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

Dissipative time crystal in a strongly interacting Rydberg gas

Xiaoling Wu State Key Laboratory of Low Dimensional Quantum Physics, Department of Physics, Tsinghua University, Beijing 100084, China    Zhuqing Wang State Key Laboratory of Low Dimensional Quantum Physics, Department of Physics, Tsinghua University, Beijing 100084, China    Fan Yang Center for Complex Quantum Systems, Department of Physics and Astronomy, Aarhus University, DK-8000 Aarhus C, Denmark    Ruochen Gao State Key Laboratory of Low Dimensional Quantum Physics, Department of Physics, Tsinghua University, Beijing 100084, China    Chao Liang State Key Laboratory of Low Dimensional Quantum Physics, Department of Physics, Tsinghua University, Beijing 100084, China    Meng Khoon Tey State Key Laboratory of Low Dimensional Quantum Physics, Department of Physics, Tsinghua University, Beijing 100084, China Frontier Science Center for Quantum Information, Beijing 100084, China Hefei National Laboratory, Hefei, Anhui 230088, China    Xiangliang Li lixl@baqis.ac.cn Beijing Academy of Quantum Information Sciences, Beijing 100193, China    Thomas Pohl thomas.pohl@itp.tuwien.ac.at Institute for Theoretical Physics, TU Wien, Wiedner Hauptstraße 8-10/136, A-1040 Vienna, Austria    Li You lyou@mail.tsinghua.edu.cn State Key Laboratory of Low Dimensional Quantum Physics, Department of Physics, Tsinghua University, Beijing 100084, China Frontier Science Center for Quantum Information, Beijing 100084, China Hefei National Laboratory, Hefei, Anhui 230088, China Beijing Academy of Quantum Information Sciences, Beijing 100193, China
Abstract

The notion of spontaneous symmetry breaking has been well established to characterize classical and quantum phase transitions of matter, such as in condensation, crystallization or quantum magnetism. Generalizations of this paradigm to the time dimension can lead to a time crystal phase, which spontaneously breaks the time translation symmetry of the system. Whereas the existence of a continuous time crystal at equilibrium has been challenged by no-go theorems, this difficulty can be circumvented by dissipation in an open system. Here, we report the experimental observation of such dissipative time crystalline order in a room-temperature atomic gas, where ground-state atoms are continuously driven to Rydberg states. The emergent time crystal is revealed by persistent oscillations of the photon transmission, and we show that the observed limit cycles arise from the coexistence and competition between distinct Rydberg components. The nondecaying autocorrelation of the oscillation, together with the robustness against temporal noises, indicate the establishment of true long-range temporal order and demonstrates the realization of a continuous time crystal.

preprint: APS/123-QED

The search for emergent many-body phases is among the central objectives of quantum physics sondhi1997continuous . While the concept of phase transitions in thermal equilibrium is well developed, the presence of driving and dissipation can result in a rich phenomenology that has no counterpart in equilibrium eisert2015quantum , such as ubiquitous self-organization effects in physics, biology, and economics haken2006information . In particular, such nonequilibrium processes can facilitate a novel dynamical phase that spontaneously breaks time translation symmetry kessler2019emergent ; buvca2019non ; dogra2019dissipation ; dreon2022self , commonly referred to as a time crystal shapere2012classical ; sacha2017time ; zhang2017observation ; choi2017observation ; rovny2018observation ; riera2020time ; randall2021many ; kyprianidis2021observation ; kessler2021observation ; mi2022time ; taheri2022all . In analogy to crystals in space, a continuous time crystal (CTC) phase has an order parameter with self-sustained oscillations, even though the system is driven in a continuous manner wilczek2012quantum ; nozieres2013time ; bruno2013impossibility ; watanabe2015absence ; iemini2018boundary ; bakker2022driven ; carollo2022exact ; krishna2023measurement ; nie2023mode . Remarkably, the spontaneous time translation symmetry breaking associated with a dissipative CTC has been recently observed with atomic Bose-Einstein condensate in an optical cavity kongkhambut2022observation . The inevitable loss of atoms in the ultracold regime, however, complicates the investigation of long-time dynamics and thereby hinders the analysis of long-range time crystalline order.

Ensembles of Rydberg atoms represent a suitable platform for exploring many-body phenomena away from equilibrium emerging from coherent driving, dissipation, and long-range dipole-dipole interactions carr2013nonequilibrium ; malossi2014full ; ding2020phase ; wu2021concise ; horowicz2021critical ; franz2022observation . Such a Rydberg gas is well controllable and can be confined in a room-temperature vapour cell at virtually no atom loss. Here, we exploit this feature and report the experimental observation of long-range time crystalline order in a continuously driven Rydberg gas [see Fig. 1a]. In our experiment, the CTC manifests itself in limit cycle oscillations of the Rydberg atom density and the atomic dipole moment, which we probe directly by the transmission of light through the gas [see Fig. 1b]. We demonstrate that the time crystal originates from the competition between atomic excitations in different Rydberg states, and characterize the parameter regimes that support this phase. We verify the observation of a CTC by experimentally demonstrating the establishment of a long-range autocorrelation function and its robustness against noisy temporal perturbations.

Refer to caption
Figure 1: Experimental protocol and mean-field phase diagram. a Schematic of the experimental setup, where a probe and a reference beam are propagating in parallel through a room-temperature 85Rb vapour cell. The probe beam overlaps with a counterpropagating coupling beam, which is redirected by a dichroic mirror (DM) and can drive the atoms to Rydberg states. The quarter-wave plates (QWP) are used to control the polarization of the beams, and the transmission difference between the probe and the reference is detected by a balanced photon detector (PD). b Single-run transmission difference for Rydberg states with a principal quantum number n=75𝑛75n=75italic_n = 75. c The left and the right panel respectively show the level scheme and the mean-field phase diagram for the case of a single Rydberg state |rket𝑟|r\rangle| italic_r ⟩ with χ=16γ𝜒16𝛾\chi=-16\gammaitalic_χ = - 16 italic_γ, which supports a stationary phase (SS) and a bistable phase (BS). d The left and the right panel respectively show the level scheme and the phase diagram for a system involving two Rydberg states |rket𝑟|r\rangle| italic_r ⟩ and |sket𝑠|s\rangle| italic_s ⟩ with χ=16γ𝜒16𝛾\chi=-16\gammaitalic_χ = - 16 italic_γ and δ=8γ𝛿8𝛾\delta=8\gammaitalic_δ = 8 italic_γ. Here, a limit cycle oscillating phase (OSC) is identified. The overlapping region between an OSC phase and a BS phase contains one stationary state and one stable limit cycle.

To understand the origin of the limit cycle in our experiment, we first consider the microscopic description of the driven-dissipative Rydberg gas. As illustrated in Fig. 1c, the applied laser fields generated a coherent coupling between the atomic ground state |gket𝑔|g\rangle| italic_g ⟩ and a Rydberg state |rket𝑟|r\rangle| italic_r ⟩, with a corresponding Rabi frequency ΩΩ\Omegaroman_Ω and frequency detuning ΔΔ\Deltaroman_Δ. The loss of coherence is described by the decay of Rydberg population with a rate γ𝛾\gammaitalic_γ. The strong interaction between Rydberg states is governed by a Hamiltonian H^I=ij(Vij/2)n^in^jsubscript^𝐻Isubscript𝑖𝑗subscript𝑉𝑖𝑗2subscript^𝑛𝑖subscript^𝑛𝑗\hat{H}_{\mathrm{I}}=\sum_{i\neq j}(V_{ij}/2)\hat{n}_{i}\hat{n}_{j}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT roman_I end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i ≠ italic_j end_POSTSUBSCRIPT ( italic_V start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT / 2 ) over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, with n^i=|riri|subscript^𝑛𝑖ketsubscript𝑟𝑖brasubscript𝑟𝑖\hat{n}_{i}=|r_{i}\rangle\langle r_{i}|over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = | italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ⟨ italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | the local Rydberg density and Vij=C6/|𝐫i𝐫j|6subscript𝑉𝑖𝑗subscript𝐶6superscriptsubscript𝐫𝑖subscript𝐫𝑗6V_{ij}=C_{6}/|\mathbf{r}_{i}-\mathbf{r}_{j}|^{6}italic_V start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT / | bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - bold_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT the van der Waals interaction between Rydberg atoms located at 𝐫isubscript𝐫𝑖\mathbf{r}_{i}bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and 𝐫jsubscript𝐫𝑗\mathbf{r}_{j}bold_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT. In a thermal Rydberg gas, however, the atomic motion averages out the associated spatial correlations between Rydberg atoms and permits a mean-field treatment of the interaction carr2013nonequilibrium , whereby the laser detuning ΔΔχnrΔΔ𝜒subscript𝑛𝑟\Delta\rightarrow\Delta-\chi n_{r}roman_Δ → roman_Δ - italic_χ italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT acquires a dependence on the uniform Rydberg density nr=n^isubscript𝑛𝑟delimited-⟨⟩subscript^𝑛𝑖n_{r}=\langle\hat{n}_{i}\rangleitalic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = ⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩, with an effective nonlinearity strength χ𝜒\chiitalic_χ. The finite decay rate γ𝛾\gammaitalic_γ usually relaxes the system to a stationary mixed state (SS), corresponding to the fixed point of the nonlinear optical Bloch equation. However, it has been shown that the interaction induced nonlinearity can cause a saddle-node bifurcation carr2013nonequilibrium , through which a bistable stationary state emerges in a finite region of the parameter space. For an attractive interaction (χ<0𝜒0\chi<0italic_χ < 0), such a bistable phase (BS) can occur at the negative detuning regime [see the right panel of Fig. 1c]. At its boundaries one finds a discontinuous nonequilibrium transition between distinct stationary states with low and high Rydberg-state concentration, respectively.

The situation changes dramatically when more than one Rydberg states come into play. In order to illustrate this effect, we consider a minimal extension, in which one additional Rydberg state |sket𝑠|s\rangle| italic_s ⟩ is coupled with an identical Rabi frequency ΩΩ\Omegaroman_Ω but different detuning ΔδΔ𝛿\Delta-\deltaroman_Δ - italic_δ, [see Fig. 1d]. These distinct Rydberg states can establish their respective bistable phases at different detunings ΔΔ\Deltaroman_Δ for a sufficiently large energy separation δ𝛿\deltaitalic_δ. More importantly, their strong interactions generate nonlinear energy shifts, ENL=χ(nr+ns)subscript𝐸NL𝜒subscript𝑛𝑟subscript𝑛𝑠E_{\mathrm{NL}}=\chi(n_{r}+n_{s})italic_E start_POSTSUBSCRIPT roman_NL end_POSTSUBSCRIPT = italic_χ ( italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ), that couple the dynamics of both Rydberg states (see Methods). The resulting competition can drive a Hopf bifurcation for χδsimilar-to𝜒𝛿\chi\sim\deltaitalic_χ ∼ italic_δ, whereby a non-stationary dynamical phase appears in between the two bistable regions, as illustrated in the right panel of Fig. 1d. In this non-stationary regime, the interaction can facilitate the excitation of one Rydberg state at the cost of the other, leading to limit cycle dynamics with persistent oscillations of the Rydberg densities without damping.

