Skip to main content
NIST Author Manuscripts logoLink to NIST Author Manuscripts
. Author manuscript; available in PMC: 2024 Oct 21.
Published in final edited form as: Nature. 2011 Mar 3;471(7336):83–86. doi: 10.1038/nature09887

Spin–orbit-coupled Bose–Einstein condensates

Y-J Lin 1, K Jiménez-García 1,2, I B Spielman 1
PMCID: PMC11493149  NIHMSID: NIHMS1526091  PMID: 21368828

Abstract

Spin–orbit (SO) coupling—the interaction between a quantum particle’s spin and its momentum—is ubiquitous in physical systems. In condensed matter systems, SO coupling is crucial for the spin-Hall effect1,2 and topological insulators35; it contributes to the electronic properties of materials such as GaAs, and is important for spintronic devices6. Quantum many-body systems of ultracold atoms can be precisely controlled experimentally, and would therefore seem to provide an ideal platform on which to study SO coupling. Although an atom’s intrinsic SO coupling affects its electronic structure, it does not lead to coupling between the spin and the centre-of-mass motion of the atom. Here, we engineer SO coupling (with equal Rashba7 and Dresselhaus8 strengths) in a neutral atomic Bose–Einstein condensate by dressing two atomic spin states with a pair of lasers9. Such coupling has not been realized previously for ultracold atomic gases, or indeed any bosonic system. Furthermore, in the presence of the laser coupling, the interactions between the two dressed atomic spin states are modified, driving a quantum phase transition from a spatially spin-mixed state (lasers off) to a phase-separated state (above a critical laser intensity). We develop a many-body theory that provides quantitative agreement with the observed location of the transition. The engineered SO coupling—equally applicable for bosons and fermions—sets the stage for the realization of topological insulators in fermionic neutral atom systems.


Quantum particles have an internal ‘spin’ angular momentum; this can be intrinsic for fundamental particles like electrons, or a combination of intrinsic (from nucleons and electrons) and orbital for composite particles like atoms. SO coupling links a particle’s spin to its motion, and generally occurs for particles moving in static electric fields, such as the nuclear field of an atom or the crystal field in a material. The coupling results from the Zeeman interaction μB between a particle’s magnetic moment μ, parallel to the spin σ, and a magnetic field B present in the frame moving with the particle. For example, Maxwell’s equations dictate that a static electric field E=E0z^ in the laboratory frame (at rest) gives a magnetic field BSO=E0(ħ/mc2)(ky,kx,0) in the frame of an object moving with momentum ħk=ħ(kx,ky,kz) where c is the speed of light in vacuum and m is the particle’s mass. The resulting momentum-dependent Zeeman interaction μBSO(k)σxkyσykx is known as the Rashba7 SO coupling. In combination with the Dresselhaus8coupling σxkyσykx, these describe two-dimensional SO coupling in solids to first order.

In materials, the SO coupling strengths are generally intrinsic properties, which are largely determined by the specific material and the details of its growth, and are thus only slightly adjustable in the laboratory. We demonstrate SO coupling in an 87Rb Bose–Einstein condensate (BEC) where a pair of Raman lasers create a momentum-sensitive coupling between two internal atomic states. This SO coupling is equivalent to that of an electronic system with equal contributions of Rashba and Dresselhaus9 couplings, and with a uniform magnetic field B in the y^z^ plane, which is described by the single-particle Hamiltonian:

H^=ħ2k^22m1ˇ[B+BSO(k^)]μ=ħ2k^22m1ˇ+Ω2σˇz+δ2σˇy+2αk^xσˇy (1)

α parametrizes the SO-coupling strength; Ω=gμBBz and δ=gμBBy result from the Zeeman fields along z^ and y^, respectively; and σˇx,y,z are the 2×2 Pauli matrices. Without SO coupling, electrons have group velocity, vx=ħkx/m independent of their spin. With SO coupling, their velocity becomes spin-dependent, vx=ħ(kx±2αm/ħ2)/m for spin | and | electrons (quantized along y^). In two recent experiments, this form of SO coupling was engineered in GaAs heterostructures where confinement into two-dimensional planes linearized the native cubic SO coupling of GaAs to produce a Dresselhaus term, and asymmetries in the confining potential gave rise to Rashba coupling. In one experiment a persistent spin helix was found6, and in another the SO coupling was only revealed by adding a Zeeman field10.

SO coupling for neutral atoms enables a range of exciting experiments, and importantly, it is essential in the realization of neutral atom topological insulators. Topological insulators are novel fermionic band insulators including integer quantum Hall states and now spin quantum Hall states that insulate in the bulk, but conduct in topologically protected quantized edge channels. The first-known topological insulators—integer quantum Hall states11—require large magnetic fields that explicitly break time-reversal symmetry. In a seminal paper3, Kane and Mele showed that in some cases SO coupling leads to zero-magnetic-field topological insulators that preserve time-reversal symmetry. In the absence of the bulk conductance that plagues current materials, cold atoms can potentially realize such an insulator in its most pristine form, perhaps revealing its quantized edge (in two dimensions) or surface (in three dimensions) states. To go beyond the form of SO coupling we created, almost any SO coupling, including that needed for topological insulators, is possible with additional lasers1214.

