Dipolar evaporation of reactive molecules to below the Fermi temperature

Abstract

The control of molecules is key to the investigation of quantum phases, in which rich degrees of freedom can be used to encode information and strong interactions can be precisely tuned1. Inelastic losses in molecular collisions2,3,4,5, however, have greatly hampered the engineering of low-entropy molecular systems6. So far, the only quantum degenerate gas of molecules has been created via association of two highly degenerate atomic gases7,8. Here we use an external electric field along with optical lattice confinement to create a two-dimensional Fermi gas of spin-polarized potassium–rubidium (KRb) polar molecules, in which elastic, tunable dipolar interactions dominate over all inelastic processes. Direct thermalization among the molecules in the trap leads to efficient dipolar evaporative cooling, yielding a rapid increase in phase-space density. At the onset of quantum degeneracy, we observe the effects of Fermi statistics on the thermodynamics of the molecular gas. These results demonstrate a general strategy for achieving quantum degeneracy in dipolar molecular gases in which strong, long-range and anisotropic dipolar interactions can drive the emergence of exotic many-body phases, such as interlayer pairing and p-wave superfluidity.

Main

The complex internal structure of molecules can be both useful and a hindrance: it represents a key resource for the development of tunable and programmable quantum devices1,9,10, but it is also responsible for strong inelastic losses during collisions11,12,13,14. Despite recent advances in molecular quantum science15,16,17,18,19,20,21,22,23,24, full control of elastic collisions between molecules has not been achieved, making it very difficult to create the low-entropy bulk molecular gases that are required for the exploration of rich many-body physics and emergent quantum phenomena1,25.

Here, we report the realization of highly tunable elastic interactions in a quantum gas of polar molecules through the application of an external electric field along a stack of two-dimensional (2D) layers generated with a one-dimensional optical lattice. The induced electric dipole moment in the laboratory frame gives rise to repulsive dipolar interactions that stabilize the molecular gas against reactive collisions and formation of collisional complexes. These long-range interactions provide a large elastic collision cross-section for identical ultracold fermionic molecules, in contrast to contact interactions26. We demonstrate the enhancement of dipolar interactions by several orders of magnitude and achieve a ratio of elastic-to-inelastic collisions that exceeds 100. This favourable interaction regime enables direct molecular thermalization and efficient evaporative cooling, allowing us to bring the molecular temperature T below the Fermi temperature TF. The onset of quantum degeneracy is signalled by deviations from the classical expansion energy as the ratio T/TF is reduced below unity7,27.

Our strategy follows previous theory proposals28,29,30 and our earlier experimental study on molecular reactions in quasi-two dimensions31. This geometry allows us to take advantage of the anisotropic character of the dipolar potential and retain only the repulsive side-to-side dipole–dipole interactions within each 2D site, while preventing the attractive head-to-tail interactions that facilitate losses at short range. Our recent advances in the production of degenerate Fermi gases of polar molecules7,8, combined with precise electric-field control using in-vacuum electrodes32 (Fig. 1), allow us to perform a systematic characterization of the properties of a 2D Fermi gas of polar molecules.

Fig. 1: Experimental setup.
figure1

a, The 2D molecular cloud is trapped at the centre of the electrode assembly (grey). 2D optical trapping is achieved with the VL (green), which is loaded using the ODT (orange) and LSL (red). Absorption images of molecules are collected through the same lens as that used to focus the LSL. b, Sketch of the experiment as seen down the z axis. The bias electric field is generated along y, perpendicular to the 2D layers of the VL. c, Matter-wave focusing data of the Rb layers in the VL, which have a spacing of 540 nm.

A long-lived 2D Fermi gas of polar molecules

The KRb 2D Fermi gas is created from an ultracold atomic mixture of fermionic 40K and bosonic 87Rb atoms. The atomic mixture is initially held in a crossed optical dipole trap (ODT) and then transferred into a single layer of a large-spacing lattice (LSL) with an 8-μm spatial period, which increases the mixture’s confinement along the vertical direction (y). The mixture is then transferred into a vertical lattice (VL) with spacing of 540 nm that confines it to a quasi-2D geometry. The intermediate LSL transfer results in the Rb cloud populating a controllable number of VL layers τ ranging between 5 and 15. We directly probe the number of occupied 2D layers via a matter-wave focusing technique on the Rb cloud (Fig. 1c)33,34. The measured τ is in excellent agreement with theoretical modelling of the in situ cloud size (see Methods).

Magneto-optical association is used to pair roughly half of the initial Rb atoms into ground-state KRb molecules35. This process is fast and coherent, and the resulting molecular cloud populates the same layers originally occupied by the Rb cloud. The leftover K and Rb atoms are selectively and quickly removed from the trap. In the VL, the trap frequencies are set to (ωx, ωy, ωz= 2π × (40, 17,000, 40) Hz. The quoted trapping frequencies are for KRb throughout the paper unless otherwise stated. We create a 2D gas with N ≈ 20,000 trapped molecules, a typical temperature T ≈ 250 nK, and T/TF ranging from 1.5 to 3 depending on τ.

