This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

thanks: These authors contributed equally to this work.thanks: These authors contributed equally to this work.

Observation of heat scaling across a first-order quantum phase transition in a spinor condensate

H.-Y. Liang    L.-Y. Qiu    Y.-B. Yang    H.-X. Yang    T. Tian    Y. Xu [email protected]    L.-M. Duan [email protected] Center for Quantum Information, IIIS, Tsinghua University, Beijing 100084, PR China
Abstract

Heat generated as a result of the breakdown of an adiabatic process is one of the central concepts of thermodynamics. In isolated systems, the heat can be defined as an energy increase due to transitions between distinct energy levels. Across a second-order quantum phase transition (QPT), the heat is predicted theoretically to exhibit a power-law scaling, but it is a significant challenge for an experimental observation. In addition, it remains elusive whether a power-law scaling of heat can exist for a first-order QPT. Here we experimentally observe a power-law scaling of heat in a spinor condensate when a system is linearly driven from a polar phase to an antiferromagnetic phase across a first-order QPT. We experimentally evaluate the heat generated during two non-equilibrium processes by probing the atom number on a hyperfine energy level. The experimentally measured scaling exponents agree well with our numerical simulation results. Our work therefore opens a new avenue to experimentally and theoretically exploring the properties of heat in non-equilibrium dynamics.

In quantum mechanics, at zero temperature, when we drive an isolated system by tuning a system parameter, if the driving rate is so slow such that the process is adiabatic, the transition between energy levels cannot occur, and heat cannot be created. Yet, when the system undergoes a second-order QPT during the process, the relaxation time diverges and thus adiabaticity cannot be maintained. As a result, transition between energy levels does occur, producing the heat Polkovnikov2008PRL ; Polkovnikov2008NP . In fact, across the transition point, the physics can be described by the quantum Kibble-Zurek mechanism (KZM) and universal scaling laws for various quantities, such as the temporal onset of excitations, the density of defects and the heat, are predicted polkovnikov2011colloquium ; de2010adiabatic . While important aspects of the quantum KZM have been experimentally observed chen2011quantum ; Chapman2016PRL ; clark2016universal ; zhang2017defect ; Nature2019Lukin ; Qiueaba7292 , the experimental measurement of the heat still remains a significant challenge.

Such non-equilibrium dynamics is of crucial importance ranging from cosmology to condensed matter kibble1980some ; zurek1985cosmological ; PhysRevLett.95.035701 ; zurek2005dynamics ; PhysRevB.72.161201 . Yet the existence of scaling laws is not limited to non-equilibrium dynamics across a second-order QPT. It has been predicted that the scaling of some quantities can also occur across a first-order QPT where multiple phases coexist panagopoulos2015off ; Zhong2017PRE ; coulamy2017dynamics ; pelissetto2017dynamic ; Shimizu_2018 . In particular, very recently, the KZM has been generalized to account for a power-law scaling of the temporal onset of spin excitations present in a spinor condensate across the first-order QPT Qiueaba7292 . The generalized KZM (GKZM) has also been experimentally observed in a spinor condensate Qiueaba7292 . Similar to a second-order QPT, it is natural to ask whether the heat can still exhibit a power-law scaling across the first-order QPT.

A spinor Bose-Einstein condensate (BEC), described by a vector order parameter, provides a controllable platform to explore non-equilibrium dynamics, and various interesting relevant phenomena have been experimentally observed sadler2006spontaneous ; PhysRevLett.103.250401 ; PhysRevLett.108.035301 ; PhysRevLett.105.090402 ; PhysRevA.84.063625 ; parker2013direct ; Yang2019PRA ; Tian2020PRL . In some parameter regime for the condensate, the spin and spatial degrees of freedom are decoupled because all spin states have the same spatial wave function under the single-mode approximation Ueda2012 . As a consequence, the physics is significantly simplified so that the spin degrees of freedom can be separately studied. For an antiferromagnetic (AFM) sodium condensate, there is a first-order QPT between a polar phase with atoms all occupying the mF=0m_{F}=0 level and an AFM phase with atoms equally occupying the mF=±1m_{F}=\pm 1 levels, where mFm_{F} is the magnetic quantum number. The system thus provides an ideal platform to study the non-equilibrium physics across a first-order QPT.

Here we theoretically and experimentally demonstrate the existence of a power-law scaling of the heat in a sodium spinor condensate for two dynamical processes: a one-way process where a system is driven from a polar phase to an AFM phase and a cyclic process where a system ends up at the initial polar phase. For the one-way process, the power-law scaling is well characterized by the GKZM. In experiments, we prepare an initial condensate in the polar phase and then slowly vary the quadratic Zeeman energy qq by controlling magnetic and microwave fields to realize the two non-equilibrium processes. Since the energy gap vanishes at the transition point, adiabaticity cannot be maintained no matter how the system is driven, leading to the appearance of excitations as well as heat, which can be used as a measure of how strongly adiabaticity is broken. Based on Refs. Polkovnikov2008PRL ; Polkovnikov2008NP , the heat density can be defined as an energy increase per atom relative to the ground state at the final quadratic Zeeman energy qfq_{f} over the entire process. In experiments, it can be evaluated by measuring the atom number occupying the mF=0m_{F}=0 level owing to a simple approximate relation between the heat density and the particle number when |qf||q_{f}| is large.

Refer to caption
Figure 1: Schematic illustration of the quench protocol and theoretical demonstration of the heat scaling with respect to the quench rate. (a) The phase diagram depicted by ρ0\langle\rho_{0}\rangle versus the quadratic Zeeman energy qq. Two linear quench protocols are considered: a one-way protocol where qq is varied from qi>0q_{i}>0 to qf<0q_{f}<0 and a cyclic protocol where qq is changed from qi>0q_{i}>0 to qm<0q_{m}<0 and then back to qf>0q_{f}>0. (b) The scaling of the heat density Qo(qf)Q_{o}(q_{f}) with the quench rate vv for a one-way process with qi=c2q_{i}=c_{2} and qf=0.3c2q_{f}=-0.3c_{2}. (c) The scaling exponents ηo\eta_{o} of Qo(qf)Q_{o}(q_{f}) when qfq_{f} is set to a range of different values. (d) The scaling of the heat density Qo(qa)Q_{o}(q_{a}) with the quench rate vv for a one-way process with qi=c2q_{i}=c_{2}. Here qa=vtaq_{a}=-vt_{a} is the corresponding quadratic Zeeman energy at the critical time tat_{a} (see the discussion in the text). In (b-d), the blue diamonds (purple crosses), the red circles (green triangles) and the yellow squares (cyan asterisks) represent the theoretical results for Qo(qf)Q_{o}(q_{f}) [Qo(qa)Q_{o}(q_{a})] obtained by the numerical simulation, the GKZM and the KZM, respectively. (e) The scaling of the heat density Qc(qf)Q_{c}(q_{f}) versus the quench rate vv for a cyclic process with qf=6c2q_{f}=6c_{2} and qm=2.5c2q_{m}=-2.5c_{2}, where blue circles and red diamonds correspond to qi=6c2q_{i}=6c_{2} and qi=c2q_{i}=c_{2}, respectively. (f) The scaling exponents of the heat density for different qfq_{f} for a cyclic process with qi=c2q_{i}=c_{2} and qm=2.5c2q_{m}=-2.5c_{2}. Here N=1.0×104N=1.0\times 10^{4}.

