License: arXiv.org perpetual non-exclusive license
arXiv:2401.00785v1 [quant-ph] 01 Jan 2024

Theoretical Study on Superradiant Raman Scattering with Rubidium Atoms in An Optical Cavity

Huihui Yu Henan Key Laboratory of Diamond Optoelectronic Materials and Devices, Key Laboratory of Material Physics Ministry of Education, School of Physics and Microelectronics, Zhengzhou University, Daxue Road 75, Zhengzhou 450052, China    Yuan Zhang yzhuaudipc@zzu.edu.cn Henan Key Laboratory of Diamond Optoelectronic Materials and Devices, Key Laboratory of Material Physics Ministry of Education, School of Physics and Microelectronics, Zhengzhou University, Daxue Road 75, Zhengzhou 450052, China Institute of Quantum Materials and Physics, Henan Academy of Sciences, Zhengzhou 450046, China    Gang Chen Henan Key Laboratory of Diamond Optoelectronic Materials and Devices, Key Laboratory of Material Physics Ministry of Education, School of Physics and Microelectronics, Zhengzhou University, Daxue Road 75, Zhengzhou 450052, China Institute of Quantum Materials and Physics, Henan Academy of Sciences, Zhengzhou 450046, China    Chongxin Shan cxshan@zzu.edu.cn Henan Key Laboratory of Diamond Optoelectronic Materials and Devices, Key Laboratory of Material Physics Ministry of Education, School of Physics and Microelectronics, Zhengzhou University, Daxue Road 75, Zhengzhou 450052, China Institute of Quantum Materials and Physics, Henan Academy of Sciences, Zhengzhou 450046, China
Abstract

Superradiant Raman scattering of Rubidium atoms has been explored in the experiment [Nature 484, 78 (2012)] to prove the concept of the superradiant laser, which attracts significant attentions in quantum metrology due to the expected ultra-narrow linewidth down to millihertz. To better understand the physics involved in this experiment, we have developed a quantum master equation theory by treating the Rubidium atoms as three-level systems, and coupling them with a dressed laser and an optical cavity. Our simulations show different superradiant Raman scattering pulses for the systems within the crossover and strong coupling regime, and the shifted and broader spectrum of the steady-state Raman scattering. Thus, our studies provide a unified view on the superradiant Raman scattering pulses, and an alternative explanation to the broad spectrum of the steady-state Raman scattering, as observed in the experiment. In future, our theory can be readily applied to study other interesting phenomena relying on the superradiant Raman scattering, such as magnetic field sensing, real-time tracking of quantum phase, Dicke phase transition of non-equilibrium dynamics and so on.

I Introduction

Superradiance, i.e. collective spontaneous emission of quantum emitters, was firstly proposed by R. H. Dicke in 1954 (RHDicke1954, ), and was then studied extensively in theories and experiments in 1980s (AVAndreev1980, ). Because this effect was often studied with the quantum emitters under the pulsed light illumination, it was considered as a transient phenomenon following the collective decay of the emitters. However, in 2009, D. Meiser et al. suggested that such a collective decay can be compensated by an incoherent atomic pumping, and predicted that the resulted superradiant laser can have a ultra-narrow linewidth down to millihertz for optical lattice clock systems (DMeiser2009, ). Due to the application potential in quantum metrology, the superradiant laser has been studied extensively in both theories (DATieri2017, ; KDebnath2018, ; YZhang2021, ) and experiments (JDBohnet2012, ; MANorcia2016, ; MANorcia2016-1, ) thereafter.

Among the experiments on the superradiant laser, the one by J. G. Bohnet et al, was often viewed as a proof-of-concept of such an effect (JDBohnet2012, ), although it relied on Raman transition of Rubidium atoms (Fig. 1) rather than real optical clock transitions. To better understand the physics involved, in this article, we develop a quantum master equation theory by treating the atoms as three-level systems and coupling them with a dressed laser and an optical cavity. By solving the master equation with cumulant mean-field approach, we simulate the systems with tens of thousands of Rubidium atoms.

Our calculations show the different superradiant Raman scattering pulses for the system in the crossover and strong coupling regime, and the shifted and broader Raman scattering spectrum for the system at steady-state. Thus, our results provide a unified view on the superradiant Raman pulses, and an alternative explanation on the broad spectrum, as observed in the experiments. In future, our theory can be readily applied to study other interesting phenomena relying on the superradiant Raman scattering, such as magnetic field sensing (JMWeiner2012, ), real-time track of quantum phase (AShankar2019, ), Dicke phase transition of non-equilibrium dynamics  (MPBaden2019, ), and so on.

The current article is organized as follows. In the following section, we present the corresponding quantum master equation, the effective master equation after eliminating the optically excited level, and their solution with the cumulant mean-field approach. In Sec. III and  IV, we present our numerical results on the transient and steady-state superradiant Raman scattering, respectively. In the end, we summarize our work and comment on the possible studies in the future.

Refer to caption
Figure 1: System and energy diagram. Panel (a) shows tens of thousands of Rubidium-87 atoms trapped in an optical lattice inside an optical cavity, a dressed (re-pumping) laser transmitting (transecting) the optical cavity, and the superradiant Raman signal out of the cavity. Panel (b) displays a simplified energy diagram, where the dressed laser excites the atoms from the hyper-fine ground level |2k=|52S1/2,F=2,mf=2ketsubscript2𝑘ketformulae-sequencesuperscript52subscript𝑆12𝐹2subscript𝑚𝑓2\left|2_{k}\right\rangle=\left|5^{2}S_{1/2},F=2,m_{f}=-2\right\rangle| 2 start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ⟩ = | 5 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_S start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT , italic_F = 2 , italic_m start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = - 2 ⟩ to the virtual level (dashed line) slightly above the excited level |3k=|52P1/2,F=2,mf=2ketsubscript3𝑘ketformulae-sequencesuperscript52subscript𝑃12𝐹2subscript𝑚𝑓2\left|3_{k}\right\rangle=\left|5^{2}P_{1/2},F=2,m_{f}=-2\right\rangle| 3 start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ⟩ = | 5 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT , italic_F = 2 , italic_m start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = - 2 ⟩ (green arrow), and the coupling with the optical cavity de-excites the atoms to the hyper-fine level |1k=|52S1/2,F=1,mf=1ketsubscript1𝑘ketformulae-sequencesuperscript52subscript𝑆12𝐹1subscript𝑚𝑓1\left|1_{k}\right\rangle=\left|5^{2}S_{1/2},F=1,m_{f}=-1\right\rangle| 1 start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ⟩ = | 5 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_S start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT , italic_F = 1 , italic_m start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = - 1 ⟩(red wavy arrow), producing superradiant Raman scattering pulses. If the atoms are re-pumped to the higher hyper-fine ground level (blue dashed arrow), the superradiant Raman scattering can be maintained.

II Quantum Master Equation

The system under consideration consists of tens of thousands of Rubidium-87 atoms trapped in an optical lattice inside an optical cavity (Fig. 1a). A dressed laser transmits through the cavity, and excites the atoms from the hyper-fine ground level to the virtual level slightly above the excited level, and the coupling with the cavity de-excites the atoms to the hyper-fine level, leading to the superradiant Raman scattering pulses (Fig. 1b). If the atoms are also re-pumped to the higher hyper-fine ground level, the superradiant Raman scattering can be maintained. Note that the current diagram is only valid for the system within the weak or crossover coupling regime, and the excited level of the bare atoms should be replaced by the atoms-photon dressed states for the system within the strong coupling regime.

To describe the aforementioned dynamics, we establish the following quantum master equation for the reduced density operator ρ^^𝜌\hat{\rho}over^ start_ARG italic_ρ end_ARG in the interaction picture:

tρ^=i[H^ac+H^d,ρ^]κ𝒟[a^]ρ^𝑡^𝜌𝑖Planck-constant-over-2-pisubscript^𝐻𝑎𝑐subscript^𝐻𝑑^𝜌𝜅𝒟delimited-[]^𝑎^𝜌\displaystyle\frac{\partial}{\partial t}\hat{\rho}=-\frac{i}{\hbar}[\hat{H}_{a% -c}+\hat{H}_{d},\hat{\rho}]-\kappa\mathcal{D}\left[\hat{a}\right]\hat{\rho}divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG over^ start_ARG italic_ρ end_ARG = - divide start_ARG italic_i end_ARG start_ARG roman_ℏ end_ARG [ over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_a - italic_c end_POSTSUBSCRIPT + over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT , over^ start_ARG italic_ρ end_ARG ] - italic_κ caligraphic_D [ over^ start_ARG italic_a end_ARG ] over^ start_ARG italic_ρ end_ARG
k=1N(γ31𝒟[σ^k13]ρ^+γ12𝒟[σ^k21]ρ^).superscriptsubscript𝑘1𝑁subscript𝛾31𝒟delimited-[]superscriptsubscript^𝜎𝑘13^𝜌subscript𝛾12𝒟delimited-[]superscriptsubscript^𝜎𝑘21^𝜌\displaystyle-\sum_{k=1}^{N}(\gamma_{31}\mathcal{D}\left[\hat{\sigma}_{k}^{13}% \right]\hat{\rho}+\gamma_{12}\mathcal{D}\left[\hat{\sigma}_{k}^{21}\right]\hat% {\rho}).- ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( italic_γ start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT caligraphic_D [ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 13 end_POSTSUPERSCRIPT ] over^ start_ARG italic_ρ end_ARG + italic_γ start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT caligraphic_D [ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT ] over^ start_ARG italic_ρ end_ARG ) . (1)

