Direct observation of effective ferromagnetic domains of cold atoms in a shaken optical lattice

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

We can then linearize the Heisenberg equation for in the small quantity obtaining a set of linear coupled equations for and :

In Situ Imaging of Cold Atomic Gases

Optical Lattices. Chapter Polarization

1. SYSTEM GEOMETRY AND PREPARATION FIG. S1.

5. Gross-Pitaevskii theory

Non-Equilibrium Physics with Quantum Gases

Magnetism, Rotons, and Beyond: Engineering atomic systems with lattice shaking.

A Mixture of Bose and Fermi Superfluids. C. Salomon

Lattice modulation experiments with fermions in optical lattices and more

Coherent manipulation of atomic wavefunctions in an optical lattice. V. V. Ivanov & A. Alberti, M. Schioppo, G. Ferrari and G. M.

B 2 P 2, which implies that g B should be

Confining ultracold atoms on a ring in reduced dimensions

Drag force and superfluidity in the supersolid striped phase of a spin-orbit-coupled Bose gas

Experimental Reconstruction of the Berry Curvature in a Floquet Bloch Band

Luigi Paolasini

J10M.1 - Rod on a Rail (M93M.2)

Cooperative Phenomena

COPYRIGHTED MATERIAL. Index

SOLUTIONS for Homework #2. 1. The given wave function can be normalized to the total probability equal to 1, ψ(x) = Ne λ x.

Roton Mode in Dipolar Bose-Einstein Condensates

Conference on Research Frontiers in Ultra-Cold Atoms. 4-8 May Generation of a synthetic vector potential in ultracold neutral Rubidium

Quantum Theory of Matter

Spin-injection Spectroscopy of a Spin-orbit coupled Fermi Gas

Landau-Fermi liquid theory

5.74 Introductory Quantum Mechanics II

October Entrance Examination: Condensed Matter Multiple choice quizzes

Supplementary Materials for

Supplementary information for Quantum delayed-choice experiment with a beam splitter in a quantum superposition

1 Supplementary Figure

Ultracold Fermi and Bose Gases and Spinless Bose Charged Sound Particles

Electrons in a periodic potential

Laser Cooling and Trapping of Atoms

But what happens when free (i.e. unbound) charged particles experience a magnetic field which influences orbital motion? e.g. electrons in a metal.

Quantum Condensed Matter Physics Lecture 9

Design and realization of exotic quantum phases in atomic gases

BEC in one dimension

Saturation Absorption Spectroscopy of Rubidium Atom

The non-interacting Bose gas

Physics with Neutrons I, WS 2015/2016. Lecture 11, MLZ is a cooperation between:

Chapter 2 Experimental Realization of One-Dimensional Bose Gases

MRSEC. Inflation and coherent dynamics in a Bose-Einstein condensate driven across a quantum critical point. Cheng Chin. Funding:

All-optical generation of surface plasmons in graphene

Artificial Gauge Fields for Neutral Atoms

2013 CAP Prize Examination

Collective behavior, from particles to fields

Floquet Topological Insulator:

Supplementary Information: Electrically Driven Single Electron Spin Resonance in a Slanting Zeeman Field

Supplementary Material for Dynamic optical lattices of sub-wavelength spacing for ultracold atoms

Likewise, any operator, including the most generic Hamiltonian, can be written in this basis as H11 H

UNIVERSITY OF SOUTHAMPTON

Quantum Quenches in Extended Systems

Valley Zeeman effect in elementary optical excitations of monolayerwse 2

Matter-Radiation Interaction

Collective Effects. Equilibrium and Nonequilibrium Physics

J07M.1 - Ball on a Turntable

NANOSCALE SCIENCE & TECHNOLOGY

Spins and spin-orbit coupling in semiconductors, metals, and nanostructures

4.2 Elastic and inelastic neutron scattering

Aditi Mitra New York University

9 Atomic Coherence in Three-Level Atoms

EE 223 Applied Quantum Mechanics 2 Winter 2016

Effects of spin-orbit coupling on the BKT transition and the vortexantivortex structure in 2D Fermi Gases

