Collective transition quenching in the presence of multiple competing decay channels

Wai-Keong Mok darielmok@caltech.edu Institute for Quantum Information and Matter, California Institute of Technology, Pasadena, CA 91125, USA    Stuart J. Masson Department of Physics, Columbia University, New York, New York 10027, USA    Dan M. Stamper-Kurn Department of Physics, University of California, Berkeley, California 94720 Challenge Institute for Quantum Computation, University of California, Berkeley, California 94720 Materials Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720    Tanya Zelevinsky Department of Physics, Columbia University, New York, New York 10027, USA    Ana Asenjo-Garcia ana.asenjo@columbia.edu Department of Physics, Columbia University, New York, New York 10027, USA
Abstract

We present a theoretical framework for ‘collective transition quenching’, a quantum many-body dissipative phenomenon that occurs in systems with multiple collective decay channels. Despite the competition, interactions suppress all but the dominant decay transition, leading to a ‘winner takes all’ dynamic where the system primarily settles into the dominant ground state. We prove that, in the presence of permutation symmetry, this problem is exactly solvable for any number of competing channels. Additionally, we develop an approximate model for the dynamics by mapping the evolution into a continuity equation for a fluid, and show analytically that the dominant transition ratio converges to unity with increasing system size as a power-law, for any branching ratio. This near-deterministic preparation of the dominant ground state has broad applicability. As an example we discuss a protocol for molecular photoassociation where collective dynamics effectively acts as a catalyst, amplifying the yield in a particular final state. Our results open new avenues for many-body strategies in the preparation and control of quantum systems.

The suppression of decay pathways into undesired quantum states is crucial for the control and manipulation of open quantum systems. Simplified theoretical models often rely on ‘closed’ transitions where the excited state decays predominantly to a specific ground state. Yet, in practice, real-world emitters seldom adhere to the idealized paradigm of two-level systems. For instance, highly excited atomic states (such as Rydberg states) have many lower-energy states accessible by spontaneous emission [1]. In photochemistry, decay to a single ground state is often inefficient due to numerous competing pathways arising from electronic, vibrational, and rotational degrees of freedom [2, 3]. Solid-state emitters (such as color centers or dye molecules) also suffer from parasitic decay from phonon sidebands [4, 5]. Achieving closed transitions in experiments is challenging, and often involves isolating two-level systems from more complex internal structures or employing repumping techniques to redirect population back into the desired states. However, these approaches are inherently limited by the natural, single-particle branching ratios of the transitions.

The natural branching ratios and dynamics of decay can, in fact, be modified by engineering the quantum system and its environment. One common approach is to tailor the dielectric environment by placing emitters within optical cavities [6], waveguides [7, 8], or other photonic structures [9, 10] such that the desired decay channel is Purcell-enhanced. Numerical studies with multilevel atoms [11, 12, 13, 14] and molecules [15] have suggested collective emission as an alternative to circumvent limitations from single-particle branching ratios. These proposals rely on many-body, transient superradiance [16, 17], a phenomenon characterized by avalanche-like behavior [18, 19, 20] where decay into a given ground state enhances the probability of subsequent emission into that same state. This process effectively steers the emitters towards a specific ground state. A comprehensive analytical treatment of this physics remains lacking, which is critical for fully understanding and exploiting its potential.

We term the phenomenon where correlated decay suppresses all but the most dominant emission path as collective transition quenching. As shown in Fig. 1(a), we consider an ensemble of emitters coupled to a reservoir. The emitters have multiple decay channels, each leading to different final states. Each decay channel can be collectively enhanced by many-body correlations that emerge dynamically, leading to competition between them. We find approximate steady-state solutions for the populations of the different ground states by modeling the quantum dynamics through a continuity equation for a fluid. We prove that, despite the competition, the population density of the dominant ground state exhibits a power-law convergence to unity for any branching ratio, with the power-law exponent characterized by the ratio between dominant and subdominant decay rates. This is supported by a rigorous analysis of the exact steady-state solution. We apply our framework to the problem of photochemistry, where sub-optimal branching ratios limit the effectiveness of molecule creation, direct laser cooling [21], and optical imaging. Specifically, in molecular photoassociation of strontium dimers, we demonstrate that collective transition quenching greatly enhances sample purity.

Refer to caption
(a)
Figure 1: Increased population of the dominant transition (with suppressed fluctuations) due to collective decay in multilevel systems. (a) N𝑁Nitalic_N multilevel emitters with d𝑑ditalic_d ground states and a single excited state decay collectively to an environment (at rates Γ1,,ΓdsubscriptΓ1subscriptΓ𝑑\Gamma_{1},\ldots,\Gamma_{d}roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , roman_Γ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT), such as a “bad” cavity, a waveguide in the “mirror configuration” (where the relative distance between the emitters is a half-integer multiple of the resonance wavelength λ0subscript𝜆0\lambda_{0}italic_λ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) or free space (for the latter, a dense ensemble of subwavelength volume is required to preserve permutational symmetry). (b) Dissipative dynamics can be modeled as a random walk between permutationally symmetric ground states (d=2𝑑2d=2italic_d = 2 depicted), labelled by |n1,,ndketsubscript𝑛1subscript𝑛𝑑\ket{n_{1},\ldots,n_{d}}| start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG ⟩, where nμsubscript𝑛𝜇n_{\mu}italic_n start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT denotes the population of the ground state |gμketsubscript𝑔𝜇\ket{g_{\mu}}| start_ARG italic_g start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG ⟩. (c) Marginal probability distribution P(n1)=n2P(n1,n2)𝑃subscript𝑛1subscriptsubscript𝑛2𝑃subscript𝑛1subscript𝑛2P(n_{1})=\sum_{n_{2}}P(n_{1},n_{2})italic_P ( italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) = ∑ start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_P ( italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) for the population density of the ground state |g1ketsubscript𝑔1\ket{g_{1}}| start_ARG italic_g start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ⟩, for N=150𝑁150N=150italic_N = 150 and d=2𝑑2d=2italic_d = 2 in the steady state, after complete depletion of the fully-inverted initial state. The distributions for collective decay with r2=1subscript𝑟21r_{2}=1italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1 (gray), and collective decay with r2=0.5subscript𝑟20.5r_{2}=0.5italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.5 (blue) are obtained via numerical simulations of Eq. (2). The binomial distribution is plotted for independent decay with r2=0.5subscript𝑟20.5r_{2}=0.5italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.5 (green).

We consider an (undriven) ensemble of N𝑁Nitalic_N identical emitters with a level structure consisting of a single excited state |eisubscriptket𝑒𝑖\ket{e}_{i}| start_ARG italic_e end_ARG ⟩ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and d𝑑ditalic_d ground states labelled |gμisubscriptketsubscript𝑔𝜇𝑖\ket{g_{\mu}}_{i}| start_ARG italic_g start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG ⟩ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, with μ{1,,d}𝜇1𝑑\mu\in\{1,\ldots,d\}italic_μ ∈ { 1 , … , italic_d } and i{1,,N}𝑖1𝑁i\in\{1,\ldots,N\}italic_i ∈ { 1 , … , italic_N }. The emitters are symmetrically coupled to a Markovian environment, with separate couplings for each transition |e|gμket𝑒ketsubscript𝑔𝜇\ket{e}\to\ket{g_{\mu}}| start_ARG italic_e end_ARG ⟩ → | start_ARG italic_g start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG ⟩. Physically, this can be realized by coupling the emitters to near-resonant cavity modes (in the “bad cavity” or “weak coupling” limit), to a single-mode waveguide in the “mirror configuration” (where the emitters are separated by a half-integer multiple of the wavelength [22, 23]), or via free-space interactions in a dense ensemble of subwavelength volume, as depicted in Fig. 1(a). We assume that photons from different transitions can be resolved either by polarization or frequency (for the waveguide configuration, we assume that the transition frequencies are approximately equal to fulfil the mirror condition). The system dynamics is modelled by the master equation