We start by considering the following Hamiltonian describing a spin-1 BEC under single-mode approximation

H^=c2𝐋^22N+mF=11(qmF2pmF)a^mFa^mF,\hat{H}=c_{2}\frac{\hat{\bf L}^{2}}{2N}+\sum_{m_{F}=-1}^{1}(qm_{F}^{2}-pm_{F})\hat{a}^{\dagger}_{m_{F}}\hat{a}_{m_{F}}, (1)

where a^mF\hat{a}_{m_{F}}^{\dagger} (a^mF\hat{a}_{m_{F}}) is the Boson creation (annihilation) operator for an atom in the hyperfine level |F=1,mF|F=1,m_{F}\rangle, 𝐋^\hat{\bf L} is the total spin operator with L^μ=i,ja^i(fμ)ija^j\hat{L}_{\mu}=\sum_{i,j}\hat{a}^{\dagger}_{i}(f_{\mu})_{ij}\hat{a}_{j} (μ=x,y,z\mu=x,y,z) and fμf_{\mu} being the spin-1 angular momentum matrix, c2c_{2} is the spin-dependent interaction (c2>0c_{2}>0 for sodium atoms) and NN is the total number of atoms. Here pp and qq describe linear and quadratic Zeeman energies, respectively. In our experiments, we initialize our condensates in the polar phase so that the dynamics is restricted to the eigenspace of Lz=0L_{z}=0, since the magnetization along zz is conserved, i.e., [L^z,H^]=0[\hat{L}_{z},\hat{H}]=0. The linear Zeeman term pp thus becomes irrelevant in the dynamics even though its value is not equal to zero in the experiment. So the ground states of the spinor condensate correspond to the polar and AFM phases when q>0q>0 and q<0q<0, respectively. The phase diagram is shown in Fig. 1(a) where the mean value ρ0\langle\rho_{0}\rangle with ρ0=a^0a^0/N\rho_{0}=\hat{a}_{0}^{\dagger}\hat{a}_{0}/N taken as an order parameter drops to zero from one at q=0q=0, indicating the occurrence of the first-order QPT.

Refer to caption
Figure 2: A comparison between the scaling of the heat density and the quasi-heat density. (a) The relative difference of the quasi-heat density Q~o(qf)\widetilde{Q}_{o}(q_{f}) compared to the heat density Qo(qf)Q_{o}(q_{f}) when |qf||q_{f}| increases for a fixed quench rate v=0.02c22v=0.02c_{2}^{2}. A comparison between the scaling of the heat density Q(qf)Q(q_{f}) and the quasi-heat density Q~(qf)\widetilde{Q}(q_{f}) for (b) a one-way process and (c-d) cyclic processes. In (b), qi=c2q_{i}=c_{2} and qf=3c2q_{f}=-3c_{2}, in (c), qi=qf=6c2q_{i}=q_{f}=6c_{2} and qm=2.5c2q_{m}=-2.5c_{2}, and in (d), qi=c2q_{i}=c_{2}, qf=6c2q_{f}=6c_{2} and qm=2.5c2q_{m}=-2.5c_{2}. Here N=1.0×104N=1.0\times 10^{4}.

We investigate the heat production in two types of quench processes: a one-way process for linearly ramping qq from qiq_{i} (qi>0q_{i}>0) to qfq_{f} (qf<0q_{f}<0) and a cyclic process for linearly ramping qq from qiq_{i} to qmq_{m} (qm<0q_{m}<0) and then back to qfq_{f}, forming a cyclic process when qf=qiq_{f}=q_{i} [see Fig. 1(a)]. In both scenarios, we calculate the energy increase at the end of the quench for different ramp rates vv. To numerically determine the energy of a spinor condensate, we diagonalize the Hamiltonian under the Fock state basis in the subspace of zero magnetization, {|N/2,0,N/2,|N/21,2,N/21,,|0,N,0}\{|{N}/{2},0,{N}/{2}\rangle,|{N}/{2}-1,2,{N}/{2}-1\rangle,...,|0,N,0\rangle\}, yielding instantaneous eigenstates |ϕn(q)|\phi_{n}(q)\rangle (n=1,2,,N/2+1)(n=1,2,...,{N}/{2}+1) of H^(q)\hat{H}(q) satisfying H^(q)|ϕn(q)=En(q)|ϕn(q)\hat{H}(q)|\phi_{n}(q)\rangle=E_{n}(q)|\phi_{n}(q)\rangle. We also solve the Schrödinger equation i|Ψ(t)/t=H^(t)|Ψ(t)i\hbar\partial|\Psi(t)\rangle/\partial t=\hat{H}(t)|\Psi(t)\rangle (we take h=1h=1 as a natural unit) to determine the evolving state of the spinor condensate. The energy per atom at the end of the quench is given by Ψ(qf)|H^(qf)|Ψ(qf)=HI/N+qfqfρ0\langle\Psi(q_{f})|\hat{H}(q_{f})|\Psi(q_{f})\rangle=\langle H_{I}\rangle/N+q_{f}-q_{f}\langle\rho_{0}\rangle with HI=c2𝐋^2/(2N)H_{I}=c_{2}\hat{\bf L}^{2}/(2N) characterizing the interactions. Since the corresponding ground state energy per atom in the AFM phase is qfq_{f}, the produced heat per atom over the one-way process is given by Qo=HI/Nqfρ0Q_{o}=\langle H_{I}\rangle/N-q_{f}\langle\rho_{0}\rangle. For the cyclic process, the produced heat per atom is Qc=HI/N+qf(1ρ0)Q_{c}=\langle H_{I}\rangle/N+q_{f}(1-\langle\rho_{0}\rangle) since the ground state energy of the polar phase is zero.