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

Vortices and superfluidity

Γ43 γ. Pump Γ31 Γ32 Γ42 Γ41

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

PRINCIPLES OF NONLINEAR OPTICAL SPECTROSCOPY

6. Interference of BECs

Bose-Bose mixtures in confined dimensions

Numerical Simulations of Faraday Waves in Binary Bose-Einstein Condensates

Interaction between atoms

SUPPLEMENTARY INFORMATION

Lattice Vibrations. Chris J. Pickard. ω (cm -1 ) 200 W L Γ X W K K W

Notes on x-ray scattering - M. Le Tacon, B. Keimer (06/2015)

Spinor Bose gases lecture outline

Superfluid 3 He. Miguel A. Morales

Lecture 4: Basic elements of band theory

ĝ r = {R v} r = R r + v.

Quantum optics of many-body systems

The Quantum Heisenberg Ferromagnet

Ginzburg-Landau Theory of Phase Transitions

Neutron interferometry. Hofer Joachim

The 4th Windsor Summer School on Condensed Matter Theory Quantum Transport and Dynamics in Nanostructures Great Park, Windsor, UK, August 6-18, 2007

SUPPLEMENTARY INFORMATION

Experimental reconstruction of the Berry curvature in a topological Bloch band

American Institute of Physics 319

Experimental Determination of Crystal Structure

Supplementary Materials for

Quality Factor Thickness (nm) Quality Factor Thickness (nm) Quality Factor 10

Spontaneous Symmetry Breaking in Bose-Einstein Condensates

Correlated 2D Electron Aspects of the Quantum Hall Effect

A. Time dependence from separation of timescales

Identical Particles. Bosons and Fermions

Supplementary Information: Enhancement of the tractor-beam pulling force on an optically bound structure

Neutrons on a surface of liquid helium.

Chapter4: Quantum Optical Control

arxiv: v1 [cond-mat.quant-gas] 30 Oct 2015

Measuring atomic NOON-states and using them to make precision measurements

Controlling Spin Exchange Interactions of Ultracold Atoms in Optical Lattices

Transcription:

DOI: 1.138/NPHYS2789 Direct observation of effective ferromagnetic domains of cold atoms in a shaken optical lattice Colin V. Parker, 1 Li-Chung Ha, 1 and Cheng Chin 1, 2 1 The James Franck Institute and Department of Physics, University of Chicago, Chicago, IL 6637, USA 2 Enrico Fermi Institute, University of Chicago, Chicago, IL 6637, USA (Dated: October 1, 213) NATURE PHYSICS www.nature.com/naturephysics 1

DOI: 1.138/NPHYS2789 PHASE-MODULATING THE LATTICE Our optical setup is shown in Fig. S1. The lattice shaking is accomplished by sinusoidally varying the phase of the retro-reflected beam. The beam passes through two acousto-optic modulators (AOMs) which, combined, leave the phase unaffected, whereupon it is focused by a mirror, then reflected to pass through the same AOMs again, for a total of four AOM passes[1]. When the AOM frequency is changed, the relative phase of the acoustic waves in the two AOMs changes, leading to variable total phase change on reflection. The sinusoidal modulation of the phase is then done by frequency modulation of the AOM signal. We calibrate the phase modulation by measuring the momentum kick acquired by a step function lattice displacement. A null result occurs for steps which are multiples of 2π, allowing us to calibrate in terms of the lattice spacing. FIG. S1: Lattice phase modulation setup The lattice beam passes through two AOMs operating at the same frequency (from the same source), passing each AOM twice. The total frequency shift is zero so the return beam can make a lattice with the incoming beam. The lens is positioned so that the AOM frequency can change while maintaining the retroreflection condition. When the AOM frequency changes, a phase shift develops in the beam path between the two AOMs leading to a change of the optical lattice phase. NUMERICAL CALCULATIONS OF THE HYBRIDIZED BAND The single-particle physics of an atom of mass m in a (1D) time-dependent optical lattice formed by retroreflected light of wavelength λ L is governed by the Hamiltonian 2 H(t) = 2 2m x + U 2 sin 2 2π [x x (t)], λ L (S1) 2 NATURE PHYSICS www.nature.com/naturephysics