The 2D molecular cloud is at the centre of an in-vacuum six-electrode assembly composed of two indium tin oxide (ITO)-coated glass plates and four tungsten rods (Fig. 1a). With this, we generate a highly tunable bias electric field EDC that induces strong dipolar interactions between molecules (Fig. 1b). The ratio γ of the voltage of the rods to the voltage of the plates can be used to cancel the curvature introduced by the parallel plate edges (flat-field configuration) or to introduce additional curvatures and gradients for molecule manipulation.

The chemically reactive KRb molecules suffer from inelastic two-body losses2,12, which result in the average molecular density n decaying over time t according to a two-body rate equation of the form:

$$\frac{{\rm{d}}n}{{\rm{d}}t}=-\beta {n}^{2}+\frac{{\rm{\partial }}n}{{\rm{\partial }}T}\frac{{\rm{d}}T}{{\rm{d}}t},$$
(1)

where β is the two-body loss rate coefficient and the second term on the right side of equation (1) accounts for temperature changes affecting the density3.

In a three-dimensional (3D) harmonic trap, β increases sharply with EDC, so that inelastic interactions dominate elastic ones3,36. However, strong confinement along the direction of EDC suppresses this detrimental loss increase31 by preventing head-to-tail collisions along EDC. Even though n is large in the occupied layers, the molecular gas shows a remarkable stability with repulsive interactions turned on. With an induced dipole moment d = 0.2 D in the flat-field configuration, KRb molecules survive for several seconds (Fig. 2a). The evolution of β as a function of EDC is shown in Fig. 2b. Close to EDC = 4.7 kV cm−1, β reaches a minimum of nearly five times below the zero-field value. The increase of β for d > 0.2 D is consistent with a quasi-2D picture of dipolar scattering37,38.

Fig. 2: Long-lived polar molecules in 2D.
figure2

a, Time evolution of the molecular density n at d = 0.2 D. b, Inelastic loss rate β as a function of dipole moment. All error bars are 1 standard error of the mean (s.e.), determined from two-body decay fits (equation (1)) The top x axis shows the bias electric field EDC at the corresponding dipole moment. c, Both β (grey circles) and the heating rate (orange squares) saturate at their minimum values near ωy = 2π × 7 kHz (vertical grey bar indicates uncertainty in the molecule temperature), consistent with the mechanism of quasi-2D dipolar scattering. Heating rate error bars are 1 s.e., determined from linear fits.

To understand the effect of optical confinement, we perform a thorough characterization of the 3D-to-2D crossover by measuring β versus ωy at EDC = 5 kV cm−1 (Fig. 2c). Here, β drops abruptly as the lattice vertical confinement is increased and reaches a plateau near ωy = 2π × 7 kHz, corresponding to the quasi-2D limit where kBT ħωy, with kB being the Boltzmann constant and ħ the reduced Planck constant, and the molecules principally occupy the lowest band of the VL. In contrast to the 3D case, where the heating rate exceeds 3 µK s−1, in quasi-two dimensions we do not record a substantial increase in temperature along the radial direction. The suppression of heating is due to cancellation of anti-evaporation in quasi-two dimensions30 and represents another advantage of this configuration.

Cross-dimensional thermalization

To characterize elastic interactions in our thermal molecular cloud, we perform cross-dimensional thermalization at various values of EDC. We diabatically change the power in one of the ODT beams to suddenly increase the energy along z. Elastic collisions between the molecules then redistribute the excess energy from z onto x. The rate Γth of the temperature equilibration between the two axes is proportional to the dipolar elastic collision rate28,39,40. We extract Γth by fitting the increase of T along x with an exponential curve.

With the loss suppressed in quasi-2D, we expect a substantial increase of the thermalization rate Γth with a d4 dependence28,30. Comparing the thermalization dynamics observed for d = 0.1 D and 0.21 D (Fig. 3a), we see Γth increase by a factor of 10. Over our investigated range of EDC, Γth changes by two orders of magnitude (Fig. 3b), showing the extreme tunability of elastic dipolar interactions in our system. We observe the cross-dimensional thermalization dynamics at lower dipolar strength (d < 0.1 D) being dominated by cross-dimensional relaxation owing to trap anharmonicity, which limits the smallest Γth that we can measure. For d ≥ 0.1 D, a fit to a power-law dependence of Γth on d yields a power of 3.3 ± 1.0, in good agreement with theoretical expectations28,30. For the highest values of d we explored, the rate Γth is comparable to the radial trapping frequency, opening the way for future studies of collective dynamics in molecular gases41.

Fig. 3: Tuning strong dipolar elastic interactions in a 2D molecular gas.
figure3

a, Cross-dimensional thermalization dynamics at d = 0.1 D (orange diamonds) and d = 0.21 D (grey squares). Error bars are 1 s.e. of 5 independent measurements. b, The trend of Γth extracted from cross-dimensional thermalization dynamics as a function of d. The solid line is a power-law fit for d ≥ 0.1 D that yields a power of 3.3(1.0). The filled grey circle corresponds to the measurement at d = 0.0 D, and it is artificially placed at d = 0.03 D for figure clarity. The dashed horizontal line at Γth = 2 s−1 is the background cross-dimensional relaxation from trap anharmonicity. All error bars are 1 s.e., determined from exponential fits.