To create SO coupling, we select two internal ‘spin’ states from within the 87Rb 5S1/2, F=1 ground electronic manifold, and label them pseudo-spin-up and pseudo-spin-down in analogy with an electron’s two spin states: |=|F=1, mF=0 and |=F=1, mF=1. A pair of λ=804.1nm Raman lasers, intersecting at θ=90 and detuned by δ from Raman resonance (Fig. 1a), couple these states with strength Ω; here ħkL=2πħ/λ and EL=ħ2kL2/2m are the natural units of momentum and energy. In this configuration, the atomic Hamiltonian is given by equation (1), with kx replaced by a quasimomentum q and an overall EL energy offset. Ω and δ give rise to effective Zeeman fields along z^ and y^, respectively. The SO-coupling term 2ELqσˇy/kL results from the laser geometry, and α=EL/kL is set by λ and θ, independent of Ω (see Methods). In contrast with the electronic case, the atomic Hamiltonian couples bare atomic states |,κx=q+kL and |,κx=qkL with different velocities, ħκx/m=ħ(q±kL)/m.

Figure 1 |. Scheme for creating SO coupling.

Figure 1 |

a, Level diagram. Two λ804.1nm lasers (thick lines) coupled states |F=1,mF=0=| and |F=1,mF=1=|, differing in energy by a ħωZ Zeeman shift. The lasers, with frequency difference ΔωL/2π=(ωZ+δ/ħ)/2π, were detuned δ from the Raman resonance. |mF=0 and |mF=+1 had a ħ(ωZωq) energy difference; because ħωq=3.8EL is large, |mF=+1 can be neglected. b, Computed dispersion. Eigenenergies at δ=0 for Ω=0 (grey) to 5EL. When Ω<4EL the two minima correspond to the dressed spin states | and |. c, Measured minima. Quasimomentum q, of |, versus Ω at δ=0, corresponding to the minima of E(q). Each point is averaged over about ten experiments; the uncertainties are their standard deviation. d, Spin-momentum decomposition. Data for sudden laser turn-off: δ0,Ω=2EL (top image pair), and Ω=6EL (bottom image pair). For Ω=2EL,| consists of |,κx0 and |,κx2kL, and | consists of |,κx2kL and |,κx0.

The spectrum, a new energy-quasimomentum dispersion of the SO-coupled Hamiltonian, is displayed in Fig. 1b at δ=0 and for a range of couplings Ω. The dispersion is divided into upper and lower branches E±(q), and we focus on E(q). For Ω<4EL and small δ (see Fig. 2a), E(q) consists of a double well in quasi-momentum15, where the group velocity E(q)/ħq is zero. States near the two minima are dressed spin states, labelled as | and |. As Ω increases, the two dressed spin states merge into a single minimum and the simple picture of two dressed spins is inapplicable. Instead, that strong coupling limit effectively describes spinless bosons with a tunable dispersion relation16 with which we engineered synthetic electric17 and magnetic fields18 for neutral atoms.

Figure 2 |. Phases of a SO-coupled BEC.

Figure 2 |

a, b, Mean field phase diagrams for infinite homogeneous SO-coupled 87Rb BECs (1.5-kHz chemical potential). The background colours indicate atom fraction in | and |. Between the dashed lines there are two dressed spin states, | and |. a, Single-particle phase diagram in the Ωδ plane. b, Phase diagram (enlargement of the grey rectangle in a), as modified by interactions. The dots represent a metastable region where the fraction of atoms f, remains largely unchanged for th=3s. c, Miscible-to-immiscible transition. Phase line for mixtures of dressed spins and images after TOF (with populations NN, mapped from | and | showing the transition from phase-mixed to phase-separated within the ‘metastable window’ of detuning.

In the absence of Raman coupling, atoms with spins | and | spatially mixed perfectly in a BEC. By increasing Ω we observed an abrupt quantum phase transition to a new state where the two dressed spins spatially separated, resulting from a modified effective interaction between the dressed spins.

We studied SO coupling in oblate 87Rb BECs with about 1.8×105 atoms in a λ=1,064-nm crossed dipole trap with frequencies (fx,fy,fz)(50,50,140)Hz. The bias magnetic field B0y^ generated a ωZ/2π4.81MHz Zeeman shift between | and |. The Raman beams propagated along y^±x^ and had a constant frequency difference ΔωL/2π4.81MHz. The small detuning from the Raman resonance δ=ħ(ΔωLωZ) was set by B0, and the state |mF=+1 was decoupled owing to the quadratic Zeeman effect (see Methods).