In this equation, the Hamiltonian H^ac=g31[ei(ωcω31)ta^k=1Nσ^k13+ei(ωcω31)ta^k=1Nσ^k31]subscript^𝐻𝑎𝑐Planck-constant-over-2-pisubscript𝑔31delimited-[]superscript𝑒𝑖subscript𝜔𝑐subscript𝜔31𝑡superscript^𝑎superscriptsubscript𝑘1𝑁superscriptsubscript^𝜎𝑘13superscript𝑒𝑖subscript𝜔𝑐subscript𝜔31𝑡^𝑎superscriptsubscript𝑘1𝑁superscriptsubscript^𝜎𝑘31\hat{H}_{a-c}=\hbar g_{31}[e^{i\left(\omega_{c}-\omega_{31}\right)t}\hat{a}^{% \dagger}\sum_{k=1}^{N}\hat{\sigma}_{k}^{13}+e^{-i\left(\omega_{c}-\omega_{31}% \right)t}\hat{a}\sum_{k=1}^{N}\hat{\sigma}_{k}^{31}]over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_a - italic_c end_POSTSUBSCRIPT = roman_ℏ italic_g start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT [ italic_e start_POSTSUPERSCRIPT italic_i ( italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT ) italic_t end_POSTSUPERSCRIPT over^ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 13 end_POSTSUPERSCRIPT + italic_e start_POSTSUPERSCRIPT - italic_i ( italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT ) italic_t end_POSTSUPERSCRIPT over^ start_ARG italic_a end_ARG ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 31 end_POSTSUPERSCRIPT ] describes the coherent energy exchange between the atoms and the optical cavity with strength g31subscript𝑔31g_{31}italic_g start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT. Here, the optical cavity is modeled as a quantum harmonic oscillator with a frequency ωcsubscript𝜔𝑐\omega_{c}italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, and the photon creation operator a^superscript^𝑎\hat{a}^{\dagger}over^ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT and annihilation operator a^^𝑎\hat{a}over^ start_ARG italic_a end_ARG. σ^k13superscriptsubscript^𝜎𝑘13\hat{\sigma}_{k}^{13}over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 13 end_POSTSUPERSCRIPT, σ^k31superscriptsubscript^𝜎𝑘31\hat{\sigma}_{k}^{31}over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 31 end_POSTSUPERSCRIPT are the lowering and raising operator associated with the 1k3ksubscript1𝑘subscript3𝑘1_{k}\to 3_{k}1 start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT → 3 start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT transition of frequency ω31subscript𝜔31\omega_{31}italic_ω start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT. The index k𝑘kitalic_k indicates the individual of the total N𝑁Nitalic_N atoms. In Eq. (1), the Hamiltonian H^d=Ω[ei(ωdω32)tk=1Nσk23+ei(ωdω32)tk=1Nσk32]subscript^𝐻𝑑Planck-constant-over-2-piΩdelimited-[]superscript𝑒𝑖subscript𝜔𝑑subscript𝜔32𝑡superscriptsubscript𝑘1𝑁superscriptsubscript𝜎𝑘23superscript𝑒𝑖subscript𝜔𝑑subscript𝜔32𝑡superscriptsubscript𝑘1𝑁superscriptsubscript𝜎𝑘32\hat{H}_{d}=\hbar\Omega[e^{i\left(\omega_{d}-\omega_{32}\right)t}\sum_{k=1}^{N% }\sigma_{k}^{23}+e^{-i\left(\omega_{d}-\omega_{32}\right)t}\sum_{k=1}^{N}% \sigma_{k}^{32}]over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = roman_ℏ roman_Ω [ italic_e start_POSTSUPERSCRIPT italic_i ( italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT ) italic_t end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 23 end_POSTSUPERSCRIPT + italic_e start_POSTSUPERSCRIPT - italic_i ( italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT ) italic_t end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 32 end_POSTSUPERSCRIPT ] describes the driving of the atoms by a dressing laser of frequency ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT with a strength ΩΩ\Omegaroman_Ω, where σk23,σk32superscriptsubscript𝜎𝑘23superscriptsubscript𝜎𝑘32\sigma_{k}^{23},\sigma_{k}^{32}italic_σ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 23 end_POSTSUPERSCRIPT , italic_σ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 32 end_POSTSUPERSCRIPT are the lowering and raising operators related to the 2k3ksubscript2𝑘subscript3𝑘2_{k}\to 3_{k}2 start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT → 3 start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT transition of frequency ω32subscript𝜔32\omega_{32}italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT. The remaining terms of Eq. (1) describe the system dissipation with the Lindblad superoperator 𝒟[o^]ρ^=12(o^o^ρ^+ρ^o^o^)o^ρ^o^𝒟delimited-[]^𝑜^𝜌12superscript^𝑜^𝑜^𝜌^𝜌superscript^𝑜^𝑜^𝑜^𝜌superscript^𝑜\mathcal{D}\left[\hat{o}\right]\hat{\rho}=\frac{1}{2}\left(\hat{o}^{\dagger}% \hat{o}\hat{\rho}+\hat{\rho}\hat{o}^{\dagger}\hat{o}\right)-\hat{o}\hat{\rho}% \hat{o}^{\dagger}caligraphic_D [ over^ start_ARG italic_o end_ARG ] over^ start_ARG italic_ρ end_ARG = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( over^ start_ARG italic_o end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_o end_ARG over^ start_ARG italic_ρ end_ARG + over^ start_ARG italic_ρ end_ARG over^ start_ARG italic_o end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_o end_ARG ) - over^ start_ARG italic_o end_ARG over^ start_ARG italic_ρ end_ARG over^ start_ARG italic_o end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT (for any operator o^^𝑜\hat{o}over^ start_ARG italic_o end_ARG), which includes the photon loss of the cavity with a rate κ𝜅\kappaitalic_κ, the spontaneous emission and the incoherent pumping of atoms with the rates γ31,γ12subscript𝛾31subscript𝛾12\gamma_{31},\gamma_{12}italic_γ start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT , italic_γ start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT. The incoherent pumping can be implemented by coupling the multiple-level atoms with a sequence of laser pulses, as implemented in the experiment (JDBohnet2012, ). Here, we consider the rather complex process phenomenologically with a dissipative process, which is the inverse of the spontaneous emission and transfers incoherently the population from the lower to higher hyper-fine ground level. For the sake of simplicity, we have also ignored other possible decay and dephasing processes, but emphasize that they can be easily incorporated into the master equation. The Raman scattering spectrum can be calculated from the expression S(ω)=2κRe[𝑑τeiωτa^(τ)a^(0)]S𝜔2𝜅Redelimited-[]differential-d𝜏superscript𝑒𝑖𝜔𝜏delimited-⟨⟩superscript^𝑎𝜏^𝑎0{\rm S}(\omega)=2\kappa{\rm Re}[\int d\tau e^{-i\omega\tau}\left\langle\hat{a}% ^{\dagger}(\tau)\hat{a}(0)\right\rangle]roman_S ( italic_ω ) = 2 italic_κ roman_Re [ ∫ italic_d italic_τ italic_e start_POSTSUPERSCRIPT - italic_i italic_ω italic_τ end_POSTSUPERSCRIPT ⟨ over^ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_τ ) over^ start_ARG italic_a end_ARG ( 0 ) ⟩ ], where the equation for the two-time correlation function (the integral kernel) can be obtained by applying quantum regression theorem (DMeiser2009, ).

