polarisation scheme. Only laser intensities and detunings have to be adapted whilst<br />

switching. Note that there is no trapping <strong>for</strong>ce in the PGC configuration.


60 Experimental setup<br />

3.7.2 Experimental configuration<br />

The 3-dimensional realisation of the MOT scheme is shown in Fig. 3.11(b). We<br />

use one horizontal (B1) and two diagonal (B2, B3) pairs of counter-propagating,<br />

circularly polarised, collimated laser beams. The horizontal x-direction is the axis<br />

of cylindrical symmetry of the magnetic field. The mutual angle of B2 and B3 in<br />

the yz-plane is 70 ◦ . This allows beam waists up to 5 mm (1/e 2 intensity radius)<br />

without significant clipping of the beams by the prism, given a MOT height larger<br />

than ∼ 5 mm. In the following, be<strong>for</strong>e turning to the experimental per<strong>for</strong>mance of<br />

the MOT, (i) the optical setup and, (ii) the magnetic field coils are described.<br />

(i) Optical setup.— The optical components <strong>for</strong> the MOT were arranged in a<br />

way to maintain access to the vapour cell <strong>for</strong> other optics. A schematic top view<br />

is shown in Fig. 3.12. The beam pairs B1, B2, and B3 are <strong>der</strong>ived from a single<br />

Gaussian beam from a single-mode optical fibre. The fibre output was collimated<br />

to a waist of 4 mm by two achromatic lens doublets, L1(f = 50 mm, dia. 30 mm)<br />

and L2 (f = 300 mm, dia. 40 mm). The lens diameters were chosen large enough<br />

to avoid diffraction fringes in the collimated beam. Two polarising cubes split off<br />

subsequently 1/3 and1/2 of the power, thus preparing three beams of equal power.<br />

The linear polarisation <strong>for</strong> the splitting is adjusted by polarisation-controlling fibre<br />

loops (PCL) and a half-wave plate (HW). Optionally, a polariser is used directly after<br />

the fibre to keep the splitting ratio constant, despite thermal polarisation drifts of<br />

the fibre output. The cubes steer the beams B2 and B3 upwards at an angle of<br />

35 ◦ with the optical table. The beams are then horizontally directed towards the<br />

upper mirrors (UM) of the trapping setup, shown in the photograph. These mirrors<br />

steer the beams downwards through the vapour cell un<strong>der</strong> the same angle, 35 ◦ .The<br />

purpose of this construction is to use the (dielectric) mirrors exclusively either with<br />

normal or 45 ◦ incidence. Care was also taken to have only light in purely linear TE<br />

or TM polarisation being reflected un<strong>der</strong> 45 ◦ incidence. This secures the polarisation<br />

of the reflected light to stay linear. Beam B1passes the cell horizontally. Be<strong>for</strong>e<br />

passing the cell, all beams become circularly polarised by quarter-wave plates (QW).<br />

In retro-reflection, the initial helicity in each beam is restored by a second QW plate.<br />

The iris diaphragm (ID) be<strong>for</strong>e the beam splitters facilitates initial spatial alignment<br />

of the MOT beams by means of observing the fluorescence of the narrowed beams<br />

in the rubidium vapour cell. A repumping laser (RL) is coupled in by the second<br />

cube and is superposed with the the trapping beams B2 and B3.<br />

The present retro-reflection concept suffers from an imbalance in the laser intensities,<br />

causing an imbalance in the light <strong>for</strong>ces, <strong>for</strong> two reasons. First, the cloud of<br />

cold atoms in the MOT is optically dense. Hence, each retro-reflected beam carries<br />

a shadow in the centre. Particularly in molasses cooling this may increase the final<br />

temperature. To avoid this effect, six independent beams of equal intensity could be<br />

used. However, this requires more optical components and twice the laser power. A<br />

simpler solution is to apply a slight directional misalignment of the retro-reflected<br />

beams, in or<strong>der</strong> to keep the cloud of atoms mostly out of the shadows.


3.7 The magneto-optical trap 61<br />

UM<br />

QW<br />

-I<br />

prism<br />

LM<br />

+I<br />

B2<br />

B1<br />

B3<br />

RL<br />

PB<br />

HW<br />

ID<br />

L2<br />

L1<br />

OF<br />

TL<br />

PCL<br />

Figure 3.12: Optical scheme MOT. Top view in the drawing (not to scale): retroreflected<br />

beam pairs (B1, B2, B3), trapping laser (TL), repumping laser (RL), optical<br />

fibre (OF), polarisation-controlling fibre loops (PCL), collimation lenses (L1, L2), iris<br />

diaphragm (ID), polarising beam splitting cubes (PB), half-wave plate (HW), upper and<br />

lower mirrors (UM, LM), quarter-wave plates (QW), opposing quadrupole field currents<br />

(±I). An “atom cloud“ (black dot) and a glass prism are also indicated in the drawing.<br />

The second imbalance stems from the uncoated UHV cuvette. While passing<br />

4 glass surfaces, the beams suffer significant reflection losses. For example, in TM<br />

(TE) polarisation, the retro-reflected diagonal beams have 18 % (25 %) less power,<br />

when meeting the cold atoms in the MOT again. Obviously, also a prepared circular<br />

polarisation of the beams will become elliptical to a certain degree due to the<br />

different losses in TM and TE polarisation. The imbalance in optical power may<br />

be overcome by making the back-travelling beams slightly convergent, in or<strong>der</strong> to<br />

increase the intensity at the place of cold atoms. Note, that with a cell made from<br />

fused silica (n =1.45) the reduction in optical power would be different, namely<br />

6.5 % (21%) <strong>for</strong> TM (TE) polarisation. For details on the consequences of the beam<br />

imbalance on the per<strong>for</strong>mance of the MOT, see e.g. Ref. [154].


62 Experimental setup<br />

(ii) Magnetic field coils.— The magnetic quadrupole field gradient is provided<br />

by a pair of coils, placed in-axis with the horizontal trapping beam, B3. The currents,<br />

±I, oppose each other. This results in a magnetic field that increases approximately<br />

linearly with the distance from the centre. Along the symmetry axis, r =(x, 0, 0),<br />

and close to the centre, the field is given by:<br />

B(r) = b x, (3.4)<br />

b = µ 0 NI<br />

3 R 2 d<br />

(R 2 + d 2 ) 5 2<br />

ˆx , (3.5)<br />

where N = 25 is the number of turns per coil, made from 0.8 mm dia. copper wire.<br />

The wire is stacked in a square grid pattern such that the (average) coil radius is<br />

R = 17 mm. The distance between the coils is 2d = 57 mm, being limited by the<br />

cuvette of 42 mm width. The resulting field gradient along in the x-direction is<br />

calculated to be |b(I)| = I × 1.9 G/cm A. The coils with an estimated induction<br />

of each ∼ 25 µH can be switched off by a power MOSFET (IRF 530) within 20 µs.<br />

After switching, the induction current is dissipated by a 4.7 Ω bypass resistor. The<br />

MOSFET switch was chosen with a breakdown voltage of 100 V to manage the<br />

induction voltage peak ∼ 50 V without damage. Also visible in the photograph of<br />

Fig. 3.12 are connections <strong>for</strong> coolant flow through the coil mounts. However, since<br />

our MOT was not operated continuously over a longer period and the current usually<br />

was I 10 A, no cooling has been required so far.<br />

The source-free character of the magnetic field, ∇·B = 0, results in a field<br />

gradient twice as large along the symmetry x-axis as compared to the orthogonal<br />

yz-plane. Also the non-orthogonal crossing angle of the trapping beams in this plane<br />

together with intensity imbalances results in a reduced vertical trapping <strong>for</strong>ce. This<br />

may have caused the observed vertically elongated MOT shape of approximately 1:2<br />

aspect ratio.<br />

In or<strong>der</strong> to compensate the earth magnetic field, usually ∼ 0.5 G, and other<br />

stray fields at the location of the MOT, we mounted a cage-like frame of three coil<br />

pairs around the setup: 2 rectangular pairs (60 cm wide, 80 cm high) and 1circular<br />

pair <strong>for</strong> the vertical axis (dia. 85 cm). Each pair consists of 80 turns of 0.8 mm dia.<br />

copper wires. The applied current ranged between ±400 mA, thus compensating<br />

indeed fields ∼ 0.5 G.<br />

3.7.3 Loading the MOT<br />

The trapping light of the MOT is tuned 1.5 Γ to the red of the F g =2−→ F e =3<br />

cycling transition of the 87 Rb D2-line (see Fig. 3.7). The atoms are trapped in the<br />

F g = 2 ground state. <strong>Atoms</strong> that are off-resonantly excited to F e = {1, 2} and<br />

thus optically pumped into the F g = 1ground state, are transferred back by the<br />

repumping laser, which is in resonance with the D1-line (F g = 1 −→ F e = 2).<br />

Typically, a power of 15 mW is distributed among the three trapping beam pairs.<br />

The intensity in the trap centre is ∼ 30 I 0 .


3.7 The magneto-optical trap 63<br />

molasses<br />

MOT<br />

molasses<br />

fluorescence signal<br />

(arb. units)<br />

MOT<br />

molasses<br />

vapour<br />

0 1 2 3<br />

time (s)<br />

Figure 3.13: Fluorescence of 87 Rb in MOT, molasses and background vapour.<br />

The fluorescence of trapped atoms during MOT loading and molasses cooling<br />

was observed by a photodiode facing down from above the UHV cuvette. In or<strong>der</strong><br />

to capture light from a large solid angle, we used a lens (f = 50 mm, dia.= 50 mm)<br />

directly above the cuvette, and imaged the cloud on the photodiode. Fig. 3.13 shows<br />

a fluorescence signal, recorded when continuously cycling between MOT loading and<br />

molasses cooling. Note that the signal started here with a molasses period. The<br />

arrow at 0.8 s indicates the start of a MOT loading period, when the current in the<br />

quadrupole coils was switched on together with the trapping laser. The fluorescence<br />

signal at that moment stems from fluorescence of rubidium background vapour in<br />

both trapping and repumping light. The latter was applied permanently. After 2 s<br />

the number of trapped atoms saturates. By using slightly higher rubidium vapour<br />

pressure, also loading times ∼ 0.5 s were achieved, which allows to increase the<br />

repetition rate of experiments. After loading the MOT, at 2.8 s, the magnetic field<br />

was switched off and the trapping light was changed into the molasses configuration<br />

by increasing the red detuning to δ = −10 Γ and by reducing the intensity to half<br />

the MOT intensity. Hence, the fluorescence signal abruptly gets weaker. The cloud<br />

diffusively expanded out of the detector’s field of view. After 1s, the signal settled<br />

at the fluorescence from the background vapour (in the molasses light), and the<br />

cycle was repeated. Taking into account the solid angle covered by the detection<br />

scheme, we evaluated an atom number of 5 × 10 7 atoms in the MOT. The horizontal<br />

rms diameter of the MOT was approximately 0.5 mm.


64Experimental setup<br />

3.7.4Time-of-flight temperature measurement<br />

In early experiments on laser cooling, the temperature of cold atom clouds was<br />

investigated by a “release & recapture” technique [155], with which the cooling<br />

light is switched off <strong>for</strong> a short ballistic expansion period of the released cloud after<br />

that only atoms in reach of the cooling light are recaptured and detected. The<br />

temperature is <strong>der</strong>ived from the recaptured atom fraction. This method has been<br />

succeeded by the more accurate time-of-flight technique (TOF) [156], which we<br />

used also in our experiments. Various other techniques have been reported so far,<br />

e.g. using recoil-induced resonances [157] or imaging of ballistically expanding atom<br />

clouds [39].<br />

The TOF technique makes also use of the ballistical expansion. During molasses<br />

cooling, the equilibrium temperature is established after a few ms. In or<strong>der</strong> to<br />

release an atom cloud, we shutter the cooling laser mechanically after typically 4 ms<br />

of cooling. A significantly longer cooling time might lead to un<strong>for</strong>tunate diffusive<br />

atom loss. The falling atoms pass a thin sheet of resonant probe light below, and<br />

well separated from the cooling region. Either the fluorescence or the absorption is<br />

recorded as a function of time, see e.g. Fig. 5.2(a). The temperature is obtained by<br />

fitting a thermal Maxwell-Boltzmann velocity distribution to the recorded signal.<br />

If the initial cloud size is known only approximately, the TOF method requires<br />

a fall height sufficiently large to make the cloud size negligible with respect to the<br />

thermal expansion during the fall time. In our setup, we can drop the atoms over<br />

a distance of 5 mm, which is sufficient to determine the temperature within 15 %<br />

accuracy. The TOF method is also a simple way to investigate atoms bouncing on<br />

an atom mirror, see Chap. 5.<br />

Fig. 3.14(a) shows TOF signals <strong>for</strong> two distinct height settings of the flat probe<br />

beam, 1.2 mm and 4.3 mm below the MOT. The origin of the time axis is the time<br />

when the mechanical shutter of the cooling light was closed. The signals were again<br />

recorded in fluorescence by means of a photodiode from above the vacuum cuvette,<br />

similar to the signal shown in Fig. 3.13. The probe had a waist of 0.4 mm vertical<br />

and 1.4 mm horizontal (1/e 2 radius). It was tuned close to resonance with the<br />

F g =2−→ F e = 3 transition of the 87 Rb D2-line.<br />

For the fluorescence technique to be efficient also with small quantities of atoms,<br />

the probe must saturate the optical transition. In a travelling-wave probe beam, the<br />

atoms would be quickly accelerated by the recoils from absorbed photons. The atoms<br />

are there<strong>for</strong>e Doppler-shifted out of resonance and lost <strong>for</strong> longer recording of the<br />

TOF signal. In or<strong>der</strong> to achieve a longer interaction time per atom, the probe was<br />

used in retro-reflection as a standing wave. The flat, horizontal probing section was<br />

<strong>for</strong>med by a cylindrical telescope with the focus of two cylindrical lenses (f = 75 mm)<br />

located below the MOT. An additional measure to enhance the interaction time was<br />

to choose a small red probe detuning ∼ Γ, which converts the probe beam into a<br />

1D molasses cooling configuration.


3.7 The magneto-optical trap 65<br />

probe fluorescence (arb. u.)<br />

(a)<br />

0 10 20 30 40 50<br />

time (ms)<br />

( ) fitted fall height (mm)<br />

(b)<br />

10<br />

6<br />

9<br />

4<br />

8<br />

2<br />

7<br />

0<br />

6<br />

0 1 2 3 4 5<br />

relative probe height (mm)<br />

( ) fitted temperature (K)<br />

Figure 3.14: Time-of-flight temperature measurement. (a) Fluorescence signals with<br />

probe beam 1.2 mm (◦) and4.3 mm (•) below the MOT. Solid lines are fits to a Maxwell-<br />

Boltzmann distribution. (b) Systematics of the temperature fit <strong>for</strong> various probe settings:<br />

fitted fall height vs. (relative) probe setting (◦, with linear fit); fitted temperatures (•).<br />

The fit of the distribution of an expanding and gravitationally accelerating cloud<br />

to the TOF signals assumes an Gaussian initial phase-space distribution in the<br />

vertical direction, Φ 0 (z 0 ,v 0 ,t 0 ), that corresponds to the optical molasses temperature<br />

[cf. Fig. 2.4(a)]:<br />

[<br />

1<br />

Φ 0 (z 0 ,v 0 ,t 0 ) = exp − 1 ( ) 2 ( ) 2<br />

2πσ z σ v 2 ( z0 v0<br />

+ )]<br />

, (3.6)<br />

σ z σ v (T )<br />

√<br />

kB T<br />

σ v (T ) =<br />

M . (3.7)<br />

The temperature is represented by the rms velocity spread σ v . The initial rms<br />

radius of the cloud is σ z . As atoms fall, the time-evolution of the distribution can<br />

be written by using trans<strong>for</strong>med coordinates:<br />

z = z 0 + v 0 t + 1 2 gt2 , v = v 0 + gt , (3.8)<br />

∫∫<br />

∫∫<br />

dzdv Φ(z, v, t) = dz 0 dv 0 Φ 0 (z 0 ,v 0 ,t 0 ) ≡ 1 . (3.9)<br />

When the probe beam is approximated by a square intensity profile of thickness d<br />

and centred at the height z p , the signal recorded from atoms that pass the probe<br />

section is<br />

s(v, t) =<br />

∫ zp+d/2<br />

z p−d/2<br />

Φ(z, v, t) dz ≃ Φ(z p ,v,t) d. (3.10)


66 Experimental setup<br />

Integration over the velocity distribution leads to the TOF signal:<br />

S(t) =<br />

≃<br />

∫ +∞<br />

−∞<br />

s(v, t) dv (3.11)<br />

1<br />

√ d exp (− 1<br />

2πσ(t) 2σ 2 (t)<br />

( 1<br />

2 gt2 − z p<br />

) 2<br />

) , (3.12)<br />

σ(t) = √ σ 2 z + σ2 v (T )t2 . (3.13)<br />

The TOF signal is thus described by a Gaussian distribution, the rms width of<br />

which is growing in time. If z p , σ z ,andd are known, the temperature T in σ v (T )is<br />

the only parameter to fit the recorded signals to. Alternatively, z p and σ z can also<br />

be treated as fit parameters. Note that d appears as an overall amplitude scaling<br />

factor.<br />

Fig. 3.14(b) shows temperature and fall height as obtained when fitting T , z p ,<br />

σ z and the signal amplitude <strong>for</strong> various relative experimental probe height settings.<br />

The statistical errors in the temperatures are also shown. The statistical errors in<br />

the heights are small and not shown. Accurate knowledge of the fall height is not<br />

required to obtain a reliable temperature. This can be tested by fixing the height<br />

with a slightly different value and fitting again with the temperature as the only fit<br />

parameter. We find that the shift in the fitted temperature remains within the error<br />

margins.<br />

The temperature fits suggest a small statistical error, T =8.5(1) µK. However,<br />

the uncertainty in the initial size of the molasses, σ z , causes a systematic error. Using<br />

σ z also as a fit parameter resulted in σ z ≃ 0.55 mm. However, images recorded with<br />

a CCD camera suggested a value of 0.25 mm. When the σ z =0.25 mm was used<br />

as a fixed parameter, the temperature fitted to ≃ 12 µK, since the contribution<br />

of thermal expansion in the expression S(t) was increased. It is obvious that a<br />

systematic error in σ z is more severe <strong>for</strong> small fall heights, <strong>for</strong> which a falling cloud<br />

haslittletimetoexpandbe<strong>for</strong>ebeingprobed.<br />

3.7.5 Molasses cooling and magnetic field compensation<br />

The equilibrium temperature that is achieved in polarisation gradient cooling is expected<br />

to scale ∝ I L /δ with the intensity and detuning of the cooling light [70].<br />

In or<strong>der</strong> to optimize the cooling process experimentally, we per<strong>for</strong>med TOF temperature<br />

measurements <strong>for</strong> various settings of the cooling laser. Fig. 3.15(a) shows<br />

TOF signals that were recorded <strong>for</strong> various red detunings ranging from 1.2 − 8.3Γ<br />

(see also Ref. [158]). In Fig. 3.15(b), the fitted temperatures are plotted vs. the inverse<br />

detuning. The inverse dependence on the detuning seems to be approximately<br />

fulfilled with our cooling setup. The linear dependence on the intensity, however,<br />

indicates an offset, see Fig. 3.15(c).


3.7 The magneto-optical trap 67<br />

(a)<br />

(b)<br />

probe fluorescence (arb.u.)<br />

119 K<br />

103 K<br />

40 K<br />

81 K<br />

66 K<br />

57 K<br />

47 K<br />

40 K<br />

temperature (K)<br />

temperature (K)<br />

100<br />

50<br />

0<br />

-1.0 -0.8 -0.6 -0.4 -0.2 0<br />

1/detuning /<br />

50<br />

25<br />

(c)<br />

0 20 40 60 80<br />

time (ms)<br />

0<br />

0 1 2 3 4 5<br />

molasses intensity I/I 0<br />

Figure 3.15: (a) TOF signals after 4 ms of molasses cooling, laser detuning ranging from<br />

1.2 − 8.3Γ (top-down), Maxwell-Boltzmann distribution fitted to the 40 µK signal (solid<br />

curve). (b) Fitted temperature vs. inverse detuning. The temperatures were relatively<br />

large, because stray magnetic fields were not compensated. (c) Temperature vs. laser<br />

intensity per cooling beam, in units of the saturation intensity, I 0 =1.67 mW/cm 2 .<br />

The lowest temperatures are achieved in molasses cooling when earth and other<br />

stray magnetic fields are compensated on the mG level. Note that the signals shown<br />

in Fig. 3.15 were recorded be<strong>for</strong>e any field compensation measure. Hence, the final<br />

temperatures were relatively high. The experimental region inside the UHV cell is<br />

not accessible <strong>for</strong> external field probes. In situ, one may investigate field dependent<br />

spectral properties of the atomic species using, e.g., electro-magnetically induced<br />

transparency (EIT) [159], i.e. the Hanle level-crossing effect [160]. Although these<br />

techniques are sensitive on the µG level, they require additional laser sources. As<br />

a simpler probe, we use the measured molasses temperature to optimize the field<br />

compensation [71], thus achieving temperatures as low as shown in Fig. 3.14. Experimentally,<br />

it proved to be also sufficient to observe the diffusion of atoms during<br />

molasses cooling and to maximize the diffusion time constant by means of the field<br />

compensation coils, see Fig. 3.13.


68 Experimental setup


4<br />

A high-power tapered<br />

semiconductor amplifier system<br />

A laser amplifier system has been characterised which provides up<br />

to 200 mW output at 780 nm wavelength after a single-mode optical<br />

fibre. The system is based on a tapered semiconductor gain element<br />

that amplifies the output of a narrow-linewidth diode laser. Gain<br />

and saturation are discussed as a function of operating temperature<br />

and injection current. The spectral properties of the amplifier<br />

were investigated with a grating spectrometer. Amplified spontaneous<br />

emission (ASE) was observed as a spectral background with<br />

a full width half maximum of 4 nm. The ASE background was suppressed<br />

to below the detection limit of the spectrometer by a proper<br />

choice of operating current and temperature, and by sending the<br />

light through a single-mode optical fibre. The final ASE spectral<br />

density was less than 0.1 nW/MHz, i.e. less than 0.2% of the optical<br />

power. Related to a rubidium optical transition linewidth of<br />

Γ/2π =6MHz, this gives a background suppression of better than<br />

−82 dB. An indication of the beam quality is provided by the fibre<br />

coupling efficiency up to 59 %. The application of the amplifier<br />

system as a laser source <strong>for</strong> atom optical experiments is discussed.<br />

This chapter is based on the preprint<br />

D. Voigt, E.C. Schil<strong>der</strong>, R.J.C. Spreeuw, and H.B. van Linden van den Heuvell,<br />

arXiv:physics/0004043, accepted <strong>for</strong> publication in Appl. Phys. B.<br />

69


70 A high-power tapered semiconductor amplifier system<br />

4.1 Introduction<br />

The techniques of laser cooling and trapping of neutral atoms require stable, narrowlinewidth<br />

and frequency-tunable laser sources [11, 12]. Commonly used systems <strong>for</strong><br />

the near-infrared wavelengths are based on external grating diode lasers (EGDL)<br />

[117]. Optical feedback from a grating narrows the linewidth to less than 1 MHz and<br />

provides tunability. High-power single-transverse-mode diode lasers can provide up<br />

to 80 mW optical output at wavelengths below 800 nm. In this power range, diode<br />

lasers thus provide a less costly alternative to Ti:Sapphire lasers. If more power<br />

is required, the output of an EGDL can be amplified. Presently, there are three<br />

common techniques based on semiconductor gain elements: (i) Injection-locking of<br />

a single-mode laser diode [138,139] by seeding light from an EGDL results typically<br />

in 60 − 80 mW optical power at 780 nm wavelength. (ii) Amplification in a doublepass<br />

through a broad-area emitting diode laser (BAL) [161–167]. This yields an<br />

optical output of typically 150 mW after spatial filtering. A disadvantage is the<br />

relatively low gain of 10 − 15, requiring high seed input power. The BAL gain can<br />

be improved using phase conjugating mirrors in the seed incoupling setup [168].<br />

(iii) Travelling-wave amplification in a semiconductor gain element with a tapered<br />

waveguide, a “tapered amplifier” (TA) [163, 169–171]. Compared to a BAL this<br />

yields higher gain and higher power after spatial filtering. This approach requires<br />

much lower input and a less complex optical setup than a BAL. However, a TA gain<br />

element is consi<strong>der</strong>ably more expensive.<br />

We have investigated a TA system that amplifies the narrow-linewidth seed beam<br />

of an EGDL and provides up to 200 mW optical output from a single-mode optical<br />

fibre. With the tapered gain element, characterised in this chapter, the system<br />

operates on the D2 (5 S 1/2 −→ 5 P 3/2 ) line of rubidium at a wavelength of 780 nm,<br />

see Fig. 3.7. Another gain element of the same type but with a different centre<br />

wavelength is used on the D1(5 S 1/2 −→ 5 P 1/2 ) line at 795 nm. The input facet<br />

of the tapered gain element has the typical width (≈ 5 µm) of a low power singletransverse-mode<br />

diode laser. A seeding beam is amplified in a single pass and<br />

expanded laterally by the taper to a width of typically 100 − 200 µm such that the<br />

light intensity at the output facet is kept below the damage threshold and the beam<br />

remains diffraction limited (see e.g. Ref. [169]). The output power can thus be much<br />

larger than from a single-mode waveguide.<br />

In previous work, TAs have been used as sources <strong>for</strong> frequency-doubling and<br />

pumping solid state lasers [172]. Apart from the achievable output power, frequency<br />

tunability of the narrow-linewidth output [173], simultaneous multifrequency generation<br />

[174], and spatial mode properties, including coupling to optical fibres [175–177]<br />

have been addressed.<br />

In this chapter the broadband spectral properties of the TA are discussed. We<br />

have minimized the background due to ASE in the gain element by adjusting the<br />

operating conditions of the amplifier, i.e. temperature, injection current and seed<br />

input power. We have also investigated the coupling efficiency of the TA output<br />

to a single-mode optical fibre, and have found that the latter acts both as a spatial<br />

and as a spectral filter. The properties of three gain elements of the same type are


4.2 Amplifier setup 71<br />

TA<br />

OF<br />

EGDL<br />

OI<br />

CL<br />

OI<br />

SA<br />

GS<br />

top view<br />

IC<br />

TA<br />

OC<br />

CL<br />

side view<br />

Figure 4.1: Setup of the tapered amplifier system. Seed laser (EGDL), tapered gain element<br />

(TA), 60 dB optical isolators (OI), single-mode optical fibre (OF), optical spectrum<br />

analyser (SA) and grating spectrometer (GS). A top and side view of the gain element is<br />

shown with input and output collimators (IC,OC). A cylindrical lens (CL) compensates<br />

astigmatism (not to scale).<br />

compared. Atom optical applications usually require good suppression of spectral<br />

background. For example, in far off-resonance optical dipole traps [10], scattering<br />

of resonant light from the background causes heating and atom loss. Consequences<br />

of ASE background in such schemes are also discussed in this chapter.<br />

4.2 Amplifier setup<br />

The amplifier system consists of a seed laser, the output of which is amplified in<br />

a single pass by the tapered gain element, as shown in Fig. 4.1. The TA output is<br />

coupled to a single-mode optical fibre (OFR, type PAF-X-5-780 fibre port, input<br />

beam dia. 0.9−1.8 mm). The seed laser is an EGDL with a linewidth of less than<br />

1MHz. It operates by a 60 mW single-mode laser diode (Hitachi, HL 7851G98)<br />

and provides 28 mW to seed the amplifier at 780 nm wavelength. Coupling of the<br />

seeding beam to the amplifier was realised by mode-matching the seed laser with<br />

the backward travelling beam emitted by the TA. The divergence angles from the<br />

seed laser emission and the backward directed TA emission are similar. Hence, sufficient<br />

mode-matching was obtained using identical collimation lenses <strong>for</strong> both (Thor-<br />

Labs, C 230 TM-B, f =4.5 mm, N.A.=0.55). Additional mode shaping, e.g. with


72 A high-power tapered semiconductor amplifier system<br />

anamorphic prism pairs, was not necessary. An optical isolator with 60 dB isolation<br />

protects the stabilised seed laser from feedback by the mode-matched beam of the<br />

amplifier (Gsänger Optoelektronik, type DLI 1). The 5 mm aperture of the isolator<br />

is sufficiently large not to clip the elliptical seed beam.<br />

The TA was a SDL 8630 E (Spectra Diode Laboratories, ser.no. TD 310). According<br />

to the manufacturer’s data sheet, the output power ranged between 0.5−0.55 W<br />

within a wavelength tuning range of 787 − 797 nm, at an operating temperature of<br />

21 ◦ C. The beam quality parameter is typically specified as M 2 < 1.4 [178]. The<br />

TA should be protected from any reflected light, because it will be amplified in the<br />

backward direction and may destroy the amplifier’s entrance facet. Hence, the output<br />

collimator has a large numerical aperture (ThorLabs, C 330 TM-B, f =3.1mm,<br />

N.A.=0.68) and the beam is sent through a second 60 dB optical isolator (Gsänger,<br />

FR 788 TS). The plane of the tapered gain element is vertically oriented, so that<br />

diffraction yields a large horizontal divergence. The beam is then collimated similarly<br />

to the seed input, but yields a focus in the vertical plane. A cylindrical lens<br />

(Melles Griot, no. 01LQC 006/076, f = 100 mm) compensates the astigmatism of<br />

the beam, so that the beam can couple to a single-mode optical fibre. The astigmatism<br />

correction is shown in Fig. 4.1(see also Ref [158]).<br />

There is a consi<strong>der</strong>able loss in optical power due to the isolator transmission.<br />

Taking also into account small reflection losses on the lens surfaces, we estimated<br />

the useful output power to be 78 % of the power emitted by the TA facet. In the<br />

remain<strong>der</strong> of this chapter, all quoted powers are as measured with a calibrated power<br />

meter behind the optical isolator (Newport, meter 840-C, detector 818-ST, calibration<br />

module 818-CM). The narrow spectral line of the seed laser and amplifier output<br />

was monitored by an optical spectrum analyser with 1GHz free spectral range and<br />

with 50 MHz resolution. The amplifier’s broad spectral background was analysed<br />

using a grating spectrometer with a resolution of 0.27 nm. Also the output of the<br />

single-mode fibre was recorded with the spectrometer.<br />

The amplifier was provided as an open heat sink device, see Fig. 4.2. We mounted<br />

it on a water cooled base and stabilised it to the desired operating temperature<br />

within a few mK by a 40 W thermo-electric cooler. Thermal isolation from the<br />

ambient air and electromagnetic shielding were provided by a metal housing. When<br />

operating the amplifier at temperatures below the dew point, we flushed the containment<br />

with dry nitrogen. It is necessary to have a compact, stable mounting<br />

of the gain element and collimators. We mounted the collimators in a commercial<br />

xy-flexure mount to allow <strong>for</strong> lateral lens adjustment (New Focus, 9051M fibre<br />

launcher). The axial z-adjustment was done by two translation stages (Newport,<br />

type UMR 3.5, travel 5 mm). All adjustments, except that of the z-direction of the<br />

output collimator, are accessible from outside. This proved to be very convenient<br />

<strong>for</strong> mode-matching the seed beam and also <strong>for</strong> compensating beam displacement of<br />

the TA output when changing temperature or current.


4.2 Amplifier setup 73<br />

metal cap<br />

optical<br />

isolator<br />

cylindrical<br />

lens<br />

dessicant<br />

containment<br />

gain element,<br />

collimators<br />

collimation<br />

actuator access<br />

amplified beam<br />

seed<br />

40 cm<br />

base block,<br />

heatsink (brass)<br />

optical breadboard<br />

(removable)<br />

coolflow<br />

channels<br />

Back view<br />

Top view<br />

Mount of<br />

gain element<br />

transverse<br />

translation<br />

output<br />

collimator<br />

input<br />

collimator<br />

axial<br />

translation<br />

7cm<br />

9.5 cm<br />

optical axis,<br />

12 cm above table<br />

Peltier cooler<br />

gain<br />

element<br />

heatsink<br />

Figure 4.2: Construction of the tapered amplifier setup.


