Main

In ultracold atomic gases, two-body contact interactions are characterized by the s-wave scattering length a, and the unitary regime is realized in the limit a → ∞, with a tuned using magnetic Feshbach resonances25. In Bose gases, tuning a to infinity also enhances three-body recombination, which leads to particle loss and heating, making unitary Bose gases inherently dynamic, non-equilibrium systems. Experimentally, these systems are studied by rapidly quenching a to infinity (Fig. 1a), which initiates the non-equilibrium dynamics. If starting with a Bose–Einstein condensate (BEC) in the k ≈ 0 momentum state, after the quench the momentum distribution broadens (the kinetic energy increases) owing to lossless correlation dynamics and to recombination heating (Fig. 1b). The interplay between these two processes raises many questions, such as whether the gas attains a strongly correlated quasi-equilibrium steady state before degeneracy is lost.

Fig. 1: Dynamics of a degenerate Bose gas quenched to unitarity.
figure 1

a, Quench protocol. The red circles depict atoms, and their sizes the interaction strength, which is limited at unitarity by the interparticle spacing; a is the s-wave scattering length and thold is the hold time at unitarity. b, Momentum distribution nk(k) for different thold values; the initial gas density is n = 5.1 μm−3, corresponding to a Fermi momentum of kn = 6.7 μm−1 and a Fermi time of tn = 27 μs. c, Populations of individual k states show rapid initial growth, saturation at (quasi-)steady-state values of \({\bar{n}}_{k}(k)\) (dashed lines) and long-time heating. The error bars reflect 1 s.e.m. (not visible when smaller than the symbol size). The solid lines are sigmoid fits used to extract the initial-growth half-way times τ(k).

The timescales of the different processes are set by the natural length scales of the system. Within the universality hypothesis14, in a homogeneous degenerate unitary gas the only relevant length scale is the interparticle spacing n1/3, where n is the particle density, which (in analogy with Fermi gases) sets the Fermi momentum ħkn = ħ(6π2n)1/3, energy \({E}_{n}={\hbar }^{2}{k}_{n}^{2}{\rm{/}}(2m)\) and time tn = ħ/En, where m is the particle mass and ħ is the reduced Planck constant. Additional, potentially relevant length scales are the sizes of the Efimov trimer states that exist as a result of resonant two-body interactions17,26,27,28,29,30,31. Three-body correlations8 and Efimov trimers9 have been observed experimentally, but all degenerate-gas dynamics have been consistent with tn being the only characteristic timescale6,9,10. This universality has so far made it impossible to disentangle the lossless from the recombination-induced dynamics. Experimental evidence has suggested that the lossless processes are faster, sufficiently so that the gas attains a degenerate steady state6,10; however, almost nothing could be established about the nature of this state. Here we isolate the effects of the lossless post-quench dynamics through momentum- and time-resolved studies of degenerate and thermal Bose gases.

We prepare a homogeneous 39K Bose gas in an optical-box trap10 with a volume of around 3 × 104 μm3 and use a Feshbach resonance centred at8 402.70(3) G. Initially, we prepare either a quasi-pure BEC or a thermal gas. In both cases, we start with a weakly interacting system, with na3 < 104, then quench the gas to unitarity (within 2 μs) and let it evolve for a time thold; in our box trap, tn is a global variable and after the quench all parts of the system evolve in the same way. After the time thold, we quench the gas back to low a, release it from the trap and measure its momentum distribution nk(k); we normalize nk so that

$$\int 4{\rm{\pi }}{k}^{2}{n}_{k}{\rm{d}}k=1$$

See Methods for further experimental details.

We first present our study of degenerate gases. In Fig. 1b we show nk(k) for an initial BEC density of n = 5.1 μm−3 and various thold. In Fig. 1c we illustrate our key experimental observation. By looking at nk values for individual k states, we discern separate stages in the evolution of nk: after a rapid initial growth, nk reaches a (quasi-)steady-state plateau, before the long-time heating takes over. All timescales are of order tn, but distinguishable. We discern such time separation for k/kn ≳ 0.8. For each k in this range, we identify the plateau occupation \({\bar{n}}_{k}\) (dashed lines in Fig. 1c) and use sigmoid fits (solid lines) to extract the characteristic time τ(k) for the initial rapid growth of nk, defined such that \({n}_{k}(k,\tau (k))={\bar{n}}_{k}(k)/2\) . We note that tn and kn correspond to the initial n; for our longest τ we observe particle loss of approximately 20%.