ρ˙=i[μ=1dχμA^μA^μ,ρ]+μ=1dΓμ𝒟[A^μ]ρ,˙𝜌𝑖superscriptsubscript𝜇1𝑑subscript𝜒𝜇superscriptsubscript^𝐴𝜇subscript^𝐴𝜇𝜌superscriptsubscript𝜇1𝑑subscriptΓ𝜇𝒟delimited-[]subscript^𝐴𝜇𝜌\dot{\rho}=-i\left[\sum_{\mu=1}^{d}\chi_{\mu}\hat{A}_{\mu}^{{\dagger}}\hat{A}_% {\mu},\rho\right]+\sum_{\mu=1}^{d}\Gamma_{\mu}\mathcal{D}[\hat{A}_{\mu}]\rho,over˙ start_ARG italic_ρ end_ARG = - italic_i [ ∑ start_POSTSUBSCRIPT italic_μ = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_χ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT over^ start_ARG italic_A end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_A end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT , italic_ρ ] + ∑ start_POSTSUBSCRIPT italic_μ = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT roman_Γ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT caligraphic_D [ over^ start_ARG italic_A end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ] italic_ρ , (1)

where χμsubscript𝜒𝜇\chi_{\mu}italic_χ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT are the coherent interaction rates, A^μ=i=1N(|gμe|)isubscript^𝐴𝜇superscriptsubscript𝑖1𝑁subscriptketsubscript𝑔𝜇bra𝑒𝑖\hat{A}_{\mu}=\sum_{i=1}^{N}(\ket{g_{\mu}}\bra{e})_{i}over^ start_ARG italic_A end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( | start_ARG italic_g start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG ⟩ ⟨ start_ARG italic_e end_ARG | ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the collective lowering operator on the transition |e|gμket𝑒ketsubscript𝑔𝜇\ket{e}\rightarrow\ket{g_{\mu}}| start_ARG italic_e end_ARG ⟩ → | start_ARG italic_g start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG ⟩, and 𝒟[A^]ρA^ρA^{A^A^,ρ}/2𝒟delimited-[]^𝐴𝜌^𝐴𝜌superscript^𝐴superscript^𝐴^𝐴𝜌2\mathcal{D}[\hat{A}]\rho\equiv\hat{A}\rho\hat{A}^{\dagger}-\{\hat{A}^{\dagger}% \hat{A},\rho\}/2caligraphic_D [ over^ start_ARG italic_A end_ARG ] italic_ρ ≡ over^ start_ARG italic_A end_ARG italic_ρ over^ start_ARG italic_A end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT - { over^ start_ARG italic_A end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_A end_ARG , italic_ρ } / 2. Collective dissipation occurs with rates ΓμsubscriptΓ𝜇\Gamma_{\mu}roman_Γ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT (identical to the single-emitter ones), and we assume the ordering Γ1Γ2ΓdsubscriptΓ1subscriptΓ2subscriptΓ𝑑\Gamma_{1}\geq\Gamma_{2}\geq\ldots\geq\Gamma_{d}roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≥ roman_Γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≥ … ≥ roman_Γ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT, such that Γ1subscriptΓ1\Gamma_{1}roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is the dominant decay rate. For convenience, we denote the ratio between dominant and subdominant decay rates as rμΓμ/Γ1subscript𝑟𝜇subscriptΓ𝜇subscriptΓ1r_{\mu}\equiv\Gamma_{\mu}/\Gamma_{1}italic_r start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ≡ roman_Γ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT / roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT.

Refer to caption
(a)
Figure 2: (a) Velocity vector field of the continuum model for r2=0.5subscript𝑟20.5r_{2}=0.5italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.5 and N=10𝑁10N=10italic_N = 10. The dashed line represents the flow trajectory in the single-particle approximation starting from the fully excited state (x1,x2)=(0,0)subscript𝑥1subscript𝑥200(x_{1},x_{2})=(0,0)( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = ( 0 , 0 ), where x1(2)subscript𝑥12x_{1(2)}italic_x start_POSTSUBSCRIPT 1 ( 2 ) end_POSTSUBSCRIPT denotes the population of |g1(2)ketsubscript𝑔12\ket{g_{1(2)}}| start_ARG italic_g start_POSTSUBSCRIPT 1 ( 2 ) end_POSTSUBSCRIPT end_ARG ⟩. Darker color indicates lower velocity. The red solid line shows the neutrally stable ground state attractor. The unphysical region (where the ground state population is larger than N𝑁Nitalic_N) is shown in gray. (b) Population density outside the dominant ground state [1Dominant transition ratio (DTR)1Dominant transition ratio (DTR)1-\text{Dominant transition ratio (DTR)}1 - Dominant transition ratio (DTR)] against number of emitters N𝑁Nitalic_N, for various decay ratios r2=Γ2/Γ1subscript𝑟2subscriptΓ2subscriptΓ1r_{2}=\Gamma_{2}/\Gamma_{1}italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = roman_Γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. The points are obtained from numerically solving the rate equation (2), while the dotted lines denote the approximate formula (9). Solid lines denote the exact asymptotic solution (10). (c) Cumulative distribution (i.e., total probability of the dominant ground state population being at most n1subscript𝑛1n_{1}italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT) Pc(n1)=n1=0n1P(n1)subscript𝑃𝑐subscript𝑛1superscriptsubscriptsuperscriptsubscript𝑛10subscript𝑛1𝑃superscriptsubscript𝑛1P_{c}(n_{1})=\sum_{n_{1}^{\prime}=0}^{n_{1}}P(n_{1}^{\prime})italic_P start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) = ∑ start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_P ( italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ). Data obtained from numerical simulation of the rate equation (2), with r2=0.4subscript𝑟20.4r_{2}=0.4italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.4. Darker colors indicate larger number of emitters, with 61N10461𝑁superscript10461\leq N\leq 10^{4}61 ≤ italic_N ≤ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT. All plots are made for d=2𝑑2d=2italic_d = 2 ground states.

Dynamical evolution can be understood as a random walk in the subspace of permutationally symmetric states, as shown in Fig. 1(b). Since Eq. (1) preserves permutation symmetry, basis states |n1,,ndketsubscript𝑛1subscript𝑛𝑑\ket{n_{1},\ldots,n_{d}}| start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG ⟩ are fully described by occupation numbers, where nμsubscript𝑛𝜇n_{\mu}italic_n start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT is the population of |gμketsubscript𝑔𝜇\ket{g_{\mu}}| start_ARG italic_g start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG ⟩. These are typically entangled, as they consist of a symmetric superposition of excitations over N𝑁Nitalic_N particles. The population of the excited state is ne=Nμnμsubscript𝑛𝑒𝑁subscript𝜇subscript𝑛𝜇n_{e}=N-\sum_{\mu}n_{\mu}italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = italic_N - ∑ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT, with ne=0subscript𝑛𝑒0n_{e}=0italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = 0 in the final, steady state. Employing the ansatz ρ={nμ}Pn1,,nd|n1,,ndn1,,nd|𝜌subscriptsubscript𝑛𝜇subscript𝑃subscript𝑛1subscript𝑛𝑑ketsubscript𝑛1subscript𝑛𝑑brasubscript𝑛1subscript𝑛𝑑\rho=\sum_{\{n_{\mu}\}}P_{n_{1},\ldots,n_{d}}\ket{n_{1},\ldots,n_{d}}\bra{n_{1% },\ldots,n_{d}}italic_ρ = ∑ start_POSTSUBSCRIPT { italic_n start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT } end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUBSCRIPT | start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG ⟩ ⟨ start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG |, Eq. (1) reduces to a rate equation [see Supplementary Material (SM) [24]]