We prepared BECs with an equal population of | and | at Ω, δ=0, then we adiabatically increased Ω to a final value up to 7EL in 70 ms, and finally we allowed the system to equilibrate for a holding time th=70ms. We abruptly (toff<1μs) turned off the Raman lasers and the dipole trap—thus projecting the dressed states onto their constituent bare spin and momentum states—and absorption-imaged them after a 30.1-ms time of flight (TOF). For Ω>4EL (Fig. 1d), the BEC was located at the single minimum q0 of E(q) with a single momentum component in each spin state corresponding to the pair {|,q0+kL,|,q0kL}. However, for Ω<4EL we observed two momentum components in each spin state, corresponding to the two minima of E(q) at q and q. The agreement between the data (symbols), and the expected minima locations (curves), demonstrates the existence of the SO coupling associated with the Raman dressing. We kept δ0 when turning on Ω by maintaining equal populations in bare spins |, | (see Fig. 1d).

We experimentally studied the low-temperature phases of these interacting SO-coupled bosons as a function of Ω and δ. The zero-temperature mean-field phase diagram (Fig. 2a, b) includes phases composed of a single dressed spin state, a spatial mixture of both dressed spin states, and coexisting but spatially phase-separated dressed spins.

This phase diagram can largely be understood as the result of non-interacting bosons condensing into the lowest-energy single particle state, and can be divided into three regimes (Fig. 2a). In the region of positive detuning marked |, there are double minima at q=q,q in E(q) with E(q)<E(q) and the bosons condense at q. In the region marked | the reverse holds. The energy difference between the two minima is Δ(Ω,δ)=E(q)E(q)δ for small δ (see Methods). In the third ‘single minimum’ regime, the atoms condense at the single minimum q0. These dressed spins act as free particles with group velocity ħKx/m (with an effective mass mm, for small Ω), where Kx=qq,,0 for the different minima.

We investigated the phase diagram using BECs with initially equal spin populations prepared as described previously, but with δ0 and th up to 3 s. We probed the atoms after abruptly removing the dipole trap, and then ramping Ω0 in 1.5 ms. This approximately mapped | and | back to their undressed counterparts | and | (see Methods). We absorption-imaged the atoms after a 30-ms TOF, during the last 20 ms of which a Stern-Gerlach magnetic field gradient along y^ separated the spin components.

Figure 3a shows the condensate fraction f=N/(N+N) in | at Ω=0.6EL as a function of δ, at th=0.1s, 1 s and 3 s, where N and N denote the number of condensed atoms in | and |, respectively. The BEC is all | for δ0 and all | for δ0, but both dressed spin populations substantially coexisted for detunings within ±wδ (obtained by fitting f to the error function where δ=±wδ corresponds to f=0.50±0.16). Figure 3b shows wδ versus Ω for hold times th. wδ decreases with th; even by our longest th of 3 s it has not reached equilibrium.

Figure 3 |. Population relaxation.

Figure 3 |

a, Condensate fraction f in | at Ω=0.6EI versus detuning δ at th=0.1, 0.5 and 3 s showing wδ decrease with increasing th. The solid curves are fits to the error function from which we obtained the width wδ. b, Metastable detuning width. Width wδ versus Ω at th=0.1, 0.5 and 3 s; the data fits well to a[b+(Ω/EL)2] (dashed curves).

Conventional F=1 spinor BECs have been studied in 23Na and 87Rb without Raman coupling1921. For our | and | states, the interaction energy depends on the local density in each spin state, and is described by:

H^I=12d3r[(c0+c22)(ρ^+ρ^)2+c22(ρ^2ρ^2)+(c2+c)ρ^ρ^]

where ρ^ and ρ^; are density operators for | and |, and normal ordering is implied. In the 87Rb F=1 manifold, the spin-independent interaction is c0=7.79×1012Hzcm3, the spin-dependent interaction22 is c2=3.61×1014Hzcm3, and c=0. Because |c0||c2|, the interaction is almost spin-independent, but c2<0, so the two-component mixture of | and | has a spatially mixed ground state (is miscible). When H^I is re-expressed in terms of the dressed spin states, cc0Ω2/(8EL2) is non-zero and corresponds to an effective interaction between | and |. This modifies the ground state of our SO-coupled BEC (mixtures of | and |) from phase-mixed to phase-separated above a critical Raman coupling strength Ωc. This transition lies outside the common single-mode approximation20.

The effective interaction between | and | is an exchange energy resulting from the non-orthogonal spin part of | and | (see Methods): a spatial mixture produces total density modulations15 with wavevector 2kL, in analogy with the spin-textures of the electronic case6. These increase the state-independent interaction energy in H^I wherever the two dressed spins spatially overlap, contributing to the c; term. (Such a term does not appear for radio-frequency-dressed states, which are always spin-orthogonal.) Because c; and c2 have opposite sign here, the dressed BEC can go from miscible to immiscible at the miscibility threshold19 for a two-component BEC c0+c2+c/2=c0(c0+c2), when Ω=Ωc (this result is in agreement with an independent theory presented in ref. 23).

