arxiv:cond-mat/ v1 17 Apr 2000

Similar documents
BEC of 6 Li 2 molecules: Exploring the BEC-BCS crossover

Cold fermions, Feshbach resonance, and molecular condensates (II)

PHYSICAL REVIEW LETTERS

Achieving steady-state Bose-Einstein condensation

Confining ultracold atoms on a ring in reduced dimensions

Bose-Einstein condensation of lithium molecules and studies of a strongly interacting Fermi gas

From laser cooling to BEC First experiments of superfluid hydrodynamics

Collapsing Bose-Einstein condensates beyond the Gross-Pitaevskii approximation

A study of the BEC-BCS crossover region with Lithium 6

Lecture 3. Bose-Einstein condensation Ultracold molecules

arxiv:cond-mat/ v1 13 Mar 1998

Introduction to Cold Atoms and Bose-Einstein Condensation. Randy Hulet

From BEC to BCS. Molecular BECs and Fermionic Condensates of Cooper Pairs. Preseminar Extreme Matter Institute EMMI. and

Bose-Einstein Condensation Lesson

Bose-Einstein condensation; Quantum weirdness at the lowest temperature in the universe

Fermi Condensates ULTRACOLD QUANTUM GASES

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 24 Jul 2001

Ultracold molecules - a new frontier for quantum & chemical physics

Ultracold Fermi Gases with unbalanced spin populations

C. A. Sackett, J. M. Gerton, M. Welling, and R. G. Hulet 1. Physics Department, MS 61. Rice University. Houston, TX 77251

Evolution of a collapsing and exploding Bose-Einstein condensate in different trap symmetries

Evidence for Efimov Quantum states

Nature of spinor Bose-Einstein condensates in rubidium

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 17 Feb 2003

arxiv:cond-mat/ v1 28 Jan 2003

Fluids with dipolar coupling

The phases of matter familiar for us from everyday life are: solid, liquid, gas and plasma (e.f. flames of fire). There are, however, many other

Studies of Ultracold. Ytterbium and Lithium. Anders H. Hansen University of Washington Dept of Physics

arxiv:cond-mat/ v1 [cond-mat.quant-gas] 26 Nov 2002

Non-equilibrium Dynamics in Ultracold Fermionic and Bosonic Gases

Lecture 4. Feshbach resonances Ultracold molecules

arxiv:cond-mat/ v2 20 Jun 2003

arxiv:cond-mat/ v1 25 Nov 1997

Fundamentals and New Frontiers of Bose Einstein Condensation

arxiv:cond-mat/ v1 [cond-mat.other] 27 Mar 2007

Introduction to Bose-Einstein condensation 4. STRONGLY INTERACTING ATOMIC FERMI GASES

Harvard University Physics 284 Spring 2018 Strongly correlated systems in atomic and condensed matter physics

Fermi-Bose and Bose-Bose K-Rb Quantum Degenerate Mixtures

Precision Interferometry with a Bose-Einstein Condensate. Cass Sackett. Research Talk 17 October 2008

arxiv:quant-ph/ v2 5 Feb 2001

Experiments with an Ultracold Three-Component Fermi Gas

arxiv:physics/ v2 [physics.atom-ph] 5 Oct 1998

6P 3/2 6S 1/2. optical pumping. degenerate Raman transition. ıν=2> ıν=1> ıν=0> ıν=2> ıν=1> ıν=0> Zeeman energy. F=3, m F =3

Optimization of transfer of laser-cooled atom cloud to a quadrupole magnetic trap

Quantum Quantum Optics Optics VII, VII, Zakopane Zakopane, 11 June 09, 11

Multipath Interferometer on an AtomChip. Francesco Saverio Cataliotti

Revolution in Physics. What is the second quantum revolution? Think different from Particle-Wave Duality

Strongly correlated systems in atomic and condensed matter physics. Lecture notes for Physics 284 by Eugene Demler Harvard University

PROGRESS TOWARDS CONSTRUCTION OF A FERMIONIC ATOMIC CLOCK FOR NASA S DEEP SPACE NETWORK

Ultracold chromium atoms: From Feshbach resonances to a dipolar Bose-Einstein condensate