P˙n1,,nd=(Nν=1dnν)μ=1dΓμ(nμ+1)Pn1,,nμ,,nd+(Nν=1dnν+1)μ=1dΓμnμPn1,,nμ1,,ndsubscript˙𝑃subscript𝑛1subscript𝑛𝑑𝑁superscriptsubscript𝜈1𝑑subscript𝑛𝜈superscriptsubscript𝜇1𝑑subscriptΓ𝜇subscript𝑛𝜇1subscript𝑃subscript𝑛1subscript𝑛𝜇subscript𝑛𝑑𝑁superscriptsubscript𝜈1𝑑subscript𝑛𝜈1superscriptsubscript𝜇1𝑑subscriptΓ𝜇subscript𝑛𝜇subscript𝑃subscript𝑛1subscript𝑛𝜇1subscript𝑛𝑑\begin{split}\dot{P}_{n_{1},\ldots,n_{d}}&=-\left(N-\sum_{\nu=1}^{d}n_{\nu}% \right)\sum_{\mu=1}^{d}\Gamma_{\mu}(n_{\mu}+1)P_{n_{1},\ldots,n_{\mu},\ldots,n% _{d}}\\ &+\left(N-\sum_{\nu=1}^{d}n_{\nu}+1\right)\sum_{\mu=1}^{d}\Gamma_{\mu}n_{\mu}P% _{n_{1},\ldots,n_{\mu}-1,\ldots,n_{d}}\end{split}start_ROW start_CELL over˙ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL start_CELL = - ( italic_N - ∑ start_POSTSUBSCRIPT italic_ν = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ) ∑ start_POSTSUBSCRIPT italic_μ = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT roman_Γ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( italic_n start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT + 1 ) italic_P start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + ( italic_N - ∑ start_POSTSUBSCRIPT italic_ν = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT + 1 ) ∑ start_POSTSUBSCRIPT italic_μ = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT roman_Γ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT - 1 , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_CELL end_ROW (2)

for the probabilities Pn1,,ndsubscript𝑃subscript𝑛1subscript𝑛𝑑P_{n_{1},\ldots,n_{d}}italic_P start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUBSCRIPT of occupying the state |n1,,ndketsubscript𝑛1subscript𝑛𝑑\ket{n_{1},\ldots,n_{d}}| start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG ⟩. This ansatz is justified even if there are coherences in the initial state (i.e., off-diagonal terms in the density matrix), since they are decoupled and hence do not affect population dynamics. While the rate equation holds for any permutationally symmetric initial state, below we choose the fully inverted state (i.e., |eNsuperscriptket𝑒tensor-productabsent𝑁\ket{e}^{\otimes N}| start_ARG italic_e end_ARG ⟩ start_POSTSUPERSCRIPT ⊗ italic_N end_POSTSUPERSCRIPT).

Although Eq. (2) can be efficiently simulated, it is non-trivial to obtain the steady state analytically. Moreover, the steady state is highly non-unique, since any combination of emitters in any ground state is a possible steady state. The trivial case of just one collective decay channel (d=1)𝑑1(d=1)( italic_d = 1 ) reduces to the problem of Dicke superradiance [16, 17]. We instead focus on the problem of multiple competing transitions.

We quantify collective transition quenching by the dominant transition ratio (DTR), defined as the mean population density of the dominant ground state in the steady state, i.e.,

DTR=n1¯N,DTR¯subscript𝑛1𝑁\text{DTR}=\frac{\overline{n_{1}}}{N},DTR = divide start_ARG over¯ start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG end_ARG start_ARG italic_N end_ARG , (3)

where n1¯¯subscript𝑛1\overline{n_{1}}over¯ start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG is the mean number of emitters in |g1ketsubscript𝑔1\ket{g_{1}}| start_ARG italic_g start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ⟩ as t𝑡t\to\inftyitalic_t → ∞. For independent emitters, the marginal probability distribution P(n1)=n2,,ndPn1,,nd𝑃subscript𝑛1subscriptsubscript𝑛2subscript𝑛𝑑subscript𝑃subscript𝑛1subscript𝑛𝑑P(n_{1})=\sum_{n_{2},\ldots,n_{d}}P_{n_{1},\ldots,n_{d}}italic_P ( italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) = ∑ start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUBSCRIPT for the dominant ground state is a binomial distribution whose average (normalized by N𝑁Nitalic_N) is the DTR =(1+μ>1rμ)1absentsuperscript1subscript𝜇1subscript𝑟𝜇1=(1+\sum_{\mu>1}r_{\mu})^{-1}= ( 1 + ∑ start_POSTSUBSCRIPT italic_μ > 1 end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, determined solely by the decay ratios and independent of N𝑁Nitalic_N [see Fig. 1(c)]. Below, we prove that the DTR always converges to 1 as N𝑁N\to\inftyitalic_N → ∞ for any number of collective decay channels, assuming Γ1>Γμ>1subscriptΓ1subscriptΓ𝜇1\Gamma_{1}>\Gamma_{\mu>1}roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > roman_Γ start_POSTSUBSCRIPT italic_μ > 1 end_POSTSUBSCRIPT. This effect can be attributed to the superradiant enhancement, which amplifies the dominant transition relative to all subdominant transitions. If Γ1=Γ2==ΓdsubscriptΓ1subscriptΓ2subscriptΓ𝑑\Gamma_{1}=\Gamma_{2}=\ldots=\Gamma_{d}roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_Γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = … = roman_Γ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT, one obtains a uniform distribution for Pn1,,ndsubscript𝑃subscript𝑛1subscript𝑛𝑑P_{n_{1},\ldots,n_{d}}italic_P start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUBSCRIPT [13], as shown in Fig. 1(c).

To gain analytical insights from the rate equation, we transform it into a continuity equation of a fluid that flows in Euclidean space. In the limit of large N𝑁Nitalic_N, we make the continuum approximation [25] by setting nμxμsubscript𝑛𝜇subscript𝑥𝜇n_{\mu}\to x_{\mu}italic_n start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT → italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT, x=(x1xd)T+d𝑥superscriptsubscript𝑥1subscript𝑥𝑑𝑇superscriptsubscript𝑑\vec{x}=(x_{1}\ldots x_{d})^{T}\in\mathbb{R}_{+}^{d}over→ start_ARG italic_x end_ARG = ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT … italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ∈ blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, and f(n1,,nμ,,nd)f(n1,,nμ1,,nd)xμf(x)𝑓subscript𝑛1subscript𝑛𝜇subscript𝑛𝑑𝑓subscript𝑛1subscript𝑛𝜇1subscript𝑛𝑑subscript𝑥𝜇𝑓𝑥f(n_{1},\ldots,n_{\mu},\ldots,n_{d})-f(n_{1},\ldots,n_{\mu}-1,\ldots,n_{d})\to% \frac{\partial}{\partial x_{\mu}}f(\vec{x})italic_f ( italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) - italic_f ( italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_n start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT - 1 , … , italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) → divide start_ARG ∂ end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG italic_f ( over→ start_ARG italic_x end_ARG ) for an arbitrary differentiable function f𝑓fitalic_f. The rate equation is then approximated by the continuity equation

tP(x,t)=(v(x)P(x,t)),𝑡𝑃𝑥𝑡𝑣𝑥𝑃𝑥𝑡\frac{\partial}{\partial t}P(\vec{x},t)=-\mathbf{\nabla}\cdot(\vec{v}(\vec{x})% P(\vec{x},t)),divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG italic_P ( over→ start_ARG italic_x end_ARG , italic_t ) = - ∇ ⋅ ( over→ start_ARG italic_v end_ARG ( over→ start_ARG italic_x end_ARG ) italic_P ( over→ start_ARG italic_x end_ARG , italic_t ) ) , (4)