Figure 2b depicts the mean field phase diagram including interactions, computed by minimizing the interaction energy HI plus the single particle detuning Δ(Ω,δ)δ. This phase diagram adds two new phases, mixed (hashed) and phase-separated (bold line), to those present in the non-interacting case. The c2(ρ^2ρ^2)/2 term in H^I implies that the energy difference between a | BEC and a | BEC is proportional to N2c2. The detuning required to compensate for this difference slightly displaces the symmetry point of the phase diagram downwards. As evidenced by the width of the metastable window 2wδ in Fig. 2b, for |δ|<wδ the spin-population does not have time to relax to equilibrium. The miscibility condition does not depend on atom number, so the phase line in Fig. 2c shows the system’s phases for |δ|<wδ: phase-mixed for Ω<Ωc and phase-separated for Ω>Ωc where Ωc8c2/c0EL0.19EL.

We measured the miscibility of the dressed spin components from their spatial profiles after TOF, for Ω=0 to 2EL and δ0 such that NTNT, where NT, is the total atom number including both the condensed and thermal components in |, |. For each TOF image, we numerically re-centred the Stern-Gerlach-separated spin distributions (Fig. 2c, and see Methods), giving condensate densities n(x,y) and n(x,y). Given that the self-similar expansion of BECs released from harmonic traps essentially magnifies the in situ spatial spin distribution, these reflect the in situ densities24.

A dimensionless metric s=1nn/(n2n2)12 quantifies the degree of phase separation (where is the spatial average over a single image). s=0 for any perfect mixture n(x,y)n(x,y), and s=1 for complete phase separation. Figure 4 displays s versus Raman coupling Ω with a hold time th=3s, showing that s0 for small Ω (as expected given our miscible bare spins) and s abruptly increases above a critical Ωc. The inset to Fig. 4 plots s as a function of time, showing that s reaches steady state in 0.14(3) s, which is much less than th. To obtain Ωc, we fitted the data in Fig. 4 to a slowly increasing function below Ωc and the power-law 1(Ω/Ωc)a above Ωc. The resulting Ωc=0.20(2)EL is in agreement with the mean field prediction Ωc=0.19EL. This demonstrates a quantum phase transition for a two-component SO-coupled BEC, from miscible when Ω<Ωc to immiscible when Ω>Ωc.

Figure 4 |. Miscible to immiscible phase transition.

Figure 4 |

Phase separation s versus Ω with th=3s; the solid curve is a fit to the function described in the text. The power-law component of the fit has an exponent a=0.75±0.07; this is not a critical exponent, but instead results from the decreasing size of the domain wall between the regions of | and | as Ω increases. Each point represents an average over 15 to 50 realizations and the uncertainties are the standard deviation. Inset, phase separation s versus th with Ω=0.6EL fitted to an exponential showing the rapid 0.14(3)-s timescale for phase separation.

Even below Ωc, s slowly increased with increasing Ω. To understand this effect, we numerically solved the two-dimensional spinor Gross-Pitaevskii equation in the presence of a trapping potential. This demonstrated that the differential interaction term c2(ρ^2ρ^2)/2 in H^I favours slightly different density profiles for each spin component, while the (c2+c)ρ^ρ^ term favours matched profiles. Thus, as c2+c approached zero from below this balancing effect decreased, causing s to increase.

An infinite system should fully phase separate (s=1) for all Ω>Ωc. In our finite system, the boundary between the phase-separated spins, set by the spin-healing length (ξs=ħ2/2m|c2+c|n, where n is the local density), can be comparable to the system size. We interpret the increase of s above Ωc as resulting from the decrease of ξs with increasing Ω.

We realized SO coupling in an 87Rb BEC, and observed a quantum phase transition from spatially mixed to spatially separated. By operating at lower magnetic field (with a smaller quadratic Zeeman shift), our method extends to the full F=1 or F=2 manifold of 87Rb or 23Na, enabling a new kind of tuning for spinor BECs, without the losses associated with Feshbach tuning25. Such modifications may allow access to the expected non-abelian vortices in some F=2 condensates26. Because our SO coupling is in the small Ω limit, this technique is practical for fermionic 40K, with its smaller fine-structure splitting and thus larger spontaneous emission rate27. When the Fermi energy lies in the gap between the lower and upper bands (for example, Fig. 1b) there will be a single Fermi surface; this situation can induce p-wave coupling between fermions28 and more recent work anticipates the appearance of Majorana fermions29.

METHODS SUMMARY

System preparation.

Our experiments began with nearly pure 87Rb BECs of approximately 1.8×105 atoms in the |F=1,mF=1 state30 confined in a crossed optical dipole trap. The trap consisted of a pair of l,064-nm laser beams propagating along x^y^(1/e2 radii of wx^+y^120μm and wz^50μm) and x^y^ (1/e2 radii of wx^y^wz^65μm).