Refer to caption
Figure 2: Transmission spectrum and experimental phase diagram. a Energy level diagram for the Rydberg excitation scheme. The ground state |g=|5S1/2,F=3,mF=3ket𝑔ketformulae-sequence5subscript𝑆12𝐹3subscript𝑚𝐹3|g\rangle=|5S_{1/2},F=3,m_{F}=3\rangle| italic_g ⟩ = | 5 italic_S start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT , italic_F = 3 , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 3 ⟩ is coupled to an intermediate state |e=|5P3/2,F=4,mF=4ket𝑒ketformulae-sequence5subscript𝑃32𝐹4subscript𝑚𝐹4|e\rangle=|5P_{3/2},F=4,m_{F}=4\rangle| italic_e ⟩ = | 5 italic_P start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT , italic_F = 4 , italic_m start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 4 ⟩ by a σ+superscript𝜎\sigma^{+}italic_σ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT polarized 780 nm probe field, and three Rydberg Zeeman sublevels |r=|nD5/2,mJ=1/2ket𝑟ket𝑛subscript𝐷52subscript𝑚𝐽12|r\rangle=|nD_{5/2},m_{J}=1/2\rangle| italic_r ⟩ = | italic_n italic_D start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = 1 / 2 ⟩, |s=|nD5/2,mJ=5/2ket𝑠ket𝑛subscript𝐷52subscript𝑚𝐽52|s\rangle=|nD_{5/2},m_{J}=5/2\rangle| italic_s ⟩ = | italic_n italic_D start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = 5 / 2 ⟩, |q=|nD3/2,mJ=1/2ket𝑞ket𝑛subscript𝐷32subscript𝑚𝐽12|q\rangle=|nD_{3/2},m_{J}=1/2\rangle| italic_q ⟩ = | italic_n italic_D start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = 1 / 2 ⟩ are further connected by a 480 nm coupling field. b Scanned transmission spectrum for the different indicated magnetic fields at elliptical polarized coupling field. The Rabi frequencies for the probe and the coupling fields are Ωp/2π=20subscriptΩ𝑝2𝜋20\Omega_{p}/2\pi=20roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / 2 italic_π = 20 MHz and Ωc/2π=0.7subscriptΩ𝑐2𝜋0.7\Omega_{c}/2\pi=0.7roman_Ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / 2 italic_π = 0.7 MHz (for the transition to |sket𝑠|s\rangle| italic_s ⟩), and the intermediate-state detuning is Δp/2π=30subscriptΔ𝑝2𝜋30\Delta_{p}/2\pi=30roman_Δ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / 2 italic_π = 30 MHz. c Scanned transmission spectrum for the different indicated polarizations at a low magnetic field B=2.35𝐵2.35B=2.35italic_B = 2.35 G. The Rabi frequencies are Ωp/2π=19subscriptΩ𝑝2𝜋19\Omega_{p}/2\pi=19roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / 2 italic_π = 19 MHz and Ωc/2π=1.3subscriptΩ𝑐2𝜋1.3\Omega_{c}/2\pi=1.3roman_Ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / 2 italic_π = 1.3 MHz (for a σ+superscript𝜎\sigma^{+}italic_σ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT circularly polarized coupling beam) with an intermediate-state detuning Δp/2π=94subscriptΔ𝑝2𝜋94\Delta_{p}/2\pi=94roman_Δ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / 2 italic_π = 94 MHz. The scanning rate is 2π×2.842𝜋2.842\pi\times 2.842 italic_π × 2.84 MHz/ms for b and c. d Evolution of the scanned transmission spectrum at increasing Rabi frequency ΩcsubscriptΩ𝑐\Omega_{c}roman_Ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, performed with n=75𝑛75n=75italic_n = 75, Ωp/2π=20MHzsubscriptΩ𝑝2𝜋20MHz\Omega_{p}/2\pi=20~{}\mathrm{MHz}roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / 2 italic_π = 20 roman_MHz and Δp/2π=30MHzsubscriptΔ𝑝2𝜋30MHz\Delta_{p}/2\pi=30~{}\mathrm{MHz}roman_Δ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / 2 italic_π = 30 roman_MHz. The scanning rate is 2π×3.172𝜋3.172\pi\times 3.172 italic_π × 3.17 MHz/ms and the coupling field is linearly polarized. e Extracted phase diagram with the oscillation contrast ratio 𝒞𝒞\mathcal{C}caligraphic_C as the order parameter.

The experimental setup for observing such an oscillatory dynamics is depicted in Fig. 1a, where 85Rb atoms are trapped in a room-temperature vapour cell. The atoms are continuously excited to the Rydberg states via a two-photon process, in which a probe beam and a coupling beam counterpropagates with each other. Here, the probe and the coupling fields drive the ground state manifold |5S1/2,F=3ket5subscript𝑆12𝐹3|5S_{1/2},F=3\rangle| 5 italic_S start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT , italic_F = 3 ⟩ to the Rydberg manifold |nDJket𝑛subscript𝐷𝐽|nD_{J}\rangle| italic_n italic_D start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT ⟩ (J=5/2, 3/2𝐽5232J=5/2,\ 3/2italic_J = 5 / 2 , 3 / 2) via intermediate states |5P3/2,F=4ket5subscript𝑃32𝐹4|5P_{3/2},F=4\rangle| 5 italic_P start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT , italic_F = 4 ⟩. The exact Zeeman level of the states can be specified by the polarization of the beams [see Fig. 2a]. In the limit cycle phase, the population of Rydberg states and the transition dipole moments are coupled with each other to form the persistent oscillation. The oscillation of the transition dipole moment between the ground and the excited state then results in an oscillating polarization on the probe transition, which can be directly monitored by the light transmission. To obtain a clear transmission signal, we use a calcite beam displacer to generate a reference beam parallel to the probe beam for differential measurement. The transmission signal for a high lying Rydberg state with a principal quantum number n=75𝑛75n=75italic_n = 75 is shown in Fig. 1b, which exhibits a stable periodic oscillation pattern. In this single experiment, three Rydberg states close in energy are involved in the driving scheme su2022optimizing , and should all participate in the synchronized oscillating dynamics.

To confirm that the oscillation is indeed induced by the involvement of multiple Rydberg states, we reduce the principal quantum number to n=69𝑛69n=69italic_n = 69 and apply a magnetic field B𝐵{B}italic_B to adjust the energy differences between distinct Zeeman levels. Then, with a careful choice of the polarization, we can study the dynamics mainly involving two Rydberg states close in energy, i.e., states |rket𝑟|r\rangle| italic_r ⟩ and |sket𝑠|s\rangle| italic_s ⟩ in Fig. 2a. The branching ratio of the coupling to these states can be tuned by the polarization of the coupling field su2022optimizing . In order to identify the region displaying limit cycles, we perform a scanning spectroscopy measurement, in which the frequency of the coupling beam is slowly scanned near the two-photon resonance with a fixed intermediate-state detuning ΔpsubscriptΔ𝑝\Delta_{p}roman_Δ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. In the scanning process, the magnetic insensitive Rydberg level |q=|69D3/2,mJ=1/2ket𝑞ket69subscript𝐷32subscript𝑚𝐽12|q\rangle=|69D_{3/2},m_{J}=1/2\rangle| italic_q ⟩ = | 69 italic_D start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT = 1 / 2 ⟩ separated from the two target states can also be laser coupled, and is chosen as a reference state with a detuning ΔqsubscriptΔ𝑞\Delta_{q}roman_Δ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT.

The observed transmission spectrum for different magnetic fields and polarizations are shown in Figs. 2b and 2c. First, we set the polarization of coupling beam to be elliptical, by which both Rydberg states can be excited, and they are distinguishable at a large magnetic field B=4.69G𝐵4.69GB=4.69\ \mathrm{G}italic_B = 4.69 roman_G [see Fig. 2b]. Crucially, oscillations appear only in the case of a relatively small magnetic field, where |rket𝑟|r\rangle| italic_r ⟩ and |sket𝑠|s\rangle| italic_s ⟩ are close enough to induce the competition. Next, we fix the magnetic field at a small value B=2.35G𝐵2.35GB=2.35~{}\mathrm{G}italic_B = 2.35 roman_G and vary the polarization of the coupling light [see Fig. 2c]. We find the oscillating phase only exists at an elliptical polarization, but disappears for a perfectly circularly polarized case σsuperscript𝜎\sigma^{-}italic_σ start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT or σ+superscript𝜎\sigma^{+}italic_σ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, where only one of the Rydberg states |rket𝑟|r\rangle| italic_r ⟩ or |sket𝑠|s\rangle| italic_s ⟩ is excited. Based on this experimental evidence, we can conclude that it is the competition between multiple Rydberg levels facilitating the limit cycle phase, in agreement with the mean-field prediction.

Refer to caption
Figure 3: Establishment of the long-range temporal order. a shows a single-shot realization of the quench dynamics, where the blue line represents the sustained oscillating signal of the transmission, and the red lines represent the intensity of the probe field monitored by an independent PD. b displays the discrete Fourier-transform of the time-dependent transmitted power recorded in a, performed for different time windows 2-32 ms, 50-80 ms, and 90-120 ms (from left to right). c Evolution of the peak frequency by varying duration of the signal starting from 2 ms. The error bar corresponds to the standard deviation of 100 independent realizations and the light blue shading represents the frequency resolution of the DFT. d shows the ACF G(τ)𝐺𝜏G(\tau)italic_G ( italic_τ ) for the time windows considered in b. The coupling scheme is the same as Fig. 2d, with {Ωp,Ωc}/2π={18,0.9}subscriptΩ𝑝subscriptΩ𝑐2𝜋180.9\{\Omega_{p},\Omega_{c}\}/2\pi=\{18,0.9\}{ roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT , roman_Ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT } / 2 italic_π = { 18 , 0.9 } MHz.

In addition to multiple Rydberg components, sufficiently large Rabi frequencies of the probe (ΩpsubscriptΩ𝑝\Omega_{p}roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT) and the coupling (ΩcsubscriptΩ𝑐\Omega_{c}roman_Ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT) are also necessary to induce the limit cycle. Figure 2d displays the transmission spectrum at different ΩcsubscriptΩ𝑐\Omega_{c}roman_Ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT with a large and fixed ΩpsubscriptΩ𝑝\Omega_{p}roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, from which rich oscillating patterns can be identified. First, we note that the oscillation completely disappears if the Rabi frequency is too small. Then, weak and long-period oscillations come into appearance when ΩcsubscriptΩ𝑐\Omega_{c}roman_Ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT slightly exceeds a certain critical point. As the Rabi frequency is further increased, the amplitude of the oscillation as well as its range increases. Choosing the oscillating contrast ratio 𝒞𝒞\mathcal{C}caligraphic_C as the order parameter, we can extract an approximate phase diagram from the transmission spectrum, as shown in Fig. 2e, where limit cycle phase is revealed by a nonvanishing contrast. The measured phase diagram indicates that the limit cycle phase exists over a wide parameter regimes and is not sensitive to experimental conditions, such as laser power and frequency, etc.