which describes a fluid flow in +dsuperscriptsubscript𝑑\mathbb{R}_{+}^{d}blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT, governed by the position-dependent velocity field v(x)𝑣𝑥\vec{v}(\vec{x})over→ start_ARG italic_v end_ARG ( over→ start_ARG italic_x end_ARG ) with μ𝜇\muitalic_μ-th component

vμ(x)=Γμ(Nν=1dxν)(xμ+1).subscript𝑣𝜇𝑥subscriptΓ𝜇𝑁superscriptsubscript𝜈1𝑑subscript𝑥𝜈subscript𝑥𝜇1v_{\mu}(\vec{x})=\Gamma_{\mu}\left(N-\sum_{\nu=1}^{d}x_{\nu}\right)(x_{\mu}+1).italic_v start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( over→ start_ARG italic_x end_ARG ) = roman_Γ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ( italic_N - ∑ start_POSTSUBSCRIPT italic_ν = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ) ( italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT + 1 ) . (5)

The flow comes to a stop as νxνNsubscript𝜈subscript𝑥𝜈𝑁\sum_{\nu}x_{\nu}\to N∑ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT → italic_N, which physically corresponds to the system approaching its steady state, where no excited state population remains. The velocity field is illustrated for d=2𝑑2d=2italic_d = 2 in Fig. 2(a). The fully excited initial state corresponds to the origin of +dsuperscriptsubscript𝑑\mathbb{R}_{+}^{d}blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT. Due to the non-uniqueness of the steady state, it is not sufficient to simply solve for tP=0subscript𝑡𝑃0\partial_{t}P=0∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_P = 0.

Solving Eq. (4) analytically for arbitrary times is a formidable task. Instead, we employ a single-particle approximation where the fluid is idealized as a point particle initialized at the origin, with velocity dynamics dx/dt=v(x)d𝑥𝑑𝑡𝑣𝑥\text{d}\vec{x}/dt=\vec{v}(\vec{x})d over→ start_ARG italic_x end_ARG / italic_d italic_t = over→ start_ARG italic_v end_ARG ( over→ start_ARG italic_x end_ARG ). This reduces the partial differential equation in Eq. (4) to a coupled system of d𝑑ditalic_d ordinary differential equations. Dividing the components of v(x)𝑣𝑥\vec{v}(\vec{x})over→ start_ARG italic_v end_ARG ( over→ start_ARG italic_x end_ARG ), we readily find

dxμdxν=ΓμΓνxμ+1xν+1dsubscript𝑥𝜇dsubscript𝑥𝜈subscriptΓ𝜇subscriptΓ𝜈subscript𝑥𝜇1subscript𝑥𝜈1\frac{\text{d}x_{\mu}}{\text{d}x_{\nu}}=\frac{\Gamma_{\mu}}{\Gamma_{\nu}}\frac% {x_{\mu}+1}{x_{\nu}+1}divide start_ARG d italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG start_ARG d italic_x start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG = divide start_ARG roman_Γ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG start_ARG roman_Γ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG divide start_ARG italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT + 1 end_ARG start_ARG italic_x start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT + 1 end_ARG (6)

for any pair of μ,ν{1,,d}𝜇𝜈1𝑑\mu,\nu\in\{1,\ldots,d\}italic_μ , italic_ν ∈ { 1 , … , italic_d }. Integrating the above expression yields

(xμ+1)Γν=(xν+1)Γμ.superscriptsubscript𝑥𝜇1subscriptΓ𝜈superscriptsubscript𝑥𝜈1subscriptΓ𝜇(x_{\mu}+1)^{\Gamma_{\nu}}=(x_{\nu}+1)^{\Gamma_{\mu}}.( italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT + 1 ) start_POSTSUPERSCRIPT roman_Γ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_POSTSUPERSCRIPT = ( italic_x start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT + 1 ) start_POSTSUPERSCRIPT roman_Γ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT . (7)

The d1𝑑1d-1italic_d - 1 independent equations of the form (7) define the particle trajectory in +dsuperscriptsubscript𝑑\mathbb{R}_{+}^{d}blackboard_R start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT starting from the origin. Generalizing Eq. (7) to other permutationally symmetric initial states is straightforward (see SM [24]).

The steady state solution has a geometrical interpretation as the intersection between the particle trajectory and the hyperplane μ=1dxμ=Nsuperscriptsubscript𝜇1𝑑subscript𝑥𝜇𝑁\sum_{\mu=1}^{d}x_{\mu}=N∑ start_POSTSUBSCRIPT italic_μ = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT = italic_N. To find an asymptotic solution, we assume that xμ1much-greater-thansubscript𝑥𝜇1x_{\mu}\gg 1italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ≫ 1. The trajectory is then described by a simpler set of equations xμΓν=xνΓμsuperscriptsubscript𝑥𝜇subscriptΓ𝜈superscriptsubscript𝑥𝜈subscriptΓ𝜇x_{\mu}^{\Gamma_{\nu}}=x_{\nu}^{\Gamma_{\mu}}italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Γ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_POSTSUPERSCRIPT = italic_x start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Γ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT. The constraint μ=1dxμ=Nsuperscriptsubscript𝜇1𝑑subscript𝑥𝜇𝑁\sum_{\mu=1}^{d}x_{\mu}=N∑ start_POSTSUBSCRIPT italic_μ = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT = italic_N can then be rewritten as μ=1dxdΓμ/Γd=Nsuperscriptsubscript𝜇1𝑑superscriptsubscript𝑥𝑑subscriptΓ𝜇subscriptΓ𝑑𝑁\sum_{\mu=1}^{d}x_{d}^{\Gamma_{\mu}/\Gamma_{d}}=N∑ start_POSTSUBSCRIPT italic_μ = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Γ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT / roman_Γ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUPERSCRIPT = italic_N. For simplicity, we consider the non-degenerate case where all ΓμsubscriptΓ𝜇\Gamma_{\mu}roman_Γ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT are distinct (see SM [24] for the degenerate case). Using the method of dominant balance [26] we obtain xdNrdsubscript𝑥𝑑superscript𝑁subscript𝑟𝑑x_{d}\approx N^{r_{d}}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ≈ italic_N start_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, valid in the regime 1NrdNmuch-less-than1superscript𝑁subscript𝑟𝑑much-less-than𝑁1\ll N^{r_{d}}\ll N1 ≪ italic_N start_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ≪ italic_N. In the regime 1NrμNμ>1much-less-than1superscript𝑁subscript𝑟𝜇much-less-than𝑁for-all𝜇11\ll N^{r_{\mu}}\ll N\;\forall\;\mu>11 ≪ italic_N start_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ≪ italic_N ∀ italic_μ > 1, an iterative method yields the self-consistent solution

x1Nμ=2dNrμ,xμ=Nrμ,μ>1.formulae-sequencesubscript𝑥1𝑁superscriptsubscript𝜇2𝑑superscript𝑁subscript𝑟𝜇formulae-sequencesubscript𝑥𝜇superscript𝑁subscript𝑟𝜇𝜇1x_{1}\approx N-\sum_{\mu=2}^{d}{N^{r_{\mu}}},\quad x_{\mu}=N^{r_{\mu}},\quad% \mu>1.italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ italic_N - ∑ start_POSTSUBSCRIPT italic_μ = 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_N start_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT = italic_N start_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , italic_μ > 1 . (8)

The dominant transition ratio thus reads

DTRx1N1μ=2d1N1rμ,DTRsubscript𝑥1𝑁1superscriptsubscript𝜇2𝑑1superscript𝑁1subscript𝑟𝜇\text{DTR}\approx\frac{x_{1}}{N}\approx 1-\sum_{\mu=2}^{d}\frac{1}{N^{1-r_{\mu% }}},DTR ≈ divide start_ARG italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_N end_ARG ≈ 1 - ∑ start_POSTSUBSCRIPT italic_μ = 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUPERSCRIPT 1 - italic_r start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG , (9)

which converges to unity as N𝑁N\to\inftyitalic_N → ∞, with the slowest convergence characterized by the power law Nr21similar-toabsentsuperscript𝑁subscript𝑟21\sim N^{r_{2}-1}∼ italic_N start_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - 1 end_POSTSUPERSCRIPT. Equation (9) provides our first theoretical prediction for collective transition quenching. Comparing with numerical simulations of the rate equation (2) for d=2𝑑2d=2italic_d = 2 in Fig. 2(b), the approximate formula (9) agrees qualitatively, with higher accuracy attained for smaller r2subscript𝑟2r_{2}italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT.