We prepared equal mixtures of |F=1,mF=1 and |1,0 using an initially off-resonant radio-frequency magnetic field Brf(t)x^. We adiabatically ramped δ0 to in l5ms, decreased the radio-frequency coupling strength Ωrf to about 150 Hz, which is much less than ħωq, in 6 ms, and suddenly turned off Ωrf, projecting the BEC into an equal superposition of |mF=1 and |mF=0. We subsequently ramped δ to its desired value in 6 ms and then linearly increased the intensity of the Raman lasers from zero to the final coupling Ω in 70 ms.

Magnetic fields.

Three pairs of Helmholtz coils, orthogonally aligned along x^+y^, x^y^ and z^, provided bias fields (Bx+y, Bxy, and Bz). By monitoring the F=1,mF=1 and |1,0 populations in a nominally resonant radio-frequency dressed state, prepared as above, we observed a short-time (less than about 10 min) root-mean-square field stability gμBBRMS/h80Hz. The field drifted slowly on longer timescales (but changed abruptly when unwary colleagues entered through our laboratory’s ferromagnetic doors). We compensated for the drift by tracking the radio-frequency and Raman resonance conditions.

The small energy scales involved in the experiment meant that it was crucial to minimize magnetic field gradients. We detected stray gradients by monitoring the spatial distribution of |mF=1|mF=0 spin mixtures after TOF. Small magnetic field gradients caused this otherwise miscible mixture to phase-separate along the direction of the gradient. We cancelled the gradients in the x^y^ plane with two pairs of anti-Helmholtz coils, aligned along x^+y^ and x^y^, to gμBB/h0.7Hzμm1.

METHODS

System preparation.

Our experiments began with nearly pure 87Rb BECs of approximately 1.8×105 atoms in the |F=1,mF=1 state30 confined in a crossed optical dipole trap. The trap consisted of a pair of 1,064-nm laser beams propagating along x^y^ (1/e2 radii of wx^+y^120μm and wz^50μm) and x^y^ (1/e2 radii of wx^y^wz^65μm).

We prepared equal mixtures of |F=1,mF=1 and |1,0 using an initially off-resonant radio-frequency magnetic field Brf(t)x^. We adiabatically ramped δ to δ0 in 15 ms, decreased the radio-frequency coupling strength Ωrf to about 150 Hz, which is much less than ħωq, in 6 ms, and suddenly turned off Ωrf projecting the BEC into an equal superposition of |mF=1 and |mF=0. We subsequently ramped δ to its desired value in 6 ms and then linearly increased the intensity of the Raman lasers from zero to the final coupling Ω in 70 ms.

Magnetic fields.

Three pairs of Helmholtz coils, orthogonally aligned along x^+y^, x^y^ and z^, provided bias fields (Bx+y, Bxy, and Bz). By monitoring the F=1,mF=1 and |1,0 populations in a nominally resonant radio-frequency dressed state, prepared as above, we observed a short-time (less than about 10 min) root-mean-square field stability gμBBRMS/h80Hz. The field drifted slowly on longer timescales (but changed abruptly when unwary colleagues entered through our laboratory’s ferromagnetic doors). We compensated for the drift by tracking the radio-frequency and Raman resonance conditions.

The small energy scales involved in the experiment meant that it was crucial to minimize magnetic field gradients. We detected stray gradients by monitoring the spatial distribution of |mF=1|mF=0 spin mixtures after TOF. Small magnetic field gradients caused this otherwise miscible mixture to phase-separate along the direction of the gradient. We cancelled the gradients in the x^y^ plane with two pairs of anti-Helmholtz coils, aligned along x^+y^ and x^y^, to gμBB/h0.7Hzμm1.

SO-coupled Hamiltonian.

Our system30 consisted of a F=1 BEC with a bias magnetic field along y^ at the intersection of two Raman laser beams propagating along x^+y^ and x^+y^ with angular frequencies ωL and ωL+ΔωL, respectively. The rank-1 tensor light shift of these beams produced an effective Zeeman magnetic field along the z direction with Hamiltonian H^R=ΩRσˇ3,zcos(2kLx^+ΔωLt), where σˇ3,x,y,z are the 3×3 Pauli matrices and we define Iˇ3 as the 3×3 identity matrix. If we take y^ as the natural quantization axis (by expressing the Pauli matrices in a rotated basis σˇ3,yσˇ3,z, σˇ3,xσˇ3,y and σˇ3,zσˇ3,x) and make the rotating wave approximation, the Hamiltonian for spin states {|mF=+1,|0,|1} in the frame rotating at ΔωL is:

H^3=ħ2k^22m1ˇ3+(3δ/2+ħωq000δ/2000δ/2)+ΩR2σˇ3,xcos(2kLx^)ΩR2σˇ3,ysin(2kLx^) (2)

As we justify below, |mF=+1 can be neglected for large enough ħωq, which gives the effective two-level Hamiltonian:

H^2=ħ2k^22m1ˇ+δ2σˇz+Ω2σˇxcos(2kLx^)Ω2σˇysin(2kLx^)

for the pseudo-spins |=|mF=0 and |=|1 where Ω=ΩR/2. After a local pseudo-spin rotation by θ(x^)=2kLx^ about the pseudo-spin z^ axis followed by a global pseudo-spin rotation σˇzσˇy, σˇyσˇx and, σˇxσˇz the 2×2 Hamiltonian takes the SO-coupled form:

H^2=ħ2k^22m1ˇ+Ω2σˇz+δ2σˇy+2ħ2kLk^x2mσˇy+EL1ˇ

The SO term linear in k^x results from the non-commutation of the spatially dependent rotation about the pseudo-spin z axis and the kinetic energy.