The experiment of J. G. Bohnet et al. is motivated by that the dynamics for the current system in weak coupling regime is similar to the superradiant laser system. To see this point, we eliminate adiabatically the excited level by defining the slowly varying term σ¯^k13ei(ωlω32)tσ^k13superscriptsubscript^¯𝜎𝑘13superscript𝑒𝑖subscript𝜔𝑙subscript𝜔32𝑡superscriptsubscript^𝜎𝑘13\hat{\bar{\sigma}}_{k}^{13}\equiv e^{i(\omega_{l}-\omega_{32})t}\hat{\sigma}_{% k}^{13}over^ start_ARG over¯ start_ARG italic_σ end_ARG end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 13 end_POSTSUPERSCRIPT ≡ italic_e start_POSTSUPERSCRIPT italic_i ( italic_ω start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT ) italic_t end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 13 end_POSTSUPERSCRIPT, and deriving the corresponding Heisenberg equation tσ¯^k13i(ωdω32+iγ31/2)σ¯^k13iΩσ^k12subscript𝑡superscriptsubscript^¯𝜎𝑘13𝑖subscript𝜔𝑑subscript𝜔32𝑖subscript𝛾312superscriptsubscript^¯𝜎𝑘13𝑖Ωsuperscriptsubscript^𝜎𝑘12\partial_{t}\hat{\bar{\sigma}}_{k}^{13}\approx i(\omega_{d}-\omega_{32}+i% \gamma_{31}/2)\hat{\bar{\sigma}}_{k}^{13}-i\Omega\hat{\sigma}_{k}^{12}∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT over^ start_ARG over¯ start_ARG italic_σ end_ARG end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 13 end_POSTSUPERSCRIPT ≈ italic_i ( italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT + italic_i italic_γ start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT / 2 ) over^ start_ARG over¯ start_ARG italic_σ end_ARG end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 13 end_POSTSUPERSCRIPT - italic_i roman_Ω over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT. Then, we solve this equation at steady-state to obtain σ¯^k13[Ω/(ωdω32+iγ31/2)]σ^k12superscriptsubscript^¯𝜎𝑘13delimited-[]Ωsubscript𝜔𝑑subscript𝜔32𝑖subscript𝛾312superscriptsubscript^𝜎𝑘12\hat{\bar{\sigma}}_{k}^{13}\approx[\Omega/(\omega_{d}-\omega_{32}+i\gamma_{31}% /2)]\hat{\sigma}_{k}^{12}over^ start_ARG over¯ start_ARG italic_σ end_ARG end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 13 end_POSTSUPERSCRIPT ≈ [ roman_Ω / ( italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT + italic_i italic_γ start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT / 2 ) ] over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT, and insert this solution into Eq. (1) to obtain an effective quantum master equation:

tρ^=i[H^ac(t),ρ^]κ𝒟[a^]ρ^subscript𝑡^𝜌𝑖Planck-constant-over-2-pisubscriptsuperscript^𝐻𝑎𝑐𝑡^𝜌𝜅𝒟delimited-[]^𝑎^𝜌\displaystyle\partial_{t}\hat{\rho}=-\frac{i}{\hbar}[\hat{H}^{\prime}_{a-c}(t)% ,\hat{\rho}]-\kappa\mathcal{D}[\hat{a}]\hat{\rho}∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT over^ start_ARG italic_ρ end_ARG = - divide start_ARG italic_i end_ARG start_ARG roman_ℏ end_ARG [ over^ start_ARG italic_H end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_a - italic_c end_POSTSUBSCRIPT ( italic_t ) , over^ start_ARG italic_ρ end_ARG ] - italic_κ caligraphic_D [ over^ start_ARG italic_a end_ARG ] over^ start_ARG italic_ρ end_ARG
k=1N(γ21𝒟[σ^k12]ρ^+γ12𝒟[σ^k21]ρ^).superscriptsubscript𝑘1𝑁subscript𝛾21𝒟delimited-[]superscriptsubscript^𝜎𝑘12^𝜌subscript𝛾12𝒟delimited-[]superscriptsubscript^𝜎𝑘21^𝜌\displaystyle-\sum_{k=1}^{N}(\gamma_{21}\mathcal{D}[\hat{\sigma}_{k}^{12}]\hat% {\rho}+\gamma_{12}\mathcal{D}\left[\hat{\sigma}_{k}^{21}\right]\hat{\rho}).- ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( italic_γ start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT caligraphic_D [ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT ] over^ start_ARG italic_ρ end_ARG + italic_γ start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT caligraphic_D [ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT ] over^ start_ARG italic_ρ end_ARG ) . (2)

In this equation, the effective Hamiltonian H^ac(t)=g21[ei(ωcω31ωd+ω32)ta^k=1Nσ^k12+h.c.]\hat{H}^{\prime}_{a-c}(t)=\hbar g_{21}\left[e^{i(\omega_{c}-\omega_{31}-\omega% _{d}+\omega_{32})t}\hat{a}^{\dagger}\sum_{k=1}^{N}\hat{\sigma}_{k}^{12}+{\rm h% .c.}\right]over^ start_ARG italic_H end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_a - italic_c end_POSTSUBSCRIPT ( italic_t ) = roman_ℏ italic_g start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT [ italic_e start_POSTSUPERSCRIPT italic_i ( italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT ) italic_t end_POSTSUPERSCRIPT over^ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT + roman_h . roman_c . ] describes the coherent energy exchange between the atomic transition 1k2ksubscript1𝑘subscript2𝑘1_{k}\to 2_{k}1 start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT → 2 start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT and the optical cavity with the strength g21=(g31Ω)/(ω32ωdiγ31/2)subscript𝑔21subscript𝑔31Ωsubscript𝜔32subscript𝜔𝑑𝑖subscript𝛾312g_{21}=-(g_{31}\Omega)/(\omega_{32}-\omega_{d}-i\gamma_{31}/2)italic_g start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT = - ( italic_g start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT roman_Ω ) / ( italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - italic_i italic_γ start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT / 2 ). The emerging Lindblad term in Eq. (2) describes the Raman transition-induced decay with the rate γ21=γ31Ω2/|ω32ωdiγ31/2|2subscript𝛾21subscript𝛾31superscriptΩ2superscriptsubscript𝜔32subscript𝜔𝑑𝑖subscript𝛾3122\gamma_{21}=\gamma_{31}\Omega^{2}/|\omega_{32}-\omega_{d}-i\gamma_{31}/2|^{2}italic_γ start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT = italic_γ start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / | italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - italic_i italic_γ start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT / 2 | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Eq. (2) has a similar form as the quantum master equation for the superradiant laser (DMeiser2009, ), except that g21subscript𝑔21g_{21}italic_g start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT and γ21subscript𝛾21\gamma_{21}italic_γ start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT can be actively controlled by adjusting the parameters Ω,ωdΩsubscript𝜔𝑑\Omega,\omega_{d}roman_Ω , italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT of the dressed laser.

To simulate systems with tens of thousands of atoms, we solve the quantum master equation (1) and (2) with the cumulant mean-field approach. In this approach, we derive the equations to^=tr{(tρ^)o^}subscript𝑡delimited-⟨⟩^𝑜trsubscript𝑡^𝜌^𝑜\partial_{t}\left\langle\hat{o}\right\rangle={\rm tr}\{(\partial_{t}\hat{\rho}% )\hat{o}\}∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ⟨ over^ start_ARG italic_o end_ARG ⟩ = roman_tr { ( ∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT over^ start_ARG italic_ρ end_ARG ) over^ start_ARG italic_o end_ARG } for the mean values o^delimited-⟨⟩^𝑜\left\langle\hat{o}\right\rangle⟨ over^ start_ARG italic_o end_ARG ⟩ (for any operator o^^𝑜\hat{o}over^ start_ARG italic_o end_ARG), and obtain a hierarchy of inter-dependent equations for the mean values of many operators due to the interaction or the collective process, and apply the cumulant expansion approximation, e.g. o^p^q^o^p^q^+p^o^q^+q^o^p^2o^p^q^delimited-⟨⟩^𝑜^𝑝^𝑞delimited-⟨⟩^𝑜delimited-⟨⟩^𝑝^𝑞delimited-⟨⟩^𝑝delimited-⟨⟩^𝑜^𝑞delimited-⟨⟩^𝑞delimited-⟨⟩^𝑜^𝑝2delimited-⟨⟩^𝑜delimited-⟨⟩^𝑝delimited-⟨⟩^𝑞\left\langle\hat{o}\hat{p}\hat{q}\right\rangle\approx\left\langle\hat{o}\right% \rangle\left\langle\hat{p}\hat{q}\right\rangle+\left\langle\hat{p}\right% \rangle\left\langle\hat{o}\hat{q}\right\rangle+\left\langle\hat{q}\right% \rangle\left\langle\hat{o}\hat{p}\right\rangle-2\left\langle\hat{o}\right% \rangle\left\langle\hat{p}\right\rangle\left\langle\hat{q}\right\rangle⟨ over^ start_ARG italic_o end_ARG over^ start_ARG italic_p end_ARG over^ start_ARG italic_q end_ARG ⟩ ≈ ⟨ over^ start_ARG italic_o end_ARG ⟩ ⟨ over^ start_ARG italic_p end_ARG over^ start_ARG italic_q end_ARG ⟩ + ⟨ over^ start_ARG italic_p end_ARG ⟩ ⟨ over^ start_ARG italic_o end_ARG over^ start_ARG italic_q end_ARG ⟩ + ⟨ over^ start_ARG italic_q end_ARG ⟩ ⟨ over^ start_ARG italic_o end_ARG over^ start_ARG italic_p end_ARG ⟩ - 2 ⟨ over^ start_ARG italic_o end_ARG ⟩ ⟨ over^ start_ARG italic_p end_ARG ⟩ ⟨ over^ start_ARG italic_q end_ARG ⟩ (for any operators o^,p^,q^^𝑜^𝑝^𝑞\hat{o},\hat{p},\hat{q}over^ start_ARG italic_o end_ARG , over^ start_ARG italic_p end_ARG , over^ start_ARG italic_q end_ARG), to remove the hierarchy and obtain a closed set of equations. If all the atoms are identical the terms σ^klmdelimited-⟨⟩subscriptsuperscript^𝜎𝑙𝑚𝑘\left\langle\hat{\sigma}^{lm}_{k}\right\rangle⟨ over^ start_ARG italic_σ end_ARG start_POSTSUPERSCRIPT italic_l italic_m end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ⟩, a^σ^k23delimited-⟨⟩superscript^𝑎superscriptsubscript^𝜎𝑘23\left\langle\hat{a}^{\dagger}\hat{\sigma}_{k}^{23}\right\rangle⟨ over^ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 23 end_POSTSUPERSCRIPT ⟩ are the same for all the atoms, and the terms σ^klmσ^klmdelimited-⟨⟩superscriptsubscript^𝜎𝑘𝑙𝑚superscriptsubscript^𝜎superscript𝑘superscript𝑙superscript𝑚\left\langle\hat{\sigma}_{k}^{lm}\hat{\sigma}_{k^{\prime}}^{l^{\prime}m^{% \prime}}\right\rangle⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_l italic_m end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_l start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ⟩ are same for all atom pairs (k,k)𝑘superscript𝑘(k,k^{\prime})( italic_k , italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ). In this way, we can reduce the number of coupled equations from the order of N3similar-toabsentsuperscript𝑁3\sim N^{3}∼ italic_N start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT to a few tens. In practice, we employ the QuantumCumulant.jl package (DPlankensteiner, ) to implement the above procedure, and summarize the corresponding codes in Appendix A. To calculate the superradiant Raman scattering spectrum, we have also reformulated the codes to derive the equations for the two-time correlation functions, and ensured that they work also in the interaction picture as employed here.

