Startseite Naturwissenschaften Theory of dynamical superradiance in organic materials
Artikel Open Access

Theory of dynamical superradiance in organic materials

  • Lukas Freter ORCID logo , Piper Fowler-Wright ORCID logo , Javier Cuerda ORCID logo , Brendon W. Lovett ORCID logo , Jonathan Keeling ORCID logo EMAIL logo und Päivi Törmä ORCID logo EMAIL logo
Veröffentlicht/Copyright: 8. Dezember 2025

Abstract

We develop the theory of dynamical superradiance – the collective exchange of energy between an ensemble of initially excited emitters and a single-mode cavity – for organic materials where electronic states are coupled to vibrational modes. We consider two models to capture the vibrational effects: first, vibrations treated as a Markovian bath for two-level emitters, via a pure dephasing term in the Lindblad master equation for the system; second, vibrational modes directly included in the system via the Holstein–Tavis–Cummings Hamiltonian. By exploiting the permutation symmetry of the emitters and weak U(1) symmetry, we develop a numerical method capable of exactly solving the Tavis–Cummings model with local dissipation for up to 140 emitters. Using the exact method, we validate mean-field and second-order cumulant approximations and use them to describe macroscopic numbers of emitters. We analyze the dynamics of the average cavity photon number, electronic coherence, and Bloch vector length and show that the effect of vibrational mode coupling goes beyond simple dephasing. Our results show that superradiance is possible in the presence of vibrational mode coupling; for negative cavity detunings, the vibrational coupling may even enhance superradiance. We identify asymmetry of the photon number rise time as a function of the detuning of the cavity frequency as an experimentally accessible signature of such vibrationally assisted superradiance.

1 Introduction

The term superradiance (SR), first introduced by Dicke in 1954 [1], refers to the collective spontaneous emission of N closely spaced quantum emitters in free space. Due to the interaction of the emitters with a common light field, a macroscopic dipole moment builds up during the emission process, resulting in a peak emission rate which scales as N 2, compared to ∼ N for independent emitters. For an introduction to SR, see Refs. [2], [3]; we also highlight recent interest in the phenomenon of SR in extended systems of different geometries [4], [5], [6], [7], [8], [9], [10], [11], [12], [13], [14], [15], [16], [17], [18]. In the special case where the emission process starts from an incoherent, fully inverted ensemble of emitters, the process has also been called superfluorescence (SF). We adopt the convention [3] that SR refers to any case where the initial state has some nonzero coherence, and SF to the specific case where the initial state has no coherence [Figure 1(b)].

Figure 1: 
Schematic of the dynamical SR system. (a) Emitters (yellow circles) are placed inside a single-mode cavity, which is subject to photon losses at a rate κ. The internal structure of the emitters follows either (i) the TC model or (ii) the HTC model. For the TC model, the emitters are ideal two-level systems with energy splitting ω
0 and are subject to nonradiative decay at rate γ and pure dephasing at rate γ

ϕ
. For the HTC model, the emitters are dressed by vibrational levels with spacing ω

ν
, with additional vibrational thermalization (rates γ
↑ and γ
↓). (b) Illustration of SF and SR initial conditions according to the polar angle θ of the state of each two-level system.
Figure 1:

Schematic of the dynamical SR system. (a) Emitters (yellow circles) are placed inside a single-mode cavity, which is subject to photon losses at a rate κ. The internal structure of the emitters follows either (i) the TC model or (ii) the HTC model. For the TC model, the emitters are ideal two-level systems with energy splitting ω 0 and are subject to nonradiative decay at rate γ and pure dephasing at rate γ ϕ . For the HTC model, the emitters are dressed by vibrational levels with spacing ω ν , with additional vibrational thermalization (rates γ and γ ). (b) Illustration of SF and SR initial conditions according to the polar angle θ of the state of each two-level system.

While the original SR and SF, as discussed above, occur for emitters in free space, there are several closely related phenomena that occur for emitters in a cavity. The first, “dynamical SR,” describes the dynamics of initially excited emitters coupling collectively to a single cavity mode [19], [20], [21], [22], [23], [24]. Dynamical SR is the subject of this paper. The second is the superradiant phase transition of the Dicke model [25], [26], [27], [28], [29], which describes a steady-state phenomenon of the many-emitter–cavity system. The third is the idea of the “superradiant laser” [30], [31], [32], which can be considered either as corresponding to the original Dicke SR phenomena but continuously refreshed to provide a steady state, or considered as using a bad cavity to synchronize emitters in an analog of Huygen’s clocks.

If the emitters coupled to the cavity mode are treated as ideal two-level systems, and one makes the rotating wave approximation, then the combined light–matter system can be described with the Tavis–Cummings (TC) model [33]. In this case, the behavior of dynamical SR is well-studied. A mean-field solution of this model predicts a train of hyperbolic secant pulses for the number of photons in the cavity starting from a fully inverted initial state [20], and finite-size quantum corrections to this solution have been considered in Refs. [34], [35]. The same dynamics have also been studied in the context of cold gases [36], [37], [38], [39], semiconductor microcavities [24], [40], impurity spins coupled to a microwave resonator [41], and free-electron lasers [42]. There further exist studies of complementary initial states, where all emitters are in the ground state and the photon mode is in a Fock state [43] or a coherent state [44]. Studies of the dynamics in low-excitation subspaces are also given in Refs. [45], [46].

Approximating physical emitters as ideal two-level systems is, however, not typically realistic; furthermore, additional degrees of freedom may lead to richer light–matter coupling phenomena. For example, emitters in solid-state environments are subject to coupling to phonon modes and other sources of dephasing. With organic molecules in particular, there are nontrivial effects of vibrational coupling to the electronic degrees of freedom [47], [48], [49], unless one explicitly works to suppress vibrational excitations [50].

Organic emitters have been used in experiments demonstrating strong light–matter coupling, polariton lasing, and Bose–Einstein condensation [51], [52], [53], [54], [55]. Hence, they are a promising platform for harnessing collective effects in light–matter coupling. Motivated by the necessity to model realistic emitters for state-of-the-art experiments, we extend in this work the treatment of dynamical SR from two-level systems [19], [20], [21], [22], [23], [24] to organic molecules, by explicitly incorporating vibrational modes in the system description.

As a first approximation, the effect of a vibrational environment on the electronic degrees of freedom can be treated as a pure dephasing term in a Markovian master equation, see Figure 1(a)i. This is a phenomenological model aiming to capture the decoherence induced by a vibrational bath that weakly couples to the electronic states, but crucially cannot account for realistic electron–phonon interactions. To go beyond the pure dephasing model, we include one vibrational mode per molecule explicitly in the system Hamiltonian, giving the Holstein–Tavis–Cummings (HTC) model [47], [56], [57], [58], [59], [60], Figure 1(a)ii. The main advancement of this work is to study dynamical SR in the HTC model. We will see that, while in both the phenomenological model and the HTC model vibrational coupling suppresses the extent and duration of the dynamical SR, there are significant differences in behavior, particularly when the vibronic coupling becomes strong.

The leading challenge of modeling dynamical SR systems is the exponential growth of the size of the Hilbert space with the number of molecules. For the open TC model with local incoherent processes, methods leveraging the weak permutation symmetry of the emitters [61], [62], [63], [64], [65], [66], [67] allow numerical solution up to N ≈ 30 emitters. Here, we extend a powerful numerical method [68], which additionally exploits the weak U(1) symmetry of the Lindblad master equation, enabling a solution of the dissipative TC model with local losses and dephasing up to N ∼ 140 emitters when starting from a fully inverted initial state.

Experiments, however, have significantly larger numbers of emitters, and so other methods are required to reach large N. For this, we use a standard mean-field approach (assuming factorization of the molecular and cavity degrees of freedom), along with a second-order cumulant approach that captures leading-order corrections to the mean-field theory. Such mean-field and cumulant approaches can generally be expected to match the exact solution in the thermodynamic limit N [69], [70], [71]; however, this is not always the case [14], [72], [73], [74], [75]. By comparing the exact solution to mean-field and second-order cumulant solutions, we show that the early-time behavior of dynamical SR in the TC model can be captured with both of these approximate methods if the process starts in an SR initial state (θ ≠ 0 in Figure 1(b)), not in the SF (θ = 0) one. Since the second-order cumulant solution converges to the mean-field solution in the large N limit, this motivates using mean-field theory to study dynamical SR with vibrationally dressed molecules in the HTC model.

Cumulant approaches have previously been used to study polariton lasing in the HTC model [76], [77]. However, we note that these works use a reduced set of cumulant equations where terms that vanish due to U(1) symmetry are omitted. While appropriate for studying dynamics from homogeneous initial conditions, such an approach cannot be used to describe SR initial conditions where there is an initial breaking of this symmetry.

We point out another related work [78] where mean-field theory was combined with the process-tensor formalism to solve the HTC model with a continuum of vibrational modes, treated as a non-Markovian bath. In the present work, we model the situation where there is one (or a few) vibrational modes strongly coupled to the electronic state and show how higher-order correlations are included via cumulants.

Our key findings are that strong vibronic coupling destroys dynamical SR, which is most clearly seen in the suppression of the buildup of electronic coherence. However, for weak and moderate vibronic coupling – closer to the typical values in organic molecules – SR is still possible. Interestingly, the vibrational coupling can even assist superradiance when the cavity frequency is smaller than the electronic transition frequency (negative cavity detuning). While a pure dephasing model can account for the suppression of SR, it fails to capture the dependence on the cavity detuning predicted in the HTC model, which is a measurable effect in experiments.

The structure of the paper is as follows. In Section 2, we introduce the dissipative TC model and our exact numerical solution method. We benchmark mean-field and second-order cumulant solutions against the exact solution for SF and SR initial conditions, and for different numbers of emitters. In Section 3, we introduce the open HTC model, compute its dynamics using mean-field theory, and compare it to the TC model with a pure dephasing term. Section 4 concludes the paper.

2 Numerical solution for the dissipative Tavis–Cummings model

2.1 Model description

We start by considering a system of N two-level systems coupled to a single cavity mode, as in Figure 1(a)i. In the dipole and rotating-wave approximations, such a system is well described by the TC Hamiltonian [33]. Setting = 1, the system Hamiltonian is

(1) H ̂ TC = ω c a ̂ a ̂ + i = 1 N ω 0 2 σ ̂ i z + g a ̂ σ ̂ i + + a ̂ σ ̂ i ,

where a ̂ and a ̂ are the annihilation and creation operators of the cavity mode of frequency ω c, σ ̂ i z , ± are Pauli matrices describing the ith emitter satisfying [ σ ̂ i z , σ ̂ j ± ] = ± 2 δ i j σ ̂ i ± , ω 0 is the energy splitting of the emitters, and g is the individual light–matter coupling (which depends on the mode volume). To include loss and dephasing processes, we employ a GKSL master equation for the density matrix ρ [79],

(2) t ρ = i [ H ̂ TC , ρ ] + D TC ,

(3) D TC = κ L [ a ̂ ] + i = 1 N γ L [ σ ̂ i ] + γ ϕ L [ σ ̂ i z ] ,

with L [ X ̂ ] = X ̂ ρ X ̂ { X ̂ X ̂ , ρ } / 2 . The rate κ describes the loss of photons from the cavity, γ individual loss of molecular excitations (e.g., by spontaneous emission into a non-confined radiation mode), and γ ϕ individual molecular dephasing. The last pure dephasing term can be understood as a Markovian approximation to the effect of vibrational modes coupling to the emitters.

To treat dynamical SR, we choose an initial state ρ(0) = |ψ(θ)⟩⟨ψ(θ)| with

(4) | ψ ( θ ) = | 0 i = 1 N e i θ σ ̂ i x / 2 | i ,

where |↑⟩ i is an excited (inverted) emitter state in the Hilbert space of the ith two-level system. The initial state has zero photons in the cavity mode, and the emitters are fully inverted and coherently tilted by an angle θ. For θ = 0, the emitters are fully inverted, giving SF, whereas for θ ≠ 0, there is some initial dipole moment present, giving SR [Figure 1(b)]. Note that a rotation of the inverted states about any axis in the xy plane (i.e., exp i n x σ ̂ i x + n y σ ̂ i y θ / 2 ; n x 2 + n y 2 = 1 ) leads to an emitter state with equivalent initial coherence as in Eq. (4), and we choose here the x-axis for simplicity (i.e., n x = 1, n y = 0). Such states of two-level systems are the analog of coherent states in bosonic systems, which are also referred to as spin-coherent states or atomic-coherent states [80], [81], and they have recently been used to study entanglement in Dicke SR [82].

2.2 Numerical method

To solve Eq. (2) numerically, we first make use of the weak permutation symmetry of the emitters [61], [62], [63], [64], [65], [66], [67] by expanding the density matrix as

(5) ρ = λ , n , n ρ λ , n , n | n n | O ̂ λ ,

where |n⟩⟨n′| is the state of the cavity mode with photon numbers n and n′, and O ̂ λ is a permutation-symmetric density matrix depending on the N-emitter state labeled by λ. Since we consider two-level systems, the permutation-symmetric states can be labeled by enumerating the ways of dividing N elements into four bins corresponding to the possible emitter matrix elements of a single two-level system ↑↑, ↓↑, ↑↓, and ↓↓ (↑: excited state, ↓: ground state). Thus, λ describes the 4-element lists m λ = (m λ,↑↑, m λ,↓↑, m λ,↑↓, m λ,↓↓) with m λ,p ≥ 0 and p m λ,p = N. Explicitly, each m λ corresponds to a matrix O ̂ λ that consists of the sum of all states obtainable by permuting pairs of emitters in a state with m λ,↑↑ inverted emitters in the left and right states of the density matrix, followed by m λ,↓↑ ground-state emitters in the left state and inverted emitters in the right state, and so on.

Using the mapping described above, one can, in principle, construct the operators O ̂ λ in the original explicit 2 N dimensional two-level-system basis. The permutation approach is, however, based on avoiding the explicit representation and performing all calculations within the compressed space. The number of unique density-matrix elements in the emitter space is given by the number of possible partitions m λ , which is equal to N + 3 3 .

Since the initial state in Eq. (4) has N excitations in the emitter state and there are no gain terms in Eq. (2), the photon Hilbert space can be truncated at N + 1 without introducing any approximations. The total number of density-matrix elements, therefore, scales as ( N + 1 ) 2 × N + 3 3 , and an exact solution of Eq. (2) becomes possible up to order of N ∼ 30 emitters.

We can further exploit the weak [83], [84] U(1) symmetry of the master equation (2), which states that it is invariant under a ̂ a ̂ e i ϕ , σ ̂ n σ ̂ n e i ϕ . Let us denote the total number of excitations in the left and right states of a given density-matrix element ρ λ,n,n as ν = n + m λ,↑↑ + m λ,↑↓ and ν′ = n′ + m λ,↑↑ + m λ,↓↑. As a consequence of the weak U(1) symmetry, only elements with a constant value of νν′ can couple in the dynamics. This motivates the introduction of the block form of the density matrix

(6) ρ = ν , ν ρ ( ν , ν ) ,

where ρ (ν, ν′) denotes the collection of all elements with left and right excitation numbers equal to ν and ν′, respectively. The most general equations of motion for the individual blocks respecting the weak U(1) symmetry then read

(7) t ρ ( ν , ν ) = k Z L k ( ν , ν ) ρ ( ν + k , ν + k ) ,

where L k ( ν , ν ) are matrices determined by the master equation. All terms with k < 0 describe gain processes, k = 0 describes the coherent dynamics and dephasing, and k > 0 describes losses. Note that blocks with different values of ν and ν′ do couple, but their difference νν′ is constant. Such a structure of master equation – combining weak U(1) symmetry with weak permutation symmetry – was used in Ref. [68] to discuss steady state superradiance in a system of Rydberg atoms.

Since Eq. (2) features jump operators that change the total excitation number by at most one, and the system we consider has no gain, only the terms k = 0 and k = 1 are nonzero in Eq. (7). Further, we are primarily interested in calculating the expectation value of the number of photons in the cavity mode n ̂ = a ̂ a ̂ , so we can restrict Eq. (7) to only those blocks with νν′ = 0 on which the photon number operator depends. In this case, we arrive at the evolution equation for the blocks ρ (ν)ρ (ν,ν)

(8) t ρ ( ν ) = L 0 ( ν ) ρ ( ν ) + L 1 ( ν ) ρ ( ν + 1 ) .

The matrices L 0 ( ν ) and L 1 ( ν ) can be constructed from Eq. (2). Crucially, since we consider a system without gain, we can solve Eq. (8) sequentially by noting that there exists a block with the highest excitation number ν max. For this block, the second term on the right-hand side in Eq. (8) vanishes, resulting in a homogeneous system of differential equations for the block ρ ( ν max ) , which can be numerically integrated. The solution ρ ( ν max ) may then be used to solve the block ρ ( ν max 1 ) . Iterating until ν = 0 yields dynamics for all blocks with νν′ = 0. We note that this method of solution is reminiscent of that used to solve the diagonal elements of the density matrix in Dicke SR [85].

From the initial state in Eq. (4), we see that the maximum excitation number is ν max = N. The number of elements of the blocks ρ (ν) for a fixed N increases monotonically from 1 at ν = 0 to N + 3 3 at ν = N. The largest block must have exactly as many elements as there are emitter states, because for every emitter state, one can always find photon numbers n and n′ such that ν = ν′ = N.

Code for the numerical scheme described is available in the PIBS (Permutational Invariance Block Solver) Python package [86]. This method allows one to solve Eq. (2) with initial conditions given in Eq. (4) up to N = 140 emitters. Currently, only the solution for the diagonal block ν = ν′ is implemented in the PIBS code, but an extension to off-diagonal blocks is straightforward.

2.3 Benchmarking Cumulant approaches

In this section, we compare the exact solution of Eq. (2) based on PIBS to approximate first- and second-order cumulant approaches. In a cumulant approximation of order M, one splits expectation values of M + 1 operators in the Heisenberg equations of motion into expectation values of at most M operators by setting the cumulants of order M + 1 to zero. For example, in a first-order cumulant approximation, the expectation values of any two operators A ̂ and B ̂ are split as A ̂ B ̂ = A ̂ B ̂ , because the second-order cumulant given by A ̂ B ̂ = A ̂ B ̂ A ̂ B ̂ [87] is set to zero. This is equivalent to mean-field theory. Going one order higher, keeping expectations of pairs of operators and setting third-order cumulants to zero, one arrives at the second-order cumulant approximation.

If the initial state respects the U(1) symmetry of the Hamiltonian in Eq. (1), one can simplify the second-order cumulant equations by dropping expectation values of operators that do not conserve the number of excitations, such as a ̂ and σ ̂ i + . If these “symmetry-breaking” terms are zero initially, then they remain zero for all times. The resulting equations may be called symmetry-preserving cumulant equations [73].

The initial state in Eq. (4) breaks the U(1) symmetry for θ ≠ 0, since σ ̂ i + = i sin ( θ ) / 2 . As a consequence, for SR initial conditions, one must retain all terms in the cumulant expansion, resulting in symmetry-broken cumulant equations. We use the Julia package QuantumCumulants.jl [88] to calculate and evolve the symmetry-broken cumulant equations.

In Figure 2, we compare the cumulant approaches to the exact solution for different numbers of emitters and initial conditions, by plotting the normalized average number of photons in the cavity n ̂ / N = a ̂ a ̂ / N as a function of time. For brevity, we refer to the first-order cumulant approximation as mean field and to the second-order cumulant approximation simply as cumulants. We do not show a mean-field solution for SF initial conditions, as the mean-field solution in that case is just zero for all times if we start from the symmetric initial state. We also note that in the mean-field treatment, n ̂ / N = | a ̂ | 2 / N becomes independent of N.

Figure 2: 
Comparison of exact dynamics of the average number of photons in the cavity calculated with PIBS and cumulant approaches. Panels (a) and (c) show the dynamics for SF initial conditions for N = 10 and N = 140. Panels (b) and (d) show the dynamics for SR initial conditions (θ = π/4) for N = 10 and N = 100. Other parameters: 


g


N


=
0.4

e
V
,

Δ
=


ω


c


−


ω


0


=
−
0.35

e
V
,

κ
=
0.01

e
V
,

γ
=
0.001

e
V
,



γ


ϕ


=
0.0075

e
V


$g\sqrt{N}=0.4\,\mathrm{e}\mathrm{V},\,{\Delta }={\omega }_{\mathrm{c}}-{\omega }_{0}=-0.35\,\mathrm{e}\mathrm{V},\,\kappa =0.01\,\mathrm{e}\mathrm{V},\,\gamma =0.001\,\mathrm{e}\mathrm{V},\,{\gamma }_{\phi }=0.0075\,\mathrm{e}\mathrm{V}$



.
Figure 2:

Comparison of exact dynamics of the average number of photons in the cavity calculated with PIBS and cumulant approaches. Panels (a) and (c) show the dynamics for SF initial conditions for N = 10 and N = 140. Panels (b) and (d) show the dynamics for SR initial conditions (θ = π/4) for N = 10 and N = 100. Other parameters: g N = 0.4 e V , Δ = ω c ω 0 = 0.35 e V , κ = 0.01 e V , γ = 0.001 e V , γ ϕ = 0.0075 e V .

Figure 2(a) and (c) shows the dynamics for N = 10 and N = 140 for SF initial conditions (θ = 0). One can see that cumulants underestimate the dephasing processes, as the amplitude of oscillations is much larger than that in the exact solution. As such, for N = 140, we only see agreement between the cumulant approximation and the exact solution at early times, specifically for t < 10 fs. At later times, there is no agreement over the range of N for which we can obtain exact results. Part of the dephasing (which is different from the dephasing due to the vibrations) causing this mismatch can be explained by the quantum corrections to the energy eigenvalues of the TC Hamiltonian for finite N [34]. As discussed there, the mean-field (classical) solution with periodic oscillations corresponds to the case of equal (harmonic) energy-level spacing. At finite N, however, the level spacing is not perfectly regular. Using the fact that the resulting anharmonicity is known to scale as 1/(ln N)3, and that the cumulant solution for N = 140 matches the exact solution up to t ≈ 10 fs [cf. Figure 2(c)], we can estimate that, e.g., for N = 108, this dephasing would set in after t ≈ 520 fs.

Figure 2(b) and (d) compares the exact solution to cumulants and mean field for an SR initial state with θ = π/4. Generally, the cumulant solution matches the exact solution better than the mean-field solution. Moreover, for increasing N, the early time period where cumulants match the exact solution increases. This indicates that for SR initial states, symmetry-broken cumulants capture the early-time behavior well. We provide quantitative evidence that this is the case in Appendix A, where we show that both mean-field and cumulants converge monotonically toward the exact solution for N up to 100.

As an additional test of mean-field results, we note that for SR initial states, the symmetry-broken cumulant solution converges to the mean-field solution for large enough N, as demonstrated in Figure 3. Since we are ultimately interested in the SR initial state, this leads us to conclude that for large N, it is reasonable to use mean-field theory.

Figure 3: 
Comparison of symmetry-broken second-order cumulant approximation to mean-field approximation for (a) N = 100 [same as Figure 2(d)], (b) N = 1,000, and (c) N = 10,000. The SR angle is set to θ = π/4, all other parameters are the same as in Figure 2.
Figure 3:

Comparison of symmetry-broken second-order cumulant approximation to mean-field approximation for (a) N = 100 [same as Figure 2(d)], (b) N = 1,000, and (c) N = 10,000. The SR angle is set to θ = π/4, all other parameters are the same as in Figure 2.

3 Superradiance with coupling to vibrations

3.1 Model description

In the previous section, we discussed the dissipative TC model with local dephasing, capturing the emitters vibronic coupling within a Markovian approximation. We now introduce a model that includes vibrations explicitly in the system dynamics. As a simplest case, we assume that there is one dominant vibrational mode of frequency ω ν , which is the same for all molecules. We extend the TC model from Eq. (1) to include this mode,

(9) H ̂ HTC = H ̂ TC + i = 1 N ω ν b ̂ i b ̂ i + S b ̂ i + b ̂ i σ ̂ i z ,

which is the HTC model [47], [56].

The vibrational mode of the ith molecule is treated as a harmonic oscillator with annihilation and creation operators b ̂ i and b ̂ i . The coupling strength between the vibrations and the electronic degrees of freedom is set by the Huang–Rhys parameter S. Including the same dissipation processes as before (see Eq. (3) [89]), and adding thermalization processes for the vibrational modes, the master equation is

(10) t ρ = i [ H ̂ HTC + H ̂ LS , ρ ] + D TC + i = 1 N γ L [ b ̂ i ] + γ L [ b ̂ i ] ,

where D TC was defined in Eq. (3). The thermalization rates are γ = γ ν n B and γ = γ ν (n B + 1), where n B = e ω ν / T 1 1 is the Bose–Einstein distribution at temperature T (we set k B = 1). The model of thermalization here, which includes a Lamb-shift term H ̂ LS = i ( γ ν / 4 ) i = 1 N b ̂ i b ̂ i b ̂ i b ̂ i , is one of momentum damping, as discussed in Ref. [90]. The energy of the HTC Hamiltonian (9) depends on the vibrational displacement x ̂ i b ̂ i + b ̂ i , and the thermalization drives the system toward x ̂ i such that the energy is minimized. In Figure 1(a)ii, the internal level structure and incoherent processes of the HTC model are shown.

Using Eq. (4), we can write the initial state including vibrations as

(11) ρ ( 0 ) = | ψ ( θ ) ψ ( θ ) | i = 1 N ρ vib , i ,

where ρ vib , i = 1 e ω ν / T e ω ν b ̂ i b ̂ i / T is a thermal state of the vibrational mode of molecule i.

The Hilbert space of the HTC model is much larger than the one considered in Section 2. Truncating the vibrational mode to N ν levels increases the Hilbert space by a factor of N ν N , which makes an exact solution quickly intractable for growing N. However, from Section 2, we have confidence in mean-field theory capturing the essential early-time behavior also for the HTC model and, therefore, solve the dynamics using the mean-field approximation.

In what follows, we compare the dynamics of the model Eq. (10) for two different cases [cf. Figure 1(a)]: (i) S = 0 and variable γ ϕ , which just reduces to the TC model Eqs. (1) and (2), and (ii) γ ϕ = 0 and variable S. We do this comparison because both S and γ ϕ model forms of dephasing by phonons, and we are interested in how these different processes in solid-state environments modify superradiance.

3.2 Results

The results shown here were calculated in the mean-field approximation using the Julia package QuantumCumulants.jl [88]. We consider SR initial conditions with θ = 10−3 π throughout, and set the number of molecules N = 108.

Figure 4(a) shows the average number of photons in the cavity mode as a function of time for different values of S. For S = 0, the HTC model reduces to the TC model, resulting in damped Rabi-oscillations (note the logarithmic scale). For S ≤ 0.3, after a build-up time where n ̂ / N grows from 0 up to roughly 10−6, there is a region where the photon number increases exponentially with time as n ̂ e t / τ . The parameter τ is the characteristic risetime, which is seen to increase with S. Increasing S beyond a value of 0.3 changes the exponential behavior after the build-up time, and a characterization based on a single risetime τ is no longer meaningful. It is, however, clear that a stronger coupling to the vibrational modes inhibits the emission of photons into the cavity mode.

Figure 4: 
Mean-field solution of the HTC model starting from SR initial conditions. (a), (b), and (c) Average number of photons in the cavity mode, electronic coherence, and magnitude of the Bloch vector, respectively, as a function of time for different coupling strengths to the vibrational mode. (c), (d), and (f) Same as (a), (b), and (c) but with the vibrational mode replaced by a pure dephasing term with rate γ

ϕ
. The curves S = 0 in (a), (b), and (c) are the same as the curves with γ

ϕ
 = 0 in (d), (e), and (f). Note that the values of S and γ

ϕ
 are not equidistantly spaced. Parameters: 


N
=
1


0


8


,

g


N


=
0.2

e
V
,

Δ
=
0
,

κ
=
0.01

e
V
,

γ
=
1


0


−
6



e
V
,

θ
=
1


0


−
3


π
,



ω


ν


=
0.15

e
V
,



γ


ν


=
0.01

e
V
,

T
=
0.026

e
V


$N=1{0}^{8},\,g\sqrt{N}=0.2\,\mathrm{e}\mathrm{V},\,{\Delta }=0,\,\kappa =0.01\,\mathrm{e}\mathrm{V},\,\gamma =1{0}^{-6}\,\mathrm{e}\mathrm{V},\,\theta =1{0}^{-3}\pi ,\,{\omega }_{\nu }=0.15\,\mathrm{e}\mathrm{V},\,{\gamma }_{\nu }=0.01\,\mathrm{e}\mathrm{V},\,T=0.026\,\mathrm{e}\mathrm{V}$



. For (a), (b), and (c): γ

ϕ
 = 0; for (d), (e), and (f): S = 0.
Figure 4:

Mean-field solution of the HTC model starting from SR initial conditions. (a), (b), and (c) Average number of photons in the cavity mode, electronic coherence, and magnitude of the Bloch vector, respectively, as a function of time for different coupling strengths to the vibrational mode. (c), (d), and (f) Same as (a), (b), and (c) but with the vibrational mode replaced by a pure dephasing term with rate γ ϕ . The curves S = 0 in (a), (b), and (c) are the same as the curves with γ ϕ = 0 in (d), (e), and (f). Note that the values of S and γ ϕ are not equidistantly spaced. Parameters: N = 1 0 8 , g N = 0.2 e V , Δ = 0 , κ = 0.01 e V , γ = 1 0 6 e V , θ = 1 0 3 π , ω ν = 0.15 e V , γ ν = 0.01 e V , T = 0.026 e V . For (a), (b), and (c): γ ϕ = 0; for (d), (e), and (f): S = 0.

In Figure 4(d), the vibrational mode has been replaced by a pure Markovian dephasing term with rate γ ϕ . For all values of γ ϕ , there is a region where the photon number increases exponentially, and the risetime increases with increasing γ ϕ . Thus, the pure dephasing model can capture the photon dynamics for small S below 0.3 but breaks down for larger values of S.

To characterize the extent to which the vibrational modes destroy superradiance, we additionally calculate the electronic coherence | σ ̂ + | | σ ̂ i + | for the same sets of parameters used in Figure 4(a) and (d), see Figure 4(b) and (e). Since we use SR initial conditions, the coherence at t = 0 is nonzero and given by | σ ̂ + | = sin ( θ ) / 2 . Figure 4(b) shows that the coherence increases exponentially at early times for small S, similar to the exponential rise in the photon number in Figure 4(a). For S ≳ 0.3, the exponential increase of the coherence is inhibited, and there are regions where the coherence decreases before it has reached its maximum. In Appendix B, we show that increasing the initial amount of coherence (increasing θ) tends to make the initial rise of photon number faster and less dependent on the vibrational coupling, and in Appendix C, we show that the threshold of SR at a certain value of S (here at S ≈ 0.3) emerges as a competition between the light–matter coupling g N and the vibronic coupling ω ν S .

In Figure 4(e), where the vibrational mode has been replaced by a pure dephasing term, we see an exponential increase of the coherence for low dephasing rates and early times, similar to the behavior in Figure 4(b) for small S. Large dephasing rates also destroy superradiance, as can be seen by the initial decrease of coherence for γ ϕ ≳ 0.06 eV. We note that the long-time behavior of the coherence is quite different in Figure 4(b) and (e). In the former case, the coherence builds up eventually and remains relatively large, whereas in the latter case, the coherence decreases. This shows that, when regarding ⟨ σ ̂ +⟩, the coupling to vibrations does not result in the same kind of decoherence as introduced by a pure dephasing term, even for small S.

Lastly, Figure 4(c) and (f) shows the magnitude of the Bloch vector J ̂ 2 = J ̂ x 2 + J ̂ y 2 + J ̂ z 2 , J ̂ α = i = 1 N σ ̂ i α / 2 , which, within mean-field theory and using N ≫ 1, is given by

(12) J ̂ 2 = N 2 4 σ ̂ z 2 + 4 | σ ̂ + | 2 .

The initial state in Eq. (4) is an eigenstate of J ̂ 2 with the maximum possible eigenvalue N/2(N/2 + 1) ≈ N 2/4. In the absence of incoherent processes and vibronic coupling, i.e., for the closed TC model in Eq. (1), the emitter state remains an eigenstate of J ̂ 2 with the same eigenvalue since [ H ̂ TC , J ̂ 2 ] = 0 and, therefore, J ̂ 2 N 2 / 4 for all times. This is seen for the curves with S = 0 or γ ϕ = 0 in Figure 4(c) and (f). We note that the effect of local molecular losses proportional to γ = 10−6 eV is negligible in the time period considered here, which would otherwise lead to a decrease in J ̂ 2 .

For nonzero S or γ ϕ , the magnitude of the Bloch vector does not retain its maximum initial value. In the case of vibrational dressing in Figure 4(c), J ̂ 2 is subject to oscillations but stays well above zero in the considered time frame. In contrast, for pure dephasing in Figure 4(f), J ̂ 2 decreases monotonically to zero, with minor bumps for small dephasing rates, showing a stark difference between the two models of treating the vibrational effects.

Such behavior can be related to discussions of how any processes that break spin conservation destroy free-space superradiance [91], [92], [93]. As noted in those works, when the dynamics approaches the equator – specifically when | σ ^ z | 1 / N – simple counting arguments show there are many more accessible subradiant states (quantum states with J ̂ 2 N ) than superradiant states. As such, any process that can change the modulus of J ̂ will lead to transitions to these subradiant states, reducing J ̂ 2 and ultimately destroying the superradiance. Notably, these processes only become relevant as one approaches the equator: when ⟨ σ ̂ z ⟩≃ 1, scattering to subradiant states is suppressed if the scattering does not directly change ⟨ σ ̂ z ⟩. References [91], [92], [93] discuss how direct interactions between different emitters, breaking the permutation symmetry, have such an effect. Here, we see similar behavior driven by individual dephasing or vibrational modes.

Interestingly, for both models, a stronger influence of the vibrational modes (i.e., larger S or larger γ ϕ ) leads to an increased time period, for which J ̂ 2 stays close to its maximum initial value. This can be explained by the decreased emission rate of photons for larger S and γ ϕ , see Figure 4(a) and (d), which means that the excitation state remains close to its initial value of σ ̂ z 1 for longer times and thus avoids states near the equator where transitions to subradiant states arise. According to Eq. (12), this results in J ̂ 2 N 2 / 4 .

In experiments, the cavity frequency can typically be readily varied; our calculations predict interesting phenomena when varying the cavity detuning Δ = ω cω 0, with a clear difference between pure dephasing and having coupling to a vibrational mode. The dynamics of the photon mode in the HTC model is plotted for Δ = −0.15 eV in Figure 5(a) and for Δ = −0.3 eV in Figure 5(b). Compared to the case of zero detuning in Figure 4(a), we see that a negative detuning favors the emission of photons for larger values of S. This is seen most clearly in Figure 5(b), where the curves up to S = 0.4 almost collapse into one curve at early times. In the inset of Figure 5(b), we observe that the photon number rises faster for 0.1 ≤ S ≤ 0.4 than for S = 0. This can be explained by first noting that, for ω ν = 0.15 eV and T = 0.026 eV used here, the phonons are almost fully in the ground state at t = 0. Further, for a negative detuning, the cavity frequency ω c = ω 0 + Δ is smaller than the level splitting of the electronic ground and excited ω 0. Thus, the cavity frequency matches transitions from the excited state manifold with zero vibrational excitations to the ground-state manifold with vibrational excitations. As S increases, the most probable number of vibrational excitations created in an optical transition increases. As such, one sees an optimal value of S in this regime where processes exciting one vibrational mode become dominant. In contrast, if the vibrational mode is treated as a pure dephasing term, such a reordering phenomenon does not occur.

Figure 5: 
Average number of photons in the cavity mode as a function of time for different coupling strengths to the vibrational mode for (a) Δ = −0.15 eV and (b) Δ = −0.3 eV. All other parameters are the same as in Figure 4, and we specifically note that the vibrational frequency is set to ω

ν
 = 0.15 eV.
Figure 5:

Average number of photons in the cavity mode as a function of time for different coupling strengths to the vibrational mode for (a) Δ = −0.15 eV and (b) Δ = −0.3 eV. All other parameters are the same as in Figure 4, and we specifically note that the vibrational frequency is set to ω ν = 0.15 eV.

For a final set of results, we calculate the risetime τ for different values of S, Δ, and γ ϕ , while making sure to stay in a region where τ is well defined. Figure 6(a) shows τ as a function of S and different values of Δ. For small negative detunings, τ increases monotonically with S, as evident in Figure 4(a). In contrast, large negative detunings Δ ≤ −0.25 eV show regions where increasing S leads to a shorter risetime, which is the behavior observed in Figure 5. This behavior is confirmed in Figure 6(b), which shows τ as a function of Δ for different S. Here, we see in particular how the minimum of τ at zero detuning for the TC model (S = 0) shifts to more negative detunings as S > 0 is increased, and further, the symmetry about the minimum value of τ is lost. This symmetry remains in the case where the vibrational mode is replaced by a pure dephasing rate, irrespective of the strength of the dephasing, see Figure 6(c). In that case, the risetime increases as γ ϕ increases, and the minimum value is always at zero detuning.

Figure 6: 
Photon number risetime extracted from an exponential fit to the linear regions as seen for example in Figure 4(a) and (c). (a) Risetime as a function of S for different detunings. (b) Risetime as a function of detuning for different values of S. (c) Same as (b), but the vibrational mode has been replaced by a pure dephasing term with rate γ

ϕ
. Note that the curve with S = 0 in (b) is the same as the curve with γ

ϕ
 = 0 in (c). All other parameters are the same as in Figure 4.
Figure 6:

Photon number risetime extracted from an exponential fit to the linear regions as seen for example in Figure 4(a) and (c). (a) Risetime as a function of S for different detunings. (b) Risetime as a function of detuning for different values of S. (c) Same as (b), but the vibrational mode has been replaced by a pure dephasing term with rate γ ϕ . Note that the curve with S = 0 in (b) is the same as the curve with γ ϕ = 0 in (c). All other parameters are the same as in Figure 4.

The risetime behavior for different values of θ is discussed in Appendix B. We find that there is an initial rapid rise to a higher photon number for large θ than for a small one. That is, the photon number rise is faster for a smaller amount of electronic excitation (smaller inversion). This is a curious behavior, opposite, for example, to usual pulsed lasing, where the pulse build-up time becomes shorter for a higher amount of excitation [55]. This is because coherence plays a crucial role in SR, and larger θ (up to π/2) means higher initial coherence.

4 Conclusions

We have studied how processes of vibrational coupling and pure dephasing – relevant to realistic models of emitters in solid-state environments – affect the dynamics of superradiance and superfluorescence for organic molecules in an optical cavity. To enable this study, we introduced an efficient method leveraging both weak permutation and weak U(1) symmetry of the many-molecule–cavity system that allows an exact solution for a large number of emitters. For the Tavis–Cummings model with dissipation and dephasing, solutions were obtained up to N ∼ 140 emitters. While we focused the exact method on the simple case of two-level emitters in a single-mode cavity, it can be extended to handle more complex multilevel systems and other near-resonant cavity modes. This offers a valuable tool to benchmark other numerically exact and approximate methods for cavity-QED systems, greatly extending the range of system sizes that can be handled exactly.

Using our exact method, we benchmarked mean-field and second-order cumulant solutions, verifying their accuracy in capturing early-time dynamics when the initial state has nonzero initial coherence. This justified the application of mean-field theory to study the effect of the vibrational environment of organic emitters on the dynamics. Here, we compared two models: a Holstein–Tavis–Cummings model including a discrete vibrational mode for each emitter, and a simplified phenomenological model where the vibronic coupling is replaced by Markovian pure dephasing. We found that for small vibrational coupling or dephasing, both models predict an exponential rise of the photon number at early times. In contrast, strong vibrational coupling leads to qualitatively different photon dynamics in the case of vibrational dressing, which the pure dephasing model cannot capture. Analysis of the electronic coherence revealed that sufficiently strong coupling to vibrations (roughly, S ≥ 0.3) suppresses dynamical superradiance. As the values of S in typical organic molecules are at a maximum around 0.1, dynamical superradiance should be possible. On the other hand, the maximal values of the order 0.1 are large enough to see specific effects of the vibrational coupling, such as the asymmetry of the risetime as a function of cavity detuning.

For experimental studies of the superradiance in the presence of vibrational coupling, the dependence of the risetime on the cavity detuning is interesting. The detuning can typically be easily varied in experiments, and an asymmetry of the risetime as a function of detuning would be a clear signature of a SR process that involves a coherent coupling to vibrational states. In contrast, if a symmetric behavior is observed, it would indicate that the coupling to the vibrational states is not coherent, and they act only as an effective dephasing environment. Another interesting experiment would be to compare the effect of coherent and incoherent pumping. The former is required for the superradiance, and in that case, one would see a more rapid rise in the photon number for smaller inversion, while for incoherent pumping leading to lasing, one would expect the opposite. The time-scales for the parameters we have chosen to use are very small, tens of femtoseconds, but for other parameter regimes, the dynamics could be brought to the regime accessible to state-of-the-art time-resolved techniques.

The power of the mean-field approaches lies in their independence of computation complexity from system size, which allows for solutions for arbitrarily large N, as well as their broad applicability. Future work may combine these and exact approaches to study the effects of multiple vibrational modes or a continuum of modes for the local emitter environments, opening pathways to explore superradiance and related phenomena across the diverse range of materials and environments encountered in real physical systems.


Corresponding authors: Jonathan Keeling, SUPA, School of Physics and Astronomy, University of St Andrews, St Andrews, KY16 9SS, UK, E-mail: ; and Päivi Törmä, Department of Applied Physics, Aalto University School of Science, FI-00076 Aalto, Espoo, Finland, E-mail:

Award Identifier / Grant number: Future Makers

Award Identifier / Grant number: 339313

Award Identifier / Grant number: Photonics Research and Innovation (PREIN) / 346529

Acknowledgments

Part of the calculations presented above were performed using computer resources within the Aalto University School of Science “Science-IT” project.

  1. Research funding: This work was supported by the Research Council of Finland under project number 339313, and by the Jane and Aatos Erkko Foundation and the Technology Industries of Finland Centennial Foundation as part of the Future Makers funding program. The work is part of the Research Council of Finland Flagship Programme, Photonics Research and Innovation (PREIN), decision number 346529, Aalto University.

  2. Author contributions: LF, PFW, JC, BWL, JK, and PT have accepted responsibility for the entire content of this manuscript and consented to its submission to the journal, reviewed all the results, and approved the final version of the manuscript. PT and JK initiated and supervised the research. LF and PFW wrote the code used in the calculations. LF performed the calculations. LF, PFW, JC, BWL, JK, and PT participated in discussing and interpreting the results. LF prepared the manuscript with contributions from all coauthors.

  3. Conflict of interest: Authors state no conflicts of interest.

  4. Data availability: The datasets generated and analyzed during the current study are available in the Zenodo repository at https://doi.org/10.5281/zenodo.17542025.

Appendix A: Convergence of mean field approaches

In this appendix, we quantify the convergence of mean-field (mf) and second-order cumulant (c2) solutions to the exact solution for the dissipative Tavis–Cummings model from Section 2 for superradiance initial conditions (θ = π/4). To this end, we define the sum-squared error between the mean-field or cumulant solution and the exact solution for a given quantity x ̂ as

(A1) e α ( x ̂ ) = i = 1 M ( x ̂ ( t i ) α x ̂ ( t i ) PIBS ) 2 ,

where α ∈ {mf, c2}. The upper bound M of the sum is given by a maximum time t max = t M up to which the error is calculated. Figure 7(a) shows the so-defined error for the normalized photon number n ̂ / N and for the two-level system population σ ̂ + σ ̂ for different values of N ranging from 5 to 100. The maximum time is t max = 130 fs, which covers the entire domain calculated in Figure 2. For increasing N, the mean-field error monotonically decreases for both calculated quantities. However, this is not the case for the cumulant solution, which does not exhibit a clear convergence behavior. This may at first seem surprising, but the apparent lack of convergence lies in the choice of t max. As can be seen in Figure 2(b) and (d), the cumulant solution agrees well with the exact solution up to some time t 0, after which the cumulant solution starts having irregular oscillations. It is in this region t > t 0 where a large error can build up in the cumulant solution, because the photon number oscillations can be out of phase with the exact solution. In contrast, the mean-field solution shows regular oscillations in the photon number for all times [cf. Figure 2(b), and (d)]. To show the convergence in the initial region t < t 0, we set t max = 10 fs in Figure 7(b) and calculate the same error quantities as in Figure 7(a). The error for the photon number and for the two-level system population are equal in that case, and for both the mean-field solution and the cumulant solution, a clear convergence behavior is seen for increasing N. These results reinforce the main message from Section 2, namely that in the large N limit, the mean-field approximation is reasonable for superradiance initial conditions.

Figure 7: 
Sum-squared error of the normalized photon number and the two-level system excitation for both mean-field and second-order cumulant solutions as defined in Eq. (A1) for (a) t
max = 130 fs and (b) t
max = 10 fs. For the mean-field solution 



⟨



n

̂


⟩

=
|

⟨



a

̂


⟩



|


2




$\langle \hat{n}\rangle =\vert \langle \hat{a}\rangle {\vert }^{2}$



 and 



⟨





σ

̂



+






σ

̂



−



⟩

=
|

⟨





σ

̂



−



⟩



|


2




$\langle {\hat{\sigma }}^{+}{\hat{\sigma }}^{-}\rangle =\vert \langle {\hat{\sigma }}^{-}\rangle {\vert }^{2}$



. The number of two-level systems ranges from N = 5 to N = 100 and the superradiance angle is θ = π/4. All other parameters are the same as in Figure 2.
Figure 7:

Sum-squared error of the normalized photon number and the two-level system excitation for both mean-field and second-order cumulant solutions as defined in Eq. (A1) for (a) t max = 130 fs and (b) t max = 10 fs. For the mean-field solution n ̂ = | a ̂ | 2 and  σ ̂ + σ ̂ = | σ ̂ | 2 . The number of two-level systems ranges from N = 5 to N = 100 and the superradiance angle is θ = π/4. All other parameters are the same as in Figure 2.

Appendix B: Influence of θ on dynamical SR

In this appendix, we investigate the influence of the angle θ on dynamical SR in the presence of vibrational dressing. All results are calculated with mean-field theory. Figure 8 shows the average number of photons and the electronic coherence as a function of time for three different values of θ [Figure 8(a) and (d) is the same as Figure 4(a) and (b)]. Clearly, an increase in θ leads to an increase in initial electronic coherence |⟨ σ ̂ +⟩| = sin(θ)/2, resulting in an increased photon emission rate. However, irrespective of θ, for strong vibrational coupling S ≳ 0.4, the coherence tends to decrease at early times, which indicates that SR is suppressed.

Figure 8: 
Average number of photons in the cavity mode as a function of time in the HTC model for (a) θ = 10−3
π, (b) θ = 10−2
π, and (c) θ = 10−1
π, for different vibrational coupling strengths. Panels (d)–(f) show the electronic coherence for the same values of θ. All other parameters are the same as in Figure 4.
Figure 8:

Average number of photons in the cavity mode as a function of time in the HTC model for (a) θ = 10−3 π, (b) θ = 10−2 π, and (c) θ = 10−1 π, for different vibrational coupling strengths. Panels (d)–(f) show the electronic coherence for the same values of θ. All other parameters are the same as in Figure 4.

Figure 9(a) and (b) shows how the angle θ changes the risetime behavior in the case of vibrational dressing for Δ = 0 and Δ = −0.3 eV, respectively. We set S = 0.2, which ensures that the system still exhibits an exponential rise behavior at early times. We see that for small values of θ, the risetime (defined by fitting Ae t/τ to the exponential regions) is approximately constant. However, increasing θ results in an enhanced initial photon emission, before the exponential regime is reached. If θ becomes too large, the exponential regime disappears.

Figure 9: 
Average number of photons as a function of time in the HTC model for different values of θ and S = 0.2, for (a) Δ = 0 and (b) Δ = −0.3 eV. Note the nonuniform spacing of θ. All other parameters are the same as in Figure 4.
Figure 9:

Average number of photons as a function of time in the HTC model for different values of θ and S = 0.2, for (a) Δ = 0 and (b) Δ = −0.3 eV. Note the nonuniform spacing of θ. All other parameters are the same as in Figure 4.

Appendix C: Influence of g N on dynamical SR

We study here in more detail the effect of the light–matter coupling strength g N on the threshold of dynamical SR reported in Section 3. To this end, we calculate the same dynamics as in Figure 4(a) and (b) in the main text for three different values of g N , while all other parameters are kept the same. The results are shown in Figure 10, where we note that Figure 10(c) and (d), corresponding to g N = 0.2 e V , is the same as Figure 4(a) and (b). Recall that for g N = 0.2 e V , we identified a threshold for dynamical SR for S ≈ 0.3, since beyond that value of S, the electronic coherence does not monotonically increase anymore. From Figure 10(b) and 10(f), we see that the so-defined SR threshold is at roughly S ≈ 0.1 for g N = 0.05 e V , whereas for g N = 0.4 e V , even for S = 0.8, the electronic coherence rises monotonically. We, therefore, conclude that the SR threshold emerges from a competition between light–matter coupling and vibronic coupling: The stronger the light–matter coupling, the larger the maximum value of S for which dynamical SR is still possible. This makes sense, since g N corresponds to the Rabi-oscillation frequency of the cavity-emitter system, i.e., larger values of g N result in a faster initial emission of photons, as seen in Figure 10(a), (c), and (e).

Figure 10: 
Mean-field solution of the average number of photons 



⟨



n

̂


⟩

/
N
=
|

⟨



a

̂


⟩



|


2


/
N


$\langle \hat{n}\rangle /N=\vert \langle \hat{a}\rangle {\vert }^{2}/N$



 and the electronic coherence 


|

⟨





σ

̂



+



⟩

|


$\vert \langle {\hat{\sigma }}^{+}\rangle \vert $



 in the HTC model for different values of S and 


g


N




$g\sqrt{N}$



. (a) and (b): 


g


N


=
0.05

e
V


$g\sqrt{N}=0.05\,\mathrm{e}\mathrm{V}$



, (c) and (d): 


g


N


=
0.2

e
V


$g\sqrt{N}=0.2\,\mathrm{e}\mathrm{V}$



 [same as Figure 4(a) and (b)], (e) and (f): 


g


N


=
0.4

e
V


$g\sqrt{N}=0.4\,\mathrm{e}\mathrm{V}$



. All other parameters are the same as in Figure 4(a)–(c).
Figure 10:

Mean-field solution of the average number of photons n ̂ / N = | a ̂ | 2 / N and the electronic coherence | σ ̂ + | in the HTC model for different values of S and  g N . (a) and (b): g N = 0.05 e V , (c) and (d): g N = 0.2 e V [same as Figure 4(a) and (b)], (e) and (f): g N = 0.4 e V . All other parameters are the same as in Figure 4(a)–(c).

References

[1] R. H. Dicke, “Coherence in spontaneous radiation processes,” Phys. Rev., vol. 93, no. 1, p. 99, 1954. https://doi.org/10.1103/physrev.93.99.Suche in Google Scholar

[2] M. Gross and S. Haroche, “Superradiance: An essay on the theory of collective spontaneous emission,” Phys. Rep., vol. 93, no. 5, p. 301, 1982. https://doi.org/10.1016/0370-1573(82)90102-8.Suche in Google Scholar

[3] K. Cong, Q. Zhang, Y. Wang, G. T. Noe, A. Belyanin, and J. Kono, “Dicke superradiance in solids [Invited],” JOSA B, vol. 33, no. 7, p. C80, 2016. https://doi.org/10.1364/josab.33.000c80.Suche in Google Scholar

[4] H. J. Carmichael and K. Kim, “A quantum trajectory unraveling of the superradiance master equation,” Opt. Commun., vol. 179, no. 1, p. 417, 2000. https://doi.org/10.1016/s0030-4018(99)00694-x.Suche in Google Scholar

[5] J. P. Clemens, L. Horvath, B. C. Sanders, and H. J. Carmichael, “Collective spontaneous emission from a line of atoms,” Phys. Rev. A, vol. 68, no. 2, p. 023809, 2003. https://doi.org/10.1103/physreva.68.023809.Suche in Google Scholar

[6] 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 Semiclassical Opt., vol. 6, no. 8, p. S736, 2004. https://doi.org/10.1088/1464-4266/6/8/017.Suche in Google Scholar

[7] M. O. Scully, E. S. Fry, C. H. R. Ooi, and K. Wódkiewicz, “Directed spontaneous emission from an extended ensemble of N atoms: Timing is everything,” Phys. Rev. Lett., vol. 96, no. 1, p. 010501, 2006. https://doi.org/10.1103/physrevlett.96.010501.Suche in Google Scholar PubMed

[8] H. Zoubi and H. Ritsch, “Metastability and directional emission characteristics of excitons in 1D optical lattices,” Europhys. Lett., vol. 90, no. 2, p. 23001, 2010. https://doi.org/10.1209/0295-5075/90/23001.Suche in Google Scholar

[9] D. Bhatti, J. von Zanthier, and G. S. Agarwal, “Superbunching and Nonclassicality as new Hallmarks of Superradiance,” Sci. Rep., vol. 5, no. 1, p. 17335, 2015. https://doi.org/10.1038/srep17335.Suche in Google Scholar PubMed PubMed Central

[10] E. Shahmoon, D. S. Wild, M. D. Lukin, and S. F. Yelin, “Cooperative resonances in light scattering from two-dimensional atomic arrays,” Phys. Rev. Lett., vol. 118, no. 11, p. 113601, 2017. https://doi.org/10.1103/physrevlett.118.113601.Suche in Google Scholar

[11] S. J. Masson, I. Ferrier-Barbut, L. A. Orozco, A. Browaeys, and A. Asenjo-Garcia, “Many-body signatures of collective decay in atomic chains,” Phys. Rev. Lett., vol. 125, no. 26, p. 263601, 2020. https://doi.org/10.1103/physrevlett.125.263601.Suche in Google Scholar PubMed

[12] S. J. Masson and A. Asenjo-Garcia, “Universality of Dicke superradiance in arrays of quantum emitters,” Nat. Commun., vol. 13, no. 1, p. 2285, 2022. https://doi.org/10.1038/s41467-022-29805-4.Suche in Google Scholar PubMed PubMed Central

[13] O. Rubies-Bigorda and S. F. Yelin, “Superradiance and subradiance in inverted atomic arrays,” Phys. Rev. A, vol. 106, no. 5, p. 053717, 2022. https://doi.org/10.1103/physreva.106.053717.Suche in Google Scholar

[14] O. Rubies-Bigorda, S. Ostermann, and S. F. Yelin, “Characterizing superradiant dynamics in atomic arrays via a cumulant expansion approach,” Phys. Rev. Res., vol. 5, no. 1, p. 013091, 2023. https://doi.org/10.1103/physrevresearch.5.013091.Suche in Google Scholar

[15] W.-K. Mok, A. Asenjo-Garcia, T. C. Sum, and L.-C. Kwek, “Dicke superradiance requires interactions beyond nearest neighbors,” Phys. Rev. Lett., vol. 130, no. 21, p. 213605, 2023. https://doi.org/10.1103/physrevlett.130.213605.Suche in Google Scholar PubMed

[16] S. J. Masson, J. P. Covey, S. Will, and A. Asenjo-Garcia, “Dicke superradiance in ordered arrays of multilevel atoms,” PRX Quantum, vol. 5, no. 1, p. 010344, 2024. https://doi.org/10.1103/prxquantum.5.010344.Suche in Google Scholar

[17] R. Holzinger and S. Yelin, “Collective superradiance: Estimating the peak emission rate and time,” 2025, arXiv:2504.09985 [quant-ph].Suche in Google Scholar

[18] R. Holzinger and S. F. Yelin, “Beyond Dicke superradiance: Universal scaling of the peak emission rate,” 2025, arXiv:2506.12649 [quant-ph].Suche in Google Scholar

[19] E. Abate and H. Haken, “Exakte Behandlung eines Maser-Modells,” Z. Naturforsch. A, vol. 19, no. 7–8, p. 857, 1964. https://doi.org/10.1515/zna-1964-7-806.Suche in Google Scholar

[20] R. Bonifacio and G. Preparata, “Coherent spontaneous emission,” Phys. Rev. A, vol. 2, no. 2, p. 336, 1970. https://doi.org/10.1103/physreva.2.336.Suche in Google Scholar

[21] S. M. Barnett and P. L. Knight, “Collective and quantum effects in models of rydberg atom maser evolution, optica acta,” Int. J. Opt., vol. 31, no. 4, pp. 435–452, 1984. https://doi.org/10.1080/713821515.Suche in Google Scholar

[22] J. Seke and F. Rattay, “Influence of the Fock-state field on the many-atom radiation processes in a damped cavity,” Phys. Rev. A, vol. 39, no. 1, p. 171, 1989. https://doi.org/10.1103/physreva.39.171.Suche in Google Scholar PubMed

[23] J. Seke, O. Hittmair, and F. Rattay, “Collapse and revival phenomena in the many-atom-jaynes-cummings model in the presence on initial fock- and coherent-state fields,” Opt. Commun., vol. 70, no. 4, p. 281, 1989. https://doi.org/10.1016/0030-4018(89)90320-9.Suche in Google Scholar

[24] P. R. Eastham, “The new physics of non-equilibrium condensates: Insights from classical dynamics,” J. Phys.: Condens. Matter, vol. 19, no. 29, p. 295210, 2007. https://doi.org/10.1088/0953-8984/19/29/295210.Suche in Google Scholar PubMed

[25] K. Hepp and E. H. Lieb, “On the superradiant phase transition for molecules in a quantized radiation field: The dicke maser model,” Ann. Phys., vol. 76, no. 2, p. 360, 1973. https://doi.org/10.1016/0003-4916(73)90039-0.Suche in Google Scholar

[26] Y. K. Wang and F. T. Hioe, “Phase transition in the Dicke model of superradiance,” Phys. Rev. A, vol. 7, no. 3, p. 831, 1973. https://doi.org/10.1103/physreva.7.831.Suche in Google Scholar

[27] K. Baumann, C. Guerlin, F. Brennecke, and T. Esslinger, “Dicke quantum phase transition with a superfluid gas in an optical cavity,” Nature, vol. 464, no. 7293, p. 1301, 2010. https://doi.org/10.1038/nature09009.Suche in Google Scholar PubMed

[28] P. Kirton, M. M. Roses, J. Keeling, and E. G. Dalla Torre, “Introduction to the Dicke model: From equilibrium to nonequilibrium, and vice versa,” Adv. Quantum Technol., vol. 2, no. 1–2, p. 1800043, 2019. https://doi.org/10.1002/qute.201800043.Suche in Google Scholar

[29] M. M. Roses and E. G. D. Torre, “Dicke model,” PLoS One, vol. 15, no. 9, p. e0235197, 2020. https://doi.org/10.1371/journal.pone.0235197.Suche in Google Scholar PubMed PubMed Central

[30] F. Haake, M. I. Kolobov, C. Fabre, E. Giacobino, and S. Reynaud, “Superradiant laser,” Phys. Rev. Lett., vol. 71, no. 7, p. 995, 1993. https://doi.org/10.1103/physrevlett.71.995.Suche in Google Scholar

[31] D. Meiser, J. Ye, D. R. Carlson, and M. J. Holland, “Prospects for a millihertz-linewidth laser,” Phys. Rev. Lett., vol. 102, no. 16, p. 163601, 2009. https://doi.org/10.1103/physrevlett.102.163601.Suche in Google Scholar

[32] J. G. Bohnet, Z. Chen, J. M. Weiner, D. Meiser, M. J. Holland, and J. K. Thompson, “A steady-state superradiant laser with less than one intracavity photon,” Nature, vol. 484, no. 7392, p. 78, 2012. https://doi.org/10.1038/nature10920.Suche in Google Scholar PubMed

[33] M. Tavis and F. W. Cummings, “Exact solution for an N-Molecule–Radiation-Field hamiltonian,” Phys. Rev., vol. 170, no. 2, p. 379, 1968. https://doi.org/10.1103/physrev.170.379.Suche in Google Scholar

[34] J. Keeling, “Quantum corrections to the semiclassical collective dynamics in the Tavis-Cummings model,” Phys. Rev. A, vol. 79, no. 5, p. 053825, 2009. https://doi.org/10.1103/physreva.79.053825.Suche in Google Scholar

[35] O. Babelon, L. Cantini, and B. Douçot, “A semi-classical study of the Jaynes–Cummings model,” J. Stat. Mech.: Theor. Exp., vol. 2009, no. 7, p. P07011, 2009. https://doi.org/10.1088/1742-5468/2009/07/p07011.Suche in Google Scholar

[36] A. V. Andreev, V. Gurarie, and L. Radzihovsky, “Nonequilibrium dynamics and thermodynamics of a degenerate fermi gas across a feshbach resonance,” Phys. Rev. Lett., vol. 93, no. 13, p. 130402, 2004. https://doi.org/10.1103/physrevlett.93.130402.Suche in Google Scholar PubMed

[37] R. A. Barankov and L. S. Levitov, “Atom-molecule coexistence and collective dynamics near a feshbach resonance of cold fermions,” Phys. Rev. Lett., vol. 93, no. 13, p. 130403, 2004. https://doi.org/10.1103/physrevlett.93.130403.Suche in Google Scholar PubMed

[38] M. H. Szymańska, B. D. Simons, and K. Burnett, “Dynamics of the BCS-BEC crossover in a degenerate fermi gas,” Phys. Rev. Lett., vol. 94, no. 17, p. 170402, 2005. https://doi.org/10.1103/physrevlett.94.170402.Suche in Google Scholar PubMed

[39] E. A. Yuzbashyan, V. B. Kuznetsov, and B. L. Altshuler, “Integrable dynamics of coupled Fermi-Bose condensates,” Phys. Rev. B, vol. 72, no. 14, p. 144524, 2005. https://doi.org/10.1103/physrevb.72.144524.Suche in Google Scholar

[40] P. R. Eastham and R. T. Phillips, “Quantum condensation from a tailored exciton population in a microcavity,” Phys. Rev. B, vol. 79, no. 16, p. 165303, 2009. https://doi.org/10.1103/physrevb.79.165303.Suche in Google Scholar

[41] B. C. Rose, et al.., “Coherent rabi dynamics of a superradiant spin ensemble in a microwave cavity,” Phys. Rev. X, vol. 7, no. 3, p. 031002, 2017. https://doi.org/10.1103/physrevx.7.031002.Suche in Google Scholar

[42] P. Kling, E. Giese, C. M. Carmesin, R. Sauerbrey, and W. P. Schleich, “High-gain quantum free-electron laser: Long-time dynamics and requirements,” Phys. Rev. Res., vol. 3, no. 3, p. 033232, 2021. https://doi.org/10.1103/physrevresearch.3.033232.Suche in Google Scholar

[43] C. Sträter, O. Tsyplyatyev, and A. Faribault, “Nonequilibrum dynamics in the strongly excited inhomogeneous Dicke model,” Phys. Rev. B, vol. 86, no. 19, p. 195101, 2012. https://doi.org/10.1103/physrevb.86.195101.Suche in Google Scholar

[44] K. Jürgens, F. Lengers, D. Groll, D. E. Reiter, D. Wigger, and T. Kuhn, “Comparison of the semiclassical and quantum optical field dynamics in a pulse-excited optical cavity with a finite number of quantum emitters,” Phys. Rev. B, vol. 104, no. 20, p. 205308, 2021. https://doi.org/10.1103/physrevb.104.205308.Suche in Google Scholar

[45] O. Tsyplyatyev and D. Loss, “Dynamics of the inhomogeneous Dicke model for a single-boson mode coupled to a bath of nonidentical spin-1/2 systems,” Phys. Rev. A, vol. 80, no. 2, p. 023803, 2009. https://doi.org/10.1103/physreva.80.023803.Suche in Google Scholar

[46] O. Tsyplyatyev and D. Loss, “Classical and quantum regimes of the inhomogeneous Dicke model and its Ehrenfest time,” Phys. Rev. B, vol. 82, no. 2, p. 024305, 2010. https://doi.org/10.1103/physrevb.82.024305.Suche in Google Scholar

[47] J. A. Ćwik, S. Reja, P. B. Littlewood, and J. Keeling, “Polariton condensation with saturable molecules dressed by vibrational modes,” Europhys. Lett., vol. 105, no. 4, p. 47009, 2014. https://doi.org/10.1209/0295-5075/105/47009.Suche in Google Scholar

[48] F. Herrera and F. C. Spano, “Theory of nanoscale organic cavities: The essential role of vibration-photon dressed states,” ACS Photonics, vol. 5, no. 1, p. 65, 2018. https://doi.org/10.1021/acsphotonics.7b00728.Suche in Google Scholar

[49] D. Wellnitz, G. Pupillo, and J. Schachenmayer, “Disorder enhanced vibrational entanglement and dynamics in polaritonic chemistry,” Commun. Phys., vol. 5, no. 1, p. 120, 2022. https://doi.org/10.1038/s42005-022-00892-5.Suche in Google Scholar

[50] D. Wang, et al.., “Turning a molecule into a coherent two-level quantum system,” Nat. Phys., vol. 15, no. 5, p. 483, 2019. https://doi.org/10.1038/s41567-019-0436-5.Suche in Google Scholar

[51] P. Törmä and W. L. Barnes, “Strong coupling between surface plasmon polaritons and emitters: A review,” Rep. Prog. Phys., vol. 78, no. 1, p. 013901, 2014. https://doi.org/10.1088/0034-4885/78/1/013901.Suche in Google Scholar PubMed

[52] T. K. Hakala, et al.., “Bose–Einstein condensation in a plasmonic lattice,” Nat. Phys., vol. 14, no. 7, p. 739, 2018. https://doi.org/10.1038/s41567-018-0109-9.Suche in Google Scholar

[53] M. De Giorgi, et al.., “Interaction and coherence of a plasmon–exciton polariton condensate,” ACS Photonics, vol. 5, no. 9, p. 3666, 2018. https://doi.org/10.1021/acsphotonics.8b00662.Suche in Google Scholar

[54] J. Keeling and S. Kéna-Cohen, “Bose–einstein condensation of exciton-polaritons in organic microcavities,” Annu. Rev. Phys. Chem., vol. 71, no. 1, p. 435, 2020. https://doi.org/10.1146/annurev-physchem-010920-102509.Suche in Google Scholar PubMed

[55] A. I. Väkeväinen, A. J. Moilanen, M. Nečada, T. K. Hakala, K. S. Daskalakis, and P. Törmä, “Sub-picosecond thermalization dynamics in condensation of strongly coupled lattice plasmons,” Nat. Commun., vol. 11, no. 1, p. 3139, 2020. https://doi.org/10.1038/s41467-020-16906-1.Suche in Google Scholar PubMed PubMed Central

[56] J. Roden, W. T. Strunz, K. B. Whaley, and A. Eisfeld, “Accounting for intra-molecular vibrational modes in open quantum system description of molecular systems,” J. Chem. Phys., vol. 137, no. 20, p. 204110, 2012. https://doi.org/10.1063/1.4765329.Suche in Google Scholar PubMed

[57] F. C. Spano, “Optical microcavities enhance the exciton coherence length and eliminate vibronic coupling in J-aggregates,” J. Chem. Phys., vol. 142, no. 18, p. 184707, 2015. https://doi.org/10.1063/1.4919348.Suche in Google Scholar PubMed

[58] F. Herrera and F. C. Spano, “Cavity-controlled chemistry in molecular ensembles,” Phys. Rev. Lett., vol. 116, no. 23, p. 238301, 2016. https://doi.org/10.1103/physrevlett.116.238301.Suche in Google Scholar PubMed

[59] N. Wu, J. Feist, and F. J. Garcia-Vidal, “When polarons meet polaritons: Exciton-vibration interactions in organic molecules strongly coupled to confined light fields,” Phys. Rev. B, vol. 94, no. 19, p. 195409, 2016. https://doi.org/10.1103/physrevb.94.195409.Suche in Google Scholar

[60] M. A. Zeb, P. G. Kirton, and J. Keeling, “Exact states and spectra of vibrationally dressed polaritons,” ACS Photonics, vol. 5, no. 1, p. 249, 2018. https://doi.org/10.1021/acsphotonics.7b00916.Suche in Google Scholar

[61] B. A. Chase and J. M. Geremia, “Collective processes of an ensemble of spin-1/2 particles,” Phys. Rev. A, vol. 78, no. 5, p. 052101, 2008. https://doi.org/10.1103/physreva.78.052101.Suche in Google Scholar

[62] M. Xu, D. A. Tieri, and M. J. Holland, “Simulating open quantum systems by applying SU(4) to quantum master equations,” Phys. Rev. A, vol. 87, no. 6, p. 062101, 2013. https://doi.org/10.1103/physreva.87.062101.Suche in Google Scholar

[63] M. Richter, M. Gegg, T. S. Theuerholz, and A. Knorr, “Numerically exact solution of the many emitter–cavity laser problem: Application to the fully quantized spaser emission,” Phys. Rev. B, vol. 91, no. 3, p. 035306, 2015. https://doi.org/10.1103/physrevb.91.035306.Suche in Google Scholar

[64] F. Damanet, D. Braun, and J. Martin, “Cooperative spontaneous emission from indistinguishable atoms in arbitrary motional quantum states,” Phys. Rev. A, vol. 94, no. 3, p. 033838, 2016. https://doi.org/10.1103/physreva.94.033838.Suche in Google Scholar

[65] P. Kirton and J. Keeling, “Suppressing and restoring the Dicke superradiance transition by dephasing and decay,” Phys. Rev. Lett., vol. 118, no. 12, p. 123602, 2017. https://doi.org/10.1103/physrevlett.118.123602.Suche in Google Scholar PubMed

[66] 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, vol. 98, no. 6, p. 063815, 2018. https://doi.org/10.1103/physreva.98.063815.Suche in Google Scholar

[67] D. Barberena, “Generalized holstein-primakoff mapping and 1/n expansion of collective spin systems undergoing single particle dissipation,” 2025, arXiv:2508.05751 [quant-ph].Suche in Google Scholar

[68] Z.-X. Gong et al.., “Steady-state superradiance with Rydberg polaritons,” 2016. arXiv:1611.00797 [quant-ph].Suche in Google Scholar

[69] P. R. Eastham and P. B. Littlewood, “Bose condensation of cavity polaritons beyond the linear regime: The thermal equilibrium of a model microcavity,” Phys. Rev. B, vol. 64, no. 23, p. 235101, 2001. https://doi.org/10.1103/physrevb.64.235101.Suche in Google Scholar

[70] T. Mori, “Exactness of the mean-field dynamics in optical cavity systems,” J. Stat. Mech.: Theor. Exp., vol. 2013, no. 6, p. P06005, 2013. https://doi.org/10.1088/1742-5468/2013/06/p06005.Suche in Google Scholar

[71] F. Carollo and I. Lesanovsky, “Exactness of mean-field equations for open dicke models with an application to pattern retrieval dynamics,” Phys. Rev. Lett., vol. 126, no. 23, p. 230601, 2021. https://doi.org/10.1103/physrevlett.126.230601.Suche in Google Scholar

[72] F. Robicheaux and D. A. Suresh, “Beyond lowest order mean-field theory for light interacting with atom arrays,” Phys. Rev. A, vol. 104, no. 2, p. 023702, 2021. https://doi.org/10.1103/physreva.104.023702.Suche in Google Scholar

[73] P. Fowler-Wright, K. B. Arnardóttir, P. Kirton, B. W. Lovett, and J. Keeling, “Determining the validity of cumulant expansions for central spin models,” Phys. Rev. Res., vol. 5, no. 3, p. 033148, 2023. https://doi.org/10.1103/physrevresearch.5.033148.Suche in Google Scholar

[74] K. C. Stitely, et al.., “Quantum fluctuation dynamics of dispersive superradiant pulses in a hybrid light-matter system,” Phys. Rev. Lett., vol. 131, no. 14, p. 143604, 2023. https://doi.org/10.1103/physrevlett.131.143604.Suche in Google Scholar PubMed

[75] F. Carollo, “Non-Gaussian dynamics of quantum fluctuations and mean-field limit in open quantum central spin systems,” Phys. Rev. Lett., vol. 131, no. 22, p. 227102, 2023. https://doi.org/10.1103/physrevlett.131.227102.Suche in Google Scholar

[76] K. B. Arnardottir, A. J. Moilanen, A. Strashko, P. Törmä, and J. Keeling, “Multimode organic polariton lasing,” Phys. Rev. Lett., vol. 125, no. 23, p. 233603, 2020. https://doi.org/10.1103/physrevlett.125.233603.Suche in Google Scholar PubMed

[77] A. J. Moilanen, K. B. Arnardóttir, J. Keeling, and P. Törmä, “Mode switching dynamics in organic polariton lasing,” Phys. Rev. B, vol. 106, no. 19, p. 195403, 2022. https://doi.org/10.1103/physrevb.106.195403.Suche in Google Scholar

[78] P. Fowler-Wright, B. W. Lovett, and J. Keeling, “Efficient many-body non-Markovian dynamics of organic polaritons,” Phys. Rev. Lett., vol. 129, no. 17, p. 173001, 2022. https://doi.org/10.1103/physrevlett.129.173001.Suche in Google Scholar

[79] H.-P. Breuer and F. Petruccione, The Theory of Open Quantum Systems, New York, Oxford University Press, 2002.10.1007/3-540-44874-8_4Suche in Google Scholar

[80] J. M. Radcliffe, “Some properties of coherent spin states,” J. Phys. A: Gen. Phys., vol. 4, no. 3, p. 313, 1971. https://doi.org/10.1088/0305-4470/4/3/009.Suche in Google Scholar

[81] L. Mandel and E. Wolf, Optical Coherence and Quantum Optics, New York, Cambridge University Press, 1995.10.1017/CBO9781139644105Suche in Google Scholar

[82] P. Rosario, L. O. R. Solak, A. Cidrim, R. Bachelard, and J. Schachenmayer, “Unraveling Dicke superradiant decay with separable coherent spin states,” 2025, arXiv:2504.13418 [quant-ph].10.1103/xcxr-sm9cSuche in Google Scholar PubMed

[83] B. Buča and T. Prosen, “A note on symmetry reductions of the Lindblad equation: Transport in constrained open spin chains,” New J. Phys., vol. 14, no. 7, p. 073007, 2012. https://doi.org/10.1088/1367-2630/14/7/073007.Suche in Google Scholar

[84] V. V. Albert and L. Jiang, “Symmetries and conserved quantities in Lindblad master equations,” Phys. Rev. A, vol. 89, no. 2, p. 022118, 2014. https://doi.org/10.1103/physreva.89.022118.Suche in Google Scholar

[85] R. Bonifacio, P. Schwendimann, and F. Haake, “Quantum statistical theory of superradiance. II,” Phys. Rev. A, vol. 4, no. 3, p. 854, 1971. https://doi.org/10.1103/physreva.4.854.Suche in Google Scholar

[86] L. Freter and P. Fowler-Wright, Lukasfreter/pibs: Version 0.1.0, Zenodo, 2025.Suche in Google Scholar

[87] C. Gardiner, Stochastic Methods: A Handbook for the Natural and Social Sciences, Heidelberg, Springer Berlin Heidelberg, 2009.Suche in Google Scholar

[88] D. Plankensteiner, C. Hotter, and H. Ritsch, “QuantumCumulants.jl: A Julia framework for generalized mean-field equations in open quantum systems,” Quantum, vol. 6, p. 617, 2022, https://doi.org/10.22331/q-2022-01-04-617.Suche in Google Scholar

[89] We note that since the dominant vibrational mode is included in the system description in the HTC model, the dephasing terms in Eq. (10) and Eq. (2) have different meanings in principle. However, since we only set γϕ > 0 in case S = 0, the two dephasing rates are equivalent.Suche in Google Scholar

[90] J. R. Mannouch, W. Barford, and S. Al-Assam, “Ultra-fast relaxation, decoherence, and localization of photoexcited states in π-conjugated polymers,” J. Chem. Phys., vol. 148, no. 3, p. 034901, 2018. https://doi.org/10.1063/1.5009393.Suche in Google Scholar PubMed

[91] R. Friedberg, S. R. Hartmann, and J. T. Manassah, “Limited superradiant damping of small samples,” Phys. Lett. A, vol. 40, no. 5, p. 365, 1972. https://doi.org/10.1016/0375-9601(72)90533-6.Suche in Google Scholar

[92] R. Friedberg and S. R. Hartmann, “Temporal evolution of superradiance in a small sphere,” Phys. Rev. A, vol. 10, no. 5, p. 1728, 1974. https://doi.org/10.1103/physreva.10.1728.Suche in Google Scholar

[93] B. Coffey and R. Friedberg, “Effect of short-range Coulomb interaction on cooperative spontaneous emission,” Phys. Rev. A, vol. 17, no. 3, p. 1033, 1978. https://doi.org/10.1103/physreva.17.1033.Suche in Google Scholar

Received: 2025-09-03
Accepted: 2025-11-24
Published Online: 2025-12-08

© 2025 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Artikel in diesem Heft

  1. Frontmatter
  2. Reviews
  3. Light-driven micro/nanobots
  4. Tunable BIC metamaterials with Dirac semimetals
  5. Large-scale silicon photonics switches for AI/ML interconnections based on a 300-mm CMOS pilot line
  6. Perspective
  7. Density-functional tight binding meets Maxwell: unraveling the mysteries of (strong) light–matter coupling efficiently
  8. Letters
  9. Broadband on-chip spectral sensing via directly integrated narrowband plasmonic filters for computational multispectral imaging
  10. Sub-100 nm manipulation of blue light over a large field of view using Si nanolens array
  11. Tunable bound states in the continuum through hybridization of 1D and 2D metasurfaces
  12. Integrated array of coupled exciton–polariton condensates
  13. Disentangling the absorption lineshape of methylene blue for nanocavity strong coupling
  14. Research Articles
  15. Demonstration of multiple-wavelength-band photonic integrated circuits using a silicon and silicon nitride 2.5D integration method
  16. Inverse-designed gyrotropic scatterers for non-reciprocal analog computing
  17. Highly sensitive broadband photodetector based on PtSe2 photothermal effect and fiber harmonic Vernier effect
  18. Online training and pruning of multi-wavelength photonic neural networks
  19. Robust transport of high-speed data in a topological valley Hall insulator
  20. Engineering super- and sub-radiant hybrid plasmons in a tunable graphene frame-heptamer metasurface
  21. Near-unity fueling light into a single plasmonic nanocavity
  22. Polarization-dependent gain characterization in x-cut LNOI erbium-doped waveguide amplifiers
  23. Intramodal stimulated Brillouin scattering in suspended AlN waveguides
  24. Single-shot Stokes polarimetry of plasmon-coupled single-molecule fluorescence
  25. Metastructure-enabled scalable multiple mode-order converters: conceptual design and demonstration in direct-access add/drop multiplexing systems
  26. High-sensitivity U-shaped biosensor for rabbit IgG detection based on PDA/AuNPs/PDA sandwich structure
  27. Deep-learning-based polarization-dependent switching metasurface in dual-band for optical communication
  28. A nonlocal metasurface for optical edge detection in the far-field
  29. Coexistence of weak and strong coupling in a photonic molecule through dissipative coupling to a quantum dot
  30. Mitigate the variation of energy band gap with electric field induced by quantum confinement Stark effect via a gradient quantum system for frequency-stable laser diodes
  31. Orthogonal canalized polaritons via coupling graphene plasmon and phonon polaritons of hBN metasurface
  32. Dual-polarization electromagnetic window simultaneously with extreme in-band angle-stability and out-of-band RCS reduction empowered by flip-coding metasurface
  33. Record-level, exceptionally broadband borophene-based absorber with near-perfect absorption: design and comparison with a graphene-based counterpart
  34. Generalized non-Hermitian Hamiltonian for guided resonances in photonic crystal slabs
  35. A 10× continuously zoomable metalens system with super-wide field of view and near-diffraction–limited resolution
  36. Continuously tunable broadband adiabatic coupler for programmable photonic processors
  37. Diffraction order-engineered polarization-dependent silicon nano-antennas metagrating for compact subtissue Mueller microscopy
  38. Lithography-free subwavelength metacoatings for high thermal radiation background camouflage empowered by deep neural network
  39. Multicolor nanoring arrays with uniform and decoupled scattering for augmented reality displays
  40. Permittivity-asymmetric qBIC metasurfaces for refractive index sensing
  41. Theory of dynamical superradiance in organic materials
  42. Second-harmonic generation in NbOI2-integrated silicon nitride microdisk resonators
  43. A comprehensive study of plasmonic mode hybridization in gold nanoparticle-over-mirror (NPoM) arrays
  44. Foundry-enabled wafer-scale characterization and modeling of silicon photonic DWDM links
  45. Rough Fabry–Perot cavity: a vastly multi-scale numerical problem
  46. Classification of quantum-spin-hall topological phase in 2D photonic continuous media using electromagnetic parameters
  47. Light-guided spectral sculpting in chiral azobenzene-doped cholesteric liquid crystals for reconfigurable narrowband unpolarized light sources
  48. Modelling Purcell enhancement of metasurfaces supporting quasi-bound states in the continuum
  49. Ultranarrow polaritonic cavities formed by one-dimensional junctions of two-dimensional in-plane heterostructures
  50. Bridging the scalability gap in van der Waals light guiding with high refractive index MoTe2
  51. Ultrafast optical modulation of vibrational strong coupling in ReCl(CO)3(2,2-bipyridine)
  52. Chirality-driven all-optical image differentiation
  53. Wafer-scale CMOS foundry silicon-on-insulator devices for integrated temporal pulse compression
  54. Monolithic temperature-insensitive high-Q Ta2O5 microdisk resonator
  55. Nanogap-enhanced terahertz suppression of superconductivity
  56. Large-gap cascaded Moiré metasurfaces enabling switchable bright-field and phase-contrast imaging compatible with coherent and incoherent light
  57. Synergistic enhancement of magneto-optical response in cobalt-based metasurfaces via plasmonic, lattice, and cavity modes
  58. Scalable unitary computing using time-parallelized photonic lattices
  59. Diffusion model-based inverse design of photonic crystals for customized refraction
  60. Wafer-scale integration of photonic integrated circuits and atomic vapor cells
  61. Optical see-through augmented reality via inverse-designed waveguide couplers
  62. One-dimensional dielectric grating structure for plasmonic coupling and routing
  63. MCP-enabled LLM for meta-optics inverse design: leveraging differentiable solver without LLM expertise
  64. Broadband variable beamsplitter made of a subwavelength-thick metamaterial
  65. Scaling-dependent tunability of spin-driven photocurrents in magnetic metamaterials
  66. AI-based analysis algorithm incorporating nanoscale structural variations and measurement-angle misalignment in spectroscopic ellipsometry
Heruntergeladen am 22.1.2026 von https://www.degruyterbrill.com/document/doi/10.1515/nanoph-2025-0427/html
Button zum nach oben scrollen