DOI: 1.138/NPHYS2789 SUPPLEMENTARY INFORMATION where λ L /2 is the lattice constant, U is the lattice depth, and x (t) is the time-dependent lattice offset. For periodic lattice offsets x (t + τ) =x (t), the Hamiltonian is characterized by temporally and spatially periodic Floquet states. In analogy with spatially periodic Bloch states, which define momentum only up to an overall lattice momentum, the Floquet states define energy only up to an overall energy E = h/τ, corresponding to the absorption or emission of one quantum of energy at the shaking frequency. To obtain numerical dispersions for this Hamiltonian, we consider only the lowest 21 momentum states around the desired momentum k, that is, k 2k L,k 18k L,...,k,...,k+18k L,k+2k L, with k L =2π/λ L. This is more than enough states to describe the lowest band, and similar results are obtained using fewer states. We first compute the time average band structure by diagonalizing the timeaveraged potential H (k) = H(k, t) for each momentum and express the time-dependent Hamiltonian as H(k, t) =H (k)+h (k, t). We then break this Hamiltonian into 1 discrete time steps and compute the Trotter product of the time evolution operator (with = 1), U(k, τ) = τ e ih(k,t) N ( ) e i th (k) (1 + i th (k, n t)). (S2) n= Finally we numerically diagonalize U(k, t) to obtain the eigenstates and energies (up a factor τ). Figure S2 shows the dispersion of the lowest Floquet band with for several shaking amplitudes, and barrier height between the minima of the dispersion. In general, the Floquet bands k, t n will be periodic time-dependent superpositions of many of the unshaken bands k n (i.e. the bands obtained with no shaking). However, in our case the band spacing is sufficiently anharmonic that only the lowest two bands are near-resonant for k near k, and modulation is sufficiently weak that higher order terms are negligible. Therefore, the lowest band k, t is effectively a superposition of the lowest bare states k and k 1, that is, k, t = α(t) k + β(t) k 1, where α(t) and β(t) are time-dependent amplitudes. Under our conditions with x = 65 nm, α.95, and β.25. The relative phase evolves by 2π over the course of a shaking cycle. EFFECTIVE PARAMAGNETIC AND FERROMAGNETIC PHASES The 1D optical lattice in the z direction has a lattice depth of U =7E R, the 133 Cs atoms at scattering length of 1.4 1.9 nm in the lattice are well in the superfluid regime. The Hamiltonian of the system is given by NATURE PHYSICS www.nature.com/naturephysics 3

DOI: 1.138/NPHYS2789.4 a 3 E (E R ).2 2 1 E/k (nk) B barrier height (E R ).2-1 1 k ( k ).1 b 2 4 6 8 x (nm) L 1 5 E/k (nk) B FIG. S2: Dispersion calculation a Calculated dispersion of the lowest Floquet band with different amounts of shaking (listed as peak-to-peak, from bottom to top): no shaking, 21 nm, 43 nm, 64 nm, and 85 nm. The sharp features correspond to weak coupling with bands above the lowest two. b Height of the barrier between the two wells as a function of the shaking amplitude. Most of the experiments are performed with shaking amplitude x = 65 nm which yields a barrier height of.11 E R or 7 nk. H = p ɛ p a pa p + g 2 a p 1 +pa p 2 pa p1 a p2, p 1,p 2,p (S3) where a p is the creation operator of a particle with momentum p = (p x,p y,p z ), a p is the annihilation operator of a particle with momentum p, g =4π 2 a/mv is the coupling constant, V is volume of the system, ɛ p = p2 x +p2 y 2m the energy dispersion along the lattice direction. + ɛ(p z) is the kinetic energy, and ɛ(p z ) is In the zero or weak modulation regime, the lattice dispersion has one minimum ɛ = 4 NATURE PHYSICS www.nature.com/naturephysics