Effective two-level system.

For atoms in |mF=1 and |mF=0 with velocities ħκx/m0 and Raman-coupled near resonance, δ0, the |mF=+1 state is detuned from resonance owing to the ħωq=3.8EL quadratic Zeeman shift. For δ/4EL1 and Ω<4EL, we have Δ(Ω,δ)δ[1(Ω/4EL)2]1/2.

Effect of the neglected state.

In our experiment, we focused on the two-level system formed by the |mF=1 and |mF=0 states. We verified the validity of this assumption by adiabatically eliminating the |mF=+1 state from the full three-level problem. To second-order in Ω, this procedure modifies the detuning δ and SO-coupling strength α in equation (1) by:

δ(2)=(Ω2)214EL+ħωq132Ω2EL
α(2)=(Ω2)2α(4EL+ħωq)2α256(ΩEL)2

In these expressions, we have retained only the largest term in a 1/ωq expansion. In our experiment, where ħωq=3.8EL, δ is substantially changed at our largest coupling Ω=7EL. To maintain the desired detuning δ in the simple two-level model (that is, Δδ+δ(2)=0 in Fig. 1c), we changed gμBB0 by as much as 3EL to compensate for δ(2). We did not correct for the change to α, which was always small.

Although both terms are small at the Ω=0.2EL transition from miscible to immiscible, slow drifts in B0 prompted us to locate Δ=0 empirically from the equal-population condition, NT=NT. As a result, δ in equation (1) implicitly includes the perturbative correction δ(2).

Origin of the effective interaction term.

The additional c term in the interaction Hamiltonian for dressed spins directly results from transforming into the basis of dressed spins, which are:

|,Kx|,κx=Kx+q+kLε|,κx=Kx+qkL

and

|,Kx|,κx=Kx+qkLε|,κx=Kx+q+kL (3)

where ħKx/m is the group velocity, Kx=qq for | and Kx=qq for |, and ε=Ω/8EL1. Thus, in second quantized notation, the dressed field operators transform according to:

ψ^(r)=ψ^(r)+εe2ikLxψ^(r)

and

ψ^(r)=ψ^(r)+εe2ikLxψ^(r)

where q14ε2kLkL and q14ε2kLkL. Inserting the transformed operators into:

H^I=12d3r[(c0+c22)(ρ^+ρ^)2+c22(ρ^2ρ^2)+c2ρ^ρ^]

gives the interaction Hamiltonian (with normal ordering implied) for dressed spins which can be understood order-by-order (both c2/c0 and ε are treated as small parameters). In this analysis, the terms proportional to c2 are unchanged to the order of c2/c0, and we only need to evaluate the transformation of the spin-independent term (proportional to c0). At O(ε) and O(ε3) all the terms in the expansion include the high-spatial-frequency prefactors e±2ikLx or e±4ikLx. For density distributions that vary slowly on the λ/2 length scale these average to zero. The O(ε2) term, however, has terms without these modulations, and is:

H^I(ε2)=12d3r(8c0ε2ψ^ψ^ψ^ψ^)

giving rise to c=c0Ω2/(8EL2).

Mean field phase diagram.

We compute the mean-field phase diagram for a ground-state BEC composed of a mixture of dressed spins in an infinite homogeneous system. This applies to our atoms in a harmonic trap in the limit of Rξs, where R is the system size, ξs=ħ2/2m|c2+c|n is the spin healing length and n is the density. We first minimize the interaction energy H^I at fixed N,, with an effective interaction c as a function of Ω. The two dressed spins are either phase-mixed, both fully occupying the system’s volume V, or phase-separated with a fixed total volume constraint V=V+V. For the phase-separated case, minimizing the free energy gives the volumes V and V, determined by N, and V. The interaction energy of a phase-mixed state is smaller than that of a phase-separated state for the miscibility condition c0+c2+c/2<c0(c0+c2), corresponding to Ω<Ωc. This condition is independent of N,: for any N, the system is miscible at Ω<Ωc. Then, at a given Ω, we minimize the sum of the interaction energy and the single-particle energy from the Raman detuning, (NN)δ/2, allowing N, to vary. For the miscible case (Ω<Ωc), the BEC is a mixture with fraction N/(N+N)(0,1) only in the range of detuning δ(δ0Wδ,δ0+Wδ), where δ0=c2n/2, Wδ=|δ0|(1Ω/Ωc)1/2 and n=(N+N)/V. For the immiscible case (Ω>Ωc),Wδ=(c2/8c0)c2n is negligibly small compared to c2n.