We employ the Dicke state picture to analyze the collective dynamics of the atomic ensemble. To this end, we define firstly the collective operators J^x(y)=1(i)2k=1N(σ^k12±σ^k21)subscript^𝐽𝑥𝑦1𝑖2superscriptsubscript𝑘1𝑁plus-or-minussuperscriptsubscript^𝜎𝑘12superscriptsubscript^𝜎𝑘21\hat{J}_{x(y)}=\frac{1(i)}{2}\sum_{k=1}^{N}\left(\hat{\sigma}_{k}^{12}\pm\hat{% \sigma}_{k}^{21}\right)over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_x ( italic_y ) end_POSTSUBSCRIPT = divide start_ARG 1 ( italic_i ) end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT ± over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT ),J^z=12k=1N(2σ^k221^k)subscript^𝐽𝑧12superscriptsubscript𝑘1𝑁2superscriptsubscript^𝜎𝑘22subscript^1𝑘\hat{J}_{z}=\frac{1}{2}\sum_{k=1}^{N}\left(2\hat{\sigma}_{k}^{22}-\hat{1}_{k}\right)over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( 2 over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT - over^ start_ARG 1 end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ), and their square J^x(y)2=14kN1^k+14kkN(σ^k12σ^k12±σ^k12σ^k21±σ^k21σ^k12+σ^k21σ^k21)superscriptsubscript^𝐽𝑥𝑦214superscriptsubscript𝑘𝑁subscript^1𝑘14superscriptsubscript𝑘superscript𝑘𝑁plus-or-minussuperscriptsubscript^𝜎𝑘12superscriptsubscript^𝜎superscript𝑘12superscriptsubscript^𝜎𝑘12superscriptsubscript^𝜎superscript𝑘21superscriptsubscript^𝜎𝑘21superscriptsubscript^𝜎superscript𝑘12superscriptsubscript^𝜎𝑘21superscriptsubscript^𝜎superscript𝑘21\hat{J}_{x\left(y\right)}^{2}=\frac{1}{4}\sum_{k}^{N}\hat{1}_{k}+\frac{1}{4}% \sum_{k\neq k^{\prime}}^{N}(\hat{\sigma}_{k}^{12}\hat{\sigma}_{k^{\prime}}^{12% }\pm\hat{\sigma}_{k}^{12}\hat{\sigma}_{k^{\prime}}^{21}\pm\hat{\sigma}_{k}^{21% }\hat{\sigma}_{k^{\prime}}^{12}+\hat{\sigma}_{k}^{21}\hat{\sigma}_{k^{\prime}}% ^{21})over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_x ( italic_y ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG 4 end_ARG ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT over^ start_ARG 1 end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG 4 end_ARG ∑ start_POSTSUBSCRIPT italic_k ≠ italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT ± over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT ± over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT + over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT ), and J^z2=14kN1^k+14kkN(4σ^k22σ^k222σ^k221^k21^kσ^k22+1^k1^k)superscriptsubscript^𝐽𝑧214superscriptsubscript𝑘𝑁subscript^1𝑘14superscriptsubscript𝑘superscript𝑘𝑁4superscriptsubscript^𝜎𝑘22superscriptsubscript^𝜎superscript𝑘222superscriptsubscript^𝜎𝑘22subscript^1superscript𝑘2subscript^1𝑘superscriptsubscript^𝜎superscript𝑘22subscript^1𝑘subscript^1superscript𝑘\hat{J}_{z}^{2}=\frac{1}{4}\sum_{k}^{N}\hat{1}_{k}+\frac{1}{4}\sum_{k\neq k^{% \prime}}^{N}(4\hat{\sigma}_{k}^{22}\hat{\sigma}_{k^{\prime}}^{22}-2\hat{\sigma% }_{k}^{22}\hat{1}_{k^{\prime}}-2\hat{1}_{k}\hat{\sigma}_{k^{\prime}}^{22}+\hat% {1}_{k}\hat{1}_{k^{\prime}})over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG 4 end_ARG ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT over^ start_ARG 1 end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG 4 end_ARG ∑ start_POSTSUBSCRIPT italic_k ≠ italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( 4 over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT - 2 over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT over^ start_ARG 1 end_ARG start_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT - 2 over^ start_ARG 1 end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT + over^ start_ARG 1 end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT over^ start_ARG 1 end_ARG start_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ). Here, 1^ksubscript^1𝑘\hat{1}_{k}over^ start_ARG 1 end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT are the identity operators. Then, we define the Dicke states as the eigen states of the equations (i=x,yzJ^i2)|J,M=J(J+1)|J,Msubscript𝑖𝑥𝑦𝑧superscriptsubscript^𝐽𝑖2ket𝐽𝑀𝐽𝐽1ket𝐽𝑀\left(\sum_{i=x,yz}\hat{J}_{i}^{2}\right)\left|J,M\right\rangle=J\left(J+1% \right)\left|J,M\right\rangle( ∑ start_POSTSUBSCRIPT italic_i = italic_x , italic_y italic_z end_POSTSUBSCRIPT over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) | italic_J , italic_M ⟩ = italic_J ( italic_J + 1 ) | italic_J , italic_M ⟩, J^z|J,M=M|J,Msubscript^𝐽𝑧ket𝐽𝑀𝑀ket𝐽𝑀\hat{J}_{z}\left|J,M\right\rangle=M\left|J,M\right\rangleover^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | italic_J , italic_M ⟩ = italic_M | italic_J , italic_M ⟩, where the integer or half-integer J=0,,N/2𝐽0𝑁2J=0,...,N/2italic_J = 0 , … , italic_N / 2 indicates the degree of symmetry (the coupling strength) and the number M𝑀Mitalic_M in the range JMJ𝐽𝑀𝐽-J\leq M\leq J- italic_J ≤ italic_M ≤ italic_J labels the degree of the excitation. Inspired by the above equations, we introduce the mean values J¯,M¯¯𝐽¯𝑀\bar{J},\bar{M}over¯ start_ARG italic_J end_ARG , over¯ start_ARG italic_M end_ARG through the equations J¯(J¯+1)=i=x,yzJ^i2¯𝐽¯𝐽1subscript𝑖𝑥𝑦𝑧delimited-⟨⟩superscriptsubscript^𝐽𝑖2\bar{J}\left(\bar{J}+1\right)=\sum_{i=x,yz}\left\langle\hat{J}_{i}^{2}\right\rangleover¯ start_ARG italic_J end_ARG ( over¯ start_ARG italic_J end_ARG + 1 ) = ∑ start_POSTSUBSCRIPT italic_i = italic_x , italic_y italic_z end_POSTSUBSCRIPT ⟨ over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ and M¯=J^z¯𝑀delimited-⟨⟩subscript^𝐽𝑧\bar{M}=\left\langle\hat{J}_{z}\right\rangleover¯ start_ARG italic_M end_ARG = ⟨ over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩. Note that all the processes of two-level systems, as described by the effective quantum master equation, have been identified in the Dicke state picture  (ShammahN, ; ZhangY2018, ). If all the atoms are identical, the expectation values of the collective operators can be calculated from the expressions J^x(y)2=14N±14N(N1)(σ^112σ^212±σ^112σ^221±σ^121σ^212+σ^121σ^221)delimited-⟨⟩superscriptsubscript^𝐽𝑥𝑦2plus-or-minus14𝑁14𝑁𝑁1plus-or-minusdelimited-⟨⟩superscriptsubscript^𝜎112superscriptsubscript^𝜎212delimited-⟨⟩superscriptsubscript^𝜎112superscriptsubscript^𝜎221delimited-⟨⟩superscriptsubscript^𝜎121superscriptsubscript^𝜎212delimited-⟨⟩superscriptsubscript^𝜎121superscriptsubscript^𝜎221\left\langle\hat{J}_{x\left(y\right)}^{2}\right\rangle=\frac{1}{4}N\pm\frac{1}% {4}N\left(N-1\right)(\left\langle\hat{\sigma}_{1}^{12}\hat{\sigma}_{2}^{12}% \right\rangle\pm\left\langle\hat{\sigma}_{1}^{12}\hat{\sigma}_{2}^{21}\right% \rangle\pm\left\langle\hat{\sigma}_{1}^{21}\hat{\sigma}_{2}^{12}\right\rangle+% \left\langle\hat{\sigma}_{1}^{21}\hat{\sigma}_{2}^{21}\right\rangle)⟨ over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_x ( italic_y ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_N ± divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_N ( italic_N - 1 ) ( ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT ⟩ ± ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT ⟩ ± ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT ⟩ + ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT ⟩ ), and J^z2=14N+14N(N1)(4σ^122σ^2224σ^122+1)delimited-⟨⟩superscriptsubscript^𝐽𝑧214𝑁14𝑁𝑁14delimited-⟨⟩superscriptsubscript^𝜎122superscriptsubscript^𝜎2224delimited-⟨⟩superscriptsubscript^𝜎1221\left\langle\hat{J}_{z}^{2}\right\rangle=\frac{1}{4}N+\frac{1}{4}N\left(N-1% \right)\left(4\left\langle\hat{\sigma}_{1}^{22}\hat{\sigma}_{2}^{22}\right% \rangle-4\left\langle\hat{\sigma}_{1}^{22}\right\rangle+1\right)⟨ over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ = divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_N + divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_N ( italic_N - 1 ) ( 4 ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT ⟩ - 4 ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT ⟩ + 1 ), as well as J^x=12N(σ^112+σ^121),J^y=i2N(σ^112σ^121),J^z=12N(2σ^1221)formulae-sequencedelimited-⟨⟩subscript^𝐽𝑥12𝑁delimited-⟨⟩superscriptsubscript^𝜎112delimited-⟨⟩superscriptsubscript^𝜎121formulae-sequencedelimited-⟨⟩subscript^𝐽𝑦𝑖2𝑁delimited-⟨⟩superscriptsubscript^𝜎112delimited-⟨⟩superscriptsubscript^𝜎121delimited-⟨⟩subscript^𝐽𝑧12𝑁2delimited-⟨⟩superscriptsubscript^𝜎1221\left\langle\hat{J}_{x}\right\rangle=\frac{1}{2}N(\left\langle\hat{\sigma}_{1}% ^{12}\right\rangle+\left\langle\hat{\sigma}_{1}^{21}\right\rangle),\left% \langle\hat{J}_{y}\right\rangle=\frac{i}{2}N(\left\langle\hat{\sigma}_{1}^{12}% \right\rangle-\left\langle\hat{\sigma}_{1}^{21}\right\rangle),\left\langle\hat% {J}_{z}\right\rangle=\frac{1}{2}N(2\left\langle\hat{\sigma}_{1}^{22}\right% \rangle-1)⟨ over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ⟩ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_N ( ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT ⟩ + ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT ⟩ ) , ⟨ over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ⟩ = divide start_ARG italic_i end_ARG start_ARG 2 end_ARG italic_N ( ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT ⟩ - ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT ⟩ ) , ⟨ over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⟩ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_N ( 2 ⟨ over^ start_ARG italic_σ end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT ⟩ - 1 ). As a complementary picture of the collective process, we can also calculate the collective spin vector (also known as Bloch vector), 𝐀=i=x,y,zJ^i𝐞i𝐀subscript𝑖𝑥𝑦𝑧delimited-⟨⟩subscript^𝐽𝑖subscript𝐞𝑖\mathbf{A}=\sum_{i=x,y,z}\left\langle\hat{J}_{i}\right\rangle\mathbf{e}_{i}bold_A = ∑ start_POSTSUBSCRIPT italic_i = italic_x , italic_y , italic_z end_POSTSUBSCRIPT ⟨ over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ bold_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT with the expectation values J^i=α=1,2J^idelimited-⟨⟩subscript^𝐽𝑖subscript𝛼12delimited-⟨⟩subscript^𝐽𝑖\left\langle\hat{J}_{i}\right\rangle=\sum_{\alpha=1,2}\left\langle\hat{J}_{i}\right\rangle⟨ over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ = ∑ start_POSTSUBSCRIPT italic_α = 1 , 2 end_POSTSUBSCRIPT ⟨ over^ start_ARG italic_J end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⟩ and the unit vectors 𝐞isubscript𝐞𝑖\mathbf{e}_{i}bold_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of Cartesian coordinate system. However, our calculations show that the vector components Ax,Aysubscript𝐴𝑥subscript𝐴𝑦A_{x},A_{y}italic_A start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_A start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT are always zero within the pulsed and continuous superradiant Raman scattering, and thus the collective spin vector is not a good picture to analyze the cause of these scattering phenomena.