Having identified parameter regions that support oscillatory states, we can now study their properties in relation to the physics of time crystals. Specifically, we investigate a quench dynamics, where the probe field is suddenly turned on and then held at a constant strength for a very long time. As a typical realization shown in Fig. 3a, the elementary oscillation pattern is rapidly established within 200μssimilar-toabsent200𝜇s\sim 200~{}\mu\mathrm{s}∼ 200 italic_μ roman_s, and then sustained in the entire driving window of 125ms125ms125~{}\mathrm{ms}125 roman_ms. A detailed analysis of the signal reveals that the early oscillations have a slowly drifting frequency during the initial transient dynamics, which evolve to periodic oscillations with a steady frequency. Figure 3b displays the discrete Fourier transform (DFT) of the transmission signal at different stages. For the time window 2-32 ms, the Fourier spectrum has several broad peaks, indicating the absence of a well-defined periodicity. In contrast, the spectrum features equidistant narrow Fourier peaks for the time window 50-80 ms and keeps its shape at later times 90-120 ms, suggesting that a stable periodic pattern has been reached. We then perform 100 independent measurements and study evolution of the oscillation frequencies. As shown in Fig. 3c, the fundamental frequency drifts towards higher values in the transient region, and then converges to a stable one. These repeated realizations of the quench dynamics confirm that the above evolution towards a stable oscillating periodicity is not a consequence of fluctuations of the experimental conditions but rather an inherent process of the interacting atoms.

We further extract the autocorrelation function (ACF) G(τ)=𝑑tI(t)I(t+τ)𝐺𝜏differential-d𝑡𝐼𝑡𝐼𝑡𝜏G(\tau)=\int dtI(t)I(t+\tau)italic_G ( italic_τ ) = ∫ italic_d italic_t italic_I ( italic_t ) italic_I ( italic_t + italic_τ ), where I(t)𝐼𝑡I(t)italic_I ( italic_t ) is the shifted transmission signal with zero mean [𝑑tI(t)=0differential-d𝑡𝐼𝑡0\int dtI(t)=0∫ italic_d italic_t italic_I ( italic_t ) = 0]. The ACF is the classical estimation of the two-time quantum correlation function medenjak2020isolated ; guo2022quantum ; greilich2023continuous , which can quantify temporal correlations of the dynamics. Figure 3d shows the ACF at the same stages as in Fig. 3b. In the transient stage, the envelope of the ACF exhibits approximately an algebraic decay with delay τ𝜏\tauitalic_τ, in accordance with a quasi long-range order. Remarkably, as the system enters the stable periodic stage, the ACF shows a persistent oscillation with an almost constant amplitude within the considered range of the delay τ𝜏\tauitalic_τ (80similar-toabsent80\sim 80∼ 80 periods). The nondecaying behavior of the ACF here corroborates the establishment of a true long-range order (LRO), which is a strong evidence of the time crystal.

Refer to caption
Figure 4: Robustness against temporal perturbations and melting of the observed time crystal. a Normalized crystalline fraction as a function of the noise strength 𝒩𝒩\mathcal{N}caligraphic_N, defined as 𝒩=ω|Pn(ω)|/ω|P0(ω)|1𝒩subscript𝜔subscript𝑃𝑛𝜔subscript𝜔subscript𝑃0𝜔1\mathcal{N}=\sum_{\omega}|P_{n}(\omega)|/\sum_{\omega}|P_{0}(\omega)|-1caligraphic_N = ∑ start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT | italic_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_ω ) | / ∑ start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT | italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ω ) | - 1, where Pn(ω)subscript𝑃𝑛𝜔P_{n}(\omega)italic_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_ω ) and P0(ω)subscript𝑃0𝜔P_{0}(\omega)italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ω ) correspond to the single-sided amplitude spectrum of the monitored probe intensity with and without noise, respectively. The vertical and horizontal error bars respectively represent the standard deviation of the normalized crystalline fraction Q𝑄Qitalic_Q and the noise strength 𝒩𝒩\mathcal{N}caligraphic_N of 10 independent realizations. The inset illustrates the transmission signal (blue) and the monitored probe intensity (red) at the data point indicated by the dashed line. b The left and the right panels respectively correspond to the modulus of the amplitude spectrum |A(ω)|𝐴𝜔|A(\omega)|| italic_A ( italic_ω ) | and the autocorrelation function G(τ)𝐺𝜏G(\tau)italic_G ( italic_τ ) at the data points indicated by the dashed line. The experimental parameters are the same as in Fig. 3.

As a defining property, a CTC is predicted to be rigid, i.e., robust against temporal perturbations. To test the rigidity of the observed limit cycle, we add intensity noise on the probe light during the quench dynamics. For a small but finite noise strength 𝒩𝒩\mathcal{N}caligraphic_N, we note that the ordered oscillation of the transmission signal will not suddenly disappear, but preserves the basic pattern found in the noise-free case [see the inset of Fig. 4a]. With increasing noise strength, the random noise gradually dominates over the ordered oscillation. To quantify such a melting process of the CTC, we introduce a relative crystalline fraction kongkhambut2022observation , defined as Q=|ωωc|δω|A(ω)|/ω|A(ω)|𝑄subscript𝜔subscript𝜔𝑐𝛿𝜔𝐴𝜔subscript𝜔𝐴𝜔Q={\sum_{|\omega-\omega_{c}|\leq\delta\omega}|A(\omega)}|/{\sum_{\omega}|A(% \omega)}|italic_Q = ∑ start_POSTSUBSCRIPT | italic_ω - italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT | ≤ italic_δ italic_ω end_POSTSUBSCRIPT | italic_A ( italic_ω ) | / ∑ start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT | italic_A ( italic_ω ) |, where ωcsubscript𝜔𝑐\omega_{c}italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and δω𝛿𝜔\delta\omegaitalic_δ italic_ω respectively denote the fundamental frequency and the resolution of the DFT amplitude A(ω)𝐴𝜔A(\omega)italic_A ( italic_ω ). As shown in Fig. 4a, the measured crystalline fraction drops almost linearly with the noise strength 𝒩𝒩\mathcal{N}caligraphic_N in the initial stage and then holds within a certain range of strong noise strengths. To investigate the temporal correlation of the melting time crystal, we extract the DFT spectrum and the ACF for measurements performed in different stages [see Fig. 4b]. In the linear drop zone, the DFT spectrum still has equally separated peaks embedded in a noisy background. The oscillation of the ACF remains persistent before a slight drop at a very large delay τ𝜏\tauitalic_τ, signaling a preserved LRO. Entering the strong noise region, while the DFT amplitude spectrum has a sharp peak at the fundamental frequency, high-frequency components are erased by the noise. Meanwhile, the noise destroys the LRO, as revealed by the decay of the autocorrelation function for τ>0𝜏0\tau>0italic_τ > 0. These observations demonstrate that the long-range temporal correlation can be partially maintained during the melting, confirming that the observed dissipative time crystal is robust to temporal fluctuations of the continuous driving.

In summary, we experimentally demonstrate dissipative time crystalline order in a room-temperature Rydberg atom ensemble, which is directly observed by the oscillatory transmission of the probe beam. The observed oscillation originates from the simultaneous coupling to distinct Rydberg states, by which the competition between different Rydberg components facilitates the emergence of limit cycles. Importantly, the persistent oscillation has true long-range temporal order and is robust against noisy perturbations, fulfilling the fundamental criteria of a CTC. By extending the accessible range of the physical parameters, e.g., the Rabi frequency and the atomic density, it will be interesting to measure the complete phase diagram including bistability phases, which can be distinguished via hysteresis measurements carr2013nonequilibrium . By employing a periodic driving, the current setup also holds promise for the study of a discrete time crystal else2016floquet ; gong2018discrete ; lazarides2020time ; cabot2022metastable ; pal2018temporal ; cosme2019time ; gambetta2019discrete ; tuquero2022dissipative . Realization of the model in a Rydberg-atom tweezer array could allow for further studies of the time-crystalline order, such as the behavior in different spatial dimensions, the role of quantum fluctuations and correlations, as well as their scaling with the system size. Our work provides a realistic interacting many-body system for systematically investigating the dissipative time crystal, and opens up a new route to quantum synchronization khasseh2019many ; buvca2022algebraic and sensing Ilias2022criticality ; cabot2023continuous .

Note added. After completion of this work, we became aware of two recent reports ding2023ergodicity ; wadenpfuhl2023emergence . Ref. ding2023ergodicity investigates transient oscillations due to inhomogeneous Rydberg excitation. Ref. wadenpfuhl2023emergence studies the synchronization of atoms with different velocities and observes oscillations in the transmission spectrum as a function of the control-field detuning in a heated vapor cell away from EIT resonance.

Acknowledgements.
We acknowledge valuable discussions with Drs. Klaus Mølmer, Sebastian Hofferberth, Yanhong Xiao, Yaofeng Chen, Feng Chen, Yuanjiang Tang, Hadi Yarloo, and Huachen Zhang, as well as the support by Prof. Luyi Yang. This work is supported by the National Key R&\&&D Program of China (Grant No. 2018YFA0306504 and No. 2018YFA0306503), the National Natural Science Foundation of China (NSFC) (Grant No. 92265205 and 12361131576), and the Innovation Program for Quantum Science and Technology (2021ZD0302100). F. Yang and T. Pohl acknowledge the support from Carlsberg Foundation through the “Semper Ardens” Research Project QCooL and from the Danish National Research Foundation (DNRF) through the Center of Excellence “CCQ” (Grant No. DNRF156). T. Pohl acknowledges the support from the Austrian Science Fund (FWF) [10.55776/COE1] and the European Union - NextGenerationEU, and the Horizon Europe ERC synergy grant SuperWave (grant no. 101071882).

Author contributions

X.W. constructed the initial experimental set up and observed the primary phenomenon, while Z.W. and X.L. gave crucial assistance. Z.W., R.G., and X.W. performed the experiments and analysed the data after thorough discussions with F.Y. The theoretical model was proposed by X.W., Z.W., F.Y., and T.P. The theoretical analysis and numerical simulations are conducted by F.Y. All authors discussed the results and contributed to the manuscript. X.L., T.P., and L.Y. supervised the study.

Competing interests

The authors declare no competing interests.