Refer to caption
Figure 3: Experimentally observed power-law scaling of the quasi-heat density Q~o\widetilde{Q}_{o} with respect to the ramp rate vv for a one-way quench. In (a-c), we consider processes with qi10Hzq_{i}\approx 10\,\textrm{Hz} and qf=21.60Hzq_{f}=-21.60\,\textrm{Hz}, 24.47Hz-24.47\,\textrm{Hz} and 28.36Hz-28.36\,\textrm{Hz}, respectively (see Appendix A for the discussion of the error arising from the calibration of qq). for a spinor condensate with about 11001100 atoms corresponding to c2=8.1±0.9Hzc_{2}=8.1\pm 0.9\,\textrm{Hz} (See Appendix B for details on how to experimentally evaluate the value of c2c_{2} and its error). In (d), qq is varied from qi=14.33Hzq_{i}=14.33\,\textrm{Hz} to qf=29.11Hzq_{f}=-29.11\,\textrm{Hz} for a BEC with about 30003000 atoms and c2=11.8±0.8Hzc_{2}=11.8\pm 0.8\,\textrm{Hz}. The log-log plot of the experimental data are shown as black squares with error bars. The fitting of the data by a power-law function (black dashed line) gives the exponents of 0.35±0.040.35\pm 0.04 in (a), 0.35±0.030.35\pm 0.03 in (b), 0.35±0.060.35\pm 0.06 in (c) and 0.33±0.060.33\pm 0.06 in (d) with 95%95\% confidence interval. Each figure has an inset showing the theoretical results of the scaling of Q~o\widetilde{Q}_{o} obtained by the numerical simulation (blue squares), the GKZM (red circles) and the KZM (green triangles). The GKZM with exponents of 0.3300.330, 0.3290.329, 0.3300.330 and 0.3310.331 gives a better account of the power-law scaling with exponents of 0.3400.340, 0.3400.340, 0.3420.342 and 0.3370.337 obtained by the numerical simulation than the corresponding exponents of 0.3050.305, 0.3050.305, 0.3060.306 and 0.3090.309 obtained by the KZM.

In Fig. 1(b), we plot our numerical simulation results of the heat density QoQ_{o} for a one-way process, remarkably showing the existence of a power-law scaling, i.e., QovηQ_{o}\propto v^{\eta} with η=0.319\eta=0.319. This power-law scaling persists even when qfq_{f} is far away from the transition point, but the exponents change as a function of qfq_{f} and increase very slowly when |qf||q_{f}| is large [see Fig. 1(c)]. The exponents are independent of qiq_{i} given that two distinct qisq_{i}^{\prime}s are connected by an adiabatic process. For a cyclic process, we also observe the power-law scaling of the heat density QcQ_{c} as shown in Fig. 1(e). While the scaling does not depend on qiq_{i} for the same reason, Fig. 1(f) shows that the scaling exponents increase slightly with increasing qfq_{f} (but they are irrelevant of qmq_{m}).

It is a well-known fact that the universal scaling laws across a second-order QPT are accounted for by the quantum KZM PhysRevLett.95.035701 ; zurek2005dynamics ; PhysRevB.72.161201 . Its essential basis is the existence of impulse and adiabatic regions. Specifically, suppose at t=0t=0, q=qc=0q=q_{c}=0 and the system is in the polar phase. When we linearly drive the system into the AFM phase, the system cannot respond (impulse region) until the response time τ(ta)=1/Δ(ta)=ta\tau(t_{a})=1/\Delta(t_{a})=t_{a}, where Δ\Delta is the relevant energy gap. When t>tat>t_{a}, adiabaticity is restored (adiabatic region). For a second-order QPT, the relevant energy gap refers to the gap between the ground state and the first excited state. Based on the KZM, the heat induced by a slow quench across the critical point is shown to exhibit a power-law dependence on the ramp rate with the scaling exponent determined by the equilibrium critical exponents polkovnikov2011colloquium ; de2010adiabatic ; You2018PRA .

For a first-order QPT, we have demonstrated the existence of impulse and adiabatic regions when a spinor condensate is linearly driven across the transition point Qiueaba7292 . Yet, in stark contrast to the KZM, the relevant energy gap is the gap between the maximally occupied state (the metastable state) and its corresponding first excited state in the first-order case. For example, when q<0q<0, the metastable state refers to the many-body metastable polar phase Qiueaba7292 . We now apply the GKZM to determine the heat scaling. To be more precise, we use the equation |1/Δ(t)|=|Δ(t)/Δ˙(t)||1/\Delta(t)|=|\Delta(t)/\dot{\Delta}(t)| to calculate the critical time tat_{a} and the corresponding qa=vtaq_{a}=-vt_{a}. Since the evolving state is frozen to the initial state when t<tat<t_{a}, the heat density can be evaluated by Qo(qf)=[nPnEn(qf)Eg(qf)]/NQ_{o}(q_{f})=[\sum_{n}P_{n}E_{n}(q_{f})-E_{g}(q_{f})]/N, where Pn=|ϕn(qa)|Ψ(qi)|2P_{n}=|\langle\phi_{n}(q_{a})|\Psi(q_{i})\rangle|^{2} is the probability that the initial state |Ψ(qi)|\Psi(q_{i})\rangle occupies the eigenstate |ϕn(qa)|\phi_{n}(q_{a})\rangle of H^(qa)\hat{H}(q_{a}) corresponding to the eigenenergy En(qa)E_{n}(q_{a}).

In Fig. 1(b), we show the power-law scaling of the heat density Qo(qf)Q_{o}(q_{f}) with respect to vv calculated by the GKZM, which agrees very well with the numerical simulation results. Figure 1(c) also displays their comparison of scaling exponents versus qfq_{f}, showing very good agreement with less than 1.5%1.5\% discrepancy. In comparison, we further compute the heat scaling based on the KZM, which exhibits conspicuous discrepancy especially for large |qf||q_{f}| as shown in Fig. 1(c). For example, when qf=20c2q_{f}=-20c_{2}, the scaling exponent obtained by the KZM has about 8.6%8.6\% difference from the numerical simulation results, while the GKZM only exhibits about 1.0%1.0\% difference. This indicates that the GKZM gives a better account of the heat scaling law at a first-order QPT.

In fact, based on the GKZM, the heat per atom produced during a quench process ending at q=qaq=q_{a} is given by qa-q_{a}, implying that the heat scaling is determined by the scaling of qaq_{a}. In Fig. 1(d), we display the scaling obtained by the numerical simulation, the GKZM and the KZM, demonstrating that the GKZM gives a closer prediction of the power-law exponent to the numerical simulation result than the KZM.

To experimentally probe the heat density is a formidable task due to the complexity of the spin interactions, which is hard to measure. Fortunately, for the one-way process, when |qf||q_{f}| is large, the heat density QoQ_{o} is dominated by the second part, Q~o=qfρ0\widetilde{Q}_{o}=-q_{f}\langle\rho_{0}\rangle, which can be experimentally evaluated by measuring ρ0\langle\rho_{0}\rangle and qfq_{f}. We call Q~o\widetilde{Q}_{o} the quasi-heat density to distinguish it from the heat density QoQ_{o}. Similarly, for the cyclic process, we define the corresponding quasi-heat density as Q~c=qf(1ρ0)\widetilde{Q}_{c}=q_{f}(1-\langle\rho_{0}\rangle). For both processes, the heat density and the quasi-heat density are related by the following equation

Qs=Q~s+HI/NQ_{s}=\widetilde{Q}_{s}+\langle H_{I}\rangle/N (2)

with s=o,cs=o,c referring to the one-way and cyclic processes, respectively.