An estimate of the ratio of elastic-to-inelastic collisions is obtained by comparing Γth to the initial rate of inelastic losses Γin, which is expressed as Γin = βn0, with n0 the initial average density of the 2D gas. From the data in Fig. 2b, at d = 0.2 D, we estimate a rate Γin = 0.83(5) s−1, whereas Γth = 21(6) s−1 for the same dipole strength. In the temperature regime of the cross-dimensional thermalization experiments, theoretical models30 predict that α = 8 elastic collisions are needed for each molecule to reach thermal equilibrium. This indicates a ratio of elastic-to-inelastic collisions α(Γth/Γin) = 200 ± 60, demonstrating that elastic processes dominate in this system.

Electric-field-controlled evaporative cooling

The large elastic-to-inelastic collision ratio is an excellent setting for evaporatively cooling to enhance the phase-space density of our molecular cloud. For non-degenerate 2D fermionic gases, phase-space density is inversely proportional to (T/TF)2, and phase-space density increases only if the change of N versus T during evaporation fulfills the criterion:

$${S}_{{\rm{e}}{\rm{v}}{\rm{a}}{\rm{p}}}=\frac{{\rm{\partial }}\log N}{{\rm{\partial }}\log T} < 2.$$
(2)

When Sevap = 2, the gas maintains a constant T/TF.

The efficiency of evaporative cooling relies on our ability to selectively remove the hottest molecules from the trap and to let the remaining molecules re-thermalize to a lower temperature. Reducing the trap depth by lowering the optical trap power for evaporation, as is routinely done in ultracold atom experiments, is not a viable solution here because we cannot lower the tight 2D confinement without affecting the stability of the cloud (Fig. 2c). Instead, by increasing γ with respect to the flat-field configuration, we introduce a tunable anti-trapping electric field along the x direction to reduce the radial confinement experienced by the molecules (Fig. 4a). By measuring the change of ωx as a function of γ, we can directly reconstruct the combined electro-optical potential and benchmark its theoretical modelling (see Extended Data Fig. 2).

Fig. 4: Evaporative cooling to the quantum degenerate regime.
figure4

a, Cuts along the x axis of the combined electro-optical potential for the flat-field configuration (left panel) and at the end of the evaporation (right panel). b, Evolution of N and T (orange squares) at different stages of the evaporation trajectory at EDC = 6.5 kV cm−1 and d = 0.25 D. The power-law fit (orange line) yields Sevap = 1.06(15). The dashed grey line is for a constant T/TF, corresponding to Sevap = 2.0. Error bars are 1 s.e. of four independent measurements. c, Summary of Sevap versus d. All error bars are 1 s.e., determined from power-law fits. d, Average of 20 band-mapped absorption images of the molecular cloud in the x–y plane after 5.84 ms of time of flight for T/TF = 2.0(1) (top) and T/TF = 0.81(15) (bottom). e, Optical density (OD, dimensionless) profiles (orange circles for T/TF = 2.0(1) and grey diamonds for T/TF = 0.81(15)) of the images in d (integrated along y), together with the Fermi–Dirac fit to the whole cloud (grey line) and the Gaussian fit to the outer wings (orange line). f, Measurement of δU/U at different values of T/TF from the Fermi–Dirac fit to the entire cloud (grey circles) and from the Gaussian fit to the outer wings of the cloud (orange squares). The solid and dashed curves show δU/U for the 2D and 3D ideal Fermi gases, respectively. All error bars are 1 s.e., determined from Gaussian or polylogarithmic fits.

For the evaporation measurement, we start with a 2D gas with layer number τ = 5 ± 1, ωy = 2π × 17 kHz, and an average T/TF = 1.5(1). After creating the molecules (see Methods), we ramp EDC to a target field while keeping γ at the flat-field value. We trigger the evaporation by increasing γ to reduce the trap depth. We do not observe any evaporation until the truncation parameter η, defined as the ratio of trap depth over thermal energy kBT, reaches a value of 4 (see Methods), in good agreement with theoretical expectations30. We further increase γ over a timescale of hundreds of milliseconds, which is long enough for the molecules to efficiently re-thermalize at a lower T as the trap depth is reduced. At the end of the evaporation ramp, we return to the flat-field configuration and ramp EDC back to its initial value. We coherently convert the ground state molecules back to the Feshbach state and image the cloud of Feshbach molecules after band-mapping from the VL. Detailed time sequences for the evolution of EDC, γ and trap depth are shown in the Methods.

At EDC = 6.5 kV cm−1, the evolution of N and T at different stages of the optimized evaporation sequence is shown in Fig. 4b. To characterize the evaporation efficiency, we fit the N versus T dependence with a power-law function to extract Sevap. For the data shown in Fig. 4b, we obtain Sevap = 1.06(15), far below the threshold of 2 required to increase phase-space density. The trend of Sevap versus d is plotted in Fig. 4c and reaches a minimum (that is, maximum increase in phase-space density) at d = 0.25 D, where the ratio of elastic-to-inelastic collisions is the largest37,38.

When we cool molecules to T < TF (Fig. 4d), we witness the onset of Fermi degeneracy, which is signalled by deviations from classical thermodynamics owing to the increasing role of the Pauli exclusion principle. Here, \({T}_{{\rm{F}}}=\frac{\hbar {\omega }_{{\rm{R}}}}{{k}_{{\rm{B}}}}{\left(\frac{2N}{\tau }\right)}^{1/2}\), with \({\omega }_{{\rm{R}}}=\sqrt{{\omega }_{x}{\omega }_{z}}\,\) the geometric mean of the radial trapping frequency. Our best result produced a 2D molecular Fermi gas with N = 1.7(1) × 103 and T/TF = 0.6(2).