References

  • (1) Sondhi, S. L., Girvin, S. M., Carini, J. P. & Shahar, D. Continuous quantum phase transitions. Rev. Mod. Phys. 69, 315 (1997).
  • (2) Eisert, J., Friesdorf, M. & Gogolin, C. Quantum many-body systems out of equilibrium. Nat. Phys. 11, 124–130 (2015).
  • (3) Haken, H. Information and self-organization: A macroscopic approach to complex systems. (2006).
  • (4) Keßler, H., Cosme, J. G., Hemmerling, M., Mathey, L. & Hemmerich, A. Emergent limit cycles and time crystal dynamics in an atom-cavity system. Phys. Rev. A 99, 053605 (2019).
  • (5) Buča, B., Tindall, J. & Jaksch, D. Non-stationary coherent quantum many-body dynamics through dissipation. Nat. Commun. 10, 1730 (2019).
  • (6) Dogra, N., Landini, M., Kroeger, K., Hruby, L., Donner, T. & Esslinger, T. Dissipation-induced structural instability and chiral dynamics in a quantum gas. Science 366, 1496–1499 (2019).
  • (7) Dreon, D., Baumgärtner, A., Li, X., Hertlein, S., Esslinger, T. & Donner, T. Self-oscillating pump in a topological dissipative atom–cavity system. Nature 608, 494–498 (2022).
  • (8) Shapere, A. & Wilczek, F. Classical time crystals. Phys. Rev. Lett. 109, 160402 (2012).
  • (9) Sacha, K. & Zakrzewski, J. Time crystals: a review. Rep. Prog. Phys. 81, 016401 (2017).
  • (10) Zhang, J., Hess, P. W., Kyprianidis, A., Becker, Petra, Lee, A, Smith, J, Pagano, Gaetano, Potirniche, I-D, Potter, Andrew C, Vishwanath, Ashvin, Yao, N.Y. & Monroe, C. Observation of a discrete time crystal. Nature 543, 217–220 (2017).
  • (11) Choi, S.et al. Observation of discrete time-crystalline order in a disordered dipolar many-body system. Nature 543, 221–225 (2017).
  • (12) Rovny, J., Blum, R. L. & Barrett, S. E. Observation of discrete-time-crystal signatures in an ordered dipolar many-body system. Phys. Rev. Lett. 120, 180603 (2018).
  • (13) Riera-Campeny, A., Moreno-Cardoner, M. & Sanpera, A. Time crystallinity in open quantum systems. Quantum 4, 270 (2020).
  • (14) Randall, J. et al. Many-body–localized discrete time crystal with a programmable spin-based quantum simulator. Science 374, 1474–1478 (2021).
  • (15) Kyprianidis, A. et al. Observation of a prethermal discrete time crystal. Science 372, 1192–1196 (2021).
  • (16) Keßler, H., Kongkhambut, P., Georges, C., Mathey, L., Cosme, J. G. & Hemmerich, A. Observation of a dissipative time crystal. Phys. Rev. Lett. 127, 043602 (2021).
  • (17) Mi, X. et al. Time-crystalline eigenstate order on a quantum processor. Nature 601, 531–536 (2022).
  • (18) Taheri, H., Matsko, A. B., Maleki, L. & Sacha, K. All-optical dissipative discrete time crystals. Nat. Commun. 13, 848 (2022).
  • (19) Wilczek, F. Quantum time crystals. Phys. Rev. Lett. 109, 160401 (2012).
  • (20) Nozières, P. Time crystals: Can diamagnetic currents drive a charge density wave into rotation?. Europhys. Lett. 103, 57008 (2013).
  • (21) Bruno, P. Impossibility of spontaneously rotating time crystals: a no-go theorem. Phys. Rev. Lett. 111, 070402 (2013).
  • (22) Watanabe, H. & Oshikawa, M. Absence of quantum time crystals. Phys. Rev. Lett. 114, 251603 (2015).
  • (23) Iemini, F., Russomanno, A., Keeling, J. & Schirò, M., Dalmonte, M. & Fazio, R. Boundary time crystals. Phys. Rev. Lett. 121, 035301 (2018).
  • (24) Bakker, L. R., Bahovadinov, M. S., Kurlov, D. V., Gritsev, V., Fedorov, A. K. & Krimer, D. O. Driven-dissipative time crystalline phases in a two-mode bosonic system with Kerr nonlinearity. Phys. Rev. Lett. 129, 250401 (2022).
  • (25) Carollo, F. & Lesanovsky, I. Exact solution of a boundary time-crystal phase transition: time-translation symmetry breaking and non-markovian dynamics of correlations. Phys. Rev. A 105, L040202 (2022).
  • (26) Krishna, M., Solanki, P., Hajdušek, M. & Vinjanampathy, S. Measurement-Induced Continuous Time Crystals. Phys. Rev. Lett. 130, 150401 (2023).
  • (27) Nie, X. & Zheng, W. Mode softening in time-crystalline transitions of open quantum systems. Phys. Rev. A 107, 033311 (2023).
  • (28) Kongkhambut, P., Skulte, J., Mathey, L., Cosme, J. G., Hemmerich, A. & Keßler, H. Observation of a continuous time crystal. Science 377, 670–673 (2022).
  • (29) Carr, C., Ritter, R., Wade, C. G., Adams, C. S. & Weatherill, K. J. Nonequilibrium phase transition in a dilute Rydberg ensemble. Phys. Rev. Lett. 111, 113901 (2013).
  • (30) Malossi, N., Valado, M. M., Scotto, S., Huillery, P., Pillet, P., Ciampini, D., Arimondo, E. & Morsch, O. Full counting statistics and phase diagram of a dissipative Rydberg gas. Phys. Rev. Lett. 113, 023006 (2014).
  • (31) Ding, D. -S., Busche, H., Shi, B. -S., Guo, G. -C. & Adams, C. S. Phase diagram and self-organizing dynamics in a thermal ensemble of strongly interacting Rydberg atoms. Phys. Rev. X 10, 021023 (2020).
  • (32) Wu, X., Liang, X., Tian, Y., Yang, F., Chen, C., Liu, Y. -C., Tey, M. K. & You, L. A concise review of Rydberg atom based quantum computation and quantum simulation. Chin. Phys. B 30, 020305 (2021).
  • (33) Horowicz, Y., Katz, O., Raz, O. & Firstenberg, O. Critical dynamics and phase transition of a strongly interacting warm spin gas. Proc. Natl. Acad. Sci. U.S.A 118, e2106400118 (2021).
  • (34) Franz, T., Geier, S., Hainaut, C., Thaicharoen, N., Braemer, A., Gärttner, M., Zürn, G. & Weidemüller, M. Observation of universal relaxation dynamics in disordered quantum spin systems. arXiv:2209.08080 (2022).
  • (35) Su, H. -J., Liou, J. -Y., Lin, I. -C. & Chen, Y. -H. Optimizing the Rydberg EIT spectrum in a thermal vapor. Opt. Express 30, 1499–1510 (2022).
  • (36) Medenjak, M., Buča, B. & Jaksch, D. Isolated Heisenberg magnet as a quantum time crystal. Phys. Rev. B 102, 041117 (2020).
  • (37) Guo, T. -C. & You, L. Quantum Phases of Time Order in Many-Body Ground States. Frontiers in Physics 10, (2022).
  • (38) Greilich, A., Kopteva, N. E., Kamenskii, A. N., Sokolov, P. S., Korenev, V. L. & Bayer, M. Robust continuous time crystal in an electron–nuclear spin system. Nat. Phys. 20, 631–636 (2023).
  • (39) Else, D. V., Bauer, B. & Nayak, C. Floquet time crystals. Phys. Rev. Lett. 117, 090402 (2016).
  • (40) Gong, Z., Hamazaki, R. & Ueda, M. Discrete time-crystalline order in cavity and circuit QED systems. Phys. Rev. Lett. 120, 040404 (2018).
  • (41) Lazarides, A., Roy, S., Piazza, F. & Moessner, R. Time crystallinity in dissipative Floquet systems. Phys. Rev. Research 2, 022002 (2020).
  • (42) Cabot, A., Carollo, F. & Lesanovsky, I. Metastable discrete time-crystal resonances in a dissipative central spin system. Phys. Rev. B 106, 134311 (2022).
  • (43) Pal, S., Nishad, N., Mahesh, T. S. & Sreejith, G. J. Temporal order in periodically driven spins in star-shaped clusters. Phys. Rev. Lett. 120, 180602 (2018).
  • (44) Cosme, J. G., Skulte, J. & Mathey, L. Time crystals in a shaken atom-cavity system. Phys. Rev. A 100, 053615 (2019).
  • (45) Gambetta, F. M., Carollo, F., Marcuzzi, M., Garrahan, JP & Lesanovsky, I Discrete time crystals in the absence of manifest symmetries or disorder in open quantum systems. Phys. Rev. Lett. 122, 015701 (2019).
  • (46) Tuquero, R. J. L., Skulte, J., Mathey, L. & Cosme, J. G. Dissipative time crystal in an atom-cavity system: Influence of trap and competing interactions. Phys. Rev. A 105, 043311 (2022).
  • (47) Khasseh, R., Fazio, R., Ruffo, S. & Russomanno, A. Many-body synchronization in a classical hamiltonian system. Phys. Rev. Lett. 123, 184301 (2019).
  • (48) Buča, B., Booker, C. & Jaksch, D. Algebraic theory of quantum synchronization and limit cycles under dissipation. SciPost Phys. 12, 097 (2022).
  • (49) Ilias, T., Yang, D., Huelga, S. F. & Plenio, M. B. Criticality-Enhanced Quantum Sensing via Continuous Measurement. PRX Quantum 3, 010354 (2022).
  • (50) Cabot, A., Carollo, F. & Lesanovsky, I. Continuous sensing and parameter estimation with the boundary time-crystal. Phys. Rev. Lett. 132, 050801 (2024).
  • (51) Ding, D. -S., Bai, Z., Liu, Z. -K., Shi, B. -S., Guo, G. -C., Li, W. & Adams, C. S. Ergodicity breaking from Rydberg clusters in a driven-dissipative many-body system. Sci. Adv 10, 9, eadl5893 (2024).
  • (52) Wadenpfuhl, K. & Adams, C. S. Emergence of synchronisation in a driven-dissipative hot Rydberg vapor. Phys. Rev. Lett. 131, 143002 (2023).