DOI: 1.138/NPHYS2789 SUPPLEMENTARY INFORMATION p 2 z/2m for low quasi-momenta p z, where the effective mass m m is given by the positive curvature of the ground band at p z = where the condensate resides. In the presence of a weak effective magnetic field b = v, where v is the relative velocity of the atoms to the lattice, the energy minimum is shifted to p z = m v. The effective magnetization defined as <p z >= m v shows a paramagnetic susceptibility, given by the effective mass m. At the critical modulation depth, the band curvature vanishes and the susceptibility diverges. Near this modulation depth, the dispersion can be approximated by ɛ = αp 2 z + βp 4 z, where the parameter α changes sign at the critical modulation depth, and β >. This transition is analogous to the paramagnetic to ferromagnetic transition in Landau theory. Beyond the critical modulation depth the lowest band develops two minima and at sufficiently low temperatures populations in one or both of the minima p = ±(,, k ) dominate. The two states are admixtures of the same quasi-momentum states in the ground and the first excited bands. Mixing with higher bands are far off-resonant and thus negligible. Introducing =,, k and =,, k and neglecting small populations in other momentum states, we perform the sum in Eq. (S3) and rewrite the Hamiltonian as H = σ=, ɛ σ N σ + g 2 N 2 + g 2 N 2 +2gN N, (S4) where N σ = a σa σ is the population in the σ state and ɛ σ is the associated kinetic energy. In the presence of a weak effective field b = v, the difference in kinetic energy is E = ɛ ɛ =2 k v and the effective magnetization <p z >= k for E > and <p z >= k for E <. The large and macroscopic magnetization in the presence of very small external field suggests a ferromagnetic interaction between the two pseudo-spin states. Equation (S4) and the effective spin Hamiltonian in Eq. (1) of the main text further suggest the ferromagnetic interaction is strong, on the same order as the chemical potential and is much larger than in regular spinor condensates. In the case of a sample with spatial inhomogeneity (i.e. domains), the kinetic energy term can become significant. To include the spatial dependence, we can expand the dispersion about the minima ±k to obtain an effective mass m o in the lattice direction and m in the other directions. Using the two-component wavefunction Ψ =(ψ,ψ ) and Eq. (S4), we obtain the Hamiltonian NATURE PHYSICS www.nature.com/naturephysics 5

DOI: 1.138/NPHYS2789 H = 2 2m xyψ xy Ψ+ 2 2m zψ o z Ψ+U(x,y,z)Ψ Ψ+ 3g 4 Ψ Ψ 2 g ( Ψ σ z Ψ ) 2,(S5) 4 where σ z is a Pauli matrix, and U(x,y,z) is the confining potential of the 3D trap. This Hamiltonian suggests that excitations in all directions can be generated by the spin interaction. Initial velocity and effective field To induce an effective energy imbalance we use controlled initial velocity changes by changing the relative beam powers of the trapping beams, which moves the trap center slightly and alters the initial velocity. The initial velocity changes are quantified by observing the change in momentum of a cloud with no lattice shaking during a series of control experiments taken between successive data shots. Over times much shorter than the trapping frequency, the effective energy imbalance of a given initial velocity can be quantified by performing a Galilean transformation to the atomic reference frame. This modifies the dispersion relation E(k) E(k) kv, where v is the initial velocity. For small v this will split the spin states by ± k v, where ±k corresponds to the positions of the minima. Over longer periods this approximation breaks down, but it remains true that initial velocity breaks the symmetry between the two states and transformation provides an approximate energy scale. In addition to the controlled velocity, there is an uncontrolled component of the initial velocity which changes on long timescales ( 5-1 shots), such that several successive shots can be acquired under similar conditions. In order to measure the initial velocity sensitivity as accurately as possible, we have performed several scans under each set of conditions ( 4) and adjusted the 5% point of each scan to zero, essentially re-calibrating the total initial velocity every few shots to factor in the slow variations in laboratory conditions. By doing this, we estimate and correct for the slowly varying shifts of the initial velocity, as the data points near the transition midpoint are acquired within a time window shorter than the slow variations. To the extent that the midpoint of the transition is characterized by random statistics of the whole sample, this re-calibration could tend to bias us toward a sharper transition by a small amount proportional to the step size of the bias between shots ( 1 nk), because the calibration can reduce some of the true uncertainty about the transi- 6 NATURE PHYSICS www.nature.com/naturephysics

