<<

Quantum Electrodynamic Control of Matter: Cavity-Enhanced Ferroelectric

Yuto Ashida∗ Department of Applied Physics, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan

Atac¸ I˙mamo˘glu and Jer´ omeˆ Faist Institute of Quantum Electronics, ETH Zurich, CH-8093 Zurich,¨ Switzerland

Dieter Jaksch Clarendon Laboratory, University of Oxford, Parks Road, Oxford OX1 3PU, United Kingdom

Andrea Cavalleri Max Planck Institute for the Structure and Dynamics of Matter, 22761 Hamburg, Germany and Clarendon Laboratory, University of Oxford, Parks Road, Oxford OX1 3PU, United Kingdom

Eugene Demler Department of Physics, Harvard University, Cambridge, MA 02138, USA

The interaction between light and matter can be utilized to qualitatively alter physical properties of mate- rials. Recent theoretical and experimental studies have explored this possibility of controlling matter by light based on driving many-body systems via strong classical electromagnetic radiation, leading to a time-dependent Hamiltonian for electronic or lattice degrees of freedom. To avoid inevitable heating effects in driven interacting systems, pump-probe setups with ultrashort laser pulses have so far been used to study transient light-induced modifications of material properties. Here, we pursue yet another direction of controlling quantum matter by modifying quantum fluctuations of its electromagnetic environment. In contrast to earlier proposals focusing on light-enhanced electron-electron interactions, we consider a dipolar quantum many-body system embedded in a cavity composed of metal mirrors, and formulate a theoretical framework to manipulate its equilibrium properties on the basis of quantum light-matter interaction. Owing to the strong coupling of polarizable mate- rial excitations to the cavity electric field, the cavity confinement leads to the hybridization of different types of the fundamental excitations, including dipolar , cavity photons, and in metal mirrors. We show that this hybridization qualitatively alters the nature of the collective excitations and can be used to selec- tively control energy-level structures in a wide range of systems supporting dipole moments. Most notably, in quantum paraelectric materials, we show that the cavity-induced softening of infrared optical phonons enhances the ferroelectric phase in comparison with the bulk materials. Our findings suggest an intriguing possibility of inducing a superradiant-type via the light-matter coupling without external pumping. We also discuss possible applications of the cavity-induced modifications in collective excitations to molecular materials and excitonic devices.

I. INTRODUCTION Studies of light-induced modifications in material proper- ties date back to experiments by Dayem and Wyatt [3,4], who observed that coherent microwave radiation leads to an One of the central goals in both quantum optics and con- increase of the critical current in superconductors. This phe- densed matter physics is to understand macroscopic phenom- nomenon has been understood from the Eliashberg electron- ena emerging from the fundamental quantum degrees of free- photon theory [5–7], in which photons with frequencies be- dom. Historically, quantum optics strived to study light- low the quasiparticle threshold are shown to modify the dis- matter interactions mainly in few-body regimes [1]. On an- tribution of quasiparticles in a way that the superconducting other front, investigated collective order parameter is enhanced. Other seminal works studying phenomena of many-body systems, and largely focused on op- control of matter by light include Floquet engineering of elec- timizing structural and electronic phases of . In partic- tronic band structures [8–12], and photo-induced pumping by ular, a number of techniques, including applied strain, elec- ultrashort laser pulses for creating metastable nonequilibrium

arXiv:2003.13695v1 [cond-mat.mes-hall] 30 Mar 2020 tronic doping, and isotope substitution, have been used to en- states [13–28]. Nonlinear optical phenomena can also be un- hance a macroscopic order such as ferroelectricity or super- derstood from the perspective of light-induced changes in op- conductivity. As an interdisciplinary frontier at the intersec- tical properties of matter [29–31]. For instance, the light- tion of optics and condensed matter physics, there have re- induced transparency can be considered as strong modifica- cently been remarkable developments in using classical elec- tion of the refractive index and absorption coefficient by a tromagnetic radiation to control transient states of matter [2]. control optical beam [32], and the optical phase conjugation The aim of this paper is to explore yet another route towards phenomena arise from the resonant parametric scattering of controlling the phase of matter by quantum light, namely, by signal photons and matter excitations [33, 34]. The essential modifying quantum fluctuations of the electromagnetic field element of all these developments is an external pump that in equilibrium – in the absence of an external drive. provides classical light field acting on quantum matter. 2

An alternative approach to controlling matter by light is to utilize quantum fluctuations of the vacuum electromagnetic field. A prototypical model to understand such quantum light- matter interaction is the Dicke model, which describes an en- semble of two-level atoms coupled to a single electromagnetic mode of a cavity [35–40]. For a sufficiently strong electric dipole, Dicke has shown that one should find instability into the superradiant phase characterized by spontaneous atomic polarization and macroscopic photon occupation [35]. This transition originates from the competition between the de- crease of the total energy due to the light-matter coupling and the energy cost of adding photons to the cavity and admixing the excited state of atoms. The Dicke model is exactly solv- able via the Bethe-ansatz equations originally discussed in the central spin model [41–43], and the existence of the superra- diant transition has been rigorously proven in the thermody- namic limit [37, 38]. Subsequently to Dicke’s original work, it has been pointed out that the superradiant transition requires such strong inter- action that one needs to include the Aˆ2 term in the Hamil- tonian, where Aˆ denotes the vector potential of electromag- netic fields. In turn, the latter term has been shown to pre- FIG. 1. (a) Schematic phase diagram of a quantum paraelectric vent the transition [44, 45]. Even though ideas have been sug- material. The horizontal axis represents the strength of a tuning pa- gested to circumvent this no-go theorem by using microwave rameter such as pressure or substitution, which governs the transi- resonators [46, 47] or including additional inter-particle in- tion between ferroelectric and paraelectric phases. The vertical axis is the temperature while the red dashed line in the paraelectric phase teractions [48, 49], the problem is still a subject of debate indicates the fact that the energy gap of an optical decreases [50]. An alternative proposal using multilevel systems with and goes down to zero at the paraelectric-to-ferroelectric transition external pump fields [51, 52] has been experimentally real- point. The main aim of this paper is to show that this structural quan- ized [53, 54]. Analogues of the Dicke transition have also tum phase transition can be induced by softening the phonon mode been explored in ultracold atoms, where matter excitations via the light-matter coupling in the cavity with no external pump. correspond to different momentum states of atoms [55, 56] (b,c) Ferroelectric and paraelectric phases in the case of ionic crys- or hyperfine states in spin-orbit-coupled Bose gases [57]. In tals. Green (blue) circles indicate the ions with the positive (neg- all these realizations, external driving is essential to obtain the ative) charges. The ferroelectric phase in (b) is characterized by a superradiant-type phases. nonzero uniform displacement causing an electric polarization (red arrows) while the mean-phonon displacement is absent in the para- The present work in this context suggests a promising route electric phase in (c). The variables φ and Ω represent the phonon towards realizing a genuine equilibrium superradiant transi- field and the bare phonon frequency, respectively (see Sec. III). tion under no external drive, which has so far remained elu- sive. Our consideration is motivated by an observation that transition into a superradiant state should be easier to attain in a system naturally close to a phase with spontaneous elec- such cavity is qualitatively different from the bulk tric polarization. We investigate cavity-induced changes in a spectrum, and can exhibit significant mode-softening, indicat- quantum paraelectric material at the verge of the ferroelec- ing enhancement of the ferroelectric instability. tric order (see Fig.1(a-c)); such materials can remain para- Building upon the analysis of collective hybridized modes, electrics even at zero temperature while the quantum phase we investigate the quantum phase transition into the ferro- transition into the ferroelectric phase can be induced by, e.g., electric phase of the heterostructure cavity configuration (cf. isotope substitution. We thus discuss a regime of a thermody- Fig.2). An important aspect of our analysis is the inclusion namically large system, in which multiple continuum modes of interactions originating from nonlinearities of the of both electromagnetic fields and matter excitations must be dipolar phonons. In the bulk geometry, the importance of such included. This makes a sharp contrast to the previous studies nonlinearities in understanding the quantum phase transition on the superradiant transitions, which have solely focused on has been discussed in Ref. [58]. In contrast, we focus on the a simplified case of many-body systems coupled to only a few cavity geometry and present a variational analysis to study ef- modes of a cavity. We formulate a simple yet general theo- fects of polariton interactions in such inhomogeneous setups. retical framework to analyze hybridization of continuum mat- One of the most notable results is that one can find the cavity- ter excitations and light modes confined in metallic cladding induced transition into a ferroelectric phase even when the layers acting as cavity mirrors (see Fig.2). Resulting polari- corresponding bulk system resides in the paraelectric phase. ton excitations consist of infrared active phonons in the quan- This finding suggests an intriguing possibility of inducing the tum paraelectric, electromagnetic fields in the cavity, and plas- superradiant quantum phase transition by utilizing the light- mons in the metallic electrodes. We show that the spectrum of matter interaction with the vacuum electromagnetic field. 3