Figure 2(a) shows the decline of the relative difference between the heat density QoQ_{o} and the quasi-heat density Q~o\widetilde{Q}_{o} when |qf||q_{f}| is increased; the relative difference decreases to less than 20%20\% when qf=3c2q_{f}=-3c_{2} when v=0.02c22v=0.02c_{2}^{2}. In fact, the scalings determined by these two energy increases agree much better than their energy differences even for not very large |qf||q_{f}|, which is experimentally achievable. For instance, when qf=3c2q_{f}=-3c_{2}, while there exists 18%18\% difference of Q~o\widetilde{Q}_{o} compared to QoQ_{o}, their scaling exponents are in excellent agreement with only less than 3%3\% discrepancy for the one-way process [see Fig. 2(b)]. For the cyclic process when qf=6c2q_{f}=6c_{2}, the scaling exponent difference is also smaller than 3%3\% [see Fig. 2(c) and (d)]. This allows us to obtain the scaling of the heat density by experimentally measuring ρ0\langle\rho_{0}\rangle for relatively large |qf||q_{f}|.

Refer to caption
Figure 4: Experimentally observed power-law scaling of the quasi-heat density Q~c\widetilde{Q}_{c} with respect to the quench rate for cyclic processes. In (a-c), qq is linearly changed from qi15Hzq_{i}\approx 15\,\textrm{Hz} to qm=15.19Hzq_{m}=-15.19\,\textrm{Hz} in (a) [qm=22.00Hzq_{m}=-22.00\,\textrm{Hz} in (b) and qm=15.49Hzq_{m}=-15.49\,\textrm{Hz} in (c)] and then back to the final value qf=60.10Hzq_{f}=60.10\,\textrm{Hz} [qf=66.45Hzq_{f}=66.45\,\textrm{Hz} in (c)] for a condensate with about 11001100 atoms corresponding to c2=8.1±0.9Hzc_{2}=8.1\pm 0.9\,\textrm{Hz}. In (d), the atom number N=3000N=3000 corresponding to c2=11.8±0.8Hzc_{2}=11.8\pm 0.8\,\textrm{Hz}, qi=17.55Hzq_{i}=17.55\,\textrm{Hz}, qm=18.10Hzq_{m}=-18.10\,\textrm{Hz} and qf=66.73Hzq_{f}=66.73\,\textrm{Hz}. The experimental data are plotted in the logarithmic scale, which are fitted by power-law functions, giving the exponents of 0.33±0.040.33\pm 0.04 in (a), 0.32±0.030.32\pm 0.03 in (b), 0.33±0.060.33\pm 0.06 in (c) and 0.34±0.050.34\pm 0.05 in (d) under 95%95\% confidence interval. The inset of each figure shows the numerical simulation results of QcQ_{c} (blue circles) and Q~c\widetilde{Q}_{c} (red squares).

In experiments, we produce a sodium BEC in an all-optical trap by evaporation of atoms Yang2019PRA . At the evaporation cooling stage, we apply a strong magnetic field gradient to remove the atoms on the hyperfine levels |F=1,mF=±1|F=1,m_{F}=\pm 1\rangle out of the trap, leaving all atoms on the |F=1,mF=0|F=1,m_{F}=0\rangle level. After that, a weak and nearly resonant microwave pulse is applied to excite the atoms from |F=1,mF=0|F=1,m_{F}=0\rangle to |F=2,mF=0|F=2,m_{F}=0\rangle (the corresponding detuning is δ6kHz\delta\simeq-6\,\textrm{kHz}). Since the atoms on the latter level suffer a significant loss due to three body decay, this process kicks many atoms out of the BEC cloud, resulting in less than 30003000 atoms remaining in the trap. The reduction of the atom number allows us to avoid the unwanted relaxation to the AFM ground state when qq is tuned to negative values. The atoms are then immersed in a uniform magnetic field with qi10Hzq_{i}\approx 10\,\textrm{Hz} for 2s2\,\textrm{s} to equilibrate to the polar phase. Afterwards, we slowly decrease the magnetic field strength so as to linearly vary qq according to the relation qB2277Hz/G2q\approx B^{2}\cdot 277\,\textrm{Hz}/\textrm{G}^{2}. When qq is changed to around 5Hz5\,\textrm{Hz}, we immediately switch on a microwave field with a frequency of 1.7701264GHz1.7701264\,\textrm{GHz} (the detuning is δ=1.5MHz\delta=-1.5\,\textrm{MHz} relative to the transition from |F=1,mF=0|F=1,m_{F}=0\rangle to |F=2,mF=0|F=2,m_{F}=0\rangle) and then gradually raise its amplitude so that qq is linearly driven to the negative regime YouLi2017Science (Appendix A). During the process, the amplitude of the microwave field is precisely controlled by a proportional-integral-derivative (PID) feedback system according to a careful calibration of qq’s values. For each quench rate vv, ρ0\rho_{0} is measured by a standard Stern-Gerlach fluorescence imaging at the end of the linear quench in each experiment and ρ0\langle\rho_{0}\rangle is evaluated by averaging over 40 repeated measurements (Appendix C).

In Fig. 3, we show the experimental results of the quasi-heat density Q~o\widetilde{Q}_{o} with respect to the ramp rate vv for the one-way quench with four different sets of quench parameters. In (a-c), we drive a BEC with about 11001100 atoms and c2=8.1±0.9Hzc_{2}=8.1\pm 0.9\,\textrm{Hz} to three distinct qfq_{f}. The experimental data clearly demonstrate the existence of a power-law scaling for these different conditions. We fit the data by a power-law function, i.e., Q~ovη~\widetilde{Q}_{o}\propto v^{\widetilde{\eta}}, giving the fitting exponents of η~=0.35±0.04\widetilde{\eta}=0.35\pm 0.04, η~=0.35±0.03\widetilde{\eta}=0.35\pm 0.03 and η~=0.35±0.06\widetilde{\eta}=0.35\pm 0.06, respectively. The results agree well with the numerical simulation results with the power-law fitting exponents of 0.3400.340, 0.3400.340 and 0.3420.342, respectively [see the insets of Fig. 3]. Here the numerically calculated exponents for Q~o\widetilde{Q}_{o} exhibit about 10%10\% difference from the exponents of QoQ_{o}, which is larger than the result shown in Fig. 2(b) due to larger ramp rates considered here to reduce the relaxation to the ground states in experiments (Appendix D). In Fig. 3(d), we further plot the experimental results for a BEC with roughly 30003000 atoms and c2=11.8±0.8Hzc_{2}=11.8\pm 0.8\,\textrm{Hz}, showing the existence of a power-law scaling with the fitting exponent of 0.33±0.060.33\pm 0.06, which is in good agreement with the numerically obtained exponent of 0.3370.337.