DOI: 1.138/NPHYS2789 SUPPLEMENTARY INFORMATION tion midpoint. However, because the midpoint of the transition involves domain formation, the re-calibration will have a much smaller effect on the transition width, as the region near the midpoint is uniquely identified by simultaneous population of both spin states. To determine the effective temperature, we assume a Boltzmann distribution, where the average momentum M in terms of initial velocity v would be ( ) k M(v) = k v tanh. (S6) k B T eff DOMAIN RECONSTRUCTION 1 left moving right moving scaled density (a.u.) 1 1 1 9 18 exp Bragg peak FIG. S3: Bragg peak distribution Calculated momentum weights at several values of the applied shaking phase, along with the experimentally determined momentum weights. The zero of phase corresponds to maximum rightward velocity. In images taken with 5 ms TOF the domain structure is visible in the center of the image, as portions of the cloud move in opposite directions. On either side of the main cloud are NATURE PHYSICS www.nature.com/naturephysics 7

DOI: 1.138/NPHYS2789 higher momentum Bragg peaks induced by the lattice, which has been abruptly turned off along during the expansion. True in situ images cannot distinguish the domains, while for longer TOF the shape of the domains is distorted too much during the expansion. To leading order atoms in both spin states respond to the lattice shaking with the same phase, so the physical motion (in the lab frame) for either must be the same. If the atoms are released when traveling to the right, the spin-up atoms can have mostly momentum k, with only small amounts of higher Bragg peaks at k ±4π/λ L. By contrast, spin-down atoms must have a large fraction at k +4π/λ L in order to be physically moving right. Thus, the relative strength of the center and side two Bragg peaks forms a signature that identifies the spin state of each section of the image. Figure S3 shows a comparison of the theoretically computed Bragg peak weights with the experimentally determined values. Again we note that the computations include only single particle physics with no confinement. We define a threedimensional vector which is proportional to the density at each Bragg peak, ω =(ω 1,ω,ω 1 ) (Fig. 3 of the letter). The vector is defined pixel-wise and is normalized to sum to one. Using images which can be identified clearly as fully spin-up or spin-down, we determine typical vectors ω = (.43,.48,.9) and ω = (.2,.71,.9). Then when analyzing multi-domain images, we compute the population fraction of different components in each pixel by projecting the vector along the ω ω axis. A histogram of the resulting projection, which we denote as W, is shown in Fig. 3c of the letter. We associate the peaks of this histogram with spin-up and spin-down, and values in between with fractional population according to the value of W. Values of W beyond the histogram peaks are assumed to be fully magnetized. The vectors ω as well as the histogram peaks depend on the phase of the lattice modulation at the TOF release time, and we calibrate both under different experimental conditions. After determining the pixel-wise spin density of a 5 ms TOF image, we shift each image back by the expected travel distance during the TOF (left or right, depending on spin) and reconstruct the number and magnetization densities at the time of release. In a typical case near the center of our cloud, the experimental uncertainty (from all sources) is 4% of the total density. After projecting onto a unit vector on the ω ω axis, we scale up by a factor of 4 to get the fraction of spin up. Our magnetization fidelity is given by the resulting uncertainty in spin fraction of 16%. The resolution of magnetization density is set by the depth-of-focus-limited imaging resolution (1.5 µm), and the expansion and distortion of the domains during TOF. The latter may be estimated by 8 NATURE PHYSICS www.nature.com/naturephysics