the vacuum fluctuations) and vibrational excitations of indi- vidual molecules has been experimentally realized [118], and its nontrivial influence on the has recently been reported [119]. In this context, one of the main novelties in the present work is an accurate analysis taking into account the dispersive na- ture of polariton excitations in the dielectric of finite thick- ness as well as the surface plasmonic effects occurring at the metal-insulator interfaces, i.e., a finite plasma frequency of the metallic cavity mirrors. We demonstrate that the proper con- sideration of these aspects is important to accurately under- stand several key features in the strong light-matter coupling regime, which are not readily attainable in the conventional bulk settings. We point out that the cavity-induced changes of the polariton dispersions provide a general route towards con- trolling energy-level structures not only in paraelectric materi- als, but also in many other systems possessing electric dipole moments, such as molecular solids, organic light-emitting de- vices, and excitonic systems. Our analysis can further be ex- FIG. 2. (Top) Setup of the heterostructure cavity configuration considered in the present manuscript. A slab of paraelectric mate- tended to include fabricated surface nanostructures [120, 121]. rial with a finite thickness is sandwiched between two semi-infinite In view of the versatility of the present formulation, we antici- metallic electrodes. Right panel represents spatial profiles of typical pate that our results will be useful to advance our understand- energy modes along the spatial direction perpendicular to the metal- ing of a variety of physical systems lying at the intersection insulator interfaces. (Bottom) Schematic figure illustrating three fun- of condensed matter physics, quantum optics, and quantum damental excitations relevant to the theory. A phonon excitation de- chemistry. scribes an optical collective mode induced by, e.g., the deformation The remainder of the paper is organized as follows. In of ionic crystals around the equilibrium configuration. More gen- Sec.II, we summarize our main results at a nontechnical level. erally, it can also represent dipole moments of molecular solids or In Sec. III, we present a theoretical framework for describing assembled organic molecules. The interaction between phonons and light-matter hybrid collective modes in both bulk and cavity cavity photons is mediated by the dipole coupling, whose strength is represented by g. A , a collective excitation of electronic settings. As a first step, we neglect phonon nonlinearities and gases in metal mirrors, leads to the electromagnetic coupling charac- determine dispersion relations of elementary excitations in the terized by the plasma frequency ωp. heterostructure cavity configuration. In Sec.IV, we include phonon nonlinearities based on the variational approach and analyze their effect on the paraelectric-to-ferroelectric phase transition in the cavity. We show that the mode confinement in Before concluding this section, we put our work in a the cavity geometry induces significant softening of polariton broader context of studies on quantum matter strongly inter- modes, indicating the enhanced propensity for ferroelectricity. acting with electromagnetic fields. Importantly, our model We present a phase diagram of the paraelectric-to-ferroelectric bears a close connection with physical systems discussed transition in the cavity geometry. In Sec.V, we present a sum- in several areas, including cavity quantum electrodynamics mary of results and suggest several interesting directions for (QED), plasmonics, and polaritonic chemistry. Firstly, in the future investigations. field of cavity QED, one is typically interested in small quan- tum systems coupled to cavity photons [35, 59–64]; promi- nent examples include individual atoms [65–67], quantum dots [68–73], and light-emitting defects in solids such as II. SUMMARY OF MAIN RESULTS nitrogen- and silicon-vacancy centers [74–80]. These systems are applicable for realizing single-photon transistors [81–83] We first summarize the main findings of this work at a non- and sources [84–87], and can thus provide a promising route technical level before proceeding to the detailed theoretical to enabling photonic quantum information processing [88– formulation in the subsequent sections. The main focus of this 90]. Achieving a strong light-matter coupling has also been paper is on many-body systems that are naturally close to the the subject of intense research in the fields of plasmonics in superradiant-type transition, i.e., paraelectric materials in the nanostructures [91–93] and polaritonic chemistry [94–99]. In vicinity of the quantum phase transition into the ferroelectric both areas, cavity setups have been discussed as a platform phase (see Fig.1(a-c)). For instance, SrTiO 3 and KTaO3 are to modify a broad range of physical properties, including su- paraelectric at ambient conditions, but isotope substitution of perconductivity [100–102], charge and energy transport [103– 18O for 16O or applying pressure can bring the materials into 110], hybridized excitations in molecular crystals [111–114] a phase with spontaneous electric polarization [122]. and light-harvesting complexes [115], and chemical reactivity The key property of such materials relevant to our analysis of organic compounds [116, 117]. Strong coupling between is the existence of a low-energy infrared optical phonon mode. the zero-point fluctuations of the electromagnetic field (i.e., A conventional way to approach the quantum critical point 4 is to vary a tuning parameter (such as pressure or chemical excitations included in our analysis is plasmons, i.e., collec- composition) in order to decrease energy of this phonon mode. tive excitations of electronic density in metal mirrors. Since Transition into the broken symmetry phase occurs when the plasmons describe fluctuations of electric charges, they are in- minimum energy of the phonon dispersion goes down to zero herently coupled to the electromagnetic oscillations, which in as indicated by the red dashed line in Fig.1(a). the bulk system result in plasma resonance at frequency ωp. Our resonator setting facilitates hybridization of the elec- An important feature of plasmons relevant to our analysis is tric dipole moment of the optical phonon with the quantized that they can form surface modes localized at the interfaces electromagnetic modes confined in the cavity; the resulting that are known as the polaritons [91]. This polariton modes further couple to plasmons in metal mirrors feature originates from the fact that electric fields can be re- (see Fig.2). A salient feature of our analysis is the emphasis flected by the metal interfaces in such a way that the fields on quantum nature of the light-matter coupling, which arises take maximum values at the boundaries. As schematically from strong zero-point fluctuations of the vacuum electromag- illustrated in Fig.3, the resulting hybridized dispersions ex- netic fields in the cavity, while a similar argument also applies hibit qualitative changes in the nature of cavity polariton ex- to an equilibrium state at finite temperature. Placing the fo- citations compared to the bulk material. Altogether, the rich- cus on equilibrium systems makes the present work distinct ness of the cavity energy modes arises from the intrinsic hy- from the previous studies on materials subject to classical light bridization of different types of excitations, including infrared waves, in which the use of strong external drive is essential. phonons in the paraelectric, electromagnetic fields in the cav- It is such quantum aspect of light-matter interaction that mo- ity, and plasmons in the metallic electrodes. tivates us to ask two questions as follows: To answer the first question (A) in a concrete manner, we begin our analysis with a detailed discussion of the quadratic (A) How does the light-matter coupling modify polariton theory, which allows us to identify all the hybridized eigen- modes of paraelectric materials in the cavity geometry? modes in the cavity geometry. Interestingly, even when the system is far from the transition point, the cavity confinement (B) Is it possible to enhance the ferroelectric instability by can qualitatively alter the nature of collective excitations (see modifying quantum electromagnetic environment of a e.g., Fig.6 in Sec. III). We emphasize that, while our theory many-body system under no external pump? is formulated for a heterostructure cavity configuration, the To address these questions, we formulate a simple yet present approach of controlling material excitations via mod- generic theoretical framework including three essential ingre- ifying quantum electromagnetic environment can readily be dients: infrared active phonons, cavity photons, and plasmons generalized to a wide range of platforms. More specifically, in metal mirrors. To be concrete, we focus on the prototypi- besides quantum paraelectrics, we also discuss several po- cal heterostructure configuration that consists of a paraelectric tential applications of such cavity-based control of polariton material sandwiched between large metallic electrodes (see modes to molecular materials and excitonic devices, such as the top panel in Fig.2). One of the notable features of this sys- tem is that energies can be confined in volumes much smaller than the free-space wavelengths of photons, leading to strong (a) Bulk (b) Cavity configuration enhancement of light-matter interaction [123]. Strong hy- bridization between the continuum of electromagnetic fields ω ω … and matter excitations, including both phonons and plasmons, necessitates analyzing new collective modes that are absent in the bulk geometry. We assume that the size of the cav- ity is large enough for collective eigenmodes to be charac- Ω terized by a conserved in-plane momentum. In this respect, the present consideration is distinct from most of the previous studies on the superradiant phases, where only one or at most k q a few modes in a cavity have been included. In our theory, phonon modes possess electric dipole mo- FIG. 3. Nontechnical summary of the cavity-induced changes in col- ments and are described by the real-valued vector quantum lective excitations of a dipolar insulator. (a) Bulk band dispersions. field φˆ. Physically, this field can correspond to the ampli- The hybridization between a photon having the linear dispersion and an optical phonon having the nondispersive band at the frequency tude of the optical phonon, which describes a distortion of an Ω (black dashed lines) leads to two gapped branches correspond- ionic crystal leading to electric polarization. We note that the ing to the upper and lower transverse phonon polaritons (black same formalism can be applied to systems with other polar- curves). (b) Collective excitations of a dipolar insulator confined in izable modes such as collective dipole moments in assembled the planar cavity as a function of the in-plane wavevector q. The organic molecules or molecular solids. Interaction between hybridization of dipolar phonons, cavity photons, and plasmons in polarizable modes and cavity electric fields is characterized metal mirrors, leads to softening of the optical phonon modes (red by the light-matter dipole coupling whose strength is repre- curves and black solid arrow) and also supports the coupled surface sented by g (see Sec. III for more detailed discussions). This modes (blue curve). The former can be applied to enhance the ferro- coupling results in the formation of the collective excitations electric instability (see Sec.IV) while the latter can be used to realize known as the phonon polaritons. Another class of elementary the photon- conversion (see Sec.V and App.D). 5 improving the efficiency of organic light-emitting devices and A. Bulk realizing photon-exciton converters (see Sec.V and App.D). The inclusion of full complexities appropriate for specific ex- Before moving on to the case of the cavity setting, we perimental systems will require further analyses beyond the first discuss the three-dimensional bulk setup with no bound- present work. ary effects and determine the bulk dispersion relations for the To address the second question (B), we demonstrate that the phonon-photon hybridized modes, i.e., the phonon polaritons. resonator setup in Fig.2 provides yet another route towards Later, we generalize the theory to the cavity configuration by controlling the phase of matter via its coupling to quantum taking into account confinement effects and also contributions fluctuations of electromagnetic fields. Building upon the iden- from plasmons in metal mirrors. tification of collective modes detailed in Sec. III, we show that The total system of the bulk setting is governed by the the hybridization with the cavity electromagnetic fields can Hamiltonian lead to significant softening of the dipolar phonon modes, and ˆ ˆ ˆ ˆ ultimately induce the paraelectric-to-ferroelectric phase tran- Htot = Hlight + Hmatter + Hl−m, (1) sition when the softened phonon mode goes down to the zero where the first (second) term describes the dynamics of the frequency (see Fig.3(b)). In particular, we find that this tran- electromagnetic field (dipolar matter field), and the last term sition to the ferroelectric phase is possible even when the cor- represents the light-matter coupling between these two fields. responding bulk system is still within the paraelectric phase. We specify each of those terms below. In free space, phonon nonlinearities render the formation of a spontaneous dipole moment energetically unfavorable. The role of the cavity here is to alleviate this adverse effect of 1. Electromagnetic field phonon-phonon interactions, thus allowing one to induce fer- roelectricity in quantum paraelectric materials that otherwise The free evolution of the electromagnetic field is described exhibit no spontaneous polarizations even at zero temperature. by the Hamiltonian Our results thus indicate an intriguing possibility of induc- " # Z ˆ 2 2 2 ing the genuine superradiant-type quantum phase transition 3 Π 0c   Hˆlight = d r + ∇ × Aˆ , (2) via the light-matter coupling with no external pump. 20 2 We note that this phenomenon is not present at the level of a quadratic model, and can only be found when phonon where 0 is the vacuum permittivity, c is the vacuum light ˆ ˆ nonlinearities are included into the analysis. In the case of speed, A(r) is the vector potential of photons, and Π(r) is bulk quantum paraelectrics with no cavity confinement, the its canonically conjugate variable. The integral is taken over a 3 crucial role of nonlinear effects at the verge of the quantum cubic whose volume is V = L and we choose the Coulomb phase transition has previously been discussed in Ref. [58]. gauge We take into account the nonlinear effects in the cavity geom- ∇ · Aˆ = 0. (3) etry on the basis of the variational approach, leading to renor- malization of the hybridized collective excitations. The shift Assuming the periodic boundary conditions and introducing of the paraelectric-to-ferroelectric transition point can directly discrete wavenumbers ki = 2πni/L with i = x, y, z and ni be related to this renormalization of energy dispersions. We being integer numbers, we represent the vector fields in terms demonstrate that the phenomenon of cavity-enhanced ferro- of their Fourier components as electricity is more pronounced when confinement effects are r ˆ X ~ ik·r  enhanced by using a cavity geometry with a thinner thickness A(r) = aˆkλe kλ + H.c. , (4) 20V ωk or a higher plasma frequency (see Fig.9(b) in Sec.IV). kλ r X 0ωk Πˆ (r) = −i ~ aˆ eik·r − H.c. , (5) 2V kλ kλ III. PHONON-PHOTON-PLASMON HYBRIDIZATION kλ where ωk = c |k| is the vacuum photon dispersion, and aˆkλ † Our main goal is to understand the qualitative physics of a (aˆkλ) is the annihilation (creation) operator of photons with dipolar insulator coupled to quantized electromagnetic fields momentum k and polarization λ. The Coulomb gauge (3) in a cavity rather than to make predictions specific to particu- leads to the transversality condition of the orthonormal polar- lar experimental setups. In this section, we outline the general ization vectors kλ: theory of the light-matter interaction and formulate a frame- k · kν = 0, (6) work to reveal generic features in the hybridization of col- †  = δ . lective dipolar modes, cavity photons, and plasmons in metal kλ kν λν (7) mirrors. While the theory does not feature full complexities of The Hamiltonian can be diagonalized in the basis of momen- an experimental system, it applies to a broad range of materi- tum k and polarization λ as als supporting a polarizable vector field as matter excitations. ˆ X † In the next section, we will apply the framework to an ionic Hlight = ~ωkaˆkλaˆkλ, (8) crystal confined in the cavity as a concrete case and analyze kλ its paraelectric-to-ferroelectric phase transition. where we omit the vacuum energy. 6