For a cyclic quench, similar to the one-way one, we initially prepare the condensate in the polar phase with the quadratic Zeeman energy qiq_{i} (e.g., qi15Hzq_{i}\approx 15\,\textrm{Hz}) provided by a uniform magnetic field and then linearly decrease qq to roughly 5Hz5\,\textrm{Hz} by decreasing the magnetic field strength. After that, we shine a microwave field to the BEC to linearly vary qq from about 5Hz5\,\textrm{Hz} to qmq_{m} (qm<0q_{m}<0, e.g., qm=22Hzq_{m}=-22\,\textrm{Hz}) and then back to 5Hz5\,\textrm{Hz} by controlling the field amplitude. We then turn off the microwave field and raise the magnetic field strength until qq slowly rises to qfq_{f} (e.g., qf=60.1Hzq_{f}=60.1\,\textrm{Hz}). Since the results do not depend on the value of qiq_{i} when it is sufficiently large so that the dynamics is adiabatic under the quench rate at q=qiq=q_{i} [see Fig. 1(e)], in experiments, we use qiq_{i} and qfq_{f} with qf>qiq_{f}>q_{i} for experimental convenience. The entire ramping process is precisely controlled to be linear according to the calibration of qq. Similarly, at the end of each quench, the quasi-heat density Q~\widetilde{Q} is measured by probing ρ0\langle\rho_{0}\rangle through the Stern-Gerlach fluorescence imaging.

For a cyclic quench, in Fig. 4, we show the experimental measurement of the quasi-heat density Q~c\widetilde{Q}_{c} as a function of the quench rate vv under different quench parameters. For a BEC cloud with about 11001100 atoms and c2=8.1±0.9Hzc_{2}=8.1\pm 0.9\,\textrm{Hz}, the results shown in Fig. 4(a-c) evidently illustrate a power-law scaling of the quasi-heat density with fitting exponents of 0.33±0.040.33\pm 0.04, 0.32±0.030.32\pm 0.03 and 0.33±0.060.33\pm 0.06, respectively. The exponents agree well with the exponents of 0.3480.348, 0.3520.352 and 0.3520.352 numerically obtained for Q~c\widetilde{Q}_{c}, which are larger than 0.3300.330, 0.3300.330 and 0.3350.335 (numerically calculated scaling exponents for QcQ_{c}) by about 5.4%5.4\%, 6.7%6.7\% and 5.1%5.1\%, respectively. This also shows that our experimental measurements cannot differentiate the slight difference between Q~c\widetilde{Q}_{c} and QcQ_{c}. Additionally, we raise the atom number to around 30003000 corresponding to c2=11.8±0.8Hzc_{2}=11.8\pm 0.8\,\textrm{Hz} and perform the experiments under the quench parameters of qi=17.55Hzq_{i}=17.55\,\textrm{Hz}, qm=18.10Hzq_{m}=-18.10\,\textrm{Hz} and qf=66.73Hzq_{f}=66.73\,\textrm{Hz}. Figure 4(d) reveals the existence of a power-law scaling with a fitting exponent of 0.34±0.050.34\pm 0.05, in good agreement with the numerical result of 0.3450.345.

Our work demonstrates the first experimental observation of a power-law scaling of heat with respect to a ramp rate for non-equilibrium dynamics. Two types of quench processes including one-way and cyclic processes are studied across a first-order QPT in a spinor condensate. The experimentally measured scaling exponents for both non-equilibrium processes agree well with our numerical simulation results.

Acknowledgements.
We thank Yingmei Liu, Ceren Dag, and Anjun Chu for helpful discussions. This work was supported by the Beijing Academy of Quantum Information Sciences, the National key Research and Development Program of China (2016YFA0301902), Frontier Science Center for Quantum Information of the Ministry of Education of China, and Tsinghua University Initiative Scientific Research Program. Y. Xu also acknowledges the support from the start-up fund from Tsinghua University, the National Natural Science Foundation of China (11974201) and the National Thousand-Young-Talents Program.

Appendix A: Calibration of the quadratic Zeeman energy qq

In the experiment, the quadratic Zeeman energy qq is contributed by both the magnetic and microwave fields so that

q=qB+qM,q=q_{B}+q_{M}, (A1)

where qBq_{B} and qMq_{M} are generated by the magnetic and microwave field, respectively. Specifically, qBq_{B} is determined by the magnetic field strength BB through qBB2277(Hz/G2)q_{B}\approx B^{2}\cdot 277\,(\textrm{Hz}/\textrm{G}^{2}). Based on the relation pB700(KHz/G)p\approx B\cdot 700\,(\textrm{KHz/G}) with pp being the linear Zeeman energy, BB can be measured by probing pp through a Rabi oscillation between the Zeeman energy level |F=1,mF=0|F=1,m_{F}=0\rangle and |F=2,mF=1|F=2,m_{F}=-1\rangle. A microwave field at large frequency detuning shifts the energy of the Zeeman levels due to the AC Stark effect and contributes to the quadratic Zeeman energy as

qM=ΔEmF=+1+ΔEmF=12ΔEmF=02q_{M}=\frac{\Delta E_{m_{F}=+1}+\Delta E_{m_{F}=-1}-2\Delta E_{m_{F}=0}}{2} (A2)

with

ΔEmF=14k=1,0,+1ΩmF,mF+k2ΔmF,mF+k,\Delta E_{m_{F}}=\frac{1}{4}\sum_{k=-1,0,+1}\frac{\Omega^{2}_{m_{F},m_{F}+k}}{\Delta_{m_{F},m_{F}+k}}, (A3)

where ΩmF,mF+k\Omega_{m_{F},m_{F}+k} is the Rabi frequency for the resonant transition from |F=1,mF|F=1,m_{F}\rangle to |F=2,mF+k|F=2,m_{F}+k\rangle and ΔmF,mF+k\Delta_{m_{F},m_{F}+k} is the microwave detuning relative to the transition between these energy levels.

In the experiment, the magnetic field strength is controlled by a voltage VBV_{B} that determines the magnitude of the current flowing through the Helmholtz coils. In Fig. A1(a), we show the measured qBq_{B} as a function of the voltage VBV_{B}, which is well fitted by a parabola function, thus allowing us to linearly change qBq_{B} by controlling VBV_{B}.