We extract T by using either a fit to the Fermi–Dirac distribution on the entire expanded cloud or a Gaussian fit restricted to the cloud’s outer wings (see Methods). As shown in Fig. 4e, for T/TF = 0.81(15) the Gaussian fit to the outer wings of the time of flight density profile overestimates the density at the centre. We quantify this through the increasing difference δU = U − Ucl between the energy U of the fermionic gas and the energy Ucl kBT, as T/TF decreases7,27. U is determined from a Gaussian fit to the whole cloud (see Methods) and Ucl is calculated based on the measured T. Owing to the different density of states in the harmonic trap, the chemical potential crosses zero for T/TF = 0.78 for the 2D case, in contrast to 0.57 for the 3D case. Correspondingly, the 2D Fermi gas shows a larger δU with respect to the 3D case at the same T/TF (ref. 42). As we reach T < TF, in excellent agreement with theoretical expectations, we observe a large increase of δU (Fig. 4f). This is a hallmark for the onset of quantum degeneracy in trapped Fermi gases27.

Conclusions

We have realized a 2D Fermi gas of reactive polar molecules where precisely tunable elastic dipolar interactions dominate all inelastic processes. This allowed us to perform evaporative cooling of molecules to the onset of Fermi degeneracy. We demonstrated a general and robust scheme for ultracold gases of polar molecules to reach quantum degeneracy. For example, using a strong 2D confinement and large dipolar interactions, this method should enable Bose–Einstein condensation in bosonic molecular gases. It has long been anticipated that quantum gases of polar molecules in two dimensions would allow access to strongly correlated many-body phases43,44,45,46,47,48,49,50,51,52. Our results set the stage for exploration of these exotic regimes.

Methods

Experimental protocol

The experiment starts with an ultracold atomic mixture of 40K and 87Rb, held in the ODT at a magnetic field of 555 G. The trap frequencies for Rb in the ODT are (ωx, ωy, ωz= 2π × (40, 180, 40) Hz. The atomic mixture is then transferred into a single layer of the LSL. The LSL beams propagate at a shallow angle of 4 degrees along z, resulting in a lattice spacing of 8 µm along y. At the end of the LSL ramp, we decrease the ODT power, so that the trap frequencies for Rb in the combined trap are (ωx, ωy, ωz= 2π × (25, 600, 25) Hz. Typically, we have 4.1 × 105 K atoms and 7.0 × 104 Rb atoms at T = 115(10) nK. About 30% of Rb is condensed. At this point, we load the mixture into the VL and adjust the ODT in order for the KRb molecules to experience radial trap frequencies at zero electric field of (ωx, ωz= 2π × (40, 40) Hz, with ωy/(2π) ranging from a few kilohertz up to 20 kHz. To compensate for the limited transmittivity of the ITO plates at the 1.064-µm VL wavelength and to avoid spurious superlattices, the VL beams have a 11-degree tilt with respect to y, resulting in a lattice spacing of 540 nm.

To create molecules, we first sweep the magnetic field adiabatically through the KRb heteronuclear Feshbach resonance at 546.6 G. The magnetic field is ramped from 555 G to 545.5 G in 4 ms, creating 2.5 × 104 Feshbach molecules that are subsequently transferred to the absolute KRb ground state by stimulated Raman adiabatic passage (STIRAP)1. By tuning the Raman lasers, we create KRb molecules at 0 kV cm−1 or 4.5 kV cm−1. For molecule creation at 4.5 kV cm−1, the field is ramped to the target value 10 ms before the Feshbach sweep. The STIRAP one-way transfer efficiency is 85(2)% at 0 kV cm−1 and 82(3)% at 4.5 kV cm−1. We do not observe any dependence of β and Sevap on the initial value of the electric field.

Matter-wave focusing and layer counting

The VL layer spacing of 540 nm is too small to resolve with conventional absorption imaging. To quantify the number of occupied layers τ, we use a matter-wave technique that maps the in situ density distribution onto the momentum distribution, which can then be imaged in time of flight. To do so, we instantaneously release the cloud from the VL and the LSL into the ODT. The cloud expands into the ODT harmonic potential for a quarter of the oscillation period along y. This corresponds to a 90-degree rotation in phase space. As a result, after the rotation, the momentum distribution along y in time of flight corresponds to the original in-situ density profile. Increasing the time of flight increases the layer separation until they can be resolved optically. From a set of averaged matter-wave density profiles, we obtain a histogram with the normalized particle number per layer from which we extract the number of layers τ. We perform this analysis on a cloud of Rb atoms without K, eliminating the K–Rb interactions during the phase space rotation time. The Rb cloud used for matter-wave amplification imaging has the same trap parameters, number and temperature of the Rb cloud used for the molecule experiment. When the K–Rb interactions are removed by setting the magnetic field to the zero-crossing of the Feshbach resonance, the full contrast is restored. For the data in Fig. 1c, obtained by averaging 20 matter-wave images of the Rb cloud, the density histogram is shown in Extended Data Fig. 1. For a fixed molecule distribution, the definition of τ depends on the physical quantity being calculated. Using \(\tau =N/\langle {N}_{i}\rangle \), where \(\langle {N}_{i}\rangle \) is the average particle number per layer over the measured distribution, we extract τ = 4.6(2) for the data in Extended Data Fig. 1. Theoretical modelling53 for the Rb cloud in the same conditions yields a consistent value of τ = 5.1(2).