ˆ⊥ ˆ⊥† 2. Matter field where N = V/v is the number of modes, φkλ (φkλ) is the an- nihilation (creation) operator of transverse phonons with mo- ˆk ˆk† We consider a dipolar insulator whose excitation is repre- mentum k and polarization λ, and φk (φk ) is the annihilation sented by a real-valued vector field φˆ(r). For instance, it can (creation) operator of longitudinal phonons. We recall that the describe an infrared optical phonon mode induced by the col- transverse polarization vectors kλ satisfy Eqs. (6) and (7). At lective deformation around the equilibrium configuration in the quadratic level (U = 0), only the transverse components ionic crystals such as MgO, SiC and NaCl [124]. In other couple to the dynamical electromagnetic field while the lon- examples, it can represent collective dipolar modes associ- gitudinal components couple to the static longitudinal electric ated with assembled organic molecules, molecular crystals, field as discussed below. In the similar manner as in the elec- ˆ ˆ and excitonic systems. We focus on a nondispersive matter tromagnetic Hamiltonian Hlight, the quadratic part H0 of the excitation with frequency Ω as appropriate for, e.g., an optical matter Hamiltonian Hˆmatter can be diagonalized as phonon in dipolar insulators or a collective dipolar mode in " # molecular materials. The effective Hamiltonian of the matter X ⊥† X k† k Hˆ = Ω φˆ φˆ⊥ + φˆ φˆ . (16) field is thus given by 0 ~ kλ kλ k k kλ k

Hˆmatter = Hˆ0 + Hˆint, (9) 3. Light-matter coupling where Hˆ0 is the quadratic part

Z d3r  πˆ 2 1  The transverse dipolar matter field φˆ⊥ can directly couple Hˆ = + MΩ2φˆ2 , (10) 0 v 2M 2 to the electromagnetic field in a bilinear form. For instance, we recall that the field φˆ⊥ can represent transverse displace- and Hˆint is the most relevant interaction ments of ionic charges with opposite signs and thus can cou- ple to the electric field. In terms of the vector potential Aˆ, the Z d3r  2 light-matter interaction is given by Hˆ = U φˆ · φˆ . (11) int v Z 3  ∗ ∗ 2  d r Z e ⊥ (Z e) 2 Hˆl−m = − πˆ · Aˆ + Aˆ + VC, Here, πˆ(r) is a conjugate field of the matter field φˆ(r), v is v M 2M the unit-cell volume in the insulator, M is the effective mass, (17) which corresponds to the reduced mass of two ions in the ∗ case of ionic crystals, and U ≥ 0 characterizes the interaction where Z e is the effective charge of a dipolar mode, which strength. Physically, the interaction term can originate from, corresponds to the charge of ions in the case of ionic crystals, e.g., phonon anharmonicity. When we will discuss in SecIV and VC is the Coulomb potential term whose explicit form is the transition between the paraelectric phase with hφi = 0 specified below. We note that, aside a constant factor, the con- and the ferroelectric phase with hφi= 6 0, the phonon fre- jugate field πˆ physically corresponds to the current operator 2 ˆ quency Ω should be considered as the tuning parameter gov- j ∝ ∂tφ associated with the dipolar field φ. erning the transition in the effective theory (cf. Fig.1). We We here invoke neither the dipole approximation nor the note that in the case of ionic crystals the present model ne- rotating wave approximation that are often used in literature, glects the coupling between electrons and soft phonons, which but can break down especially in strong-coupling regimes is typically considered as weak [125]. [81, 126]. The light-matter interaction (17) including both 2 The matter field φˆ(r) and its conjugate variable πˆ(r) πˆ · Aˆ and A terms allows one to explicitly retain the gauge can be quantized through the canonical quantization. In the invariance of the theory. It is also noteworthy that the light- Coulomb gauge, it is useful to decompose them in terms matter coupling (17) is nonvanishing even in the absence of of the transverse and longitudinal components as φˆ(r) = photons. This interaction with the vacuum electromagnetic φˆk(r) + φˆ⊥(r) and πˆ(r) = πˆ k(r) + πˆ ⊥(r), leading to the field originates from the zero-point fluctuations of the electric expressions field and can be viewed as the light-matter interaction medi- ated by virtual photons. In this paper, we focus on such a low r photon, quantum regime with no external pump. X   φˆ⊥(r) = ~ φˆ⊥ eik·r + H.c. , (12) Our choice of the Coulomb gauge leads to the nondynam- 2MNΩ kλ kλ kλ ical contribution VC resulting from the constraint relating the r longitudinal component of the electric field to that of the mat- ⊥ ~MΩ X  ˆ⊥ ik·r  πˆ (r) = −i φ e kλ − H.c. , (13) ter field. To see this explicitly, we first decompose the electric 2N kλ kλ field in terms of the transverse and longitudinal components r   as ˆk ~ X ˆk ik·r k φ (r) = φke + H.c. , (14) 2MNΩ |k| ˆ ˆ⊥ ˆk k E = E + E , (18) r   MΩ X k k ˆ⊥ ˆk πˆ k(r) = −i ~ φˆ eik·r − H.c. , (15) which satisfy ∇ · E = 0 and ∇ × E = 0. On one hand, the 2N k |k| k transverse part can be directly related to the vector potential 7 as then be simplified as    ⊥P  ˆ ⊥ 1 X  1 0 πˆ−kλ ˆ⊥ ∂A Hˆ = πˆ⊥P AˆP E = − . (19) tot,0 2 kλ kλ 0 1 AˆP ∂t kλ −kλ 1  2  πˆ⊥  On the other hand, the longitudinal part is subject to the con- X ⊥ ˆ  Ω −gΩ −kλ + πˆkλ Akλ 2 2 ˆ . 2 −gΩ ωk + g A−kλ straint resulting from the Coulomb gauge kλ Z∗e (28) Eˆk = − φˆk, (20) 0v We here introduce the coupling strength g as s ˆk (Z∗e)2 where we recall that φ is the longitudinal part of the matter g = . k (29) field satisfying ∇ × φˆ = 0. This longitudinal component 0Mv leads to the expression of the Coulomb potential term V as C The dispersion relations of the transverse modes can now be follows [127]: obtained by diagonalizing the matrix in the second term in Z 2 ∗ 2 Z 3 2 Eq. (28). The resulting Hamiltonian for the transverse com- 0 3  k (Z e) d r  ˆk VC = d r Eˆ = φ . (21) ponents is 2 20v v ˆ ⊥ X ⊥† ⊥ Htot,0 = ~ωksγˆksλγˆksλ, (30) ksλ 4. Bulk dispersions where s = ± denotes an upper or a lower bulk dispersion with eigenfrequencies One can diagonalize the quadratic part of the total Hamil- v u q tonian, u 2 2 2 2 2 2 2 2 2 tωk + Ω + g ± (ωk + Ω + g ) − 4ωkΩ ωk± = , Hˆtot,0 = Hˆlight + Hˆ0 + Hˆl−m, (22) 2 (31) which allows one to obtain an analytical expression of the bulk dispersions. The presence of light-matter interaction leads to

the formation of the hybridized modes consisting of photons ]

and phonons, known as the phonon polaritons. This hybridiza- Ω [ tion strongly modifies the underlying dispersion relations such that the avoided crossing occurs between two polariton disper- sions. At the quadratic level, we can decompose the quadratic Hamiltonian Hˆtot,0 in Eq. (22) into two parts that solely in- clude either transverse or longitudinal components of the field operators as follows: ω

ˆ ˆ ⊥ ˆ k Htot,0 = Htot,0 + Htot,0. (23)

To diagonalize the transverse part of the quadratic Hamilto- ˆ ⊥ nian Htot,0, it is useful to introduce the conjugate pairs of vari- ables for the field operators πˆ ⊥ and Aˆ in the Fourier space:

r   πˆ⊥ = −i ~ φˆ⊥ − φˆ⊥† , (24) kλ 2Ω kλ −kλ r k [Ω/c] Ω   πˆ⊥P = ~ φˆ⊥ + φˆ⊥† , (25) kλ 2 kλ −kλ FIG. 4. Bulk dispersions of the phonon-photon hybridized sys- r tem plotted against the norm of the three-dimensional wavevector ˆ ~  †  Akλ = aˆkλ +a ˆ−kλ , (26) k = |k|. The solid curves represent the eigenvalues of the transverse 2ωk modes (cf. Eq. (31)) for the upper (s = +) and the lower (s = −) r P ~ωk  †  phonon polaritons, which are obtained by diagonalizing the trans- Aˆ = −i aˆ − aˆ , (27) ˆ ⊥ ˆ kλ 2 kλ −kλ verse part Htot,0 of the quadratic light-matter Hamiltonian Htot,0. The solid horizontal line represents the nondispersive longitudinal where the superscript P indicates that an operator is a conju- phonon mode (cf. Eq. (32)) obtained by diagonalizing the longitudi- ˆ k gate variable of the corresponding operator, i.e., each pair of nal part Htot,0. The dotted lines show the photon dispersion ω = ck conjugate variables satisfies the canonical commutation rela- and the nondispersive bare phonon frequency at ω = Ω. The param- tions. The transverse part of the quadratic Hamiltonian can eter is chosen as g = 3.5 Ω. 8