To determine the microwave quadratic Zeeman energy qMq_{M}, we experimentally evaluate the Rabi frequencies Ω0,1\Omega_{0,-1}, Ω0,1\Omega_{0,1} and Ω1,1\Omega_{-1,-1} by probing the Rabi oscillations for a resonant transition between the two corresponding energy levels. Other Rabi frequencies can be calculated according to the following relations

Ω0,1=3Ω1,0,Ω1,2=6Ω1,0,\displaystyle\Omega_{0,1}=\sqrt{3}\Omega_{-1,0},\ \Omega_{1,2}=\sqrt{6}\Omega_{-1,0}, (A4)
Ω1,1=Ω1,1=32Ω0,0,\displaystyle\Omega_{1,1}=\Omega_{-1,-1}=\frac{\sqrt{3}}{2}\Omega_{0,0}, (A5)
Ω0,1=3Ω1,0,Ω1,2=6Ω1,0.\displaystyle\Omega_{0,-1}=\sqrt{3}\Omega_{1,0},\ \Omega_{-1,-2}=\sqrt{6}\Omega_{1,0}. (A6)

We now fix the microwave’s frequency at 1.7701264GHz1.7701264\,\textrm{GHz} with a detuning Δ0,0=1500kHz\Delta_{0,0}=-1500\,\textrm{kHz} for the transition from |F=1,mF=0|F=1,m_{F}=0\rangle to |F=2,mF=0|F=2,m_{F}=0\rangle. qMq_{M} is then calculated based on Eq. (A2). Since the magnetic field still exists with qB5.0Hzq_{B}\approx 5.0\,\textrm{Hz} when the microwave pulse is applied, the total quadratic Zeeman energy q=qM+qBq=q_{M}+q_{B}. Because the Rabi frequencies depend on the microwave field amplitude, we can control the amplitude to vary qMq_{M} and qq. In our experiment, we stabilize the amplitude of the microwave pulse by a PID system and calibrate the values of qq at several different microwave amplitudes controlled by the feedback voltage VfV_{f}. The measured qq with respect to VfV_{f} is displayed in Fig. A1(b) with a parabola fitting to the data allowing for a linear ramp of qq by tuning VfV_{f}.

Refer to caption
Figure A1: (Color online) Experimental calibration of the quadratic Zeeman energy qq. (a) Experimentally measured qBq_{B} (black crosses) as a function of the voltage VBV_{B}. (b) Experimentally measured qq (black crosses) as a function of the feedback voltage VfV_{f} of the PID system. The experimental data are fitted by the red dashed parabola curves.

To estimate the error between the calibrated qq and the true value of qq, we apply a sudden quench method to measure the transition point qc=0q_{c}=0. Specifically, we initialize the condensate in the polar phase with qi>0q_{i}>0 and then suddenly change qq to qfq_{f}. After 500ms500\,\textrm{ms}’ evolution, we probe the atom populations on the mF=0m_{F}=0 state. If qf>0q_{f}>0, then all atoms should remain on the mF=0m_{F}=0 state; otherwise, if qf<0q_{f}<0, atoms will show up on the mF=±1m_{F}=\pm 1 level. In the experiment, the sudden quench of qq is realized by switching on the microwave field with the frequency of 1.7701264GHz1.7701264\,\textrm{GHz} and a certain amplitude controlled by VfV_{f}. Since qfq_{f} decreases as VfV_{f} increases [see Fig. A1(b)], we can find a maximum VfmaxV_{f}^{\textrm{max}} so that all the atoms stay on the mF=0m_{F}=0 state and a minimum VfminV_{f}^{\textrm{min}} so that atoms begin to show up on the mF=±1m_{F}=\pm 1 states. The calibrated value of the transition point qcq_{c} is thus qcali=[q(Vfmin)+q(Vfmax)]/2q_{\textrm{cali}}=[q(V_{f}^{\textrm{min}})+q(V_{f}^{\textrm{max}})]/2, resulting in a calibration error of δq=qcaliqc=qcali\delta q=q_{\textrm{cali}}-q_{c}=q_{\textrm{cali}} for each set of data. We summarize the calibration error δq\delta q for 21 days’ measurements in Table A1.

q(Vfmax)q(V_{f}^{max})(Hz) q(Vfmin)q(V_{f}^{min})(Hz) qcaliq_{cali}(Hz) δq\delta q(Hz)
-1.03 -1.11 -1.07
0.43 0.35 0.39
0.15 0.07 0.11
-0.62 -0.85 -0.74
-0.83 -1.06 -0.95
0.35 0.28 0.32
-1.16 -1.37 -1.27
0.13 0.06 0.10
-0.12 -0.17 -0.15
0.20 0.13 0.17
-1.43 -1.63 -1.53 0.45±0.70-0.45\pm 0.70
-1.23 -1.44 -1.34
0.40 0.33 0.37
0.29 0.22 0.26
-1.04 -1.25 -1.15
0.31 0.20 0.26
-1.13 -1.34 -1.24
0.32 0.25 0.29
0.01 -0.27 -0.13
-1.32 -1.53 -1.43
-0.64 -0.84 -0.74
Table A1: Summary of the measured values of q(Vfmax)q(V_{f}^{\textrm{max}}) and q(Vfmin)q(V_{f}^{\textrm{min}}) obtained through sudden quench experiments performed during 21 days. The calibrated value of qq in each row is obtained by qcali=(q(Vfmax)+q(Vfmin))/2q_{\textrm{cali}}=(q(V_{f}^{\textrm{max}})+q(V_{f}^{\textrm{min}}))/2, giving the calibration error δq=q¯cali+σ\delta q=\bar{q}_{\textrm{cali}}+\sigma where q¯cali\bar{q}_{\textrm{cali}} is the mean value of qcaliq_{\textrm{cali}} and σ\sigma is the standard deviation.

Appendix B: Measurement of c2c_{2}

The spin-dependent interaction coefficient c2c_{2} can be measured by observing the spin oscillation, i.e., time evolution of ρ0\rho_{0} when the condensate is initially prepared to the state |ρ1,ρ0,ρ1|\rho_{1},\rho_{0},\rho_{-1}\rangle with ρ1=ρ1\rho_{1}=\rho_{-1} and 0<ρ0<10<\rho_{0}<1. According to the mean-field theory, where the quantum fluctuations are neglected and the operators are replaced by their expectation values, the spin-mixing dynamical equations for the spin-1 condensate are given by You2005PRA

ρ˙0=2c2ρ0(1ρ0)2m2sinθ,\dot{\rho}_{0}=\frac{2c_{2}}{\hbar}\rho_{0}\sqrt{(1-\rho_{0})^{2}-m^{2}}\,\textrm{sin}\theta, (B1)
θ˙=2q+2c2(12ρ0)+(2c2)(1ρ0)(12ρ0)m2(1ρ0)2m2cosθ,\dot{\theta}=-\frac{2q}{\hbar}+\frac{2c_{2}}{\hbar}(1-2\rho_{0})+(\frac{2c_{2}}{\hbar})\frac{(1-\rho_{0})(1-2\rho_{0})-m^{2}}{\sqrt{(1-\rho_{0})^{2}-m^{2}}}\,\textrm{cos}\theta, (B2)

where m=ρ1ρ1m=\rho_{1}-\rho_{-1} is the magnetization and θ=θ++θ2θ0\theta=\theta_{+}+\theta_{-}-2\theta_{0} is the relative phase. By fixing the quadratic Zeeman energy qq and m0m\approx 0, we simulate the time evolution of ρ0\rho_{0} to find the value of c2c_{2} that best fits the experimental results. c2c_{2} is then obtained by averaging over 5 measurements for 5 distinct qq. In Fig. B1(a) and (b), we show the experimental and theoretical results of the time evolution of ρ0\rho_{0} with the initial state being |ρ1,ρ0,ρ1=|0.28,0.46,0.26|\rho_{1},\rho_{0},\rho_{-1}\rangle=|0.28,0.46,0.26\rangle for N=1100N=1100 and |ρ1,ρ0,ρ1=|0.30,0.44,0.26|\rho_{1},\rho_{0},\rho_{-1}\rangle=|0.30,0.44,0.26\rangle for N=3000N=3000, respectively. In Table B1, we summarize the five measured results for different atom numbers, giving the mean value of c2c_{2} of 8.1±0.9Hz8.1\,\pm 0.9\,\textrm{Hz} for N=1100N=1100 and 11.8Hz±0.8Hz11.8\,\textrm{Hz}\pm 0.8\,\textrm{Hz} for N=3000N=3000.