74A high-power tapered semiconductor amplifier system<br />

(a)<br />

(b)<br />

spectral density (mW/nm)<br />

30<br />

20<br />

10<br />

0<br />

5 0 C<br />

10 0 C<br />

13 0 C<br />

16 0 C<br />

780 790 800<br />

wavelength (nm)<br />

wavelength (nm)<br />

power (mW)<br />

790<br />

788<br />

786<br />

150<br />

100<br />

50<br />

(c)<br />

5 10 15<br />

temperature ( 0 C)<br />

5 10 15<br />

temperature ( 0 C)<br />

Figure 4.3: Temperature dependence of the unseeded amplifier at 1.2 A injection current.<br />

(a) spectra, (b) centre wavelength, (c) output power after optical isolator. Solid lines<br />

indicate linear fits.<br />

4.3 Unseeded operation of the amplifier<br />

When the TA receives no seed input, it operates as a laser diode. Thus, when the<br />

injection current, I TA , is increased from zero, the optical output power indicates the<br />

lasing threshold [see Fig. 4.4(a,b)]. Generally, both the operating wavelength and<br />

the optical power of a laser diode depend on the temperature. These properties<br />

are shown in Fig. 4.3. The emission spectrum of the lasing gain element is almost<br />

Gaussian shaped, with a width of 4 nm (1/e 2 intensity). It appears as a background<br />

of ASE also in the spectra when operating the gain element as an amplifier (see<br />

below). The oscillatory structures on the spectra are artifacts of the spectrometer.<br />

In the fitted Gaussian spectra, we evaluated the centre wavelength at each temperature<br />

setting. It increases with temperature with a slope of 0.28 nm/K, typical <strong>for</strong><br />

semiconductor lasers.<br />

The temperature dependence of the output power is shown in Fig. 4.3(c). We<br />

operated the TA within the specifications of the manufacturer’s data sheet that<br />

recommends to keep the optical power at the output facet below 550 mW. As the<br />

temperature increases, the conversion efficiency (mW/A) decreases and the threshold<br />

current increases. This can be seen in Fig. 4.4(a,b) (open symbols) where the<br />

optical output power P is plotted vs. the current I TA <strong>for</strong> two temperature settings.<br />

The threshold current of the unseeded TA increases from 0.78 A (5 ◦ C) to 0.86 A<br />

(14 ◦ C). From the slopes above threshold, we find that the conversion efficiency<br />

decreases from 0.7 W/A (5 ◦ C) to 0.5 W/A (14 ◦ C). In or<strong>der</strong> to measure the unperturbed<br />

output of the unseeded TA, one has to prevent light emitted from the


4.4 Amplification of a seed beam 75<br />

(a) T =14 0 C<br />

(c) T =14 0 C<br />

P (mW)<br />

P (mW)<br />

300<br />

150<br />

300<br />

150<br />

5.3 mW seed<br />

1.5 mW<br />

no seed<br />

0<br />

0.4 0.6 0.8 1.0 1.2<br />

I TA<br />

(A)<br />

(b) T =5 0 C<br />

8.6 mW<br />

1.7 mW<br />

no seed<br />

P (mW)<br />

300<br />

200<br />

100<br />

1.3 A<br />

I TA<br />

=1.2A<br />

1.0 A<br />

0.9 A<br />

0.8 A<br />

0<br />

0.4 0.6 0.8 1.0 1.2<br />

I TA<br />

(A)<br />

0<br />

0 5 10<br />

P seed<br />

(mW)<br />

Figure 4.4: Amplifier output vs. injection current and seed power. Lasing thresholds <strong>for</strong><br />

the unseeded amplifier are 0.86 A (14 ◦ C) and 0.78A (5 ◦ C), indicated by dashed lines.<br />

entrance facet from being reflected. Even a very weak reflection, e.g. from a power<br />

meter, would be amplified in the <strong>for</strong>ward direction. For the unseeded TA, we also<br />

measured the light propagating backward from the amplifier’s entrance facet. It<br />

reaches typically a power of 10 − 25 mW <strong>for</strong> injection currents from 1 − 1.4 A.<br />

Hence, a good isolation of the seeding laser is necessary.<br />

4.4 Amplification of a seed beam<br />

Amplification of a seed beam is evident in the output power of the TA. In Fig.<br />

4.4(a,b), the output power <strong>for</strong> different values of the seed power, P seed , is plotted <strong>for</strong><br />

two temperature settings. For the larger seed inputs of 8.6 mW and 5.3 mW, respectively,<br />

the amplifier was well saturated. The saturation is evident from Fig. 4.4(c)<br />

where P seed was varied <strong>for</strong> injection currents from 0.8−1.3 A.WithP seed ≈ 4mWthe<br />

device appeared to be saturated <strong>for</strong> all current settings. For P seed between 2−4 mW,<br />

the amplification ranged from 70 − 140, e.g. 320 mW output with 4 mW seed.<br />

The spectral properties of the TA and in particular the suppression of ASE background<br />

are discussed in the following. Fig. 4.5 shows the power spectral density of<br />

the TA output be<strong>for</strong>e an optical fibre <strong>for</strong> 16 ◦ Cand5 ◦ C operating temperature. In<br />

both cases the amplifier was saturated with 28 mW seed input. For comparison also<br />

the corresponding spectra of the unseeded amplifier are shown. In saturation, the<br />

broad ASE background is distinguished from a narrow peak of the amplified seed signal.<br />

The width of the peak is given by the bandwidth of the spectrometer, 0.27 nm


76 A high-power tapered semiconductor amplifier system<br />

(a) T =16 0 C<br />

(b) T =5 0 C<br />

30<br />

20<br />

P =323mW<br />

= 72dB<br />

ASE = 5.6 %<br />

30<br />

20<br />

P = 410 mW<br />

= 77dB<br />

ASE = 1.4 %<br />

dBm / nm<br />

10<br />

0<br />

dBm / nm<br />

10<br />

0<br />

-10<br />

-10<br />

775 780 785 790 795<br />

wavelength (nm)<br />

775 780 785 790 795<br />

wavelength (nm)<br />

Figure 4.5: Spectrum of the amplifier output (be<strong>for</strong>e the fibre). The seed power was<br />

28mW, the injection current 1.2 A. Dashed curves are <strong>for</strong> unseeded operation. ASE is<br />

the background fraction of the total optical power P and ε is the ASE suppression <strong>for</strong> the<br />

power spectral density in units of mW/Γ (see text).<br />

FWHM. Note that <strong>for</strong> the characterisation of the TA system, a different spectrometer<br />

was used than the PC-card spectrometer mentioned in Chap. 3. The linear dynamic<br />

range of the photomultiplier tube (PMT) which was used as a detector with<br />

the grating spectrometer, did not cover the entire dynamic range of 40 dB. There<strong>for</strong>e,<br />

we used a calibrated neutral density filter (CASIX, type NDG 0100) when recording<br />

the large signal of the locked laser line. A filter transmission of 6.0 % (780 nm) and<br />

5.1% (795 nm) was measured and linearly interpolated between these wavelengths.<br />

The small gap in the right slope of the peak in Fig. 4.5(b) (at 10 dBm/nm) indicates<br />

where the signal recorded with the filter was joined to the ASE spectrum recorded<br />

without the filter. Within the dynamic range where it was used, we verified that<br />

the response of the PMT was linear to within 1%. By means of an optical spectrum<br />

analyser and Doppler-free spectroscopy on rubidium, we could also verify that the<br />

amplified beam was spectrally narrow, comparable to that of the EGDL.<br />

The influence of the operating temperature is obvious first from the increased<br />

output power at lower temperature: 323 mW (16 ◦ C) and 410 mW (5 ◦ C), respectively.<br />

Second, both the peak level and total amount of ASE background are better<br />

suppressed at lower temperature. We attribute this to the shift of the gain profile<br />

of the TA toward the seed wavelength of 780 nm at a lower temperature [179]. The<br />

fraction of ASE background in the TA output is obtained by integrating the power<br />

spectral densities in Fig. 4.5, yielding 5.6 % (16 ◦ C) and 1.4 % (5 ◦ C), respectively.<br />

More than the total ASE fraction, the important figure <strong>for</strong> atom-optical applications<br />

is the fraction of ASE within the natural linewidth of the atomic transition<br />

used. We define this ratio ε by comparing the power in the peak with the ASE power<br />

in a bandwidth given by a typical atomic natural linewidth, e.g. Γ/2π =6MHz


4.5 Spatial and spectral filtering using an optical fibre 77<br />

<strong>for</strong> rubidium. At 16 ◦ C,theASEpeakvalueof+2.5dBm/nm is then reexpressed<br />

as 22 nW/Γ, or 7.9 nW/Γ at5 ◦ C, respectively. With 323 mW in the narrow line,<br />

this leads to a suppression ratio ε = −72 dB, or −77 dB with 410 mW, respectively.<br />

By an appropriate choice of the operating temperature one can thus optimize the<br />

spectral properties of the TA output. Even better suppression is achieved using an<br />

optical fibre as a spectral filter.<br />

4.5 Spatial and spectral filtering by an optical fibre<br />

For many applications, laser beam quality is an important property. A convenient<br />

method to obtain spatial filtering is to send the light through a single-mode optical<br />

fibre. An additional advantage of the fibre is a decoupling of the optical alignment<br />

between different parts of the experimental setup. Here, the coupling efficiency is<br />

discussed and the spectrum of the transmitted light is compared with the spectrum<br />

be<strong>for</strong>e the fibre. We observe that spatial filtering by the fibre is accompanied by<br />

spectral filtering. Evidently, the contribution of ASE in the TA beam is spatially<br />

distinguishable from the amplified seed signal.<br />

The spatial mode properties of the saturated TA output were slightly different<br />

<strong>for</strong> different injection currents. Fig. 4.6(a,b) represents the fibre transmission vs. the<br />

current. The fibre coupling had been optimized <strong>for</strong> a current of 1A and the TA was<br />

saturated. A maximum transmission of 46 % was achieved. For comparison, with an<br />

unamplified EGDL, after circularising the beam using an anamorphic prism pair, a<br />

typical fibre transmission of 75 % was obtained. The slope in the transmission curve<br />

is probably due to a beam displacement caused by the current-dependent thermal<br />

load of the gain element. Such a displacement was also observed when the operating<br />

temperature was changed. With the fibre coupling thus optimized, light from the<br />

unseeded TA had less transmission than the amplified seed signal. Fig. 4.6(c,d)<br />

shows <strong>for</strong> a fixed current of 1A that the fibre transmission was almost independent<br />

of the seed input power, i.e. the beam shape did not change.<br />

Also the light after the fibre was analysed using the grating spectrometer <strong>for</strong> an<br />

operating temperature of 5 ◦ C, see Fig. 4.7(a). For the saturated amplifier a spectral<br />

ASE background cannot be distinguished after the fibre, since the peak is identical<br />

with the spectrometer response function. (This response function was obtained by<br />

recording the spectrum of the narrow-linewidth EGDL laser. A similar response was<br />

also obtained using a HeNe laser.) Thus, we can only assign an upper limit of 0.2 %<br />

<strong>for</strong> the ASE contribution. The suppression ratio is ε


78 A high-power tapered semiconductor amplifier system<br />

P (mW)<br />

400<br />

200<br />

(a)<br />

fibre in<br />

out<br />

P (mW)<br />

400<br />

200<br />

(c)<br />

fibre in<br />

out<br />

transmission<br />

0<br />

0.50<br />

0.25<br />

0.00<br />

0.6 0.8 1.0 1.2<br />

I TA<br />

(A)<br />

(b)<br />

seeded<br />

no seed<br />

0.6 0.8 1.0 1.2<br />

I TA<br />

(A)<br />

transmission<br />

0<br />

0.47<br />

0.46<br />

0 10 20 30<br />

P seed<br />

(mW)<br />

(d)<br />

0 10 20 30<br />

P seed<br />

(mW)<br />

Figure 4.6: Transmission through a single-mode optical fibre. (a) Fibre input () and<br />

output () with 28mW seed, and without seed (△ ▽); (b) fibre transmission with (•)<br />

and without (◦) seed; (c) fibre input () and output () as a function of the seed power;<br />

(d) corresponding fibre transmission.<br />

(a) T =5 0 C<br />

(b)<br />

30<br />

< 82dB<br />

1.0<br />

16 o C<br />

dBm / nm<br />

20<br />

10<br />

0<br />

no seed<br />

background fraction<br />

0.5<br />

5 o C<br />

-10<br />

fiber out<br />

fiber in<br />

0.0<br />

775 780 785 790<br />

wavelength (nm)<br />

0 1 2 3 15 30<br />

P seed<br />

(mW)<br />

Figure 4.7: Spectral filtering by a single-mode optical fibre. (a) Saturation with 28mW<br />

seed power at 1.2 A current, 130 mW power after the fibre, 410 mW be<strong>for</strong>e. ASE background<br />

is not distinguishable from the spectrometer response function after the fibre.<br />

(b) The ASE fraction depends on the saturation: fibre input ( △) and output (▽) at<br />

1.2 A current. For comparison: fibre input with 1.45 A current (•).


4.6 Variations of individual gain elements 79<br />

Optimal ASE suppression required a careful alignment of the mode-matched seed<br />

input, i.e. optimization of the TA saturation, whereas achieving maximum output<br />

power was less critical. It is also obvious from the figure, that larger gain of the TA<br />

with larger operating current improved the output spectrum (•).<br />

Summarising the results of Sec. 4.4 and 4.5, the spectral properties of the TA can<br />

be optimized by choosing an appropriate operating temperature, spectral filtering<br />

with an optical fibre and saturation of the gain element.<br />

4.6 Variations of individual gain elements<br />

We compared the TD 310 gain element with two other gain elements of the same<br />

type (SDL 8630 E). One gain element (ser. no. TD 430, 777 nm) was used in the<br />

setup described above. A second (ser. no. TD 387, 790 nm) was implemented in a<br />

commercial TA system (TUI Optics, TA 100) and operated on both the D2 and the<br />

D1transition of rubidium at 780 nm and 795 nm, respectively.<br />

For the different gain elements, we found consi<strong>der</strong>able differences in their beam<br />

quality and consequently their fibre coupling efficiency. Whereas TD 310 and TD 387<br />

showed a dominant double-lobed mode structure in the far field and permitted only a<br />

fibre transmission of 46 %, the TD 430 beam showed a less pronounced lobe structure.<br />

Fig. 4.8 gives an impression of the collimated TD 430 beam profile, as imaged with<br />

a CCD video camera. Fibre coupling was achieved using the output collimator and<br />

the cylindrical lens to shape a “circular”, though slightly converging beam at the<br />

location of the fibre port. With this gain element, we could couple 59 % to the<br />

fibre and obtained 200 mW after the fibre, with an ASE suppression of better than<br />

−84 dB. Note that already at its first usage, the TD 310 displayed a shadow in the<br />

near field of its amplified output beam. After approximately 100 hours of operation<br />

the gain element quickly degraded and became inoperable.<br />

The amplification properties also showed striking differences among the gain<br />

elements. TD 430 has similar saturation properties as TD 310. For example, when<br />

seeded by a master oscillator, optimal spectral purity of the output was achieved<br />

when also the (amplified) output power was at maximum. In contrast, TD 387<br />

operates as a laser oscillator rather than an amplifier, yielding saturated output<br />

power already without seed. Although the coatings of our gain elements were not<br />

specified by the manufacturer, the difference in behaviour suggests that TD 387<br />

may have a larger reflectivity on the entrance facet, see e.g. Ref. [180]. Hence, the<br />

TD 387 requires (permanent) monitoring by a spectrometer in or<strong>der</strong> to optimize<br />

Figure 4.8: Far field beam profile of the TD 410 gain element.


80 A high-power tapered semiconductor amplifier system<br />

seed incoupling and ASE suppression. The current of the TD 387 cannot be tuned<br />

continuously, because it shows discrete “locking-ranges”, resembling the injectionlocking<br />

behaviour of single-mode diode lasers.<br />

4.7 Far off-resonance dipole potentials with<br />

spectral background<br />

In this section the consequences of a broad spectral ASE background <strong>for</strong> light scattering<br />

in optical dipole traps are estimated. A background that covers atomic resonances<br />

leads to extra resonant scattering. Usually the detuning δ <strong>for</strong> a dipole trap<br />

is chosen as large as possible, given the available laser intensity I L . The reason is<br />

that off-resonance scattering scales as Γ ′ OR ∝ I L/δ 2 at low saturation and large detuning,<br />

whereas the dipole potential is only inversely proportional to the detuning,<br />

U dip ∝ I L /δ (see e.g. Ref. [2] and Appendix A.3).<br />

In the presence of resonant background the total scattering rate of the atoms is<br />

Γ ′ =Γ ′ OR +Γ′ R , (4.1)<br />

where Γ ′ R represents the resonant scattering. For a fixed depth of the optical dipole<br />

potential this results in a maximum useful laser detuning, δ max , at which the scattering<br />

rate of the atoms, Γ ′ , is minimized. With low atomic saturation by a weak<br />

spectral background, we can write<br />

Γ ′ R ≈ Γπ εI L<br />

. (4.2)<br />

4 I 0<br />

The saturation intensity is, e.g., I 0 =1.67 mW/cm 2 <strong>for</strong> the D2-line of rubidium.<br />

Hence, with the restriction of a fixed potential U dip , the two scattering contributions<br />

scale as Γ ′ OR ∝ 1/δ and Γ′ R ∝ δ, respectively. This results in the optimum detuning<br />

and minimum scattering rate,<br />

δ max = ±Γ/ √ 2πε , (4.3)<br />

Γ ′ = 2 √ 2πε U dip / . (4.4)<br />

As an example we consi<strong>der</strong> atoms cooled to a temperature of a few µK inoptical<br />

molasses and require an optical potential depth of U dip /h ≈ 1MHz. If the allowable<br />

scattering rate is, e.g. Γ max < 100 s −1 , this yields a required background suppression<br />

ε


4.8 Conclusions 81<br />

4.8 Conclusions<br />

We have investigated a tapered semiconductor amplifier system, that provides 150−<br />

200 mW narrow linewidth output from a single-mode optical fibre, where the fibre<br />

transmission is up to 59 %, depending on the actual gain element in use. The system<br />

requires less than 5 mW seed input to saturate with an amplification up to 140 at<br />

this seed level. The output of the amplifier includes a broad spectral background<br />

of amplified spontaneous emission. We have found three means of reducing this<br />

background: (i) Choosing the operating temperature such that the gain profile of<br />

the amplifier is shifted toward the amplified wavelength, (ii) filtering the output<br />

beam spectrally with a single-mode optical fibre, and (iii) saturating the amplifier<br />

with sufficient seed input power. With these measures, the ASE background is below<br />

the resolution of our spectrometer. That is, the ASE fraction is less than 0.2 % of the<br />

optical power in the beam and the peak level is less than 0.1nW/MHz. Relating<br />

the power spectral density of the background to the natural transition linewidth<br />

of rubidium (Γ/2π = 6 MHz), the ASE suppression is better than −82 dB. The<br />

atom-optical application of such an amplifier system with far off-resonance dipole<br />

potentials was discussed. A broad ASE background implies here an optimum laser<br />

detuning with which light scattering by atoms is minimized. A tapered amplifier<br />

system may be a lower-cost alternative to a Ti:Sapphire laser. The available singletransverse-mode<br />

optical power and spectral properties are similar to those of broadarea<br />

semiconductor laser amplifiers.


82 A high-power tapered semiconductor amplifier system


5<br />

The evanescent-wave atom mirror<br />

<strong>Cold</strong> atoms (10 µK) from a vapour-cell magneto-optical trap were<br />

used to study elastic, normal-incidence bouncing on an evanescentwave<br />

mirror. Bouncing atoms were released 6 mm above the mirror<br />

and were detected by a time-of-flight technique. The fraction<br />

of bouncing atoms depends on the effective mirror surface in relation<br />

to the size of the ballistically expanding atom cloud. This<br />

fraction was investigated as a function of optical power, laser detuning,<br />

evanescent-wave polarisation, and cloud temperature. The<br />

observed bouncing fractions up to 8% were in agreement with calculated<br />

predictions, necessarily including the contribution of the<br />

Van <strong>der</strong> Waals interaction to the mirror potential.<br />

83


84The evanescent-wave atom mirror<br />

5.1 Introduction<br />

The operating principle of an evanescent-wave mirror <strong>for</strong> atoms [3–5] and loading<br />

of an optical surface trap by an inelastic mirror were discussed in Chap. 2. As an<br />

intermediate step towards this goal, we first studied an elastic mirror. Bouncing<br />

atoms from a falling cloud of cold atoms were detected by a time-of-flight (TOF)<br />

probing technique. We used the TOF method to investigate the bouncing fraction<br />

of atoms from the cloud. This fraction is determined by the effective mirror surface<br />

which is to be compared with the size of the ballistically expanding atom cloud.<br />

In the course of this chapter, TOF experiments with elastically bouncing atoms<br />

are presented. We investigated the fraction of atoms bouncing on the mirror as<br />

a function of power, detuning and polarisation of the evanescent-wave, i.e. as a<br />

function of the effective mirror surface. Also the temperature of the falling atom<br />

cloud was varied. Finally, the dependence on the evanescent-wave decay length will<br />

be discussed.<br />

5.2 Fraction of bouncing atoms<br />

5.2.1 Effective mirror surface<br />

The potential of the evanescent-wave mirror, U = U dip + U grav + U VdW , is described<br />

by the optical dipole potential, the gravitational potential and the Van <strong>der</strong> Waals<br />

interaction, see Chap. 2. The z-direction was here chosen as the vertical direction,<br />

and the xz-plane as the plane of incidence <strong>for</strong> the evanescent-wave laser, having<br />

the evanescent-wave propagating in the x-direction along a horizontal glass surface<br />

(remind Fig. 2.1, or see Fig. 5.2).<br />

The fraction η of bouncing atoms from a falling cloud depends on both the<br />

effective surface of the evanescent-wave mirror and the spatial extension of the cloud.<br />

The bouncing threshold is given by the kinetic energy of the incident atoms, U th =<br />

p 2 i /2M. Within the transverse Gaussian profile of the dipole potential, the effective<br />

mirror surface is enclosed by the circumference at which U th equals the maximum<br />

in the total potential U. Beyond this circumference, atoms slip across the potential<br />

maximum (if existing), hit the prism and are lost by heating or sticking to the<br />

surface. For a purely optical potential, this would be the circumference, where the<br />

potential at the prism surface equals U th .


5.2 Fraction of bouncing atoms 85<br />

If we take a Gaussian laser beam power P L and waist w 0 <strong>for</strong> the evanescent-wave,<br />

the intensity is given by<br />

I L (r ′ ) = I L (0) exp (− 2r′2<br />

) , I<br />

w0<br />

2 L (0) = 2P L<br />

, (5.1)<br />

πw0<br />

2<br />

where r ′ is the transverse distance from the beam centre.<br />

The optical potential of the mirror can be written in Cartesian and polar coordinates<br />

(x = r cos φ, y = r sin φ, z):<br />

( x<br />

2<br />

U dip (x, y, z) = U 0 exp (−2<br />

(χw 0 ) + y2<br />

2<br />

= U 0 exp (− 2r2<br />

w 2 0<br />

w 2 0<br />

)<br />

− 2κz) , (5.2)<br />

( )<br />

cos 2 φ<br />

+sin 2 φ − 2κz) , (5.3)<br />

In this notation, U 0 ≈ (Γ 2 /8δ)T j I L /I 0 is the maximum optical light shift in the<br />

centre of the mirror at the glass surface. For geometrical reasons, the evanescentwave<br />

waist is stretched in the x-direction by a factor χ =1/cos θ i ,comparedwith<br />

the waist w 0 of the used laser beam. In our configuration this factor is 1.334,<br />

as calculated <strong>for</strong> an evanescent-wave angle of incidence close to the critical angle,<br />

θ c =41.43 ◦ .<br />

In the hypothetical case of a purely optical potential, i.e. neglecting the Van <strong>der</strong><br />

Waals interaction, the threshold circumference R th (φ) can be calculated analytically<br />

by the condition:<br />

This leads to<br />

χ 2<br />

U th = U dip (R th (φ),φ,0) . (5.4)<br />

R th (φ) = R 0<br />

( cos 2 φ<br />

χ 2<br />

R 0 = w 0<br />

√<br />

1<br />

2 ln (<br />

U0<br />

U th<br />

)<br />

) −<br />

1<br />

+sin 2 2<br />

φ , (5.5)<br />

∝<br />

√<br />

ln<br />

( )<br />

Tj I L<br />

. (5.6)<br />

δ<br />

Here, R 0 is the threshold as it would occur <strong>for</strong> a circular evanescent wave (χ =0).<br />

In a more realistic calculation, taking into account also the Van <strong>der</strong> Waals interaction,<br />

R th (φ) can be found as follows. One first solves numerically, <strong>for</strong> the value<br />

z max ,whereU(r, φ, z max ) has its maximum as a function of z. The value of R th (φ)<br />

is found by solving also numerically<br />

U th = U(R th (φ),φ,z max ) . (5.7)


86 The evanescent-wave atom mirror<br />

1.0<br />

0.5<br />

y (mm)<br />

0.5<br />

x(mm)<br />

1.0<br />

Figure 5.1: Effective surface of the evanescent-wave mirror with Van <strong>der</strong> Waals interaction<br />

(dark shading) and without (short-dashed line). The rms width, σ i =1.13 mm, of<br />

the bouncing cloud (light shading) corresponds with a temperature of 10 µK andafall<br />

height 6.2 mm. The TE-polarised laser is chosen, with 15 mW power, 200 Γ detuning and<br />

awaistw 0 = 335 µm (dotted line). The angle of incidence is θ i = θ c +8.7 mrad.<br />

5.2.2 Ballistic spreading and bouncing fraction<br />

Having introduced the effective mirror surface, the ballistic spreading of the thermal<br />

atom cloud and the bouncing fraction are addressed in the following. If we assume<br />

a Gaussian phase-space distribution of the initial cloud, after molasses cooling to<br />

a temperature T = 10(2) µK [see Fig. 2.4(a)], the normalised spatial distribution,<br />

F i (r), of the atoms reaching the mirror at time t i can be written as:<br />

F i (r)<br />

1<br />

= exp (− r2<br />

) , 2π F<br />

2πσi<br />

2 2σi<br />

2 i (r) rdr = 1, (5.8)<br />

0<br />

√<br />

σ i = σ0 2 +(σ v t i ) 2 , (5.9)<br />

σ v =<br />

√<br />

kB T<br />

M . (5.10)<br />

In our experiment, the rms velocity spread is σ v ≃ 3cm/s. The spatial extension of<br />

the molasses cloud is estimated as σ 0 =0.25(5) mm. Thus, when reaching the prism<br />

after a fall of 35.5 ms, the cloud has expanded to a rms radius σ i =1.13(15) mm.<br />

The bouncing fraction is calculated by integrating the transverse atom distribution<br />

over the effective mirror surface:<br />

η =<br />

∫ 2π<br />

0<br />

dφ<br />

∫ Rth (φ)<br />

0<br />

∫ ∞<br />

F i (r) rdr. (5.11)


5.2 Fraction of bouncing atoms 87<br />

For a circular evanescent-wave, R th (φ) =R 0 , the bouncing fraction can be expressed<br />

analytically:<br />

( ) 2<br />

( w0<br />

)−<br />

η 0 = 1− exp (− R2 0<br />

U0 2σ<br />

) = 1−<br />

i<br />

. (5.12)<br />

2σi<br />

2 U th<br />

For a cloud that is large compared to the effective mirror surface, σ i ≫ R th ,the<br />

density F i (r) is approximately constant within this area. The bouncing fraction is<br />

then proportional to the effective surface, so that η ∝ ln(I L /δ). In Fig. 5.1, the size<br />

of the atom cloud, evanescent-wave and effective mirror surface are shown. The<br />

latter is shown with and without Van <strong>der</strong> Waals interaction. The chosen parameters<br />

represent typical experimental parameter settings. The fall height of z 0 =6.2 mm<br />

corresponds to a bouncing threshold U th =2.2 Γ (Γ/2π =6.1MHz <strong>for</strong> the rubidium<br />

D2 line). The calculated bouncing fraction is η =0.9 % with Van <strong>der</strong> Waals<br />

interaction, or 1.7 % when neglecting it.<br />

5.2.3 Optimizing the bouncing fraction<br />

Obviously, by increasing the laser power <strong>for</strong> a given laser waist w 0 , the effective<br />

mirror surface increases. This means that the turning point in the centre of the<br />

mirror moves further and further away from the surface. However, this would be inefficient<br />

use of laser power from the perspective of optimising the bouncing fraction.<br />

Increasing the waist will yield a larger effective mirror surface. We can optimise η<br />

by keeping the mirror potential above threshold across the largest area permitted by<br />

the laser power. In principle, since U dip ∝ I L /δ, the detuning could also be varied.<br />

However, the detuning also determines the rate of light scattering by atoms during<br />

a bounce so that it is usually kept fixed.<br />

If the Van <strong>der</strong> Waals interaction is neglected and assuming a transversely circular<br />

evanescent wave, the waist optimisation can be done analytically. Starting from<br />

Eq. (5.6), we substitute Eqs. (5.1), (2.8) and (2.9), and express R 0 as a function of<br />

w 0 ,withP L and δ as parameters in place of U 0 . The optimum waist, w max ,isthen<br />

found <strong>for</strong> the optimum threshold radius, R max ,as:<br />

w max = √ √<br />

T j P L Γ<br />

2 R max , R max =<br />

2 1<br />

. (5.13)<br />

8πI 0 δ e U th<br />

Note that πwmax 2 ∝ P L , as one expects, and the optimum optical potential at the<br />

prism surface is U max = e U th . The maximum bouncing fraction depends of course<br />

on the extension of the atom cloud, e.g. with Eq. (5.12) <strong>for</strong> a circular mirror this is:<br />

( ) 2 wmax<br />

η max = 1− exp (− ) . (5.14)<br />

2σ i<br />

In the course of this chapter, a numeric example of η max will be briefly discussed,<br />

related to our experimental results.


88 The evanescent-wave atom mirror<br />

5.3 Time-of-flight detection of bouncing atoms<br />

5.3.1 Mirror configuration<br />

In Fig. 5.2(a) the configuration of the evanescent-wave mirror is shown schematically.<br />

The evanescent wave is centred at the 10 × 4mm 2 sized horizontal surface of a<br />

right-angle glass prism (Melles Griot, no. 01PRB 009). <strong>Atoms</strong> were collected in<br />

a magneto-optical trap (MOT), located 6.2 mm above the prism surface. After<br />

molasses cooling, this provided a sample of ∼ 10 7 atoms at 10(2) µK temperature.<br />

The polarisation of the evanescent-wave laser beam was chosen with respect to the<br />

plane of incidence as either TE or TM. The angle of incidence, θ i − θ c , was adjusted<br />

within 0 − 40 mrad using a telescope as described in the next chapter, see Fig. 6.1.<br />

In or<strong>der</strong> to per<strong>for</strong>m time-of-flight temperature measurements, a flat, horizontal,<br />

near-resonance absorption probe beam (AP) intersected the trajectory of falling<br />

and bouncing atoms. The probe transmission was recorded by a photodiode (PD).<br />

The waist of the probe was 0.4 mm vertical and 1.4 mm horizontal (1/e 2 intensity<br />

radius). The probe was tuned 8 Γ below the D2-line transition F g =2−→ F e =3<br />

of 87 Rb (see Fig. 3.7). The optical power in the UHV cell was 0.1 µW, so that<br />

the maximum probe intensity was 0.1mW/cm 2 , well below the saturation intensity<br />

I 0 =1.67 mW/cm 2 , as required <strong>for</strong> absorption probing, i.e. the saturation parameter<br />

was s 0 =2.5 × 10 −4 ≪ 1. In or<strong>der</strong> to distinguish the TOF signal of bouncing atoms<br />

clearly from that of falling atoms, the probe was used in a higher position (1.6 mm<br />

below the MOT) than that used <strong>for</strong> the temperature measurement in Chap. 3.<br />

As a laser source <strong>for</strong> the the evanescent wave, an injection-locked single mode<br />

laser diode (Hitachi, HL 7851G98) was used, which provided 30 mW power after a<br />

single-mode optical fibre. It was seeded by an external grating diode laser, locked<br />

to the D2-line transition F g =2−→ F e = 3. The blue detuning of the evanescent<br />

wave was adjusted by frequency shifting the seed laser beam with an acousto-optic<br />

modulator. For large detuning, δ/2π >200 MHz, we unlocked the seed laser and<br />

set its frequency manually, according to the reading of an optical spectrum analyser<br />

with 1GHz free spectral range.<br />

The evanescent-wave laser beam was collimated with a waist w 0 = 335(7) µm<br />

at the mirror location. Threefold reflection losses at the walls of the vapour cell<br />

and the prism entrance surface reduced the available optical power. For an angle<br />

of incidence close to the critical angle θ c , this loss was 25 % and 5 % <strong>for</strong> TE and<br />

TM polarisation, respectively. For TE polarisation the loss was dominated by the<br />

reflections at the cell walls, whereas <strong>for</strong> TM polarisation, the angle of incidence was<br />

close to the Brewster angle. Note that the Brewster effect allowed us to keep a TMpolarised<br />

evanescent wave permanently switched on. For TE polarisation it had to<br />

be off be<strong>for</strong>e and after the bounce. Otherwise, second-or<strong>der</strong> reflections from the cell<br />

walls pushed away falling and rising atoms. In the following, all powers are given<br />

in mW, as measured in the laser beam be<strong>for</strong>e entering the UHV cell. Note that a<br />

fused silica cuvette (see Chap. 3) has the subtle advantage of a lower refractive index<br />

(n =1.45), i.e. less reflections, compared with the glass cell (n =1.51) used here.


5.3 Time-of-flight detection of bouncing atoms 89<br />

(a)<br />

(b)<br />

z(m) <br />

0 0.5<br />

1.0<br />

AP<br />

z<br />

y<br />

TM<br />

x<br />

EW<br />

MOT<br />

PD<br />

potential U/<br />

h<br />

4<br />

3<br />

2<br />

1<br />

U dip( mg= 0)<br />

U ( mg= 0)<br />

( m g= ± 1)<br />

( m = ± 2)<br />

g<br />

U th<br />

24<br />

18<br />

12<br />

6<br />

U/ h (MHz)<br />

0<br />

0<br />

0.5<br />

1.0 1.5 0<br />

surface distance z/ 0<br />

Figure 5.2: <strong>Evanescent</strong>-wave configuration and mirror potential. (a) Time-of-flight detection<br />

by a probe beam. (b) Mirror potential U <strong>for</strong> sublevels m g = {0 ...± 2}, optical<br />

potential U dip <strong>for</strong> m g =0, bouncing threshold U th = p 2 i /2M <strong>for</strong> an incident atom of momentum<br />

p i . For parameters of the TE-polarised evanescent-wave see text and Fig. 5.4.<br />

5.3.2 Mirror parameters<br />

<strong>Atoms</strong> were released from the MOT in the F g = 2 ground state. The centre-ofmass<br />

of the cloud reaches the mirror after t i =35.5 ms,6.2mmbelow. During<br />

that time, gravity accelerates the atoms to a velocity of v i =35cm/s, or 59 v rec in<br />

units of the photon recoil velocity (v rec =5.88 mm/s). Hence, the mirror potential<br />

has to exceed a bouncing threshold U th =2.2 Γ. The mirror potential U(x, y, z)<br />

is determined by the available laser power P L , the applied detuning δ, and the<br />

evanescent-wave decay length ξ(θ i ). Such a potential is plotted in Fig. 5.2(b) <strong>for</strong> the<br />

centre of the mirror (x = y = 0). The TE-polarised evanescent-wave was assumed<br />

to have 15 mW power and 200 Γ detuning. The decay length was ξ =0.89 µm<br />

<strong>for</strong> an angle of θ i = θ c +8.7 mrad. The reduction of the potential maximum by<br />

the Van <strong>der</strong> Waals interaction, compared with a purely optical potential U dip ,is<br />

also shown. In addition, due to the Clebsch-Gordan coefficients, the m g -sublevels of<br />

the F g = 2 ground state experience different light shifts. Plotted in the figure are<br />

the eigenvalues of the atom-light dipole interaction (including the Van <strong>der</strong> Waals<br />

interaction). In general, the corresponding eigenstates are linear superpositions of<br />

the sublevels. This is due to the possible longitudinal polarisation component of<br />

the evanescent-wave causing the evanescent wave to be elliptically polarised (see<br />

Appendix A.2). However, in a purely TE-polarised evanescent wave, the eigenstates<br />

can be identified by the sublevels.


90 The evanescent-wave atom mirror<br />

(a)<br />

(b)<br />

1.00<br />

1.00<br />

probe transmission<br />

0.95<br />

0.90<br />

fall<br />

rise<br />

(x10)<br />

fall again (x10)<br />

probe transmission<br />

0.95<br />

0.90<br />

(x10)<br />

0.85<br />

0.85<br />

0 20 40 60 80 100 120<br />

time (ms)<br />

0 20 40 60<br />

time (ms)<br />

Figure 5.3: Time-of-flight signals from bouncing atoms. (a) Absorption signals in the<br />

probe beam <strong>for</strong> a fall height of 4.1 mm; vertical dashed lines indicate the time when the<br />

probe was switched on. (b) Signals <strong>for</strong> a fall height of 4.8 mm and <strong>for</strong> two different heights<br />

of the probe beam, 3 mm (solid curve) and 4 mm (dashed) above the prism.<br />

A relatively large evanescent-wave decay length ∼ λ 0 is desired in our experiments,<br />

in or<strong>der</strong> to match an optical trapping potential and to adjust the rate of<br />

optical pumping by the evanescent-wave (see Chap. 2). The influence of the Van <strong>der</strong><br />

Waals interaction on the mirror is small here. The Van <strong>der</strong> Waals contribution was<br />

there<strong>for</strong>e investigated using a smaller decay length, see Ref. [103].<br />

5.3.3 Time-of-flight signals<br />

Similar to the TOF signal of falling atoms, also signals from rising atoms after a<br />

bounce can be recorded by the probe absorption. The probe is destructive since light<br />

scattering accelerates and heats the atoms. Hence, in the experiments, the probe<br />

was switched on after the average bouncing time t i , when most atoms that reach the<br />

effective mirror surface were reflected. Fig. 5.3(a) shows typical TOF signals from<br />

falling atoms, rising atoms, and atoms that fall back again. For clarity, the latter two<br />

signals were magnified ×10. The sequence of falling and rising atoms can be used<br />

to determine the bouncing time t i , and thus the fall height of the atoms. However,<br />

it then has to be clear, that the atoms were not launched with a vertical velocity<br />

component, e.g. due to imbalanced radiation pressure in the preceding molasses<br />

cooling. We there<strong>for</strong>e took also images of the MOT−prism configuration with our<br />

CCD camera. For the experiments discussed in the following, this confirmed a fall<br />

height of z 0 =6.2(4) mm, different to the heights in Fig. 5.3. Since the probe was<br />

applied at a relatively large height, it cut into the molasses cloud, causing a nonzero<br />

absorption at t = 0. As a cross-check, two signals <strong>for</strong> a different height of the<br />

probe beam are shown in Fig. 5.3(b). The signature of the bounce is found in the<br />

symmetrical temporal shift of corresponding signals from falling and rising atoms.


5.4Investigation of bouncing atoms 91<br />

bouncing fraction (%)<br />

10<br />

5<br />

U th<br />

optical potential U 0<br />

/ h<br />

5 50<br />

Detuning /:<br />

27<br />

38<br />

120<br />

200<br />

287<br />

0<br />

0.1 1<br />

power/detuning (mW/)<br />

Figure 5.4: Bouncing fraction vs. evanescent-wave power and detuning: TE-polarisation,<br />

power between 0 − 28 mW, detuning in units of Γ=2π × 6 MHz, angle θ i = θ c +8.7 mrad,<br />

laser waist w 0 = 335 µm. Predictions with (solid line) and without (dashed) Van <strong>der</strong><br />

Waals interaction. The (optical) threshold potential is indicated by an arrow.<br />

The bouncing signals are quantitatively discussed in the next section. Note that<br />

the signal of atoms falling down again after ∼ 90 ms, is weaker than <strong>for</strong> the rising<br />

atoms. Scattering of evanescent-wave photons during a bounce causes heating, and<br />

also radiation pressure, see Chap. 6. Hence, the probe beam of limited width may<br />

have covered only part of the meanwhile spread and transversely moving atoms when<br />

recording late TOF signals.<br />

5.4Investigation of bouncing atoms<br />

5.4.1 Optical power and detuning of the evanescent wave<br />

From TOF absorption signals, as shown in Fig. 5.3, the fraction of bouncing atoms<br />

is obtained by evaluating the ratio of the integral signals <strong>for</strong> rising and falling atoms.<br />

The dependence of the bouncing fraction on the strength of the optical potential is<br />

shown in Fig. 5.4. Here, <strong>for</strong> various evanescent-wave detunings, the power was also<br />

varied. Since the effective mirror surface was relatively small compared to the size of<br />

the atom cloud, the approximately logarithmic dependence, η ∝ ln P L /δ, is clearly<br />

observed. Theoretical predictions are shown <strong>for</strong> a purely optical potential (dashed<br />

line) and including the Van <strong>der</strong> Waals interaction (solid line). Both predictions<br />

were calculated without any adjustable fit parameter and take into account also<br />

the differences in optical potential <strong>for</strong> the various magnetic sublevels (see Fig. 5.2).<br />

The bouncing fraction was averaged over the calculated values <strong>for</strong> the sublevels,<br />

assuming an unpolarised atomic sample.


92 The evanescent-wave atom mirror<br />

bounce fraction (%)<br />

10<br />

8<br />

6<br />

4<br />

2<br />

(a)<br />

(various detunings)<br />

TM<br />

TE<br />

potential U / h<br />

(b)<br />

z(m) <br />

0<br />

15<br />

0.5 1.0<br />

10<br />

5<br />

TE<br />

TM<br />

U th<br />

90<br />

60<br />

30<br />

U/ h (MHz)<br />

0<br />

0.01 0.1 1<br />

0<br />

0.5 1.0 1.5<br />

power/detuning (mW/) surface distance z/ 0<br />

0<br />

0<br />

Figure 5.5: TM- and TE-polarised mirror. (a) Measured bouncing fractions and predictions<br />

with (solid lines) and without (dashed) Van <strong>der</strong> Waals interaction. (b) Mirror potentials<br />

<strong>for</strong> the F g =2eigenstates in the centre of the mirror, as calculated <strong>for</strong> P L =15mW<br />

and δ = 200 Γ; bouncing threshold U th . Optical potentials are indicated as thick lines.<br />

The extrapolation to zero bouncing fraction yields the bouncing threshold. If<br />

there were no Van <strong>der</strong> Waals interaction, this threshold would be the incident kinetic<br />

energy of the atoms, U th =2.2 Γ. The data clearly show that the Van <strong>der</strong> Waals<br />

interaction must be taken into account. For a more quantitative investigation of<br />

this phenomenon, see the work previously reported by Landragin et al. [103] and<br />

recent experimental and theoretical work, based on an evanescent-wave mirror as an<br />

interferometrical scheme [181–183]. In those experiments the evanescent-wave decay<br />

length was chosen consi<strong>der</strong>ably shorter (ξ


5.4Investigation of bouncing atoms 93<br />

It should be noted that our detection method un<strong>der</strong>-estimated the fraction of<br />

bouncing atoms. If the probe was placed too high, it cut into the optical molasses.<br />

If the probe was too low, it cut into the fast part of the bouncing signal. Our<br />

probe position of 3 mm above the prism was a compromise. Also, a horizontal<br />

misalignment of the MOT with respect to the mirror or, equivalently, launching<br />

of the atoms with a horizontal velocity causes atoms to miss the mirror. Hence,<br />

systematically a lower than expected bouncing fraction may have been observed.<br />

For convenience, and to be sure of sufficient intensity with our limited laser power<br />

of 30 mW, we restricted the optical potentials to detunings between 30 − 200 Γ. We<br />

did not optimize the laser waist to the optimum value from Eq. (5.13), which would<br />

have been w max ≈ 1−3 mm, resulting in a bouncing fraction of η max ≈ 20−80 %. In<br />

experiments that require much larger detuning ∼ 17000 Γ (100 GHz) due to photon<br />

scattering rates, more power is necessary, e.g. 150 mW from a tapered amplifier<br />

system. The optimal laser waist is then w max ∼ 200 µm, resulting in a bouncing<br />

fraction of η max ∼ 0.8 %. There<strong>for</strong>e it may be necessary to guide the atoms down to<br />

the mirror, in or<strong>der</strong> to prevent the cloud from ballistic expansion and to keep the<br />

bouncing fraction up.<br />

5.4.2 Polarisation of the evanescent wave<br />

The effective mirror surface is larger <strong>for</strong> the TE-polarised wave as compared with<br />

the TM wave. This is because of the larger intensity enhancement factor, T j ,see<br />

Eqs. (2.5) and (2.6). Figure 5.5(b) shows a typical set of potentials that contribute<br />

to the mirror. Note, that in the case of TM polarisation, there are five different<br />

eigenvalues of the optical light shift. In the configuration consi<strong>der</strong>ed here, with<br />

θ i = θ c +8.7 mrad, the optical potential ratio is U TM /U TE = T TM /T TE =2.2 ≃ n 2 .<br />

This is shown in Fig. 5.5(a), where the bouncing fraction <strong>for</strong> both polarisations is<br />

plotted vs. the laser parameters. The vertical error bars are the statistical errors from<br />

Gaussian fits which were per<strong>for</strong>med, as an approximation of the Maxwell-Boltzmann<br />

distribution, to obtain the integrated TOF signal. The horizontal separation of the<br />

two curves gives the ratio in optical potential. Note, that also the difference in<br />

reflection loss <strong>for</strong> light passing the UHV cell and entering the prism has to be<br />

taken into account. With a loss of 4.9 % and 24.7 % <strong>for</strong> TM and TE polarisation,<br />

respectively, this leads to a curve separation of ≃ 1.3 n 2 =2.9.<br />

The two data points <strong>for</strong> large P L /δ setting, that clearly do not match the prediction,<br />

we assign to an accidental systematic effect. It did not occur <strong>for</strong> the data<br />

from Fig. 5.4, taken with similar parameters. Possibly, the atoms were released with<br />

a larger temperature than previously measured (see below). There are, however, effects<br />

that can reduce the bouncing fraction <strong>for</strong> small detuning. Radiation pressure,<br />

which is investigated in Chap. 6, can cause some atoms to drift out of the detection<br />

range of the probe. Also loss of atoms by optical hyperfine pumping into F g =1<br />

has to be consi<strong>der</strong>ed. An estimate <strong>for</strong> this pumping loss is given in Chap. 6. It was<br />

also discussed by Landragin, see Ref. [188].


94The evanescent-wave atom mirror<br />

(a)<br />

10<br />

temperature T (K)<br />

20 10 6.7 5 4<br />

(b)<br />

10<br />

decay length (m)<br />

1.17 0.82 0.67 0.58<br />

bouncing fraction (%)<br />

5<br />

bouncing fraction (%)<br />

5<br />

0<br />

0 0.05 0.10 0.15 0.20 0.25<br />

1/T (1/K)<br />

0<br />

0 5 10 15 20 25<br />

angle i<br />

- c<br />

(mrad)<br />

Figure 5.6: Ballistic spread and decay length: TM-polarised evanescent-wave with P L =<br />

28 mW and δ = 113 Γ. (a) Bouncing fraction as a function of temperature and predictions<br />

with (solid line) and without (dashed line) Van <strong>der</strong> Waals interaction, <strong>for</strong> θ i = θ c +3.9 mrad.<br />

(b) Bouncing fraction <strong>for</strong> various angles. Vertical dashed lines indicate the (almost)<br />

diffraction-limited collimation of the evanescent-wave laser. The solid curve <strong>for</strong> θ i >θ c is<br />

the prediction with the Van <strong>der</strong> Waals interaction, the dashed horizontal line without.<br />

5.4.3 Ballistic spreading of the falling atom cloud<br />

The fraction of bouncing atoms is related to the size of the ballistically expanding<br />

cloud when it hits the mirror. This was investigated by adjusting the equilibrium<br />

temperature of the cloud during molasses cooling, be<strong>for</strong>e releasing it. For this purpose,<br />

the red detuning of the cooling laser was varied between 4.5 − 12.5 Γ. The<br />

temperature <strong>for</strong> each detuning was determined by fitting a Maxwell-Boltzmann distribution<br />

to the TOF signal of falling atoms (cf. Fig. 3.15). The resulting bouncing<br />

fractions are shown in Fig. 5.6(a). Again, the predictions were calculated without<br />

adjustable parameters. The influence of the Van <strong>der</strong> Waals interaction is small here,<br />

due to the relatively large decay length, ξ =1.32 µm. Between 6 −20 µK, when molasses<br />

cooling worked reliably, the data are in good agreement with the prediction.<br />

Temperatures of 6 µK and20µK result in rms cloud radii at the prism surface of<br />

0.9 mmand1.6 mm, respectively. When trying to achieve lower temperatures by<br />

larger molasses detuning, the cooling <strong>for</strong>ces may have been too weak, such that imbalances<br />

caused horizontal drift of the falling atoms. (The error margins concerning<br />

the measured temperature were estimated by ±20 %.)<br />

Theapproximate1/T dependence of the bouncing fraction is valid only <strong>for</strong> larger<br />

temperatures. In the limit of very low temperatures, the fraction “saturates” at a<br />

value that is determined by the initial width of the Gaussian atom cloud. In our<br />

configuration, this would be a fraction of 77 % with and 80 % without Van <strong>der</strong> Waals<br />

interactions.


5.4Investigation of bouncing atoms 95<br />

5.4.4 Bouncing fraction vs. decay length<br />

Measurements with varied decay length are shown in Fig. 5.6(b). The angle θ i was<br />

varied from below the critical angle up to 25 mrad above, using an optical alignment<br />

scheme that will be explained in detail in the next chapter. Close to the critical<br />

angle, θ i ≈ θ c , the bouncing fraction dropped off. This did not occur abruptly<br />

due to the nonzero divergence of the laser beam, which caused a spreading in the<br />

angle of incidence θ i and in the decay parameter κ(θ i ). We measured a far-field<br />

half-angle divergence of ∆θ i 1mrad <strong>for</strong> the collimated beam, which was close to<br />

the diffraction limit <strong>for</strong> a Gaussian beam of waist w 0 = 335 µm:<br />

∆θ dif = λ 0<br />

πw 0<br />

= 0.74 mrad . (5.15)<br />

For details on the angle calibration, see in the next chapter. Note that, however<br />

the waist was located at the prism surface to have a plane wave incident at the<br />

glass-vacuum interface, the diffraction limit causes a spread in the evanescent-wave<br />

decay parameter, κ. Below the critical angle, θ i θ c , there is no total internal<br />

reflection (TIR) except <strong>for</strong> a small fraction of the beam power. This produces a<br />

small bouncing signal in this regime. Similarly, above the critical angle, θ i θ c ,<br />

some light is transmitted. Thus, the bouncing signal is also smaller than <strong>for</strong> larger<br />

angles with pure TIR. Another effect, that reduces the observed bouncing fraction<br />

<strong>for</strong> angles close to θ c , may be radiation pressure as mentioned above.<br />

A measurement with κ as an experimentally adjustable parameter may offer<br />

a way of investigating the intrinsic scaling of the atom-surface interaction, e.g.,<br />

the 1/z 3 or 1/z 4 behaviour of the Van <strong>der</strong> Waals or Casimir-Pol<strong>der</strong> interaction,<br />

respectively. By means of κ, the surface distance of the turning point of bouncing<br />

atoms can be adjusted (cf. the atomic beam experiments in Ref. [185]).<br />

A predicted curve <strong>for</strong> the bouncing fraction including the Van <strong>der</strong> Waals interaction<br />

is shown in the figure. The (asymptotical) value <strong>for</strong> θ i −→ θ c equals the<br />

prediction without Van <strong>der</strong> Waals interaction, which is independent of κ, i.e. constant<br />

over the whole range of angles (dashed curve). Obviously, within the used<br />

range of angles, the measurements were not sensitive enough to distinguish between<br />

the predicted curves. The scatter in the observed bouncing fractions may be due<br />

the lack of a more accurate TOF reference signal of the falling atoms. However,<br />

the statistical errors of the fits to the bouncing signals were small. These statistical<br />

errors are shown in the figure.<br />

The sensitivity to the variation with the angle may be improved by an extension<br />

of the angle range. The optical access in the setup allowed us to increase the angle<br />

only up to 25 mrad. This was not a principal limitation of the setup and has been<br />

improved meanwhile. However, the decay length approaches a minimum value of<br />

110 nm and, <strong>for</strong> large angles, the variation in the bouncing fraction is negligible.<br />

An improvement towards a smaller decay length may be the use of a prism with<br />

larger refractive index. For example with SF 11 (n =1.76), the decay length drops<br />

to 86 nm and the variation in the bouncing fraction should be larger.


96 The evanescent-wave atom mirror<br />

5.5 Conclusions<br />

<strong>Cold</strong> atoms that bounce elastically and in normal incidence on an evanescent-wave<br />

mirror were detected by time-of-flight measurements. These measurements confirm<br />

the expected properties of the mirror potential in terms of the observed number of<br />

bouncing atoms as a fraction of the amount of atoms that were initially released<br />

on the mirror. We observed bouncing fractions up to 8 %, which is limited by the<br />

ballistic spread of the atoms in relation to the effective mirror surface. We varied<br />

the effective surface by means of the laser detuning and power, and adjusted the<br />

temperature of the released atom cloud between 6 − 20 µK. The measurements<br />

clearly show the significance of the Van <strong>der</strong> Waals atom–surface interaction and<br />

confirm the calculation using no adjustable parameters. The measurements were<br />

not sufficiently sensitive to reveal properties of the Van <strong>der</strong> Waals interaction in<br />

more detail.<br />

We observed larger bouncing fractions with a TM-polarised mirror than with<br />

a TE-polarised mirror, using the same laser power. This is in agreement with a<br />

calculation based on the Fresnel coefficients, which show that the evanescent-wave<br />

optical potential in the TM mode is larger by a factor ≈ 2.9 compared to the TE<br />

mode.<br />

When tuning the evanescent wave angle of incidence through the critical angle,<br />

we could observe a small bouncing fraction also <strong>for</strong> angles below the critical angle.<br />

This is due to the diffraction limited collimation of the laser beam. With the<br />

laser power available, the optimization of the effective mirror surface may result<br />

in a consi<strong>der</strong>ably larger bouncing fraction 20 %. However, our goal are experiments<br />

with much larger detuning (see Chap. 2), which requires a relatively small<br />

evanescent-wave spot.


6<br />

Radiation pressure<br />

exerted by evanescent waves<br />

Radiation pressure, that is exerted on cold rubidium atoms while<br />

bouncing on an evanescent-wave atom mirror, was directly observed.<br />

It was analysed by imaging the motion of the atoms after<br />

the bounce. The number of absorbed photons was measured<br />

<strong>for</strong> laser detunings ranging from 190 MHz to 1.4 GHz and <strong>for</strong><br />

evanescent-wave angles from 0.9 mrad to 24 mrad above the critical<br />

angle of total internal reflection. Depending on these settings, velocity<br />

changes parallel with the mirror surface were observed, ranging<br />

from 1 to 18 cm/s. This corresponds with 2 to 31 photon recoils<br />

per atom. These results were independent of the evanescent-wave<br />

optical power.<br />

This chapter is based on the publication<br />

D. Voigt, B.T. Wolschrijn, R. Jansen, N. Bhattacharya, R.J.C. Spreeuw,<br />

and H.B. van Linden van den Heuvell, Phys. Rev. A 61, 063412 (2000).<br />

97


98 Radiation pressure exerted by evanescent waves<br />

6.1 Introduction<br />

Most experimental work on evanescent-wave mirrors so far, has been concentrated<br />

on the reflective properties [4,5], i.e. the change of the atomic motion perpendicular<br />

to the surface [189]. This is dominated by the dipole <strong>for</strong>ce due to the strong gradient<br />

of the electric field amplitude. In this chapter, measurement of the <strong>for</strong>ce parallel to<br />

the surface are presented. It was mentioned already in the original proposal of an<br />

evanescent-wave mirror, Ref. [3], that there should be such a <strong>for</strong>ce. The propagating<br />

component of the wave vector leads to a spontaneous scattering <strong>for</strong>ce, “radiation<br />

pressure” [152, 153]. To our knowledge, we presented the first direct observation of<br />

radiation pressure exerted by evanescent waves on cold atoms. Previously, a <strong>for</strong>ce<br />

parallel to the surface was observed <strong>for</strong> micrometer-sized dielectric spheres moving in<br />

an evanescent-wave [190]. The basic phenomenon of radiation pressure, the photon<br />

recoil momentum, was mentioned already in 1917 by Einstein, in his work on the<br />

quantisation of the electro-magnetic field [191]. It was first observed experimentally<br />

in 1933 by Frisch [192], by means of the deflection of an atomic beam with freely<br />

propagating light. In 1976, it was proposed by Roosen and Imbert to use also a<br />

beam deflection to probe the radiation pressure of an evanescent wave [193].<br />

In our experiment, we observed the trajectory of a cloud of cold rubidium atoms<br />

bouncing on a horizontal evanescent-wave mirror. The radiation pressure appeared<br />

as a change in horizontal velocity during the bounce. We studied the average number<br />

of scattered photons per atom as a function of the detuning and angle of incidence<br />

of the evanescent wave. The latter varies the “steepness” of the optical potential.<br />

It was discussed in Chap. 2, that due to its short extension at the or<strong>der</strong> of the<br />

optical wavelength, λ 0 , an evanescent-wave mirror constitutes a promising tool <strong>for</strong><br />

loading low-dimensional optical atom traps in the vicinity of a dielectric surface<br />

[82,83,86,87]. It is this application which drives our interest in experimental control<br />

of the photon scattering of bouncing atoms.<br />

In the following section, this scattering is discussed as a source of radiation<br />

pressure by the evanescent wave. Section 6.3 describes the actual experimental configuration<br />

and the imaging method used to observe bouncing atoms. Section 6.4<br />

investigates the radiation pressure in dependence on the angle of incidence and the<br />

laser detuning, including a discussion of several systematic errors.<br />

6.2 Photon scattering by bouncing atoms<br />

In Chap. 2, the phenomenon of an evanescent wave was introduced. By total internal<br />

reflection of a laser beam, that is incident in the xz-plane, the evanescent wave was<br />

established in the horizontal xy-plane at the vacuum side of a glass surface, see<br />

Figs. 2.1and 6.1. The wave vector of the evanescent wave, k = (k x , 0,iκ), was<br />

found with a propagating component along the surface, k x = k 0 n sin θ i >k 0 ,where<br />

k 0 =2π/λ 0 is the vacuum wavenumber, n is the refractive index, and θ i is the angle of<br />

incidence. The optical dipole potential of the atom mirror, U dip (z) =U 0 exp(−2κz),<br />

is realised by choosing a blue laser detuning with respect to an atomic resonance


6.3 Observation of bouncing atoms 99<br />

and using the exponentially decaying field amplitude that is due to the imaginary<br />

wave vector component perpendicular to the surface, κ(θ i )=k 0<br />

√<br />

n2 sin 2 θ i − 1.The<br />

decay length was defined as ξ(θ i )=1/κ(θ i ).<br />

The number of scattered photons per bounce, N scat =Γp i /δκ, was obtained in<br />

Eq. (2.19) by integrating the scattering rate of an atom along the vertical bouncing<br />

trajectory (z(t),v(t)). Note, that N scat is independent of U 0 , i.e., an atom climbs the<br />

exponential mirror potential up to the turning point, no matter what the maximum<br />

optical potential at the glass surface is. The “steepness” of the optical potential is<br />

determined by κ. The steeper the potential, the shorter the time an atom spends<br />

in the light field and the smaller N scat . This behaviour is shown schematically in<br />

Fig. 6.3 <strong>for</strong> two different angles θ i .<br />

We expect that an absorbed photon gives a recoil momentum to the atom,<br />

p rec = k x ˆx , (6.1)<br />

which is directed along the propagating component of the evanescent wave. This was<br />

discussed, e.g. in Ref. [193]. Experimentally, we observed this effect by the altered<br />

horizontal velocity of atom clouds after the bounce. The spontaneous emission of<br />

photons during the scattering cycles leads also to heating of the cloud and thus to<br />

thermal expansion [194, 195]. Note, that the expression (6.1) is valid exactly only<br />

<strong>for</strong> a TE-polarised evanescent-wave. In TM polarisation, the wave is elliptically<br />

polarised and both the Poynting vector and the radiation pressure may be directed<br />

away from the propagation direction of the wave [101,193]. However, with the angle<br />

of incidence close to the critical angle θ c also the TM wave is nearly linearly polarised<br />

and the expression (6.1) may be used.<br />

In principle, N scat is changed if other than optical <strong>for</strong>ces are present. For example,<br />

the Van <strong>der</strong> Waals attraction, that was neglected in the <strong>der</strong>ivation of Eq. (2.19),<br />

tends to “soften” the potential and thus to increase N scat . We investigated this<br />

numerically and found it to be below the resolution of our detection method.<br />

6.3 Observation of bouncing atoms<br />

The radiation pressure experiment was per<strong>for</strong>med using the same optical configuration<br />

of the evanescent-wave mirror as with the bouncing fraction experiments<br />

described in the previous chapter. Also the laser systems were identical. In or<strong>der</strong><br />

to investigate radiation pressure as a function of detuning δ in a range as large as<br />

possible with the present laser power of 28 mW, we used a TM-polarised evanescent<br />

wave. In the previous chapter it was verified by means of the bouncing fraction,<br />

that this polarisation yields a stronger dipole potential than a TE-polarised beam<br />

of the same power, see also Eqs. (2.5) and (2.6).<br />

Two particular differences with the <strong>for</strong>mer setup were, however, (i) the specific<br />

use of an optical scheme to reproducibly adjust the evanescent-wave angle of incidence<br />

and, (ii) imaging of bouncing atoms instead of recording time-of-flight signals,<br />

see Fig. 6.1. A minor difference was that the magneto-optical trap (MOT) was operated<br />

at slightly larger height (6.6 mm) above the prism surface.


100 Radiation pressure exerted by evanescent waves<br />

(a)<br />

CCD<br />

FP<br />

x<br />

MOT<br />

z<br />

y<br />

(b)<br />

O<br />

L1<br />

f<br />

F<br />

f<br />

L2<br />

M<br />

2f<br />

S<br />

<br />

i<br />

i<br />

n<br />

a<br />

EW<br />

EW<br />

TM<br />

Figure 6.1: (a) <strong>Evanescent</strong>-wave mirror with fluorescence imaging. Magneto-optical<br />

trap (MOT), 6.6 mm above a prism, TM-polarised evanescent-wave beam (EW), camerafacinginthey-direction<br />

(CCD), resonant fluorescence probe beam from above (FP).<br />

(b) Confocal relay telescope <strong>for</strong> adjusting the angle of incidence θ i . The lenses L1 and<br />

L2 have equal focal length, f =75mm. The “object” spot (O) is imaged to the fixed<br />

evanescent wave-spot (S). A translation of L1 by a distance ∆a changes the angle of<br />

incidence by ∆θ i . M is a steering mirror, F is the focal plane in the telescope.<br />

(i) Angle adjustment.— Our intention was to probe the number of scattered<br />

photons, N scat ∝ ξ(θ i )/δ, as a function of decay length and detuning. There<strong>for</strong>e it<br />

was desirable to adjust the evanescent-wave angle θ i in a well defined manner with<br />

preserved calibration. In particular, a displacement of the evanescent-wave spot<br />

due to the angle adjustment was not admissible. Such a displacement leads to a<br />

systematic error in our measurements, see Section 6.4.2.<br />

The optical setup with which we adjusted the evanescent-wave angle is shown<br />

in Fig. 6.1(b). The basic idea is to image an (hypothetical) object (O) inthelaser<br />

beam (EW) to a fixed spot (S) at the prism surface, where the evanescent wave is established.<br />

The laser beam emerged from a single-mode optical fibre, was collimated<br />

and directed through a relay telescope to the prism. The angle of incidence, θ i ,was<br />

controlled by the vertical displacement ∆a of the first telescope lens, L1. This lens<br />

directs the beam, whereas the second lens, L2, images it to S. A displacement ∆a<br />

leads to a variation in θ i ,givenby:<br />

∆θ i = ∆a<br />

nf . (6.2)<br />

The refractive index n occurs here by Snel’s law <strong>for</strong> the beam entering the prism.<br />

Due to the 2f lens spacing the beam is again collimated at the evanescent-wave<br />

location, with a minimum waist of 335 µm atthespotS (1/e 2 intensity radius).<br />

The focal length of both lenses (30 mm dia.) was f = 75 mm, which allowed <strong>for</strong><br />

angles up to 25 mrad beyond θ c . When using larger lenses (40 mm dia., f = 80 mm),<br />

also angles up to 50 mrad were possible.


6.3 Observation of bouncing atoms 101<br />

Figure 6.2: Fluorescence images of a bouncing atom cloud. The first image was taken<br />

5 ms after releasing the atoms from the MOT. The contour of the right-angle prism (width<br />

10 mm) and the direction of the EW laser beam are indicated in the first frame. For<br />

comparison the horizontal placement of the MOT is also indicated in the frame (vertical<br />

dashed line).<br />

(ii) Imaging.— Compared with time-of-flight methods, the strength of imaging<br />

cold atoms lays in its potential of resolving possible horizontal motion of bouncing<br />

atom clouds. More specifically, changes in the horizontal motion are consi<strong>der</strong>ed in<br />

this chapter. (Also the mutual alignment of the MOT and the evanescent-wave was<br />

facilitated using such images.)<br />

<strong>Atoms</strong> that have bounced on the evanescent-wave mirror were detected by induced<br />

fluorescence from a pulsed probe beam in resonance with the F g =2−→<br />

F e = 3 transition of the D2 line. The probe beam had a diameter of 10 mm and<br />

was directed vertically downward. The fluorescence was recorded from the side by<br />

a digital frame-transfer CCD camera (Princeton Instruments) with a commercial<br />

objective of 50 mm focal length. The integration time was chosen between 0.1ms<br />

and 1ms, and was matched to the duration of the probe pulse. Each camera image<br />

consisted of 400×400 pixels, that were hardware-binned on the CCD array in groups<br />

of four pixels. The field of view was 10 × 10mm 2 with a spatial resolution of 51 µm<br />

per pixel. With 15 µm pixel width, this corresponded to a magnification of 0.6.<br />

A typical timing sequence of the experiment was as follows. The MOT was<br />

loaded from the background vapour during 1s. After 4 ms of polarisation gradient<br />

cooling in optical molasses the atoms were released in the F g = 2 ground state by<br />

closing a shutter in the cooling laser beams. The image capture was triggered with<br />

a variable time delay after releasing the atoms. During the entire sequence, the<br />

evanescent-wave laser was permanently on. In addition, a permanent repumping


102 Radiation pressure exerted by evanescent waves<br />

beam counteracted optical pumping of the probed atoms to the F g = 1ground<br />

state. We observed no significant influence on the per<strong>for</strong>mance of the evanescentwave<br />

mirror by the repumping light.<br />

We measured the trajectories of bouncing atoms by taking a series of images<br />

with incremental time delays. A typical series with increments of 10 ms between the<br />

images is shown in Fig. 6.2. Our detection destroys the atom cloud, so a new sample<br />

was prepared <strong>for</strong> each image. The exposure time was 0.5 ms. Each image has been<br />

averaged over 10 shots. The image at 35 ms shows the cloud just be<strong>for</strong>e the average<br />

bouncing time, t i =36.7 ms, that corresponds to the fall height of 6.6 mm. In later<br />

frames we see the atom cloud bouncing up from the surface. Close to the prism, the<br />

fast vertical motion caused blurring of the image. Another cause of vertical blur is<br />

motion due to radiation pressure by the probe pulse. The horizontal motion of the<br />

clouds was not affected by the probe. We checked this by comparing with images<br />

taken with consi<strong>der</strong>ably shorter probe pulses of 0.1ms duration.<br />

6.4The observation of radiation pressure<br />

6.4.1 Results<br />

Radiation pressure in the evanescent wave was observed by analysing the horizontal<br />

motion of the clouds. From the camera images, we determined the centre-of-mass<br />

(COM) position of the clouds to about ±1pixel accuracy. Such COM trajectories<br />

are shown in Fig. 6.3(b) <strong>for</strong> various angle settings of the evanescent-wave. We see<br />

clearly, that a steep optical potential, i.e. a small decay length, causes less radiation<br />

pressure than a shallow potential. For further quantitative investigation, in Fig. 6.4,<br />

the horizontal position was plotted vs. the time elapsed since release. We find that<br />

the horizontal motion is uni<strong>for</strong>m be<strong>for</strong>e and after the bounce. The horizontal velocity<br />

changes suddenly during the bounce as a consequence of scattering evanescent-wave<br />

photons. The change in velocity is obtained from a linear fit.<br />

In Fig. 6.5, it is shown how the radiation pressure depends on the laser detuning<br />

δ and on the angle of incidence θ i . The fitted horizontal velocity change has been<br />

expressed in units of the evanescent-wave photon recoil, p rec = k 0 n sin θ i , with<br />

k 0 /M =5.88 mm/s andn sin θ i ranging between 1and 1.03.<br />

In Fig. 6.5(a), the detuning was varied from 188 − 1400 MHz, or 31 − 233 Γ.<br />

Two sets of data are shown, taken <strong>for</strong> two different angles, θ i = θ c +0.9 mradand<br />

θ c +15.2 mrad. This corresponds to a decay length of ξ(θ i )=2.8 µm and0.67 µm,<br />

respectively. We find that the number of scattered photons is inversely proportional<br />

to δ, as expected. The predictions based on Eq. (2.19) are indicated in the figure<br />

(solid lines).<br />

In Fig. 6.5(b), the detuning was kept fixed at 44 Γ and the angle of incidence was<br />

varied between 0.9 mrad and 24.0 mrad above the critical angle θ c . This leads to a<br />

variation of ξ (θ i )from2.8 µm to0.53 µm. Here also, we find a linear dependence<br />

on ξ (θ i ). The observed radiation pressure ranges from 2 to 31photon recoils per<br />

atom. Note, that we separate this subtle effect from the faster vertical motion, in<br />

which atoms enter the optical potential with a momentum of p i ≃ 63 p rec .


6.4The observation of radiation pressure 103<br />

(a)<br />

<br />

i<br />

EW<br />

U(z)<br />

U(z)<br />

steep<br />

p i<br />

2<br />

2M<br />

shallow<br />

z<br />

vertical Z (mm)<br />

8<br />

6<br />

4<br />

2<br />

(b)<br />

0<br />

0 1 2 3<br />

horizontal X (mm)<br />

Figure 6.3: Radiation pressure as a function of mirror steepness. (a) A large evanescentwave<br />

angle θ i causes a steep potential, U(z). The incident atom momentum, p i =63p rec<br />

(v i =37cm/s), corresponds with a fall height of 6.6 mm. A bouncing atoms spends more<br />

time in a shallow potential, there<strong>for</strong>e scattering more photons. (b) Cloud trajectories<br />

of bouncing atoms observed by camera images <strong>for</strong> various angle settings. The symbols<br />

correspond with those of Fig. 6.4. The dashed arrow indicates increasing mirror steepness.<br />

4<br />

3<br />

(a)<br />

4<br />

3<br />

(b)<br />

=1.87 m<br />

X (mm)<br />

2<br />

X (mm)<br />

2<br />

1<br />

1<br />

=0.67 m<br />

0<br />

0<br />

30 40 50 60 70<br />

time (ms)<br />

30 40 50 60 70<br />

time (ms)<br />

Figure 6.4: Horizontal motion of bouncing atom clouds. The centre of mass position<br />

is plotted vs. time since release. Bouncing occurred at 36.7 ms (vertical dashed line).<br />

(a) The evanescent-wave decay length was varied as ξ(θ i )/λ 0 = {2.40, 1.32, 1.01, 0.86,<br />

0.76, 0.68}, from large to small velocity change. The detuning was 44 Γ and the power<br />

was 19 mW. (b) Comparison of two values of evanescent-wave power, 19 mW (, •) and<br />

10.5 mW (□, ◦). The detuning was 31 Γ and the evanescent-wave decay lengths were<br />

1.87 µm (2.40 λ 0 )and0.67 µm (0.86 λ 0 ). Solid lines indicate linear fits.


104Radiation pressure exerted by evanescent waves<br />

(a)<br />

(b)<br />

photon recoils N scat<br />

30<br />

20<br />

10<br />

photon recoils N scat<br />

30<br />

20<br />

10<br />

0<br />

0 0.01 0.02 0.03<br />

detuning /<br />

0<br />

0 1 2 3<br />

decay length (m)<br />

Figure 6.5: Radiation pressure on bouncing atoms expressed as number of absorbed<br />

photons, N scat . (a) Detuning δ varied <strong>for</strong> ξ =2.8 µm (◦) and0.67 µm (•). (b) evanescentwave<br />

decay length ξ varied <strong>for</strong> δ =44Γ.Thelaserpowerwas19 mW. The thin solid line is<br />

a linear fit through the first four data points. Theoretical predictions due to Eq.(2.19): twolevel<br />

atom (thick solid lines), rubidium excited-state hyperfine structure and saturation<br />

taken into account (dashed lines).<br />

In Fig. 6.4(b), we compare trajectories <strong>for</strong> 19(1) mW and 10.5(5) mW power in<br />

the evanescent wave. As expected from Eq. (2.19), there is no significant difference<br />

in horizontal motion. For a decay length of ξ =2.8 µm both power settings lead to<br />

essentially the same radiation pressure, that is 25(3) scattered photons <strong>for</strong> 19 mW<br />

and 23(2) photons <strong>for</strong> 10.5 mW. The corresponding observations <strong>for</strong> 0.67 µm decay<br />

length were 13(2) and 11(1) photons, respectively.<br />

In the previous chapter it was discussed how the optical power determines the<br />

effective mirror surface and thus the fraction of bouncing atoms. Here this was<br />

also visible in the horizontal width of imaged atom clouds. For a given evanescent<br />

wave power, there is an upper limit <strong>for</strong> the detuning, above which no bouncing can<br />

occur. For the data in Fig. 6.5(a), this threshold is calculated as δ th =6.5 GHz<br />

<strong>for</strong> ξ =0.67 µm and 8.1GHz <strong>for</strong> ξ =2.8 µm. The difference in the threshold<br />

detuning is due to the Van <strong>der</strong> Waals interaction. With our laser power, Eq. (2.19)<br />

thus predicts <strong>for</strong> our mirror a minimal (average) number of 0.25 scattered photons<br />

per atom. A second threshold condition, <strong>for</strong> fixed detuning, is indeed given by the<br />

Van <strong>der</strong> Waals interaction, which yields a lower limit <strong>for</strong> the minimally useful decay<br />

length ξ. For Fig. 6.5(b) this lower limit is calculated as ξ th = 116 nm, i.e. <strong>for</strong> an<br />

angle θ th = θ c +0.59 rad.


6.4The observation of radiation pressure 105<br />

6.4.2 Systematic errors and discussion<br />

According to Eq. (2.19), the radiation pressure should be inversely proportional to<br />

both δ and κ(θ i ). As shown in Fig. 6.5, we find deviations from this expectation<br />

in our experiment, particularly in the κ-dependence. A linear fit to the data <strong>for</strong><br />

ξ


106 Radiation pressure exerted by evanescent waves<br />

1.0<br />

calibrated angle i<br />

- c<br />

(mrad)<br />

-4 -3 -2 -1 0 1 2<br />

laser transmission<br />

0.5<br />

(x10)<br />

0.0<br />

22.4 22.6 22.8 23.0<br />

relative position lens L1 (mm)<br />

Figure 6.6: Calibration of the evanescent-wave angle. Light from the evanescent-wave<br />

laser beam, that was transmitted trough the prism surface, was detected using a power<br />

meter, in coarse (•) and fine (◦) meter range. The ×10 magnified fine reading is shown<br />

with an artificial offset (⋄). The critical angle setting (dashed line) is blurred by the<br />

diffraction limited beam collimation. The arrow indicates a setting θ i = θ c +0.9 mrad<br />

(ξ =2.8 µm), used as the smallest angle among others in the experiments.<br />

(iii) Diffuse light.— Light from the evanescent-wave can diffusely scatter and<br />

propagate into the vacuum due to roughness of the prism surface. We presume this<br />

is the reason <strong>for</strong> the extrapolated offset of ≈ 3 photon recoils in the radiation pressure<br />

[Fig. 6.5(b)]. A preferential light scattering in the direction of the propagating<br />

evanescent-wave component can be explained, if the power spectrum of the surface<br />

roughness is narrow compared to 1/λ 0 [195]. The effect of surface roughness on<br />

bouncing atoms has previously been observed [194] as a broadening of atom clouds<br />

by the roughness of the dipole potential. In our case, we observe a change in centreof-mass<br />

motion of the clouds due to an increase in the spontaneous scattering <strong>for</strong>ce.<br />

Such a contribution to the radiation pressure due to surface roughness vanishes in<br />

the limit of large detuning δ. Thus, we find no significant offset in Fig. 6.5(a). Scattered<br />

light might also be the reason <strong>for</strong> the small difference in radiation pressure<br />

<strong>for</strong> the two distinct evanescent-wave power settings, shown in Fig. 6.4(b). Lower<br />

intensity of the diffuse light implies slightly less radiation pressure.<br />

(iv) Van <strong>der</strong> Waals interaction.— As stated above, the Van <strong>der</strong> Waals interaction<br />

softens the mirror potential. This makes bouncing atoms move longer in<br />

the light field, thus enhancing photon scattering. This was investigated numerically<br />

by integrating the scattering rate along an atom’s path, including the Van <strong>der</strong><br />

Waals contribution to the mirror potential. Even with the shortest decay parameter<br />

of 0.53 µm in the present experiment, the (average) number of scattered photons


6.4The observation of radiation pressure 107<br />

would increase only about 0.8 % compared with Eq. (2.19). This was not resolved<br />

experimentally. For example, with a detuning of 1GHz and 2.5 mW power, an enhancement<br />

from N scat =1.09 to a value of 1.13 due to the Van <strong>der</strong> Waals interaction<br />

is calculated <strong>for</strong> a decay length of 370 nm (θ i = θ c +49.5 mrad).<br />

However, this result was obtained by averaging the scattered photons over the<br />

effective mirror surface. At the edges of this surface, that is at the bouncing threshold<br />

circumference, R th (φ) from Eq. (5.7), the turning point of a bouncing atom<br />

approaches the maximum of the mirror potential. There<strong>for</strong>e the calculated number<br />

of scattered photons is large, i.e., it diverges <strong>for</strong> an atom at exactly the threshold<br />

circumference.<br />

(v) Excited hyperfine state contributions.— In the two-level model, the scattering<br />

rate was expressed in the dipole potential as Γ ′ =(Γ/δ) U dip . This is no<br />

longer true if we take into account the excited state manifold F e = {0, 1, 2, 3} of<br />

87 Rb. All levels except F e = 0 contribute to the mirror potential and the scattering<br />

rate. Due to hyperfine pumping to F g = 1, part of the atoms are lost, such that a<br />

lower net radiation pressure results. Nevertheless, we observed no influence of the<br />

permanently present repumping laser on the number of scattered photons, probably<br />

because it did not saturate the repumping transition, F g =1−→ F e =2.<br />

Assuming a predominantly linear polarisation of the TM-polarised evanescentwave<br />

with θ i ≈ θ c , we can define a hyperfine correction, β HF ,tothenumberof<br />

scattered photons, N HF = β HF N scat (cf. Appendix A.3):<br />

β HF = δ 23<br />

5<br />

( ) 2<br />

∑ ∑ d 2 2,F e<br />

〈2,m g , 1, 0|F e ,m g 〉〈2,m g − j ′ , 1,j ′ |F e ,m g 〉<br />

δ 2,Fe<br />

F e<br />

+1<br />

∑<br />

j ′ =−1<br />

m g<br />

∑<br />

F e<br />

(d 2,Fe 〈2,m g , 1, 0|F e ,m g 〉) 2<br />

δ 2,Fe<br />

.(6.3)<br />

This correction averages over equally occupied ground state sublevels m g . The<br />

numerator is proportional to the partial photon scattering rate which leaves the<br />

atom in the same ground level, F g = 2. Note that the scattering amplitudes through<br />

different intermediate F e states are first added coherently, then squared [196]. The<br />

detuning <strong>for</strong> each level is assigned as δ 2,Fe . The summation over j ′ accounts <strong>for</strong> the<br />

three possible polarisations emitted in the scattering process. The denominator is<br />

proportional to the light shift, adding contributions from all excited F e levels. With<br />

an evanescent-wave detuning of δ 2,3 = 44 Γ, the correction results in a number of<br />

N HF scattered photons typically 9 % lower than expected <strong>for</strong> a two-level atom.<br />

(vi) Saturation effects.— In or<strong>der</strong> to investigate the influence of saturation on<br />

the number of scattered photons, we solved the optical Bloch equations numerically<br />

<strong>for</strong> the steady-state excited state population, σ ee<br />

(st) , see Appendix A.3. A bouncing<br />

atom encounters the evanescent wave as a light pulse with a typical duration between<br />

3and10µs. This is short compared to the natural excited state lifetime, τ =26ns.<br />

The steady-state assumption is thus justified.


108 Radiation pressure exerted by evanescent waves<br />

The temporal variation of the Rabi frequency Ω R (t) is expressed using the vertical<br />

bouncing trajectory, v z (t) =v i tanh (κv i t):<br />

ζ(t) = 1 κ ln (cosh(κv it)) . (6.4)<br />

Since the potential at the turning point is the maximum potential encountered by<br />

the atom, we have deliberately chosen the turning point as the origin, ζ =0,ofa<br />

trans<strong>for</strong>med height coordinate, ζ ≡ z − (ln U 0 /U i )/2κ. The Rabi frequency is then<br />

given as a function of ζ, asΩ R (ζ) =Ω R (0) exp(−κζ), or as function of t:<br />

Ω R (t) = Ω R (0)<br />

1<br />

cosh(κv i t)<br />

= 2<br />

√<br />

δ Ui<br />

sech(κv it) . (6.5)<br />

We can thus integrate the time-dependent scattering rate, Γ ′ (t) = Γσ ee<br />

(st) (t), <strong>for</strong><br />

a bouncing atom [cf. Eq. (A.16)]. With an evanescent-wave detuning of 44 Γ, we<br />

find approximately 7 % fewer scattered photons compared with the unsaturated<br />

expression of Eq. (2.19). Note, that the bounces occur sufficiently slowly to preserve<br />

adiabaticity. In Fig. 6.5, we show predicted curves, corrected <strong>for</strong> hyperfine structure<br />

and saturation (dashed solid lines).<br />

6.5 Conclusions<br />

We have directly observed radiation pressure that was exerted on rubidium atoms<br />

while bouncing on an evanescent-wave atom mirror. We did so by analysing the<br />

bouncing trajectories. The radiation pressure was directed parallel to the propagating<br />

component of the evanescent wave, that is, parallel to the glass surface. We<br />

observed 2−31photon recoils per atom per bounce and found the radiation pressure<br />

to be independent of the optical power in the evanescent wave, as expected from the<br />

exponential character of the evanescent wave.<br />

The inverse proportionality to both the evanescent-wave detuning and the angle<br />

of incidence is in reasonable agreement with a simple two-level-atom calculation,<br />

using steady-state expressions in the limit of low saturation <strong>for</strong> the evanescent-wave<br />

optical potential and the photon scattering rate. The agreement improved when also<br />

the excited state hyperfine structure and saturation effects were taken into account.<br />

The measured number of photon recoils as a function of the evanescent-wave decay<br />

length indicates an offset of approximately 3 recoils in the limit of a very steep<br />

evanescent-wave potential. We assume, that this is due to light that is diffusely<br />

scattered due to roughness of the prism surface but retains a preferential <strong>for</strong>ward<br />

direction parallel with the evanescent-wave propagating component.<br />

With improved resolution, it should be possible to resolve the discrete nature<br />

of the number of photon recoils and also their magnitude, k x > k 0 [197]. Our<br />

technique could also be used to observe quantum-electrodynamic effects <strong>for</strong> atoms<br />

in the vicinity of a surface, such as radiation pressure in the xy-plane but out of the<br />

x-direction of the propagating evanescent-wave component [101].


7<br />

Inelastic evanescent-wave mirrors<br />

Inelastic bouncing of cold (10 µK) rubidium atoms from an<br />

evanescent-wave mirror was observed by tuning the evanescentwave<br />

laser close to an open optical transition. The number of<br />

photons that were off-resonantly scattered by a bouncing atom<br />

was 1. The resulting optical hyperfine pumping by the evanescent<br />

wave causes dissipation. <strong>Atoms</strong> that un<strong>der</strong>go a change in hyperfine<br />

ground state jump off the mirror with reduced kinetic energy. Inelastic<br />

mirrors <strong>for</strong> 87 Rb were realised on both fine structure lines,<br />

D1 and D2. <strong>Cold</strong> atom clouds were released from 6 mm above the<br />

mirror and both elastically and inelastically bouncing atoms were<br />

detected by absorption imaging. The observed inelastic bouncing<br />

height ranged between 0.5 − 1.1 mm. The optical pumping efficiency,<br />

was adjusted between 30 − 100 % by varying the laser detuning.<br />

Using absorption imaging also the velocity distribution of<br />

inelastically bouncing atoms was investigated.<br />

109


110 Inelastic evanescent-wave mirrors<br />

7.1 Introduction<br />

<strong>Evanescent</strong>-wave mirrors <strong>for</strong> atoms [3–5] are usually designed to preserve the coherence<br />

and to provide specular reflection of atomic matter waves [194, 195]. This<br />

requires the amount of photons scattered by the bouncing atoms to be low within the<br />

typical bouncing time scale of 3−10 µs, i.e. the scattering rate should be ≪ 10 6 s −1 .<br />

Since this rate varies as ∝ 1/δ with the detuning, evanescent-wave mirrors are commonly<br />

realised with large blue detuning.<br />

In experiments that allow <strong>for</strong> a few scattered photons, evanescent waves are<br />

preferentially tuned to optical cycling (“closed”) transitions. This avoids atom loss<br />

by change of hyperfine state. To observe radiation pressure by several scattered<br />

photons, we there<strong>for</strong>e used the closed F g =2−→ F e = 3 transition on the D2 line of<br />

rubidium, see Chap. 6. On the other hand, in various applications photon scattering<br />

by bouncing atoms on an open transition is an essential part of the physical process<br />

un<strong>der</strong> investigation. Examples are:<br />

(i) loading schemes <strong>for</strong> a low-dimensional optical trap [82–84,86,87], as discussed<br />

in Chap. 2. A spontaneous optical Raman transition provides a dissipative, phasespace<br />

compressing mechanism to transfer atoms into the trap. This results in an<br />

accumulation of atoms that are decoupled from the mirror potential in a layer close<br />

to the surface. Also a two-dimensional magnetic waveguide <strong>for</strong> cold atoms using an<br />

optical loading mechanism has been proposed [85].<br />

(ii) Spontaneous Raman transitions are essential <strong>for</strong> reflection cooling of atoms<br />

by an evanescent-wave [16, 17, 106]. A single inelastic bounce can be consi<strong>der</strong>ed<br />

as a fundamental “Sisyphus” process [106], in analogy with polarisation gradient<br />

cooling [70]. Net cooling is achieved by multiple inelastic bouncing. Reflection<br />

cooling is not restricted to evanescent-wave mirrors. It has also been demonstrated<br />

in a gravito-optical trap, where cooling occurred by reflections at a hollow, conically<br />

shaped dipole trapping potential [198].<br />

(iii) Diffraction of cold atoms by an evanescent-wave grating should also be mentioned<br />

at this place. It represents an atom-optical tool, e.g. as beam splitting mechanism<br />

<strong>for</strong> atom interferometers. It was first demonstrated using atomic beams at<br />

grazing incidence [199] and later with cold atoms at normal incidence [200]. Stimulated<br />

Raman transitions between magnetic or hyperfine sublevels are inherent to<br />

the diffraction process [201].<br />

We have realised inelastic mirrors on both the D1(795 nm) and D2 (780 nm)<br />

rubidium fine structure lines, see Fig. 3.7. Our particular interest is in the D1line,<br />

since the hyperfine structure, F g = F e = {1, 2}, allows to prepare dark, far offresonance<br />

optical trapping potentials, see Chap. 2. Note that there is no hyperfine<br />

cycling transition on the D1line, so that (purely) elastic bouncing was not possible<br />

using this line with our mo<strong>der</strong>ate evanescent-wave detunings.<br />

We observed bouncing atoms directly with an absorption imaging technique that<br />

allowed us to trace the evolution of the atomic density after the bounce. We thus<br />

observed the inelastic bouncing height, in contrast to the time-of-flight detection that<br />

was employed in earlier experiments by Desbiolles et al. [106]. The bouncing height,<br />

velocity distribution and inelastic transfer efficiency are discussed qualitatively.


7.2 Principle of inelastic evanescent-wave mirrors 111<br />

F =2<br />

g<br />

U 2<br />

2<br />

F =1<br />

g<br />

z=6mm<br />

0<br />

MOT<br />

DP (RP) OP<br />

inelastic<br />

height<br />

height z / 0<br />

U 1<br />

0<br />

4<br />

6 4 2 0 0 2<br />

Figure 7.1: The inelastic evanescent-wave mirror. Hyperfine ground state potentials<br />

U 1,2 (thick curves), bouncing threshold (dashed horizontal line), optical wavelength<br />

λ 0 = 780 nm, and transition linewidth Γ=2π × 6.0 MHz. <strong>Atoms</strong> are released from a<br />

height z 0 and depumped (DP) into F g =1. Optical pumping (OP) transfers a fraction<br />

of the atoms back into F g =2, causing inelastic bouncing. Elastically bouncing atoms<br />

remain in F g =1. <strong>Atoms</strong> are exclusively detected in F g =2. <strong>Atoms</strong> in F g =1can be<br />

repumped (RP) to be also detected.<br />

potential U / h<br />

7.2 Principle of inelastic evanescent-wave mirrors<br />

Inelastic bouncing from an evanescent-wave mirror occurs when bouncing atoms<br />

dissipate potential energy, which they have acquired from their kinetic energy by<br />

climbing the mirror potential. The spontaneous process involved, is optical pumping<br />

either by the evanescent wave of the mirror or by an additional near-resonance<br />

evanescent pumping field [86]. If no additional light is provided, optical pumping<br />

can only occur when the mirror laser is tuned to an open optical transition. Working<br />

with 87 Rb, this is realised with atoms falling down in the F g = 1ground state. Using<br />

a mirror on the D1line, the evanescent wave is applied with a detuning δ 1 above<br />

the F g =1−→ F e = 2 transition. <strong>Atoms</strong> are off-resonantly excited to F e = {1, 2}<br />

by the evanescent wave and decay into F g = 2. The laser detuning relative to the<br />

F g =2−→ F e = 2 transition will be denoted as δ 2 in the following. The same<br />

notation holds <strong>for</strong> an inelastic mirror on the D2 line, since there is no dipole-allowed<br />

transition from F g = 1to the F e = 3 excited state.<br />

The bouncing process is illustrated in Fig. 7.1. Similar to the experiments discussed<br />

in the previous chapters, a sample of cold atoms (≈ 10 µK) is released in the<br />

F g = 2 ground state, 6 mm above the evanescent-wave mirror. A depumping pulse,<br />

in resonance with the open F g =2−→ F e = 2 transition, transfers all falling atoms<br />

into F g = 1. The potentials, U 1,2 , are shown in the figure [see also Eq. (2.10)]. Close


112 Inelastic evanescent-wave mirrors<br />

to the mirror, the figure is scaled in units of the optical wavelength, λ 0 = 780 nm<br />

<strong>for</strong> the D2 line. The potential in that region is determined by the evanescent-wave<br />

dipole potential and the Van <strong>der</strong> Waals interaction. The broken axis between the<br />

potential curves represents the separation by the ground state hyperfine splitting,<br />

δ GHF = 1139 Γ. The turning point of the atoms is determined by the initial gravitational<br />

potential, here U grav (z 0 )=Mgz 0 =2.1 Γ. (In Chap. 5, this potential was discussed<br />

as the bouncing threshold U th , when varying the mirror parameters.) Optical<br />

pumping by the evanescent wave transfers a fraction of the bouncing atoms back into<br />

F g =2. Sinceδ 1 ≪ δ GHF , the detuning <strong>for</strong> atoms in F g =2isδ 2 = δ 1 + δ GHF ≫ δ 1 .<br />

There<strong>for</strong>e the potential ratio is β = U 2 /U 1 ≈ δ 1 /δ 2 ≪ 1. Pumpedatomsendup<br />

in a lower potential and bounce inelastically, whereas atoms that remain in F g =1<br />

can complete the bounce elastically. <strong>Atoms</strong> in F g = 2 are detected by an absorption<br />

probe on the F g =2−→ F e = 3 cycling transition. Elastically bouncing atoms in<br />

F g = 1are detected by first repumping them into F g =2.<br />

The calculated potentials in the figure are valid <strong>for</strong> the centre of the mirror<br />

(x = y = 0) and correspond to the evanescent-wave parameters of the bouncing<br />

sequence shown in Fig. 7.3. Note that U 2 is below the threshold. Hence, most of<br />

the atoms that are pumped while on their way towards the surface hit the glass, are<br />

heated and lost.<br />

7.3 Configuration of the inelastic mirror<br />

The configuration shown in Fig. 7.2(a) is similar to that of the elastic mirror in<br />

Fig. 6.1, with a few modifications. The hypotenuse of the right-angle prism is used<br />

to couple in the additional depumping beam from below. Bouncing atoms are now<br />

observed by absorption imaging. For investigating low atomic densities this is more<br />

sensitive than fluorescence imaging, especially when a relatively strong background<br />

is present [151]. In particular, there may be a consi<strong>der</strong>able background illumination<br />

in the imaging field-of-view by evanescent light that is diffusely scattered due to<br />

roughness of the prism surface. The imaging scheme is illustrated in Fig. 7.2(b). The<br />

collimated absorption probe is directed through the sample of falling or bouncing<br />

atoms, the shadow of which is imaged on the CCD camera by a relay telescope<br />

(L1and L2, Melles Griot, glass doublets, no. 06 LAI 011/076, dia. 30 mm). In the<br />

present experiments, unity magnification was chosen. Hence, atoms were imaged<br />

with a resolution of 15 µm, equal to the CCD pixel size. A different magnification<br />

is possible by introducing a microscope objective between CCD and lens L2. For<br />

more details on this setup, see Ref. [202].<br />

The main function of the relay telescope is to translate the image to a more<br />

accessible place. It has the additional advantage that it allows the insertion of a<br />

beam stop or a phase plate in the focal plane of the telescope, with the purpose<br />

of dark field imaging or phase contrast imaging, respectively [7]. In the present<br />

experiments, the atomic density was too low <strong>for</strong> the use of imaging techniques that<br />

are nondestructive to the atomic sample [149, 150]. For a discussion of the various<br />

techniques see, e.g., Ref. [47].


7.3 Configuration of the inelastic mirror 113<br />

(a)<br />

z<br />

(b)<br />

CCD<br />

x<br />

MOT<br />

y<br />

AP<br />

S<br />

L1<br />

L2<br />

CCD<br />

DP<br />

EW<br />

AP<br />

TM<br />

prism<br />

100 mm 200 mm 100 mm<br />

Figure 7.2: Configuration of the inelastic mirror. (a) Falling atoms from a MOT are<br />

depumped (DP) into F g =1. Inelastically bouncing atoms are detected in F g =2by an<br />

absorption probe (AP). A repumping beam (not shown) optionally transfers elastically<br />

bouncing atoms into the detectable F g =2state. (b) Absorption imaging: A collimated<br />

probe beam is directed through the atomic sample (S) onto the CCD detector. The sample<br />

is imaged by a relay telescope of unity magnification. The focal length of the achromatic<br />

lenses (L1,L2) is 100 mm.<br />

The probe had a waist of approximately 5 mm (1/e 2 intensity radius), with<br />

apowerof∼ 100 µW. The frequency was chosen in resonance with the cycling<br />

transition F g = 2 −→ F e = 3 on the D2 line. The saturation parameter was<br />

s 0 0.2. The probe exposure time τ ex was chosen between 20 − 70 µs, so that an<br />

atom scatters ≈ 200 photons. Longer exposure is not useful since ∼ 400 photon<br />

recoils from the probe are sufficient to Doppler-shift the atom out of resonance<br />

(400 k 0 v rec ≈ Γ/2). Furthermore the image would be blurred by atomic motion.<br />

The maximum velocity in the experiments, v i ≈ 60 v rec , together with an imaging<br />

resolution of 15 µm allows a maximum exposure time of τ ex ∼ 40 µs. In or<strong>der</strong> to<br />

achieve quantitatively accurate absorption data, one should keep s 0 ≪ 1.


114Inelastic evanescent-wave mirrors<br />

7.4Observation of inelastically bouncing atoms<br />

7.4.1 Inelastic bouncing height<br />

In the experiments discussed here, an inelastic mirror was first realised with the<br />

evanescent field on the open transition F g =1−→ F e =2oftheD2lineof 87 Rb.<br />

The evanescent-wave was TM-polarised with a waist of 0.5 mm (1/e 2 )and26mW<br />

power from an injection-locked single-mode diode laser. The angle of incidence was<br />

varied between 1.8 mrad and 18 mrad beyond the critical angle and the detuning δ 1<br />

between 70 Γ and 230 Γ.<br />

Asampleof∼ 10 7 atoms was loaded within 2 s in the MOT, followed by 5 ms<br />

of molasses cooling to a temperature of 10 µK. A typical image sequence displaying<br />

inelastic bounces is shown in Fig. 7.3 <strong>for</strong> an angle of θ i = θ c +1.8 mrad (decay length<br />

2.5 λ 0 ) and a detuning of δ 1 = 150 Γ. Due to the destructive character of the probe,<br />

each frame was taken in a new realisation of the experiment. The time indicated <strong>for</strong><br />

each frame is the time elapsed since shuttering the cooling light. Sequence (a) shows<br />

the falling and expanding thermal cloud. The irregular shape of the cloud in the<br />

first frame, taken immediately after release, may be a consequence of imbalanced<br />

molasses cooling <strong>for</strong>ces. Some saturated CCD pixels appear as white spots.<br />

In or<strong>der</strong> to observe inelastic bounces, 4 − 27 ms after releasing the cloud, the<br />

depumping laser was switched on <strong>for</strong> about 2 ms. The upward directed radiation<br />

pressure of the depumping beam transfers a few photon recoils to the atoms. This<br />

reduces the incident velocity v i at the mirror by approximately 3 %. The continued<br />

image sequence, (b), with inelastically bouncing atoms starts at 35 ms, when the<br />

cloud centre-of-mass hits the mirror. In the following frames, the bouncing cloud<br />

leaves the mirror and reaches its maximum height of 0.8 mm at t ≈ 47 ms. This was<br />

14 % of the MOT height z 0 and in reasonable agreement with the potential ratio<br />

β ≈ δ 1 /δ 2 =0.12.<br />

The transfer of atoms into F g = 2 preferentially occurs while atoms are near the<br />

turning point, where a relatively long time is spent in a region of a strong optical<br />

field. Hence, we can indeed expect a well established peak in the vertical column<br />

density of inelastically bouncing atoms, as it is obvious from the sequence shown. In<br />

addition, a tail of atoms is visible, stretching out to a height expected <strong>for</strong> elastically<br />

bouncing atoms only. This tail is caused by atoms, that were transferred into F g =2<br />

either be<strong>for</strong>e reaching the turning point or after having partly reaccelerated off the<br />

mirror potential. The bouncing dynamics, together with the stochastic nature of<br />

optical pumping, thus cause a broad redistribution in atomic velocities. The velocity<br />

distribution translates into the imaged spatial distribution after 10 − 20 ms of free<br />

flight.<br />

The background visible in the images is due to imperfections of the detection<br />

setup. The observed density of bouncing atoms is significantly lower than that of the<br />

falling atoms. (This is partly due to the projection on the image plane.) There<strong>for</strong>e<br />

the image contrast was enhanced in the sequence (b) by reducing the gray scale<br />

display range by an or<strong>der</strong> in magnitude. Hence, the background appeared in these<br />

images. The white region at the <strong>for</strong>mer location of the MOT is an artifact, possibly


7.4Observation of inelastically bouncing atoms 115<br />

(a)<br />

0ms 15ms 19ms 23ms 27ms 31ms<br />

(b)<br />

35 ms 39 ms 43 ms 47 ms 51 ms 55 ms<br />

Figure 7.3: Time sequence of inelastically bouncing atoms. (a) <strong>Atoms</strong> fall from the<br />

MOT (in the F g =1state). Gray scale indicates atomic density. The field of view is 8 mm<br />

in height, and the prism surface is indicated by a horizontal dotted line. The first frame<br />

was taken immediately after switching off the molasses cooling light. Subsequent frames<br />

each represent a new realisation of the experiment. After 35 ms the cloud centre-of-mass<br />

hits the prism. (b) Inelastically bouncing atoms were detected in F g =2.<br />

due to a memory-effect of the CCD array, caused by the intense illumination from<br />

molasses cooling light, several 10 ms be<strong>for</strong>e an image capture. Interference fringes<br />

and circular patterns stem from the probe laser and were due to reflections at the<br />

uncoated UHV cell and due to diffraction from dust particles. Each absorption<br />

image was the result of three image captures, taken shortly after each other. First,<br />

the atoms were probed. A second frame was similarly taken without loading the<br />

MOT as a zero-absorption reference. From both images the background illumination<br />

was subtracted, as captured by the third frame without using the probe pulse.<br />

Division of signal and reference image results in the absorption image. The noise<br />

level in the atomic signals was reduced by averaging 5 realisations <strong>for</strong> each image.<br />

In principle, no fringes should occur, unless the reflecting optical surfaces move in<br />

the time between recording the signal image and the reference image. Accumulating<br />

more realisations may be a remedy to average out drifting fringe patterns.


116 Inelastic evanescent-wave mirrors<br />

7.4.2 Atom density and transfer efficiency<br />

An inelastic mirror on the D1line (795 nm) of 87 Rb was realised using the tapered<br />

amplifier system with the TD 387 gain element as a laser source <strong>for</strong> the evanescentwave,<br />

see Chap. 4. Due to the larger available power of 73 mW, as compared to<br />

the injection-locked diode laser, the mirror could be established with a larger laser<br />

waist of 0.8 mm (1/e 2 ). The laser was again TM-polarised. The detuning was<br />

70 − 300 Γ above the F g =1−→ F e = 2 resonance, and the angle was θ i = θ c +<br />

16.6 mrad, which resulted in a decay length of 0.82 λ 0 (0.65 µm). This atom mirror<br />

was used to investigate the efficiency of transferring bouncing atoms into F g =2asa<br />

function of evanescent-wave detuning and decay length. A detailed study including<br />

a numerical analysis will be presented elsewhere, see Refs. [202,203]. In this section,<br />

the experimental results are discussed.<br />

Density of bouncing atoms.— In or<strong>der</strong> to quantitatively investigate bouncing<br />

atoms, we converted the absorption images into the corresponding atomic column<br />

density distributions in the xz-plane (see Appendix A.4). The 2D gray-scale density<br />

plots of Fig. 7.4 represent column densities. In the left image only inelastically<br />

bouncing atoms (N 2 ) were detected. In the right image repumping light was supplied<br />

be<strong>for</strong>e detection, so that all atoms were detected (N 1 +N 2 ). In these measurements,<br />

the evanescent-wave detuning was δ 1 = 200 Γ and the images were averaged over 10<br />

experimental runs. The vertical 1D (linear) density ρ z (z) was obtained by summing<br />

lines of the 2D image <strong>for</strong> different values of x. In Fig. 7.4(a) also the density of elastically<br />

bouncing atoms is shown, as <strong>der</strong>ived from the combined signal in Fig. 7.4(b).<br />

The linear densities were normalised to the atom numbers, N 1 and N 2 , by integrating<br />

the column density distributions and using the resonant rubidium absorption<br />

cross section σ 0 =3λ 2 0/2π, see Appendix A.4. Our absorption probe was linearly π-<br />

polarised and we assumed that the atoms were randomly distributed over the F g =2<br />

ground state magnetic sublevels m g = {0 ...± 2}. The absorption cross section was<br />

there<strong>for</strong>e averaged over these m g -levels using the Clebsch-Gordan coefficients <strong>for</strong> the<br />

F g =2−→ F e = 3 transition and the reduced dipole matrix element, here d 2,3 =1:<br />

σ = 1 5 σ 0d 2 2,3<br />

∑<br />

m g<br />

〈2,m g , 1, 0|3,m g 〉 2 = 7<br />

15 σ 0 = 1 3.5 × 10 −10 cm 2 . (7.1)<br />

Whereas the column density represents the measured quantity, the physically interesting<br />

quantity is the 3D spatial density, ρ(r) =ρ x (x)ρ y (y)ρ z (z). It is shown<br />

as ρ(0, 0,z) by an alternative density scaling in Fig. 7.4 and was calculated un<strong>der</strong><br />

the assumption that the horizontal distributions were Gaussians. Due to the aspect<br />

ratio, χ ≈ 1.3, of the elliptical effective mirror surface, the rms width of the cloud<br />

in the x-direction is wi<strong>der</strong> by a factor χ compared to the y-direction.<br />

The asymmetrical vertical distribution, ρ z (z), originates from the distribution of<br />

velocities at which atoms leave the surface. It is evident from the peaked structure,<br />

that there is a strong preference <strong>for</strong> atoms to be pumped when they are slow, i.e.<br />

close to the turning point on the mirror. For a hypothetic monochromatic sample<br />

(without spreading in v i ), the distribution would be sharply edged, since the turning


7.4Observation of inelastically bouncing atoms 117<br />

linear density z<br />

(z) (10 5 /mm)<br />

1.5<br />

1.0<br />

0.5<br />

0<br />

1.5<br />

1.0<br />

0.5<br />

0<br />

(a) 200 <br />

N 1,2<br />

= 1.5x10 5<br />

(b) 200 <br />

N 1<br />

+N 2<br />

= 3.0x10 5<br />

( N 2 )<br />

( N+ N)<br />

1 2<br />

6 5 4 3 2 1 0<br />

height above prism (mm)<br />

4<br />

2<br />

0<br />

4<br />

2<br />

0<br />

spatial density (z) (10 8 /cm 3 )<br />

z<br />

( N 2 ) ( N+ N)<br />

x<br />

1 2<br />

6mm<br />

z<br />

(z) (10 5 /mm)<br />

1.5<br />

1.0<br />

0.5<br />

0<br />

(c) 100 <br />

N 1<br />

= 0.7x10 5<br />

N 2<br />

= 2.2x10 5<br />

6 5 4 3 2 1 0<br />

height above prism (mm)<br />

3<br />

2<br />

1<br />

0<br />

(z) (10 8 /cm 3 )<br />

Figure 7.4: Vertical column density of atoms, 4 ms after bouncing. <strong>Evanescent</strong>-wave<br />

tuned to the D1 line of 87 Rb: (a) Inelastically bouncing atoms detected in F g = 2<br />

(N 2 , thick curve). (b) All atoms (N 1 + N 2 ) were detected with additional repumping<br />

of elastically bouncing atoms. The elastic contribution (N 1 ) is obtained by subtraction<br />

[thin curve in (a)]. The absorption images corresponding to the line sums in (a) and (b),<br />

are also shown. The prism surface is indicated by a dotted line. (c) Densities obtained<br />

with a different evanescent-wave detuning (see Ref. [203]).<br />

point defines the smallest possible inelastic velocity ≈ √ βv i . Obviously, the tail of<br />

fast atoms has also an edge, since the fastest atoms can just reach the MOT height<br />

z 0 . The velocity distribution of inelastically bouncing atoms cannot be described in<br />

terms of a thermal Maxwell-Boltzmann distribution, as is the case with elastically<br />

bouncing atoms. The large spread in velocities suggests “heating” of the cloud,<br />

if one would assign a temperature at all. More useful may be an investigation of<br />

atomic phase-space density. Another feature of the bouncing dynamics is revealed<br />

by a closer look at the evolution of fast atoms in the sequence of Fig. 7.3(b). From<br />

51ms on, fast atoms were still rising and separating from slower atoms that fall<br />

down again. Indeed, numerical analysis indicates that the larger velocities were<br />

slightly more populated than medium velocities, see Ref. [203].<br />

Transfer efficiency.— When investigating radiation pressure on elastically<br />

bouncing atoms in Chap. 6, a simple analytical model <strong>for</strong> two-level atoms led to<br />

Eq. (2.19) <strong>for</strong> the number of scattered photons on the cycling transition, N scat ∝ 1/δ.<br />

The same result can be used to estimate the transfer efficiency into F g =2byinelastic<br />

bouncing. For comparison, the elastic and inelastic contributions to the atom<br />

density are shown in Fig. 7.4(c) <strong>for</strong> a smaller detuning of δ 1 = 100 Γ. It is obvious


118 Inelastic evanescent-wave mirrors<br />

that a larger fraction of atoms were pumped, namely 80 % of the atoms ended up<br />

in F g = 2, compared to 50 % <strong>for</strong> 200 Γ, so that less atoms completed the bounce<br />

elastically. Note that, although the peak linear density in F g = 2 is larger <strong>for</strong> 100 Γ,<br />

the peak spatial density is similar to that <strong>for</strong> 200 Γ. This is due to the larger effective<br />

mirror surface, i.e. a larger bouncing fraction <strong>for</strong> smaller detuning (see Chap. 5).<br />

The 3D density is there<strong>for</strong>e distributed broa<strong>der</strong> in the lateral directions <strong>for</strong> smaller<br />

detuning. The transfer efficiency can be estimated as<br />

N 2<br />

= 1− q Nscat , (7.2)<br />

N 1 + N 2<br />

where q ≈ 0.5 is the branching ratio to the ground states, defined as the fraction that<br />

goes into F g = 2 (cf. Ref [16]). Thus with N scat , we can expect an efficiency of 37 %<br />

and 60 % using a detuning of 200 Γ and 100 Γ, respectively. A more adequate model<br />

should include the excited state hyperfine structure, using expressions similar to that<br />

of Eq. (6.3) <strong>for</strong> the radiation pressure hyperfine correction β HF . Also depumping of<br />

atoms back into F g = 1has to be consi<strong>der</strong>ed, see Ref. [202].<br />

Finally, note that due to the lack of any cycling transition on the D1line, elastic<br />

bouncing with a larger number of scattered photons, N scat ≫ 1, as presented with<br />

the radiation pressure investigations of Chap. 6 is not possible. Of course, in the<br />

limit of large detuning, δ 1,2 ≫ δ GHF , all atoms bounce elastically (β → 1). However,<br />

also the radiation pressure is then negligible.<br />

7.5 Conclusions<br />

Inelastic mirrors <strong>for</strong> cold rubidium atoms were realised using evanescent-wave optical<br />

potentials tuned near an open optical transitions of the D1(795 nm) or D2 (780 nm)<br />

line of 87 Rb, thus introducing spontaneous Raman transitions between hyperfine<br />

ground states. Bouncing atom clouds were directly observed by absorption imaging.<br />

The evolution of the peak atomic density reveals the inelasticity of the reflection<br />

on the mirror, e.g. loss of kinetic energy ranging between 81 − 92 %. The dynamics<br />

of the internal state transfer of bouncing atoms causes a broadened non-thermal<br />

atomic velocity distribution, that is observed as a tail of fast atoms in absorption<br />

images of bouncing atoms. This suggests that, although the observed single inelastic<br />

bounce represents a fundamental step of a “Sisyphus” reflection cooling mechanism,<br />

it involves heating (and thus a reduction in phase-space density). Only a succession<br />

of multiple bounces leads to a net cooling effect and, finally, establishes a thermal<br />

barometric density distribution of atoms at a temperature lower than the initial<br />

one [17]. Note that in the proposed low-dimensional trapping scheme of Chap. 2<br />

the phase-space density already piles up by a single bouncing process. This is<br />

due to spatially selective pumping in combination with a trapping potential that<br />

accumulates atoms in the vicinity of the surface. Further experimental investigations<br />

which are in progress, have to show whether pumping by the evanescent-wave mirror<br />

alone, can be used to efficiently optimise a trap loading scheme. For the envisaged<br />

very far detuned evanescent waves it may be necessary to introduce an additional<br />

near-resonance evanescent-wave contribution in or<strong>der</strong> to adjust optical scattering<br />

rates.


A<br />

Appendix<br />

A.1 Useful atom-optical numbers <strong>for</strong> 87 Rb<br />

Spectroscopy: [143, 144]<br />

D1 line λ 0 (5s 2 S 1/2 → 5p 2 P 1/2 ) 795.0 nm<br />

natural lifetime τ ≡ 1/Γ 27.70(4) ns<br />

natural linewidth Γ/2π 5.75 MHz<br />

saturation intensity I 0 ≡ πhcΓ/3λ 3 0 1.49 mW/cm 2<br />

D2 line λ 0 (5s 2 S 1/2 → 5p 2 P 3/2 ) 780.2 nm<br />

τ 26.24(4) ns<br />

Γ/2π 6.07 MHz<br />

I 0 1.67 mW/cm 2<br />

Laser cooling (D2 line):<br />

Doppler temperature T D ≡ Γ/2k B 146 µK<br />

Doppler capture velocity Γ/k L 4.7 m/s<br />

typ. Doppler velocity v D ≡ √ 2k B T D /M 16.7 cm/s<br />

recoil temperature T R ≡ (k L ) 2 /Mk B<br />

=2E R /k B 361 nK<br />

recoil velocity v R ≡ k L /M 5.88 mm/s<br />

recoil frequency ω R ≡ E R / = kL 2 /2M 2π × 3.77 kHz<br />

thermal De Broglie Λ ≡ h/ √ 2πMk B T 15.5 nm ( T D )<br />

wavelength 312 nm ( T R )<br />

Gravitation: Mg/k B 1.03 mK/cm<br />

Mg/h 21.4 MHz/cm<br />

Mg/µ B 15.3 G/cm<br />

Atomic collisions: [63] a 2,2 (s-wave scattering, 109(10) a 0<br />

a 1,−1 length a F,m ) 106(6) a 0<br />

General constants: h/k B 48.0 µK/MHz<br />

k B /h 20.8 kHz/µK<br />

µ B /h 1.40 MHz/G<br />

µ B /k B 67.2 µK/G<br />

0.67 K/T<br />

g 0.98 (cm/s)/ms<br />

119


120 Appendix<br />

A.2 Fresnel coefficients <strong>for</strong> evanescent waves<br />

Fresnel coefficients are usually <strong>der</strong>ived <strong>for</strong> the (complex) reflection and transmission<br />

coefficient of light incident with an angle θ i θ c , i.e. with a complex “transmission angle”.<br />

Snel’s law, sin θ t = n sin θ i > 1, then is written as<br />

cos θ t = i √ n 2 sin 2 θ i − 1 .<br />

(A.1)<br />

The wave vectors and polarisations maintain their common <strong>for</strong>m using the complex<br />

angle θ t . The wave vectors are<br />

k i = nk 0 (sin θ i , 0, cos θ i ) , (A.2)<br />

k t = k 0 (sin θ t , 0, cos θ t ) . (A.3)<br />

The polarisations are<br />

s i = s t = (0, 1, 0) , (A.4)<br />

p i = (− cos θ i , 0, sin θ i ) , (A.5)<br />

p t = (− cos θ t , 0, sin θ t ) . (A.6)<br />

Note that p t is not normalised in the usual way, p ∗ t·p t ≠ 1. Instead, it obeys the<br />

normalisation p t·p t = 1. The reflection and transmission coefficients also keep the<br />

common <strong>for</strong>m,<br />

r s = n cos θ i − cos θ t<br />

n cos θ i +cosθ t<br />

= n cos θ i − i √ n 2 sin 2 θ i − 1<br />

n cos θ i + i √ n 2 sin 2 θ i − 1 ,<br />

(A.7)<br />

t s =<br />

2n cos θ i<br />

n cos θ i +cosθ t<br />

=<br />

2n cos θ i<br />

n cos θ i + i √ n 2 sin 2 θ i − 1 ,<br />

(A.8)<br />

r p = cos θ i − n cos θ t<br />

cos θ i + n cos θ t<br />

= cos θ i − in √ n 2 sin 2 θ i − 1<br />

cos θ i + in √ n 2 sin 2 θ i − 1 ,<br />

(A.9)<br />

t p =<br />

2n cos θ i<br />

cos θ i + n cos θ t<br />

=<br />

2n cos θ i<br />

cos θ i + in √ n 2 sin 2 θ i − 1 .<br />

(A.10)<br />

However the transmission occurs into the evanescent wave, t s and t p are proportionality<br />

factors between the incident and the evanescent field amplitude. Indeed we<br />

find |r s,p | =1.


A.3 Light <strong>for</strong>ces and scattering rate 121<br />

A.3 Light <strong>for</strong>ces and scattering rate<br />

A.3.1<br />

Two-level atoms<br />

A detailed description of the atom-light interaction can be found, e.g., in Ref. [2].<br />

We assume an atom with a ground state |g〉 andanexcitedstate|e〉 of lifetime<br />

τ =1/Γ, where Γ is the natural transition linewidth, e.g. with Γ/2π =6.1MHz <strong>for</strong><br />

the rubidium D2 line. The states are separated by ω 0 , and the detuning of a laser<br />

frequency ω L is defined as δ = ω L − ω 0 . A useful expression in the description of the<br />

atom-light coupling is the Rabi frequency <strong>for</strong> a given laser intensity I L :<br />

Ω R = Γ<br />

√<br />

IL<br />

2I 0<br />

.<br />

(A.11)<br />

It describes the resonant (δ = 0) cycling frequency between the ground and excited<br />

state population. The saturation intensity is defined as I 0 = πhcΓ/3λ 3 0,withthe<br />

optical wavelength λ 0 .<br />

Scattering rate and spontaneous <strong>for</strong>ce.— The atomic scattering rate is obtained<br />

by solving the “Optical Bloch Equations” (OBE). These describe the evolution<br />

of the density operator ˆσ of an atom coupled to the light field. We obtain<br />

the OBE’s <strong>for</strong> the Bloch vector, (u, v, w), by the elimination of the fast evolution<br />

∝ exp(iω L t) of the laser oscillation in the “rotating-wave approximation”, and using<br />

the trans<strong>for</strong>m {ˆσ ge , ˆσ eg , ˆσ gg , ˆσ ee } = {σ ge exp (−iω L t),σ eg exp (iω L t), σ gg ,σ ee }:<br />

u = 1 2 (ˆσ ge +ˆσ eg ) , ˙u = δv− Γ u, (A.12)<br />

2<br />

v = 1 2i (ˆσ ge − ˆσ eg ) , ˙v = −δu− Ω R w − Γ v, (A.13)<br />

2<br />

w = 1 2 (ˆσ ee − ˆσ gg ) , ẇ = Ω R v − Γw − Γ 2 . (A.14)<br />

The component w describes half the population inversion between the atomic states.<br />

A useful notation is also the saturation parameter:<br />

s 0 = 1 2<br />

Ω 2 R<br />

δ 2 + ( )<br />

Γ 2<br />

=<br />

2<br />

1<br />

1+ ( 2δ<br />

Γ<br />

) 2<br />

I L<br />

I 0<br />

.<br />

(A.15)<br />

As a steady-state solution <strong>for</strong> the scattering rate Γ ′ , i.e. the excited state population<br />

, we find:<br />

σ (st)<br />

ee<br />

Γ ′ = Γσ (st)<br />

ee = Γ 2<br />

s 0<br />

1+s 0<br />

= Γ 2<br />

1+ ( 2δ<br />

Γ<br />

1<br />

) 2<br />

+<br />

I L<br />

I0<br />

I L<br />

I 0<br />

.<br />

(A.16)<br />

The recoil, k L , from absorbed photons causes radiation pressure or, the “spontaneous<br />

light <strong>for</strong>ce”, F sp = k L Γ ′ . This <strong>for</strong>ce saturates <strong>for</strong> s 0 ≫ 1as F sp = k L Γ/2.


122 Appendix<br />

Far off-resonance dipole potentials.— In the limit of large detuning,<br />

|δ| ≫Γ, we can approximate the saturation parameter by s 0 ≈ (Γ/2δ) 2 I L /I 0 .Ifalso<br />

|δ| ≫ Ω R , the eigenstates of the atom-light interaction approach the uncoupled<br />

states, |g〉 and |e〉, and the coupling to the field effectively causes a “light shift” of<br />

these states or, a “dipole potential”. If in this limit also the saturation parameter<br />

is small, s 0 ≪ 1, the (ground state) light shift and the scattering rate are given as:<br />

1<br />

U dip ≈ 1 2 s 0δ ≈ Ω2 R<br />

4δ<br />

= Γ2<br />

8δ<br />

Γ ′ ≈ 1 2 s 0Γ ≈ Ω2 R Γ<br />

4δ 2 = Γ3<br />

8δ 2<br />

I L<br />

I 0<br />

,<br />

I L<br />

I 0<br />

.<br />

(A.17)<br />

(A.18)<br />

The ratio of light shift and scattering rate is now simply U dip /Γ ′ ≈ δ/Γ. Since<br />

the light shift is usually spatially varying, its gradient represents the “dipole <strong>for</strong>ce”,<br />

F dip (r) =−∇U dip (r).<br />

A.3.2<br />

Multilevel atoms — rubidium hyperfine structure<br />

In the interaction of a multilevel atom with a laser field, the polarisation state of<br />

the light has to be consi<strong>der</strong>ed together with the coupling strengths of the various<br />

optical transitions between atomic sublevels. For example, the D2 line of 87 Rb is<br />

a J g =1/2 −→ J e =3/2 transition. The coupling to the nuclear spin, I =3/2,<br />

results in the hyperfine structure with the ground and excited states F g = {1, 2}<br />

and F e = {0, 1, 2, 3}, respectively, shown in Fig. 3.7.<br />

Spherical polarisation basis.— The expression (A.11) <strong>for</strong> the Rabi frequency<br />

has its origin in the coupling of the atomic dipole moment to the electric field,<br />

written as E =(1/2)ˆɛ E exp (−iωt)+c.c.:<br />

Ω R = 2 d · ˆɛ E<br />

<br />

. (A.19)<br />

The matrix element of the dipole operator D is here d = 〈g|D|e〉, the electric field<br />

amplitude is E, and the field polarisation is given by the unit vector ˆɛ.<br />

Itmaybeusefultoworkinasphericalbasis{ˆɛ − , ˆɛ 0 , ˆɛ + }, that is defined in the<br />

cartesian basis {ˆx, ŷ, ẑ} as:<br />

⎛<br />

ˆɛ − = √ 1 ⎝<br />

2<br />

1<br />

0<br />

−i<br />

⎞<br />

⎛<br />

⎠ , ˆɛ 0 = ⎝<br />

0<br />

1<br />

0<br />

⎞<br />

⎛<br />

⎠ , ˆɛ + = √ 1 ⎝<br />

2<br />

−1<br />

0<br />

−i<br />

⎞<br />

⎠ .<br />

(A.20)<br />

These basis vectors describe σ − , π, andσ + -polarised light with respect to the deliberately<br />

chosen y-direction. The dipole operator can now be expressed in the<br />

spherical basis, D j = D · ˆɛ j ,wherej = {0, ±1}.


A.3 Light <strong>for</strong>ces and scattering rate 123<br />

Reduced dipole matrix elements.— Writing ground and exited state of a<br />

rubidium atom as |F g ,m g 〉 and |F e ,m e 〉, the Wigner-Eckart theorem is applied to<br />

factorise the dipole matrix element:<br />

〈F g ,m g |D j |F e ,m e 〉 = 〈F g |D|F e 〉〈F e ,m e , 1,j|F g ,m g 〉 . (A.21)<br />

The first term is the “reduced dipole matrix element”, D Fg,Fe . It is independent of<br />

the atomic orientation, i.e. polarisation and sublevel structure. The second term is a<br />

Clebsch-Gordan coefficient, describing the coupling of the sublevels to the spherical<br />

polarisation component j of the light field.<br />

The reduced matrix elements are calculated starting from the matrix element<br />

D 2,3 <strong>for</strong> the closed transition of the rubidium D2 line, which is equivalent to the<br />

reduced matrix element of a two-level atom. This can be expressed using the<br />

Eqs. (A.11), (A.19), the saturation intensity, and the relation I L =(1/2)ε 0 c|E| 2 :<br />

D 2,3 =<br />

√<br />

Γ 3ε 0λ 3 0<br />

8π 2 = 2.53 × 10 −29 Cm. (A.22)<br />

The reduced matrix elements of the other hyperfine transitions are calculated by<br />

{<br />

D Fg,Fe = D 2,3 (−1)<br />

√(2J Fg+Je+I+1 Fg F<br />

e + 1)(2F g +1)<br />

e 1<br />

J e J g I<br />

}<br />

6j<br />

, (A.23)<br />

d Fg,F e<br />

= D F g,F e<br />

D 2,3<br />

. (A.24)<br />

where d Fg,F e<br />

is a dimensionless expression relative to the closed transition. The<br />

Racah “6j” symbol and the Clebsch-Gordan coefficients can be calculated, using<br />

e.g. the Mathematica software package (Wolfram Research).<br />

Light-shift Hamiltonian <strong>for</strong> rubidium.— In or<strong>der</strong> to calculate the light-shift<br />

Hamiltonian <strong>for</strong> a given polarisation ˆɛ, it is useful to define a “reduced light-shift<br />

Hamiltonian” Λ(F g ,F e , ˆɛ) with matrix elements Λ mg,m ′ (F g g,F e , ˆɛ), which result from<br />

the angular part of Eq. (A.21), see e.g. Ref. [79].<br />

We there<strong>for</strong>e define a tensor C(F g ,F e ) with the Clebsch-Gordan coefficients as<br />

elements, C me,mg,j(F g ,F e )=〈F g ,m g , 1,j|F e ,m e 〉. The elements of the polarisability<br />

tensor A(F g ,F e ) are therewith defined as:<br />

A j ′ ,m ′ g ,mg,j (F g ,F e ) = ∑ m e<br />

C me,m g,j(F g ,F e ) C me,m ′ g ,j′(F g,F e ) . (A.25)<br />

They describe the coupling to the light field in terms of an excitation of an atom<br />

from |F g ,m g 〉 to |F e ,m e + j〉 by the component ɛ j of the polarisation ˆɛ, followed by<br />

a (stimulated) deexcitation to |F g ,m ′ g = m g + j − j ′ 〉 by the component ɛ j ′.


124Appendix<br />

The reduced light-shift Hamiltonian is now defined in the spherical polarisation<br />

basis as:<br />

Λ(F g ,F e , ˆɛ) = ˆɛ † ·A(F g ,F e ) · ˆɛ. (A.26)<br />

In low-saturation and <strong>for</strong> large detuning, the light-shift Hamiltonian <strong>for</strong> a rubidium<br />

atom in the ground state F g can be written similarly to the 2-level expression of<br />

Eq. (A.17):<br />

U Fg = Γ2<br />

8<br />

I L<br />

∑ d 2 F g,F e<br />

Λ(F g ,F e , ˆɛ)<br />

. (A.27)<br />

I 0 δ<br />

F Fg,Fe e<br />

This expression has to be calculated <strong>for</strong> both the D1and the D2 line, summing over<br />

F e<br />

(D1) = {1, 2} and F e<br />

(D2) = {0, 1, 2, 3}, respectively. Also the reduced dipole matrix<br />

elements have to be calculated <strong>for</strong> both lines. The total light shift is obtained as<br />

U (tot)<br />

F g<br />

= U (D1)<br />

F g<br />

+ U (D2)<br />

F g<br />

. However, either the detunings δ (D1)<br />

F g,F e<br />

or δ (D2)<br />

F g,F e<br />

are usually<br />

small compared to the splitting of 7.2 THz (or 15 nm) between the D-lines. Thus<br />

calculating the dominant contribution may be sufficient.<br />

A.3.3<br />

Transition matrix elements <strong>for</strong> 87 Rb<br />

D1<br />

-2<br />

-1<br />

0<br />

+1<br />

+2<br />

-1<br />

0<br />

+1<br />

5<br />

30 20<br />

15<br />

5<br />

10<br />

10<br />

15<br />

5<br />

15<br />

15<br />

15<br />

20<br />

10<br />

5 10<br />

5 5 5 5<br />

5<br />

15<br />

5<br />

20<br />

15<br />

5<br />

15<br />

20<br />

15<br />

15 15<br />

5 30 30<br />

15<br />

30<br />

-1<br />

0 +1<br />

F =1<br />

g<br />

-2<br />

-1<br />

0 +1<br />

F =2<br />

g<br />

+2<br />

Figure A.1: Transition matrix elements of the D1 line: 60 (d (D1)<br />

F g,F e<br />

C me,m g,j(F g ,F e )) 2 .


A.3 Light <strong>for</strong>ces and scattering rate 125<br />

D2<br />

-3<br />

-2<br />

-1<br />

0<br />

+1<br />

+2<br />

+3<br />

0<br />

-2 -1 0 +1 +2<br />

60<br />

-1 0 +1<br />

30 20 10<br />

20<br />

40<br />

10<br />

15<br />

15 10<br />

20 4<br />

1 3<br />

1 5<br />

15 15<br />

3<br />

5 10<br />

25<br />

4<br />

32<br />

25 25 25<br />

20 25<br />

6<br />

20<br />

25 3 3<br />

12 12<br />

24 24<br />

36 32<br />

40<br />

4<br />

20<br />

60<br />

20<br />

15<br />

20<br />

5<br />

15<br />

15 15<br />

5<br />

30<br />

6<br />

-1<br />

0 +1<br />

F =1<br />

g<br />

-2<br />

-1<br />

0 +1<br />

F =2<br />

g<br />

+2<br />

Figure A.2: Transition matrix elements of the D2 line: 60 (d (D2)<br />

F g,F e<br />

C me,m g,j(F g ,F e )) 2 .


126 Appendix<br />

A.4Analysis of absorption images<br />

Due to the vertical symmetry axis of our mirror configuration, a factorised spatial<br />

density of N atoms is assumed,<br />

∫∫∫<br />

ρ(r) = Nρ x (x)ρ y (y)ρ z (z) , N = ρ(r)dr 3 , (A.28)<br />

ρ x (x) =<br />

)<br />

1<br />

√ exp<br />

(− x2<br />

, (A.29)<br />

2πσx 2σx<br />

2<br />

ρ y (y)<br />

similarly with σ y = σ x<br />

χ .<br />

(A.30)<br />

The absorption measurements project the distributions onto the xz-plane, while<br />

integrating in the y-direction (the line-of-sight). The coordinates in the xz-plane<br />

can be defined as r ′ =(x, z). An absorption image, A(r ′ )=I d (r ′ )/I L (r ′ ), is the<br />

ratio of the detected probe laser intensity, I d (r ′ ), and the incident intensity, I L (r ′ ).<br />

For a coordinate r ′ in the detection plane, the absorption law is written as:<br />

dI(r)<br />

dy<br />

= −ρ(r) σ(δ) I(r) , (A.31)<br />

σ(δ) = 3λ2 0<br />

2π<br />

1<br />

1+4 ( )<br />

δ 2<br />

. (A.32)<br />

Γ<br />

Here, σ(δ) is the detuning dependent absorption cross section <strong>for</strong> unity Clebsch-<br />

Gordan coefficients [34]. By integration along the y-direction, the relation between<br />

the absorption image, A(r ′ ), and the atomic density is found in the xz-projection:<br />

D(r ′ ) = − ln A(r ′ ) = Nσ(δ) ρ x (x)ρ z (z) . (A.33)<br />

From the image data, a line sum along x can be <strong>for</strong>med, which leads to the vertical<br />

atomic density, ρ z (z):<br />

L(z) =<br />

∫ +∞<br />

−∞<br />

ln A(r ′ ) dz = Nσ(δ) ρ z (z) . (A.34)<br />

The projected density, D(r ′ ), allows to read out the horizontal Gaussian width, σ x .<br />

Hence, by the Eqs. (A.29) and (A.30), the 3D density from Eq. (A.28) is known.


References<br />

[1] C.S. Adams, M. Sigel, and J. Mlynek, Atom optics, Phys.Rep.240, 143 (1994).<br />

[2] C.N. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg, Atom-Photon Interactions<br />

(Wiley, New York, 1992).<br />

[3] R.J. Cook and R.K. Hill, An electromagnetic mirror <strong>for</strong> neutral atoms,<br />

Opt. Comm. 43, 258 (1982).<br />

[4] V.I. Balykin, V.S. Letokhov, Yu.B. Ovchinnikov, and A.I. Sidorov,<br />

Reflection of an atomic beam from a gradient of an optical field,<br />

Pis’ma Zh. Eksp. Teor. Fiz. 45, 282 (1987) [JETP Lett. 45, 353 (1987)].<br />

[5] M.A. Kasevich, D.S. Weiss, and S. Chu, Normal-incidence reflection of slow atoms<br />

from an optical evanescent wave, Opt. Lett. 15, 607 (1990).<br />

[6] M. Born and E. Wolf, Principles of Optics (Cambridge University Press, 1999).<br />

[7] E. Hecht, Optics (Addison-Wesley, 1987).<br />

[8] T.W. Hänsch and A.L. Schawlow, Cooling of gases by laser radiation,<br />

Opt. Comm. 13, 68 (1975).<br />

[9] D. Wineland and H. Dehmelt, Proposed 10 14 δν < ν laser fluorescence spectroscopy<br />

on Tl + mono-ion oscillator III, Bull. Am. Phys. Soc. 20, 637 (1975).<br />

[10] R. Grimm, M. Weidemüller, and Yu.B. Ovchinnikov, Optical dipole traps <strong>for</strong> neutral<br />

atoms, Adv. At. Mol. Opt. Phys. 42, 95 (2000).<br />

[11] H. Metcalf and P. van <strong>der</strong> Straten, Cooling and trapping of neutral atoms,<br />

Phys. Rep. 244, 203 (1994).<br />

[12] C.S. Adams, and E. Riis, Laser cooling and trapping of neutral atoms,<br />

Prog. Quant. Electr. 21, 1 (1997).<br />

[13] S. Chu, The manipulation of neutral particles, Rev.Mod.Phys.70, 685 (1998).<br />

[14] C.N. Cohen-Tannoudji, Manipulating atoms with photons, ibid., p. 707.<br />

[15] W.D. Phillips, Laser cooling and trapping of neutral atoms, ibid., p. 721.<br />

[16] D.V. Laryushin, Yu.B. Ovchinnikov, V.I. Balykin, and V.S. Letokhov,<br />

Reflection cooling of sodium atoms in an evanescent light wave,<br />

Opt. Comm. 135, 138 (1997). [Pis’ma Zh. Éksp. Teor. Fiz. 62, 102 (1995).]<br />

[17] Yu.B. Ovchinnikov, I. Manek, and R. Grimm, Surface trap <strong>for</strong> Cs atoms based on<br />

evanescent-wave cooling, Phys. Rev. Lett. 79, 2225 (1997).<br />

[18] P. Berman, ed., Atom interferometry (Academic Press, 1997).<br />

[19] M.J. Snadden, J.M. McGuirk, P. Bouyer, K.G. Haritos, and M.A. Kasevich,<br />

Measurement of the earth’s gravity gradient with an atom interferometer-based gravity<br />

gradiometer, Phys. Rev. Lett. 81, 971 (1998).<br />

[20] A. Peters, K.Y. Chung and S. Chu, Measurement of gravitational acceleration by<br />

dropping atoms, Nature400, 849 (1999).<br />

[21] D.S. Weiss, B.C. Young, and S. Chu, Precision measurement of the photon recoil of<br />

an atom using atom interferometry, Phys. Rev. Lett. 70, 2706 (1993).<br />

[22] R. Blatt, J. Eschner, D. Leibfried, and F. Schmidt-Kaler, eds., Laser spectroscopy<br />

XIV (World Scientific, Singapore, 1999).<br />

127


128 References<br />

[23] J.T. Bahns, P.L. Gould, and W.C. Stwalley, Formation of cold (T ≤ 1 K) molecules,<br />

Ad. At. Mol. Opt. Phys. 42, 171 (2000).<br />

[24] R. Wynar, R.S. Freeland, D.J. Han, C. Ryu, and D.J. Heinzen, Molecules in a Bose-<br />

Einstein condensate, Science 287, 1016 (2000).<br />

[25] P.R. Berman, ed., Cavity quantum electrodynamics, supplement to Adv. At. Mol.<br />

Opt. Phys. (Academic Press, 1994).<br />

[26] J. Weiner, V.S. Bagnato, S. Zilio, and P.S. Julienne, Experiments and theory in cold<br />

and ultracold collisions, Rev.Mod.Phys.71, 1 (1999).<br />

[27] Q. Niu, X.-G. Zhao, G.A. Georgakis, and M.G. Raizen, Atomic Landau-Zener tunnelling<br />

and Wannier-Stark lad<strong>der</strong>s in optical potentials, Phys. Rev. Lett. 76, 4504 (1996).<br />

[28] M.B. Dahan, E. Peik, J. Reichel, Y. Castin, and C. Salomon, Bloch oscillations of<br />

atoms in an optical potential, Phys. Rev. Lett. 76, 4508 (1996).<br />

[29] P. Laurent, P. Lemonde, G. Santarelli, M. Abgrall, J. Kitching, Y. Sortais, S. Bize,<br />

M. Santos, C. Nicolas, S. Zhang, G. Schehr, A. Clairon, A. Mann, A. Luiten,<br />

S. Chang, and C. Salomon, <strong>Cold</strong> atom clocks on earth and in space, in Ref. [22].<br />

[30] G. Timp, R.E. Behringer, D.M. Tennant, J.E. Cunningham, M. Prentiss, and<br />

K.K. Berggren, Using light as a lens <strong>for</strong> submicron, neutral-atom lithography,<br />

Phys. Rev. Lett. 69, 1636 (1992).<br />

[31] B. Brezger, T. Schulze, P.O. Schmidt, R. Mertens, T. Pfau, and J. Mlynek, Polarisation<br />

gradient light masks in atom lithography, Europhys. Lett. 46, 148 (1999).<br />

[32] C.H. Bennett and D. DiVincenzo, Quantum in<strong>for</strong>mation and computation,<br />

Nature 404, 247 (2000).<br />

[33] E.L. Raab, M. Prentiss, A. Cable, S. Chu, and D.E. Pritchard, Trapping of neutral<br />

sodium atoms with radiation pressure, Phys. Rev. Lett. 59, 2631 (1987).<br />

[34] A.E. Siegman, Lasers, (University Science Books, Mill Valley, CA, 1986).<br />

[35] T.H. Maiman, Stimulated optical radiation in ruby, Nature187, 493 (1960).<br />

[36] S. Bose, Plancks Gesetz und Lichtquantenhypothese, Z.Phys.26, 178 (1924).<br />

[37] A. Einstein, Sitzber. Kgl. Preuss. Akad. Wiss. 261 (1924).<br />

[38] J.T.M. Walraven, Atomic hydrogen in magnetostatic traps, in Proceedings of the<br />

SUSSP 44 conference on Quantum dynamics of simple systems, ed. by G.-L. Oppo,<br />

S.M. Barnett, E. Riis, and M. Wilkinson (IOP, Bristol, 1996).<br />

[39] M.H. An<strong>der</strong>son, J.R. Ensher, M.R. Matthews, C.E. Wieman, and E.A. Cornell,<br />

Observation of Bose-Einstein condensation in a dilute atomic vapour,<br />

Science 269, 198 (1995).<br />

[40] K.B. Davis, M.-O. Mewes, M.R. Andrews, N.J. van Druten, D.S. Durfee,<br />

D.M. Kurn, and W. Ketterle, Bose-Einstein condensation in a gas of sodium atoms,<br />

Phys. Rev. Lett. 75, 3969 (1995).<br />

[41] C.C. Bradley, C.A. Sackett, and R.G. Hulet, Bose-Einstein condensation of lithium:<br />

observation of limited condensate number, Phys. Rev. Lett. 78, 985 (1997). [See also<br />

C.C. Bradley et al., Phys. Rev. Lett. 75, 1687 (1995).]<br />

[42] D.G. Fried, T.C. Killian, L. Willmann, D. Landhuis, S.C. Moss, D. Kleppner,<br />

and Thomas J. Greytak, Bose-Einstein condensation of atomic hydrogen,<br />

Phys. Rev. Lett. 81, 3811 (1998).<br />

[43] A.I. Safonov, S.A. Vasilyev, I.S. Yasnikov, I.I. Lukashevich, and S. Jaakkola,<br />

Observation of quasicondensate in two-dimensional atomic hydrogen,<br />

Phys. Rev. Lett. 81, 4545 (1998).


References 129<br />

[44] B. DeMarco and D.S. Jin, Onset of Fermi degeneracy in a trapped atomic gas,<br />

Science 285, 1703 (1999).<br />

[45] A. Griffin, D. Snoke, and S. Stringari, eds., Bose-Einstein condensation (Cambridge<br />

University Press, 1995).<br />

[46] F. Dalfovo, S. Giorgini, L.P. Pitaevskii, and S. Stringari, Theory of Bose-Einstein<br />

condensation in trapped gases, Rev.Mod.Phys.71, 463 (1999).<br />

[47] M. Inguscio, S. Stringari, and C. Wieman, eds., Bose-Einstein condensation in atomic<br />

gases, Proc. of the Int. School of Physics “Enrico Fermi” (IOS, Amsterdam, 1999).<br />

[48] A.S. Parkins and D.F. Walls, The physics of trapped dilute-gas Bose-Einstein<br />

condensates, Phys.Rep.303, 1 (1998).<br />

[49] M.-O. Mewes, M.R. Andrews, D.M. Kurn, D.S. Durfee, C.G. Townsend, and<br />

W. Ketterle, Output coupler <strong>for</strong> Bose-Einstein condensed atoms, Phys. Rev. Lett. 78,<br />

582 (1997).<br />

[50] B.P. An<strong>der</strong>son and M.A. Kasevich, Macroscopic quantum interference from atomic<br />

tunnel arrays, Science 282, 1686 (1998).<br />

[51] I. Bloch, T.W. Hänsch, and T. Esslinger, Atom laser with a cw output coupler,<br />

Phys. Rev. Lett. 82, 3008 (1999).<br />

[52] E.W. Hagley, L. Deng, M. Kozuma, J. Wen, K. Helmerson, S.L. Rolston, and W.D.<br />

Phillips, A well-collimated quasi-continuous atom laser, Science 283, 1706 (1999).<br />

[53] H.F. Hess, Evaporative cooling of magnetically trapped and compressed spin-polarised<br />

hydrogen, Phys.Rev.B34, 3476 (1986).<br />

[54] O.J. Luiten, M.W. Reynolds, and J.T.M. Walraven, Kinetic theory of the evaporative<br />

cooling of a trapped gas, Phys.Rev.A53, 381 (1996).<br />

[55] W. Ketterle and N.J. van Druten, Evaporative cooling of trapped atoms,<br />

Adv.At.Mol.Opt.Phys.37, 181 (1996).<br />

[56] S. Bhongale and M. Holland, Loading a continuous-wave atom laser by optical pumping<br />

techniques, Phys.Rev.A62, 043604 (2000).<br />

[57] E. Mandonnet, A. Minguzzi, R. Dum, I. Carusotto, Y. Castin, and J. Dalibard,<br />

Evaporative cooling of an atomic beam, Eur.J.Phys.D10, 18 (2000).<br />

[58] H. Wiseman and M. Collett, An atom laser based on dark-state cooling,<br />

Phys. Lett. A 202, 246 (1995).<br />

[59] R.J.C. Spreeuw, T. Pfau, U. Janicke, and M. Wilkens, Laser-like scheme <strong>for</strong> atomic<br />

matter waves, Europhys. Lett. 32, 469 (1995).<br />

[60] M. Olshanii, Y. Castin, and J. Dalibard, A model <strong>for</strong> an atom laser, inLaser Spectroscopy<br />

XII, eds. M. Inguscio, M. Allegrini, and A. Sasso (World Scientific, 1996).<br />

[61] S. Inouye, T. Pfau, S. Gupta, A.P. Chikkatur, A. Görlitz, D.E. Pritchard, and W. Ketterle,<br />

Phase-coherent amplification of atomic matter waves, Nature402, 641 (1999).<br />

[62] J. Dalibard, Collisional dynamics of ultra-cold atomic gases, in Ref. [47].<br />

[63] D.J. Heinzen, Ultracold atomic interactions, in Ref. [47].<br />

[64] Y. Castin, J.I. Cirac, and M. Lewenstein, Reabsorption of light by trapped atoms,<br />

Phys. Rev. Lett. 80, 5305 (1998).<br />

[65] U. Janicke and M. Wilkens, Prospects of matter wave amplification in the presence<br />

of a single photon, Europhys. Lett. 35, 561 (1996).<br />

[66] J.I. Cirac, M. Lewenstein, and P. Zoller, Collective laser cooling of trapped atoms,<br />

Europhys. Lett. 35, 647 (1996).


130 References<br />

[67] D.W. Sesko, T.G. Walker, and C.E. Wieman, Behaviour of neutral atoms in a spontaneous<br />

<strong>for</strong>ce trap, J.Opt.Soc.Am.B8, 946 (1991).<br />

[68] T.W. Hijmans, G.V. Shlyapnikov, and A.L. Burin, Influence of radiative interatomic<br />

collisions on dark-state cooling, Phys.Rev.A54, 4332 (1996).<br />

[69] M.H. An<strong>der</strong>son, W. Petrich, J.R. Ensher, and E. Cornell, Reduction of lightassisted<br />

collisional loss rate from a low-pressure vapour-cell trap, Phys.Rev.A50,<br />

R3597 (1994).<br />

[70] J. Dalibard and C. Cohen-Tannoudji, Laser cooling below the Doppler limit by polarisation<br />

gradients: simple theoretical models, inLaser cooling and trapping, ed.by<br />

S. Chu and C. Wieman, J. Opt. Soc. Am. B 6 special issue (1989), p. 2023.<br />

[71] P.D. Lett, W.D. Phillips, S.L. Rolston, C.E. Tanner, R.N. Watts, and C.I. Westbrook,<br />

Optical molasses, ibid., p. 2084.<br />

[72] W. Ketterle, K.B. Davis, M.A. Joffe, A. Martin, and D.E. Pritchard, High densities of<br />

cold atoms in a dark spontaneous-<strong>for</strong>ce optical trap, Phys. Rev. Lett. 70, 2253 (1993).<br />

[73] M. Drewsen, Ph. Laurent, A. Nadir, G. Santarelli, A. Clairon, Y. Castin, D. Grison,<br />

and C. Salomon, Investigation of sub-Doppler cooling effects in a cesium magnetooptical<br />

trap, Appl.Phys.B59, 283 (1994).<br />

[74] C.G. Townsend, N.H. Edwards, C.J. Cooper, K.P. Zetie, C.J. Foot, A.M. Steane,<br />

P. Szriftgiser, H. Perrin, and J. Dalibard, Phase-space density in the magneto-optical<br />

trap, Phys.Rev.A52, 1423 (1995).<br />

[75] S.E. Hamann, D.L. Haycock, G. Klose, P.H. Pax, I.H. Deutsch, and P.S. Jessen,<br />

Resolved-sideband Raman cooling to the ground state of an optical lattice,<br />

Phys. Rev. Lett. 80, 4149 (1998).<br />

[76] I. Bouchoule, H. Perrin, A. Kuhn, M. Morinaga, and C. Salomon, Neutral atoms<br />

prepared in Fock states of a one-dimensional harmonic potential, Phys.Rev.A59,<br />

R8 (1999).<br />

[77] A.J. Kerman, V. Vuletić, C. Chin, and S. Chu, Beyond optical molasses: 3DRaman<br />

sideband cooling of atomic cesium to high phase-space density, Phys. Rev. Lett. 84,<br />

439 (2000).<br />

[78] D.-J. Han, S. Wolf, S. Oliver, C. McCormick, M.T. DePue, and D.S. Weiss,<br />

3DRaman sideband cooling of cesium atoms at high density, Phys. Rev. Lett.<br />

85, 724 (2000).<br />

[79] P.S. Jessen and I.H. Deutsch, Optical lattices, Adv. At. Mol. Opt. Phys. 37, 95 (1996).<br />

[80] L. Guidoni and P. Verkerk, Optical lattices: cold atoms or<strong>der</strong>ed by light (Ph.D.<br />

tutorial), J.Opt.B.: Quant.Semiclass.Opt.1, R23 (1999).<br />

[81] T. Ido, Y. Isoya, and H. Katori, Optical-dipole trapping of Sr atoms at a high phasespace<br />

density, Phys.Rev.A61, 061403(R) (2000).<br />

[82] P. Desbiolles and J. Dalibard, Loading atoms in a bi-dimensional light trap,<br />

Opt. Comm. 132, 540 (1996).<br />

[83] W.L. Power, T. Pfau, and M. Wilkens, Loading atoms into a surface trap: simulations<br />

of an experimental scheme, Opt. Comm. 143, 125 (1997).<br />

[84] T. Pfau and J. Mlynek, A 2Dquantum gas of laser cooled atoms, OSA Trends Optics<br />

Photonics Series 7, 33 (1997).<br />

[85] E.A. Hinds, M.G. Boshier, and I.G. Hughes, Magnetic waveguide <strong>for</strong> trapping cold<br />

atom gases in two dimensions, Phys. Rev. Lett. 80, 645 (1998).


References 131<br />

[86] H. Gauck, M. Hartl, D. Schneble, H. Schnitzler, T. Pfau, and J. Mlynek, Quasi-2Dgas<br />

of laser cooled atoms in a planar matter waveguide, Phys. Rev. Lett. 81, 5298 (1998).<br />

[87] R.J.C. Spreeuw, D. Voigt, B.T. Wolschrijn, and H.B. van Linden van den Heuvell,<br />

Creating a low-dimensional quantum gas using dark states in an inelastic evanescentwave<br />

mirror, Phys.Rev.A61, 053604 (2000).<br />

[88] V. Bagnato, D. Kleppner, Bose-Einstein condensation in low-dimensional traps,<br />

Phys. Rev. A 44, 7439 (1991).<br />

[89] D.S. Petrov, M. Holzmann, and G.V. Shlyapnikov, Bose-Einstein condensation in<br />

quasi 2Dtrapped gases, Phys. Rev. Lett. 84 2551 (2000).<br />

[90] J.M. Kosterlitz and D.J. Thouless, Or<strong>der</strong>ing, metastability and phase transitions in<br />

two-dimensional systems, J.Phys.C6, 1181 (1973).<br />

[91] D. Müller, D.Z. An<strong>der</strong>son, R.J. Grow, P.D.D. Schwindt, and E.A. Cornell, Guiding<br />

neutral atoms around curves with lithographically patterned current-carrying wires,<br />

Phys. Rev. Lett. 83, 5194 (1999).<br />

[92] N.H. Dekker, C.S. Lee, V. Lorent, J.H. Thywissen, S.P. Smith, M. Drndić,<br />

R.M. Westervelt, and M. Prentiss, Guiding neutral atoms on a chip, Phys. Rev. Lett.<br />

84, 1124 (2000).<br />

[93] J. Reichel, W. Hänsel, and T.W. Hänsch, Atomic micromanipulation with magnetic<br />

surface traps, Phys. Rev. Lett. 83, 3398 (1999).<br />

[94] R. Folman, P. Krüger, D. Cassettari, B. Hessmo, T. Maier, and J. Schmiedmayer,<br />

Controlling cold atoms using nanofabricated surfaces: atom chips, Phys. Rev. Lett.<br />

84, 4749 (2000).<br />

[95] A. Browaeys, J. Poupard, A. Robert, S. Nowak, W. Rooijakkers, E. Arimondo,<br />

L. Marcassa, D. Boiron, C.I. Westbrook, and A. Aspect, Two body loss rate in a<br />

magneto-optical trap of metastable He, Eur.Phys.J.D8, 199 (2000).<br />

[96] N. Herschbach, P.J.J. Tol, W. Hogervorst, and W. Vassen, Suppression of Penning<br />

ionisation by spin polarisation of cold He 2 3 S atoms, Phys.Rev.A61, 050702(R)<br />

(2000).<br />

[97] H. Wiseman, A. Martins, and D. Walls, An atom laser based on evaporative cooling,<br />

Quant. Semiclass. Opt. 8, 737 (1996).<br />

[98] A.M. Guzmán, M. Moore, and P. Meystre, Theory of a coherent atomic-beam<br />

generator, Phys.Rev.A53, 977 (1996).<br />

[99] M. Holland, K. Burnett, C. Gardiner, J.I. Cirac, and P. Zoller, Theory of an atom<br />

laser, Phys.Rev.A54, R1757 (1996).<br />

[100] J.-Y. Courtois, J.-M. Courty, and J.C. Mertz, Internal dynamics of multilevel atoms<br />

near a vacuum-dielectric interface, Phys.Rev.A53, 1862 (1996).<br />

[101] C. Henkel and J.-Y. Courtois, Recoil and momentum diffusion of an atom close to<br />

a vacuum-dielectric interface, Eur.Phys.J.D3, 129 (1998).<br />

[102] L. Spruch, Retarded, or Casimir, long-range potentials, Physics Today, November<br />

1986, p. 37.<br />

[103] A. Landragin, J.-Y. Courtois, G. Labeyrie, N. Vansteenkiste, C.I. Westbrook,<br />

and A. Aspect, Measurement of the Van <strong>der</strong> Waals <strong>for</strong>ce in an atomic mirror,<br />

Phys. Rev. Lett. 77, 1464 (1996).<br />

[104] J.D. Miller, R.A. Cline, D.J. Heinzen, Far-off-resonance optical trapping of atoms,<br />

Phys. Rev. A 47, R4567 (1993).


132 References<br />

[105] T. Müller-Seydlitz, M. Hartl, B. Brezger, H. Hänsel, C. Keller, A. Schnetz,<br />

R.J.C. Spreeuw, T. Pfau, and J. Mlynek, <strong>Atoms</strong> in the lowest motional band of a<br />

three-dimensional optical lattice, Phys. Rev. Lett. 78, 1038 (1997).<br />

[106] P. Desbiolles, M. Arndt, P. Szriftgiser, and J. Dalibard, Elementary Sisyphus process<br />

close to a dielectric, Phys.Rev.A54, 4292 (1996).<br />

[107] A. Aspect, E. Arimondo, R. Kaiser, N. Vansteenkiste, and C. Cohen-Tannoudji,<br />

Laser cooling below the one-photon recoil energy by velocity-selective coherent population<br />

trapping, Phys. Rev. Lett. 61, 826 (1988).<br />

[108] J. Lawall, S. Kulin, B. Saubamea, N. Bigelow, M. Leduc, and C. Cohen-Tannoudji,<br />

Three-dimensional laser cooling of helium beyond the single-photon recoil limit,<br />

Phys. Rev. Lett. 75, 4195 (1995).<br />

[109] P.S. Jessen, C. Gerz, P.D. Lett, W.D. Phillips, S.L. Rolston, R.J.C. Spreeuw,<br />

and C.I. Westbrook, Observation of quantized motion of Rb in an optical field,<br />

Phys. Rev. Lett. 69, 49 (1992).<br />

[110] P. Verkerk, B. Lounis, C. Salomon, C. Cohen-Tannoudji, J.-Y. Courtois, and<br />

G. Grynberg, Dynamics and spatial or<strong>der</strong> of cold cesium atoms in a periodic optical<br />

potential, Phys. Rev. Lett. 68, 3861 (1992).<br />

[111] D. Boiron, A. Michaud, J.M. Fournier, L. Simard, M. Sprenger, G. Grynberg,<br />

and C. Salomon, <strong>Cold</strong> and dense cesium clouds in far-detuned dipole traps,<br />

Phys. Rev. A 57, R4106 (1998).<br />

[112] S. Friebel, R. Scheunemann, J. Walz, T.W. Hänsch, and M. Weitz, Laser cooling in<br />

aCO 2 -laser optical lattice, Appl.Phys.B67, 699 (1998).<br />

[113] N. Vansteenkiste, P. Vignolo, and A. Aspect, Optical reversibility theorems <strong>for</strong><br />

polarisation; application to remote control of polarisation, J.Opt.Soc.Am.A10,<br />

2240 (1993).<br />

[114] W. Seifert, R. Kaiser, A. Aspect, and J. Mlynek, Reflection of atoms from a dielectric<br />

wave guide, Opt. Comm. 111, 566 (1994).<br />

[115] Z.T. Lu, K.L. Corwin, M.J. Renn, M.H. An<strong>der</strong>son, E.A. Cornell, and C.E. Wieman,<br />

Low-velocity intense source of atoms from a magneto-optical trap, Phys. Rev. Lett.<br />

77, 3331 (1996).<br />

[116] K. Dieckmann, R.J.C. Spreeuw, M. Weidemüller, and J.T.M. Walraven, Two-dimensional<br />

magneto-optical trap as a source of slow atoms, Phys.Rev.A58, 3891 (1998).<br />

[117] C.E. Wieman and L. Hollberg, Using diode lasers <strong>for</strong> atomic physics,<br />

Rev. Sci. Instr. 62, 1 (1991).<br />

[118] K.B. MacAdam, A. Steinbach, and C. Wieman, A narrow-band tunable diode laser<br />

system with grating feedback and a saturated absorption spectrometer <strong>for</strong> Cs and Rb,<br />

Am.J.Phys.60, 1098 (1992).<br />

[119] G.C. Bjorklund, M.D. Levenson, W. Lenth, and C. Ortiz, Frequency modulation<br />

(FM) spectroscopy, Appl.Phys.B32, 145 (1983).<br />

[120] B. Cheron, H. Gilles, J. Havel, O. Moreau, and H. Sorel, Laser frequency stabilization<br />

using Zeeman effect, J. Phys. III, 4, 401 (1994).<br />

[121] K.L. Corwin, Z.-T. Lu, C.F. Hand, R.J. Epstein, and C.E. Wieman, Frequencystabilised<br />

diode laser with the Zeeman shift in an atomic vapour, Appl.Opt.37, 3295<br />

(1998).<br />

[122] M. Gertsvolf and M. Rosenbluh, Injection locking of a diode laser locked to a Zeeman<br />

frequency stabilised laser oscillator, Opt. Comm. 170, 269 (1999).


References 133<br />

[123] C.S. Adams, J. Vorberg, and J. Mlynek, Single-frequency operation of a diodepumped<br />

lanthanum-neodymium-hexaaluminate laser by using a twisted-mode cavity,<br />

Opt. Lett. 18, 420 (1993).<br />

[124] C. Monroe, W. Swann, H. Robinson, and C. Wieman, Very cold trapped atoms in a<br />

vapour cell, Phys. Rev. Lett. 65, 1571 (1990).<br />

[125] K. Lindquist, M. Stephens, and C. Wieman, Experimental and theoretical study of<br />

the vapour-cell Zeeman optical trap, Phys.Rev.A46, 4082 (1992).<br />

[126] H. Ludvigsen, A. Äijälä, A. Pietiläinen, H. Talvitie, and E. Ikonen, Laser cooling of<br />

rubidium atoms in a vapour cell, Physica Scripta 49, 424 (1994).<br />

[127] CRC Handbook of Chemistry and Physics, 73 rd edition, (CRC Press, 1992).<br />

[128] C.J. Myatt, N.R. Newbury, R.W. Ghrist, S. Loutzenhiser, and C.E. Wieman,<br />

Multiply loaded magneto-optical trap, Opt. Lett. 21, 290 (1996).<br />

[129] A. Noble and M. Kasevich, UHV optical window seal to Conflat TM knife edge,<br />

Rev. Sci. Instr. 65, 3042 (1994).<br />

[130] Le Carbone-Lorraine, Cefilac Étanchéïté, catalogues Hélicoflex high per<strong>for</strong>mance<br />

sealing and Seals <strong>for</strong> ultra-high vacuum and cryogenics (Saint-Etienne, France).<br />

[131] K. Dieckmann et al., AMOLF Institute, Amsterdam, private communication.<br />

[132] C. Wieman, G. Flowers, and S. Gilbert, Inexpensive laser cooling and trapping experiment<br />

<strong>for</strong> un<strong>der</strong>graduate laboratories, Am.J.Phys.63, 317 (1995).<br />

[133] W. Demtrö<strong>der</strong>, Laser spectroscopy (Springer, Heidelberg, 1996).<br />

[134] M. Mürtz, J.S. Wells, L. Hollberg, T. Zibrova, and N. Mackie, Extended-cavity<br />

grating-tuned operation of mid-infrared InAsSb diode lasers, Appl.Phys.B66, 277<br />

(1998).<br />

[135] L. Ricci, M. Weidemüller, T. Esslinger, A. Hemmerich, C. Zimmermann, V. Vuletic,<br />

W. König, and T.W. Hänsch, A compact grating-stabilised diode laser system <strong>for</strong><br />

atomic physics, Opt. Comm. 117, 541 (1995).<br />

[136] T. Hof, D. Fick, H.J. Jänsch, Application of diode lasers as a spectroscopic tool at<br />

670 nm, Opt. Comm. 124, 283 (1996).<br />

[137] H. Leinen, D. Gläßner, H. Metcalf, R. Wynands, D. Haubrich, and D. Meschede,<br />

GaN blue diode lasers: a spectroscopist’s view, Appl.Phys.B70, 567 (2000).<br />

[138] G.R. Hadley, Injection locking of diode lasers, IEEE J. Quant. Electr. 22, 419 (1986).<br />

[139] P. Spano, S. Piazzolla, and M. Tamburrini, Frequency and intensity noise in<br />

injection-locked semiconductor lasers: theory and experiment, IEEE J. Quant. Electr.<br />

22, 427 (1986).<br />

[140] S. Wolf, Entwicklung und Charakterisierung eines frequenzstabilen Diodenlasers,<br />

diploma thesis, Universität Konstanz (1994, unpublished).<br />

[141] A.S. Arnold, J.S. Wilson, and M.G. Boshier, A simple extended-cavity diode laser,<br />

Rev. Sci. Instr. 69, 1236 (1998).<br />

[142] P. Jessen, private communication.<br />

[143] G.P. Barwood, P. Gill, and W.R.C. Rowley, Frequency measurements on optically<br />

narrowed Rb-stabilised laser diodes at 780 nm and 795 nm, Appl.Phys.B53, 142<br />

(1991).<br />

[144] U. Volz and H. Schmoranzer, Precision lifetime measurement on alkali atoms and<br />

on helium by beam-gas-laser spectroscopy, Physica Scripta T65, 48 (1996).<br />

[145] T.W. Hänsch and B. Couillaud, Laser frequency stabilisation by polarisation spectroscopy<br />

of a reflecting reference cavity, Opt. Comm. 35, 441 (1980).


134References<br />

[146] W.D. Lee and J.C. Campbell, Narrow-linewidth frequency stabilized<br />

Al x Ga 1−x As/GaAs laser, Appl.Phys.B60, 1544 (1992).<br />

[147] Z.D. Liu, D. Bloch, and M. Ducloy, Absolute active frequency locking of a diode laser<br />

with optical feedback generated by Doppler-free collinear polarisation spectroscopy,<br />

Appl. Phys. B 65, 274 (1994).<br />

[148] S.J.H. Petra, Development of frequency stabilised laser diodes <strong>for</strong> building a magnetooptical<br />

trap, diploma thesis, Universiteit van Amsterdam (1998, unpublished).<br />

[149] M.R. Andrews, M.-O. Mewes, N.J. van Druten, D.S. Durfee, D.M. Kurn, and<br />

W. Ketterle, Direct, non-destructive observation of a Bose condensate, Science 273,<br />

84 (1996).<br />

[150] C.C. Bradley, C.A. Sackett, and R.G. Hulet, Analysis of in situ images of Bose-<br />

Einstein condensates of lithium, Phys.Rev.A55, 3951 (1997).<br />

[151] R.J.C. Spreeuw, Absorption imaging of cold 2Datoms, unpublished.<br />

[152] R.J. Cook, Theory of resonance-radiation pressure, Phys.Rev.A22, 1078 (1980).<br />

[153] J.P. Gordon and A. Ashkin, Motion of atoms in a radiation trap,<br />

Phys. Rev. A 21,1606 (1980).<br />

[154] A.M. Steane, M. Chowdhury, and C.J. Foot, Radiation <strong>for</strong>ce in the magneto-optical<br />

trap, J. Opt. Soc. Am. B 9, 2142 (1992).<br />

[155] S. Chu, L. Hollberg, J.E. Bjorkholm, A. Cable, and A. Ashkin, Three-dimensional<br />

viscous confinement and cooling of atoms by resonance radiation pressure,<br />

Phys. Rev. Lett. 55, 48 (1985).<br />

[156] P.D. Lett, R.N. Watts, C.I. Westbrook, and W.D. Phillips, Observation of atoms<br />

laser cooled below the Doppler limit, Phys. Rev. Lett. 61, 169 (1988).<br />

[157] D.R. Meacher, D. Boiron, H. Metcalf, C. Salomon, and G. Grynberg, Method <strong>for</strong><br />

velocimetry of cold atoms, Phys.Rev.A50, R1992 (1994).<br />

[158] E.C. Schil<strong>der</strong>, An atomic trampoline, diploma thesis, Universiteit van Amsterdam<br />

(1999, unpublished).<br />

[159] R. Wynands and A. Nagel, Precision spectroscopy with coherent dark states,<br />

Appl. Phys. B 68, 1 (1999).<br />

[160] W. Hanle, Über magnetische Beeinflussung <strong>der</strong> Polarisation <strong>der</strong> Resonanzfluoreszenz,<br />

Z.Phys.30, 93 (1924).<br />

[161] G.L. Abbas, S. Yang, V.W.S. Chan, and J.G. Fujimoto, Injection behaviour and<br />

modelling of 100 mW broad area diode lasers, IEEE J. Quant. Electr. 24, 609 (1988).<br />

[162] L. Goldberg and M.K. Chun, Injection locking characteristics of a 1 W broad stripe<br />

laser diode, Appl. Phys. Lett. 53, 1900 (1988).<br />

[163] L. Goldberg, D. Mehuys, M.R. Surette, and D.C. Hall, High-power, near-diffractionlimited<br />

large-area travelling-wave semiconductor amplifiers, IEEE J. Quant. Electr.<br />

29, 2028 (1993).<br />

[164] A.C. Fey-den Boer, H.C.W. Beijerinck, and K.A.H. van Leeuwen, High-power broadarea<br />

diode lasers <strong>for</strong> laser cooling, Appl.Phys.B64, 415 (1997).<br />

[165] E. Gehrig, B. Beier, K.-J. Boller, and R. Wallenstein, Experimental characterisation<br />

and numerical modelling of an AlGaAs oscillator broad area double pass amplifier<br />

system, Appl.Phys.B66, 287 (1998).<br />

[166] M. Praeger, V. Vuletić, T. Fischer, T.W. Hänsch, and C. Zimmermann, Abroad<br />

emitter diode laser system <strong>for</strong> lithium spectroscopy, Appl.Phys.B67, 163 (1998).