⊥ ⊥† and γˆksλ (γˆksλ) represents an annihilation (creation) opera- configurations as a quantum electrodynamic setting towards tor of a transverse phonon-photon hybridized mode with mo- controlling collective matter excitations and, in particular, to mentum k, polarization λ, and mode s. The solid curves in point out a possibility of inducing the superradiant-type quan- Fig.4 show these two branches of the bulk phonon polaritons. tum phase transition via the vacuum electromagnetic environ- The splitting between the upper and lower dispersive modes ment. More specifically, analyzing the hybridization of cav- is characterized by the coupling strength g. The upper branch ity electromagnetic fields and matter excitations, we demon- exhibits a quadratic dispersion at low k = |k| and is gapped, strate that the present cavity architecture enriches the under- i.e., it takes a finite value pΩ2 + g2 at k = 0, while the lower lying dispersion relations in a way qualitatively different from one linearly grows from zero and saturates at Ω even in the the corresponding bulk case. Including phonon nonlinearities, limit of k → ∞. we will further show that the heterostructure configuration en- Meanwhile, the longitudinal part of the total quadratic ables one to significantly soften the phonon modes, which ul- ˆ k timately induces the structural phase transition. Hamiltonian Htot,0 only contains the longitudinal matter field φˆk and can be readily diagonalized as 1. Hamiltonian ˆ k k X ˆk† ˆk Htot,0 = ~Ω φk φk, (32) k To reveal essential features of the present cavity setting, we where Ωk = pΩ2 + g2 is the longitudinal phonon frequency consider an ideal metal with the Drude property whose plasma p 2 [128]. We note that the longitudinal mode is independent of frequency is denoted as ωp = nee /(me0), where e is the momentum k and is thus nondispersive as shown in the the elementary electric charge, ne is the electron density in solid horizontal line in Fig.4. the metal, and me is the electronic mass. We assume that all the materials are nonmagnetic such that the relative magnetic permeability µ is always taken to be µ = 1. As illustrated B. Cavity configuration in Fig.5, we consider a setup with the center of a dipolar insulator being positioned at z = 0 while the insulator-metal z = ±d/2 We next consider the cavity setting and take into account interfaces being positioned at . We again choose the ˆ ˆ confinement effects as well as the hybridization of electro- Coulomb gauge ∇·A = 0 and thus the Hamiltonian Hlight of magnetic fields with plasmons in metal mirrors in addition the free electromagnetic field is the same as in the bulk case to collective dipolar modes. We focus on the prototypical (cf. Eq. (2)) while the matter Hamiltonian is modified as heterostructure configuration consisting of a dipolar insula- Hˆ = Hˆ + Hˆ , (33) tor sandwiched between two semi-infinite ideal metals (see matter 0 int Z 3  2  Fig.5). We consider eigenmodes propagating with the two- d r πˆ 1 2 ˆ2 Hˆ0 = + MΩ φ , (34) dimensional in-plane momentum q while the formalism de- I v 2M 2 veloped in this section can be extended to deal with more Z d3r  2 Hˆ = U φˆ · φˆ , (35) complex setups such as fabricated metal surface structures. int v Thin-film configurations considered here have been pre- I viously applied to light-emitting devices [129, 130], the where we define the integral over the insulator region as electron-tunneling emission [131], and the transmitting R 3 R 2 R d/2 I d r ≡ d r −d/2 dz. The light-matter coupling term is waveguide [132]. One of the main novelties introduced by also modified as this paper is to reveal the great potential of heterostructure Z 2 3 0ωp 2 Hˆl−m = d r Aˆ M 2 Z 3  ∗ ∗ 2  d r Z e ⊥ (Z e) 2 + − πˆ · Aˆ+ Aˆ +VC, (36) I v M 2M where the first term describes the plasmon coupling in the R 3 R 2 R −d/2 R ∞  metal regions defined as M d r ≡ d r −∞ + d/2 dz, and the second term is the coupling between the insulator and the cavity electromagnetic field. We recall that the Coulomb gauge leads to the additional constraint on the longitudinal components of the electric and matter fields in the insulator region |z| < d/2: Z∗e FIG. 5. Schematic figure illustrating the cavity configuration con- Eˆk = − φˆk. (37) sidered in this paper. A dipolar insulator whose center is positioned 0v at z = 0 is sandwiched between two semi-infinite metals with inter- As in the bulk case, this constraint results in the Coulomb faces at z = ±d/2. potential term VC originating from the longitudinal dipolar 9 modes as follows: Meanwhile, the parity label P indicates the symmetry of the Z 2 ∗ 2 Z 3 2 mode function with respect to the spatial inversion along the 0 3  k (Z e) d r  ˆk VC = d r Eˆ = φ . (38) z direction. From the parity symmetry of the Maxwell equa- 2 I 20v I v tions, it follows that the z component of the mode function Summarizing, the present cavity light-matter system con- must have the opposite parity to that of the x, y components sists of two canonically conjugate pairs of variables (Πˆ , Aˆ) (see AppendixB). We thus assign the parity label P to each mode function according to and (πˆ, φˆ); the former describes the dynamical electromag- netic degrees of freedom and only have transverse compo- ⊥  ⊥  uqnλ(−z) = P uqnλ(z) , (43) nents while the latter describes matter excitations and have x,y x,y u⊥ (−z) = −P u⊥ (z) . (44) both transverse and longitudinal parts. Their time evolution is qnλ z qnλ z governed by the Hamiltonian consisting of Eqs. (2), (33) and Finally, we can impose the orthonormal conditions on the (36), and is subject to the constraint (37). mode functions as We note that the use of the Coulomb gauge remains to be Z a convenient choice even in the present case of the inhomo- ⊥† dz u u⊥ = δ δ . (45) geneous geometry. An alternative naive choice would be the qnλ qmν nm λν ∇ · [(r, ω)Aˆ(r, ω)] = 0  condition with being the relative We can expand the matter degrees of freedom in the similar permittivity in the medium, which might be viewed as a quan- manner as in the electromagnetic field. Because the matter tum counterpart of the gauge constraint in the macroscopic field resides only in the insulator region, it is natural to expand Maxwell equations. However, it has been pointed out that in the transverse components of the dipolar matter field as inhomogeneous media this gauge choice suffers from the diffi- culty of performing the canonical quantization of the variables r ˆ⊥ ~ X  ˆ⊥ ⊥ iq·ρ  due to the frequency-domain formulation of the gauge con- φ (r)= φqnλfqnλ(z)e + H.c. , 2MNdΩ straint [127]. In contrast, the Coulomb gauge chosen here can qnλ be still imposed on a general inhomogeneous system without (46) such difficulties; we thus employ it throughout this paper. where Nd = A/v = N/d is the number of modes per thick- ⊥ ness, and {fqnλ} is a complete set of transverse orthonormal ⊥ iq·ρ 2. Elementary excitations functions in the insulator region satisfying ∇ · (fqnλe ) = 0. We note that a position variable r in the matter field φˆ The quadratic part of the total cavity Hamiltonian can still should be interpreted as the parameters restricted to the region be diagonalized, which allows us to identify the hybridized |z| < d/2. In the discrete variable λ = (Λ,P ), the polariza- eigenmodes and the corresponding elementary excitation en- tion label takes either TM or TE modes, Λ ∈ {TM, TE}, and ergies. To this end, we first expand the electromagnetic vector the parity label P = ± is defined in the same manner as in potential in terms of the transverse elementary modes. Owing Eqs. (43) and (44). Another possible matter excitations are to the spatial invariance on the xy plane, we obtain the longitudinal modes, for which we can expand the matter s field as ˆ X ~ ⊥  r A(r)= aˆqnλUqnλ(r)+H.c. , (39) ˆk ~ X  ˆk k iq·ρ  20Aωqnλ φ (r) = φqnλfqnλ(z)e + H.c. , qnλ 2MNdΩ qnλ ⊥ iq·ρ ⊥ Uqnλ(r) = e uqnλ(z), (40) (47) where A = L2 is the area of interest on the xy plane, q is the k T where {fqnλ} is a complete set of longitudinal orthonor- two-dimensional in-plane wavevector q = (qx, qy) , and ρ = T mal functions in the insulator region, which satisfies ∇ × (x, y) is the position vector on the xy plane. The transverse k iq·ρ ⊥ ⊥ (fqnλe ) = 0. We denote this mode by the label λ = mode function Uqnλ satisfies ∇ · Uqnλ = 0. Because the spatial invariance along the z direction is lost (L,P ), where the polarization label Λ is fixed to be the longi- in the heterostructure configuration, an elementary excitation tudinal one Λ = L while the parity P = ± of the mode func- tion is defined in the same way as in the transverse case. The is now labeled by a discrete number n ∈ N (instead of qz in the bulk case) as well as the two-dimensional vector q. We longitudinal mode associates with an excitation of the longi- also introduce a discrete variable λ = (Λ,P ) that includes the tudinal electric field via Eq. (37). ⊥ polarization label Λ ∈ {TM, TE} and the parity label P = An explicit form of the transverse mode functions {uqnλ} ⊥ k ±. Here, the polarization label Λ in the discrete subscript λ and {fqnλ}, the longitudinal mode functions {fqnλ}, and indicates either the transverse magnetic (TM) or electric (TE) the corresponding excitation energies {ωqnλ} can be deter- polarization associated with the zero magnetic or electric field mined from solving the eigenvalue problem with imposing in the z direction: suitable boundary conditions at the interfaces as detailed in AppendixB. In the transverse modes, we note that both mat- u⊥ (z) = 0 for Λ = TM, (41) qnλ z ter and light fields as well as the hybridized eigenmodes share ∇ × U ⊥ (r) = 0 for Λ = TE. (42) u⊥ qnλ z the same spatial profile represented by qnλ owing to the 10

(a) (TM, +), (L, ±) (b) (TM, ー) (c) (TE, ±) ] ] ] Ω Ω Ω [ [ [ TM TM1 2 TE 1

SS (=TM 0 ) AS2 TE 0 ω ω L ± ω

RTE RTM n ∈even RTMn ∈odd n

AS1 q [Ω/c] q [Ω/c] q [Ω/c]

(d) (TM, +), RTM n ∈even (e) (TM, ー), RTM n ∈odd (f) (TE, ±), RTE n … ] … ] ] … Ω Ω Ω [ RTM4 [ [ RTM5 … RTM2 RTM3 RTE2