Remarkably, we can solve for the DTR exactly in the asymptotic N𝑁N\rightarrow\inftyitalic_N → ∞ limit. In the SM [24], we prove that the dynamics governed by the rate equation (2) is integrable, and derive the complete set of N+1𝑁1N+1italic_N + 1 independent conserved quantities. We thus overcome the problem of non-unique steady states. However, because of the complexity of the exact solution, we only use it to compute the DTR, which reads

DTR=n1¯N=1μ=2dΓ~(1rμ)N1rμ,DTR¯subscript𝑛1𝑁1superscriptsubscript𝜇2𝑑~Γ1subscript𝑟𝜇superscript𝑁1subscript𝑟𝜇\text{DTR}=\frac{\overline{n_{1}}}{N}=1-\sum_{\mu=2}^{d}\frac{\tilde{\Gamma}(1% -r_{\mu})}{N^{1-r_{\mu}}},DTR = divide start_ARG over¯ start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG end_ARG start_ARG italic_N end_ARG = 1 - ∑ start_POSTSUBSCRIPT italic_μ = 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT divide start_ARG over~ start_ARG roman_Γ end_ARG ( 1 - italic_r start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ) end_ARG start_ARG italic_N start_POSTSUPERSCRIPT 1 - italic_r start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG , (10)

where Γ~()~Γ\tilde{\Gamma}(\cdot)over~ start_ARG roman_Γ end_ARG ( ⋅ ) is the gamma function. The exact solution yields the same scaling as the approximate formula (9), and agrees excellently with numerical simulations, as shown in Fig. 2(b). This agreement explains why the approximation of Eq. (9) becomes more accurate for smaller rμsubscript𝑟𝜇r_{\mu}italic_r start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT, since Γ~(1rμ)=1+O(rμ)~Γ1subscript𝑟𝜇1𝑂subscript𝑟𝜇\tilde{\Gamma}(1-r_{\mu})=1+O(r_{\mu})over~ start_ARG roman_Γ end_ARG ( 1 - italic_r start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ) = 1 + italic_O ( italic_r start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ). The error in the prefactor of Eq. (9) likely arises from the single-particle approximation.

Our formalism can also be easily extended to include non-collective decay channels, modeled as a leakage at a rate ΓleaksubscriptΓleak\Gamma_{\text{leak}}roman_Γ start_POSTSUBSCRIPT leak end_POSTSUBSCRIPT [24]. Going back to the fluid model, this adds an extra component v0=Γleak(Nμ=1dxμx0)subscript𝑣0subscriptΓleak𝑁superscriptsubscript𝜇1𝑑subscript𝑥𝜇subscript𝑥0v_{0}=\Gamma_{\text{leak}}\left(N-\sum_{\mu=1}^{d}x_{\mu}-x_{0}\right)italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_Γ start_POSTSUBSCRIPT leak end_POSTSUBSCRIPT ( italic_N - ∑ start_POSTSUBSCRIPT italic_μ = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) to the velocity field and modifies ν=1dxνν=0dxνsuperscriptsubscript𝜈1𝑑subscript𝑥𝜈superscriptsubscript𝜈0𝑑subscript𝑥𝜈\sum_{\nu=1}^{d}x_{\nu}\to\sum_{\nu=0}^{d}x_{\nu}∑ start_POSTSUBSCRIPT italic_ν = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT → ∑ start_POSTSUBSCRIPT italic_ν = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT in Eq. (5). A similar analysis for large N𝑁Nitalic_N yields x0(Γleak/Γ1)lnNsubscript𝑥0subscriptΓleaksubscriptΓ1𝑁x_{0}\approx(\Gamma_{\text{leak}}/\Gamma_{1})\ln Nitalic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ ( roman_Γ start_POSTSUBSCRIPT leak end_POSTSUBSCRIPT / roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) roman_ln italic_N, in the same regime as the validity of Eq. (9). This justifies the omission of non-collective decay in our model, since x0xμNrμmuch-less-thansubscript𝑥0subscript𝑥𝜇similar-tosuperscript𝑁subscript𝑟𝜇x_{0}\ll x_{\mu}\sim N^{r_{\mu}}italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≪ italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ∼ italic_N start_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, and the approximate DTR reads

DTR1μ=2d1N1rμΓleakΓ1lnNN.DTR1superscriptsubscript𝜇2𝑑1superscript𝑁1subscript𝑟𝜇subscriptΓleaksubscriptΓ1𝑁𝑁\text{DTR}\approx 1-\sum_{\mu=2}^{d}\frac{1}{N^{1-r_{\mu}}}-\frac{\Gamma_{% \text{leak}}}{\Gamma_{1}}\frac{\ln N}{N}.DTR ≈ 1 - ∑ start_POSTSUBSCRIPT italic_μ = 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUPERSCRIPT 1 - italic_r start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG - divide start_ARG roman_Γ start_POSTSUBSCRIPT leak end_POSTSUBSCRIPT end_ARG start_ARG roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG divide start_ARG roman_ln italic_N end_ARG start_ARG italic_N end_ARG . (11)

In the non-competing scenario with only one collective decay channel (d=1)𝑑1(d=1)( italic_d = 1 ), the DTR always converges to unity as lnN/Nsimilar-toabsent𝑁𝑁\sim\ln N/N∼ roman_ln italic_N / italic_N (even if |e|g1ket𝑒ketsubscript𝑔1\ket{e}\to\ket{g_{1}}| start_ARG italic_e end_ARG ⟩ → | start_ARG italic_g start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ⟩ is not the most dominant transition). This recovers the well-established result of Dicke superradiance in the presence of local decay [15, 27].

The avalanche-like behavior of the dominant transition not only impacts the population of the dominant ground state, but also lowers the fluctuations of the probability distribution dramatically. For d=2𝑑2d=2italic_d = 2, we numerically observe (for small r2subscript𝑟2r_{2}italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) that the relative fluctuation δn1/n1¯𝛿subscript𝑛1¯subscript𝑛1\delta n_{1}/\overline{n_{1}}italic_δ italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / over¯ start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG in the dominant ground state population vanishes as N𝑁N\to\inftyitalic_N → ∞ faster than N1/2similar-toabsentsuperscript𝑁12\sim N^{-1/2}∼ italic_N start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT (expected from independent decay). This causes the steady state distribution to become sharply peaked at density n1/N=1subscript𝑛1𝑁1n_{1}/N=1italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_N = 1, as indicated by the cumulative probability distribution in Fig. 2(c) and in the SM [24]. By the Bhatia-Davis inequality [28], the relative fluctuation can be bounded as δn1/n1¯(1DTR)/DTR𝛿subscript𝑛1¯subscript𝑛11DTRDTR\delta n_{1}/\overline{n_{1}}\leq\sqrt{(1-\text{DTR})/\text{DTR}}italic_δ italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / over¯ start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ≤ square-root start_ARG ( 1 - DTR ) / DTR end_ARG which vanishes like N(r21)/2similar-toabsentsuperscript𝑁subscript𝑟212\sim N^{(r_{2}-1)/2}∼ italic_N start_POSTSUPERSCRIPT ( italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - 1 ) / 2 end_POSTSUPERSCRIPT. While this bound is not tight, it justifies the single-particle approximation made in our theoretical analysis.