In Fig. 1c we also see that the curves for different k values are not aligned in time; nk(2kn) shows signs of heating before nk(kn) reaches its steady-state value. This finding illustrates why lossless and recombination dynamics could not be separated quantitatively by considering all k at the same evolution time6,10, such as by looking at the kinetic energy per particle E(thold)10. Instead, we separately obtain \({\bar{n}}_{k}\) for different k values and piece together the function \({\bar{n}}_{k}(k)\). Doing so does not give the momentum distribution at any specific time, but allows us to infer what the steady-state nk(k) would be if the gas did not suffer from losses and heating. We assume that at early times \(\left({t}_{{\rm{hold}}}={\mathscr{O}}({t}_{n})\right)\), all non-zero-k states are primarily fed from the macroscopically occupied BEC (Fig. 1b).

In Fig. 2 we plot the dimensionless τ/tn and \({\bar{n}}_{k}{k}_{n}^{3}\) versus the dimensionless k/kn, for three BEC densities. By expressing all quantities in dimensionless form, all of our data fall onto universal curves (within experimental errors).

Fig. 2: Universal post-quench dynamics and the steady-state momentum distribution in the degenerate Bose gas.
figure 2

a, b, The momentum-dependent half-way time τ(k) for reaching the steady state (a) and the steady-state momentum distribution \({\bar{n}}_{k}(k)\) (b), for three different BEC densities n. Expressing all quantities in dimensionless form, using the Fermi time tn and momentum kn as the natural scales, collapses all of our data onto universal curves. The error bars show fitting errors (not visible when smaller than the symbol size). The solid line in b is an exponential fit, \({\bar{n}}_{k}{k}_{n}^{3}=1.53{\rm{e}}{\rm{x}}{\rm{p}}(-3.62k/{k}_{n})\).

In the experimentally accessible range of momenta, our data are consistent with the scaling τ/tn ∝ kn/k at low k and τ/tn ∝ (kn/k)2 at high k. These scalings were predicted for the emergence of a prethermal steady state20,21,22,23,24. According to this prediction, at short times after the quench, the excitations are similar to the Bogoliubov modes in a weakly interacting BEC—which are phonons at low k and particles at high k—but with the usual mean-field energy replaced by an energy of order En. The speed of sound is then of order ħkn/m and the crossover between the two regimes is at \(k={\mathscr{O}}({k}_{n})\). Finally, τ(k) is set by the dephasing time, which is given approximately by the inverse of the excitation energy.

The form of the universal \({\bar{n}}_{k}{k}_{n}^{3}\) curve was not anticipated and poses a new theoretical challenge. Empirically, over three orders of magnitude of \({\bar{n}}_{k}{k}_{n}^{3}\), our data are well captured by a simple exponential, Aexp(−Bk/kn), with A = 1.53(5) and B = 3.62(2) (where the errors are 1σ fitting errors). This function implies a condensed fraction of

$$\eta =1-\int 4\pi {k}^{2}{\bar{n}}_{k}{\rm{d}}k=19(4){\rm{ \% }}$$

Up to k ≈ 3kn we do not observe the asymptotic form nk ~ 1/k4 that is expected32 at very high k; however, even if nk changed to this more slowly decaying form immediately outside of our experimental range, η would change by less than 3%. Previous theoretical work22,24 has predicted values of η in the prethermal state that are close to our estimate, but the exponential form of \({\bar{n}}_{k}{k}_{n}^{3}\) has not previously been predicted. Explaining this experimental observation may require explicit consideration of the quench back to low a.