Figure 2b shows the mean field phase diagram as a function of (Ω,δ), where δ/EL is displayed with a quasi-logarithmic scaling, using the sign function (δ/EL)[log10(|δ/EL|+|δmin/EL|)log10|δmin/EL|], in order to display δ within the range of interest. This scaling function smoothly evolves from logarithmic, that is, approximately sgn(δ/EL)log10|δ/EL| for |δ|δmin, to linear, that is, approximately δ for |δ|δmin, where δmin/EL=0.001EL=1.5Hz.

In our measurement of the dressed spin fraction f (see Fig. 3a), δ=0 is determined from the NT=NT condition. We identify this condition as δ=δ0 and apply it for all hold times th. Because |δ0|3Hz is below our approximately 80-Hz root-mean-square field noise, we are unable to distinguish δ0 from 0.

Recombining TOF images of dressed spins.

To probe the dressed spin states (equation (3)), each of which is a spin and momentum superposition, we adiabatically mapped them into bare spins, |,κx=q+kL and |,κx=qkL, respectively. Then, in each image outside an 90μm radius disk containing the condensate for each spin distribution, we fitted nT,T(x,y) to a gaussian modelling the thermal background and subtracted that fit from nT,T(x,y) to obtain the condensate two-dimensional density n,(x,y). Thus, for each dressed spin we readily obtained the temperature, total number NT,T, and condensate densities n,(x,y).

To analyse the miscibility from the TOF images where a Stern-Gerlach gradient separated individual spin states, we re-centred the distributions to obtain n(x,y) and n(x,y). This took into account the displacement due to the Stern-Gerlach gradient and the non-zero velocities ħκx/m of each spin state (after the adiabatic mapping). The two origins were determined in the following way: we loaded the dressed states at a desired coupling Ω but with detuning δ chosen to put all atoms in either | or |. Because q,=(1Ω2/32EL2)kL (see Fig. 1c), these velocities ħκx/m=ħ(q+kL)/m,ħ(qkL)/m depend slightly on Ω, and our technique to determine the origin of the distributions accounts for this effect.

Calibration of Raman coupling.

Both Raman lasers were derived from the same Ti:sapphire laser at λ804.1nm, and were offset from each other by a pair of acousto-optic modulators driven by two phase-locked frequency synthesizers near 80 MHz. We calibrated the Raman coupling strength Ω by fitting the three-level Rabi oscillations between the mF=1,0 and +1 states driven by the Raman coupling to the expected behaviour.

Acknowledgements

We thank E. Demler, T.-L. Ho and H. Zhai for conceptual input; and we appreciate conversations with J.V. Porto and W. D. Phillips. This work was partially supported by ONR, ARO with funds from the DARPA OLE programme, and the NSF through the Physics Frontier Center at the Joint Quantum Institute. K.J.-G. acknowledges CONACYT.

Footnotes

Author Information Reprints and permissions information is available at www.nature.com/reprints. The authors declare no competing financial interests. Readers are welcome to comment on the online version of this article at www.nature.com/nature.

Full Methods and any associated references are available in the online version of the paper at www.nature.com/nature.