Finally, the convergence time T𝑇Titalic_T to the steady state can be estimated within the fluid model under the single-particle approximation. By noting that dt=dx1/v1d𝑡dsubscript𝑥1subscript𝑣1\text{d}t=\text{d}x_{1}/v_{1}d italic_t = d italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, we find (see SM [24])

T=1Γ1𝒞dx1(Nx1μ>1xμ)(x1+1)2r2Γ1lnNN,𝑇1subscriptΓ1subscript𝒞dsubscript𝑥1𝑁subscript𝑥1subscript𝜇1subscript𝑥𝜇subscript𝑥112subscript𝑟2subscriptΓ1𝑁𝑁T=\frac{1}{\Gamma_{1}}\int_{\mathcal{C}}\frac{\text{d}x_{1}}{(N-x_{1}-\sum_{% \mu>1}x_{\mu})(x_{1}+1)}\approx\frac{2-r_{2}}{\Gamma_{1}}\frac{\ln N}{N},italic_T = divide start_ARG 1 end_ARG start_ARG roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ∫ start_POSTSUBSCRIPT caligraphic_C end_POSTSUBSCRIPT divide start_ARG d italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ( italic_N - italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - ∑ start_POSTSUBSCRIPT italic_μ > 1 end_POSTSUBSCRIPT italic_x start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ) ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 1 ) end_ARG ≈ divide start_ARG 2 - italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG divide start_ARG roman_ln italic_N end_ARG start_ARG italic_N end_ARG , (12)

where 𝒞𝒞\mathcal{C}caligraphic_C is the trajectory defined by Eq. (7). The presence of competing collective decay channels affects the well-known superradiant timescale of Γ1TlnN/Nsimilar-tosubscriptΓ1𝑇𝑁𝑁\Gamma_{1}T\sim\ln N/Nroman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_T ∼ roman_ln italic_N / italic_N [17] only by a constant factor.

Refer to caption
(a)
Figure 3: Targeted photoassociation of a molecular dimer. Weakly bound molecules are created by optical excitation and spontaneously decay to various rovibrational ground-electronic states. The dominant transition ratio (DTR) is plotted against the initial number of weakly bound Sr2subscriptSr2\text{Sr}_{2}Sr start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT molecules in the (1)1u(ν=1,J=1)1subscript1𝑢formulae-sequencesuperscript𝜈1superscript𝐽1(1)1_{u}(\nu^{\prime}=-1,J^{\prime}=1)( 1 ) 1 start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ( italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = - 1 , italic_J start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 1 ) state. Decay is considered to four states, (ν=1,J=0,2)formulae-sequence𝜈1𝐽02(\nu=-1,J=0,2)( italic_ν = - 1 , italic_J = 0 , 2 ) and (ν=2,J=0,2)formulae-sequence𝜈2𝐽02(\nu=-2,J=0,2)( italic_ν = - 2 , italic_J = 0 , 2 ) with branching ratios 0.54, 0.27, 0.13, and 0.06, respectively. Points show the numerical solution from the Monte-Carlo simulation of the rate equation (2) for the non-competing scenario where only the dominant transition is collectively enhanced (\square) and the competing scenario with all four collective channels (\circ). The dashed line denotes the analytical prediction (11) for the non-competing scenario with Γleak/Γ1=0.852subscriptΓleaksubscriptΓ10.852\Gamma_{\text{leak}}/\Gamma_{1}=0.852roman_Γ start_POSTSUBSCRIPT leak end_POSTSUBSCRIPT / roman_Γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.852. The solid line denotes the prediction (10) for the competing scenario. Predictions are valid for N1much-greater-than𝑁1N\gg 1italic_N ≫ 1.

As a possible application of collective transition quenching, we analyze the creation of ultracold diatomic molecules via photoassociation. In one-color photoassociation, laser-cooled atoms form weakly bound molecules that spontaneously decay into various more tightly-bound states [29]. The vibrational and rotational branching ratios are dictated by Franck-Condon factors and angular-momentum Hönl-London factors, respectively. Specifically, we examine ultracold strontium dimers that lack Feshbach resonances and must be produced optically [30, 31]. Strontium dimers have narrow optical transitions and a structureless ground state, making them well-suited for metrology [32, 33, 34, 35]. We consider photoassociation via the state (1)1u(ν=1,J=1)1subscript1𝑢formulae-sequencesuperscript𝜈1superscript𝐽1(1)1_{u}(\nu^{\prime}=-1,J^{\prime}=1)( 1 ) 1 start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ( italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = - 1 , italic_J start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 1 ) [33], where νsuperscript𝜈\nu^{\prime}italic_ν start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT and Jsuperscript𝐽J^{\prime}italic_J start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT are the excited state vibrational and rotational quantum numbers, respectively, and negative vibrational numbers count down from the dissociation threshold. This state has a dominant decay path to the ν=1,J=0formulae-sequence𝜈1𝐽0\nu=-1,J=0italic_ν = - 1 , italic_J = 0 ground state, with an overall branching ratio of 54%similar-toabsentpercent54\sim 54\%∼ 54 % (molecular parameters are detailed in the SM [24]). Our protocol uses a short, strong pulse to create N𝑁Nitalic_N weakly bound molecules from a dense sample of ultracold atoms. Alternatively, effective decay directly from unbound atoms into ground-state molecules can be engineered via a continuous off-resonant drive [36].

The avalanche-like behavior dramatically enhances the fraction of molecules in the dominant ground state, as shown in Fig. 3. If only one transition is collective, the dominant transition ratio rapidly approaches unity, as shown in Ref. [15]. This can be realized by engineering the dielectric environment to be frequency-selective (either via a cavity [37] or a photonic crystal [38]). We note that collective transition quenching occurs in the weak coupling limit of cavity QED, distinguishing it from the so-called “polaritonic chemistry” in strong coupling [39, 40]. Figure 3 demonstrates that a nearly-pure sample of molecules is created even under maximal competition, where all decay channels are collective. Collective enhancement of all transitions can occur in broadband cavities (or for molecules with hyperfine structure, which yields smaller frequency differences between levels) or in single-mode fibers [41] within the “mirror configuration” [23]. This effect can also occur in free space with dense, subwavelength molecular clouds. However, in this scenario, collisions can lead to significant losses. For polar molecules, collisions can be prevented via electric field [42] or microwave shielding [43, 44, 45, 46]. Alternatively, they can be suppressed by trapping molecules in optical tweezers to form ordered arrays [47, 48]. As these systems are extended, their dynamics are not constrained to the permutationally symmetric subspace. Nevertheless, we expect collective transition quenching to occur for molecules placed close to waveguides (at arbitrary distances) or arranged in ordered two- and three-dimensional arrays of subwavelength lattice constant [49, 50], albeit with reduced scaling.

In summary, collective decay holds promise for a wide range of quantum systems, from closing open transitions in atoms [12, 14] and preventing parasitic decay in solid-state emitters [51, 52, 5] to directing emission into dielectric nanostructures instead of unwanted modes [53, 20]. While we have primarily focused on an initially inverted ensemble, our treatment captures collective transition quenching from a large range of multiply excited states. Future research directions include studying the potential for many-body-enhanced metrology [54], due to the sensitivity of the ground state populations to the decay ratios. Other interesting avenues include the observation of symmetry breaking, which can occur if multiple dominant transitions have identical decay rates, akin to mirror symmetry breaking predicted in waveguide QED [20].