We now turn to thermal gases, which reveal some simplifications, but also more surprises. A simplification is that, while in a thermal gas the three-body recombination and the lossless dynamics are both slowed down compared to the degenerate-gas case, the three-body recombination is slowed down more4,5,33. As shown in Fig. 3a, now E(thold) exhibits two separate stages in the post-quench dynamics: a rapid initial growth (here for thold ≲ 100 μs) and long-time heating (for thold ≫ 100 μs). The shape of the curve is similar to those for individual k states in Fig. 1c and the long-time energy growth matches the theory of recombination heating4,10. These results reinforce our interpretation of the two-step dynamics, both for degenerate and for thermal gases. We now focus on the early-time dynamics. As we show in Fig. 3b, nk(k) is essentially identical at 60 μs and 126 μs, meaning that on this timescale a steady state is established for all k.

Fig. 3: Thermal Bose gas quenched to unitarity.
figure 3

a, The kinetic energy per particle E shows rapid growth for thold ≲ 100 μs and substantial heating only for thold ≫ 100 μs; the black line is the prediction for recombination heating. Here, and in bd, the initial gas density and temperature are n = 5.6 μm−3 and T = 150 nK, respectively. b, Momentum distribution nk(k) for different hold times at unitarity. The initial redistribution of particles from low k to high k (indicated by the dotted arrow) is essentially complete within 60 μs, and nk is almost identical at 126 μs and 60 μs. c, Population changes in different k-space shells, 4πk2δnk(k); the population in k0 = 6.0 μm−1 (dotted line) remains essentially unchanged. d, Vertical cuts through the plot in c. Solid lines are sigmoid fits used to extract the half-way time τ(k). e, f, τ(k) and the change in the spectral energy density (between the initial, pre-quench state and the post-quench steady state), \(\bar{{\rm{\delta }}\varepsilon }(k)\propto {k}^{4}\bar{{\rm{\delta }}{n}_{k}}(k)\). Here we show data for n = 5.6 μm−3 and T = 150 nK (blue) and for n = 1.3 μm−3 and T = 70 nK (red). For the data in a and d, 1 s.e.m. error bars are smaller than the symbol size. In e and f, the error bars (in most cases smaller than the symbol size) show fitting errors.

In a thermal gas, even before the quench to unitarity, nk is substantial for all k ≲ 1/λ, where \(\lambda =h/\sqrt{2{\rm{\pi }}m{k}_{{\rm{B}}}T}\) is the thermal wavelength, T is the initial temperature (before the quench to unitarity), kB is the Boltzmann constant and h = 2πħ. We therefore look at the redistribution of particles in k space, in particular, the change δnk(k) with respect to thold = 0 and the corresponding change δε in the spectral energy density ε = ħ2/(2m) × 4πk4nk. An additional challenge in understanding the thermal-gas case is that we have two relevant length scales, n1/3 and λ, and it is not a priori clear whether the dynamic and thermodynamic properties can be expressed in terms of dimensionless universal functions.

In Fig. 3c we show time-resolved population changes in different spherical shells in k space, 4πk2δnk. For some special k0 (dotted line), the population remains essentially constant. In Fig. 3d we show vertical cuts through Fig. 3c for k < k0, k = k0 and k > k0. Away from k0, we use sigmoid fits (solid lines) to extract τ(k), both for diminishing and for growing populations. Near k0 we see only a small wiggle in δnk, to which we cannot assign a single timescale.

In Fig. 3e, f we show τ(k) and the steady state \(\bar{{\rm{\delta }}\varepsilon }(k)\) for two different combinations of n and T. The \(\bar{{\rm{\delta }}\varepsilon }(k)\) curve conveys the redistribution of particles from k < k0 to k > k0 and the resulting energy growth

$$\bar{{\rm{\Delta }}E}=\int \bar{{\rm{\delta }}\varepsilon }{\rm{d}}k$$

The dispersive shape of τ(k) was not anticipated and invites further theoretical work. Here, we empirically investigate whether these curves can be scaled into universal dimensionless functions.

For the horizontal scaling we find that the natural scale for k is 1/λ, independent of n. In Fig. 4a we plot τ(k) versus , for 15 combinations of n and T (corresponding to phase-space densities 3 between 0.2 and 2). Similarly, in Fig. 4d we plot \(\overline{{\rm{\delta }}\varepsilon }(k){\rm{/}}\lambda \) versus , so that the area under each curve is still \(\overline{{\rm{\Delta }}E}(n,T)\). In both cases, we see horizontal alignment of all of the curves, with k0 = 4.4/λ.