Before presenting the numerical results, we comment shortly on the used parameters. The optical cavity has a frequency ωc=2π×(3.77×1014+6.8×109)subscript𝜔𝑐2𝜋3.77superscript10146.8superscript109\omega_{c}=2\pi\times(3.77\times{10}^{14}+6.8\times{10}^{9})italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 2 italic_π × ( 3.77 × 10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT + 6.8 × 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT ) Hz and a damping rate κ=2π×11𝜅2𝜋11\kappa=2\pi\times 11italic_κ = 2 italic_π × 11 MHz. The atoms couple with the cavity with a strength g31=2π×506subscript𝑔312𝜋506g_{31}=2\pi\times 506italic_g start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT = 2 italic_π × 506 kHz, and have the transition frequencies ω32=2π×(3.77×10142×109)subscript𝜔322𝜋3.77superscript10142superscript109\omega_{32}=2\pi\times(3.77\times{10}^{14}-2\times{10}^{9})italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT = 2 italic_π × ( 3.77 × 10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT - 2 × 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT ) GHz, ω21=2π×6.8subscript𝜔212𝜋6.8\omega_{21}=2\pi\times 6.8italic_ω start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT = 2 italic_π × 6.8 GHz, and the decay rate γ31=2π×5.75subscript𝛾312𝜋5.75\gamma_{31}=2\pi\times 5.75italic_γ start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT = 2 italic_π × 5.75 MHz. The dressed laser with a frequency ωd=ω32+2π×2subscript𝜔𝑑subscript𝜔322𝜋2\omega_{d}=\omega_{32}+2\pi\times 2italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT + 2 italic_π × 2 GHz couples with the atoms with a strength Ω=2π×5Ω2𝜋5\Omega=2\pi\times 5roman_Ω = 2 italic_π × 5 MHz.

Refer to caption
Figure 2: Superradiant Raman scattering pulses for systems within the crossover coupling regime. Panel (a) shows the dynamics of the intra-cavity photon number, and of the atomic ensemble in the Dicke states space (inset). Here, the average Dicke numbers are defined with respect to the two hyper-fine ground levels. Panel (b) shows the evolution of the maximum, width and delay time of the pulses as function of the number of atoms. Panel (c,d) show the similar results as in the panel (b) except that the strength ΩΩ\Omegaroman_Ω and frequency detuning Δ=ωlω32Δsubscript𝜔𝑙subscript𝜔32\Delta=\omega_{l}-\omega_{32}roman_Δ = italic_ω start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT of the dressed laser are varied. Here, the atoms are assumed to be initially on the higher hyper-fine ground level, and the key reference parameters are N=104𝑁superscript104N=10^{4}italic_N = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT, Ω=2π×5Ω2𝜋5\Omega=2\pi\times 5roman_Ω = 2 italic_π × 5 MHz, and ωd=ω32+2π×2GHzsubscript𝜔𝑑subscript𝜔322𝜋2GHz\omega_{d}=\omega_{32}+2\pi\times 2{\rm GHz}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT + 2 italic_π × 2 roman_G roman_H roman_z.