Refer to caption
Figure B1: (Color online) Spin oscillation measurements of c2c_{2} by fitting the experimentally observed time evolution of ρ0\rho_{0} (blue squares) by theoretical simulations (red circles). In (a) [(b)], the initial state is prepared to |0.28,0.46,0.26|0.28,0.46,0.26\rangle (|0.30,0.44,0.26|0.30,0.44,0.26\rangle) for a system with about 11001100 (30003000) atoms and q=14.31Hzq=14.31\,\textrm{Hz} (q=14.22Hzq=14.22\,\textrm{Hz}).
q(Hz)q(Hz) θ(×π)\theta(\times\pi) c2(Hz)c_{2}(Hz) c¯2(Hz)\bar{c}_{2}(Hz) N
14.31 0.99 8.0
17.15 0.97 7.4
24.09 1.01 9.8 8.1±0.9Hz8.1\pm 0.9Hz 1100
27.69 1.01 8.2
31.91 1.15 7.3
9.44 0.95 12.6
11.62 0.91 11.4
14.22 0.95 11.4 11.8±0.8Hz11.8\pm 0.8Hz 3000
17.17 0.97 10.8
20.45 1.05 12.9
Table B1: Summary of the datasets for c2c_{2}’s measurements.
Refer to caption
Figure B2: (Color online) Log-log plots of the measured ρ0\rho_{0} (blue diagonal crosses) and their mean values (black squares). (a) A one-way quench process with N1100N\approx 1100, qi=9.96Hzq_{i}=9.96\,\textrm{Hz} and qf=24.47Hzq_{f}=-24.47\,\textrm{Hz}. (b) A cyclic quench process with N1100N\approx 1100, qi=15.04Hzq_{i}=15.04\,\textrm{Hz}, qm=15.19Hzq_{m}=-15.19\,\textrm{Hz} and qf=60.1Hzq_{f}=60.1\,\textrm{Hz}. The red dashed lines denote the linear fits of ρ0\langle\rho_{0}\rangle.

Appendix C: Experimental measurement of ρ0\rho_{0}

In experiments, we measure ρ0\rho_{0} by the standard Stern-Gerlach fluorescence imaging for different ramp rates. Fig. B2(a) shows the measured data (labelled by blue diagonal crosses) of ρ0\rho_{0} for a one-way process where qq is linearly varied from 9.96Hz9.96\,\textrm{Hz} to 24.47Hz-24.47\,\textrm{Hz}. In this case, the atom number NN is restricted to about 11001100 corresponding to a fluorescence count in the range of 2.5×1092.5\times 10^{9} and 2.8×1092.8\times 10^{9}. ρ0\langle\rho_{0}\rangle at each ramp rate is calculated by averaging over measurements repeated 40 times, which is plotted as black squares in the figure. The error bars of ρ0\langle\rho_{0}\rangle (the error in the mean) originate from the quantum fluctuations and the measurement fluctuations and is evaluated as σ/40\sigma/\sqrt{40} StatisticBook , where σ\sigma is the standard deviation of the 40 samples. For the one-way quench process, Q~o=qfρ0\widetilde{Q}_{o}=-q_{f}\,\langle\rho_{0}\rangle and thus the error bar of Q~o\widetilde{Q}_{o} can be evaluated by

δQ~o=(ρ0var+ρ02)(qfvar+qf2)(ρ0qf)2,\delta\widetilde{Q}_{o}=\sqrt{(\langle\rho_{0}\rangle^{\textrm{var}}+\langle\rho_{0}\rangle^{2})(q_{f}^{\textrm{var}}+q_{f}^{2})-(\langle\rho_{0}\rangle\,q_{f})^{2}}, (C1)

where the superscript var denotes the variance of a quantity. Since there is an error of 0.45Hz-0.45\,\textrm{Hz} for qfq_{f}, we make a correction of +0.45Hz+0.45\,\textrm{Hz} to qfq_{f} with the variance qfvar=0.702(Hz)2=0.49(Hz)2q_{f}^{var}=0.70^{2}\,(\textrm{Hz})^{2}=0.49\,(\textrm{Hz})^{2} according to the calibration of qq in Table. A1. In Fig. B2(b), we also display the original data of ρ0\langle\rho_{0}\rangle for a cyclic process, where qq is changed from 15.04Hz15.04\,\textrm{Hz} to 15.19Hz-15.19\,\textrm{Hz} and then back to 60.1Hz60.1\,\textrm{Hz} at different ramp rates for a system with roughly 11001100 atoms.

Appendix D: Finite quench rate effects on scaling exponents

In Fig. D1, we provide the numerical simulation results of QoQ_{o}, showing about 10%10\% difference for Q~o\widetilde{Q}_{o} compared with QoQ_{o}. This discrepancy is larger than the result shown in Fig. 2(b) in the main text. We attribute this discrepancy to the finite ramp rates. As shown in Table D1, the relative difference is larger for a range of quench rates with larger values due to larger contribution of interactions to total energy for a fixed qfq_{f}. In our experiments, both qfq_{f} and quench rate vv that can be taken are limited by the applied microwave field which can induce the relaxation of the condensate to the AFM ground state when its amplitude is strong or it is applied for a long time. To reduce the relaxation effect, we take the minimum qfq_{f} as 29.11Hz-29.11\,\textrm{Hz} and the slowest ramp rate as 10Hz/s10\,\textrm{Hz/s} corresponding to about 3s3\,\textrm{s} for an quench process during which the microwave field is shined to vary the quadratic Zeeman energy.

Refer to caption
Figure D1: (Color online) Comparison of the theoretical results of QoQ_{o} and Q~o\widetilde{Q}_{o} corresponding to Fig. 3 in the main text. The quench parameters are the same as in Fig. 3. The results by numerical simulation, the KZM and the GKZM are plotted as blue squares, green triangles and red circles, respectively. The inset shows the theoretical results of Q~o\widetilde{Q}_{o} as a comparison to those of QoQ_{o}.
v(c22)v(c_{2}^{2}) η\eta η~\widetilde{\eta} |ηη~|/η|\eta-\widetilde{\eta}|/\eta
0.0010.010.001\sim 0.01 0.370 0.373 0.81%\%
0.0110.10.011\sim 0.1 0.351 0.356 1.42%\%
0.110.1\sim 1 0.315 0.343 8.89%\%
Table D1: Finite quench rate effects on the scaling exponents for QoQ_{o} and Q~o\widetilde{Q}_{o} in the one-way quench process with qi=c2q_{i}=c_{2}, qf=3c2q_{f}=-3c_{2} and N=1100N=1100. Different ranges of ramp rates vv give different scaling exponents η\eta, η~\widetilde{\eta} and their relative difference.