Fig. 4: Universal dynamic and thermodynamic functions for the thermal Bose gas quenched to unitarity.
figure 4

a, d, Plotting the half-way time τ for reaching the post-quench steady state (a) and the change in the spectral energy density \(\overline{{\rm{\delta }}\varepsilon }\,{\rm{/}}\,\lambda \) (where λ is the thermal wavelength; d) versus horizontally aligns all of our curves for 15 different combinations of the initial gas density n and temperature T (see key). The vertical grey line corresponds to k0. b, Supposing that the characteristic timescale for the dynamics is \({t}_{{\rm{s}}}\propto {t}_{n}^{{\alpha }_{t}}{t}_{\lambda }^{{\beta }_{t}}\), where tλ = ħ/(kBT), we obtain the best data collapse, corresponding to the minimum of σ/σ0 (see text for details), for αt ≈ βt ≈ 1/2 (dashed cross indicates αt = βt = 1/2). This suggests that \({t}_{{\rm{s}}}=\sqrt{{t}_{n}{t}_{\lambda }}\). e, Similarly, for the energy scale \({E}_{{\rm{s}}}\propto {E}_{n}^{{\alpha }_{E}}{({k}_{{\rm{B}}}T)}^{{\beta }_{E}}\), we find αE ≈ 1 and βE ≈ 0, which suggests that Es = En. c, f, The dimensionless \(\tau \,{\rm{/}}\,\sqrt{{t}_{n}{t}_{\lambda }}\) (c) and \(\overline{{\rm{\delta }}\varepsilon }/(\lambda {E}_{n})\) (f) are, to within experimental errors, universal functions of the dimensionless . All error bars (not visible when smaller than the symbol size) show fitting errors.

A more challenging question is whether these n- and T-dependent curves may be collapsed vertically, by scaling them by some time ts(nT) and energy Es(nT). We conjecture that \({t}_{{\rm{s}}}\propto {t}_{n}^{{\alpha }_{t}}{t}_{\lambda }^{{\beta }_{t}}\), where tλ = ħ/(kBT), and similarly \({E}_{{\rm{s}}}\propto {E}_{n}^{{\alpha }_{E}}{({k}_{{\rm{B}}}T)}^{{\beta }_{E}}\), and determine for which αt,E and βt,E we get the best collapse. We treat αt,E and βt,E as independent, but physically (if there are no other relevant scales) we expect αt + βt = αE + βE = 1.

We quantify the degree of the data collapse by a single number σ, which is obtained by calculating the standard deviation of the data for all n and T at a fixed and then summing over . In Fig. 4b, e, we show plots of σ/σ0 for τ and \(\overline{{\rm{\delta }}\varepsilon }\,{\rm{/}}\,\lambda \); here, σ0 corresponds to no scaling.

For the temporal scaling, in Fig. 4b we find the lowest σ near αt = βt = 1/2, which suggests that \({t}_{{\rm{s}}}=\sqrt{{t}_{n}{t}_{\lambda }}\). In Fig. 4c we plot \(\tau \,{\rm{/}}\,\sqrt{{t}_{n}{t}_{\lambda }}\) and see that all of our data collapse onto a universal curve (within experimental scatter). For this scaling we have an intuitive interpretation. In a thermal gas, particles do not overlap, so to feel the unitary interactions after the quench they must first meet. The ts that we find, \(\sqrt{{t}_{n}{t}_{\lambda }}\propto {n}^{-1/3}\lambda m/\hbar \), matches the expected scaling for the characteristic time until meeting, which is given by the ratio of the interparticle spacing n−1/3 and the characteristic thermal velocity ħ/().

In Fig. 4e we find that the optimal values of αE and βE are αE ≈ 1 and βE ≈ 0, suggesting that Es = En. This scaling implies that, surprisingly, whereas \(\overline{{\rm{\delta }}\varepsilon }(k)\) naturally depends on n and T, its integral \(\overline{{\rm{\Delta }}E}\) is independent of T; in Fig. 4f we see that this scaling collapses all of our data onto a universal curve.