Methods

Mean-field description of the dynamics

To qualitatively understand the observed oscillation, we can consider a simple V-type three-level model shown in Fig. 1d. Then, the time translation invariant Hamiltonian (in the rotating frame) of the model is given by

H^=^𝐻absent\displaystyle\hat{H}=\ over^ start_ARG italic_H end_ARG = Ω2i(σ^igr+σ^igs+H.c.)i(Δrn^ir+Δsn^is)\displaystyle\frac{\Omega}{2}\sum_{i}\left(\hat{\sigma}_{i}^{gr}+\hat{\sigma}_% {i}^{gs}+\mathrm{H.c.}\right)-\sum_{i}\left(\Delta_{r}\hat{n}_{i}^{r}+\Delta_{% s}\hat{n}_{i}^{s}\right)divide start_ARG roman_Ω end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_r end_POSTSUPERSCRIPT + over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_s end_POSTSUPERSCRIPT + roman_H . roman_c . ) - ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( roman_Δ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT + roman_Δ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT )
+12ijVij(n^irn^jr+2n^irn^js+n^isn^js),12subscript𝑖𝑗subscript𝑉𝑖𝑗superscriptsubscript^𝑛𝑖𝑟superscriptsubscript^𝑛𝑗𝑟2superscriptsubscript^𝑛𝑖𝑟superscriptsubscript^𝑛𝑗𝑠superscriptsubscript^𝑛𝑖𝑠superscriptsubscript^𝑛𝑗𝑠\displaystyle+\frac{1}{2}\sum_{i\neq j}V_{ij}\left(\hat{n}_{i}^{r}\hat{n}_{j}^% {r}+2\hat{n}_{i}^{r}\hat{n}_{j}^{s}+\hat{n}_{i}^{s}\hat{n}_{j}^{s}\right),+ divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_i ≠ italic_j end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT + 2 over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT + over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT ) , (1)

where σ^iαβ=|αiβi|superscriptsubscript^𝜎𝑖𝛼𝛽ketsubscript𝛼𝑖brasubscript𝛽𝑖\hat{\sigma}_{i}^{\alpha\beta}=|\alpha_{i}\rangle\langle\beta_{i}|over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_β end_POSTSUPERSCRIPT = | italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ⟨ italic_β start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | is the atomic transition operator (α,β=g,r,sformulae-sequence𝛼𝛽𝑔𝑟𝑠\alpha,\beta=g,r,sitalic_α , italic_β = italic_g , italic_r , italic_s), and n^iα=|αiαi|superscriptsubscript^𝑛𝑖𝛼ketsubscript𝛼𝑖brasubscript𝛼𝑖\hat{n}_{i}^{\alpha}=|\alpha_{i}\rangle\langle\alpha_{i}|over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT = | italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ ⟨ italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | (α=r,s𝛼𝑟𝑠\alpha=r,sitalic_α = italic_r , italic_s) denotes the local Rydberg density. Here, we assume an equal interaction strength Vijsubscript𝑉𝑖𝑗V_{ij}italic_V start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT for different Rydberg levels |rket𝑟|r\rangle| italic_r ⟩ and |sket𝑠|s\rangle| italic_s ⟩. Taking the Rydberg decay into consideration, the evolution of an operator 𝒪^^𝒪\hat{\mathcal{O}}over^ start_ARG caligraphic_O end_ARG is governed by

t𝒪^=i[H^,𝒪^]+r[𝒪^]+s[𝒪^],subscript𝑡delimited-⟨⟩^𝒪delimited-⟨⟩𝑖^𝐻^𝒪superscriptsubscript𝑟delimited-[]^𝒪superscriptsubscript𝑠delimited-[]^𝒪\partial_{t}\langle\hat{\mathcal{O}}\rangle=\langle i[\hat{H},\hat{\mathcal{O}% }]+\mathcal{L}_{r}^{*}[\hat{\mathcal{O}}]+\mathcal{L}_{s}^{*}[\hat{\mathcal{O}% }]\rangle,∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ⟨ over^ start_ARG caligraphic_O end_ARG ⟩ = ⟨ italic_i [ over^ start_ARG italic_H end_ARG , over^ start_ARG caligraphic_O end_ARG ] + caligraphic_L start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT [ over^ start_ARG caligraphic_O end_ARG ] + caligraphic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT [ over^ start_ARG caligraphic_O end_ARG ] ⟩ , (2)

where α[𝒪^]=(γ/2)i(2σ^iαg𝒪^σ^igα{n^iα,𝒪^})superscriptsubscript𝛼delimited-[]^𝒪𝛾2subscript𝑖2superscriptsubscript^𝜎𝑖𝛼𝑔^𝒪superscriptsubscript^𝜎𝑖𝑔𝛼superscriptsubscript^𝑛𝑖𝛼^𝒪\mathcal{L}_{\alpha}^{*}[\hat{\mathcal{O}}]=(\gamma/2)\sum_{i}\left(2\hat{% \sigma}_{i}^{\alpha g}\hat{\mathcal{O}}\hat{\sigma}_{i}^{g\alpha}-\{\hat{n}_{i% }^{\alpha},\hat{\mathcal{O}}\}\right)caligraphic_L start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT [ over^ start_ARG caligraphic_O end_ARG ] = ( italic_γ / 2 ) ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 2 over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_g end_POSTSUPERSCRIPT over^ start_ARG caligraphic_O end_ARG over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_α end_POSTSUPERSCRIPT - { over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT , over^ start_ARG caligraphic_O end_ARG } ) is the Lindbladian describing the decay of the Rydberg state |αket𝛼|\alpha\rangle| italic_α ⟩ (α=r,s𝛼𝑟𝑠\alpha=r,sitalic_α = italic_r , italic_s). The equations of motion for the first moments are then determined by

tn^jr=subscript𝑡delimited-⟨⟩superscriptsubscript^𝑛𝑗𝑟absent\displaystyle\partial_{t}\langle\hat{n}_{j}^{r}\rangle=∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ⟩ = iΩ2[σ^jgrσ^jrg]γn^jr,𝑖Ω2delimited-[]delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑔𝑟delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑟𝑔𝛾delimited-⟨⟩superscriptsubscript^𝑛𝑗𝑟\displaystyle\ i\frac{\Omega}{2}[\langle\hat{\sigma}_{j}^{gr}\rangle-\langle% \hat{\sigma}_{j}^{rg}\rangle]-\gamma\langle\hat{n}_{j}^{r}\rangle,italic_i divide start_ARG roman_Ω end_ARG start_ARG 2 end_ARG [ ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_r end_POSTSUPERSCRIPT ⟩ - ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r italic_g end_POSTSUPERSCRIPT ⟩ ] - italic_γ ⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ⟩ ,
tn^js=subscript𝑡delimited-⟨⟩superscriptsubscript^𝑛𝑗𝑠absent\displaystyle\partial_{t}\langle\hat{n}_{j}^{s}\rangle=∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT ⟩ = iΩ2[σ^jgsσ^jsg]γn^js,𝑖Ω2delimited-[]delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑔𝑠delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑠𝑔𝛾delimited-⟨⟩superscriptsubscript^𝑛𝑗𝑠\displaystyle\ i\frac{\Omega}{2}[\langle\hat{\sigma}_{j}^{gs}\rangle-\langle% \hat{\sigma}_{j}^{sg}\rangle]-\gamma\langle\hat{n}_{j}^{s}\rangle,italic_i divide start_ARG roman_Ω end_ARG start_ARG 2 end_ARG [ ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_s end_POSTSUPERSCRIPT ⟩ - ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s italic_g end_POSTSUPERSCRIPT ⟩ ] - italic_γ ⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT ⟩ ,
tσ^jgr=subscript𝑡delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑔𝑟absent\displaystyle\partial_{t}\langle\hat{\sigma}_{j}^{gr}\rangle=∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_r end_POSTSUPERSCRIPT ⟩ = iΩ2[n^jrσ^jgg+σ^jsr]+i(Δr+iγ2)σ^jgr𝑖Ω2delimited-[]delimited-⟨⟩superscriptsubscript^𝑛𝑗𝑟delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑔𝑔delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑠𝑟𝑖subscriptΔ𝑟𝑖𝛾2delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑔𝑟\displaystyle\ i\frac{\Omega}{2}[\langle\hat{n}_{j}^{r}\rangle-\langle\hat{% \sigma}_{j}^{gg}\rangle+\langle\hat{\sigma}_{j}^{sr}\rangle]+i\left(\Delta_{r}% +i\frac{\gamma}{2}\right)\langle\hat{\sigma}_{j}^{gr}\rangleitalic_i divide start_ARG roman_Ω end_ARG start_ARG 2 end_ARG [ ⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ⟩ - ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_g end_POSTSUPERSCRIPT ⟩ + ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s italic_r end_POSTSUPERSCRIPT ⟩ ] + italic_i ( roman_Δ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_i divide start_ARG italic_γ end_ARG start_ARG 2 end_ARG ) ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_r end_POSTSUPERSCRIPT ⟩
ikjVjk[n^krσ^jgr+n^ksσ^jgr],𝑖subscript𝑘𝑗subscript𝑉𝑗𝑘delimited-[]delimited-⟨⟩superscriptsubscript^𝑛𝑘𝑟superscriptsubscript^𝜎𝑗𝑔𝑟delimited-⟨⟩superscriptsubscript^𝑛𝑘𝑠superscriptsubscript^𝜎𝑗𝑔𝑟\displaystyle-i\sum_{k\neq j}V_{jk}[\langle\hat{n}_{k}^{r}\hat{\sigma}_{j}^{gr% }\rangle+\langle\hat{n}_{k}^{s}\hat{\sigma}_{j}^{gr}\rangle],- italic_i ∑ start_POSTSUBSCRIPT italic_k ≠ italic_j end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT [ ⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_r end_POSTSUPERSCRIPT ⟩ + ⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_r end_POSTSUPERSCRIPT ⟩ ] ,
tσ^jgs=subscript𝑡delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑔𝑠absent\displaystyle\partial_{t}\langle\hat{\sigma}_{j}^{gs}\rangle=∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_s end_POSTSUPERSCRIPT ⟩ = iΩ2[n^jsσ^jgg+σ^jrs]+i(Δs+iγ2)σ^jgs𝑖Ω2delimited-[]delimited-⟨⟩superscriptsubscript^𝑛𝑗𝑠delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑔𝑔delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑟𝑠𝑖subscriptΔ𝑠𝑖𝛾2delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑔𝑠\displaystyle\ i\frac{\Omega}{2}[\langle\hat{n}_{j}^{s}\rangle-\langle\hat{% \sigma}_{j}^{gg}\rangle+\langle\hat{\sigma}_{j}^{rs}\rangle]+i\left(\Delta_{s}% +i\frac{\gamma}{2}\right)\langle\hat{\sigma}_{j}^{gs}\rangleitalic_i divide start_ARG roman_Ω end_ARG start_ARG 2 end_ARG [ ⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT ⟩ - ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_g end_POSTSUPERSCRIPT ⟩ + ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r italic_s end_POSTSUPERSCRIPT ⟩ ] + italic_i ( roman_Δ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + italic_i divide start_ARG italic_γ end_ARG start_ARG 2 end_ARG ) ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_s end_POSTSUPERSCRIPT ⟩
ikjVjk[n^ksσ^jgs+n^krσ^jgs],𝑖subscript𝑘𝑗subscript𝑉𝑗𝑘delimited-[]delimited-⟨⟩superscriptsubscript^𝑛𝑘𝑠superscriptsubscript^𝜎𝑗𝑔𝑠delimited-⟨⟩superscriptsubscript^𝑛𝑘𝑟superscriptsubscript^𝜎𝑗𝑔𝑠\displaystyle-i\sum_{k\neq j}V_{jk}[\langle\hat{n}_{k}^{s}\hat{\sigma}_{j}^{gs% }\rangle+\langle\hat{n}_{k}^{r}\hat{\sigma}_{j}^{gs}\rangle],- italic_i ∑ start_POSTSUBSCRIPT italic_k ≠ italic_j end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT [ ⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_s end_POSTSUPERSCRIPT ⟩ + ⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_s end_POSTSUPERSCRIPT ⟩ ] ,
tσ^jrs=subscript𝑡delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑟𝑠absent\displaystyle\partial_{t}\langle\hat{\sigma}_{j}^{rs}\rangle=∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r italic_s end_POSTSUPERSCRIPT ⟩ = iΩ2[σ^jgsσ^jrg]i(ΔrΔsiγ)σ^jrs,𝑖Ω2delimited-[]delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑔𝑠delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑟𝑔𝑖subscriptΔ𝑟subscriptΔ𝑠𝑖𝛾delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑟𝑠\displaystyle\ i\frac{\Omega}{2}[\langle\hat{\sigma}_{j}^{gs}\rangle-\langle% \hat{\sigma}_{j}^{rg}\rangle]-i(\Delta_{r}-\Delta_{s}-i{\gamma})\langle\hat{% \sigma}_{j}^{rs}\rangle,italic_i divide start_ARG roman_Ω end_ARG start_ARG 2 end_ARG [ ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_s end_POSTSUPERSCRIPT ⟩ - ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r italic_g end_POSTSUPERSCRIPT ⟩ ] - italic_i ( roman_Δ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT - roman_Δ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_i italic_γ ) ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r italic_s end_POSTSUPERSCRIPT ⟩ ,