III Superradiant Raman Scattering Pulses

In the following, we study the superradiant Raman scattering pulses, and find the different behaviors for the systems in the crossover regime (Fig. 2) and the strong coupling regime (Fig. 3). In the former regime, the collective coupling Ng31𝑁subscript𝑔31\sqrt{N}g_{31}square-root start_ARG italic_N end_ARG italic_g start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT is slightly larger than the photon loss rate κ𝜅\kappaitalic_κ, and the dressed laser drives directly the bare atoms. In the latter regime, Ng31𝑁subscript𝑔31\sqrt{N}g_{31}square-root start_ARG italic_N end_ARG italic_g start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT is significantly larger than κ𝜅\kappaitalic_κ, and the dressed laser derives the atoms-cavity system through the formed atom-photon dressed states. Here, we do not discuss the response for the system in the weak coupling regime, since it is similar to that for the system in the crossover regime.

We firstly focus on the systems in the crossover regime. We find that once the dressing laser excites the atoms initially on the higher hyper-fine ground level, the intra-cavity photon number increases firstly and then decreases, forming a pulse [Fig. 2(a)]. Accompanying with this evolution, the population transfers from the higher to the lower hyper-fine ground level (not shown). Since the excited level is not populated at all, this is clearly a Raman process. As a complement to the population change, the ensemble occupies initially the top-right corner of the Dicke states space, and declines vertically with a finite distance from the right boundary to the lower boundary [inset of Fig. 2(a)], which indicates the superradiant origin of the observed pulses. To prove this point further, we examine also the maximum, width and delay time of the pulses as function of number of atoms [Fig. 2(b)]. We find that the maximum increases quadratically, the width and delay time decrease inversely, which are the characteristics of the superradiant pulses (MANorcia2016, ). The same results are also achieved with the effective quantum master equation (not shown).

The above results verify that the observed phenomena are superradiant pulses. However, in contrast to the systems with the true atomic transition, here, the collective coupling with the cavity mode is mediated by Raman scattering, and can be controlled by the dressed laser. To demonstrate this point, we study further how the dressed laser parameters Ω,ωdΩsubscript𝜔𝑑\Omega,\omega_{d}roman_Ω , italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT affect the superradiant pulses [Fig. 2(c,d)]. We find that as the driving strength ΩΩ\Omegaroman_Ω increases, the pulse maximum increases quadratically (Ω2similar-toabsentsuperscriptΩ2\sim\Omega^{2}∼ roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT), while the pulse width and delay time decrease quadratically (Ω2similar-toabsentsuperscriptΩ2\sim\Omega^{-2}∼ roman_Ω start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT). On the contrary, as the detuning ωdω32subscript𝜔𝑑subscript𝜔32\omega_{d}-\omega_{32}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT increases, the pulse maximum decreases quadratically, while the pulse width and delay time increase quadratically. All these can be understood by examining effective master equation (2). In this equation, the effective coupling strength g21subscript𝑔21g_{21}italic_g start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT is proportional roughly to the ratio Ω/(ωdω32)Ωsubscript𝜔𝑑subscript𝜔32\Omega/(\omega_{d}-\omega_{32})roman_Ω / ( italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT ), and the effective Purcell-enhanced decay rate Γ4g212/κΓ4superscriptsubscript𝑔212𝜅\Gamma\approx 4g_{21}^{2}/\kapparoman_Γ ≈ 4 italic_g start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_κ scales quadratically with this ratio. Although the effective decay rate γ21subscript𝛾21\gamma_{21}italic_γ start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT scales also quadratically with this ratio, it is relatively small than the collective decay rates (J+M)(JM+1)Γ𝐽𝑀𝐽𝑀1Γ(J+M)(J-M+1)\Gamma( italic_J + italic_M ) ( italic_J - italic_M + 1 ) roman_Γ of the Dicke states |J,Mket𝐽𝑀\left|J,M\right\rangle| italic_J , italic_M ⟩ with JN/2𝐽𝑁2J\approx N/2italic_J ≈ italic_N / 2.

Refer to caption
Figure 3: Superradiant Raman scattering pulses for systems within the strong coupling regime. Panel (a-c) show the similar results as in Fig. 2(a,b,d) except that the decay time of the tail is considered instead of the pulse width. Here, the decay time is defined as the time to reach the half of the maximum value. Panel (d) shows the photon number for the systems with 5×1075superscript1075\times 10^{7}5 × 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT atoms. The key reference parameters are N=106𝑁superscript106N=10^{6}italic_N = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT, Ω=2π×5Ω2𝜋5\Omega=2\pi\times 5roman_Ω = 2 italic_π × 5 MHz, and ωd=ω32+2π×2GHzsubscript𝜔𝑑subscript𝜔322𝜋2GHz\omega_{d}=\omega_{32}+2\pi\times 2{\rm GHz}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT + 2 italic_π × 2 roman_G roman_H roman_z.

We then consider the systems in the strong coupling regime. In this case, after the switch-on of the dressed laser, the intra-cavity photon number increases dramatically in short time, but decays slowly in long time, resulting to a distorted pulse [Fig. 3(a)]. Following this dynamics, the population transfer from the higher to lower hyper-fine ground level is fast in short time, but becomes slow in long time (not shown). Here, the atomic ensemble decays also vertically, but the evolution is much closer to the right boundary [inset of Fig. 3(a)]. The further analysis indicates the pulse maximum scales linearly with number of atoms, the pulse delay time and the slow decay time of the pulse decrease inversely [Fig. 3(b)]. With the effective master equation (2), we obtain the similar results as before (Fig. A4 in the Appendix). Thus, the observed distorted pulse goes beyond the scope of the effective model, and originates from the full model.

We study further the dependence of the distorted pulses on the dressed laser parameters Ω,ωdΩsubscript𝜔𝑑\Omega,\omega_{d}roman_Ω , italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT. The pulse maximum, delay and decay time behave similarly for different dressed strength ΩΩ\Omegaroman_Ω (Fig. A5 of the Appendix) and frequency detunning ωdω32subscript𝜔𝑑subscript𝜔32\omega_{d}-\omega_{32}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT as before except that the pulse maximum scales inversely with the frequency detunning [Fig. 3(c)]. For the systems with more atoms, we find that the decay at the longer time evolves into the decayed oscillations [Fig. 3(d)]. Since this oscillation resembles the Rabi oscillations in the strong coupling regime (MANorcia2016-1, ), we attribute the distorted pulses to the behavior in the strong coupling regime. We note that the distorted pulses agree qualitatively with Fig. 1(c) of the experimental paper (JDBohnet2012, ), while the quadratic dependence of the pulse maximum agrees qualitatively with Fig. 1(e) of that paper. This comparison indicates that these results in the experiment might be caused by the systems in the different regimes.

Refer to caption
Figure 4: Continuous superradiant Raman scattering for systems within the crossover coupling regime. Panel (a) shows the evolution of the intra-cavity photon number and the atomic ensemble in the Dicke state space (inset). Panel (b) compares the shifted and broad Raman scattering spectrum computed with the full model with the centered and narrower spectrum calculated with the effective model (inset). Panel (c,d) show the steady-state intra-cavity photon number (left axes), the spectral frequency shift and linewidth (right axes) as function of the incoherent pumping rate and the strength of the dressed laser, respectively. In the panel (c), the vertical lines show NΓ/2𝑁Γ2N\Gamma/2italic_N roman_Γ / 2 and NΓ𝑁ΓN\Gammaitalic_N roman_Γ with the Purcell-enhanced decay rate Γ4g122/κΓ4superscriptsubscript𝑔122𝜅\Gamma\approx 4g_{12}^{2}/\kapparoman_Γ ≈ 4 italic_g start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_κ. Here, we assume that the atoms are incoherently pumped from the lower to higher hyper-fine ground level.

IV Continuous Superradiant Raman Scattering

In the following, we consider the systems in the presence of the incoherent atomic pumping. Similar to the above discussion, here, we also distinguish the systems in the crossover (Fig. 4) and strong coupling regime (Fig. A7), and focus on the former case firstly. In the presence of the pumping, the intra-cavity photon number demonstrates also several peaks after the first main peak, and evolves eventually to a constant value about 0.135 [Fig. 4(a)]. In accordance with this, the atomic ensemble decays vertically after climbing the lower and upper boundary, and reaches eventually to a point slightly above the left-most corner after repeating the similar dynamics [inset of Fig. 4(a)]. When the atomic ensemble reaches the steady-state, we can calculate the corresponding Raman scattering spectrum [Fig. 4(b)]. The spectrum shows a peak with a frequency about 2π×12.42𝜋12.42\pi\times 12.42 italic_π × 12.4 kHz shifted from the frequency of the optical cavity mode, and a linewidth of 2π×3.242𝜋3.242\pi\times 3.242 italic_π × 3.24 kHz. If we consider the effective model, we obtain the similar results as Fig. 4(a) for the dynamics, but a sharp spectrum with a frequency around the cavity mode, and a linewidth of 2π×0.822𝜋0.822\pi\times 0.822 italic_π × 0.82 Hz. The observed shifted and broad peak agrees qualitatively with Fig. 4(a) of the experimental paper (JDBohnet2012, ).