This lack of T dependence suggests that \(\overline{{\rm{\Delta }}E}\,{\rm{/}}\,{E}_{n}\) in a thermal gas should also be equal to \(\bar{E}\,{\rm{/}}\,{E}_{n}\) in a degenerate gas (where \(\overline{{\rm{\Delta }}E}=\bar{E}\)). Bearing in mind the caveat that we do not observe very high-k tails experimentally, from the data in Fig. 4f we estimate that \(\overline{{\rm{\Delta }}E}\,{\rm{/}}\,{E}_{n}=0.7(1)\) for a thermal gas; from the exponential \({\bar{n}}_{k}{k}_{n}^{3}\) in Fig. 2b, we obtain a consistent value of \(\bar{E}\,{\rm{/}}\,{E}_{n}=0.74(4)\) for a degenerate gas.

Our experiments establish a comprehensive view of the prethermal dynamics and thermodynamics of homogeneous Bose gases quenched to unitarity, at low and high temperatures. They provide quantitative benchmarks and new questions for the theory of unitary Bose gases. Open problems include explaining the forms of our experimentally observed universal dynamic and thermodynamic functions, and elucidating the connections between these universal features and previously observed signatures8,9 of non-universal Efimov physics. Experimentally, an important future challenge is to probe the coherence and the potential superfluid properties of the prethermal state of a degenerate unitary Bose gas.

While this paper was under review, we learned of two other experiments that observe universality in the many-body dynamics of out-of-equilibrium quantum systems34,35.

Methods

Optical-box trap and sample preparation

As described previously36,37, our box trap is formed by blue-detuned, 532-nm laser beams. It is cylindrical in shape, with a diameter of about 30 μm and a length of about 45 μm. We deduce n from the measured atom number, and take into account the fact that the trap walls are not infinitely steep36, owing to the diffraction limit on the sharpness of the laser beams, so the effective trap volume depends slightly on the energy per particle in the initially prepared sample.

Our clouds are in the lowest hyperfine ground state and we initially prepare them at a field of approximately 399.1 G. At this field, the scattering length is ai ≈ 400a0, where a0 is the Bohr radius.

Quench protocol and measurement details

At the end of thold we quench a back to ai using an exponential field ramp with a time constant of 1 μs. We use the fastest ramp that is technically possible to minimize the conversion of atoms into molecules9,10. We then release the gas from the trap and simultaneously (within about 3 ms) completely turn off interactions (a → 0). After letting the cloud expand for 6–12 ms of time of fight, we take an absorption image of it. We typically repeat each measurement about 20 times. To reconstruct nk(k) from the two-dimensional absorption images, which give the momentum distribution integrated along the line of sight, we average each image azimuthally, then average over the experimental repetitions, and finally perform the inverse Abel transform. Owing to the initial cloud size and non-infinite time of flight, our measurements of nk(k) are not quantitatively reliable for k < 2 μm−1.

Extrapolation of \({\bar{{\boldsymbol{n}}}}_{{\boldsymbol{k}}}{{\boldsymbol{k}}}_{{\boldsymbol{n}}}^{{\bf{3}}}\) in a degenerate gas

We also use our experimental data to estimate how the function \({\bar{n}}_{k}{k}_{n}^{3}\) extrapolates to lower k/kn, without presuming its functional form. For k/kn < 0.8, we do not see clear steady-state plateaux in nk(thold), such as indicated by the dashed lines in Fig. 1c. However, we can extrapolate τ ∝ tnkn/k according to the dashed line in Fig. 2a; then, assuming that heating effects are not yet substantial at thold = τ(k) and following our definition of τ, we estimate \({\bar{n}}_{k}=2{n}_{k}(\tau )\), where nk(τ) is the nk measured at the extrapolated τ. These extrapolated values of \({\bar{n}}_{k}{k}_{n}^{3}\) are shown by open symbols in Extended Data Fig. 1. They fall on the same exponential curve that fits our directly measured values of \({\bar{n}}_{k}{k}_{n}^{3}\) (solid symbols), lending further support for this unexpected functional form.