Acknowledgements.— We thank Kon Leung, Luis A. Orozco, and Maximilian Schemmer for helpful discussions. We acknowledge support by the National Science Foundation through the CAREER Award (No. 2047380) and the QLCI program (grant No. OMA-2016245), the Air Force Office of Scientific Research through their Young Investigator Prize (grant No. 21RT0751) and through grant No. FA9550-1910328, ARO through the MURI program (Grant No. W911NF-20-1-0136), DARPA (Grant No. W911NF2010090), as well as by the David and Lucile Packard Foundation and the Brown Science Foundation. The Institute for Quantum Information and Matter is an NSF Physics Frontiers Center.

References

  • Cong et al. [2022] I. Cong, H. Levine, A. Keesling, D. Bluvstein, S.-T. Wang, and M. D. Lukin, Hardware-efficient, fault-tolerant quantum computation with Rydberg atoms, Phys. Rev. X 12, 021049 (2022).
  • Carr et al. [2009] L. D. Carr, D. DeMille, R. V. Krems, and J. Ye, Cold and ultracold molecules: science, technology and applications, New J. Phys. 11, 055049 (2009).
  • Bohn et al. [2017] J. L. Bohn, A. M. Rey, and J. Ye, Cold molecules: Progress in quantum engineering of chemistry and quantum matter, Science 357, 1002 (2017).
  • Atatüre et al. [2018] M. Atatüre, D. Englund, N. Vamivakas, S.-Y. Lee, and J. Wrachtrup, Material platforms for spin-based photonic quantum technologies, Nat. Rev. Mater. 3, 38 (2018).
  • Lange et al. [2024] C. M. Lange, E. Daggett, V. Walther, L. Huang, and J. D. Hood, Superradiant and subradiant states in lifetime-limited organic molecules through laser-induced tuning, Nat. Phys. 20, 836 (2024).
  • Walther et al. [2006] H. Walther, B. T. H. Varcoe, B.-G. Englert, and T. Becker, Cavity quantum electrodynamics, Rep. Prog. Phys. 69, 1325 (2006).
  • Vetsch et al. [2010] E. Vetsch, D. Reitz, G. Sagué, R. Schmidt, S. T. Dawkins, and A. Rauschenbeutel, Optical interface created by laser-cooled atoms trapped in the evanescent field surrounding an optical nanofiber, Phys. Rev. Lett. 104, 203603 (2010).
  • Solano et al. [2017] P. Solano, P. Barberis-Blostein, F. K. Fatemi, L. A. Orozco, and S. L. Rolston, Super-radiance reveals infinite-range dipole interactions through a nanofiber, Nat. Commun. 8, 1857 (2017).
  • Goban et al. [2015] A. Goban, C.-L. Hung, J. D. Hood, S.-P. Yu, J. A. Muniz, O. Painter, and H. J. Kimble, Superradiance for atoms trapped along a photonic crystal waveguide, Phys. Rev. Lett. 115, 063601 (2015).
  • Zhou et al. [2023] X. Zhou, H. Tamura, T.-H. Chang, and C.-L. Hung, Trapped atoms and superradiance on an integrated nanophotonic microring circuit (2023), arXiv:2312.14318 .
  • Molander and Jr [1982] W. A. Molander and C. R. S. Jr, Altered branching ratios and relaxation oscillations in superfluorescent cascades, J. Phys. B At. Mol. Opt. Phys. 15, 2109 (1982).
  • Sutherland and Robicheaux [2017] R. T. Sutherland and F. Robicheaux, Superradiance in inverted multilevel atomic clouds, Phys. Rev. A 95, 033839 (2017).
  • Piñeiro Orioli et al. [2022] A. Piñeiro Orioli, J. K. Thompson, and A. M. Rey, Emergent dark states from superradiant dynamics in multilevel atoms in a cavity, Phys. Rev. X 12, 011054 (2022).
  • Masson et al. [2024] S. J. Masson, J. P. Covey, S. Will, and A. Asenjo-Garcia, Dicke superradiance in ordered arrays of multilevel atoms, PRX Quantum 5, 010344 (2024).
  • Wellnitz et al. [2020] D. Wellnitz, S. Schütz, S. Whitlock, J. Schachenmayer, and G. Pupillo, Collective dissipative molecule formation in a cavity, Phys. Rev. Lett. 125, 193201 (2020).
  • Dicke [1954] R. H. Dicke, Coherence in spontaneous radiation processes, Phys. Rev. 93, 99 (1954).
  • Gross and Haroche [1982] M. Gross and S. Haroche, Superradiance: An essay on the theory of collective spontaneous emission, Phys. Rep. 93, 301 (1982).
  • Clemens et al. [2004] J. P. Clemens, L. Horvath, B. C. Sanders, and H. J. Carmichael, Shot-to-shot fluctuations in the directed superradiant emission from extended atomic samples, J. Opt. B: Quantum Semiclass. Opt. 6, S736 (2004).
  • Masson and Asenjo-Garcia [2022] S. J. Masson and A. Asenjo-Garcia, Universality of Dicke superradiance in arrays of quantum emitters, Nat. Commun. 13, 2285 (2022).
  • Cardenas-Lopez et al. [2023] S. Cardenas-Lopez, S. J. Masson, Z. Zager, and A. Asenjo-Garcia, Many-body superradiance and dynamical mirror symmetry breaking in waveguide QED, Phys. Rev. Lett. 131, 033605 (2023).
  • Fitch and Tarbutt [2021] N. J. Fitch and M. R. Tarbutt, Laser-cooled molecules, Adv. At. Mol. Opt. Phys. 70, 157 (2021).
  • Chang et al. [2012] D. E. Chang, L. Jiang, A. V. Gorshkov, and H. J. Kimble, Cavity qed with atomic mirrors, New J. Phys. 14, 063003 (2012).
  • Sørensen et al. [2016] H. L. Sørensen, J.-B. Béguin, K. W. Kluge, I. Iakoupov, A. S. Sørensen, J. H. Müller, E. S. Polzik, and J. Appel, Coherent backscattering of light off one-dimensional atomic strings, Phys. Rev. Lett. 117, 133604 (2016).
  • [24] See Supplementary Materials for additional calculation details, which includes Refs [55, 56, 57, 58, 59].
  • Degiorgio and Ghielmetti [1971] V. Degiorgio and F. Ghielmetti, Approximate solution to the superradiance master equation, Phys. Rev. A 4, 2415 (1971).
  • Bender and Orszag [1999] C. M. Bender and S. A. Orszag, Perturbation series, in Advanced Mathematical Methods for Scientists and Engineers I: Asymptotic Methods and Perturbation Theory (Springer New York, New York, NY, 1999) pp. 319–367.
  • Malz et al. [2022] D. Malz, R. Trivedi, and J. I. Cirac, Large-N𝑁{N}italic_N limit of Dicke superradiance, Phys. Rev. A 106, 013716 (2022).
  • Bhatia and Davis [2000] R. Bhatia and C. Davis, A better bound on the variance, Am. Math. Mon. 107, 353 (2000).
  • Jones et al. [2006] K. M. Jones, E. Tiesinga, P. D. Lett, and P. S. Julienne, Ultracold photoassociation spectroscopy: Long-range molecules and atomic scattering, Rev. Mod. Phys. 78, 483 (2006).