Our measurements of the molecular cloud thus involve averaging over layers that are not equally populated. To determine T/TF, we use the layer-averaged Fermi temperature \({T}_{{\rm{F}}}=\frac{\hbar {\omega }_{{\rm{R}}}}{{k}_{{\rm{B}}}}\sqrt{2}\sqrt{{N}_{i}}=\frac{\hbar {\omega }_{{\rm{R}}}}{{k}_{{\rm{B}}}}{\left(\frac{2N}{\tau }\right)}^{1/2}\), where \(\tau =N/{\sqrt{{N}_{i}}}^{2}\,\) is an effective number of layers that accounts for the nonlinear dependence of TF on Ni. For the data in Extended Data Fig. 1, we extract τ = 4.9(0.2). For the T/TF data in the paper we thus use the closest estimate τ = 5 ± 1, where the uncertainty accounts for possible systematic errors arising from non-uniform conversion of Rb to KRb and variation of evaporation efficiency between the layers (since the density, and hence thermalization rate, varies between the layers).

To determine the 2D density for the measurement of β, we need to use a time-averaged layer number that considers the density dependence of the loss in each layer3. The layer-averaged 2D density is defined as n = N/(4πσ2τ), where σ is the root-mean-square cloud size in the radial direction. Through numerical simulation, we obtain the decay over time for a cloud with the layer distribution plotted in Extended Data Fig. 1 and compare it with the decay of a gas with a uniform layer distribution and same number N. In this case, we define τ as the value for which β in the uniform case matches β in the non-uniform case. For Extended Data Fig. 1, we obtain τ = 8 ± 1.

Electric field potential

The anti-trapping potential that is used for dipolar evaporation introduces an anti-curvature that changes the trap frequency ωx as a function of γ. Owing to the geometry of our electrodes, when γ = 0.4225 the electric field potential at the molecule position is as homogeneous as possible (fourth-order cancellation of the electric-field curvature). By increasing (decreasing) γ, ωx decreases (increases), as shown in Extended Data Fig. 2 at EDC = 5 kV cm−1. Our results follow the expected trend from finite-element simulations of the combined electro-optical potential at different γ.

Electric-field evaporation ramps

For the evaporation experiments, we lower the trap depth by increasing γ over time. The trap depth at each time point of the evaporation is estimated by simulations of the combined electro-optical potential, which is benchmarked with the measurement of ωx versus γ displayed in Extended Data Fig. 2. For the data shown in Fig. 4, the evaporation ramp takes about 800 ms and EDC, γ and trap depth evolve over time according to the plots in Extended Data Fig. 3. We also plot the trend of the parameter η and T/TF at different time points of the evaporation sequence.

Thermometry of 2D Fermi gas

The 2D in situ density n of the molecular Fermi gas is given by the Fermi–Dirac distribution54:

$$n(x,z)=-\frac{1}{{\lambda }_{{\rm{d}}{\rm{B}}}^{2}}{{\rm{L}}{\rm{i}}}_{1}\,\left\{-\exp \,[-{\beta }_{{\rm{t}}{\rm{h}}}\,\left(\frac{1}{2}m\,{\omega }_{x}^{2}{x}^{2}+\frac{1}{2}m\,{\omega }_{z}^{2}{z}^{2}-\mu \right)]\right\},$$
(3)

where m is the molecular mass, \({\beta }_{{\rm{t}}{\rm{h}}}=\frac{1}{{k}_{{\rm{B}}}T}\), µ the chemical potential, \({\lambda }_{{\rm{d}}{\rm{B}}}=\sqrt{2{\rm{\pi }}{\hbar }^{2}{\beta }_{{\rm{t}}{\rm{h}}}/m}\) the thermal de Broglie wavelength, and Lin the polylogarithmic function of order n. The column-integrated density profile is:

$${n}_{{\rm{i}}{\rm{n}}{\rm{t}}}(x)=-\frac{1}{{\lambda }_{{\rm{d}}{\rm{B}}}^{2}}\sqrt{\frac{2{\rm{\pi }}}{{\beta }_{{\rm{t}}{\rm{h}}}m{\omega }_{z}^{2}}}{{\rm{L}}{\rm{i}}}_{3/2}\,\left\{-\exp \,[{\beta }_{{\rm{t}}{\rm{h}}}\,\left(\mu -\frac{1}{2}m{\omega }_{x}^{2}{x}^{2}\right)]\right\}.$$
(4)

After a certain time of flight t, the x coordinate scales by a factor \(1/\sqrt{1+{\omega }_{x}^{2}{t}^{2}}\). The density is consequently divided by \(\sqrt{1+{\omega }_{x}^{2}{t}^{2}}\) for proper renormalization. The chemical potential µ is defined through the relation:

$$\frac{N}{\tau }={\left(\frac{{k}_{{\rm{B}}}T}{\hbar {\omega }_{{\rm{R}}}}\right)}^{2}{{\rm{L}}{\rm{i}}}_{2}(\,-{{\rm{e}}}^{{\beta }_{{\rm{t}}{\rm{h}}}\mu }).$$
(5)

Combining equation (5) with the definition of TF in 2D, we obtain:

$${\left(\frac{T}{{T}_{{\rm{F}}}}\right)}^{2}=-\frac{1}{2\,{{\rm{L}}{\rm{i}}}_{2}(\,-{{\rm{e}}}^{{\beta }_{{\rm{t}}{\rm{h}}}\mu })},$$
(6)

which allows us to extract the ratio T/TF from the polylogarithmic fit.