ω RTE

ω ω 1

RTM1 RTM0

RTE0 q [Ω/c] q [Ω/c] q [Ω/c]

FIG. 6. Elementary excitations of the cavity setting in different sectors (Λ,P ) with Λ =TM, TE, L denoting the transverse magnetic (TM), transverse electric (TE), or longitudinal (L) polarization and P = ± being the parity of transverse components of the electric field against the inversion of the direction perpendicular to the plane. The top panels plot excitation energies ω against in-plane momentum q in the sectors (a) (TM,+) and (L,±), (b) (TM,−), and (c) (TE,±), respectively. The bottom panels (d), (e), and (f) show the corresponding magnified views below the phonon frequency Ω. In (a) and (b), the TM1,2 modes represent the two highest modes in the TM sector, the SS (=TM0) mode represents the symmetric surface (SS) mode, the AS1,2 modes represent the antisymmetric surface (AS) modes whose dipolar moments point out of the plane at low q. The longitudinal L± modes have the same nondispersive energy independent of the momentum q and the parity P = ±. The assignment of the label TM0 to the SS mode is in accordance with the convention used in studies of plasmonics [133]. In (d) and (e), the RTMn modes with n = 0, 1, 2 ... represent the resonant TM modes close to the dipolar-phonon frequency Ω. The RTMn≥1 modes extend over the insulator region while the most-softened mode RTM0 is essentially the surface mode localized at the interfaces analogous to the SS mode. In (c), the TE0,1 modes represent the two highest modes in the TE sector. In (f), the RTEn modes with n = 0, 1, 2 ... represent the resonant TE modes; all of them extend over the insulator region. The parameters are d = c/Ω, g = 3.5 Ω, and ωp = 5 Ω. For the sake of visibility, the plasma frequency is set to be a rather low value while its specific choice will not qualitatively affect the influence of cavity on the ferroelectricity as discussed in Sec.IV.

† bilinear form of the light-matter coupling. Meanwhile, the where γˆqnλ (γˆqnλ) represents an annihilation (creation) op- longitudinal modes consist of only the matter field and their erator of a hybridized elementary excitation with in-plane spatial profiles are characterized in terms of the mode func- wavevector q and discrete labels n and λ. The spatial profile k of each excitation is characterized by an eigenmode function tions fqnλ. It is also noteworthy that an excitation energy ⊥ k ωqnλ of an elementary eigenmode now explicitly depends on uqnλ or fqnλ depending on the polarization, whose explicit the polarization Λ ∈ {TM, TE, L} in contrast to the bulk functional forms are given in AppendixB. case in Eq. (31). In terms of the obtained elementary eigen- Figure6 shows the elementary excitations in the cavity set- modes, the quadratic part of the total light-matter Hamilto- ting for each sector of λ = (Λ,P ), and TableI summarizes ˆ ˆ ˆ ˆ nian, Htot,0 = Hlight + H0 + Hl−m, can be diagonalized as the corresponding physical properties. Figures6(a) and (b) plot the in-plane dispersions in the sectors λ = (TM, +) and ˆ X † (L, ±) λ = (TM, −) Htot,0 = ~ωqnλγˆqnλγˆqnλ, (48) , and , respectively. Their magnified plots qnλ around the phonon frequency Ω are also given in Figs.6(d) 11

TABLE I. Summary of physical properties for each of elementary excitations. The labels for each band in the first column are consistent with those indicated in Fig.6. The second and third columns represent the parity P = ± and the polarization Λ ∈ {TM, TE, L} of each mode, respectively. The fourth column indicates the direction of the dipole moments in the insulator at low in-plane wavevector q with respect to the plane parallel to the interfaces. The fifth column shows whether or not the divergence of the electric field is vanishing. The sixth and seventh columns represent whether the longitudinal wavenumber κη takes a real value or a purely imaginary value at low and high q, respectively. The subscript η ∈ {q, n, λ} with λ = (Λ,P ) represents a set of all the variables specifying each eigenmode. The final two columns indicate the physical nature of each hybridized band at low and high q, where we abbreviate elementary excitations as plasmons (PL), photons (PT), phonons (PN), polaritons (P), and specify surface modes by S. κ Mode Band P Λ Dipole at low q Divergence η low q high q low q high q

TM2 + TM In-plane zero Real Real PL PT

TM1 − TM In-plane zero Real Real PL-P PT

RTMn≥1 sgn(n) TM In-plane zero Real Real PN-P PN

RTM0 + TM In-plane zero Real Imaginary PN-P S-PN-PL-P

L± ± L Out-of-plane nonzero Real Real PN-P PN-P

SS (=TM0) + TM In-plane zero Real Imaginary PN-P S-PN-PL-P

AS2 − TM Out-of-plane zero Imaginary Imaginary PN-P S-PN-PL-P

AS1 − TM Out-of-plane zero Imaginary Imaginary PT S-PN-PL-P

TE1 − TE In-plane zero Real Real PL PT

TE0 + TE In-plane zero Real Real PL-P PT

RTEn≥0 sgn(n) TE In-plane zero Real Real PN-P PN

and (e). These rich band structures in the TM polarization frequency value of the RTM0 mode at q → ∞ remains sub- have different physical origins. Below we provide descrip- stantially below Ω with a nonvanishing gap. Accordingly, the tions for each of elementary modes. RTM0 mode can be interpreted as the low-energy analogue of First of all, a salient feature of the elementary excitations the surface polariton mode labeled as the SS mode in Fig6(a); in the cavity setting is the emergence of a large number of both modes result from the intrinsic hybridization of dipolar soft-phonon modes lying below the bare phonon frequency phonons, cavity electromagnetic fields, and plasmons while Ω, which we term as the resonant TM (RTM) modes (see the major contribution at high q comes from the phonon (plas- Fig.6(d,e)). The RTM modes are labeled by an integer num- mon) part in the case of the RTM0 (SS) mode. ber n = 0, 1, 2 ... as RTMn. The parity P of the correspond- To gain further insights, in Fig.7 we plot the spatial pro- ⊥ ing mode function uqnλ coincides with the parity of n. The files of the electric field E of the RTM0 mode. Figures7(a) RTMn modes with n ≥ 1 lie between Ω and the lower bulk and (b) show the electrical flux lines of the RTM0 mode at low dispersion (plotted as the lower black dotted curve in Fig.6), and high in-plane momenta q, respectively, with the color in- and share the common structures. For instance, all of the dicating the magnitude of the out-of-plane component Ez of RTMn≥1 modes start from frequencies slightly below Ω and the electric field. We note that our convention of denoting the saturate to it at high q ≡ |q|. Thus, these modes predom- z-component as the out-of-plane one is in accordance with the inantly consist of the hybridization of dipolar phonons and planar cavity setting whose metal-insulator interfaces are par- cavity electromagnetic fields at low q while the phonon con- allel to the xy plane. Without loss of generality, here we focus tribution eventually becomes dominant at high q. The field on the mode propagating in the x direction, i.e. q = qx, lead- amplitudes of these RTMn≥1 modes extend over the insula- ing to the condition Ey = 0 (see AppendixB). Figures7(c) tor region owing to the real-valuedness of the longitudinal and (d) show the low-q and high-q spatial profiles of the in- wavenumber κη over the entire plane of the in-plane wavevec- plane (red solid curve) and the out-of-plane (blue solid curve) tor q. Here, the subscript η is used to summarize all the in- components of the electric field at x = 0. At low q before dices as η = {q, n, λ}. The origin of the emergence of many crossing the lower bulk dispersion, the field predominantly RTM modes below the bare phonon frequency can be under- points to the in-plane direction and extends over the insula- stood from simple physical arguments based on the frequency- tor region |z| ≤ d/2 (Fig.7(c)). In contrast, after crossing the dependent dielectric functions (see AppendixC). bulk dispersion, the longitudinal wavenumber κη turns out to Remarkably, the lowest resonant mode RTM0, which is the take a purely imaginary value and the field profiles qualita- most-softened phonon mode among all the elementary exci- tively change. Specifically, at high q the fields are localized tations in all the sectors, has a qualitatively different physical around the interfaces and the z component of the electric field nature compared to the RTMn≥1 modes discussed above. Its exhibits significant discontinuities due to the surface charges longitudinal wavenumber κη changes from a real value to a concentrated on the metal interfaces (Fig.7(d)); these features purely imaginary one as the dispersion crosses the lower bulk are characteristic of surface polariton modes. As shown later, dispersion in the bulk (see Fig.6(d)). Moreover, the saturated it is the softening of the RTM0 mode discussed here that will 12

(a) RTM0 low q (b) RTM0 high q ] ] d d [ [ z z

x [1/q] x [1/q] (c) (d) out-of-plane in-plane out-of-plane field field in-plane z [d] z [d]

FIG. 7. (a,b) Electrical flux lines and (c,d) spatial profiles of the mode function (i.e., the electric field) for the most-softened phonon mode denoted as the RTM0 mode. Without loss of generality, we consider the mode propagating in the x direction with the in-plane momentum q = qx and the vanishing electric field in the y direction Ey = 0. (a,b) The black solid curved arrows indicate the electrical flux lines on the xz plane at (a) low and (b) high in-plane momenta q. The color indicates the magnitude of the electric field Ez perpendicular to the interfaces. (c,d) The spatial profiles of the in-plane (red solid curve) and the out-of-plane (blue solid curve) components of the electric field at x = 0 in (a) and (b) are plotted in (c) and (d) in the arbitrary unit, respectively. The black dashed vertical lines indicate the interfaces at |z| = d/2. The parameters are d = c/Ω, g = 3.5 Ω, and ωp = 5 Ω. We set q = 0.1 Ω/c in (a,c) and q = 7 Ω/c in (b,d). trigger the structural instability of materials, thus ultimately This hybridized mode reduces to the bare plasmon (photon) inducing the paraelectric-to-ferroelectric phase transition. excitation in the limit of q → 0 (q → ∞). The next highest When the plasma frequency is close to the phonon fre- band (denoted as TM1 in Fig.6(b)) initiates from a frequency quency and the hybridization with plasmons becomes partic- slightly below the plasma frequency and has a physical origin ularly prominent, the above crossover behavior of the RTM0 similar to the TM2 mode, but there still exists a difference that mode also manifests itself as the emergence of roton-type ex- the TM1 mode sustains the nonvanishing hybridization with ⊥ citations. To see this explicitly, in Fig.8 we show the energy the electromagnetic fields at low q. The mode functions uη dispersion of the RTM0 mode at different light-matter cou- of the TM1,2 modes and the corresponding dipole moments pling strengths g with a low plasma frequency. It is evident predominantly point to the directions parallel to the interfaces that the energy dispersion takes the minimum value at finite at low q. The longitudinal wavenumber κη remains real over in-plane momentum q∗ > 0, indicating the formation of roton the entire in-plane momentum space, and thus the amplitudes excitations. Physically, this minimum roughly corresponds to of the TM1,2 modes extend over the insulator region (see also the vanishing point of the longitudinal wavenumber κη = 0, Fig. 11 in AppendixB). at which the crossover from low to high in-plane momentum We next discuss the symmetric and antisymmetric surface p 2 2 behavior occurs (compare Figs.7(c) with (d)). While a spe- modes lying in the frequency regime Ω + g < ω < ωp, ∗ cific value of the finite in-plane momentum q depends on which are denoted as SS(=TM0) and AS2 in Figs.6(a) and system parameters such as the thickness of the insulator, it is (b), respectively. They essentially arise from the split of the a general feature that roton excitations eventually disappear coupled surface phonon polaritons at two metal-insulator in- (i.e., q∗ → 0) as the plasmon component in the eigenmode terfaces [134]. We remark that the assignment of the label becomes less significant by, for instance, increasing either the TM0 to the SS mode is in accordance with the convention of light-matter coupling strength g or the plasma frequency ωp. studies in plasmonics [133]. At a high in-plane wavevector q, We now proceed to discuss physical properties of the other the SS and AS2 modes become bound modes that are local- elementary modes. The highest band (denoted as TM2 in ized around the interfaces; the field amplitudes take maxima Fig.6(a)) starting from the plasma frequency ωp at q = 0 is es- at the interfaces and exponentially decay away from them. In sentially the upper branch of the standard plasmon polaritons. the limit of q → ∞, the energy dispersions for both of the 13