In the mean-field treatment lee2011antiferromagnetic ; qian2012phase ; vsibalic2016driven ; he2022superradiance , we neglect the correlation between different atoms, which allows for a factorization of the high-order moments, e.g., n^krσ^jgr=n^krσ^jgrdelimited-⟨⟩superscriptsubscript^𝑛𝑘𝑟superscriptsubscript^𝜎𝑗𝑔𝑟delimited-⟨⟩superscriptsubscript^𝑛𝑘𝑟delimited-⟨⟩superscriptsubscript^𝜎𝑗𝑔𝑟\langle\hat{n}_{k}^{r}\hat{\sigma}_{j}^{gr}\rangle=\langle\hat{n}_{k}^{r}% \rangle\langle\hat{\sigma}_{j}^{gr}\rangle⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_r end_POSTSUPERSCRIPT ⟩ = ⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ⟩ ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_g italic_r end_POSTSUPERSCRIPT ⟩. Assuming a uniform spatial distribution marcuzzi2014universal , e.g., n^jr=nrdelimited-⟨⟩superscriptsubscript^𝑛𝑗𝑟subscript𝑛𝑟\langle\hat{n}_{j}^{r}\rangle=n_{r}⟨ over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ⟩ = italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, we arrive at the mean-field equations:

n˙rsubscript˙𝑛𝑟\displaystyle\dot{n}_{r}over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT =iΩ2(σgrσrg)γnr,absent𝑖Ω2subscript𝜎𝑔𝑟subscript𝜎𝑟𝑔𝛾subscript𝑛𝑟\displaystyle=i\frac{\Omega}{2}({\sigma}_{gr}-{\sigma}_{rg})-\gamma{n}_{r},= italic_i divide start_ARG roman_Ω end_ARG start_ARG 2 end_ARG ( italic_σ start_POSTSUBSCRIPT italic_g italic_r end_POSTSUBSCRIPT - italic_σ start_POSTSUBSCRIPT italic_r italic_g end_POSTSUBSCRIPT ) - italic_γ italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ,
n˙ssubscript˙𝑛𝑠\displaystyle\dot{n}_{s}over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT =iΩ2(σgsσsg)γns,absent𝑖Ω2subscript𝜎𝑔𝑠subscript𝜎𝑠𝑔𝛾subscript𝑛𝑠\displaystyle=i\frac{\Omega}{2}({\sigma}_{gs}-{\sigma}_{sg})-\gamma{n}_{s},= italic_i divide start_ARG roman_Ω end_ARG start_ARG 2 end_ARG ( italic_σ start_POSTSUBSCRIPT italic_g italic_s end_POSTSUBSCRIPT - italic_σ start_POSTSUBSCRIPT italic_s italic_g end_POSTSUBSCRIPT ) - italic_γ italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ,
σ˙grsubscript˙𝜎𝑔𝑟\displaystyle\dot{\sigma}_{gr}over˙ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_g italic_r end_POSTSUBSCRIPT =iΩ2(2nr+ns+σsr1)+i(ΔrENL+iγ2)σgr,absent𝑖Ω22subscript𝑛𝑟subscript𝑛𝑠subscript𝜎𝑠𝑟1𝑖subscriptΔ𝑟subscript𝐸NL𝑖𝛾2subscript𝜎𝑔𝑟\displaystyle=i\frac{\Omega}{2}(2{n}_{r}+{n}_{s}+{\sigma}_{sr}-1)+i\left(% \Delta_{r}-E_{\mathrm{NL}}+i\frac{\gamma}{2}\right){\sigma}_{gr},= italic_i divide start_ARG roman_Ω end_ARG start_ARG 2 end_ARG ( 2 italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + italic_σ start_POSTSUBSCRIPT italic_s italic_r end_POSTSUBSCRIPT - 1 ) + italic_i ( roman_Δ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT roman_NL end_POSTSUBSCRIPT + italic_i divide start_ARG italic_γ end_ARG start_ARG 2 end_ARG ) italic_σ start_POSTSUBSCRIPT italic_g italic_r end_POSTSUBSCRIPT ,
σ˙gssubscript˙𝜎𝑔𝑠\displaystyle\dot{\sigma}_{gs}over˙ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_g italic_s end_POSTSUBSCRIPT =iΩ2(2ns+nr+σrs1)+i(ΔsENL+iγ2)σgs,absent𝑖Ω22subscript𝑛𝑠subscript𝑛𝑟subscript𝜎𝑟𝑠1𝑖subscriptΔ𝑠subscript𝐸NL𝑖𝛾2subscript𝜎𝑔𝑠\displaystyle=i\frac{\Omega}{2}(2{n}_{s}+{n}_{r}+{\sigma}_{rs}-1)+i\left(% \Delta_{s}-E_{\mathrm{NL}}+i\frac{\gamma}{2}\right){\sigma}_{gs},= italic_i divide start_ARG roman_Ω end_ARG start_ARG 2 end_ARG ( 2 italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_σ start_POSTSUBSCRIPT italic_r italic_s end_POSTSUBSCRIPT - 1 ) + italic_i ( roman_Δ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT roman_NL end_POSTSUBSCRIPT + italic_i divide start_ARG italic_γ end_ARG start_ARG 2 end_ARG ) italic_σ start_POSTSUBSCRIPT italic_g italic_s end_POSTSUBSCRIPT ,
σ˙rssubscript˙𝜎𝑟𝑠\displaystyle\dot{\sigma}_{rs}over˙ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_r italic_s end_POSTSUBSCRIPT =iΩ2(σgsσrg)i(ΔrΔsiγ)σrs,absent𝑖Ω2subscript𝜎𝑔𝑠subscript𝜎𝑟𝑔𝑖subscriptΔ𝑟subscriptΔ𝑠𝑖𝛾subscript𝜎𝑟𝑠\displaystyle=i\frac{\Omega}{2}({\sigma}_{gs}-{\sigma}_{rg})-i(\Delta_{r}-% \Delta_{s}-i{\gamma}){\sigma}_{rs},= italic_i divide start_ARG roman_Ω end_ARG start_ARG 2 end_ARG ( italic_σ start_POSTSUBSCRIPT italic_g italic_s end_POSTSUBSCRIPT - italic_σ start_POSTSUBSCRIPT italic_r italic_g end_POSTSUBSCRIPT ) - italic_i ( roman_Δ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT - roman_Δ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT - italic_i italic_γ ) italic_σ start_POSTSUBSCRIPT italic_r italic_s end_POSTSUBSCRIPT ,

where ENL=χ(nr+ns)subscript𝐸NL𝜒subscript𝑛𝑟subscript𝑛𝑠E_{\mathrm{NL}}=\chi({n}_{r}+{n}_{s})italic_E start_POSTSUBSCRIPT roman_NL end_POSTSUBSCRIPT = italic_χ ( italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) is the nonlinear energy shift with χ=kjVjk𝜒subscript𝑘𝑗subscript𝑉𝑗𝑘\chi=\sum_{k\neq j}V_{jk}italic_χ = ∑ start_POSTSUBSCRIPT italic_k ≠ italic_j end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT italic_j italic_k end_POSTSUBSCRIPT. The stable behavior of the system is determined by the property of the fixed points of the above nonlinear Bloch equations. With a standard stability analysis strogatz2018nonlinear , we can analyze types of bifurcations and obtain the phase diagram shown in Fig. 1d, where we set Δr=ΔsubscriptΔ𝑟Δ\Delta_{r}=\Deltaroman_Δ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = roman_Δ and Δs=ΔδsubscriptΔ𝑠Δ𝛿\Delta_{s}=\Delta-\deltaroman_Δ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = roman_Δ - italic_δ. In Extended Data Fig. 1, we calculate the mean-field dynamics in different regimes. When the system is in the limit cycle phase, the mean-field components exhibit persistent periodic oscillations after an initial synchronization [see Extended Data Fig. 1a]. Here, oscillation of the imaginary part of the transition dipole σgrsubscript𝜎𝑔𝑟\sigma_{gr}italic_σ start_POSTSUBSCRIPT italic_g italic_r end_POSTSUBSCRIPT causes an oscillating absorption coefficient of the probe field, which can be directly monitored by the transmission signal as in our experiment. Instead, if the system is in the stationary phase [see Extended Data Fig. 1b], all observables will eventually converge to their steady-state values.

The primary aim of the mean-field model is to develop an intuitive understanding of the experiment and to show that the inclusion of multiple Rydberg-levels promotes the observed persistent oscillations. Here, we also compare the model prediction with the experiment on a more quantitative level. We focus specifically on the emerging frequency of the oscillation and its dependence on the system parameters. For typical parameters of the experiment, we identify oscillatory solutions with high as well as low frequencies. While high oscillatory usually overestimate the frequency by more than one order of magnitude, the slow oscillatory solutions can yield typical frequencies on the same order as in our experiment and exhibit the same dependence on the system parameters. As illustrated in Extended Data Fig. 2a, the measured oscillation frequency decreases significantly with a slight increase of the driving Rabi frequency, in agreement with the mean-field prediction shown in Extended Data Fig. 2b. Such a behavior is not immediately obvious, as one would normally expect a positive correlation between the oscillation frequency and the driving Rabi frequency, e.g., in a single-atom case.

We therefore believe that our model captures the essential physics of the observed time crystal formation and can serve as a starting point for developing a more complicated theory, where more details (e.g., incorporation of the intermediate state, the random thermal motion, Doppler effects induced by atoms with different velocities, and the laser-beam inhomogeneity) can be taken into account to yield more quantitative predictions.