Thermodynamic Measurements in a Strongly Interacting Fermi Gas

PROGRESS TOWARDS CONSTRUCTION OF A FERMION ATOMIC CLOCK FOR NASA S DEEP SPACE NETWORK

condensate H.-J. Miesner, D.M. Stamper-Kurn, M.R. Andrews, D.S. Durfee, S. Inouye, and W. Ketterle

Quantum Gases. Subhadeep Gupta. UW REU Seminar, 11 July 2011

Superfluidity in interacting Fermi gases

Quantum superpositions and correlations in coupled atomic-molecular BECs

A Mixture of Bose and Fermi Superfluids. C. Salomon

Monte Carlo Simulation of Bose Einstein Condensation in Traps

A Mixture of Bose and Fermi Superfluids. C. Salomon

High-Temperature Superfluidity

Landau damping of transverse quadrupole oscillations of an elongated Bose-Einstein condensate

1. Cold Collision Basics

PHYS598 AQG Introduction to the course

Prospects for influencing scattering lengths with far-off-resonant light

Philipp T. Ernst, Sören Götze, Jannes Heinze, Jasper Krauser, Christoph Becker & Klaus Sengstock. Project within FerMix collaboration

BCS Pairing Dynamics. ShengQuan Zhou. Dec.10, 2006, Physics Department, University of Illinois

Bose-Einstein Condensate: A New state of matter

Introduction to cold atoms and Bose-Einstein condensation (II)

Optimization of evaporative cooling

Workshop on Coherent Phenomena in Disordered Optical Systems May 2014

What are we going to talk about: BEC and Nonlinear Atom Optics

Raman-Induced Oscillation Between an Atomic and Molecular Gas

BEC and superfluidity in ultracold Fermi gases

Ultracold scattering properties of the short-lived Rb isotopes

Cold atoms. 1: Bose-Einstein Condensation. Emil Lundh. April 13, Department of Physics Umeå University

Physics 598 ESM Term Paper Giant vortices in rapidly rotating Bose-Einstein condensates

Cold Metastable Neon Atoms Towards Degenerated Ne*- Ensembles

Ultracold atoms and molecules

Supported by NIST, the Packard Foundation, the NSF, ARO. Penn State

Anomalous Quantum Reflection of Bose-Einstein Condensates from a Silicon Surface: The Role of Dynamical Excitations

Week 13. PHY 402 Atomic and Molecular Physics Instructor: Sebastian Wüster, IISERBhopal, Frontiers of Modern AMO physics. 5.

Low-dimensional Bose gases Part 1: BEC and interactions

Reference for most of this talk:

arxiv:physics/ v2 27 Mar 2006

Dipolar Bose-Einstein condensates at finite temperature

Magnetism and Hund s Rule in an Optical Lattice with Cold Fermions

arxiv:cond-mat/ v1 [cond-mat.other] 7 Dec 2004

Observation of Feshbach resonances in ultracold

Vortices in Atomic Bose-Einstein Condensates in the large gas parameter region. Abstract

Laser Cooling and Trapping of Atoms

Quantum Mechanica. Peter van der Straten Universiteit Utrecht. Peter van der Straten (Atom Optics) Quantum Mechanica January 15, / 22

Realization of Bose-Einstein Condensation in dilute gases

Output coupling of a Bose-Einstein condensate formed in a TOP trap

The Phase of a Bose-Einstein Condensate by the Interference of Matter Waves. W. H. Kuan and T. F. Jiang

Hyperfine structure in photoassociative spectra of 6 Li 2 and 7 Li 2

arxiv:cond-mat/ v2 5 Apr 1999

Quantum correlations and atomic speckle

(Noise) correlations in optical lattices

F. Chevy Seattle May 2011

Roton Mode in Dipolar Bose-Einstein Condensates

Transcription:

Stable 85 Rb Bose-Einstein Condensates with Widely Tunable Interactions S. L. Cornish, N. R. Claussen, J. L. Roberts, E. A. Cornell and C. E. Wieman JILA, National Institute of Standards and Technology and the University of Colorado, and the arxiv:cond-mat/429v1 17 Apr 2 Department of Physics, University of Colorado, Boulder, Colorado 839-44 (February 4, 219) Abstract Bose-Einstein condensation has been achieved in a magnetically trapped sample of 85 Rb atoms. Long-lived condensates of up to 1 4 atoms have been produced by using a magnetic-field-induced Feshbach resonance to reverse the sign of the scattering length. This system provides many unique opportunities for the study of condensate physics. The variation of the scattering length near the resonance has been used to magnetically tune the condensate selfinteraction energy over a very wide range. This range extended from very strong repulsive to large attractive self-interactions. When the interactions were switched from repulsive to attractive, the condensate shrank to below our resolution limit, and after 5ms emitted a burst of high-energy atoms. Typeset using REVTEX 1