hybridization of the collective dipolar mode, electromagnetic 0.94 g/Ω=0.5 fields in the cavity, and plasmons in metal mirrors, and can emerge only in the case of the TM polarization, in which the continuity of the z component of the electric field can be

] mitigated. Thus, the surface modes including the SS(=TM0),

Ω p

[ AS , and RTM modes can be excited only by -polarized 0.935 1,2 0 light. It is worthwhile to mention that in the context of plas- 0.73 monics the nanoscale localization of such surface modes has g/Ω=1.5 previously found applications to manipulate light waves be- low the diffraction limit [135, 136]. 0.72 The longitudinal modes L± in Fig.6(a) have the parity ω P = ± and lie at the constant frequency Ωk = pΩ2 + g2. 0.71 These nondispersive modes arise from the vanishing insulator permittivity and is characterized by the nonzero divergence of iq·ρ k g/Ω=4 the mode function ∇ · (e f η) 6= 0. Both the in-plane and 0.4 out-of-plane amplitudes of the longitudinal electric field ex- tend over the insulator region while the amplitudes vanish in the metal regions due to the boundary conditions. At low q, 0.39 the out-of-plane component becomes dominant in the insula- tor region due to the surface charges. 0 1 2 3 4 5 Finally, Fig.6(c) plots the TE-polarized bands of both par- q [Ω/c] ities P = ±. The corresponding magnified plot around the phonon frequency is also shown in Fig.6(f). These TE modes

FIG. 8. The RTM0 mode at different light-matter coupling strengths have simpler structures than the TM ones, i.e., each of all g with a low plasma frequency. The strong hybridization with plas- the modes lies slightly above the higher or lower branches of mons leads to the emergence of the minimum of the energy disper- the bulk dispersions (black dotted curves) due to energy en- sion at finite in-plane momentum q∗ > 0, indicating the formation of hancements by the cavity confinement. We label the higher roton-type excitations. The parameters are d = c/Ω and ωp = 2 Ω. set of the TE modes as TE0,1 while the lower set of a large number of resonant TE modes as RTEn=0,1,2.... Because the mode functions (and thus the electric fields) for all the modes SS and AS2 modes become flat and saturate to the same, fi- are continuous even at the interfaces, the origin of the en- nite frequency below ωp. Meanwhile, at q = 0 the AS2 mode hanced energies in these TE modes can simply be understood starts from a frequency lower than that of the SS mode, and as the acquisition of nonzero longitudinal momentum asso- thus the former exhibits a larger dispersion. The SS mode can ciated with the cavity confinement. Indeed, the correspond- have a real longitudinal wavenumber κη at low q and thus it ing longitudinal wavenumber κη remains real for all the TE can be radiative, i.e., it can be coupled to light modes outside modes over the entire region of the in-plane momentum q. the cavity when metal mirrors have finite thicknesses. As a This real-valuedness of κη also indicates that the TE modes result, the SS mode can contribute to optical properties of the are essentially radiative photonic modes that can directly cou- system. With increasing q, the longitudinal wavenumber κη ple to light modes outside the system when the thicknesses of changes from a real value to a purely imaginary value and the the metal mirrors are taken to be finite. SS mode changes into the nonradiative mode. In contrast, the Several remarks are in order. Firstly, we here emphasize AS2 mode is nonradiative over the entire in-plane momentum the importance of explicitly including plasmon contributions and can potentially be applied to nanoscale plasmonic waveg- into the analysis. This treatment ensures that the tails of the uides owing to its localized and dispersive nature. The dif- evanescent fields properly extend into the surrounding metal ference between two surface modes at low q also results in mirrors. In many of the previous studies on materials or as- their qualitatively different low-energy polarization behavior; sembled molecules confined in the quantum electromagnetic the dipole moment of the SS mode points to the in-plane di- environment, the mirrors are assumed to be perfectly reflec- rection while that of the AS2 mode is purely out-of-plane. tive such that the electric fields must exactly vanish at the in- The lowest mode in the TM polarization, which is denoted terfaces and are absent in the metal regions. In such analy- as AS1, lies slightly below the lower bulk dispersion (black ses, the physical nature of a multitude of the localized surface dotted curve) and converges to the linear photon dispersion modes discussed here cannot be addressed appropriately. For at q → 0 in contrast to the other modes featuring quadratic instance, the most-softened phonon mode RTM0 is one impor- asymptotes. This mode essentially shares the similar proper- tant example of such localized modes; an analysis assuming ties with the AS2 mode; both of them are localized modes over the perfect reflectivity cannot capture its unique features, in- the entire in-plane momentum q and point out of the plane at cluding the qualitative changes associated with increasing in- low q due to the surface charges. The AS1 mode asymptoti- plane momentum q (cf. Fig.7), the significant nonvanishing cally becomes a purely photon mode at q → 0. softening from the bare phonon frequency Ω even in the limit We note that these surface modes result from the intrinsic q → ∞ (cf. Fig.6(d)), as well as the emergence of roton-type 14 excitations (cf. Fig.8). Importantly, the correct treatment of effects through nonlinear interactions between the energeti- this RTM0 mode turns out to be crucial especially when we cally dense elementary excitations. discuss the superradiant-type quantum phase transition as de- tailed in the next section. Meanwhile, we remark that the re- flectivity does not qualitatively change physical properties of A. Variational principle the TE modes, which are analogous to the modes supported in Fabry-Perot cavities [137]. We are interested in how the coupling between the dipo- Secondly, we note that the present results can readily be lar phonon modes and the light modes confined in the cav- generalized to include further complexities in practical exper- ity affects physical properties of quantum paraelectrics on the imental situations. In particular, effects from losses due to border of ferroelectric order. The paraelectric-to-ferroelectric electron damping can be included by introducing phenomeno- phase transition is essentially a structural transition of the lat- logical imaginary parts in the permittivity of the metal mirrors. tice structure in, e.g., ionic crystals. The phonon field φˆ plays One can also use a mathematically equivalent, yet more rigor- the role of the order parameter, which spontaneously breaks ous approach by, e.g., considering coupling to bosonic reser- the symmetry in the ordered, ferroelectric phase. In ionic voir modes or relying on the macroscopic Green’s function crystals, this order is supported by the imbalanced charge con- formalism and the microscopic Hopfield approach. Yet, the figuration with the relative displacement, which accompanies effects found here can in principle outweigh the losses owing a uniform and nonzero electric polarization (cf. Fig.1(b)). to the strong cavity confinement. Thus, the transition is necessarily associated with the soft- Finally, we remark that the rich structures of the elemen- ening of an optical phonon mode while the acoustic phonon tary excitations demonstrated above can further be tuned by mode is irrelevant. varying the thickness d of the dipolar insulator. The increase We thus consider an effective model described by the total of d generally leads to a generation of more delocalized bulk Hamiltonian modes extending over the insulator. Thus, aside from the lo- Hˆ = Hˆ + Hˆ + Hˆ , (49) calized surface modes, the eigenmodes become denser and tot matter light l−m ˆ eventually converge to the corresponding bulk dispersions in where Hmatter now includes both the quadratic terms (34) and the limit of d → ∞. In contrast, the decrease of d leads to the most relevant interaction term (35), Hˆlight is the Hamil- the filtering of the delocalized modes; for instance, a degener- tonian for the free electromagnetic field (2), and Hˆl−m de- ate frequency of the TE0 and SS modes at zero-in-plane mo- scribes the light-matter coupling in the cavity setting as given mentum shifts up for a smaller d. Thus, with decreasing d, by Eq. (36). We note that here it is crucial to go beyond the lin- the localized nature of the modes plays more and more cru- ear regime and to include nonlinear effects because the system cial roles in determining the physical properties of the present is at the verge of the phase transition point and fluctuations can heterostructure configuration. We assess the effects of chang- be significant. ing the thickness in more detail in the next section by in- To qualitatively understand physical consequences of the cluding the most relevant nonlinear effect in a self-consistent cavity confinement, we perform the variational analysis on manner. The nonlinearity renormalizes the value of the bare the present nonlinear model. Specifically, we first replace the phonon frequency Ω and thus effectively modifies the under- 2 bare phonon frequency Ω in Hˆ0 (cf. Eq. (34)) by a system lying elementary excitations. We will see that a thinner in- tuning parameter r, where r ≥ 0 (r < 0) corresponds to the sulator is more preferable to realize the cavity-enhanced fer- paraelectric (ferroelectric) phase at the level of the noninter- roelectricity. The reason is that a thin heterostructure con- acting linear model. In practice, this tuning parameter can figuration can alleviate the adverse effect of the interaction be varied in various manners such as by changing densities, among the energetically dense phonon excitations and is thus chemical isotopic substitutions, or applying hydrostatic pres- advantageous for inducing the ultimate softening of the RTM0 sure. Accordingly, we denote the quadratic part of the total mode, which causes the ferroelectric instability or, namely, the Hamiltonian as Hˆ (r) and the full interacting Hamiltonian superradiant-type phase transition. tot,0 as Hˆtot(r) = Hˆtot,0(r) + Hˆint. We then represent the corre- sponding exact thermal equilibrium state as ˆ ˆ −βHˆtot(r) −β(Htot,0(r)+Hint) IV. CASE STUDY OF CAVITY-ENHANCED e e ρˆ(r) = = , (50) FERROELECTRIC PHASE TRANSITION Z(r) Z(r) ˆ where Z(r) ≡ Tr[e−βHtot(r)]. We then find the best approxi- As a concrete case study of the theoretical formulation in mation for this interacting equilibrium state by using a tuning the previous section, we here discuss the influence of the variable as a variational parameter r : quantum electrodynamic environment on the paraelectric-to- eff ferroelectric phase transition, which is the structural quantum min DKL [ˆρ0(reff )|ρˆ(r)] ≥ 0, (51) reff phase transition commonly observed in ionic crystals. Most notably, we find that in the cavity setting the transition point where DKL[·|·] is the Kullback-Leibler divergence [138], and can be shifted in favor of the ferroelectricity owing to the soft- an effective Gaussian state is defined as ˆ ening of the dipolar phonon mode. This softening results from e−βHtot,0(reff ) ρˆ0(reff ) = (52) the interplay between the modified dispersions and feedback Z0(reff ) 15