From the Gaussian fit to the whole cloud, we obtain the Gaussian width σ and a release temperature Trel, defined as:

$${T}_{{\rm{r}}{\rm{e}}{\rm{l}}}=\frac{m{\omega }_{x}^{2}{\sigma }^{2}}{{k}_{{\rm{B}}}(1+\,{\omega }_{x}^{2}{t}^{2})}.$$
(7)

The release temperature Trel is proportional to the energy density U of the Fermi gas, U = 2kBTrel, which saturates to a non-zero value as T → 0. In contrast, the energy density Ucl = 2kBT of a classical gas approaches zero as T → 0.

When the Gaussian fit is constrained to only the outer wings (that is, the high-momentum states) of the cloud, we can extract a new width σout from which, using equation (7), we obtain a corrected temperature Tout through the relation:

$${T}_{{\rm{o}}{\rm{u}}{\rm{t}}}=\frac{m{\omega }_{x}^{2}{\sigma }_{{\rm{o}}{\rm{u}}{\rm{t}}}^{2}}{{k}_{{\rm{B}}}(1+{\omega }_{x}^{2}{t}^{2})}.$$
(8)

As the excluded region from the centre of the Gaussian fit is expanded, Tout decreases from an initial value of Trel and approaches T. This is shown in Extended Data Fig. 4, where we plot the ratio Tout/Trel at different exclusion regions in units of σ. For the range of T/TF studied here, we find that an exclusion region of 1.5σ leaves us enough signal-to-noise ratio for the fit to properly converge and to return a value of Tout that is only 5% higher than T. In the main text, the Gaussian fit on the outer wings is performed with an excluded region of 1.5σ.

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request. Source data are provided with this paper.

References

  1. 1.

    Bohn, J. L., Rey, A. M. & Ye, J. Cold molecules: progress in quantum engineering of chemistry and quantum matter. Science 357, 1002–1010 (2017).

    ADS  MathSciNet  CAS  MATH  Google Scholar 

  2. 2.

    Ospelkaus, S. et al. Quantum-state controlled chemical reactions of ultracold potassium-rubidium molecules. Science 327, 853–857 (2010).

    ADS  CAS  Google Scholar 

  3. 3.

    Ni, K.-K. et al. Dipolar collisions of polar molecules in the quantum regime. Nature 464, 1324–1328 (2010).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  4. 4.

    Guo, M. et al. Dipolar collisions of ultracold ground-state bosonic molecules. Phys. Rev. X 8, 041044 (2018).

    CAS  Google Scholar 

  5. 5.

    Gregory, P. D. et al. Sticky collisions of ultracold RbCs molecules. Nat. Commun. 10, 3104 (2019).

    ADS  PubMed  PubMed Central  Google Scholar 

  6. 6.

    Moses, S. A. et al. Creation of a low-entropy quantum gas of polar molecules in an optical lattice. Science 350, 659–662 (2015).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  7. 7.

    De Marco, L. et al. A degenerate Fermi gas of polar molecules. Science 363, 853–856 (2019).

    ADS  PubMed  PubMed Central  Google Scholar 

  8. 8.

    Tobias, W. G. et al. Thermalization and sub-Poissonian density fluctuations in a degenerate molecular Fermi gas. Phys. Rev. Lett. 124, 033401 (2020).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  9. 9.

    André, A. et al. A coherent all-electrical interface between polar molecules and mesoscopic superconducting resonators. Nat. Phys. 2, 636–642 (2006).

    Google Scholar 

  10. 10.

    DeMille, D., Doyle, J. M. & Sushkov, A. O. Probing the frontiers of particle physics with tabletop-scale experiments. Science 357, 990–994 (2017).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  11. 11.

    Gregory, P. D., Blackmore, J. A., Bromley, S. L. & Cornish, S. L. Loss of ultracold 87Rb133Cs molecules via optical excitation of long-lived two-body collision complexes. Phys. Rev. Lett. 124, 163402 (2020).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  12. 12.

    Hu, M. G. et al. Direct observation of bimolecular reactions of ultracold KRb molecules. Science 366, 1111–1115 (2019).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  13. 13.

    Kirste, M. et al. Quantum-state resolved bimolecular collisions of velocity-controlled OH with NO radicals. Science 338, 1060–1063 (2012).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  14. 14.

    Christianen, A., Zwierlein, M. W., Groenenboom, G. C. & Karman, T. Photoinduced two-body loss of ultracold molecules. Phys. Rev. Lett. 123, 123402 (2019).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  15. 15.