Atom-atom interactions have a profound influence on most of the properties of Bose- Einstein condensation (BEC) in dilute alkali gases. These interactions are well described in a mean-field model by a self-interaction energy that depends only on the density of the condensate (n) and the s-wave scattering length (a) [1]. Strong repulsive interactions produce stable condensates with a size and shape determined by the self-interaction energy. In contrast, attractive interactions (a < ) lead to a condensate state where the number of atoms is limited to a small critical value determined by the magnitude of a [2]. The scattering length also determines the formation rate, the spectrum of collective excitations, the evolution of the condensate phase, the coupling with the noncondensed atoms, and other important properties. In the vast majority of condensate experiments the scattering length has been fixed at the outset by the choice of atom. However, it was proposed that the scattering length could be controlled by utilizing the strong variation expected in the vicinity of a magneticfield-induced Feshbach resonance in collisions between cold ( µk) alkali atoms [3]. Recent experiments oncold 85 RbandCsatomsandNacondensates havedemonstratedthevariation of the scattering length via this approach[4 6]. However, extraordinarily high inelastic losses in the Na condensates were found to severely limit the extent to which the scattering length could be varied and precluded an investigation of the interesting negative scattering length regime [7]. These findings prompted the subsequent proposal of several exotic coherent loss processes that remain untested in other alkali species [8 1]. Here we report the successful use of a Feshbach resonance to readily vary the selfinteraction of long-lived condensates over a large range. In 85 Rb there exists a Feshbach resonance in collisions between two atoms in the F=2, m f = 2 hyperfine ground state at a magnetic field B 155 Gauss(G) [4,11]. Near this resonance the scattering length varies dispersively as a function of magnetic field and, in principle, can have any value between and + (see inset in Fig. 1). This has allowed us to reach novel regimes of condensate physics. These include producing very large repulsive interactions (n pk a 3 1 2 ), where effects beyond the mean-field approximation should be readily observable. We can 2

also make transitions between repulsive and attractive interactions (or vice-versa). This now makes it possible to study condensates in the negative scattering length regime, including the anticipated collapse of the condensate [12], with a level of control that has not been possible in other experiments [13]. In fact, this ability to change the sign of the scattering length is essential for the existence of our 85 Rb condensate. Away from the resonance the large negative scattering length ( 4a [4,14]) limits the maximum number of atoms in a condensate to 8 [2]. However, we have produced condensates with up to 1 4 atoms by operating in a region of the Feshbach resonance where a is positive. The experimental apparatus was similar to the double magneto-optical trap (MOT) system used in our earlier work [15]. Atoms collected in a vapor cell MOT were transferred to a second MOT in a low-pressure chamber. Once sufficient atoms have accumulated in the low-pressure MOT, both the MOT s were turned off and the atoms were loaded into a purely magnetic, Ioffe-Pritchard baseball trap. During the loading sequence the atom cloud was compressed, cooled and optically pumped, resulting in a typical trapped sample of about 3 1 8 85 Rb atoms in the F=2, m f = 2 state at a temperature of T = 45µK. The lifetime of atoms in the magnetic trap due to collisions with background atoms was 45s. Forced radio-frequency evaporation was employed to cool the sample of atoms. Unfortunately, 85 Rb is plagued with pitfalls for the unwary evaporator familiar only with 87 Rb or Na. In contrast to those atoms, the elastic collision cross section for 85 Rb exhibits strong temperature dependence due to a zero in the s-wave scattering cross section at a collision energy of E/k B 35µK [16]. This decrease in the elastic collision cross section with temperature means that the standard practice of adiabatic compression to increase the initial elastic collision rate does not work. 85 Rb also suffers from unusually high inelastic collision rates. We recently investigated these losses and observed a mixture of two and three-body processes which varied with B [17]. The overall inelastic collision rate displayed several orders of magnitude variation across the Feshbach resonance, with a dependence on B similar to that of the elastic collision rate. However, the inelastic rate increased more rapidly than the elastic rate towards the peak of the Feshbach resonance, and was found to be 3