  • Stellmer et al. [2012] S. Stellmer, B. Pasquiou, R. Grimm, and F. Schreck, Creation of ultracold Sr2subscriptSr2\mathrm{Sr}_{2}roman_Sr start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT molecules in the electronic ground state, Phys. Rev. Lett. 109, 115302 (2012).
  • Reinaudi et al. [2012] G. Reinaudi, C. B. Osborn, M. McDonald, S. Kotochigova, and T. Zelevinsky, Optical production of stable ultracold Sr288superscriptsubscriptSr288{}^{88}\mathrm{Sr}_{2}start_FLOATSUPERSCRIPT 88 end_FLOATSUPERSCRIPT roman_Sr start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT molecules, Phys. Rev. Lett. 109, 115303 (2012).
  • Zelevinsky et al. [2008] T. Zelevinsky, S. Kotochigova, and J. Ye, Precision test of mass-ratio variations with lattice-confined ultracold molecules, Phys. Rev. Lett. 100, 043201 (2008).
  • Leung et al. [2023] K. H. Leung, B. Iritani, E. Tiberi, I. Majewska, M. Borkowski, R. Moszynski, and T. Zelevinsky, Terahertz vibrational molecular clock with systematic uncertainty at the 1014superscript1014{10}^{-14}10 start_POSTSUPERSCRIPT - 14 end_POSTSUPERSCRIPT level, Phys. Rev. X 13, 011047 (2023).
  • Kondov et al. [2019] S. S. Kondov, C.-H. Lee, K. H. Leung, C. Liedl, I. Majewska, R. Moszynski, and T. Zelevinsky, Molecular lattice clock with long vibrational coherence, Nat. Phys. 15, 1118 (2019).
  • Tiberi et al. [2024] E. Tiberi, M. Borkowski, B. Iritani, R. Moszynski, and T. Zelevinsky, Searching for new fundamental interactions via isotopic shifts in molecular lattice clocks (2024), arXiv:2403.07097 .
  • Inouye et al. [1999] S. Inouye, A. P. Chikkatur, D. M. Stamper-Kurn, J. Stenger, D. E. Pritchard, and W. Ketterle, Superradiant rayleigh scattering from a bose-einstein condensate, Science 285, 571 (1999).
  • Kampschulte and Denschlag [2018] T. Kampschulte and J. H. Denschlag, Cavity-controlled formation of ultracold molecules, New J. Phys. 20, 123015 (2018).
  • Pérez-Ríos et al. [2017] J. Pérez-Ríos, M. E. Kim, and C.-L. Hung, Ultracold molecule assembly with photonic crystals, New J. Phys. 19, 123035 (2017).
  • Feist et al. [2018] J. Feist, J. Galego, and F. J. Garcia-Vidal, Polaritonic chemistry with organic molecules, ACS PhotonicsACS Photonics 5, 205 (2018).
  • Hertzog et al. [2019] M. Hertzog, M. Wang, J. Mony, and K. Börjesson, Strong light–matter interactions: a new direction within chemistry, Chem. Soc. Rev. 48, 937 (2019).
  • Márquez-Mijares et al. [2023] M. Márquez-Mijares, B. Lepetit, and E. Brion, Nanofibre-based trap for rb2 molecule, Physica Scripta 98, 115404 (2023).
  • Valtolina et al. [2020] G. Valtolina, K. Matsuda, W. G. Tobias, J.-R. Li, L. De Marco, and J. Ye, Dipolar evaporation of reactive molecules to below the Fermi temperature, Nature 588, 239 (2020).
  • Anderegg et al. [2021] L. Anderegg, S. Burchesky, Y. Bao, S. S. Yu, T. Karman, E. Chae, K.-K. Ni, W. Ketterle, and J. M. Doyle, Observation of microwave shielding of ultracold molecules, Science 373, 779 (2021).
  • Schindewolf et al. [2022] A. Schindewolf, R. Bause, X.-Y. Chen, M. Duda, T. Karman, I. Bloch, and X.-Y. Luo, Evaporation of microwave-shielded polar molecules to quantum degeneracy, Nature 607, 677 (2022).
  • Lin et al. [2023] J. Lin, G. Chen, M. Jin, Z. Shi, F. Deng, W. Zhang, G. Quéméner, T. Shi, S. Yi, and D. Wang, Microwave shielding of bosonic narb molecules, Phys. Rev. X 13, 031032 (2023).
  • Bigagli et al. [2024] N. Bigagli, W. Yuan, S. Zhang, B. Bulatovic, T. Karman, I. Stevenson, and S. Will, Observation of bose–einstein condensation of dipolar molecules, Nature  (2024).
  • Anderegg et al. [2019] L. Anderegg, L. W. Cheuk, Y. Bao, S. Burchesky, W. Ketterle, K.-K. Ni, and J. M. Doyle, An optical tweezer array of ultracold molecules, Science 365, 1156 (2019).
  • Holland et al. [2023] C. M. Holland, Y. Lu, and L. W. Cheuk, On-demand entanglement of molecules in a reconfigurable optical tweezer array, Science 382, 1143 (2023).
  • Sierra et al. [2022] E. Sierra, S. J. Masson, and A. Asenjo-Garcia, Dicke superradiance in ordered lattices: Dimensionality matters, Phys. Rev. Res. 4, 023207 (2022).
  • Mok et al. [2024] W.-K. Mok, A. Poddar, E. Sierra, C. C. Rusconi, J. Preskill, and A. Asenjo-Garcia, Universal scaling laws for correlated decay of many-body quantum systems (2024), arXiv:2406.00722 .
  • Sipahigil et al. [2016] A. Sipahigil, R. E. Evans, D. D. Sukachev, M. J. Burek, J. Borregaard, M. K. Bhaskar, C. T. Nguyen, J. L. Pacheco, H. A. Atikian, C. Meuwly, R. M. Camacho, F. Jelezko, E. Bielejec, H. Park, M. Lončar, and M. D. Lukin, An integrated diamond nanophotonics platform for quantum-optical networks, Science 354, 847 (2016).
  • Scheibner et al. [2007] M. Scheibner, T. Schmidt, L. Worschech, A. Forchel, G. Bacher, T. Passow, and D. Hommel, Superradiance of quantum dots, Nat. Phys. 3, 106 (2007).
  • Liedl et al. [2024] C. Liedl, F. Tebbenjohanns, C. Bach, S. Pucher, A. Rauschenbeutel, and P. Schneeweiss, Observation of superradiant bursts in a cascaded quantum system, Phys. Rev. X 14, 011020 (2024).
  • Ding et al. [2022] D.-S. Ding, Z.-K. Liu, B.-S. Shi, G.-C. Guo, K. Mølmer, and C. S. Adams, Enhanced metrology at the critical point of a many-body Rydberg atomic system, Nat. Phys. 18, 1447 (2022).
  • Shammah et al. [2018] N. Shammah, S. Ahmed, N. Lambert, S. De Liberato, and F. Nori, Open quantum systems with local and collective incoherent processes: Efficient numerical simulations using permutational invariance, Phys. Rev. A 98, 063815 (2018).
  • [56] K. H. Leung, PhD Thesis, 2023, Columbia University. The strontium molecular lattice clock: Vibrational spectroscopy with hertz-level accuracy, .
  • [57] M. Borkowski, private communication.
  • [58] I. Majewska, PhD Thesis, 2021, University of Warsaw. Theoretical description of ultracold strontium molecules in an optical lattice: Control of photodissociation and interpretation of molecular clock experiments, .
  • Zelevinsky et al. [2006] T. Zelevinsky, M. M. Boyd, A. D. Ludlow, T. Ido, J. Ye, R. Ciuryło, P. Naidon, and P. S. Julienne, Narrow line photoassociation in an optical lattice, Phys. Rev. Lett. 96, 203201 (2006).