−βHˆtot,0(r ) with Z0(reff ) ≡ Tr[e eff ]. We introduce the variational free energy as D E Fv ≡ F0(reff ) + Hˆtot(r) − Hˆtot,0(reff ) , (53) eff where we denote an expectation value with respect to the effective density matrix as h· · · ieff = Tr[ˆρ0(reff ) ··· ], and 1 F0(reff ) = − β ln Z0(reff ) is the Gaussian variational free en- ergy. The variational condition (51) can then be simplified as [139]

min Fv ≥ F (r), (54) reff

1 where F (r) = − β ln Z(r) is the exact free energy. The opti- ∗ mal effective tuning parameter reff , which is defined by ∗ reff = arg min Fv, (55) reff gives the best approximation of the interacting model in terms of the linear theory. Within the present variational formal- ism, the phase transition point is determined from the condi- ∗ tion reff = 0 that can eventually be met as the bare tuning parameter r is decreased from a value corresponding to the paraelectric phase. In other words, the interaction feedback is introduced as the renormalization of the effective phonon ∗ 2 ˆ ∗ frequency reff = Ωeff in the variational theory Htot,0(reff ), and the transition occurs when this renormalized frequency of paraelectric materials goes down to zero. We note that in the present analysis the expectation value can be decoupled as in the standard mean-field analysis, and the theoretical proce- dure can be exact in the spherical model, where the number of components in the order parameter is taken to be infinite.

FIG. 9. (a) Variational free energies Fv plotted against the vari- B. Results ational parameter reff with decreasing tuning parameter r from top to bottom panels. The red and blue solid curves correspond to the We show the obtained variational free energy Fv against results for the cavity setting with the insulator thicknesses d = 1 the variational parameter reff at different insulator thicknesses and d = 3, respectively, while the green solid curves correspond to the bulk results. The minima of the variational energies determine d in Fig.9(a). From top to bottom panels, we decrease the ∗ 2 ∗ the equilibrium values reff = Ωeff of the effective phonon frequency bare tuning parameter r. The thermal equilibrium value reff of the effective phonon frequency is determined from identi- including the nonlinear feedback. (b) Phase diagram in the cavity set- ting. As the tuning parameter r is decreased, the transition from the fying the minimum of each variational free-energy landscape. paraelectric phase (PE) to the ferroelectric phase (FE) occurs when In all the panels, a thinner insulator thickness d in the cav- ∗ the equilibrium effective phonon frequency reff goes down to zero. ity setting leads to a smaller equilibrium phonon frequency The transition point in the cavity setting shifts in favor of FE com- ∗ reff , indicating the phonon softening owing to the coupling pared to the bulk transition point indicated by the black dashed line. with light modes strongly confined in the cavity. With further The inset shows the dependence on the plasma frequency ωp at the decreasing the thickness d or the tuning parameter r, the equi- insulator thickness d = 1. We set c = ~ = β = 1 (corresponding to, librium phonon frequency eventually goes down to zero from e.g., ~c/(kBT ) ' 38 µm for T = 60 K), and use ωp = 5, U = 1, 2 ∗ and g = 3.5. We set the momentum cutoff to be Λc = 10 . above, reff → 0+, at which the paraelectric-to-ferroelectric phase transition occurs (see e.g., the red solid curve in the bottom panel of Fig.9(a)). The corresponding phase diagram of the cavity setting with and should be a generic feature of various dipolar insulators varying thickness d is plotted in Fig.9(b), where the cavity- confined in the cavity. Most notably, this robust effect occurs induced softening results in favoring of the ferroelectric phase even when the underlying phase in the bulk setting is para- compared to the bulk transition point indicated by the black electric, thus indicating a possibility of inducing the structural dashed line. This favoring of the ferroelectric phase in the quantum phase transition via the cavity confinement as indi- quantum electromagnetic environment qualitatively persists cated by the vertical black downarrow in Fig.9(b). The higher almost independently of specific choices of the parameters plasma frequency leads to faster spatial decay of low-energy 16 modes inside the metallic regions and thus corresponds to pare materials at the verge of the ferroelectric phase. Strong tighter mode confinements. Hence, increasing the plasma fre- nonlinearity is also available in these materials. To be con- quency has qualitatively similar effect to decreasing the para- crete, if we consider SrTiO3, one can use the transverse op- electric slab thickness (see the inset in Fig.9(b)). tical phonon mode, for which the phonon nonlinear coupling To gain further insights into the phonon softening, in is U/M 2 ' 2.2 × 103 THz2 A˚ −2 kg−1 [142]. In the unit of Fig. 10 we plot the most-softened phonon mode (i.e., the c = ~ = β = 1 with, e.g., the temperature T = 60 K, this RTM0 mode) with varying the insulator thickness d but at the nonlinear coupling strength can be read as U ' 1.5, which is fixed bare tuning parameter r. It is this softening that desta- close to the value used in the present numerical calculations bilizes the paraelectric phase when it goes down to zero en- [143]. We remark that, in this choice, the thickness d = 1 ergy, triggering the phase transition. Thus, the unstable dipo- corresponds to an order of several tens of micrometers. lar phonon mode at the verge of the transition is precisely the transverse one and points to the in-plane direction (cf. Fig.7(a,c)). We emphasize that the results in Fig. 10 are ob- V. SUMMARY AND DISCUSSIONS tained by using the renormalized effective phonon frequency ∗ p ∗ Ωeff = reff for each thickness d, which is determined from Our work demonstrated the great potential of a heterostruc- identifying the minimum of the variational free energy includ- ture configuration as a platform towards controlling the quan- ing the nonlinear interaction in a self-consistent manner (cf. tum phase of matter by modifying quantum electrodynamic Fig.9(a)). We note that all the contributions from the ‘spec- environments in the absence of external pump. In particular, tator’ modes, i.e., the modes other than the RTM0, are also the present consideration revealed that the strong hybridiza- included in the variational analysis to determine the renormal- tion between the continuum of matter excitations and cavity ized phonon frequency. As shown in Fig. 10, to realize the ul- electromagnetic fields enriches the spectrum of elementary timate softening of the RTM0 mode, a sparser band structure excitations in a way qualitatively different from the bulk ma- as realized with a thinner layer is preferable since the energy terial. More specifically, the main results of this paper can be cost due to the nonlinear repulsive interaction between the en- summarized as follows: ergetically dense elementary modes can be mitigated. Finally, towards experimentally testing the present results, (i) Analysis of hybridized collective modes in a cavity. we suggest that it is promising to consider paraelectric ma- We analyzed the collective modes in the heterostruc- terials such as SrTiO3 and KTaO3 [122, 140], as well as ture configuration shown in Fig.2, i.e., the dielectric SnTe and Pb1−zSnzTe [141], where a tuning parameter can slab having an infrared active phonon at frequency Ω be varied by using pressure, chemical composition, and iso- surrounded by two metallic electrodes. Hybridization tope substitution. This flexibility allows one to naturally pre- of the phonon modes of the dielectric, the cavity elec- tromagnetic modes, and the plasmons of the surround- ing metallic mirrors gives rise to a rich structure of the eigenmodes; in particular, we found the emergence of a 2 number of energy modes at frequencies slightly below the bare phonon frequency Ω. The origin of these eigen- 1.5 d decrease modes can be understood from the strong frequency de- pendence of the dielectric function close to the phonon

ω 1 frequency (see AppendixC).