DOI: 1.138/NPHYS2789 SUPPLEMENTARY INFORMATION R = µ/mr for short TOFs, where R is the radius of the condensate. For 133 Cs in 5 ms this gives R =1.5 µm, small compared to our cloud size of 1 µm by2µm. CORRELATION ANALYSIS In the correlation analysis, we distinguish multi-domain samples as those with a nearly equal amount of population in different states. We choose those where the sum of magnetization divided by the sum of density is smaller than.275, where.5 is the theoretical value for a fully polarized sample. This corresponds roughly to having less than 77.5% of the sample in the same state. For the analysis of fully magnetized sample, we select images with this ratio larger than.35, or more than 8.5% in the same state. Although this threshold may seem low, our determination of the magnetization has associated error, and empirically these numbers correctly distinguish the single- and multi-domain cases. QUENCH DYNAMICS To fully investigate the emergence of domains from a single mode condensate, we measure the spatial and momentum distribution of the atoms after a sudden (5 ms) quench across the ferromagnetic transition. Figure S4a shows images at various hold times following the quench and for different TOF, revealing that immediately following the quench the atoms have not yet moved appreciably from their original momentum distribution, and are therefore in unstable equilibrium at zero momentum. Over the course of about 1 ms, the atoms displace from this maximum into the minima on either side in a complex and dissipative manner, eventually completely depopulating the zero momentum state, see Fig. S4a. Note that this occurs faster than the in-plane trap oscillation periods of 5 and 12 ms. The dissipative dynamics indicate that energy must flow into other degrees of freedom, for example the kinetic energy in the transverse (non-lattice) directions. Observation of fast mixing between the spin and motional degrees of freedom demonstrates that our spin-spin interactions are strong and will drive the system towards equilibrium on short timescales. Given the quantum nature of our magnetic domains, which are characterized by complex order parameters e ik x Ψ(x) and e ik x Ψ(x), where Ψ(x) is the bosonic field operator, we expect spatial interference if they were made to overlap. We do indeed see interference NATURE PHYSICS www.nature.com/naturephysics 9

DOI: 1.138/NPHYS2789 FIG. S4: Quench dynamics and magnetic domain interference a Single shot images taken in situ and with 5 ms and 3 ms TOF, at several hold times following a 5 ms quench into the ferromagnetic state. b Spatial power spectrum along the lattice direction from images with 5 ms TOF, averaged over 2 shots. A peak appears at k =.27k L for the first 1 ms. The shaking amplitude was x = 65 nm and the scattering length was 1.9 nm. at intermediate times hold times for in situ and 5 ms TOF images. Figure S4b shows the fast Fourier transform of the atomic density averaged over multiple 5 ms TOF images, showing a peak at wavevector.27k L =.9k, where k L =2π/λ L =2π/164 nm is the lattice momentum. This signal, at half the expected wavevector for interference between the two domains, is consistent instead with interference between either domain and the remnant population at zero momentum. The interference grows in strength as the hold time increases and the system relaxes from the quench, reaching a peak at 1 ms. This supports 1 NATURE PHYSICS www.nature.com/naturephysics

DOI: 1.138/NPHYS2789 SUPPLEMENTARY INFORMATION our interpretation, as at longer times the system nears equilibrium, domains have formed and there is no remnant population at zero momentum. Because our three dimensional condensate is thicker than the depth of focus of the imaging system, we lack the resolution to detect interference at 2k. We also note that the interference is weaker for in situ images compared with those taken at 5 ms TOF. This suggests that as the condensate begins to relax toward the two minima, it has already begun to break up in real space to reduce density corrugation. With a time-of-flight image, the domains pass over one and other, allowing us to visualize the quantum inference more clearly. [1] Zhang, X. Observation of Quantum Criticality with Ultracold Atoms in Optical Lattices. Ph.D. thesis, University of Chicago (212). NATURE PHYSICS www.nature.com/naturephysics 11