Experimental details

The detailed experimental setup is depicted in Extended Data Fig. 3. Our experiments are performed in a 7.5-cm-long room-temperature Rb vapour cell miller2016radio ; ripka2018room ; li2022telecom with natural abundance, where the proportions of 85Rb and 87Rb isotopes are about 72.2% and 27.8%. The laser sources of 780 nm and 480 nm light fields are [Toptica DL pro @780 nm] and [Toptica TA-SHG pro @480 nm], and both of them can work in the frequency-scanning mode or the frequency-locking mode. For 780 nm laser, we use an acoustic optical modulator (AOM) before the coupler of the fiber to control the output laser intensity, which is monitored by an independent PD. Then, a calcite beam displacer (Thorlabs BD40) is used to generate two parallell 780 nm beams for differential measurements, one of which is chosen as the probe beam (overlapped with a 480 nm coupling beam) with the other being a reference. The maximum powers of the probe (1/e21superscript𝑒21/e^{2}1 / italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT waist radius 1000μm1000𝜇𝑚1000\ \mu m1000 italic_μ italic_m) and the coupling (1/e21superscript𝑒21/e^{2}1 / italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT waist radius 900μm900𝜇𝑚900\ \mu m900 italic_μ italic_m) beams are approximately 1.2mW1.2mW1.2\ \mathrm{mW}1.2 roman_mW and 460mW460mW460\ \mathrm{mW}460 roman_mW, with which the maximum Rabi frequencies are approximately Ωp(max)2π×20.5MHzsimilar-tosuperscriptsubscriptΩ𝑝max2𝜋20.5MHz\Omega_{p}^{(\mathrm{max})}\sim 2\pi\times 20.5\ \mathrm{MHz}roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( roman_max ) end_POSTSUPERSCRIPT ∼ 2 italic_π × 20.5 roman_MHz and Ωc(max)2π×1.9MHzsimilar-tosuperscriptsubscriptΩ𝑐max2𝜋1.9MHz\Omega_{c}^{(\mathrm{max})}\sim 2\pi\times 1.9\ \mathrm{MHz}roman_Ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( roman_max ) end_POSTSUPERSCRIPT ∼ 2 italic_π × 1.9 roman_MHz (dependent on the laser polarization and the specific Zeeman level). The transmission spectrum is obtained by scanning the frequency of the coupling field (the detuning Δc(t)subscriptΔ𝑐𝑡\Delta_{c}(t)roman_Δ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t )) at a fixed intermediate-state detuning ΔpsubscriptΔ𝑝\Delta_{p}roman_Δ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT around the two-photon resonance Δp+Δc=0subscriptΔ𝑝subscriptΔ𝑐0\Delta_{p}+\Delta_{c}=0roman_Δ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT + roman_Δ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0.

For the measurement of the quench dynamics, the frequencies of both laser sources are locked to an ultralow expansion (ULE) cavity with the help of the Pound-Drever-Hall (PDH) technique. Then, by controlling the radio frequency (RF) driving on the AOM, we can turn on and off the 780 nm output laser abruptly within 2μ2𝜇2\ \mu2 italic_μs, during which the 480 nm laser is always kept on. When exploring the rigidity of the dissipative time crystal, the intensity noise added to the probe field is achieved by using a waveform generator (Keysight 33600A Series) to modulate the RF amplitude.

To adjust the energy spacing of different Rydberg Zeeman sublevels, we use a homemade 7.5-cm-long coil with 35 windings to generate a homogeneous external magnetic field parallel to the laser beams. Meanwhile, quarter-wave plates (QWP) are used to control the polarizations of the 780 nm and 480 nm light fields, with which we can tune the coupling strengths to different Zeeman levels (see Fig. 2 of the main text).

While the data shown in this paper is based on the excitation of 85Rb isotopes, similar oscillatory behavior is also observed for 87Rb isotopes, indicating that the oscillating dynamics is a universal phenomenon in Rydberg atomic ensembles. However, due to the lower natural abundance of 87Rb isotopes, the limit cycle usually requires a larger Rabi frequency or an additional heating to increase the Rydberg excitation density.

Influence of the principal quantum number

It is necessary to excite atoms to a high principal quantum number n𝑛nitalic_n for the observation of oscillatory signals. Figures 4a-d show the transmission spectrum at different n𝑛nitalic_n, where the probe (coupling) beam is elliptically (linearly) polarized with no external magnetic field. At a relatively small n=55𝑛55n=55italic_n = 55, as expected, there are only two distinct transmission peaks, representing the resonant positions or energy levels of two Rydberg nD𝑛𝐷nDitalic_n italic_D-state manifolds with total angular momentum quantum numbers J=3/2𝐽32J=3/2italic_J = 3 / 2 and J=5/2𝐽52J=5/2italic_J = 5 / 2. The peak of the latter is higher due to the larger dipole matrix element between the Rydberg state and the intermediate state. In other words, the Rabi frequency ΩcsubscriptΩ𝑐\Omega_{c}roman_Ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT of coupling beam for |55D5/2ket55subscript𝐷52|55D_{5/2}\rangle| 55 italic_D start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT ⟩ is larger than |55D3/2ket55subscript𝐷32|55D_{3/2}\rangle| 55 italic_D start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT ⟩. Consequently, the Rydberg population is also higher for |55D5/2ket55subscript𝐷52|55D_{5/2}\rangle| 55 italic_D start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT ⟩, leading to non-equilibrium dynamical phase transitions in bistable regions where the signal abruptly changes (green arrows).

When n𝑛nitalic_n is increased to 65 [see Extended Data Fig. 4b], the overall peaks become lower due to the smaller coupling Rabi frequency which scales as n3/2similar-toabsentsuperscript𝑛32\sim n^{-3/2}∼ italic_n start_POSTSUPERSCRIPT - 3 / 2 end_POSTSUPERSCRIPT. The energy difference between two Rydberg states also becomes smaller. Here, multiple phase transitions in the left side of |65D5/2ket65subscript𝐷52|65D_{5/2}\rangle| 65 italic_D start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT ⟩ are identified, and the emergent oscillatory signal appears in the right side of |65D5/2ket65subscript𝐷52|65D_{5/2}\rangle| 65 italic_D start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT ⟩ (red arrow). Dramatic changes to the above features occur when n𝑛nitalic_n is further increased [see Extended Data Fig. 4a-d]. As |nD3/2ket𝑛subscript𝐷32|nD_{3/2}\rangle| italic_n italic_D start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT ⟩ and |nD5/2ket𝑛subscript𝐷52|nD_{5/2}\rangle| italic_n italic_D start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT ⟩ become closer, the two Rydberg states come to merge and the oscillation signal gets amplified over a wide range of the detuning ΔcsubscriptΔ𝑐\Delta_{c}roman_Δ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. In general, the appearance and the enhancement of the oscillating dynamics with an increasing n𝑛nitalic_n can be attributed to: (i) the stronger Rydberg interactions, and (ii) the participation of more Rydberg states (within |nD3/2ket𝑛subscript𝐷32|nD_{3/2}\rangle| italic_n italic_D start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT ⟩ and |nD5/2ket𝑛subscript𝐷52|nD_{5/2}\rangle| italic_n italic_D start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT ⟩ manifold) in the dynamics.

Long-term stability and random phase fluctuation.

As described in the main text, the oscillating signal will synchronize to a stable frequency in each single-shot realization of the quench dynamics. However, two independent realizations separated by a very long time can still possess very different stable frequencies, due to the drift of experimental conditions, e.g., room temperatures and magnetic noises. For the measurement performed in Fig. 3c, we record 200 independent realizations of the quench dynamics, which are separated into 50 groups, with a 125 ms duration between successive realizations within each group and a 10 s duration between nearest groups. As shown in Extended Data Fig. 5a, the extracted peak frequencies of the DFT spectrum (performed for the time window 50-100 ms, already in the stable periodic region) for these 200 realizations have a long-term drift, approximately 6 times of the DFT resolution. To control the influence of frequency drifts, we select the first 100 realizations for Fig. 3c. The evolution of the peak frequency obeys the same law if we choose the full 200 realizations, albeit with a slightly larger error bar.

Here, we also find that a stable oscillation pattern can have a random initial phase. To avoid the phase uncertainty caused by fluctuations of the oscillation frequency, we further postselect the data of the same frequency from the first 100 realizations (indicated by the red box in Extended Data Fig. 5a. As shown in Extended Data Fig. 5b, the phase of the Fourier amplitude A(ω)𝐴𝜔A(\omega)italic_A ( italic_ω ) at the peak frequency ωcsubscript𝜔𝑐\omega_{c}italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is randomly distributed over [0,2π)02𝜋[0,2\pi)[ 0 , 2 italic_π ).

Such a random phase fluctuation has been related to a spontaneous time translation symmetry breaking in previous works. In fact, it can also be related to a phase diffusion process associated with the unbounded growth of quantum fluctuations in a time-crystalline order chan2015limit ; carollo2022exactM : while a single-run measurement always exhibits a persistent oscillation, quantum fluctuations can lead to damping of the average signal [see Extended Data Fig. 5c]. Since we cannot completely rule out classical fluctuations (e.g., random atomic motions and phase noise of the driving lasers), the relative contribution of quantum fluctuations is difficult to determine in the current experiment. In the main text, we focus on the autocorrelation function, which is also an efficient way to characterize a time crystal.

Role of the ionization process

In a hot Rydberg gas, atomic collisions can lead to strong ionization and plasma formation. The interaction between atoms and the resulting ions can even dominate over the Rydberg-Rydberg interaction in certain regimes weller2016charge ; wade2018terahertz ; weller2019interplay . In this subsection, we present several experimental efforts to identify the role of ionic interactions.

In Ref. weller2016charge , the authors ruled out Rydberg-Rydberg interactions as a possible mechanism to induce the bistability, based on the fact that the observed red energy shift of the transmission spectrum could not be explained by a repulsive vdW interaction between |nSket𝑛𝑆|nS\rangle| italic_n italic_S ⟩ Rydberg states of Rb atoms, but instead was consistent with the attractive interactions between Rydberg atoms and ions. In our experiments, Rb atoms are excited to |nD5/2ket𝑛subscript𝐷52|nD_{5/2}\rangle| italic_n italic_D start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT ⟩ Rydberg state, where the vdW interactions are attractive with a dispersive coefficient C6<0subscript𝐶60C_{6}<0italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT < 0. This attractive vdW interaction is qualitatively consistent with the red shifted transmission spectrum we observe in the experiment. Hence, the spectral shape alone is insufficient to rule out Rydberg-Rydberg atom interactions as the dominant mechanism.

Second, we note that our experimental conditions are quite different from the mentioned experiments where ion-induced interaction has been identified as the dominant one. For example, we do not heat the atomic vapor and therefore operate at significantly lower densities of our room-temperature gases. Moreover, our control Rabi frequency Ωc1similar-tosubscriptΩ𝑐1\Omega_{c}\sim 1~{}roman_Ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∼ 1MHz is much lower compared to other bistability experiments, which implies a lower Rydberg excitation fraction in a hot gas.