We have further studied the intra-cavity photon number, the spectral shift and linewidth as function of the incoherent pumping rate [Fig. 4(c)]. As the pumping rate γ12subscript𝛾12\gamma_{12}italic_γ start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT increases, the photon number increases gradually, and approaches a maximum for γ12=NΓ/2subscript𝛾12𝑁Γ2\gamma_{12}=N\Gamma/2italic_γ start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT = italic_N roman_Γ / 2, and then drops dramatically to a value close to zero for γ12=NΓsubscript𝛾12𝑁Γ\gamma_{12}=N\Gammaitalic_γ start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT = italic_N roman_Γ. Here, Γ4g122/κΓ4superscriptsubscript𝑔122𝜅\Gamma\approx 4g_{12}^{2}/\kapparoman_Γ ≈ 4 italic_g start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_κ is the effective Purcell-enhanced decay rate. This result agrees qualitatively with Fig. 2 of the experimental article (JDBohnet2012, ). Accompanying with this, the spectral shift increases linearly within one kilohertz, while the linewidth increases linearly from 2π×32𝜋32\pi\times 32 italic_π × 3 kHz to 2π×62𝜋62\pi\times 62 italic_π × 6 kHz. In contrast, if the effective model is adopted, the photon number behaves similarly but the linewidth behaves in a opposite way, and achieves a minimal value of 2π×0.622𝜋0.622\pi\times 0.622 italic_π × 0.62 Hz (Fig. A4 of the Appendix). This comparison indicates again the difference of the full and effective model. To reveal the physics leading to the difference, we further study the influence of the driving strength of the dressed laser [Fig. 4(d)]. We find that the photon number, the frequency shift, and the linewidth all increase quadratically with the strength. To explain the cause of the frequency shift, we note that the atom will experience a frequency shift Δω=Ω2/4ΔΔ𝜔superscriptΩ24Δ\Delta\omega=\Omega^{2}/4\Deltaroman_Δ italic_ω = roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 4 roman_Δ, i.e. an AC Stark shift, if it is driven by a laser with a strength ΩΩ\Omegaroman_Ω and a frequency detuning ΔΔ\Deltaroman_Δ. The Ω2superscriptΩ2\Omega^{2}roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT-dependence of such a shift coincides with what shown in Fig. 4(d), and the 1/Δ1Δ1/\Delta1 / roman_Δ-dependence is further illustrated in Fig.  A6. Thus, it is highly possible that the frequency shift as observed here is caused by AC Stark shift.

In Fig. A7 of the Appendix, we have studied further the continuous superradiant Raman scattering for systems within the strong coupling regime and in the presence of incoherent atomic pumping. These results agree qualitatively with the those for the system within the crossover regime, except that the frequency shift is orders of magnitude smaller and the linewidth broadening is also much smaller.

V Conclusion

In summary, to understand the experiment (JDBohnet2012, ) of J. G. Bohnet et al. proving the concept of the superradiant laser (DMeiser2009, ), we have developed a quantum master equation to describe the superradiant Raman scattering of three-level atoms coupled with an optical cavity and a dressed laser, and obtained an effective quantum master equations for the system with two-level atoms by adiabatically eliminating the higher excited level. We have solved these equations with cumulant mean-field approach to study the dynamics of the systems with many atoms, and modified codes in the QuantumCumulant.jl package to calculate the Raman scattering spectrum.

Through the numerical simulations, we distinguish the normal and distorted superradiant Raman pulses for the systems within the crossover and strong coupling regime in the absence of incoherent pumping, which agree qualitatively with the experimental results and provide a unified view on these results. More importantly, our calculations demonstrate the shifted and broad spectrum for the continuous Raman scattering in the presence of the incoherent pumping, which agrees also qualitatively with the experimental results and provides also an alternative explanation on the observed broad spectrum rather the expected narrow spectrum. In any case, our study points out the rich physics involved in the superradiant Raman scattering, and further study could explore the application of this phenomenon in the magnetic field sensing (JMWeiner2012, ), the real-time track of quantum phase (AShankar2019, ), the Dicke phase transition of non-equilibrium dynamics (MPBaden2019, ), and so on.

Acknowledgements.
Huihui Yu carried out the numerical calculations under the supervision of Yuan Zhang who developed the theory and the numerical programs. They contribute equally to the work. All authors contributed to the analyses and the writing of the manuscript. This work is supported by the National Key R&D Program of China under Grant No. 2021YFA1400900, the National Natural Science Foundation of China under Grants No. 12004344, 12174347, 12074232, 12125406, 62027816, U21A2070, and the Cross-disciplinary Innovative Research Group Project of Henan Province No. 232300421004, as well as the Danish National Research Foundation through the Center of Excellence for Complex Quantum Systems (Grant agreement No. DNRF156).

References

  • (1) R. H. Dicke, Coherence in Spontaneous Radiation Processes, Phys. Rev. 93, 99 (1954).
  • (2) A. V. Andreev, I. Emel’yanovV and Y. A. II’inskir, Collective Spontaneous Emission (Dicke Superradiance), Sov. Phys. Usp. 23, 493 (1980).
  • (3) D. Meiser, J Ye, D. R. Carlson, and M. J. Holland Prospects for a Millihertz-Linewidth Laser. Phys. Rev. Lett. 102(16), 163601 (2009).
  • (4) D. A. Tieri, M. Xu, D. Meiser, J. Cooper, and M. J. Holland, Theory of the Crossover from Lasing to Steady State Superradiance, arXiv:1702.04830.
  • (5) K. Debnath, Y. Zhang, K. Mølmer, Lasing in the Superradiant Crossover Regime, Phys. Rev. A 98(6), 063837 (2018).
  • (6) Y. Zhang, C. Shan, K. Mølmer, Ultranarrow Superradiant Lasing by Dark Atom-Photon Dressed States, Phys. Rev. Lett. 126(12), 123602 (2021).
  • (7) J. G. Bohnet, Z. Chen, J. M. Weiner, D. Meiser, M. J. Holland, J. K. Thompson, A Steady-state Superradiant Laser with Less than One Intracavity Photon, Nature 484(7392), 78-81 (2012).
  • (8) M. A. Norcia, M. N. Winchester, J. R. K. Cline, J. K. Thompson Superradiance on the Millihertz Linewidth Strontium Clock Transition. Sci. Adv. 2(10), e1601231 (2016).
  • (9) M. A. Norcia, J. K. Thompson, Cold-strontium Laser in the Superradiant Crossover Regime. Phys. Rev. X. 6, 011025 (2016).
  • (10) J. M. Weiner, K. C. Cox, J. G. Bohnet, Z. Chen, J. K. Thompson, Superradiant Raman laser magnetometer. Appl. Phys. Lett. 101, 261107 (2012)
  • (11) A. Shankar,G. P. Greve, B. Wu, J. K. Thompson, M. J. Holland, Continuous Real-Time Tracking of a Quantum Phase Below the Standard Quantum Limit Phys. Rev. Lett. 122, 233602 (2019)
  • (12) M. P. Baden, K. J. Arnold, A. L. Grimsmo, S. Parkins, M. D. Barrett, Realization of the Dicke Model Using Cavity-Assisted Raman Transitions, Phys. Rev. Lett. 113, 020408 (2014)
  • (13) D. Plankensteiner, C. Hotter, H. Ritsch, QuantumCumulants.jl: A Julia Framework for Generalized Mean-field Equations in Open Quantum Systems Quantum. 6, 617 (2022).
  • (14) Shammah, N., Ahmed, S., Lambert, N., De Liberato, S., Nori, F. Open quantum systems with local and collective incoherent processes: Efficient numerical simulations using permutational invariance. Phys. Rev. A, 98(6), 063815 (2018).
  • (15) Zhang, Y., Zhang, Y.-X., Mølmer, K. Monte-Carlo simulations of superradiant lasing. New J. Phys., 20(11), 112001 (2018).

Appendix A Julia Codes

Refer to caption
Refer to caption
Figure A1: Julia codes to solve the quantum master equation (1) with the mean-field approach. The details of the codes are explained in the text.

In this Appendix, we explain the Julia codes to solve the quantum master equations. First, we focus on the system with three-level atoms, and present the codes to solve Eq. (1) (Fig. A1). The 1st and 2nd line import the necessary packages. The 3rd line defines the complex numbers and the time argument. The 4th line defines the Hilbert space for the cavity, the single three-level atom, and the atomic ensemble, as well the total system. The 5th line defines the photon annihilation operator, the transition and projection operators of the atoms. The 6th line defines the system Hamiltonian in the interaction picture. The 7th line defines the list of operators and rates, which are used later on to specify the Lindblad dissipative superoperators. The 8th line defines a list of three operators, and derive the equations for the mean values of these operators, as well as obtain the closed set of mean-field equations by applying second-order cumulant expansion approximation. The 9th line specifies the related parameters. The 10th line defines a list of the parameters, and the 11th line specifies a list of their values. Note that for the superradiant Raman scattering pulses, the incoherent pumping rate γ12subscript𝛾12\gamma_{12}italic_γ start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT is assumed as zero.