References

  • (1) Polkovnikov, A. Microscopic Expression for Heat in the Adiabatic Basis. Phys. Rev. Lett. 101, 220402 (2008).
  • (2) Polkovnikov, A. & Gritsev, V. Breakdown of the adiabatic limit in low-dimensional gapless systems. Nat. Phys. 4, 477-481 (2008).
  • (3) Polkovnikov, A., Sengupta, K., Silva, A. & Vengalattore, M. Colloquium: Nonequilibrium dynamics of closed interacting quantum systems. Rev. Mod. Phys. 83, 863 (2011).
  • (4) De Grandi, C. & Polkovnikov, A. Adiabatic Perturbation Theory: From Landau-Zener Problem to Quenching Through a Quantum Critical Point. In Quantum Quenching, Annealing and Computation pp. 75-114 (Springer, Berlin, Heidelberg, 2010).
  • (5) Chen, D., White, M., Borries, C. & DeMarco, B. Quantum Quench of an Atomic Mott Insulator. Phys. Rev. Lett. 106, 235304 (2011).
  • (6) Anquez, M. et al. Quantum Kibble-Zurek Mechanism in a Spin-1 Bose-Einstein Condensate. Phys. Rev. Lett. 116, 155301 (2016).
  • (7) Clark, L. W., Feng, L. & Chin, C. Universal space-time scaling symmetry in the dynamics of bosons across a quantum phase transition. Science 354, 606-610 (2016).
  • (8) Zhang, J., Cucchietti, F. M., Laflamme, R. & Suter, D. Defect production in non-equilibrium phase transitions: experimental investigation of the Kibble–Zurek mechanism in a two-qubit quantum simulator. New J. Phys. 19, 043001 (2017).
  • (9) Keesling, A. et al. Quantum Kibble–Zurek mechanism and critical dynamics on a programmable Rydberg simulator. Nature 568, 207-211 (2019).
  • (10) Qiu, L.-Y. et al. Observation of generalized Kibble-Zurek mechanism across a first-order quantum phase transition in a spinor condensate. Sci. Adv. 6, eaba7292 (2020).
  • (11) Kibble, T. W. Some implications of a cosmological phase transition. Phys. Rep. 67, 183-199 (1980).
  • (12) Zurek, W. H. Cosmological experiments in superfluid helium? Nature 317, 505-508 (1985).
  • (13) Damski, B. The Simplest Quantum Model Supporting the Kibble-Zurek Mechanism of Topological Defect Production: Landau-Zener Transitions from a New Perspective. Phys. Rev. Lett. 95, 035701 (2005).
  • (14) Zurek, W. H., Dorner, U. & Zoller, P. Dynamics of a quantum phase transition. Phys. Rev. Lett. 95, 105701 (2005).
  • (15) Polkovnikov, A. Universal adiabatic dynamics in the vicinity of a quantum critical point. Phys. Rev. B 72, 161201 (2005).
  • (16) Panagopoulos, H. & Vicari, E. Off-equilibrium scaling behaviors across first-order transitions. Phys. Rev. E 92, 062107 (2015).
  • (17) Liang, N. & Zhong, F. Renormalization-group theory for cooling first-order phase transitions in Potts models. Phys. Rev. E 95, 032124 (2017).
  • (18) Coulamy, I. B., Saguia, A. & Sarandy, M. S. Dynamics of the quantum search and quench-induced first-order phase transitions. Phys. Rev. E 95, 022127 (2017).
  • (19) Pelissetto, A. & Vicari, E. Dynamic off-equilibrium transition in systems slowly driven across thermal first-order phase transitions. Phys. Rev. Lett. 118, 030602 (2017).
  • (20) Shimizu, K., Hirano, T., Park, J., Kuno, Y. & Ichinose, I. Dynamics of first-order quantum phase transitions in extended Bose–Hubbard model: from density wave to superfluid and vice versa. New J. Phys. 20, 083006 (2018).
  • (21) Sadler, L. E., Higbie, J. M., Leslie, S. R., Vengalattore, M. & Stamper-Kurn, D. M. Spontaneous symmetry breaking in a quenched ferromagnetic spinor Bose–Einstein condensate. Nature 443, 312-315 (2006).
  • (22) Leslie, L. S., Hansen, A., Wright, K. C., Deutsch, B. M. & Bigelow, N. P. Creation and Detection of Skyrmions in a Bose-Einstein Condensate. Phys. Rev. Lett. 103, 250401 (2009).
  • (23) Choi, J., Kwon, W. J. & Shin, Y. Observation of Topologically Stable 2D Skyrmions in an Antiferromagnetic Spinor Bose-Einstein Condensate. Phys. Rev. Lett. 108, 035301 (2012).
  • (24) Kronjäger, J., Becker, C., Soltan-Panahi, P., Bongs, K. & Sengstock, K. Spontaneous Pattern Formation in an Antiferromagnetic Quantum Gas. Phys. Rev. Lett. 105, 090402 (2010).
  • (25) Guzman, J. et al. Long-time-scale dynamics of spin textures in a degenerate F=1F=1 87Rb spinor Bose gas. Phys. Rev. A 84, 063625 (2011).
  • (26) Parker, C. V., Ha, L.-C. & Chin, C. Direct observation of effective ferromagnetic domains of cold atoms in a shaken optical lattice. Nat. Phys. 9, 769-774 (2013).
  • (27) Yang, H.-X. et al. Observation of dynamical quantum phase transitions in a spinor condensate. Phys. Rev. A 100, 013622 (2019).
  • (28) Tian, T. et al. Observation of dynamical quantum phase transitions with correspondence in an excited state phase diagram. Phys. Rev. Lett. 124, 043001 (2020).
  • (29) Kawaguchi, Y. & Ueda, M. Spinor Bose-Einstein condensates. Phys. Rep. 520, 253-381 (2012).
  • (30) Xue, M., Yin, S. & You, L. Universal driven critical dynamics across a quantum phase transition in ferromagnetic spinor atomic Bose-Einstein condensates. Phys. Rev. A 98, 013619 (2018).
  • (31) Luo, X.-Y. et al., Deterministic entanglement generation from driving through quantum phase transitions. Science 355, 620 (2017).
  • (32) Zhang, W., Zhou, D. L., Chang, M.-S., Chapman, M. S., and You, L., Coherent spin mixing dynamics in a spin-1 atomic condensate, Phys. Rev. A 72, 013602 (2005).
  • (33) Ross, S. M., Introductory Statistics, Academic Press, Fourth Edition (2017).