References

  • 1.Kato YK, Myers RC, Gossard AC & Awschalom DD Observation of the spin Hall effect in semiconductors. Science 306, 1910–1913 (2004). [DOI] [PubMed] [Google Scholar]
  • 2.Konig M. et al. Quantum spin Hall insulator state in HgTe quantum wells. Science 318, 766–770 (2007). [DOI] [PubMed] [Google Scholar]
  • 3.Kane CL & Mele EJ Z2 topological order and the quantum spin Hall effect. Phys. Rev. Lett 95,146802 (2005). [DOI] [PubMed] [Google Scholar]
  • 4.Bernevig BA, Hughes TL & Zhang S-C Quantum spin Hall effect and topological phase transition in HgTe quantum wells. Science 314,1757–1761 (2006). [DOI] [PubMed] [Google Scholar]
  • 5.Hsieh D. et al. A topological Dirac insulator in a quantum spin Hall phase. Nature 452,970–974 (2008). [DOI] [PubMed] [Google Scholar]
  • 6.Koralek JD et al. Emergence of the persistent spin helix in semiconductor quantum wells. Nature 458, 610–613 (2009). [DOI] [PubMed] [Google Scholar]
  • 7.Bychkov YA & Rashba EI Oscillatory effects and the magnetic susceptibility of carriers in inversion layers. J. Phys. C 17, 6039 (1984). [Google Scholar]
  • 8.Dresselhaus G. Spin-orbit coupling effects in zinc blende structures. Phys. Rev 100, 580–586(1955). [Google Scholar]
  • 9.Liu X-J, Borunda MF, Liu X. & Sinova J. Effect of induced spin-orbit coupling for atoms via laser fields. Phys. Rev. Lett 102,046402 (2009). [DOI] [PubMed] [Google Scholar]
  • 10.Quay CHL et al. Observation of a one-dimensional spin-orbit gap in a quantum wire. Nature Phys. 6,336–339 (2010). [Google Scholar]
  • 11.von Klitzing K, Dorda G. & Pepper M. New method for high-accuracy determination of the fine-structure constant based on quantized Hall resistance. Phys. Rev. Lett 45,494–497 (1980). [Google Scholar]
  • 12.Ruseckas J, Juzeliunas G, Öhberg P. & Fleischhauer M. Non-abelian gauge potentials for ultracold atoms with degenerate dark states. Phys. Rev. Lett 95, 010404 (2005). [DOI] [PubMed] [Google Scholar]
  • 13.Stanescu TD, Zhang C. & Galitski V. Nonequilibrium spin dynamics in a trapped Fermi gas with effective spin-orbit interactions. Phys. Rev. Lett 99, 110403(2007). [DOI] [PubMed] [Google Scholar]
  • 14.Dalibard J, Gerbier F, Juzeliūnas G. & Öhberg P. Artificial gauge potentials for neutral atoms. Preprint at (http://arxiv.org/abs/1008.5378) (2010). [Google Scholar]
  • 15.Higbie J& Stamper-Kurn DM Generating macroscopic quantum-superposition states in momentum and internal-state space from Bose-Einstein condensates with repulsive interactions. Phys. Rev. A 69, 053605 (2004). [Google Scholar]
  • 16.Lin Y-J et al. Bose-Einstein condensate in a uniform light-induced vector potential. Phys. Rev. Lett 102,130401 (2009). [DOI] [PubMed] [Google Scholar]
  • 17.Lin Y-J et al. A synthetic electric force acting on neutral atoms. Nature Phys. (in the press); preprint at (http://arxiv.org/abs/1008.4864) (2010). [Google Scholar]
  • 18.Lin YJ, Compton RL, Jimenez-Garcia K, Porto JV & Spielman IB Synthetic magnetic fields for ultracold neutral atoms. Nature 462, 628–632 (2009). [DOI] [PubMed] [Google Scholar]
  • 19.Stenger J. et al. Spin domains in ground-state Bose-Einstein condensates. Nature 396,345–348(1998). [Google Scholar]
  • 20.Chang M-S et al. Observation of spinor dynamics in optically trapped 87Rb Bose-Einstein condensates. Phys. Rev. Lett 92, 140403 (2004). [DOI] [PubMed] [Google Scholar]
  • 21.Ho T-L Spinor Bose condensates in optical traps. Phys. Rev. Lett 81, 742–745 (1998). [Google Scholar]
  • 22.Widera A. et al. Precision measurement of spin-dependent interaction strengths for spin-1 and spin-2 87Rb atoms. N.J. Phys 8,152 (2006). [Google Scholar]
  • 23.Ho T-L & Zhang S. Bose-Einstein condensates in non-abelian gauge fields. Preprint at (http://arxiv.org/abs/1007.0650) (2010). [Google Scholar]
  • 24.Hall DS, Matthews MR, Ensher JR, Wieman CE & Cornell EA Dynamics of component separation in a binary mixture of Bose-Einstein condensates. Phys. Rev. Lett 81,1539–1542 (1998). [Google Scholar]
  • 25.Erhard M, Schmaljohann H, Kronjäger J, Bongs K. & Sengstock K. Measurement of a mixed-spin-channel Feshbach resonance in 87Rb. Phys. Rev. A 69,032705(2004). [Google Scholar]
  • 26.Kobayashi M, Kawaguchi Y, Nitta M. & Ueda M. Collision dynamics and rung formation of non-abelian vortices. Phys. Rev. Lett 103,115301 (2009). [DOI] [PubMed] [Google Scholar]
  • 27.Goldman N. et al. Realistic time-reversal invariant topological insulators with neutral atoms. Phys. Rev. Lett 105, 255302 (2010). [DOI] [PubMed] [Google Scholar]
  • 28.Zhang C,Tewari S, Lutchyn RM& DasSarma S. px + ipy superfluid from s-wave interactions of fermionic cold atoms. Phys. Rev. Lett 101,160401 (2008). [DOI] [PubMed] [Google Scholar]
  • 29.Sau JD, Tewari S, Lutchyn RM, Stanescu TD & Das Sarma S. Non-Abelian quantum order in spin-orbit-coupled semiconductors: search for topological Majorana particles in solid-state systems. Phys. Rev. B 82, 214509 (2010). [Google Scholar]
  • 30.Lin Y-J, Perry AR, Compton RL, Spielman IB & Porto JV Rapid production of87Rb Bose-Einstein condensates in a combined magnetic and optical potential. Phys. Rev. A 79,063631 (2009). [Google Scholar]

RESOURCES