    Danzl, J. G. et al. Quantum gas of deeply bound ground state molecules. Science 321, 1062–1066 (2008).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  16. 16.

    Seeßelberg, F. et al. Extending rotational coherence of interacting polar molecules in a spin-decoupled magic trap. Phys. Rev. Lett. 121, 253401 (2018).

    ADS  PubMed  PubMed Central  Google Scholar 

  17. 17.

    Will, S. A., Park, J. W., Yan, Z. Z., Loh, H. & Zwierlein, M. W. Coherent microwave control of ultracold 23Na40K molecules. Phys. Rev. Lett. 116, 225306 (2016).

    ADS  PubMed  PubMed Central  Google Scholar 

  18. 18.

    Barry, J. F., McCarron, D. J., Norrgard, E. B., Steinecker, M. H. & Demille, D. Magneto-optical trapping of a diatomic molecule. Nature 512, 286–289 (2014).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  19. 19.

    Truppe, S. et al. Molecules cooled below the Doppler limit. Nat. Phys. 13, 1173–1176 (2017).

    CAS  Google Scholar 

  20. 20.

    Anderegg, L. et al. An optical tweezer array of ultracold molecules. Science 365, 1156–1158 (2019).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  21. 21.

    Ding, S., Wu, Y., Finneran, I. A., Burau, J. J. & Ye, J. Sub-Doppler cooling and compressed trapping of YO molecules at μK temperatures. Phys. Rev. X 10, 021049 (2020).

    CAS  Google Scholar 

  22. 22.

    Yang, H. et al. Observation of magnetically tunable Feshbach resonances in ultracold 23Na40K + 40K collisions. Science 363, 261–264 (2019).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  23. 23.

    Son, H., Park, J. J., Ketterle, W. & Jamison, A. O. Collisional cooling of ultracold molecules. Nature 580, 197–200 (2020).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  24. 24.

    Segev, Y. et al. Collisions between cold molecules in a superconducting magnetic trap. Nature 572, 189–193 (2019).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  25. 25.

    Baranov, M. A., Dalmonte, M., Pupillo, G. & Zoller, P. Condensed matter theory of dipolar quantum gases. Chem. Rev. 112, 5012–5061 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. 26.

    Aikawa, K. et al. Reaching Fermi degeneracy via universal dipolar scattering. Phys. Rev. Lett. 112, 010404 (2014).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  27. 27.

    DeMarco, B. & Jin, D. S. Onset of Fermi degeneracy in a trapped atomic gas. Science 285, 1703–1706 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. 28.

    Quéméner, G. & Bohn, J. L. Dynamics of ultracold molecules in confined geometry and electric field. Phys. Rev. A 83, 012705 (2011).

    ADS  Google Scholar 

  29. 29.

    Micheli, A. et al. Universal rates for reactive ultracold polar molecules in reduced dimensions. Phys. Rev. Lett. 105, 073202 (2010).

    ADS  PubMed  PubMed Central  Google Scholar 

  30. 30.

    Zhu, B., Quéméner, G., Rey, A. M. & Holland, M. J. Evaporative cooling of reactive polar molecules confined in a two-dimensional geometry. Phys. Rev. A 88, 063405 (2013).

    ADS  Google Scholar 

  31. 31.

    de Miranda, M. H. G. et al. Controlling the quantum stereodynamics of ultracold bimolecular reactions. Nat. Phys. 7, 502–507 (2011).

    Google Scholar 

  32. 32.

    Covey, J. P. Enhanced Optical and Electric Manipulation of a Quantum Gas of KRb Molecules. Thesis, Univ. Colorado, https://doi.org/10.1007/978-3-319-98107-9 (2018).

  33. 33.

    Shvarchuck, I. et al. Bose–einstein condensation into nonequilibrium states studied by condensate focusing. Phys. Rev. Lett. 89, 270404 (2002).

    CAS  Google Scholar 

  34. 34.

    Hueck, K. et al. Two-dimensional homogeneous Fermi gases. Phys. Rev. Lett. 120, 060402 (2018).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  35. 35.

    Ni, K.-K. et al. A high phase-space-density gas of polar molecules. Science 322, 231–235 (2008).

    ADS  CAS  Google Scholar 

  36. 36.

    Quéméner, G. & Bohn, J. L. Strong dependence of ultracold chemical rates on electric dipole moments. Phys. Rev. A 81, 022702 (2010).

    ADS  Google Scholar 

  37. 37.

    Bohn, J. L., Cavagnero, M. & Ticknor, C. Quasi-universal dipolar scattering in cold and ultracold gases. New J. Phys. 11, 055039 (2009).

    ADS  Google Scholar 

  38. 38.

    Ticknor, C. Quasi-two-dimensional dipolar scattering. Phys. Rev. A 81, 042708 (2010).

    ADS  Google Scholar 

  39. 39.

    Bohn, J. L. & Jin, D. S. Differential scattering and rethermalization in ultracold dipolar gases. Phys. Rev. A 89, 022702 (2014).

    ADS  Google Scholar 

  40. 40.

    Aikawa, K. et al. Anisotropic relaxation dynamics in a dipolar Fermi gas driven out of equilibrium. Phys. Rev. Lett. 113, 263201 (2014).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  41. 41.