significantly lower in the high field wing of the resonance than on the low field side. This knowledge of the loss rates, together with the known field dependence of the elastic cross section [4,14], has enabled us to successfully devise an evaporation path to reach BEC. This begins with evaporative cooling at a field (B = 25G) well above the Feshbach resonance. To maintain a relatively low density, and thereby minimize the inelastic losses, a relatively weak trap is used. The low initial elastic collision rate means that about 12s are needed to reach T 2µK, putting a stringent requirement on the trap lifetime. As the atoms cool, the elastic rate increases and it becomes advantageous to trade some of this increase for a reduced inelastic collision rate by moving to a smaller scattering length at B = 162.3G. The remainder of the evaporation is performed at this field (with a radial (axial) trap frequency of 17.5 Hz (6.8 Hz)). In contrast to field values away from the Feshbach resonance, the scattering length is positive at this field and stable condensates may therefore be produced. The density distribution of the trapped atom cloud was probed using absorption imaging with a 1µs laser pulse 1.6ms after the rapid (.2ms) turn-off of the magnetic trap. The shadowoftheatomcloudwasmagnifiedbyaboutafactorof1andimagedontoaccdarray to determine the spatial size and thenumber of atoms. The emergence of thebec transition was observed at T 15nK. Typically, we were able to produce pure condensates of up to 1 4 atoms with peak number densities of n pk 1 1 12 cm 3. The lifetime of the condensate at B = 162.3G was about 1s [18]. It is notable that our evaporation trajectory suffered a near-catastrophic decline prior to the observation of the BEC transition. We reached the required BEC phase space density with about 1 6 atoms, but then lost a factor of about 5 in the atom number before the characteristic two-component density distribution was visible. We suspect that as the BEC transition is reached with a high number density, the condensate acts as a hole through which atoms are lost due to inelastic collisions. This continues until the number density, and inelastic loss rates, have been reduced to the level where the condensate can stand out clearly from the thermal cloud. This point is obviously a delicate balance between the different collision processes in the cloud, all of which are strongly field dependent near the Feshbach resonance. Although we are able to decrease 4

the loss rate by moving to higher fields, the ratio of elastic to inelastic collisions actually decreases and it becomes harder to form condensates. For example, at B = 164.3G we can only produce condensates of a few thousand atoms. Conversely, moving to lower fields we are unable to utilize the higher elastic cross-section because the mean free path of an atom becomes smaller than the cloud size. This restriction together with the larger loss rate means that at B = 16.3G, for example, we are unable to form condensates. One of the features of the high inelastic loss rates reported in the Na experiments was an anomalously high decay rate when the condensate was swept rapidly through the Feshbach resonance [7]. In light of this work, it was essential to determine to what extent the 85 Rb condensate was perturbed in being swept across the Feshbach peak during the trap turnoff. These measurements also provide an additional test of the coherent loss mechanisms proposed in the literature [8 1]. We applied a linear ramp to the current in the baseball coil to sweep the magnetic field experienced by the atoms from B = 162.3G across the Feshbach peak to B 132G and then immediately turned off the trap and imaged the atom cloud. From the images we determined the fraction of condensate atoms lost as a function of the inverse ramp speed (Fig. 2). The loss for the fastest ramp, which corresponds to the direct turn-off of the magnetic trap, is less than 9%. This was determined in a separate experiment where the condensate was imaged directly in the magnetic trap both before and after the ramp. For comparison, the experiment was repeated using a cloud of thermal atoms much hotter than the BEC transition temperature. The results for the thermal atoms are consistent with the known inelastic loss rates in the vicinity of the Feshbach resonance [17]. The strong and poorly characterized temperature dependence of these known loss rates near the Feshbach peak makes it difficult to determine what fraction of the observed condensate loss can be attributed to the usual inelastic loss processes and we cannot, therefore, rule out a coherent aspect to the loss process. The solid line in Fig. 2 is the predicted loss of atoms due to the mechanism proposed by van Abeelen and Verhaar using the observed mean condensate density [9]. The predicted loss rate is about a factor of 5 lower and even at this level agreement is perhaps accidental: such loss models [9] based on the Timmermans theory 5

[8] of coupled atomic and molecular Gross-Pitaevskii equations are unlikely to be applicable to the present experiment [19]. We also changed the self-interaction energy by varying the magnetic field and observed the resulting change in the condensate shape. By applying a linear ramp to the magnetic field we have varied the magnitude of a in the condensate by almost three orders of magnitude. The duration of this ramp was sufficiently long (5ms) to ensure that the condensate responded adiabatically. Fig. 3 shows a series of condensate images for various magnetic fields such that a ranges from to very large and positive. They illustrate how we are able to smoothly move from the weakly interacting limit to deep into the strongly interacting or Thomas-Fermi (TF) regime. Moving towards the Feshbach peak the condensate size increases due to the increased self-interaction energy. The density distribution approaches the parabolic distribution with an aspect ratio of λ TF =ω z /ω r expected in the TF regime [2]. Moving in the opposite direction we approach the noninteracting regime where the condensate density distribution is a Gaussian whose dimensions are set by the harmonic oscillator lengths (a i = ( h/mω i ) 1/2 where i = r,z) [2]. From the images, the full widths at half-maximum (FWHM) of the column density distributions were determined. The scattering length was then derived by assuming a TF column density distribution with the same FWHM ( (Na) 1/5 ). In Fig. 1 we plot the scattering length versus the magnetic field. The ability to tune the atom-atom interactions in a condensate presents several exciting avenues for future research. For magnetic fields in the vicinity of the Feshbach peak the validity of the dilute-gas approximation comes under question. At a field of B = 156.6G we have already observed values of the gas parameter as high as n pk a 3 1 2. Such values are unprecedented in gaseous BEC. The lifetime of the condensate at this field is reduced to 11(15) ms due to the greater inelastic collision rate, but is still sufficient for many experiments. Under these conditions, effects beyond the mean field approximation, such as shifts in the frequencies of the collective excitations[21], should be readily observable. When we increased the magnetic field beyond B = 166.8G, where a was expected to change sign, a sudden departure from the smooth behavior in Figs. 1 and 3 was observed. 6

The condensate exhibited exotic dynamical behavior in response to the change in sign of the scattering length (Fig. 4) and there was an abrupt decrease in the condensate lifetime. These preliminary results on switching from repulsive to attractive interactions suggest a violent, but highly reproducible, destruction of the condensate. The ability to control the precise moment of the onset of a < instability is a distinct advantage over existing methods for studying this regime which rely on analyzing ensemble averages of post-collapse condensates [13]. The dynamical response of the condensate to a sudden change in the sign of the interactions can now be investigated in a controlled manner, probing the rich physics of the predicted condensate collapse process. We are pleased to acknowledge useful discussions with Murray Holland, Jim Burke, Josh Milstein and Marco Prevedelli. This research has been supported by the NSF and ONR. One of us (S. L. Cornish) acknowledges the support of a Lindemann Fellowship. 7

REFERENCES Quantum Physics Division, National Institute of Standards and Technology. [1] See, forexample, thereviewbyf.dalfovo, S.Giorgini, L.P. Pitaevskii, ands.stringari, Rev. Mod. Phys. 71, 463 (1999). [2] P. A. Ruprecht, M. J. Holland, K. Burnett, and M. Edwards, Phys. Rev. A 51, 474 (1995). [3] W. C. Stwalley, Phys. Rev. Lett. 37, 1628 (1976); E. Tiesinga, B. J. Verhaar, and H. T. C. Stoof, Phys. Rev. A 47, 4114 (1993); E. Tiesinga, A. Moerdijk, B. J. Verhaar, and H. T. C. Stoof, Phys. Rev. A 46, R1167 (1992). [4] J. L. Roberts et al., Phys. Rev. Lett. 81, 519 (1998). [5] V. Vuletić, A. J. Kerman, C. Chin and S. Chu, Phys. Rev. Lett. 82, 146 (1999). [6] S. Inouye et al., Nature 392, 151 (1998). [7] J. Stenger et al., Phys. Rev. Lett. 82, 2422 (1999). [8] E. Timmermans et al., Phys. Rep. 315, 199 (1999). [9] F. A. van Abeelen and B. J. Verhaar, Phys. Rev. Lett. 83, 155 (1999). [1] V. A. Yurovsky, A. Ben-Reuven, P. S. Julienne, and C. J. Williams, Phys. Rev. A 6, R765 (1999). [11] P. Courteille et al., Phys. Rev. Lett. 81, 69 (1998). [12] Y. Kagan, E. L. Surkov, and G. V. Shlyapnikov, Phys. Rev. Lett. 79, 264 (1997); C. A. Sackett, H. T. C. Stoof and R. G. Hulet Phys. Rev. Lett. 8, 231 (1998); M. Ueda and A. J. Leggett, Phys. Rev. Lett. 8, 1576 (1998); H. Saito and M. Ueda, e-print cond-mat/2393 [13] C. C. Bradley, C. A. Sackett, J. J. Tollett and R. G. Hulet, Phys. Rev. Lett. 75, 1687, 8