References 135<br />

[167] I. Shvarchuck, K. Dieckmann, M. Zielonkowski, and J.T.M. Walraven,<br />

Broad-area diode laser system <strong>for</strong> a rubidium Bose-Einstein condensation experiment,<br />

Appl. Phys. B 71, 475 (2000).<br />

[168] K. Iida, H. Horiuchi, O. Matoba, T. Omatsu, T. Shimura, and K. Kuroda,<br />

Injection locking of broad-area diode laser through a double phase-conjugate mirror,<br />

Opt. Comm. 146, 6 (1998).<br />

[169] J.N. Walpole, Semiconductor amplifiers and lasers with tapered gain regions,<br />

Opt. Quant. Electr. 28, 623 (1996).<br />

[170] J.N. Walpole, E.S. Kintzer, S.R. Chinn, C.A. Wang, and L.J. Missaggia, High-power<br />

strained-layer InGaAs/AlGaAs tapered travelling wave amplifier, Appl. Phys. Lett.<br />

61, 740 (1992).<br />

[171] D. Mehuys, D.F. Welch, and L. Goldberg, 2.0 W cw, diffraction-limited tapered<br />

amplifier with diode injection, Electr. Lett. 28, 1944 (1992).<br />

[172] C. Zimmermann, J. Walz, and T.W. Hänsch, Solid-state laser source <strong>for</strong><br />

(anti)hydrogen spectroscopy, Hyperfine Interactions 100, 145 (1996).<br />

[173] D. Wandt, M. Laschek, F. v. Alvensleben, A. Tünnermann, and H. Welling, Continuously<br />

tunable 0.5 W single-frequency diode laser source, Opt. Comm. 148, 261<br />

(1998).<br />

[174] G. Ferrari, M.-O. Mewes, F. Schreck, and C. Salomon, High-power multiplefrequency<br />

narrow-linewidth laser source based on a semiconductor tapered amplifier,<br />

Opt. Lett. 24, 151 (1999).<br />

[175] J.C. Livas, S.R. Chinn, E.S. Kintzer, J.N. Walpole, C.A. Wang, and L.J. Missaggia,<br />

Single-mode optical fibre coupling of high-power tapered gain region devices,<br />

IEEE Phot. Techn. Lett. 6, 422 (1994).<br />

[176] Z.L. Liau, J.N. Walpole, D.E. Mull, C.L. Dennis, and L.J. Missaggia,<br />

Accurate fabrication of anamorphic microlenses and efficient collimation of tapered<br />

unstable-resonator diode lasers, Appl.Phys.Lett.64, 3368 (1994).<br />

[177] Z.L. Liau, J.N. Walpole, J.C. Livas, E.S. Kintzer, D.E. Mull, L.J. Missaggia, and<br />

W.F. DiNatale, Fabrication of two-sided anamorphic microlenses and direct coupling<br />

of tapered high-power diode laser to single-mode fibre, IEEE Phot. Techn. Lett. 7,<br />

1315 (1995).<br />

[178] Spectra Diode Laboratories, Product Catalog (San Jose, Ca., 1996).<br />

[179] W.W. Chow and R.R. Craig, Amplified spontaneous emission effects in semiconductor<br />

laser amplifiers, IEEE. J. Quant. Electr. 26, 1363 (1990).<br />

[180] E.S. Kintzer, J.N. Walpole, S.R. Chinn, C.A. Wang, and L.J. Missaggia,<br />

High-power, strained-layer amplifiers and lasers with tapered gain regions,<br />

IEEE Phot. Techn. Lett. 5, 605 (1993).<br />

[181] L. Cognet, V. Savalli, G.Zs.K. Horvath, D. Holleville, R. Marani, N. Westbrook,<br />

C.I. Westbrook, and A. Aspect, Atomic interference in grazing incidence diffraction<br />

from an evanescent wave mirror, Phys. Rev. Lett. 81, 5044 (1998).<br />

[182] R. Marani, L. Cognet, V. Savalli, N. Westbrook, C.I. Westbrook, and A. Aspect,<br />

Using atomic interference to probe atom-surface interactions, Phys. Rev. A 61,<br />

053402 (2000).<br />

[183] M. Gorlicki, S. Feron, V. Lorent, and M. Ducloy, Interferometric approaches to<br />

atom-surface Van <strong>der</strong> Waals interactions in atomic mirrors, Phys.Rev.A61, 013603<br />

(2000).


136 References<br />

[184] H.B.G. Casimir and D. Pol<strong>der</strong>, The influence of retardation on the London-Van <strong>der</strong><br />

Waals <strong>for</strong>ces, Phys.Rev.73, 360 (1948).<br />

[185] V. Sandoghdar, C.I. Sukenik, E.A. Hinds, and S. Haroche, Direct measurement of<br />

the Van <strong>der</strong> Waals interaction between an atom and its images in a micron-sized<br />

cavity, Phys. Rev. Lett. 68, 3432 (1992).<br />

[186] C.I. Sukenik, M.G. Boshier, D. Cho, V. Sandoghdar, and E.A. Hinds, Measurement<br />

of the Casimir-Pol<strong>der</strong> <strong>for</strong>ce, Phys. Rev. Lett. 70, 560 (1993).<br />

[187] S.K. Lamoreaux, Demonstration of the Casimir <strong>for</strong>ce in the 0.6 to 6 µm range,<br />

Phys. Rev. Lett. 78, 5 (1997).<br />

[188] A. Landragin, Réflexion d’atomes sur un miroir àondeévanescente: mesure de la<br />

<strong>for</strong>ce de van <strong>der</strong> Waals et diffraction atomique, thesis (Université deParisXI,Orsay,<br />

1997).<br />

[189] W. Seifert, C.S. Adams, V.I. Balykin, C. Heine, Yu. Ovchinnikov, and J. Mlynek,<br />

Reflection of metastable argon atoms from an evanescent wave, Phys.Rev.A49,<br />

3814 (1994).<br />

[190] S. Kawata and T. Sugiura, Movement of micrometer-sized particles in the evanescent<br />

field of a laser beam, Opt. Lett. 17, 772 (1992).<br />

[191] A. Einstein, Zur Quantentheorie <strong>der</strong> Strahlung, Phys. Zeitschr. 18, 121 (1917).<br />

[192] O.R. Frisch, Nachweis des Einsteinschen Strahlungsrückstoßes, Z. Phys. 86, 42<br />

(1933).<br />

[193] G. Roosen and C. Imbert, Étude de la pression de radiation de l’onde évanescente<br />

par absorption de lumière resonnante, Opt. Comm. 18, 247 (1976).<br />

[194] A. Landragin, G. Labeyrie, C. Henkel, R. Kaiser, N. Vansteenkiste,<br />

C.I. Westbrook, and A. Aspect, Specular versus diffuse reflection of atoms from an<br />

evanescent-wave mirror, Opt. Lett. 21, 1591 (1996).<br />

[195] C. Henkel, K. Mølmer, R. Kaiser, N. Vansteenkiste, C.I. Westbrook, and A. Aspect,<br />

Diffuse atomic reflection at a rough mirror, Phys.Rev.A55, 1160 (1997).<br />

[196] R.A. Cline, J.D. Miller, M.R. Matthews, and D.J. Heinzen, Spin relaxation of optically<br />

trapped atoms by light scattering, Opt. Lett. 19, 207 (1994).<br />

[197] T. Matsudo, Y. Takahara, H. Hori, and T. Sakurai, Pseudomomentum transfer<br />

from evanescent waves to atoms measured by saturated absorption spectroscopy,<br />

Opt. Comm. 145, 64 (1998).<br />

[198] Yu.B. Ovchinnikov, I. Manek, A.I. Sidorov, G. Wasik, and R. Grimm, Gravito-optical<br />

atom trap based on a conical hollow beam, Europhys. Lett. 43, 510 (1998).<br />

[199] M. Christ, A. Scholz, M. Schiffer, R. Deutschmann, and W. Ertmer, Diffraction<br />

and reflection of a slow metastable neon beam by an evanescent light grating,<br />

Opt. Comm. 107, 211 (1994).<br />

[200] A. Landragin, L. Cognet, G.Zs.K. Horvath, C.I. Westbrook, N. Westbrook, and<br />

A. Aspect, A reflection grating <strong>for</strong> atoms at normal incidence, Europhys. Lett. 39,<br />

485 (1997).<br />

[201] C. Henkel, H. Wallis, N. Westbrook, C.I. Westbrook, A. Aspect, K. Sengstock, and<br />

W. Ertmer, Theory of atomic diffraction from evanescent waves, Appl.Phys.B69,<br />

277 (1999).<br />

[202] R. Jansen, Elastic and inelastic bouncing of rubidium atoms on an evanescent-wave<br />

potential, diploma thesis, Universiteit van Amsterdam (2000, unpublished).<br />

[203] B.T. Wolschrijn et al., to be published.


Summary<br />

In atom optics, atomic matter wave are manipulated using, e.g., mirrors or lenses, in<br />

analogy to light optics. Of particular interest are evanescent-wave (EW) mirrors <strong>for</strong><br />

atoms. This is because of the short characteristic length, of the or<strong>der</strong> of the optical<br />

wavelength, at which a reflecting optical potential can be realised by an EW. This<br />

thesis is about our studies of photon scattering by cold (10 µK) rubidium atoms<br />

that bounce vertically on an EW mirror. Depending on the mirror configuration,<br />

we observed elastically and inelastically bouncing atoms. In the elastic case photon<br />

scattering leads to radiation pressure. The inelastic mirror is a consequence of<br />

optical hyperfine pumping of the atoms. It has no counterpart in light optics.<br />

Un<strong>der</strong>standing of scattering is important <strong>for</strong> our envisaged application of the<br />

inelastic mirror to load a low-dimensional optical dipole trap. The concept is to<br />

optically pump bouncing atoms with high spatial selectivity close to the turning<br />

point on the mirror. A spontaneous Raman transition transfers atoms into the<br />

trap. Dissipation allows the phase-space density to increase, possibly leading to<br />

a low-dimensional quantum degenerate gas by purely optical means. Ultimately a<br />

continuously operating “atom laser” might be realised, as a bright source of coherent<br />

matter waves. Being an open system out of thermal equilibrium, it would be in close<br />

analogy to an optical laser.<br />

In chapter 2, previous work on EW trap loading schemes with metastable noble<br />

gas atoms is extended to alkali atoms and, specifically, to 87 Rb. Heating of atoms<br />

by scattered photons is a severe loss mechanism in optical traps. The proposed lowdimensional<br />

trap allows spontaneously emitted photons to escape into a large solid<br />

angle without being reabsorbed. Due to the relatively small ground state hyperfine<br />

splitting of the alkali atoms, addressing of these states separately with bouncing and<br />

trapping potentials is difficult. As a solution the use of “dark states” is proposed. It<br />

is shown that the light scattering rate can be reduced by several or<strong>der</strong>s, to 20 s −1<br />

per atom. This requires the use of a circularly-polarised EW. It is discussed how to<br />

realise such a field using either a single beam or multiple beams. In the latter case,<br />

the EW may provide both a bouncing and a trapping potential, which aligns the<br />

atoms in parallel horizontal lines close to the surface.<br />

Experiments were per<strong>for</strong>med using a magneto-optical trap (MOT) with molasses<br />

cooling as a source of ≈ 10 7 cold atoms. The MOT is operated in an ultra-high vacuum<br />

(UHV) rubidium vapour cell. Optical access was achieved using a rectangular<br />

glass cuvette. The various laser frequencies were provided by a system of stabilised<br />

diode lasers, if necessary amplified by injection-locked diode lasers or travellingwave<br />

tapered semiconductor amplifiers (TA). An overview of the setup is given in<br />

chapter 3, including a detailed description of our UHV sealing techniques <strong>for</strong> glass<br />

cuvettes using either knife-edged metal gaskets or an epoxy resin.<br />

137


138 Summary<br />

Chapter 4 is dedicated to the characterisation of our TA systems, which provide<br />

150 − 200 mW power after single-mode optical fibres, with fibre coupling efficiencies<br />

up to 59 %. The relevance of broad spectral background due to amplified spontaneous<br />

emission (ASE) <strong>for</strong> the application with far-off resonance dipole potentials is<br />

discussed. Related to the rubidium optical transition linewidth of Γ/2π =6MHz,<br />

we observed an ASE suppression of better than −82 dB.<br />

In chapter 5, the efficiency of the EW mirror is investigated using a time-of-flight<br />

detection of bouncing atoms. Atom clouds were released from the MOT, 6 mm<br />

above the mirror. We observed bouncing fractions up to 9 %. These fractions result<br />

from the relation between the effective mirror surface and the ballistic expansion<br />

of falling atom clouds. At a temperature of 10 µK, the rms width of the cloud<br />

was approximately 1mm at the mirror, twice as large as in the MOT. The limited<br />

effective mirror surface was due to the transverse Gaussian intensity profile of the<br />

EW laser beam. The bouncing fraction was investigated as a function of laser power,<br />

detuning and polarisation. Also the temperature of the released cloud was varied<br />

between 6 − 20 µK. The measurements clearly show the significance of the Van <strong>der</strong><br />

Waals atom-surface interaction that reduces the effective mirror surface.<br />

In chapter 6, radiation pressure is studied, which is exerted on bouncing atoms.<br />

An EW does not propagate away from the surface. It propagates, however, along<br />

the surface. There<strong>for</strong>e the radiation pressure is directed parallel to the surface.<br />

Using fluorescence imaging with a camera, we studied this radiation pressure in<br />

terms of the horizontal velocity change of bouncing atoms. We observed 2 − 31<br />

photon recoils per bounce, and found the radiation pressure to be independent of<br />

the EW power, as expected from the exponential shape of the mirror potential. A<br />

simplifying two-level atom calculation <strong>for</strong> the number of scattered photons reveals<br />

an inverse proportionality to both laser detuning and EW decay parameter. This<br />

is in agreement with our observations. However, <strong>for</strong> steep EW potentials, in which<br />

atoms bounce very quickly within ≈ 1 µs and scatter only few photons, we observe<br />

an excess scattering of approximately 3 recoils. We assume that this is due to<br />

diffusely scattered light by the roughness of the prism surface.<br />

Bouncing occurred elastically when the EW was tuned to a “closed” optical transition.<br />

Despite of scattering photons, atoms then follow a single optical potential.<br />

In chapter 7, inelastic bouncing is investigated using an “open” transition, such that<br />

optical hyperfine pumping by the EW could transfer atoms into the different hyperfine<br />

ground state. The optical potential in this final state is lower. Bouncing atoms<br />

dissipate ≈ 90 % of their potential energy on the mirror. Using absorption imaging,<br />

we directly observed the density distribution of inelastically bouncing atom clouds.<br />

The high spatial selectivity of the pumping mechanism is revealed by a pronounced<br />

peak density <strong>for</strong> atoms that dissipate the maximum possible amount of energy. A<br />

broad (non-thermal) tail of faster atoms represents atoms being pumped further<br />

away from the turning point. By adjusting the laser detuning, we observed the<br />

inelasticity ranging between 81 − 92 %. This type of a single inelastic bounce can<br />

be interpreted as the fundamental step of a “Sisyphus” reflection cooling technique.<br />

Due to the relatively large fall height (6 mm) a single reflection leads, however, to<br />

heating. Multiple reflections would be necessary to achieve a net cooling.


Samenvatting<br />

Met laserlicht kunnen krachten op neutrale atomen worden uitgeoefend. Met deze<br />

krachten kunnen de atomen worden gemanipuleerd, en worden afgekoeld, d.w.z.<br />

afgeremd. Sinds Dehmelt, Hänsch, Schawlow en Wineland in 1975 de eerste voorstellen<br />

voor laserkoeling deden, werd de “atoomoptica” met vele technieken verrijkt.<br />

Zo werd on<strong>der</strong> an<strong>der</strong>e de Nobelprijs voor natuurkunde in 1997 uitgereikt aan Chu,<br />

Cohen-Tannoudji en Phillips voor hun bijdragen aan dit vakgebied.<br />

Analoog aan de elektromagnetische lichtgolf in de optica kan de atomaire materiegolf<br />

in de atoomoptica beïnvloed worden met spiegels, lenzen, tralies en straaldelers.<br />

Dit maakt het mogelijk om bijvoorbeeld uiterst nauwkeurige atoominterferometrische<br />

experimenten uit te voeren, zoals het meten van de fijnstructuurconstante.<br />

Met lasergekoelde atomen kunnen ook zeer precieze atoomklokken worden<br />

gerealiseerd, die bijvoorbeeld in de satellietnavigatie toegepast kunnen worden.<br />

In de atoomlithografie kunnen structuren op de schaal van een nanometer aangebracht<br />

worden.<br />

Waar in de optica lasers veelvuldig gebruikt worden als intensieve, goed gebundelde<br />

en vooral monochromatische (coherente) lichtbronnen, wordt in de atoomoptica<br />

tot dusver gewerkt met relatief zwakke “atomaire gloeilampen”. In het eenvoudigste<br />

geval betreft het een straal atomen die uit een oven ontsnapt en met<br />

behulp van een serie diafragma’s wordt gecollimeerd. Desalniettemin bestaan er<br />

voorstellen voor het realiseren van een bron van coherente materiegolven, een zogenaamde<br />

“atoomlaser”.<br />

Atoomoptische componenten bestaan meestal uit laserlicht van welbepaalde intensiteit,<br />

polarisatie en frequentie. In dit proefschrift worden spiegels voor rubidiumatomen<br />

(isotoop 87<br />

37Rb) beschreven, waarbij de bewegingsrichting van de atomen door<br />

een repulsieve “evanescente” lichtgolf elastisch wordt omgekeerd. In het bijzon<strong>der</strong><br />

wordt de lichtverstrooiing aan de atomen on<strong>der</strong>zocht. Deze verstrooiing is een dissipatief<br />

proces waardoor het mogelijk wordt dat atomen ook inelastisch gereflecteerd<br />

worden. Hiervoor bestaat in de lichtoptica geen analogie.<br />

Zoals Gauck en collega’s in Konstanz hebben laten zien, kan de inelastische<br />

reflectie ook “volledig inelastisch” gemaakt worden, oftewel de atomen kunnen in<br />

een tweedimensionale optische val aan het spiegeloppervlak worden geaccumuleerd.<br />

Als deze val efficiënt genoeg geladen kan worden moet het in principe mogelijk<br />

zijn een laagdimensionaal quantumgas te verkrijgen, vergelijkbaar met het Bose-<br />

Einstein condensaat in drie dimensies. De faseovergang van een thermisch gas naar<br />

het condensaat wordt in twee dimensies vervangen door de tot dusver nog niet<br />

waargenomen Kosterlitz-Thouless overgang.<br />

139


140 Samenvatting / Zusammenfassung<br />

Voorafgaande aan de beschrijving van onze experimenten wordt hier het principe<br />

van een evanescente spiegel voor atomen toegelicht.<br />

Een evanescente lichtgolf ontstaat, als licht volledig wordt gereflecteerd aan de<br />

grensvlak tussen twee diëlektrische media. In onze experimenten gaat het om een<br />

horizontaal glasoppervlak in vacuüm. In het vacuüm boven het glasoppervlak vinden<br />

wij een lichtgolf die parallel aan het oppervlak propageert. In tegenstelling<br />

tot de horizontale golfvectorcomponent is de verticale component imaginair. Dat<br />

betekent dat de elektrische veldsterkte exponentieel afneemt met de afstand tot het<br />

oppervlak. De karakteristieke afvallengte is van de orde van de gebruikte optische<br />

golflengte, bijvoorbeeld 780 nm voor rubidiumatomen. Hoe ver<strong>der</strong> de invalshoek van<br />

het licht voorbij de kritische hoek voor totale reflectie wordt ingesteld, hoe korter<br />

de afvallengte wordt.<br />

De krachten die een lichtgolf, bijvoorbeeld een evanescent veld, op een atoom kan<br />

uitoefenen, worden in het algemeen ingedeeld in twee categoriën, namelijk “spontane<br />

krachten” en “dipoolkrachten”.<br />

De spontane kracht is gebaseerd op herhaalde absorptie en spontane emissie van<br />

fotonen. De frequentie van de lichtgolf moet hiervoor afgestemd zijn op een optische<br />

resonantie van de atomen, of tenminste niet ver daarvan af. Omdat in de regel<br />

een gerichte laserstraal gebruikt wordt, zijn de terugstoten die de atomen door de<br />

absorptie van fotonen krijgen ook gericht, zodat er een netto kracht resulteert. Deze<br />

kracht wordt ook wel “stralingsdruk” genoemd. De emissie van fotonen is spontaan<br />

en gemiddeld ongericht. Zodoende is deze kracht dissipatief en kan gebruikt worden<br />

voor het koelen van atomen. Ze leidt echter ook tot diffusie.<br />

De dipoolkracht daarentegen is een conservatieve kracht en kan door een “optische<br />

potentiaal” worden beschreven. Deze potentiaal ontstaat door de wisselwerking<br />

tussen het elektrische veld van het licht en de geïnduceerde elektrische dipoolmomenten<br />

van de atomen. De dipoolkracht heeft een dispersief karakter: als de<br />

lichtfrequentie kleiner is dan die van de atomaire overgang (rood verstemd) worden<br />

de atomen aangetrokken, als de frequentie groter is (blauw verstemd) worden ze<br />

juist afgestoten. Om ook daadwerkelijk een konservatieve kracht te verkrijgen, is het<br />

nodig om de frequentieverstemming groot te kiezen, zodat absorptie (stralingsdruk)<br />

geen rol speelt.<br />

Een ver naar het blauw verstemd evanescent lichtveld vormt dus een potentiaalbarrière<br />

voor atomen. Als de kinetische energie van de invallende atomen kleiner<br />

is dan de hoogte van het maximum van deze barrière zullen de atomen van richting<br />

omkeren. Zo wordt het evanescente veld een spiegel of een “trampoline” voor<br />

atomen.<br />

In onze experimenten wordt om te beginnen een 1mm grote koude wolk van<br />

ca. 100 miljoen rubidiumatomen ingevangen in een magneto-optische val. De wolk<br />

heeft een temperatuur van 10 µK boven het absolute nulpunt van −273.15 ◦ C. Bij<br />

deze temperatuur hebben de atomen een snelheid van enkele centimeter per seconde.<br />

Ter vergelijking: de snelheid bij kamertemperatuur bedraagt enkele hon<strong>der</strong>den meter<br />

per seconde. De atomen worden losgelaten uit de magneto-optische val en vallen<br />

vanaf een hoogte van ongeveer 5 tot 7 mm op de spiegel. Tijdens de val expandeert<br />

de wolk ballistisch en wordt in doorsnede verdubbeld. Het intensiteitsprofiel van


Samenvatting / Zusammenfassung 141<br />

het evanescente veld is gaussvormig, en zo ook het effektieve spiegeloppervlak. Dit<br />

oppervlak wordt begrenst door de contour waar de minimale barrièrehoogte nog net<br />

bereikt wordt. In het algemeen wordt slechts een deel van de wolk gereflecteerd.<br />

Deze spiegelefficientie hebben wij on<strong>der</strong>zocht als functie van de temperatuur <strong>der</strong><br />

atomen en van de laserparameters. Het effektieve spiegeloppervlak wordt bepaald<br />

door het vermogen van de laser, de frequentieverstemming en de polarisatie. Met<br />

ons evanescente veld bereikten wij spiegelefficienties tot 8 %, in overeenstemming<br />

met voorspellende berekeningen. We nemen waar dat de aantrekkende Van <strong>der</strong><br />

Waalskracht tussen de atomen en het glas de optische potentiaal verlaagt waardoor<br />

het effektieve spiegeloppervlak significant kleiner wordt.<br />

De stralingsdruk van het evanescente veld kan on<strong>der</strong>zocht worden door de beweging<br />

van de atomen met een camera te registreren. Deze druk wordt veroorzaakt<br />

door de voortplanting van de lichtgolf langs het glasoppervlak. De terugstoten van<br />

de fotonen zorgen alleen voor veran<strong>der</strong>ing van de snelheid van de atomen in deze<br />

richting, evenwijdig aan het oppervlak. Eén enkele terugstoot geeft een snelheidsveran<strong>der</strong>ing<br />

van 6 mm/s. In het experiment nemen wij horizontale snelheidsveran<strong>der</strong>ingen<br />

van de gereflecteerde atomen waar van 1tot 18 cm/s, overeenkomend<br />

met de absorptie van 2 tot 31fotonen. De verticale snelheid na een val van 6.6 mm<br />

bedraagt 36 cm/s.<br />

Rubidium heeft een hyperfijnstructuur met twee grondtoestanden en meer<strong>der</strong>e<br />

optisch aangeslagen toestanden. Desalniettemin kunnen de waarnemingen kwalitatief<br />

verklaard worden met behulp van een eenvoudig model met slechts éen grondtoestand<br />

en éen aangeslagen toestand. Er werd namelijk een “gesloten” overgang<br />

aangeslagen zodat de atomen na ca. 30 ns weer in de oorspronkelijke hyperfijntoestand<br />

terugvielen en dezelfde potentiaalcurve in het evanescente veld konden volgen.<br />

Hiermee is de aaname van het tweeniveausysteem gerechtvaardigd. Het gemiddelde<br />

aantal opgenomen terugstoten per atoom is evenredig aan de afvallengte van<br />

het veld en aan de snelheid van de atomen. In een steile potentiaal zullen de atomen<br />

snel omkeren zodat ze weinig tijd hebben om fotonen te absorberen. Afhankelijk<br />

van de ingestelde afvallengte duurt de reflectie 3 tot 10 µs. Een langzaam atoom<br />

dringt min<strong>der</strong> diep in het evanescente veld door als een snel atoom en heeft dus ook<br />

min<strong>der</strong> tijd nodig om om te keren. Daarnaast is het aantal geabsorbeerde fotonen<br />

omgekeerd evenredig met de blauwverstemming van de laser. Het aantal is daarentegen<br />

niet afhankelijk van de laserintensiteit. Dit hangt samen met het exponentiële<br />

verloop van de optische potentiaal, waardoor het atoom steeds eenzelfde traject tot<br />

zijn omkeerpunt aflegt onafhankelijk van hoever het nog is tot het maximum.<br />

We hebben de hyperfijnstructuur van het rubidium gebruikt om ook inelastische<br />

reflectie te kunnen waarnemen. Hiervoor stemden we de frequentie van het<br />

evanescente veld af op een “open” optische overgang. Een atoom kan dan vanuit<br />

grondtoestand |1〉 via een aangeslagen toestand en spontane emissie worden overgepompt<br />

naar de an<strong>der</strong>e hyperfijngrondtoestand |2〉. Intoestand|2〉 is de hoogte van<br />

de potentiaalbarrière voor het atoom echter nog maar 10 % van de oorspronkelijke<br />

waarde omdat de blauwverstemming ten opzichte van de overgang vanuit toestand<br />

|2〉 tien keer zo groot is. Door het spontaan emitteren van een foton dissipeert het<br />

atoom ca. 90 % van zijn potentiële energie zodat het na deze inelastische reflectie nog


142 Samenvatting / Zusammenfassung<br />

maar 10 % van zijn oorspronkelijke hoogte kan bereiken. Opnamen met de camera<br />

laten deze inelastische beweging van het zwaartepunt van de gereflecteerde wolk van<br />

atomen duidelijk zien. De waarschijnlijkheid voor het overpompen is het grootst<br />

als een atoom zich vlakbij het omkeerpunt bevindt, waar zijn snelheid het kleinst<br />

is en de lichtintensiteit het grootst. Daar is bovendien het energieverlies dat bij<br />

overpompen optreedt het grootst. Op grond van deze sterke ruimtelijke selectiviteit<br />

worden de meeste atomen dan ook maximaal inelastisch gereflecteerd.<br />

Dankzij deze selectiviteit kan een inelastische spiegel gebruikt worden om atomen<br />

die eerst in een relatief grote magneto-optische val geprepareerd worden, efficient te<br />

concentreren in een dunne tweedimensionale optische val dicht boven het glasoppervlak.<br />

Voor dit invangen moet aan de potentiaal van toestand (2) een potentiaalput<br />

toegevoegd worden waarin de overgepompte atomen kunnen worden gebonden. Een<br />

belangrijke consequentie van het dissipatieve karakter van de optische pompovergang<br />

is, dat de dichtheid van de atomen in zo’n val —bij gelijke temperatuur—<br />

enkele orden van grootte meer kan bedragen dan de oorspronkelijke dichtheid in de<br />

magneto-optische val. Bij voldoende ladingsefficientie moet het dan ook mogelijk<br />

zijn om met behulp van een eenmalig optisch proces, zon<strong>der</strong> ver<strong>der</strong>e afkoeling, een<br />

laagdimensionaal quantumgas te verkrijgen. Dit wordt hieron<strong>der</strong> besproken.<br />

We spreken van een quantumgas als de dichtheid van het gas zo hoog is en de<br />

temperatuur zo laag, dat de golffuncties van de atomen elkaar overlappen. (Bijvoorbeeld:<br />

de De Broglie golflengte van rubidium is bij 10 µK ongeveer 60 nm.) In dit<br />

geval moet voor de beschrijving van de atomen de klassieke Boltzmann statistiek<br />

vervangen worden door de quantumstatistiek van niet on<strong>der</strong>scheidbare bosonen of<br />

fermionen. In het geval van bosonen, zoals rubidium ( 87<br />

37Rb) met zijn heeltallige spin,<br />

ontstaat bij een temperatuur lager dan een kritische waarde het in 1924 reeds voorspelde<br />

Bose-Einstein condensaat. In het condensaat bevindt zich een macroscopisch<br />

aantal atomen in éen enkele toestand. Experimenteel werd zo’n condensaat pas in<br />

1995 aangetoond door An<strong>der</strong>son en medewerkers uit Boul<strong>der</strong>.<br />

Het is tot dusver niet gelukt deze quantumontaarding enkel door middel van optisch<br />

koelen te bereiken. Dit is te wijten aan de spontaan geemitteerde fotonen die<br />

voor de koeling essentieel zijn. Omdat deze fotonen telkens opnieuw geabsorbeerd<br />

worden, wordt de te bereiken temperatuur gelimiteerd door de terugstoten. In een<br />

magneto-optische val kan men zo slechts een gemiddelde atomaire afstand bereiken<br />

die ongeveer hon<strong>der</strong>d keer groter is dan de De Broglie golflengte. Een condensaat<br />

werd daarom tot nu toe altijd verkregen door gebruik te maken van verdampingskoelen<br />

in een magnetische val. Bij deze koelmethode ontsnappen voortdurend de<br />

snelste atomen uit de val waardoor een klein restant van enkele duizenden tot miljoenen<br />

atomen kan afkoelen tot het condensatiepunt.<br />

Een inelastische atoomspiegel zou mogelijkerwijs de kloof tussen de dichtheid in<br />

de magneto-optische val en die van een quantumgas kunnen overbruggen. Dankzij<br />

de tweedimensionale geometrie zouden spontaan geëmitteerde fotonen bovendien<br />

probleemloos kunnen ontsnappen. Het laden van de optische val kan in principe ook<br />

continu plaatsvinden. Op deze manier zou een continue atoomlaser als open systeem<br />

kunnen bestaan, qua concept vergelijkbaar met de lichtlaser: de voorgekoelde thermische<br />

atomen dienen als versterkend medium en de optische val vormt de trilholte


Samenvatting / Zusammenfassung 143<br />

voor materiegolven. Door optisch pompen worden atomen toegevoegd aan de resonante<br />

modes van deze trilholte. Dit proces is optisch gezien weliswaar spontaan,<br />

maar wordt door de bosonenstatistiek van de atomen gestimuleerd. De tot dusver<br />

geconstrueerde atoomlasers zijn gebaseerd op het weglekken van atomen uit een<br />

met behulp van verdampingskoelen geproduceerd condensaat en zijn op hun best<br />

quasicontinu.<br />

Een atoomlaser zou vergelijkbare vooruitgang in de precisie van atoomoptische<br />

experimenten kunnen opleveren als de ontwikkeling van de lichtlaser. Sinds deze<br />

in 1960 voor het eerst door Maiman werd geconstrueerd is hij zelfs in het dagelijks<br />

leven doorgedrongen. Zo zijn de bij onze experimenten gebruikte halfgelei<strong>der</strong>lasers<br />

eigenlijk bedoeld voor CD spelers.<br />

Samengevat levert dit proefschrift een experimentele bijdrage aan de atoomoptica<br />

van evanescente spiegels voor koude atomen. De lichtverstrooiing van atomen in het<br />

evanescente veld werd on<strong>der</strong>zocht voor zowel elastische als inelastische spiegels. Met<br />

het oog op alkaliatomen, en 87 Rb in het bijzon<strong>der</strong>, werd een concept ontwikkeld om<br />

atomen met gebruik van de inelastische spiegel in een laagdimensionale val over te<br />

brengen.