The results of these different conditions can be seen from the measured spectra. The widths of our spectra [see Figs. 2b and 2c] are about an order of magnitude narrower than other bistability experiments, where the broad linewidth is attributed to the ionic interaction induced broadening.

To further search for the presence of plasma, we have performed additional measurements of the fluorescence spectrum. As discussed in Ref. weller2019interplay , the presence of a plasma should lead to radiative emission at a broad range of wavelength due to recombination and population of different Rydberg levels. First, we confirmed that our spectrometer reliably detects the fluorescence from the vapor cell, by monitoring the 780 nm photons when scanning the detuning. As shown in Figs. 6a and 6d, the counts of 780 nm photons decrease significantly around the transmission peaks, which is a direct consequence of the reduced occupation of the intermediate state |5P3/2ket5subscript𝑃32|5P_{3/2}\rangle| 5 italic_P start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT ⟩ due to the electromagnetic induced transparency (EIT). Figures 6a, c, e, and f, show the entire spectrum from 200 nm to 1000 nm at different detunings of the driving field labeled in a and d. Within the sensitivity limit of our spectrometer, we observe no pronounced fluorescence except from the two driving lasers at 780 nm (probe field) and 480 nm (coupling field), and the same background values for the on-resonant [c and f] and the far-detuned [b and e] excitation. This is consistent with fluorescence imaging of the vapor cell using a qCMOS camera (HAMAMATSU, C15550-20UP) for bandpass-filtered wavelengths between 500 nm and 700 nm after background subtraction, where no obvious fluorescence signal is detected.

Although the current experiment does not show clear evidence of ionic excitation or plasma formation, we cannot completely rule out them due to the limited resolution of the fluorescence measurement. Therefore, the ionic interaction might still be relevant to the dynamics. In fact, the inclusion of ionic interactions can also be described by the mean-field model we developed and does not destroy the mechanism towards establishment of the limit cycle oscillation. As ions are generated from Rydberg atoms, their density should be proportional to the total Rydberg density, which can give rise to a nonlinear energy shift ENL=χ(nr+ns)subscript𝐸NL𝜒subscript𝑛𝑟subscript𝑛𝑠E_{\mathrm{NL}}=\chi({n}_{r}+{n}_{s})italic_E start_POSTSUBSCRIPT roman_NL end_POSTSUBSCRIPT = italic_χ ( italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT + italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) due to the ion-induced Stark shift weller2019interplay , similar to Rydberg-interaction induced mean-field level shift. Thus, the nonlinear competition between distinct Rydberg states also exists when taking the ionization process into consideration, and we only need to modify the effective nonlinear coefficient χ𝜒\chiitalic_χ according to the details of the ionic interaction. A more accurate and complete model considering charges needs to be investigated in the future.

Data availability

The data are available from the corresponding author on reasonable request.

Code availability

The codes are available upon reasonable request from the corresponding author.

Refer to caption
Extended Data Figure 1: Mean-field dynamics. The calculations are performed with Ω=3γΩ3𝛾\Omega=3\gammaroman_Ω = 3 italic_γ, χ=16γ𝜒16𝛾\chi=-16\gammaitalic_χ = - 16 italic_γ, and δ=8γ𝛿8𝛾\delta=8\gammaitalic_δ = 8 italic_γ. For the limit cycle phase shown in a with Δ=3γΔ3𝛾\Delta=-3\gammaroman_Δ = - 3 italic_γ, the order parameters exhibit persistent oscillations. For the stationary phase shown in b with Δ=3γΔ3𝛾\Delta=3\gammaroman_Δ = 3 italic_γ, the order parameters converge to steady-state values.
Refer to caption
Extended Data Figure 2: Comparison between experiment and theory. a Experimentally extracted oscillation frequency as a function of the effective two-photon Rabi frequency Ωeff=ΩcΩp/2ΔpsubscriptΩeffsubscriptΩ𝑐subscriptΩ𝑝2subscriptΔ𝑝\Omega_{\mathrm{eff}}=\Omega_{c}\Omega_{p}/2\Delta_{p}roman_Ω start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = roman_Ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / 2 roman_Δ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. In the experiment, we only vary the intensity of the coupling field (ΩcsubscriptΩ𝑐\Omega_{c}roman_Ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT), and keep all other conditions the same as in Fig. 2d. b Theoretically predicted oscillation frequency as a function of the Rabi frequency ΩΩ\Omegaroman_Ω. The parameters chosen are close to the experiment. For example, the decay rate is evaluated to be γ/2π=18.0𝛾2𝜋18.0\gamma/2\pi=18.0~{}italic_γ / 2 italic_π = 18.0kHz, consisting of the contribution from the Rydberg spontaneous decay (γdsubscript𝛾𝑑\gamma_{d}italic_γ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT) and the transient time broadening γt=v¯/Dsubscript𝛾𝑡subscript¯𝑣perpendicular-to𝐷\gamma_{t}=\bar{v}_{\perp}/Ditalic_γ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT / italic_D with D𝐷Ditalic_D the diameter of the beam and v¯=πkT/2msubscript¯𝑣perpendicular-to𝜋𝑘𝑇2𝑚\bar{v}_{\perp}=\sqrt{{\pi kT}/{2m}}over¯ start_ARG italic_v end_ARG start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = square-root start_ARG italic_π italic_k italic_T / 2 italic_m end_ARG the transverse mean velocity. The other parameters δ/2π=0.306𝛿2𝜋0.306\delta/2\pi=0.306~{}italic_δ / 2 italic_π = 0.306MHz, Δ/2π=0.186Δ2𝜋0.186\Delta/2\pi=-0.186~{}roman_Δ / 2 italic_π = - 0.186MHz, and χ/2π=0.951𝜒2𝜋0.951\chi/2\pi=-0.951~{}italic_χ / 2 italic_π = - 0.951MHz are also typical in our experiment.
Refer to caption
Extended Data Figure 3: A detailed schematic diagram of the experimental setup. The 780 nm probe and reference beams generated from a calcite beam displacer are propagating in parallel through a room-temperature Rb vapor cell, with the probe beam overlapping with a counterpropagating coupling beam. Their transmission signals are detected on a balanced photon detector (PD) for differential measurement. The output intensity of the 780 nm laser from the fiber (monitored by an independent PD) is controlled by an AOM, whose RF driving is modulated by a waveform generator. QWPs are used to control the polarization of the lasers illuminating the atoms.
Refer to caption
Extended Data Figure 4: Transmission spectrum at different principal quantum number n𝑛nitalic_n. As n𝑛nitalic_n increases, |nD3/2ket𝑛subscript𝐷32|nD_{3/2}\rangle| italic_n italic_D start_POSTSUBSCRIPT 3 / 2 end_POSTSUBSCRIPT ⟩ and |nD5/2ket𝑛subscript𝐷52|nD_{5/2}\rangle| italic_n italic_D start_POSTSUBSCRIPT 5 / 2 end_POSTSUBSCRIPT ⟩ states move closer, and their corresponding transmission peaks begin to merge. Oscillation appears and is enhanced for a sufficiently large n𝑛nitalic_n.
Refer to caption
Extended Data Figure 5: Long-term stability of the oscillation frequency and breaking of continuous time translation symmetry. a The long-term frequency drift in 200 independent realizations of the quench dynamics by extracting peak frequencies in the stable periodic region (50-100 ms). The data encircled by the red box are of the same frequency, and are postselected for b, which displays the distribution of the Fourier amplitudes (at the same peak frequency) on the complex plane. c Single-shot and average transmission signals over 36 realizations with a same stable frequency indicated by the red box in a.
Refer to caption
Extended Data Figure 6: Fluorescence measurements in our Rydberg gas. a and d are transmission signals at different principal quantum number n=55𝑛55n=55italic_n = 55 and n=75𝑛75n=75italic_n = 75, respectively. Orange squares represent the counts of 780 nm phontons collected by the spectrometer. b-c and e-f show the full spectra from 200 nm to 1000 nm at different dutunings indicated in a and d.

References

  • (1) Lee, T. E., Häffner, H. & Cross, M. C. Antiferromagnetic phase transition in a nonequilibrium lattice of Rydberg atoms. Phys. Rev. A 84, 031402 (2011).
  • (2) Qian, J., Dong, G., Zhou, L. & Zhang, W. Phase diagram of Rydberg atoms in a nonequilibrium optical lattice. Phys. Rev. A 85, 065401 (2012).
  • (3) Šibalić, N., Wade, C. G., Adams, C. S., Weatherill, K. J. & Pohl, T. Driven-dissipative many-body systems with mixed power-law interactions: Bistabilities and temperature-driven nonequilibrium phase transitions. Phys. Rev. A 94, 011401 (2016).
  • (4) He, Y., Bai, Z., Jiao, Y., Zhao, J. & Li, W. Superradiance-induced multistability in one-dimensional driven Rydberg lattice gases. Phys. Rev. A 106, 063319 (2022).
  • (5) Marcuzzi, M., Levi, E., Diehl, S., Garrahan, J. P. & Lesanovsky, I. Universal nonequilibrium properties of dissipative rydberg gases. Phys. Rev. Lett. 113, 210401 (2014).
  • (6) Strogatz, S. H. Nonlinear dynamics and chaos with student solutions manual: With applications to physics, biology, chemistry, and engineering. (2018).
  • (7) Miller, S. A., Anderson, D. A. & Raithel, G. Radio-frequency-modulated Rydberg states in a vapor cell. New J. Phys. 18, 053017 (2016).
  • (8) Ripka, F., Kübler, H., Löw, R. & Pfau, T. A room-temperature single-photon source based on strongly interacting Rydberg atoms. Science 362, 446–449 (2018).
  • (9) Li, W., Du, J., Lam, M. & Li, W. Telecom-wavelength spectra of a Rydberg state in a hot vapor. Opt. Lett. 47, 4399–4402 (2022).
  • (10) Chan, C. -K., Lee, T. E. & Gopalakrishnan, S. Limit-cycle phase in driven-dissipative spin systems. Phys. Rev. A 91, 051601 (2015).
  • (11) Carollo, F. & Lesanovsky, I. Exact solution of a boundary time-crystal phase transition: time-translation symmetry breaking and non-markovian dynamics of correlations. Phys. Rev. A 105, L040202 (2022).
  • (12) Weller, D., Urvoy, A., Rico, A., Löw, R. & Kübler, H. Charge-induced optical bistability in thermal Rydberg vapor. Phys. Rev. A 94, 063820 (2016).
  • (13) Wade, C. G., Marcuzzi, M., Levi, E., Kondo, J. M., Lesanovsky, I., Adams, C. S. & Weatherill, K. J. A terahertz-driven non-equilibrium phase transition in a room temperature atomic vapour. Nat. Commun. 9, 3567 (2018).
  • (14) Weller, D., Shaffer, J. P., Pfau, T., Löw, R. & Kübler, H. Interplay between thermal Rydberg gases and plasmas. Phys. Rev. A 99, 043418 (2019).