(1995); C. A. Sackett, J. M. Gerton, M. Welling, and R. G. Hulet, Phys. Rev. Lett. 82, 876 (1999). [14] J. P. Burke, Jr., NIST Gaithersburg, private communication. [15] C. J. Myatt et al., Opt. Lett. 21, 29 (1996). [16] J. P. Burke, Jr., J. L. Bohn, B. D. Esry, and C. H. Greene, Phys. Rev. Lett. 8, 297 (1998). [17] J. L. Roberts, N. R. Claussen, S. L. Cornish and C. E. Wieman, Magnetic Field Dependence of Ultracold Inelastic Collisions near a Feshbach Resonance, e-print physics/327. [18] This is in dramatic contrast to the very short lifetimes obtained with Na condensates anywhere near the vicinity of a Feshbach resonance [7]. We believe that the primary difference is that we have much lower density. [19] Timmermans, in effect, models the Feshbach resonance as an avoided crossing between atoms and molecules, with a density-dependent splitting given by α n, where α is a characteristic strength of the resonance [8]. Implicit in this model is the assumption that the splitting is small compared to the maximum mean-field energy allowed by the Gross- Pitaevskii equation assumption of diluteness (i.e. α n < ( h 2 /m)n 2/3 ). In contrast to the Na cases considered [7], this assumption is clearly not valid for our case; ramping toward the Feshbach resonance we reach a value of a for which n pk a 3 > 1 well before the model would predict atoms are converted to molecules. [2] G. Baym and C. J. Pethick, Phys. Rev. Lett. 76, 6 (1996). [21] L. Pitaevskii and S. Stringari, Phys. Rev. Lett. 81, 4541 (1998). 9

FIGURES FIG. 1. Scattering length in units of the Bohr radius (a ) as a function of the magnetic field. The data are derived from the condensate widths. The solid line illustrates the expected shape of the Feshbach resonance using a peak position and resonance width consistent with our previous measurements [4,17]. For reference, the shape of the full resonance has been included in the inset. FIG. 2. Fraction of atoms lost following a rapid sweep of the magnetic field through the peak of the Feshbach resonance as a function of the inverse speed of the field ramp. Data are shown for a condensate ( ) with a peak density of n pk = 1. 1 12 cm 3 and for a thermal cloud ( ) with a temperature T = 43nK and a peak density of n pk = 4.5 1 11 cm 3. The solid line is the predicted loss of atoms using the model of Ref. [9]. FIG. 3. (color) False color images and horizontal cross sections of the condensate column density for various magnetic fields. The condensate number was varied to maintain an optical depth (OD) of 1.5. The magnetic field values and corresponding scattering lengths are (a) B = 165.2G, a = 7a o (b) B = 162.3G, a = 29a o (c) B = 158.4G, a = 14a o (d) B = 157.2G, a = 3a o (e) B = 156.4G, a = 8a o. FIG. 4. Absorption image of the condensate remnant acquired 33.3ms after the scattering length is made negative. The ridge of hot atoms, extending out along the trap axis, is visible only during a brief time span centered at one-half of a radial trap period after the initial implosion. These atoms were ejected as a burst from the main sample during the implosion, then refocussed onto the trap axis by the trapping potential. Experimental and theoretical analyses of these refocussed ejectile plumes may open an observational window into the violent conditions at the center of the implosion. The field of view is 75µm 32µm. 1

a / a 1 8 6 4 2 4-4 145 16 175 157 159 161 163 165 Magnetic Field (G)

1 Atom Loss (%) 75 5 25 1 1 1 1 Inverse Ramp Speed (µs/g)

OD 1 (a) OD 1 (b) OD 1 (c) OD 1 (d) OD 1 (e) 5 1 Horizontal Position (µm)