Zusammenfassung<br />

Mit Laserlicht lassen sich Kräfte auf neutrale Atome ausüben, mit welchen diese<br />

manipuliert, vor allem aber auch gekühlt, beziehungsweise abgebremst werden können.<br />

Ausgehend von den ersten Vorschlägen zur Laserkühlung von Dehmelt, Hänsch,<br />

Schawlow und Wineland, 1975, wurde die “Atomoptik” in den letzten Jahren um<br />

viele Techniken bereichert. Unter an<strong>der</strong>em wurde 1997 <strong>der</strong> Physik-Nobelpreis an<br />

Chu, Cohen-Tannoudji und Phillips für ihre Beiträge zu diesem Fachgebiet verliehen.<br />

In Analogie zur elektromagnetischen Lichtwelle wird in <strong>der</strong> Atomoptik die atomare<br />

Materiewelle mittels Spiegeln, Linsen, Beugungsgittern und Strahlteilern beeinflußt.<br />

Dies ermöglicht beispielsweise hochpräzise atominterferometrische Experimente,<br />

wie zur Bestimmung <strong>der</strong> Feinstrukturkonstanten. Mit lasergekühlten Atomen<br />

lassen sich auch sehr genaue Atomuhren verwirklichen, die in satellitengestützten<br />

Navigationssystemen eingesetzt werden könnten. Eine Ergänzung zur Lichtoptik<br />

findet sich in <strong>der</strong> Atomlithographie mit dem Schreiben von Strukturen auf Nanometerskala.<br />

Während in <strong>der</strong> Optik vielfach Laser als intensive, gut gebündelte und vor allem<br />

monochromatische (kohärente) Lichtquellen Verwendung finden, wird in <strong>der</strong> Atomoptik<br />

bislang mit relativ schwachen “atomaren Glühlampen” gearbeitet. Im einfachsten<br />

Fall ist dies ein durch eine Serie von Diaphragmen kollimierter Strahl von<br />

Atomen, <strong>der</strong> aus einem Ofen entweicht. Es gibt allerdings Ansätze zur Verwirklichung<br />

kohärenter Materiewellenquellen, sogenannten “Atomlasern”.<br />

Das “Substrat” atomoptischer Komponenten ist zumeist Laserlicht von genau<br />

festgelegter Intensität, Polarisation und Lichtfrequenz. Die vorliegende Dissertation<br />

befaßt sich mit Spiegeln für Rubidiumatome (Isotop 87<br />

37Rb), bei denen ein repulsives<br />

“evaneszentes” Lichtfeld für die elastische Bewegungsumkehr <strong>der</strong> Atome sorgt. Insbeson<strong>der</strong>e<br />

wird die Lichtstreuung <strong>der</strong> Atome im evaneszenten Feld untersucht. Diese<br />

ermöglicht es als dissipativer Prozeß, Atome auch inelastisch zu reflektieren. Hierfür<br />

besteht keine Analogie mit lichtoptischen Spiegeln.<br />

Wie von Gauck und Kollegen in Konstanz demonstriert, ist es auch möglich,<br />

die inelastische Reflexion “vollständig inelastisch” zu gestalten, sprich, die Atome<br />

in einer zweidimensionalen optischen Falle an <strong>der</strong> Spiegeloberfläche anzusammeln.<br />

Bei hinreichen<strong>der</strong> Effizienz dieses Transfers in die Falle sollte es prinzipiell möglich<br />

sein, ein niedrigdimensionales Quantengas zu erzielen, vergleichbar mit dem Bose-<br />

Einstein Kondensat in drei Dimensionen. Dem Phasenübergang vom thermischen<br />

Gas zum Kondensat beim Unterschreiten einer kritischen Temperatur entspricht<br />

in zwei Dimensionen beispielsweise <strong>der</strong> bislang noch nicht beobachtete Kosterlitz-<br />

Thouless-Übergang.<br />

144


Samenvatting / Zusammenfassung 145<br />

Vor <strong>der</strong> Beschreibung unserer Experimente wird im folgenden zunächst das Funktionsprinzip<br />

eines evaneszenten <strong>Atoms</strong>piegels erläutert.<br />

Ein evaneszentes Lichtfeld entsteht, wenn Licht an <strong>der</strong> Grenzschicht zweier dielektrischer<br />

Medien, vom optisch dichteren Medium aus, vollständig reflektiert wird.<br />

In unseren Experimenten geschieht dies an einer horizontalen Glasoberfläche im<br />

Vakuum. Auf <strong>der</strong> Vakuumseite findet sich ein Lichtfeld, das in Richtung parallel<br />

zur Oberfläche propagiert. Im Gegensatz zur horizontalen Komponente des Wellenvektors<br />

ist die vertikale Komponente komplex imaginär. Dadurch nimmt die elektrische<br />

Feldstärke mit zunehmendem Abstand von <strong>der</strong> Oberfläche exponentiell ab.<br />

Die charakteristische Abfallänge des Feldes ist von <strong>der</strong> Größenordnung <strong>der</strong> verwendeten<br />

optischen Wellenlänge, zum Beispiel 780 nm für Rubidiumatome. Je weiter<br />

<strong>der</strong> Lichteinfallswinkel den kritischen Winkel <strong>der</strong> Totalreflexion überschreitet, desto<br />

kürzer wird diese Abfallänge.<br />

Kräfte, die ein Lichtfeld — auch ein evaneszentes — auf Atome ausüben kann,<br />

werden im allgemeinen in zwei Kategorien eingeteilt, “spontane Kräfte” und “Dipolkräfte”.<br />

Die spontane Kraft basiert auf <strong>der</strong> wie<strong>der</strong>holten Absorption und Spontanemission<br />

von Photonen eines Lichtfeldes, dessen Frequenz auf eine optische Resonanz <strong>der</strong><br />

Atome abgestimmt, zumindest aber nicht weit davon verstimmt ist. Da die Absorption<br />

in aller Regel aus einem gerichteten Laserstrahl erfolgt, sind auch die vom<br />

Atom aufgenommenen Photonenrückstöße gerichtet und resultieren in einer Kraft,<br />

auch “Strahlungsdruck” genannt. Die Emission <strong>der</strong> Photonen erfolgt spontan und<br />

ist im Mittel ungerichtet. Daher ist diese Lichtkraft dissipativ und kann zum Kühlen<br />

verwendet werden. Sie führt aber auch zur Diffusion <strong>der</strong> Atome.<br />

Die Dipolkraft ist hingegen eine konservative Kraft und kann durch ein “optisches<br />

Potential” beschrieben werden. Dieses Potential resultiert aus <strong>der</strong> Wechselwirkung<br />

des induzierten atomaren elektrischen Dipolmoments mit dem elektrischen Feld. Die<br />

Dipolkraft ist dispersiv. Das heißt, sie ist attraktiv, wenn die Lichtfrequenz kleiner<br />

als die atomare Resonanz ist (Rotverstimmung) und repulsiv im Falle einer größeren<br />

Frequenz (Blauverstimmung). Um experimentell tatsächlich eine konservative Kraft<br />

zu erzielen, wählt man eine große Verstimmung. Damit wird die Lichtabsorption,<br />

beziehungsweise die spontane Kraft weitgehend unterdrückt.<br />

Ein weit blauverstimmtes evaneszentes Lichtfeld stellt demnach eine Potentialbarriere<br />

für Atome dar. Wenn die kinetische Energie eines einfallenden <strong>Atoms</strong> die<br />

maximale Höhe dieser Barriere nicht überschreitet, wird die Bewegungsrichtung<br />

umgekehrt. Das evaneszente Feld ist dann ein “Spiegel” o<strong>der</strong> auch ein “Trampolin”<br />

für Atome.<br />

Im Experiment präparieren wir zunächst in einer magneto-optischen Falle eine<br />

ca. 1mm durchmessende kalte Rubidiumwolke von etwa 100 Millionen Atomen. Die<br />

Temperatur <strong>der</strong> Wolke ist 10 µK, dicht am Temperaturnullpunkt von −273.15 ◦ C.<br />

Dieser Temperatur entspricht eine mittlere Geschwindigkeit <strong>der</strong> Atome von wenigen<br />

cm/s, im Vergleich zu einigen 100 m/s bei Raumtemperatur. (Die magneto-optische<br />

Falle, das Vakuumsystem und die verwendeten Laser werden in den Kapiteln 3 und 4<br />

dieser Dissertation im Detail beschrieben.)


146 Samenvatting / Zusammenfassung<br />

Die Rubidiumatome werden aus 5−7 mmHöhe aus <strong>der</strong> magneto-optischen Falle<br />

auf den <strong>Atoms</strong>piegel fallen gelassen. Während des freien Falls expandiert die Wolke<br />

ballistisch und verdoppelt dabei ihren Durchmesser. Da das evaneszente Feld ein<br />

horizontal gaußförmiges Intensitätsprofil aufweist, ist die effektive Spiegeloberfläche<br />

durch das für eine Reflexion <strong>der</strong> Atome minimal notwendige optische Potential begrenzt,<br />

und es wird im allgemeinen nur ein Teil <strong>der</strong> Atome aus <strong>der</strong> Wolke tatsächlich<br />

reflektiert.<br />

Dies haben wir als Spiegeleffizienz in Abhängigkeit <strong>der</strong> Laserparameter und <strong>der</strong><br />

Temperatur <strong>der</strong> Atome untersucht (Kapitel 5). Laserleistung, -verstimmung und<br />

-polarisation geben die effektive Spiegeloberfläche vor. Mit unserem evaneszenten<br />

Feld erreichten wir Spiegeleffizienzen bis zu 8 %, in Übereinstimmung mit berechneten<br />

Vorhersagen. Es zeigte sich, daß die attraktive Van <strong>der</strong> Waals-Wechselwirkung<br />

zwischen Atomen und Glassubstrat das optische Potential erniedrigt und dadurch<br />

die effektive Spiegelfläche signifikant verringert.<br />

Der Strahlungsdruck des evaneszenten Feldes läßt sich untersuchen, wenn man<br />

die Bewegung <strong>der</strong> reflektierten Atome mit einer Kamera beobachtet (Kapitel 6).<br />

Die Ursache des Strahlungsdrucks ist die Propagation <strong>der</strong> Lichtwelle entlang <strong>der</strong><br />

Glasoberfläche. Die Rückstöße absorbierter Photonen än<strong>der</strong>n die mittlere Geschwindigkeit<br />

<strong>der</strong> Atome allein in dieser Richtung parallel zur Oberfläche. Ein einzelner<br />

Photonenrückstoß än<strong>der</strong>t die Geschwindikeit um 6 mm/s. Im Experiment<br />

beobachteten wir an reflektierten Atomen seitliche Geschwindigkeiten von 1cm/s<br />

bis zu 18 cm/s, entsprechend 2 bis 31Rückstößen. Im Vergleich dazu beträgt die<br />

vertikale Geschwindigkeit nach einem Fall aus 6.6 mmHöhe 36 cm/s,<br />

Ein vereinfachendes Modell für ein Zweiniveau-Atom kann diese Beobachtungen<br />

qualitativ erklären, trotz <strong>der</strong> tatsächlichen Hyperfeinstruktur <strong>der</strong> Rubidiumatome<br />

mit zwei Grund- und mehreren Anregungszuständen. Die Atome wurden auf einem<br />

“geschlossenen” Übergang optisch angeregt, so daß sie nach jeweils ca. 30 ns wie<strong>der</strong><br />

in den ursprünglichen Grundzustand zurück gelangten und weitgehend ungestört<br />

ein und <strong>der</strong>selben Potentialkurve im evaneszenten Feld folgten. Dies rechtfertigt<br />

die Vereinfachung auf nur zwei Zustände. Es zeigt sich, daß die gemittelte Zahl<br />

<strong>der</strong> pro Atom aufgenommenen Rückstöße proportional zur Abfallänge des Feldes<br />

und zur Geschwindigkeit <strong>der</strong> Atome ist. Auf einer steilen Potentialbarriere vollzieht<br />

sich die Bewegungsumkehr sehr schnell, und ein Atom kann nur wenige Photonen<br />

absorbieren. Je nach eingestellter Abfallänge dauert die Reflexion nur 3 − 10 µs.<br />

Ein langsameres Atom dringt zudem weniger tief in das evaneszente Feld ein als<br />

ein schnelleres Atom und benötigt weniger Zeit für die Reflexion. Im übrigen ist<br />

die Zahl <strong>der</strong> absorbierten Photonen umgekehrt proportional zur Laserverstimmung.<br />

Sie hängt aber nicht von <strong>der</strong> Laserintensität ab. Dies resultiert aus dem exponentiellen<br />

Verlauf des optischen Potentials, in dem ein Atom unabhängig vom möglicherweise<br />

geän<strong>der</strong>ten Potentialmaximum stets in gleicher Weise zu seinem Umkehrpunkt<br />

gelangt.<br />

Die Hyperfeinstruktur des Rubidiums haben wir genutzt, um auch inelastische<br />

Reflexionen zu beobachten, indem wir das evaneszente Feld bezüglich eines<br />

“offenen” optischen Übergangs abstimmten (Kapitel 7). Ein vom Ausgangszustand<br />

|1〉 angeregtes Atom kann dann mittels “optischen Pumpens” spontan in


Samenvatting / Zusammenfassung 147<br />

den an<strong>der</strong>en Hyperfeingrundzustand |2〉 übergehen. Im Zustand |2〉 besitzt die<br />

Potentialbarriere für das Atom jedoch nur 10 % <strong>der</strong> ursprünglichen Höhe, weil die<br />

Frequenzverstimmung bezüglich <strong>der</strong> Resonanzen im Zustand |2〉 ungefähr zehnfach<br />

größer ist. Durch das spontan emittierte Photon “dissipiert” das umgepumpte<br />

Atom ca. 90 % seiner potentiellen Energie und erreicht nach dieser inelastischen<br />

Reflexion nur noch 10 % <strong>der</strong> Fallhöhe. Kamera-Aufnahmen <strong>der</strong> reflektierten Atome<br />

zeigen in <strong>der</strong> Schwerpunktsbewegung <strong>der</strong> Wolke die Inelastizität des Spiegels. Die<br />

Wahrscheinlichkeit für optisches Pumpen ist am größten, während ein Atom sich<br />

— beinahe im Stillstand — in großer Lichtintensität nahe dem Umkehrpunkt befindet.<br />

Dort ist zudem die inelastische Energieabnahme am größten. Die meisten<br />

Atome finden sich aufgrund <strong>der</strong> starken räumlichen Selektivität des Pumpvorgangs<br />

maximal abgebremst.<br />

Dank dieser Selektivität kann ein inelastischer Spiegel genutzt werden, um Atome,<br />

die zunächst in einer relativ ausgedehnten magneto-optischen Falle präpariert<br />

wurden, effizient in einer sehr dünnen, zweidimensionalen optischen Falle dicht bei<br />

<strong>der</strong> Glasoberfläche anzusammeln (Kapitel 2). Dem Potential für Atome im inelastisch<br />

reflektierten Endzustand |2〉 muß hierzu ein “Potentialtopf”, die optische<br />

Falle, aufgeprägt werden. In dieser werden die umgepumpten Atome gebunden. Eine<br />

wichtige Konsequenz aus dem dissipativen Charakter des Pumpvorgangs ist, daß<br />

die Dichte <strong>der</strong> so angesammelten Atome die ursprüngliche Dichte in <strong>der</strong> magnetooptischen<br />

Falle — bei gleichbleiben<strong>der</strong> Temperatur — um mehrere Größenordnungen<br />

übersteigen kann. Bei hinreichen<strong>der</strong> Transfereffizienz in die Falle sollte es möglich<br />

sein, in einem einzigen, optisch spontanen Vorgang und ohne weiteres Nachkühlen<br />

ein niedrigdimensionales “Quantengas” zu erzeugen. Dies wird im Folgenden abschließend<br />

erläutert.<br />

Ein Quantengas liegt vor, wenn bei hinreichend großer Dichte und niedriger<br />

Temperatur die Wellenfunktionen <strong>der</strong> Atome einan<strong>der</strong> überlappen. (Bei 10 µK<br />

Temperatur ist die De Broglie-Wellenlänge für Rubidium beispielsweise 60 nm.) In<br />

diesem Fall muß man in <strong>der</strong> Beschreibung <strong>der</strong> Atome von <strong>der</strong> klassischen Boltzmann-<br />

Statistik zur Quantenstatistik nicht unterscheidbarer Bosonen o<strong>der</strong> Fermionen übergehen.<br />

Für Bosonen, 87<br />

37Rb ist mit ganzzahligem Spin ein solches, <strong>for</strong>mt sich beim<br />

Unterschreiten einer kritischen Temperatur das schon 1924 vorhergesagte Bose-<br />

Einstein Kondensat. Im Kondensat findet sich eine große Zahl von Atomen in<br />

einem einzigen, makroskopisch besetzten Zustand, einem beinahe mit bloßem Auge<br />

sichtbaren Quantenobjekt. Erstmals gelang <strong>der</strong> experimentelle Nachweis in einem<br />

atomaren Gas An<strong>der</strong>son und Kollegen, 1995, in Boul<strong>der</strong>.<br />

Mit optischen Kühlverfahren ist es bislang nicht gelungen, ein Quantenengas<br />

zu erzeugen. Das Problem liegt an den für die Dissipation essentiellen spontan<br />

emittierten Photonen. Deren Rückstöße und auch die Wie<strong>der</strong>absorption durch<br />

umgebende Atome limitieren die erreichbare Temperatur. In einer magneto-optischen<br />

Falle erreicht man beispielsweise einen mittleren Atomabstand entsprechend<br />

<strong>der</strong> 100-fachen De Broglie-Wellenlänge. Kondensate wurden deshalb generell durch<br />

Verdampfungskühlen in einer magnetischen Falle erzeugt, wobei die schnellsten<br />

Atome aus <strong>der</strong> Falle entschnappen, “verdampfen”, und so ein kleiner Rest von einigen<br />

Tausend bis Millionen Atomen bis zur Kondensation abkühlt.


148 Samenvatting / Zusammenfassung<br />

Ein inelastischer <strong>Atoms</strong>piegel könnte die Dichtheitslücke von <strong>der</strong> magneto-optischen<br />

Falle zum Quantengas überspannen. Dank <strong>der</strong> zweidimensionalen Geometrie<br />

können spontane Photonen zudem schadlos entweichen. Das Laden <strong>der</strong> optischen<br />

Falle kann prinzipiell auch kontinuierlich erfolgen. Auf diese Weise ist in konzeptioneller<br />

Analogie zum Lichtlaser ein kontinuierlicher Atomlaser als offenes System<br />

denkbar: Als Verstärkungsmedium dienen (vorgekühlte) thermische Atome, und<br />

die optische Falle <strong>for</strong>mt den Materiewellenresonator. Optisches Pumpen führt den<br />

Resonatormoden Atome zu. Dieser Prozeß ist zwar optisch spontan, wird aber<br />

durch die Quantenstatistik <strong>der</strong> Atome bosonisch stimuliert. Bislang demonstrierte<br />

Atomlaser basieren auf dem “Auslaufen” von Atomen aus einem einmalig durch<br />

Verdampfungskühlen in einer Magnetfalle erzeugten Kondensat und sind bestenfalls<br />

quasikontinuierlich.<br />

Ein Atomlaser könnte in <strong>der</strong> Präzision atomoptischer Experimente vergleichbare<br />

Fortschritte bringen wie die Entwicklung des Lichtlasers, <strong>der</strong> 1960 erstmals von<br />

Maiman demonstriert wurde und inzwischen bis ins Alltagsleben vorgedrungen ist.<br />

CD-Spieler enthalten beispielsweise eine vergleichbare Laserdiode, wie wir sie zum<br />

Kühlen unserer Atome verwenden.<br />

Zusammenfassend liefert die vorliegende Dissertation einen experimentellen Beitrag<br />

zur Atomoptik mit evaneszenten Spiegeln für kalte Atome. Es wurde die<br />

Lichtverstreuung von Atomen im evaneszenten Feld sowohl für elastische wie inelastische<br />

Spiegel untersucht. Mit Blick auf Alkali-Atome, 87 Rb im beson<strong>der</strong>en, wurde<br />

unter Nutzung des inelastischen Spiegels ein Konzept entwickelt, um Atome unter<br />

minimalen Verlusten in eine niedrigdimensionale optische Falle zu transferieren.


Nawoord<br />

Op deze plaats wil ik graag alle mensen van harte bedanken die mij tijdens de<br />

laatste jaren bij het werk en in privé altijd hebben aangemoedigd en geholpen. Een<br />

on<strong>der</strong>zoek zoals in dit boek beschreven is al gauw te complex om door éen enkele<br />

promovendus te kunnen worden uitgevoerd. Zo wil ik vooral mijn promotoren Ben<br />

van Linden van den Heuvell en Robert Spreeuw bedanken voor hun vertrouwen<br />

waarmee zij mij, in 1995, in een volstrekt kale experimenteerkamer het on<strong>der</strong>zoek<br />

lieten beginnen. Stefan Petra kwam er al gauw bij als eerste stagestudent, gevolgd<br />

door Esther Schil<strong>der</strong>, Rik Jansen en nu Aal<strong>der</strong>t van Amerongen, samen met mijn<br />

collega promovendi Bas Wolschrijn, Ronald Cornelussen, en onze postdoc Nandini<br />

Bhattacharya. Fre<strong>der</strong>ik de Jong zat er al aan het laseren zon<strong>der</strong> inversie en had<br />

altijd goed advies waar en bij wie je moet wezen. Cor Snoek wist altijd ergens een<br />

oud apparaat met tandwielen erin te vinden waartegen menig nieuw apparaat het<br />

zou moeten afleggen. Maar het leukst waren zijn verhalen erbij, uit het verleden van<br />

het laboratorium. Jullie allen bedankt voor de goede samenwerking en de geduld<br />

met de lasertjes.<br />

Mijn begelei<strong>der</strong>s Ben en Robert waren letterlijk altijd gereed voor enthousiast<br />

overleg en advies. Robert wist op elk experimenteel en natuurkundig probleem een<br />

mogelijke oplossing. Hij keerde elke <strong>for</strong>mule en elk apparaat binnenste buiten om<br />

tot de kern van de zaak door te dringen.<br />

Ben kun je zó een probleem van welke makelij dan ook naar het hoofd gooien, het<br />

wordt onmiddellijk naar Mathematica vertaald en het antwoord komt meteen terug,<br />

zij het vaak wat impliciet. Maar dit maakte de navolgende uitleg des te leuker en<br />

vergrootte het inzicht in het probleem.<br />

Een experiment werkt niet zon<strong>der</strong> apparatuur. Daarom een dank aan al de collega’s<br />

die in de werkplaatsen hun vakkennis zo enthousiast hebben ingebracht. Floris<br />

van <strong>der</strong> Woude, Willem van Aartsen, Fred van Anrooij, Martin Bijlsma en Jeroen<br />

Jacobs hebben ons met van alles voorzien, van “kikkers” tot de allerfijnste mechanica<br />

zoals Floris’ stokpaardje, de gleufscharnieren. Soms kwamen zij weliswaar de<br />

wanhoop nabij als ik weer eens niet begreep dat een centimeter duizend hon<strong>der</strong>dste<br />

millimeter bevat en geen éen meer of min<strong>der</strong>.<br />

Alof Wassink, Johan te Winkel, Flip de Leeuw, Edwin Baaij en Theo van<br />

Lieshout hebben ons vanuit de elektronica afdeling voorzien van alles wat er tussen<br />

de millivolten en kilovolten kan gebeuren. In het kort, elektronica die werkt!<br />

Brengen scherven geluk? Bij het vermijden ervan hebben Bert Zwart, Eddy<br />

Inoeng en Michiel Groeneveld ons met hun vaardigheden in de glasbewerking zeer<br />

geholpen.<br />

Niets werkt meer zon<strong>der</strong> computers. Als u dit proefschrift leest, is dit het bewijs<br />

dat ze tot het laatste moment van mijn on<strong>der</strong>zoek ook daadwerkelijk werkten. Dit is<br />

te danken aan de inzet van Derk Bouhuijs en Marc Brugman van de systeemgroep.<br />

149


Ze worden niet eens boos als je ’s avonds nog met een probleem van “eigen schuld”<br />

komt opdagen. Nee, ze lossen het zelfs op ook! Eer<strong>der</strong> kregen we ook al veel hulp<br />

van Henk Pot, Jaap Berkhout, Paul Langemeier en Thijs Post.<br />

Apparatuur wil ook besteld worden, waarbij we vooral veel steun kregen van<br />

Irma Brouwer, Dick Jensen, Andries Porsius, Ineke Baaij, en Jenne Zon<strong>der</strong>van. In<br />

het bijzon<strong>der</strong> wil ik Mariet Bos bedanken, die geduldig achter elk probleem rond<br />

mijn status als promotiebursaal is aangegaan.<br />

Natuurlijk was er ook interactie met collega’s van buiten onze groep. In het<br />

bijzon<strong>der</strong> wil ik Kai Dieckmann, Igor Shvarchuk, Matthias Weidemüller en Martin<br />

Zielonkowski danken, met wie we menig idee en on<strong>der</strong>deel hebben geruild tijdens<br />

het opbouwen van onze experimenten. Ook de ontmoetingen tijdens het “quantumcollectief”<br />

met on<strong>der</strong> an<strong>der</strong>e Pepijn Pinkse, Allard Mosk, Peter Fedichev, Tom<br />

Hijmans, Pavel Bushev, Merrit Reynolds, Jook Walraven, Gora Shlyapnikov, Paul<br />

Tol, Norbert Herschbach, Wim Vassen en Wim Hogervorst, had ik niet willen<br />

missen. Net zo min als de gezellige sfeer rond de koffietafel met Lotty Gilliéron,<br />

Ton Raassen, Peter Uylings, Arnold Dönszelmann, Jørgen Hansen, Gilles Verbockhaven,<br />

Ton Schuitemaker, Charlie Al<strong>der</strong>hout en Ronald Winter. Voor mij waren<br />

de doorgenomen gesprekson<strong>der</strong>werpen hier ook een eerste verkenning van de Ne<strong>der</strong>landse<br />

samenleving. Er waren nog vele an<strong>der</strong>en, die het werken aan het Van <strong>der</strong><br />

Waals Zeeman Instituut zo leuk maakten, om met Erik-Paul, Heidi, Huib, Klaas,<br />

Allan, Marc, Eline, Gijs, Frank en Mischa maar enkele namen te noemen.<br />

Tot slot moest ook nog dit boek geschreven worden. Voor het geduldig en heel erg<br />

constructief becommentariëren van telkens weer nieuwe versies van het manuscript<br />

wil ik aan Robert en Ben en aan Jante Salverda mijn dank uitspreken. Robert<br />

leverde met een publicatie de grondslag van hoofstuk twee. De technische tekeningen<br />

zijn als vereenvoudigde versies afkomstig van de ontwerpen van Floris van <strong>der</strong><br />

Woude. Aan Bas Wolschrijn zijn de leuke plaatjes van inelastisch stuiterende atomen<br />

in hoofdstuk 7 te danken, en de leesbaarheid van de Ne<strong>der</strong>landse samenvatting komt<br />

door Jantes vertaalkunsten.<br />

Ook was er leven buiten de experimenteerkamer: Hier möchte ich vor allem<br />

Jante für ihre Geduld und Unterstützung danken, die sie stets für mich und meine<br />

unplanbaren Vorhaben, wie “Windsurfen gehen, wenn Wind ist”, aufbringt. Wenn<br />

Wind war, bekam ich tatkräftige Unterstützung von Arnd, bei <strong>der</strong> Suche nach den<br />

einzig wahren, kohärenten, holländischen Materiewellen.<br />

Meinen Eltern, Sieglinde und Edgar, gilt hier schließlich <strong>der</strong> größte Dank. Ohne<br />

sie hätte ich meinen Weg bis zur Promotion in Amsterdam so nicht gehen können.<br />

150


Curriculum Vitae<br />

Dirk Voigt<br />

geboren 23 Mei 1969 te Tübingen (Duitsland)<br />

In 1988 deed ik Abitur (eindexamen) aan het Thomas-Mann gymnasium te Lübeck.<br />

Na voldaan te hebben aan de Duitse dienstplicht begon ik in oktober 1989 de studie<br />

natuurkunde aan de Rheinisch-Westphälische Technische Hochschule (RWTH), de<br />

universiteit te Aken. In 1991 legde ik de examens van het Vordiplom af en begon<br />

ik met de Hauptstudium. Naast de verplichte vakken zoals theoretische en experimentele<br />

fysica volgde ik als keuzevakken lasertechniek, atoom- en molecuulfysica en<br />

vaste stoffysica. Hiernaast begeleidde ik als studentassistent studenten bij werkcolleges<br />

van het vak experimentele fysica.<br />

Voor de 12 maanden durende Diplomarbeit (stage) koos ik voor het vakgebied<br />

quantumoptica. In de groep van prof. dr. J. Mlynek aan de Universität Konstanz<br />

werkte ik mee aan atoom-optische experimenten met bundels van metastabiele helium<br />

atomen. In het bijzon<strong>der</strong> hield ik me bezig met een techniek van coherent<br />

aanslaan van deze atomen om een verstrengeld atoom−foton paar te verkrijgen.<br />

In Juni 1995 verkreeg ik aan de RWTH het diploma natuurkunde met het titel<br />

Diplom-Physiker.<br />

In November 1995 begon ik in het Van <strong>der</strong> Waals-Zeeman Instituut aan de<br />

Universiteit van Amsterdam als “promotiebursaal met bijbaan” on<strong>der</strong> leiding van<br />

prof. dr. H.B. van Linden van den Heuvell en dr. R.J.C. Spreeuw een promotieon<strong>der</strong>zoek<br />

aan spiegels van evanescent licht voor koude atomen. In het begin<br />

hield het experimentele gedeelte vooral het opzetten van diodelaser-, vacuüm- en<br />

computerapparatuur in. De resultaten van het on<strong>der</strong>zoek staan beschreven in dit<br />

proefschrift. De “bijbaan” omvatte het begeleiden van studenten gedurende het<br />

atoomfysica practicum en tijdens het verrichten van hun stage in onze groep.<br />

Tijdens mijn promotieon<strong>der</strong>zoek volgde ik in 1998 de Enrico Fermi zomerschool<br />

Bose-Einstein condensation in atomic gases te Varenna (Italië). Daarnaast bezocht<br />

ik een aantal conferenties met mondelinge presentaties of posterbijdragen: in 1999<br />

de 14 e internationale conferentie voor laserspectroscopie (ICOLS99) te Innsbruck,<br />

in 2000 de 17 e int. conf. voor atoomfysica (ICAP2000) te Florence, en de jaarlijkse<br />

verga<strong>der</strong>ingen van de sectie atoomfysica en quantumelektronica van de Ne<strong>der</strong>landse<br />

Natuurkundige Vereniging en van de sectie quantumoptica van de Deutsche<br />

Physikalische Gesellschaft.<br />

151


Publications<br />

• C.R. Ekstrom, C. Kurtsiefer, D. Voigt, O. Dross, T. Pfau, and J. Mlynek,<br />

Coherent excitation of a He ∗ beam observed in atomic momentum distributions,<br />

Opt. Comm. 123, 505 (1996).<br />

• C. Kurtsiefer, O. Dross, D. Voigt, C.R. Ekstrom, T. Pfau, and J. Mlynek,<br />

Observations of correlated atom-photon pairs on the single particle level,<br />

Phys.Rev.A55, R2539 (1997).<br />

• R.J.C. Spreeuw, D. Voigt, B.T. Wolschrijn, and H.B. van Linden van den<br />

Heuvell, Creating a low-dimensional quantum gas using dark states in an inelastic<br />

evanescent-wave mirror, Phys.Rev.A61, 053604 (2000).<br />

• D. Voigt, B.T. Wolschrijn, R. Jansen, N. Bhattacharya, R.J.C. Spreeuw, and<br />

H.B.vanLindenvandenHeuvell,Observation of radiation pressure exerted<br />

by evanescent waves, Phys.Rev.A61, 063412 (2000).<br />

• D. Voigt, E.C. Schil<strong>der</strong>, R.J.C. Spreeuw, and H.B. van Linden van den Heuvell,<br />

Characterisation of a high-power tapered semiconductor amplifier system,<br />

accepted <strong>for</strong> publication in Appl. Phys. B [preprint arXiv:physics/0004043].<br />

• D. Voigt, B.T. Wolschrijn, R.A. Cornelussen, R. Jansen, N. Bhattacharya,<br />

H.B.vanLindenvandenHeuvell,andR.J.C.Spreeuw,Elastic and inelastic<br />

evanescent-wave mirrors <strong>for</strong> cold atoms, to be published in Comptes Rendus<br />

de l’Académie des Sciences [preprint arXiv:physics/0011005].<br />

152

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!