    Babadi, M. & Demler, E. Collective excitations of quasi-two-dimensional trapped dipolar fermions: transition from collisionless to hydrodynamic regime. Phys. Rev. A 86, 063638 (2012).

    ADS  Google Scholar 

  42. 42.

    Giorgini, S., Pitaevskii, L. P. & Stringari, S. Theory of ultracold atomic Fermi gases. Rev. Mod. Phys. 80, 1215 (2008).

    ADS  CAS  Google Scholar 

  43. 43.

    Büchler, H. P. et al. Strongly correlated 2D quantum phases with cold polar molecules: controlling the shape of the interaction potential. Phys. Rev. Lett. 98, 060404 (2007).

    ADS  PubMed  PubMed Central  Google Scholar 

  44. 44.

    Góral, K., Santos, L. & Lewenstein, M. Quantum phases of dipolar bosons in optical lattices. Phys. Rev. Lett. 88, 170406 (2002).

    ADS  PubMed  PubMed Central  Google Scholar 

  45. 45.

    Capogrosso-Sansone, B., Trefzger, C., Lewenstein, M., Zoller, P. & Pupillo, G. Quantum phases of cold polar molecules in 2D optical lattices. Phys. Rev. Lett. 104, 125301 (2010).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  46. 46.

    Cooper, N. R. & Shlyapnikov, G. V. Stable topological superfluid phase of ultracold polar fermionic molecules. Phys. Rev. Lett. 103, 155302 (2009).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  47. 47.

    Gorshkov, A. V. et al. Tunable superfluidity and quantum magnetism with ultracold polar molecules. Phys. Rev. Lett. 107, 115301 (2011).

    ADS  PubMed  PubMed Central  Google Scholar 

  48. 48.

    Potter, A. C., Berg, E., Wang, D. W., Halperin, B. I. & Demler, E. Superfluidity and dimerization in a multilayered system of fermionic polar molecules. Phys. Rev. Lett. 105, 220406 (2010).

    ADS  PubMed  PubMed Central  Google Scholar 

  49. 49.

    Yao, N. Y. et al. Many-body localization in dipolar systems. Phys. Rev. Lett. 113, 243002 (2014).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  50. 50.

    Barbiero, L., Menotti, C., Recati, A. & Santos, L. Out-of-equilibrium states and quasi-many-body localization in polar lattice gases. Phys. Rev. B 92, 180406 (2015).

    ADS  Google Scholar 

  51. 51.

    Zinner, N. T. & Bruun, G. M. Density waves in layered systems with fermionic polar molecules. Eur. Phys. J. D 65, 133–139 (2011).

    ADS  CAS  Google Scholar 

  52. 52.

    Peter, D., Müller, S., Wessel, S. & Büchler, H. P. Anomalous behavior of spin systems with dipolar interactions. Phys. Rev. Lett. 109, 025303 (2012).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  53. 53.

    Dalfovo, F., Giorgini, S., Pitaevskii, L. P. & Stringari, S. Theory of Bose–Einstein condensation in trapped gases. Rev. Mod. Phys. 71, 463–512 (1999).

    ADS  CAS  Google Scholar 

  54. 54.

    Inguscio, M., Ketterle, W. & Salomon, C. Ultra-cold Fermi gases. In Proceedings of the International School of Physics ‘Enrico Fermi’ (2007).

Download references

Acknowledgements

We acknowledge funding from NIST, DARPA DRINQS, ARO MURI and NSF Phys-1734006. We thank J. L. Bohn, A. M. Kaufman, and C. Miller for careful reading of the manuscript and T. Brown for technical assistance.

Author information

Affiliations

Authors

Contributions

All authors contributed to carrying out the experiments, interpreting the results, and writing the manuscript.

Corresponding authors

Correspondence to Giacomo Valtolina or Jun Ye.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks Georgy Shlyapnikov and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Layer occupancy.

Histogram of the average number per layer (relative population) for the data shown in Fig. 1c.

Extended Data Fig. 2 Trend of ωx/(2π) versus γ.

Grey points are the experimental measurements at EDC = 5 kV cm−1, the solid grey line is a linear fit to guide the eye, and the dashed line is the prediction (Sim) from the finite-element model. All error bars are 1 standard deviation of the mean.

Extended Data Fig. 3 Evaporation sequence.

a, Ramp in EDC. b, Ramp in γ. c, Trap depth versus time from the finite-element model of electro-optical potential. d, Evolution of η, calculated by taking the ratio of the trap depth and temperature at each time point. e, Evolution of T/TF during the ramp. All error bars are 1 standard error of the mean.

Extended Data Fig. 4 Fermi gas thermometry.

Trend of Tout/Trel as a function of the excluded region from the centre of the Gaussian fit for T/TF = 0.81(15) (orange diamonds) and T/TF = 2.0(1) (black circles). Solid lines are Gaussian fits to simulated density profiles for T/TF = 2.0 (black) and T/TF = 0.8 (orange). All error bars are 1 standard error of the mean.

Rights and permissions

Reprints and Permissions

About this article

Verify currency and authenticity via CrossMark

Cite this article

Valtolina, G., Matsuda, K., Tobias, W.G. et al. Dipolar evaporation of reactive molecules to below the Fermi temperature. Nature 588, 239–243 (2020). https://doi.org/10.1038/s41586-020-2980-7

Download citation

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing