Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
License: arXiv.org perpetual non-exclusive license
arXiv:2211.01320v2 [cond-mat.other] 20 Dec 2023

Robustness of the Floquet-assisted superradiant phase and possible laser operation

Lukas Broers Center for Optical Quantum Technologies, University of Hamburg, Hamburg, Germany Institute for Quantum Physics, University of Hamburg, Hamburg, Germany    Ludwig Mathey Center for Optical Quantum Technologies, University of Hamburg, Hamburg, Germany Institute for Quantum Physics, University of Hamburg, Hamburg, Germany The Hamburg Center for Ultrafast Imaging, Hamburg, Germany
Abstract

We demonstrate the robustness of the recently established Floquet-assisted superradiant phase of the parametrically driven dissipative Dicke model, inspired by light-induced dynamics in graphene. In particular, we show the robustness of this state against key imperfections and argue for the feasibility of utilizing it for laser operation. We consider the effect of a finite linewidth of the driving field, modelled via phase diffusion. We find that the linewidth of the light field in the cavity narrows drastically across the FSP transition, reminiscent of a line narrowing at the laser transition. We then demonstrate that the FSP is robust against inhomogeneous broadening, while displaying a reduction of light intensity. We show that the depleted population inversion of near-resonant Floquet states leads to hole burning in the inhomogeneously broadened Floquet spectra. Finally, we show that the FSP is robust against dissipation processes, with coefficients up to values that are experimentally available. We conclude that the FSP presents a robust mechanism that is capable of realistic laser operation.

I Introduction

In the superradiant phase transition of the Dicke model Hepp and Lieb (1973); Wang and Hioe (1973) the ground state of a set of identical two-level systems (TLS) that are coupled to a cavity is accompanied by symmetry breaking and the emergence of a coherent photon state. Realizations of the Dicke model, and consequently the superradiant phase transition, have been proposed Domokos and Ritsch (2002); Dimer et al. (2007); Nagy et al. (2010) and demonstrated experimentally Black et al. (2003); Zhiqiang et al. (2017); Baumann et al. (2010); Klinder et al. (2015). The realization of the Dicke model in cavity-BEC setups leads to intricate non-equilibrium superradiant phases, which can appear in the presence of parametric driving of the coupling parameter Bastidas et al. (2012); Chitra and Zilberberg (2015); Gong et al. (2018); Cosme et al. (2018); Zhu et al. (2019); Keßler et al. (2021); Tuquero et al. (2022); Cosme et al. (2019); Skulte et al. (2021); Kongkhambut et al. (2021); Skulte et al. (2022). Meanwhile, generalizations of the Dicke model have been studied to find rich phase diagrams that display superradiant phases, regular lasing and the unconventional counter-lasing Hioe (1973); Keeling et al. (2010); Bhaseen et al. (2012); Baksic and Ciuti (2014); Genway et al. (2014); Soriente et al. (2018); Gutiérrez-Jáuregui and Carmichael (2018); Chelpanova et al. (2021); Stitely et al. (2022). These types of Dicke models are also referred to as driven, due to the tunability of the atom-photon processes Kirton et al. (2019); Damanet et al. (2019); Kirton and Keeling (2018).

Refer to caption
Figure 1: Mechanism of the Floquet-assisted superradiant phase. (a) The bare energy levels ωzsubscript𝜔𝑧\omega_{z}italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT of a collection of two-level systems are dressed via strong driving field strengths Edsubscript𝐸dE_{\mathrm{d}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT and deformed into Floquet states. Depending on the details of driving and dissipation, this process leads to population inversion in the Floquet states as indicated by the gray hatched areas. At Floquet energies that are resonant with the cavity frequency ωcsubscript𝜔𝑐\omega_{c}italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, this population inversion is depleted and transferred into a coherent state in the cavity. (b) The phase diagram of the light field in the cavity shows the trivial phase (TP), the Dicke superradiant phase (DS) for small driving field strengths and the Floquet-assisted superradiant phase (FSP) for large driving field strengths around Edonsetsuperscriptsubscript𝐸donsetE_{\mathrm{d}}^{\mathrm{onset}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_onset end_POSTSUPERSCRIPT. The coupling strength λcFSPsuperscriptsubscript𝜆𝑐FSP\lambda_{c}^{\mathrm{FSP}}italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_FSP end_POSTSUPERSCRIPT at which the FSP emerges depends on the cavity loss rate κ𝜅\kappaitalic_κ that increases with κ𝜅\kappaitalic_κ.

The first studies on the superradiant phase were preceeded by studies on the closely related phenomenon of superradiance, introduced in the seminal work by Dicke Dicke (1954). Unlike the superradiant phase, superradiance is a transient process of collective coherent spontaneous emission that can be engineered into continuous operation which results in superradiant lasing with ultranarrow linewidths Haake et al. (1993); Kuppens et al. (1994); Meiser et al. (2009); Bohnet et al. (2012); Shi et al. (2019); Laske et al. (2019); Tang et al. (2021); Meiser and Holland (2010); Norcia and Thompson (2016); Norcia et al. (2016); Weiner et al. (2017); Debnath et al. (2018); Jäger et al. (2019); Liu et al. (2020); Zhang et al. (2021); Wu et al. (2021). The mechanisms behind regular lasing as well as superradiant lasing rely on incoherent driving which is realized via pumping into higher levels in order to create population inversion. Superradiance has also been studied in the presence of coherent driving of the two-level systems Mollow (1972); Wu et al. (1977); Narducci et al. (1978); Walls et al. (1978); Walls (1980). This lead to studies on two-photon dressed-state lasers Lezama et al. (1990); Lewenstein et al. (1990); Lu and Berman (1991); Gauthier et al. (1992); Zakrzewski et al. (1991a, b, c); Horak et al. (1995); Gheri et al. (1995) in particular in quantum dot system Muller et al. (2007); Hauss et al. (2008); Neilinger et al. (2015); Chestnov et al. (2017), as a type of lasing without inversion Scully et al. (1989); Fleischhauer et al. (1992); Zhu (1992); Mompart and Corbalán (2000). Superradiance has also been studied in solid-state systems Cong et al. (2016); Bradac et al. (2017); Breeze et al. (2018); Angerer et al. (2018). This relates to Floquet-engineering, which explores the possibilities of dynamically controlling properties such as band populations and topology Oka and Aoki (2009); Seetharam et al. (2015); Desbuquois et al. (2017); Basov et al. (2017); Jiang et al. (2021); Peano and Thorwart (2010a, b); Stehlik et al. (2016); Oka and Kitamura (2019); Zhang et al. (2017); Johansen et al. (2022); Rudner and Lindner (2020).

Generally speaking, one crucial feature of both regular and unconventional lasing is the generation of monochromatic scalable coherent emission that displays line narrowing below the intrinsic linewidth. Further, practical laser operation is expected to be stable in the presence of environmental factors. Such factors include inhomogeneous broadening which in solid-state systems occurs due to material defects, while in gaseous setups it occurs due to the velocity distribution of the atoms. This leads to spectral linewidths that exceed intrinsic linewidths due to shifted energy levels that would be degenerate in the absence of inhomogeneous broadening.

In recent work we have presented the Floquet-assisted superradiant phase (FSP) Broers and Mathey (2023) in a parametrically driven Dicke model in the presence of solid-like dissipation. The underlying mechanism of the FSP consists of strong coherent driving of the TLSs which leads to Floquet states with energies that deviate from the bare transitions. These Floquet states can be tuned into resonance with a cavity while they simulatenously experience effective population inversion due to the interplay of coherent driving and dissipation. This model is motivated by the demonstration of negative optical conductivities as a consequence of population inverted Floquet states in coherently driven graphene Broers and Mathey (2021). Coupling the TLSs to the cavity results in the population inversion being depleted in order to sustain the oscillating coherent state in the cavity. The mechanism of the FSP is illustrated in Fig. 1. The regime in which inversion occurs as calculated in previous work Broers and Mathey (2023) is indicated with grey hatched lines.

In this paper, we demonstrate the robustness of the FSP. We introduce a phase diffusion process and hence a finite linewidth in the driving field. We show that the transition into the FSP is accompanied by significant narrowing of the linewidth of the cavity light field such that the emergence of the FSP is stable in the presence of finite phase coherence of the driving field. Further, we demonstrate that the FSP is stable in the presence of inhomogeneous broadening. Finally, we show that the FSP is robust with respect to the dissipation of the TLSs up to decay rates of the order of those in recent light-driven graphene experiments. We then identify the cavity loss rate as the most sensitive parameter, as the FSP vanishes comparatively rapidly as a function of cavity losses. Overall, our results suggest the possibility of laser operation based on the FSP in a solid-state system, such as a two-band material, due to the form of the dissipative processes, the magnitude of the dissipation that we consider, the robustness against inhomogeneous broadening, and the strong line narrowing across the transition. This type of laser operation is distinct from regular lasing and superradiant lasing, but comparable to a modified type of dressed-state lasing in solids.

This work is structured as follows. In section II we describe the master equation of the parametrically driven dissipative Dicke model. In section III we introduce phase diffusion into the driving term of the TLS which leads to a broadened linewidth, that is overcome by the drastic line narrowing of the light field in the cavity. In section IV we introduce inhomogeneous broadening into the TLSs which modifies the transition, as not all TLSs participate in the FSP. In section V we show the FSP transition as a function of the cavity loss rate as well as the TLS dissipation rates to show the robustness of the FSP. In section VI we conclude our results and present an outlook for possible implementations of the FSP.

II Parametrically Driven Dicke Model

We consider a dissipative Dicke model that is parametrically driven with circularly polarized light. The Hamiltonian of this system is

1H1Planck-constant-over-2-pi𝐻\displaystyle\frac{1}{\hbar}Hdivide start_ARG 1 end_ARG start_ARG roman_ℏ end_ARG italic_H =j=1Nωzj2σzj+σ+jA(t)+σjA+(t)absentsuperscriptsubscript𝑗1𝑁superscriptsubscript𝜔𝑧𝑗2subscriptsuperscript𝜎𝑗𝑧superscriptsubscript𝜎𝑗subscript𝐴𝑡superscriptsubscript𝜎𝑗subscript𝐴𝑡\displaystyle=\sum_{j=1}^{N}\frac{\omega_{z}^{j}}{2}\sigma^{j}_{z}+\sigma_{+}^% {j}A_{-}(t)+\sigma_{-}^{j}A_{+}(t)= ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT divide start_ARG italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG italic_σ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + italic_σ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT - end_POSTSUBSCRIPT ( italic_t ) + italic_σ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( italic_t ) (1)
+ωcaa+λNj=1Nσxj(a+a)subscript𝜔𝑐superscript𝑎𝑎𝜆𝑁superscriptsubscript𝑗1𝑁subscriptsuperscript𝜎𝑗𝑥𝑎superscript𝑎\displaystyle+\omega_{c}a^{\dagger}a+\frac{\lambda}{\sqrt{N}}\sum_{j=1}^{N}% \sigma^{j}_{x}(a+a^{\dagger})+ italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a + divide start_ARG italic_λ end_ARG start_ARG square-root start_ARG italic_N end_ARG end_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_σ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT ( italic_a + italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT )

with the driving field

A±(t)=Edωdexp{±iωdt}.subscript𝐴plus-or-minus𝑡subscript𝐸dsubscript𝜔dplus-or-minus𝑖subscript𝜔d𝑡A_{\pm}(t)=\frac{E_{\mathrm{d}}}{\omega_{\mathrm{d}}}\exp\{\pm i\omega_{% \mathrm{d}}t\}.italic_A start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ( italic_t ) = divide start_ARG italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT end_ARG roman_exp { ± italic_i italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT italic_t } . (2)

ωdsubscript𝜔d\omega_{\mathrm{d}}italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT is the driving frequency and Edsubscript𝐸dE_{\mathrm{d}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT is the effective driving strength that takes the dimension of frequency squared. Edsubscript𝐸dE_{\mathrm{d}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT can be related to electric field strengths of driving terms in other systems that motivate the general form of this Hamiltonian. ωcsubscript𝜔𝑐\omega_{c}italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is the cavity frequency. The frequencies ωzjsuperscriptsubscript𝜔𝑧𝑗\omega_{z}^{j}italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT are frequencies of the TLSs. We first choose these to be equal, i.e. ωzj=ωzsuperscriptsubscript𝜔𝑧𝑗subscript𝜔𝑧\omega_{z}^{j}=\omega_{z}italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT = italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT, and later consider a distribution of frequencies ωzjsuperscriptsubscript𝜔𝑧𝑗\omega_{z}^{j}italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT, to model inhomogeneous broadening. σkjsuperscriptsubscript𝜎𝑘𝑗\sigma_{k}^{j}italic_σ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT is the k𝑘kitalic_kth Pauli-matrix acting on the j𝑗jitalic_jth TLS, where σ±j=(σxj±iσyj)/2superscriptsubscript𝜎plus-or-minus𝑗plus-or-minussuperscriptsubscript𝜎𝑥𝑗𝑖superscriptsubscript𝜎𝑦𝑗2\sigma_{\pm}^{j}=(\sigma_{x}^{j}\pm i\sigma_{y}^{j})/2italic_σ start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT = ( italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ± italic_i italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ) / 2. a()superscript𝑎a^{(\dagger)}italic_a start_POSTSUPERSCRIPT ( † ) end_POSTSUPERSCRIPT is the annihilation (creation) operator of the photon mode in the cavity. λ𝜆\lambdaitalic_λ is the coupling strength between the TLSs and the cavity. We consider a mean-field ansatz which separates the model into the two sub-Hamiltonians

1Hj1Planck-constant-over-2-pisuperscript𝐻𝑗\displaystyle\frac{1}{\hbar}H^{j}divide start_ARG 1 end_ARG start_ARG roman_ℏ end_ARG italic_H start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT =ωzj2σzj+σ+jA(t)+σjA+(t)+λNa+aσxjabsentsuperscriptsubscript𝜔𝑧𝑗2superscriptsubscript𝜎𝑧𝑗superscriptsubscript𝜎𝑗subscript𝐴𝑡superscriptsubscript𝜎𝑗subscript𝐴𝑡𝜆𝑁expectation𝑎superscript𝑎superscriptsubscript𝜎𝑥𝑗\displaystyle=\frac{\omega_{z}^{j}}{2}\sigma_{z}^{j}+\sigma_{+}^{j}A_{-}(t)+% \sigma_{-}^{j}A_{+}(t)+\frac{\lambda}{\sqrt{N}}\braket{a+a^{\dagger}}\sigma_{x% }^{j}= divide start_ARG italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT + italic_σ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT - end_POSTSUBSCRIPT ( italic_t ) + italic_σ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( italic_t ) + divide start_ARG italic_λ end_ARG start_ARG square-root start_ARG italic_N end_ARG end_ARG ⟨ start_ARG italic_a + italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT end_ARG ⟩ italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT (3)
1Hc1Planck-constant-over-2-pisubscript𝐻𝑐\displaystyle\frac{1}{\hbar}H_{c}divide start_ARG 1 end_ARG start_ARG roman_ℏ end_ARG italic_H start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT =ωcaa+λNj=1Nσxj(a+a).absentsubscript𝜔𝑐superscript𝑎𝑎𝜆𝑁superscriptsubscript𝑗1𝑁expectationsubscriptsuperscript𝜎𝑗𝑥𝑎superscript𝑎\displaystyle=\omega_{c}a^{\dagger}a+\frac{\lambda}{\sqrt{N}}\sum_{j=1}^{N}% \braket{\sigma^{j}_{x}}(a+a^{\dagger}).= italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_a + divide start_ARG italic_λ end_ARG start_ARG square-root start_ARG italic_N end_ARG end_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ⟨ start_ARG italic_σ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_ARG ⟩ ( italic_a + italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) . (4)

Here σxjexpectationsuperscriptsubscript𝜎𝑥𝑗\braket{\sigma_{x}^{j}}⟨ start_ARG italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT end_ARG ⟩ and a+aexpectation𝑎superscript𝑎\braket{a+a^{\dagger}}⟨ start_ARG italic_a + italic_a start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT end_ARG ⟩ are the expectation values of the respective operators. The dynamics generated by Eq. 4 are solved by a coherent state characterised by α=a𝛼expectation𝑎\alpha=\braket{a}italic_α = ⟨ start_ARG italic_a end_ARG ⟩ which acts as the order-paramater of superradiant phases. The dynamics of α𝛼\alphaitalic_α are governed by the equation of motion

α˙=(iωc+κ)αiλNj=1Nσxj,˙𝛼𝑖subscript𝜔𝑐𝜅𝛼𝑖𝜆𝑁superscriptsubscript𝑗1𝑁expectationsuperscriptsubscript𝜎𝑥𝑗\dot{\alpha}=-(i\omega_{c}+\kappa)\alpha-i\frac{\lambda}{\sqrt{N}}\sum_{j=1}^{% N}\braket{\sigma_{x}^{j}},over˙ start_ARG italic_α end_ARG = - ( italic_i italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT + italic_κ ) italic_α - italic_i divide start_ARG italic_λ end_ARG start_ARG square-root start_ARG italic_N end_ARG end_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ⟨ start_ARG italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT end_ARG ⟩ , (5)

where κ𝜅\kappaitalic_κ is the cavity loss rate. Additionally, we express the dynamics of the j𝑗jitalic_jth TLS via the Lindblad master equation which we write as

ρ˙jsuperscript˙𝜌𝑗\displaystyle\dot{\rho}^{j}over˙ start_ARG italic_ρ end_ARG start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT =i[ρj,Hj]+lγlj(LljρjLlj,12{Llj,Llj,ρj}).absent𝑖superscript𝜌𝑗superscript𝐻𝑗subscript𝑙superscriptsubscript𝛾𝑙𝑗superscriptsubscript𝐿𝑙𝑗superscript𝜌𝑗superscriptsubscript𝐿𝑙𝑗12superscriptsubscript𝐿𝑙𝑗superscriptsubscript𝐿𝑙𝑗superscript𝜌𝑗\displaystyle=i[\rho^{j},H^{j}]+\sum_{l}\gamma_{l}^{j}(L_{l}^{j}\rho^{j}L_{l}^% {j,\dagger}-\frac{1}{2}\{L_{l}^{j,\dagger}L_{l}^{j},\rho^{j}\}).= italic_i [ italic_ρ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT , italic_H start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ] + ∑ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT italic_γ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ( italic_L start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT italic_ρ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT italic_L start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j , † end_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG { italic_L start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j , † end_POSTSUPERSCRIPT italic_L start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT , italic_ρ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT } ) . (6)

The Lindblad operators Llj=σ+j,σj,σzjsuperscriptsubscript𝐿𝑙𝑗superscriptsubscript𝜎𝑗superscriptsubscript𝜎𝑗superscriptsubscript𝜎𝑧𝑗L_{l}^{j}=\sigma_{+}^{j},\sigma_{-}^{j},\sigma_{z}^{j}italic_L start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT = italic_σ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT , italic_σ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT , italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT are weighted by the dissipation coefficients γljsuperscriptsubscript𝛾𝑙𝑗\gamma_{l}^{j}italic_γ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT with l{+,,z}𝑙𝑧l\in\{+,-,z\}italic_l ∈ { + , - , italic_z } and act in the instantaneous eigenbasis of the j𝑗jitalic_jth TLS analogously to the method applied in previous works Nuske et al. (2020); Broers and Mathey (2023, 2021, 2022). The coefficients γjsuperscriptsubscript𝛾𝑗\gamma_{-}^{j}italic_γ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT and γ+jsuperscriptsubscript𝛾𝑗\gamma_{+}^{j}italic_γ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT describe spontaneous decay and excitation, respectively. The coefficients γzj=γzsuperscriptsubscript𝛾𝑧𝑗subscript𝛾𝑧\gamma_{z}^{j}=\gamma_{z}italic_γ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT = italic_γ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT describe dephasing. This choice of dissipation has been shown to describe the dynamics in light-driven two-band solids Nuske et al. (2020). Note that the temperature T𝑇Titalic_T is encoded in the ratio of

γjγ+jγj+γ+j=tanh(ϵjkBT),superscriptsubscript𝛾𝑗superscriptsubscript𝛾𝑗superscriptsubscript𝛾𝑗superscriptsubscript𝛾𝑗superscriptitalic-ϵ𝑗subscript𝑘𝐵𝑇\frac{\gamma_{-}^{j}-\gamma_{+}^{j}}{\gamma_{-}^{j}+\gamma_{+}^{j}}=\tanh\left% (\frac{\epsilon^{j}}{k_{B}T}\right),divide start_ARG italic_γ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT - italic_γ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT end_ARG start_ARG italic_γ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT + italic_γ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT end_ARG = roman_tanh ( divide start_ARG italic_ϵ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG ) , (7)

where the instantaneous eigenenergy scale ϵjsuperscriptitalic-ϵ𝑗\epsilon^{j}italic_ϵ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT of Hjsuperscript𝐻𝑗H^{j}italic_H start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT is roughly of the order of the j𝑗jitalic_jth TLS level spacing ωzjsuperscriptsubscript𝜔𝑧𝑗\omega_{z}^{j}italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT. Note that this suggests an ideal range of operation for the FSP. In terms of experimental feasibility it is desirable to keep Edsubscript𝐸dE_{\mathrm{d}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT small but ϵjkBTsuperscriptitalic-ϵ𝑗subscript𝑘𝐵𝑇\frac{\epsilon^{j}}{k_{B}T}divide start_ARG italic_ϵ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG large. This compromise is met up to room temperature for characteristic frequencies of the order of tens to hundreds of terahertz, which is in agreement with the motivational work on light-driven graphene McIver et al. (2020). As an example throughout this work, we use ωd=2π×48THzsubscript𝜔d2𝜋48terahertz\omega_{\mathrm{d}}=2\pi\times 48$\mathrm{THz}$italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = 2 italic_π × 48 roman_THz and ωz=2π×24THzsubscript𝜔𝑧2𝜋24terahertz\omega_{z}=2\pi\times 24$\mathrm{THz}$italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 2 italic_π × 24 roman_THz such that at room temperature tanh(ϵjkBT)1superscriptitalic-ϵ𝑗subscript𝑘𝐵𝑇1\tanh(\frac{\epsilon^{j}}{k_{B}T})\approx 1roman_tanh ( divide start_ARG italic_ϵ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG ) ≈ 1. Therefore, we take γ+j0superscriptsubscript𝛾𝑗0\gamma_{+}^{j}\approx 0italic_γ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ≈ 0.

We further use ωc=2π×12THzsubscript𝜔𝑐2𝜋12terahertz\omega_{c}=2\pi\times 12$\mathrm{THz}$italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 2 italic_π × 12 roman_THz, γj+γ+j=γ+γ+=2THzsuperscriptsubscript𝛾𝑗superscriptsubscript𝛾𝑗subscript𝛾subscript𝛾2terahertz\gamma_{-}^{j}+\gamma_{+}^{j}=\gamma_{-}+\gamma_{+}=2$\mathrm{THz}$italic_γ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT + italic_γ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT = italic_γ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT + italic_γ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT = 2 roman_THz, γz=4THzsubscript𝛾𝑧4terahertz\gamma_{z}=4$\mathrm{THz}$italic_γ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 4 roman_THz, κ=2π×120MHz𝜅2𝜋120megahertz\kappa=2\pi\times 120$\mathrm{MHz}$italic_κ = 2 italic_π × 120 roman_MHz and λc=ωcωz/22π×8.5THzsubscript𝜆𝑐subscript𝜔𝑐subscript𝜔𝑧22𝜋8.5terahertz\lambda_{c}=\sqrt{\omega_{c}\omega_{z}}/2\approx 2\pi\times 8.5$\mathrm{THz}$italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = square-root start_ARG italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG / 2 ≈ 2 italic_π × 8.5 roman_THz, where λcsubscript𝜆𝑐\lambda_{c}italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is the critical coupling strength of the standard Dicke model. The critical coupling λcsubscript𝜆𝑐\lambda_{c}italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT does not directly relate to the FSP, but it gives a readily comparable scale of the system parameters. Unless stated otherwise, we take the driving field strength to be adjusted to the onset value of the FSP. This means that the Floquet energies of the driven TLSs are resonant with the cavity frequency ωcsubscript𝜔𝑐\omega_{c}italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, such that

Ed=ωd22(1ωcωd)2(1ωzωd)2,subscript𝐸dsuperscriptsubscript𝜔d22superscript1subscript𝜔𝑐subscript𝜔d2superscript1subscript𝜔𝑧subscript𝜔d2E_{\mathrm{d}}=\frac{\omega_{\mathrm{d}}^{2}}{2}\sqrt{\left(1-\frac{\omega_{c}% }{\omega_{\mathrm{d}}}\right)^{2}-\left(1-\frac{\omega_{z}}{\omega_{\mathrm{d}% }}\right)^{2}},italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = divide start_ARG italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG square-root start_ARG ( 1 - divide start_ARG italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ( 1 - divide start_ARG italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (8)

as we have discussed in previous work Broers and Mathey (2023). In our example, with ωc/ωd=1/4subscript𝜔𝑐subscript𝜔d14\omega_{c}/\omega_{\mathrm{d}}=1/4italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = 1 / 4 and ωz/ωd=1/2subscript𝜔𝑧subscript𝜔d12\omega_{z}/\omega_{\mathrm{d}}=1/2italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = 1 / 2, this amounts to

Edonset=58ωd2.superscriptsubscript𝐸donset58superscriptsubscript𝜔d2E_{\mathrm{d}}^{\mathrm{onset}}=\frac{\sqrt{5}}{8}\omega_{\mathrm{d}}^{2}.italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_onset end_POSTSUPERSCRIPT = divide start_ARG square-root start_ARG 5 end_ARG end_ARG start_ARG 8 end_ARG italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (9)

The vertical line in Fig. 1 indicates the onset driving field strength.

III Phase Diffusion

Refer to caption
Refer to caption
Figure 2: Light field fluctuations across the FSP transition. Panels (a) and (b) show the root-mean-square and the standard deviation of the light field amplitude across the FSP transition as a function of the phase diffusion standard deviation sϕsubscript𝑠italic-ϕs_{\phi}italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT. The horizontal line indicates the coupling strength at which the FSP transition occurs in the absence of phase diffusion, i.e. sϕ=0subscript𝑠italic-ϕ0s_{\phi}=0italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 0. Panel (c) shows the amplitude of a single-trajectory of the light field for sϕ=0subscript𝑠italic-ϕ0s_{\phi}=0italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 0 (blue) and sϕ=0.4subscript𝑠italic-ϕ0.4s_{\phi}=0.4italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 0.4 (black), as well as an inset of the driving field for the same values of sϕsubscript𝑠italic-ϕs_{\phi}italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT.

As a first metric for robustness, we consider the influence of a finite linewidth of the driving field on the linewidth of the cavity light field in the FSP. For this purpose we introduce a Gaussian random walk ϕ(t)italic-ϕ𝑡\phi(t)italic_ϕ ( italic_t ) that models phase diffusion in the driving field 111We write the discretized random walk as ϕn=i=0n𝒩(0,sϕ2nT),subscriptitalic-ϕ𝑛superscriptsubscript𝑖0𝑛𝒩0superscriptsubscript𝑠italic-ϕ2subscript𝑛𝑇\displaystyle\phi_{n}=\sum_{i=0}^{n}\mathcal{N}\left(0,\frac{s_{\phi}^{2}}{n_{% T}}\right),italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT caligraphic_N ( 0 , divide start_ARG italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_n start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_ARG ) , (10) where 𝒩(μ,ν)𝒩𝜇𝜈\mathcal{N}(\mu,\nu)caligraphic_N ( italic_μ , italic_ν ) indicates a randomly sampled value from the normal distribution with mean μ𝜇\muitalic_μ and variance ν𝜈\nuitalic_ν. nTsubscript𝑛𝑇n_{T}italic_n start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT is the amount of time-steps per driving period, such that ϕ(t)italic-ϕ𝑡\phi(t)italic_ϕ ( italic_t ) is determined by interpolation of ϕnsubscriptitalic-ϕ𝑛\phi_{n}italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT., rather than the monochromatic driving field described in Eq. 2. The standard deviation of ϕ(t)italic-ϕ𝑡\phi(t)italic_ϕ ( italic_t ) after one driving period 2πωd12𝜋superscriptsubscript𝜔d12\pi\omega_{\mathrm{d}}^{-1}2 italic_π italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT is given by sϕsubscript𝑠italic-ϕs_{\phi}italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT. This corresponds to a linewidth of Δω=ωdsϕ2πΔ𝜔subscript𝜔dsubscript𝑠italic-ϕ2𝜋\Delta\omega=\omega_{\mathrm{d}}\frac{s_{\phi}}{2\pi}roman_Δ italic_ω = italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT divide start_ARG italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_π end_ARG in the driving field which we now write as

A±(t)=Edωdexp{±i(ωdt+ϕ(t))}.subscript𝐴plus-or-minus𝑡subscript𝐸dsubscript𝜔dplus-or-minus𝑖subscript𝜔d𝑡italic-ϕ𝑡A_{\pm}(t)=\frac{E_{\mathrm{d}}}{\omega_{\mathrm{d}}}\exp\{\pm i(\omega_{% \mathrm{d}}t+\phi(t))\}.italic_A start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ( italic_t ) = divide start_ARG italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT end_ARG roman_exp { ± italic_i ( italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT italic_t + italic_ϕ ( italic_t ) ) } . (11)

In Fig. 2 (a), we show the root-mean-square of the light field amplitude |α|rmssubscript𝛼rms|\alpha|_{\mathrm{rms}}| italic_α | start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT across the FSP transition at the onset value of Ed=Edonsetsubscript𝐸dsuperscriptsubscript𝐸donsetE_{\mathrm{d}}=E_{\mathrm{d}}^{\mathrm{onset}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_onset end_POSTSUPERSCRIPT, see Eq. 9, as a function of λ𝜆\lambdaitalic_λ and sϕsubscript𝑠italic-ϕs_{\phi}italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT. We see that the FSP transition is stable against the fluctuations of the driving field, i.e. as a function of sϕsubscript𝑠italic-ϕs_{\phi}italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT, however the FSP regime is shifted towards larger values of λ𝜆\lambdaitalic_λ, and the transition regime displays stronger fluctuations. Since the phase diffusion broadens the linewidth of the driving field, the TLSs develop amplitude at many frequencies including the cavity frequency ωcsubscript𝜔𝑐\omega_{c}italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT which leads to residual occupations in the cavity that contribute to the broadened FSP transition.

In Fig. 2 (b), we show the standard deviation |α|stdsubscript𝛼std|\alpha|_{\mathrm{std}}| italic_α | start_POSTSUBSCRIPT roman_std end_POSTSUBSCRIPT of the light field amplitude as a function of λ𝜆\lambdaitalic_λ and sϕsubscript𝑠italic-ϕs_{\phi}italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT. Close to the FSP transition, the amplitude that we show in Fig. 2 (a) displays large fluctuations, that are suppressed both in the FSP and the trivial phase. With increasing phase diffusion, the steady state in the FSP shows an increasing standard deviation and the sharp feature in |α|stdsubscript𝛼std|\alpha|_{\mathrm{std}}| italic_α | start_POSTSUBSCRIPT roman_std end_POSTSUBSCRIPT across the transition broadens. For intermediate values of sϕsubscript𝑠italic-ϕs_{\phi}italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT, the increased standard deviation indicates the shifted location of the FSP transition.

In Fig. 2 (c), we show the amplitude of the light field in the FSP at λ=λc𝜆subscript𝜆𝑐\lambda=\lambda_{c}italic_λ = italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for sϕ=0subscript𝑠italic-ϕ0s_{\phi}=0italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 0 and sϕ=0.4subscript𝑠italic-ϕ0.4s_{\phi}=0.4italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 0.4 on long time scales. The case of sϕ=0.4subscript𝑠italic-ϕ0.4s_{\phi}=0.4italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 0.4 corresponds to a significantly broadened driving field and hence the amplitude of the light field in the FSP jitters considerably. This case corresponds to the vertical dashed lines in Figs. 2 (a) and (b). The inset shows the effect of the phase diffusion on the driving field.

Refer to caption
Figure 3: Linewidth narrowing in the FSP. Panel (a) shows the power spectrum averaged over 50 phase diffusion trajectories as a function of the coupling strength λ𝜆\lambdaitalic_λ for sϕ=0.4subscript𝑠italic-ϕ0.4s_{\phi}=0.4italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 0.4 on a logarithmic scale. Across the FSP transition indicated by the dashed line, the linewidth of the light field in the cavity narrows drastically. Panel (b) shows the light field amplitude across the FSP transition for the individual trajectories in light colors and their mean in solid dark blue. Panel (c) shows the power spectra for λ=0.02λc𝜆0.02subscript𝜆𝑐\lambda=0.02\lambda_{c}italic_λ = 0.02 italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, rescaled by a factor of 26262626 for comparison, and λ=λc𝜆subscript𝜆𝑐\lambda=\lambda_{c}italic_λ = italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT.

Next we present the line narrowing of the light field across the FSP transition in the presence of phase diffusion. For this purpose we consider the power spectrum

S(ω)=|α^(ω)|2|α^(ω)|2𝑑ω,𝑆𝜔superscript^𝛼𝜔2subscriptsuperscript^𝛼𝜔2differential-d𝜔S(\omega)=\frac{|\hat{\alpha}(\omega)|^{2}}{\int_{\mathbb{R}}|\hat{\alpha}(% \omega)|^{2}d\omega},italic_S ( italic_ω ) = divide start_ARG | over^ start_ARG italic_α end_ARG ( italic_ω ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ∫ start_POSTSUBSCRIPT blackboard_R end_POSTSUBSCRIPT | over^ start_ARG italic_α end_ARG ( italic_ω ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d italic_ω end_ARG , (12)

where α^(ω)^𝛼𝜔\hat{\alpha}(\omega)over^ start_ARG italic_α end_ARG ( italic_ω ) is the Fourier transform of α(t)𝛼𝑡\alpha(t)italic_α ( italic_t ).

In Fig. 3 (a) we show S(ω)𝑆𝜔S(\omega)italic_S ( italic_ω ) at the onset driving field strength with sϕ=0.4subscript𝑠italic-ϕ0.4s_{\phi}=0.4italic_s start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT = 0.4 and averaged over 50 phase diffusion trajectories. Prior to the FSP transition that occurs at approximately λ=0.37λc𝜆0.37subscript𝜆𝑐\lambda=0.37\lambda_{c}italic_λ = 0.37 italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, there is a broadened signal at the cavity frequency with a linewidth that is given by the cavity loss rate κ𝜅\kappaitalic_κ. Across the transition, the linewidth narrows drastically as the occupation of the cavity mode increases. We show the corresponding light field amplitude |α|𝛼|\alpha|| italic_α | in Fig. 3 (b) as a transposed plot for the same set of sampled phase diffusion trajectories in light colors, and their mean as a dark solid line.

In Fig. 3 (c) we show the power spectrum S(ω)𝑆𝜔S(\omega)italic_S ( italic_ω ) prior to (λ=0.02λc𝜆0.02subscript𝜆𝑐\lambda=0.02\lambda_{c}italic_λ = 0.02 italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT) and past (λ=λc𝜆subscript𝜆𝑐\lambda=\lambda_{c}italic_λ = italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT) the FSP transition, rescaled for comparison. The line narrowing is clearly visible as the full-width half-maximum reduces drastically. The narrowing of the linewidth past the transition corresponds to the coherence times increasing beyond those of the driving field and those given intrinsically by the cavity. This is a hallmark of lasing states.

IV Inhomogeneous Broadening

As a second metric for robustness we consider the consequences of inhomogeneous broadening on the FSP transition. In order to include inhomogeneous broadening, we consider a distribution of detuned two-level spacings ωzjsuperscriptsubscript𝜔𝑧𝑗\omega_{z}^{j}italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT that have the average frequency ωzsubscript𝜔𝑧\omega_{z}italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. In this model the index of the TLSs is arbitrary, so we label them by their energy for convenience. The relevant quantity for broadening effects is the mode density of the TLSs. We consider N=100𝑁100N=100italic_N = 100 TLSs with normal distributed energy detuning around ωzsubscript𝜔𝑧\omega_{z}italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT with a relative standard deviation of sωsubscript𝑠𝜔s_{\omega}italic_s start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT. For this purpose, we set the j𝑗jitalic_jth energy level to

ωzj=ωz(1+sωerf1[2jN+11]),superscriptsubscript𝜔𝑧𝑗subscript𝜔𝑧1subscript𝑠𝜔superscripterf1delimited-[]2𝑗𝑁11\omega_{z}^{j}=\omega_{z}\left(1+s_{\omega}\mathrm{erf}^{-1}\left[\frac{2j}{N+% 1}-1\right]\right),italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT = italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( 1 + italic_s start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT roman_erf start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT [ divide start_ARG 2 italic_j end_ARG start_ARG italic_N + 1 end_ARG - 1 ] ) , (13)

where erf1superscripterf1\mathrm{erf}^{-1}roman_erf start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT is the inverse error function, such that the energy levels are evenly distributed as desired without random sampling.

Refer to caption
Figure 4: Effect of inhomogeneous broadening on the FSP transition. The FSP transition at the onset driving field strength as a function of the coupling strength λ𝜆\lambdaitalic_λ and the inhomogeneous broadening parameter sωsubscript𝑠𝜔s_{\omega}italic_s start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT. The vertical dashed line corresponds to the case that we show in Fig. 5.

In Fig. 4, we show the amplitude of the light field across the FSP transition as a function of the inhomogeneous broadening standard deviation sωsubscript𝑠𝜔s_{\omega}italic_s start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT. We find that the transition remains sharp for increasing sωsubscript𝑠𝜔s_{\omega}italic_s start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT. However, the value of λ𝜆\lambdaitalic_λ at which the FSP transition occurs increases with sωsubscript𝑠𝜔s_{\omega}italic_s start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT while the amplitude |α|𝛼|\alpha|| italic_α | of the light field decreases. This is a consequence of a subset of TLSs not contributing to the FSP due to being far detuned from the cavity frequency. To compensate for this lack of contributing TLSs, the coupling strength needs to increase in order to enter the FSP. If the detuning due to inhomogeneous broadening is larger than the intrinsic linewidth of a given TLS, it is not affected by the effective driving that is present due to the interaction with the cavity which contains a non-zero light field.

In order to gain insight into the Floquet states of the system as well as their occupation, we consider the TLSs embedded in a larger space spanned by the creation (annihilation) operators b1,2j,()superscriptsubscript𝑏12𝑗b_{1,2}^{j,(\dagger)}italic_b start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j , ( † ) end_POSTSUPERSCRIPT. They are connected to the Pauli matrices via

σxjsubscriptsuperscript𝜎𝑗𝑥\displaystyle\sigma^{j}_{x}italic_σ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT =b2j,b1j+b1j,b2jabsentsuperscriptsubscript𝑏2𝑗subscriptsuperscript𝑏𝑗1superscriptsubscript𝑏1𝑗subscriptsuperscript𝑏𝑗2\displaystyle=b_{2}^{j,\dagger}b^{j}_{1}+b_{1}^{j,\dagger}b^{j}_{2}= italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j , † end_POSTSUPERSCRIPT italic_b start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j , † end_POSTSUPERSCRIPT italic_b start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (14)
σyjsubscriptsuperscript𝜎𝑗𝑦\displaystyle\sigma^{j}_{y}italic_σ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT =i(b2j,b1jb1j,b2j)absent𝑖superscriptsubscript𝑏2𝑗subscriptsuperscript𝑏𝑗1superscriptsubscript𝑏1𝑗subscriptsuperscript𝑏𝑗2\displaystyle=i(b_{2}^{j,\dagger}b^{j}_{1}-b_{1}^{j,\dagger}b^{j}_{2})= italic_i ( italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j , † end_POSTSUPERSCRIPT italic_b start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j , † end_POSTSUPERSCRIPT italic_b start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) (15)
σzjsubscriptsuperscript𝜎𝑗𝑧\displaystyle\sigma^{j}_{z}italic_σ start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT =b2j,b2jb1j,b1j.absentsuperscriptsubscript𝑏2𝑗subscriptsuperscript𝑏𝑗2superscriptsubscript𝑏1𝑗subscriptsuperscript𝑏𝑗1\displaystyle=b_{2}^{j,\dagger}b^{j}_{2}-b_{1}^{j,\dagger}b^{j}_{1}.= italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j , † end_POSTSUPERSCRIPT italic_b start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j , † end_POSTSUPERSCRIPT italic_b start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT . (16)

We note that the operators b1,2subscript𝑏12b_{1,2}italic_b start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT can be either fermionic operators or hardcore bosons. While we introduce them as auxiliary operators in this work, a natural platform to implement the dynamical state that we put forth here, is in a two-band material. In that implementation, the electron operators of the two bands coincide with the Schwinger operators that we use here. In this Schwinger representation, individual two-time correlation functions bnj,(t2)bnj(t1)expectationsubscriptsuperscript𝑏𝑗𝑛subscript𝑡2subscriptsuperscript𝑏𝑗𝑛subscript𝑡1\braket{b^{j,\dagger}_{n}(t_{2})b^{j}_{n}(t_{1})}⟨ start_ARG italic_b start_POSTSUPERSCRIPT italic_j , † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) italic_b start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) end_ARG ⟩ are accessible, rather than propagators of particle conserving operators Eq. 14–16. We can calculate the steady state distribution 222Eq. 17 can be transformed such that it is always t2>t1subscript𝑡2subscript𝑡1t_{2}>t_{1}italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. Then the correlation function in Eq. 18 can be calculated by strict forward propagation using the master equation and applying the operators bnjsuperscriptsubscript𝑏𝑛𝑗b_{n}^{j}italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT and bnj,superscriptsubscript𝑏𝑛𝑗b_{n}^{j,\dagger}italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j , † end_POSTSUPERSCRIPT at t1subscript𝑡1t_{1}italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and t2subscript𝑡2t_{2}italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, respectively.

nj(ω)=1(tbta)2tatbtatb𝒢j(t2,t1)eiω(t2t1)𝑑t2𝑑t1subscript𝑛𝑗𝜔1superscriptsubscript𝑡bsubscript𝑡a2superscriptsubscriptsubscript𝑡asubscript𝑡bsuperscriptsubscriptsubscript𝑡asubscript𝑡bsuperscript𝒢𝑗subscript𝑡2subscript𝑡1superscript𝑒𝑖𝜔subscript𝑡2subscript𝑡1differential-dsubscript𝑡2differential-dsubscript𝑡1n_{j}(\omega)=\frac{1}{(t_{\mathrm{b}}-t_{\mathrm{a}})^{2}}\int_{t_{\mathrm{a}% }}^{t_{\mathrm{b}}}\int_{t_{\mathrm{a}}}^{t_{\mathrm{b}}}\mathcal{G}^{j}(t_{2}% ,t_{1})e^{i\omega(t_{2}-t_{1})}dt_{2}dt_{1}italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_ω ) = divide start_ARG 1 end_ARG start_ARG ( italic_t start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT roman_a end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∫ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT roman_a end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT roman_a end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT end_POSTSUPERSCRIPT caligraphic_G start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_e start_POSTSUPERSCRIPT italic_i italic_ω ( italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT italic_d italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_d italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (17)

for each of the N𝑁Nitalic_N TLSs with the correlation function

𝒢j(t2,t1)=n=12bnj,(t2)bnj(t1).superscript𝒢𝑗subscript𝑡2subscript𝑡1superscriptsubscript𝑛12expectationsuperscriptsubscript𝑏𝑛𝑗subscript𝑡2subscriptsuperscript𝑏𝑗𝑛subscript𝑡1\mathcal{G}^{j}(t_{2},t_{1})=\sum_{n=1}^{2}\braket{b_{n}^{j,\dagger}(t_{2})b^{% j}_{n}(t_{1})}.caligraphic_G start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) = ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟨ start_ARG italic_b start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j , † end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) italic_b start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) end_ARG ⟩ . (18)

Here tasubscript𝑡at_{\mathrm{a}}italic_t start_POSTSUBSCRIPT roman_a end_POSTSUBSCRIPT is a time that is large enough for the system to have formed a steady state and tbtasubscript𝑡bsubscript𝑡at_{\mathrm{b}}-t_{\mathrm{a}}italic_t start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT roman_a end_POSTSUBSCRIPT is an interval that is large enough to ensure sufficient frequency resolution. The distribution nj(ω)subscript𝑛𝑗𝜔n_{j}(\omega)italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_ω ) reveals the energies and occupation of the Floquet states of the j𝑗jitalic_jth TLS. Collecting the steady state distributions of all N𝑁Nitalic_N TLSs gives the collective distribution of the entirety of TLSs

n(ω)=N1j=1Nnj(ω),𝑛𝜔superscript𝑁1superscriptsubscript𝑗1𝑁subscript𝑛𝑗𝜔n(\omega)=N^{-1}\sum_{j=1}^{N}n_{j}(\omega),italic_n ( italic_ω ) = italic_N start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_ω ) , (19)

which displays the frequency resolved distribution of the ensemble of TLSs, rather than that of individual TLSs. Taking the difference of the collective distribution across opposite frequencies gives the relative collective distribution

Δn(ω)=n(ω)n(ω).Δ𝑛𝜔𝑛𝜔𝑛𝜔\Delta n(\omega)=n(\omega)-n(-\omega).roman_Δ italic_n ( italic_ω ) = italic_n ( italic_ω ) - italic_n ( - italic_ω ) . (20)

Δn(ω)Δ𝑛𝜔\Delta n(\omega)roman_Δ italic_n ( italic_ω ) displays the frequency resolved imbalance of the ensemble and reveals the effective population inversion of the entire system and how it is depleted in the FSP.

Refer to caption
Figure 5: Two-level steady state distributions in the presence of inhomogeneous broadening. Panels (a) through (c) show the steady state distributions nj(ω)subscript𝑛𝑗𝜔n_{j}(\omega)italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_ω ) for N=100𝑁100N=100italic_N = 100 TLSs at Ed=0subscript𝐸d0E_{\mathrm{d}}=0italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = 0 (a) and Ed=Edonsetsubscript𝐸dsuperscriptsubscript𝐸donsetE_{\mathrm{d}}=E_{\mathrm{d}}^{\mathrm{onset}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_onset end_POSTSUPERSCRIPT (b, c) as well as λ=0𝜆0\lambda=0italic_λ = 0 (a, b) and λ=λc𝜆subscript𝜆𝑐\lambda=\lambda_{c}italic_λ = italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (c). The distributions are concentrated at the Floquet energies of the broadened energy levels. The horizontal lines show the average level spacing ωzsubscript𝜔𝑧\omega_{z}italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and the cavity frequency ωcsubscript𝜔𝑐\omega_{c}italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. Panel (d) shows the relative collective distribution Δn(ω)Δ𝑛𝜔\Delta n(\omega)roman_Δ italic_n ( italic_ω ) for λ=λc𝜆subscript𝜆𝑐\lambda=\lambda_{c}italic_λ = italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (black line, white filling) and λ=0𝜆0\lambda=0italic_λ = 0 (dark blue filling) the difference between the two (hatched filling) is the effective population inversion of Floquet states that is depleted to sustain the FSP.

In Figs. 5 (a) through (c), we show the steady state distributions nj(ω)subscript𝑛𝑗𝜔n_{j}(\omega)italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_ω ) for sω=0.4subscript𝑠𝜔0.4s_{\omega}=0.4italic_s start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT = 0.4 and different combinations of Edsubscript𝐸dE_{\mathrm{d}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT and λ𝜆\lambdaitalic_λ. The horizontal lines indicate the centered two-level spacing ωzsubscript𝜔𝑧\omega_{z}italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and the cavity frequency ωcsubscript𝜔𝑐\omega_{c}italic_ω start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. The solid line indicates the centered spacing of the Floquet levels that for the j𝑗jitalic_jth TLS have the energy

ϵFj=±(ωd2Ed2ωd2+(ωdωzj)24).superscriptsubscriptitalic-ϵ𝐹𝑗plus-or-minussubscript𝜔d2superscriptsubscript𝐸d2superscriptsubscript𝜔d2superscriptsubscript𝜔dsuperscriptsubscript𝜔𝑧𝑗24\epsilon_{F}^{j}=\pm\left(\frac{\omega_{\mathrm{d}}}{2}-\sqrt{\frac{E_{\mathrm% {d}}^{2}}{\omega_{\mathrm{d}}^{2}}+\frac{(\omega_{\mathrm{d}}-\omega_{z}^{j})^% {2}}{4}}\right).italic_ϵ start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT = ± ( divide start_ARG italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG - square-root start_ARG divide start_ARG italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG ( italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG end_ARG ) . (21)

In Fig. 5 (a), we show the case of Ed=0subscript𝐸d0E_{\mathrm{d}}=0italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = 0 and λ=0𝜆0\lambda=0italic_λ = 0. In this equilibrium state the lower levels are all fully populated with energies given by Eq. 13. In Fig. 5 (b), we show the case of Ed=Edonsetsubscript𝐸dsuperscriptsubscript𝐸donsetE_{\mathrm{d}}=E_{\mathrm{d}}^{\mathrm{onset}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_onset end_POSTSUPERSCRIPT and λ=0𝜆0\lambda=0italic_λ = 0. As expected, the states are dominantly distributed around the Floquet energies given by Eq. 21. At this driving field strength the Floquet states are population inverted. In Fig. 5 (c), we show the case Ed=Edonsetsubscript𝐸dsuperscriptsubscript𝐸donsetE_{\mathrm{d}}=E_{\mathrm{d}}^{\mathrm{onset}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_onset end_POSTSUPERSCRIPT and λ=λc𝜆subscript𝜆𝑐\lambda=\lambda_{c}italic_λ = italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. This case is inside the FSP and the Floquet states that are close to resonant with the cavity frequency are modified due to the presence of the finite photon field in the cavity. Hence, a gap opens and the populations of the Floquet states are modified.

In Fig. 5 (d), we show the relative collective distribution Δn(ω)Δ𝑛𝜔\Delta n(\omega)roman_Δ italic_n ( italic_ω ) for sω=0.4subscript𝑠𝜔0.4s_{\omega}=0.4italic_s start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT = 0.4, Ed=Edonsetsubscript𝐸dsuperscriptsubscript𝐸donsetE_{\mathrm{d}}=E_{\mathrm{d}}^{\mathrm{onset}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_onset end_POSTSUPERSCRIPT and for both λ=0𝜆0\lambda=0italic_λ = 0 (dark blue filling) and λ=λc𝜆subscript𝜆𝑐\lambda=\lambda_{c}italic_λ = italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (black line, white filling), which correspond to Figs. 5 (b) and (c), respectively. We see that inside the FSP and close to the cavity resonance, the Floquet states are not only modified by a gap opening, but the effective population inversion is largely depleted. The inhomogeneous broadening leads to off-resonant TLSs that are unaffected and do not contribute to the FSP mechanism. This type of hole-burning is responsible for the reduced light field amplitude as a function of sωsubscript𝑠𝜔s_{\omega}italic_s start_POSTSUBSCRIPT italic_ω end_POSTSUBSCRIPT in Fig. 4.

V Two-Level and Cavity Losses

As a third metric for robustness, we show the effect of the TLS dissipation rates and the cavity loss rate on the FSP. In Fig. 6 we show the magnitude of the light field in the cavity as a function of the effective driving field strength Edsubscript𝐸dE_{\mathrm{d}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT and the coupling strength λ𝜆\lambdaitalic_λ, as well as the dissipation coefficients. The quantity we show is the intensity per TLS of the light field |α|2N1superscript𝛼2superscript𝑁1|\alpha|^{2}N^{-1}| italic_α | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_N start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT multiplied with the cavity loss rate κ𝜅\kappaitalic_κ. This serves the purpose of quantifying the coherent output of the system rather than the photon count inside the cavity.

In Fig. 6 (a) and (b) we show the effect of the TLS dissipation coefficients. We choose the example γ=γz=2(γ+γ+)𝛾subscript𝛾𝑧2subscript𝛾subscript𝛾\gamma=\gamma_{z}=2(\gamma_{-}+\gamma_{+})italic_γ = italic_γ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 2 ( italic_γ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT + italic_γ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ), i.e. we keep the ratio of γ+γ+subscript𝛾subscript𝛾\gamma_{-}+\gamma_{+}italic_γ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT + italic_γ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT and γzsubscript𝛾𝑧\gamma_{z}italic_γ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT constant for convenience. We see that the FSP transition is robust for increasing values up to γ0.15ωdless-than-or-similar-to𝛾0.15subscript𝜔d\gamma\lesssim 0.15\omega_{\mathrm{d}}italic_γ ≲ 0.15 italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT for λ=0.75λc𝜆0.75subscript𝜆𝑐\lambda=0.75\lambda_{c}italic_λ = 0.75 italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. In our example, that is inspired by light-driven graphene, this corresponds to γ45THz𝛾45terahertz\gamma\approx 45$\mathrm{THz}$italic_γ ≈ 45 roman_THz compared to the characteristic energies of the Hamiltonian of the order of ωd=2π×48THzsubscript𝜔d2𝜋48terahertz\omega_{\mathrm{d}}=2\pi\times 48$\mathrm{THz}$italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = 2 italic_π × 48 roman_THz. This demonstrates a robustness to dissipation up to rates which are comparable to the coefficients used to describe experimental setups of driven graphene Breusing et al. (2011); Gierz et al. (2013); Brida et al. (2013); Plötzing et al. (2014); McIver et al. (2020); Heide et al. (2021). This further supports the viability of graphene as a potential platform for hosting the FSP.

Refer to caption
Figure 6: Effects of strong dissipation on the FSP. The output intensity of the light field per TLS κ|α|2/N𝜅superscript𝛼2𝑁\kappa|\alpha|^{2}/Nitalic_κ | italic_α | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_N as functions of combinations of Edsubscript𝐸dE_{\mathrm{d}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT, λ𝜆\lambdaitalic_λ, κ𝜅\kappaitalic_κ and γ𝛾\gammaitalic_γ. In panel (a), we show the output intensity per TLS as a function of γ𝛾\gammaitalic_γ and Edsubscript𝐸dE_{\mathrm{d}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT for λ=0.75λc𝜆0.75subscript𝜆𝑐\lambda=0.75\lambda_{c}italic_λ = 0.75 italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and κ=ωd/400𝜅subscript𝜔d400\kappa=\omega_{\mathrm{d}}/400italic_κ = italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT / 400. In panel (b), we show the output intensity per TLS as a function of γ𝛾\gammaitalic_γ and λ𝜆\lambdaitalic_λ for Ed=Edonsetsubscript𝐸dsuperscriptsubscript𝐸donsetE_{\mathrm{d}}=E_{\mathrm{d}}^{\mathrm{onset}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_onset end_POSTSUPERSCRIPT and κ=ωd/400𝜅subscript𝜔d400\kappa=\omega_{\mathrm{d}}/400italic_κ = italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT / 400. In panel (c), we show the output intensity per TLS as a function of κ𝜅\kappaitalic_κ and Edsubscript𝐸dE_{\mathrm{d}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT for λ=0.75λc𝜆0.75subscript𝜆𝑐\lambda=0.75\lambda_{c}italic_λ = 0.75 italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and γ=ωd/24π𝛾subscript𝜔d24𝜋\gamma=\omega_{\mathrm{d}}/24\piitalic_γ = italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT / 24 italic_π. In panel (d), we show the output intensity per TLS as a function of κ𝜅\kappaitalic_κ and λ𝜆\lambdaitalic_λ for Ed=Edonsetsubscript𝐸dsuperscriptsubscript𝐸donsetE_{\mathrm{d}}=E_{\mathrm{d}}^{\mathrm{onset}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_onset end_POSTSUPERSCRIPT and γ=ωd/24π𝛾subscript𝜔d24𝜋\gamma=\omega_{\mathrm{d}}/24\piitalic_γ = italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT / 24 italic_π.

In Fig. 6 (c) and (d) we show the effect of the cavity loss rate κ𝜅\kappaitalic_κ on the FSP. We find a small range of available values for κ𝜅\kappaitalic_κ at which the FSP emerges and further the threshold at which the FSP becomes unsustainable is comparatively small at values of κ0.02ωdless-than-or-similar-to𝜅0.02subscript𝜔d\kappa\lesssim 0.02\omega_{\mathrm{d}}italic_κ ≲ 0.02 italic_ω start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT for λ=0.75λc𝜆0.75subscript𝜆𝑐\lambda=0.75\lambda_{c}italic_λ = 0.75 italic_λ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. In our example this corresponds to κ5THz𝜅5terahertz\kappa\approx 5$\mathrm{THz}$italic_κ ≈ 5 roman_THz. This demonstrates the sensitivity of the FSP with respect to cavity losses, which we identify as the most sensitive parameter.

For the purpose of realizing the FSP, for driving field strengths close to Edonsetsuperscriptsubscript𝐸donsetE_{\mathrm{d}}^{\mathrm{onset}}italic_E start_POSTSUBSCRIPT roman_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_onset end_POSTSUPERSCRIPT the FSP will occur for large values of κ𝜅\kappaitalic_κ, if λ𝜆\lambdaitalic_λ is sufficiently large. Similarly, the given magnitude of λ𝜆\lambdaitalic_λ provides an upper limit on κ𝜅\kappaitalic_κ to achieve the realization of the FSP. We note that the FSP can be realized for arbitrarily small values of λ𝜆\lambdaitalic_λ, given sufficiently small κ𝜅\kappaitalic_κ. Therefore, for a given platform, a high-finesse cavity might enable the realization. We note, however, that the optimal output intensity is achieved for intermediate magnitudes of both κ𝜅\kappaitalic_κ and λ𝜆\lambdaitalic_λ. Stronger dissipation results in a larger inversion of the TLSs, and therefore in a higher output intensity. So for the optimal operation of FSP, intermediate values of the dissipation rates are desirable.

VI Conclusion

We have demonstrated the robustness of the Floquet-assisted superradiant phase (FSP) in the parametrically driven Dicke model with a dissipative model, that is designed for the electron dynamics in a solid. The photonic steady state in the FSP is robust against inhomogeneous broadening, reasonably strong dissipative processes of the two-level systems and phase diffusion in the driving field. Across the FSP transition, the linewidth of the light field narrows drastically and overcomes the linewidth of the driving field as well as the intrinsic linewidth of the cavity that is given by its loss rate, which is a hallmark of laser mechanisms.

In our model the dissipation is performed in the instantaneous eigenbasis. This choice is motivated by the capabilities of capturing dynamics of two-band solids, as has been demonstrated in previous work Nuske et al. (2020). The dependence of the FSP on the ratio of characteristic frequencies and temperature, as well as the driving field strength in units of the driving frequency squared suggests the most promising range of driving frequencies to be of the order of tens to hundreds of terahertz. This energy scale also coincides with the situation in two-band solids such as light-driven graphene McIver et al. (2020); Nuske et al. (2020); Broers and Mathey (2021). The values of the dissipation coefficients up to which we find the FSP to be stable also agrees with realistic estimates of coherence times in available graphene samples Breusing et al. (2011); Gierz et al. (2013); Brida et al. (2013); Plötzing et al. (2014); McIver et al. (2020); Heide et al. (2021).

We conclude from the robustness and the small values of the critical coupling of the FSP that it can be utilized for laser operation. The FSP is in principle accessible in two-band solids such as light-driven graphene under realistic conditions. The details of the collective effect of solid-state dispersion relations on the emergence of the FSP will be the subject of future research. Such a demonstration would then constitute a Floquet-assisted solid-state laser system in the terahertz frequency domain. As such, we propose to expand the range of dressed-state laser mechanisms into the domain of Floquet-engineered electron bands in solids, that is accessible with current pump-probe technology.

Acknowledgements

This work is funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – SFB-925 – project 170620586, and the Cluster of Excellence ’Advanced Imaging of Matter’ (EXC 2056), Project No. 390715994.

References