The 12th line specifies the initial values of the mean-fields. Here, we assume that the atoms are initially on the excited state. The 13th line defines the ordinary differential equation (ODE) system with the derived equations, the ODE problem with the initial values, the time range and the parameters, and then solve the ODE equations with Runge-Kutta method. The 14th line extracts the time list, the intra-cavity field amplitude, the intra-cavity photon number, the populations on the two hyper-fine ground levels and the excited level. The 15th line saves the data into a text file. The 16th to 19th line repeat the same procedure as the 9th to 15th line, except that the incoherent pumping rate γ12subscript𝛾12\gamma_{12}italic_γ start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT is not zero, and the continuous superradiant Raman scattering is obtained.

Refer to caption
Refer to caption
Refer to caption
Figure A2: Julia codes to calculate the superradiant Raman scattering spectrum for the system at the steady-state.

The QuantumCumulant.jl package has provided convenient functions to derive the equations for the two-time correlation function, and to calculate the system spectrum. However, this works only in the case with the time-independent Hamiltonian. In order to compute the superradiant Raman scattering, in the current study, we have reformulated the master equation in an interaction picture, where the Hamiltonian becomes time-dependent. As a result, we need to reformulate the aforementioned codes (Fig. A2).

In Fig. A2(a), the 1st line derives the equations for the two-time correlation functions o^1(τ)o^2(0)delimited-⟨⟩subscript^𝑜1𝜏subscript^𝑜20\left\langle\hat{o}_{1}(\tau)\hat{o}_{2}(0)\right\rangle⟨ over^ start_ARG italic_o end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_τ ) over^ start_ARG italic_o end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 0 ) ⟩. Unfortunately, the derived equations can not been directly passed to the functions in the ModelingToolkit.jl package to define the mathematical model and the ODE system. The 2nd line defines the abbreviation for this package, the time argument, and the empty list for the variables. The 3rd to 5th line analyze the derived equations, and collect the variables. The 6th line analyzes the list of variables, and removes the abundance. The 7th line evaluates the mean-values at the steady-state, which are used later on to calculate the initial conditions for the two-time correlation function. The 8th to 20th line replace o^2(0)delimited-⟨⟩subscript^𝑜20\left\langle\hat{o}_{2}(0)\right\rangle⟨ over^ start_ARG italic_o end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 0 ) ⟩ with the steady-state variables. The 21th line defines the list of variables, which includes also the steady-state variables. The 22th line copies all the derived equations. The 23th to 39th line construct the object to the two-time correlation functions. The 40th line defines the mathematical model with the derived equations, and the 41th defines the ODE system. The 42th line specifies the initial values of the correlation functions, the list of values for the parameters. The 43th line defines the ODE problem and solve it with Runge-Kutta method. The 44th line defines a list of time point. The 45th line imports the "correlation2spectrum" function from the QuantumOptics.jl package. The 46th line carries out the Fourier transform to the correlation functions, and returns the frequency and the spectrum. The 47th line saves the data into a text file.

Refer to caption
Figure A3: Julia codes to solve the effective quantum master equation (2).

Furthermore, we present the codes to solve the effective quantum master equation (2) (Fig. A3). The most codes are similar as those in Fig. A1 except for several small differences. In the 4th line, we define the atoms as two-level systems. In the 6th line, we defines effective Hamiltnoian in the interaction picture. In the 7th line, we account for the effective decay process in the Lindblad term. The 10th line, we calculate the effective coupling strength g12subscript𝑔12g_{12}italic_g start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT and the effective decay rate γ21subscript𝛾21\gamma_{21}italic_γ start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT. In the 15th line, we extract only the population of the two hyper-fine ground levels.

Appendix B Extra Numerical Results

In this Appendix, we provide extra results to complement those in the main text.

Refer to caption
Figure A4: Simulations with the effective model. Panel (a) shows the evolution of the intra-cavity photon number for the system with 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT atoms within the strong coupling regime. Panel (b) shows the steady-state intra-activity photon number (left axis) and the spectral linewidth (right axis) as function of the incoherent pumping rate for the system within the crossover regime. The two vertical dashed lines represent the effective decay rate γ21subscript𝛾21\gamma_{21}italic_γ start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT (left) and the collective enhanced decay rate NΓ𝑁ΓN\Gammaitalic_N roman_Γ (right), while the vertical line represents the Purcell-enhanced decay rate ΓΓ\Gammaroman_Γ.

B.1 Calculations with Effective Model

Figure A4 shows the simulations calculated with the effective model. For the system with 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT atoms initially prepared on the upper hyper-fine ground level in the absence of incoherent pumping, the simulation shows a superradiant pulse [Fig. A4(a)] just like the system within the weak coupling regime, which contradicts with the simulations calculated with the full model [Fig. (3)(a)] and with the experimental results [Fig. 1(c) of Ref. (JDBohnet2012, )].

For the system with 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT atoms in the presence of the incoherent pumping, the intra-cavity photon number increases firstly and then decreases, which agrees with what calculated with the full model [Fig. 4(c), left axis of Fig. A4(b)] and the experimental result [Fig. 2(a) of Ref.  (JDBohnet2012, )]. At the same time, the spectral linewidth behaves oppositely, and approaches the minimal value around hertz for moderate pumping [right axis of Fig. A4(b)], which contradicts with what obtained with the full model [Fig. 4(c)]. Although the small linewidth agrees with what expected, it seems not agree with the experimental result [Fig. 4(a) of Ref.  (JDBohnet2012, )].

Refer to caption
Figure A5: Evolution of the maximum, delay and decay time of the pulses as function of the dressed strength.

B.2 Extra Results for Superradiant Raman Scattering Pulses within the Strong Coupling Regime

Figure 3 demonstrates how the maximum photon number, delay and decay time of the superradiant Raman pulses change with number of atoms and frequency detuning within the strong coupling regime. In Fig. A5, we complement these by showing the results as function of the strength ΩΩ\Omegaroman_Ω of the driving laser. We find that the maximal photon number scales as Ω2superscriptΩ2\leavevmode\nobreak\ \Omega^{2}roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT while the pulse decay and decay time scale as Ω2superscriptΩ2\leavevmode\nobreak\ \Omega^{-2}roman_Ω start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT. These results agree with those in the crossover regime (Fig. 2c).

Refer to caption
Figure A6: Frequency shift (relative to the cavity mode) in the steady-state superradiant Raman spectrum for different frequency detunning of the dressed laser to the atomic transition between the upper hyper-fine ground state and the excited state.

B.3 Frequency Shift in Scattering Spectrum for Different Frequency detuning

In Fig. 4 of the main text, we have examined the steady-state superradiant Raman scattering for the systems within the crossover regime. In Fig. A6, we complement these results with the frequency shift ΔωΔ𝜔\Delta\omegaroman_Δ italic_ω (relative to the cavity mode) as function of the frequency detuning ΔΔ\Deltaroman_Δ of the dressed laser to the atomic transition (between the upper hyper-fine ground state and the excited state). We see that the frequency shift is negative for the negative frequency detuning (Δω<0Δ𝜔0\Delta\omega<0roman_Δ italic_ω < 0 for Δ<0Δ0\Delta<0roman_Δ < 0), changes the sign for the positive frequency detuning (Δω>0Δ𝜔0\Delta\omega>0roman_Δ italic_ω > 0 for Δ>0Δ0\Delta>0roman_Δ > 0), and tend to diverge when the frequency detuning approaches zero (ΔωΔ𝜔\Delta\omega\to\inftyroman_Δ italic_ω → ∞ for Δ0Δ0\Delta\to 0roman_Δ → 0). These results seem to hint the relation Δω1/Δproportional-toΔ𝜔1Δ\Delta\omega\propto 1/\Deltaroman_Δ italic_ω ∝ 1 / roman_Δ, and suggest that the observation might be attributed to the AC Stark shift effect.

B.4 Continuous Superradiant Raman Scattering for Systems within Strong Coupling Regime

Figure A7 shows the continuous superradiant Raman scattering for systems with 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT atoms within the strong coupling regime and in the presence of incohernt pumping, which are calculated with the full model. These results agree qualitatively with the those for the system within the crossover regime (Fig. 4), except that the frequency shift is orders of magnitude less and the linewidth broadening is significantly larger. Furthermore, the intra-cavity photon number does not show any oscillations before reaching the steady-state value.

Refer to caption
Figure A7: Continuous superradiant Raman scattering for systems with 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT atoms within the strong coupling regime. The results are similar as those in Fig. 4 for the system with 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT atoms within the crossover regime, except that the steady-state photon number is much larger, the spectral shift is smaller, and the spectral linewidth is much larger. Furthermore, the intra-cavity photon number does not show the oscillations.