0.5 (ii) Cavity-enhanced ferroelectric phase transition. To go beyond the simple quadratic model of the photon- phonon-plasmon hybridization, we included phonon 0 nonlinearities in the dielectric material through the vari- 0 1 2 3 4 5 6 7 q ational method. We found that the collective modes are softened as the thickness of the dielectric slab is de- creased. The origin of this softening is the screening of FIG. 10. Softening of the phonon mode relevant to the paraelectric- the repulsive component of electric dipoles in the side- to-ferroelectric phase transition, which is labeled as the RTM0 mode in Fig.6. The solid colored curves represent the in-plane dispersions by-side configuration by metallic mirrors. The soften- of the softened phonon mode with decreasing the insulator thickness ing leads to the enhancement of the ferroelectric insta- d from top to bottom curves. Each dispersion is obtained by set- bility, which allows the cavity system to change into ting the phonon-frequency parameter Ω2 equal to the effective value ferroelectric even when the bulk material remains para- ∗ reff determined from the variational analysis for each thickness (cf. electric. We suggest that such cavity-enhanced transi- Fig.9(a)). The black dashed line represents the effective phonon tion could be experimentally observed in several quan- frequency in the bulk case. We set c = ~ = β = 1, and the param- tum paraelectric materials at the verge of the ferroelec- eters are chosen to be ωp = 5, U = 1, and g = 3.5. We set the 2 tric phase. momentum cutoff Λc = 10 . The tuning parameter is fixed to be 2 r/Λc = −45. From top to the bottom solid curves, the thickness is The mechanism proposed in this paper is not restricted to varied from d = 4 to d = 1 with step δd = 1. specific setups, but should be applicable to a wide class of sys- 17 tems supporting excitations that possess electric dipole mo- cally chiral nature of these modes may qualitatively alter the ments. For instance, our theory can be used to analyze the character of collective modes in the cavity configuration. One coupling of cavity fields to excitonic molecules [144, 145] can also consider a situation with both electric and magnetic and also to artificial quantum systems such as ultracold po- couplings being important. This can be relevant to multifer- lar molecules [146], trapped ions [147], and Rydberg atoms roic systems [175], in which ferroelectric and magnetic orders [148]. The present formulation should also be applicable to are intertwined. analyze polariton modes in the hexagonal boron nitride (hBN) Finally, it is also intriguing to consider further features and [149–151], which is a dielectric insulator material associated functionalities that can arise from fabricating additional struc- with a large optical phonon frequency. We point out again that tures at the metal-insulator interfaces. For instance, periodic the predicted changes of the elementary excitations can occur texturing or distortion of the surfaces can lead to the forma- without external driving in contrast to, e.g., the earlier studies tion of gaps in the in-plane dispersion as realized in photonic of exciton-poalritons in microcavities [152–156]. crystals. This will provide additional flexibility for manipula- As by-product of the analyses given in this paper, we en- tion of the hybridized light-matter excitations. It also merits vision several potential applications of the cavity-induced further study to consider how changes in the phonon spectrum changes of the dispersive excitation modes. For instance, one in the metal-paraelectric heterostructure can affect many-body can consider a molecular crystal sandwiched between metallic states in the metallic mirrors, such as, superconducting states. mirrors, where polaritons originate from the hybridization of This should be particularly relevant to low-density supercon- singlet and cavity electromagnetic modes. The corre- ductivity mediated by plasmons as discussed in the case of sponding triplet excitons have negligible interaction with light doped strontium titanate, where the soft transverse optical and are thus unaffected by the cavity confinement. The het- phonon plays a crucial role [176]. We hope that our work erostructure cavity configuration then allows one to change stimulates further studies in these directions. relative energies of singlet and triplet excitations, which can be applied to improve the efficiency of organic light-emitting devices and realize photon-exciton converters. We expect that the modification of the hybridized dispersion in the proposed cavity setup can also be used to control photochemical reac- ACKNOWLEDGMENTS tions in molecular systems. The similar physics may be ob- served in the presence of a single metal-dielectric interface. We are grateful to Richard Averitt, Dmitri Basov, An- Further discussions can be found in AppendixD. toine Georges, Bertrand I. Halperin, Mikhail Lukin, Dirk Several open questions remain for future studies. First, we van der Marel, Giacomo Mazza, Marios Michael, Prineha recall that the whole system considered in this paper is as- Narang, Angel Rubio, and Sho Sugiura for fruitful discus- sumed to be in thermal equilibrium, i.e., the effective temper- sions. Y.A. acknowledges support from the Japan Society for ature is identical for both of the matter and the cavity. This as- the Promotion of Science through Grant Nos. JP16J03613 and sumption could be violated if, for example, the metal mirrors JP19K23424, and Harvard University for hospitality. J.F. ac- and the insulator are further coupled to external bath modes knowledges funding from the Swiss National Science Foun- at different temperatures. Moreover, driving the cavity into dation grant 200020-192330. D.J. acknowledges support by a high-intensity regime or coherent pumping of atoms should EPSRC grant No. EP/P009565/1. D.J. and A.C. acknowledge induce the cavity-mediated interactions among the underlying funding from the European Research Council under the Eu- elementary excitations [56, 157–160]. It remains an intrigu- ropean Union’s Seventh Framework Programme (FP7/2007- ing question how our formalism can be generalized to such 2013)/ERC Grant Agreement No. 319286 (Q-MAC). A.C. ac- nonequilibrium setups. knowledges support from the Deutsche Forschungsgemein- Second, a quantitative amount of the predicted modification schaft (DFG) via the Cluster of Excellence ‘The Ham- of the elementary excitations and also that of the transition- burg Centre for Ultrafast Imaging’ (EXC 1074–Project ID point shift can in practice depend on several specifics such 194651731) and from the priority program SFB925. E.D. as the quality factor of a cavity, the phonon-scattering rate acknowledges support from Harvard-MIT CUA, AFOSR- in matter, and the plasma frequency and the conductivity in MURI Photonic Quantum Matter (award FA95501610323), metals. It merits further study to explore the impact of those and DARPA DRINQS program (award D18AC00014). experimental complexities on the present results. Meanwhile, on the border of the phase transition, there are also interesting open questions on how quantum critical behavior is modified due to cavity confinement and how it can be affected by dis- Appendix A: Macroscopic Maxwell equations sipative nature of a cavity [122, 161–167]. It would also be interesting to explore combining cavity-enhanced phase tran- sitions with topological aspects of open systems [168–173]. In the linear regime, elementary excitations of the electro- Third, it will be intriguing to consider the situation in which magnetic fields in nonabsorbing matter can in general be ob- the primary coupling between matter and light is mediated tained by solving the macroscopic Maxwell equations with by the magnetic field. A concrete example can be Damon- using boundary conditions appropriate for a physical setting Eschbach modes in ferromagnetic films [174]. The intrinsi- of interest. Specifically, the dynamics of the electromagnetic 18

fields is governed by the transverse electric (TE) polarization defined by the condi- tions, ∂B ∇ × E + = 0, (A1) h ⊥ i ∂t ∇ × U η (r) = 0 (TM modes), (B2) 1 ∂D z ∇ × B = , (A2) h ⊥ i U η (r) = 0 (TE modes), (B3) µ0 ∂t z ∇ · D = 0, (A3) and P = ± specifying the parity symmetry of the eigenmode ∇ · B = 0, (A4) function under the reflection, ⊥ ⊥ where E is the electric field, B is the magnetic field, µ0 is the [U η (x, y, −z)]x,y = P [U η (x, y, z)]x,y, (B4) vacuum permeability, and D is the macroscopic electric field ⊥ ⊥ [U (x, y, −z)]z = −P [U (x, y, z)]z. (B5) in matter. Transforming to the frequency basis, the macro- η η scopic electric field D is related to E via Because of the spatial invariance in the x and y directions, we can decompose the eigenfunction as D (r) = 0(ω, r)E (r) (A5) ⊥ iq·ρ ⊥ U η (r) = e uη (z), (B6) with 0 being the vacuum permittivity and (ω, r) being the where ρ = (x, y)T. The eigenvalue problem (A7) in the frequency- and position-dependent relative permittivity re- present case can then be given by flecting material properties. We choose the Coulomb gauge 2 −ω  (ωη, z) η u⊥ (z)+e−iq·ρ∇ × ∇ × eiq·ρu⊥ (z) = 0 ∇ · A (r) = 0. (A6) c2 η η (B7) In terms of the vector potential, the equation of motion can be simplified as with the relative permittivity being defined by

2 X −ω  (ω, r) (ω, z) = p(ω) ϑ(sz − d/2) A (r) + ∇ × [∇ × A (r)] = 0, (A7) c2 s=± ! √ X where c = 1/ 0µ0 is the vacuum light speed. In the bulk +ξ(ω) ϑ(d/2 − sz) − 1 . (B8) case, one can simply solve this eigenvalue problem by em- s=± ploying the spatial invariance and expressing the eigensolu- Here, we use the Heaviside step function tions in terms of the plane waves (cf. Eq. (4)). ( 1 (x > 0) ϑ(x) = , (B9) 0 (x < 0) Appendix B: Eigenmodes in the cavity configuration and assume the relative permittivity functions in the metals From now on, we focus on the heterostructure configuration (|z| > d/2) and the insulator (|z| < d/2) as as illustrated in Fig.5, which is symmetric under the spatial 2 ωp parity transformation along the direction perpendicular to the p(ω) = 1 − , (B10) ω2 plane, i.e., under the reflection (x, y, z) → (x, y, −z). g2 ξ(ω) = 1 + . (B11) Ω2 − ω2

1. Transverse modes We recall that ωp is the plasma frequency in the metal mir- rors, Ω is the frequency of the nondispersive matter excitation such as an optical phonon mode in ionic crystals, and g char- We first consider the transverse modes. In this case, the acterizes the strength of the light-matter coupling between the vector field can be quantized and expanded as matter dipolar mode and cavity electromagnetic fields. It is s also useful to express the electric and magnetic fields in terms X   Aˆ(r) = ~ aˆ U ⊥ (r) + H.c. , (B1) of the eigenmodes as follows: 2 Aω η η η 0 η r X ωη Eˆ(r) = ~ (ˆa E (r) + H.c.) , (B12) 2 A η η where A is the area of the system, η ∈ {q, n, λ} labels each η 0 eigenmode in the cavity setting, aˆη is its annihilation oper- ⊥ Eη (r) = iU η (r) , (B13) ator, ωη is an eigenfrequency, and U η is the corresponding ⊥ s transverse eigenmode function satisfying ∇ · U η = 0. Here, ˆ X ~ q n ∈ B(r) = (ˆaηBη (r) + H.c.) , (B14) is the two-dimensional in-plane wavevector, N la- 20Aωη bels a discrete eigenmode in each sector λ = (Λ,P ) with η ⊥ Λ ∈ {TM, TE} specifying the transverse magnetic (TM) or Bη (r) = ∇ × U η (r) . (B15) 19

TABLE II. Eigenmodes in the cavity configuration. The labels of the bands are consistent with those shown in Fig.6. The polarization of each mode is specified by Λ ∈ {TM, TE, L}, where TM and TE indicate the transverse magnetic and electric polarizations, respectively, and L ⊥ k denotes the longitudinal polarization. The parity of the mode functions uη and fη are specified by P = ± (see also the main text). We also indicate whether or not the divergence of the electric field ∇ · Eη vanishes. The last column shows the explicit functional form of the mode function for each eigenmode, where ei indicates the unit vector in the direction i = x, y, z.

Bands Λ P ∇ · Eη Mode function ⊥ I  q  u (z) = u cos(κηz)ex − i sin(κηz)ez [ϑ(d/2 − z) + ϑ(d/2 + z) − 1] η η κη TM2, RTMn=0,2,4..., SS (=TM0) TM + zero   M P −sνη z q +u e ex + is ez ϑ(sz − d/2) η s=±1 νη ⊥ I  q  u (z) = u sin(κηz)ex + i cos(κηz)ez [ϑ(d/2 − z) + ϑ(d/2 + z) − 1] η η κη TM1, RTMn=1,3,5..., AS1,2 TM − zero   M P −sνnz q +u e sex + i ez ϑ(sz − d/2) η s=±1 νn ⊥ I uη (z) = uη cos(κηz)ey [ϑ(d/2 − z) + ϑ(d/2 + z) − 1] TE1, RTEn=0,2,4... TE + zero M P −sνη z +uη s=±1 e eyϑ(sz − d/2) ⊥ I uη (z) = uη sin(κηz)ey [ϑ(d/2 − z) + ϑ(d/2 + z) − 1] TE0, RTEn=1,3,5... TE − zero M P −sνnz +uη s=±1 se eyϑ(sz − d/2)   k I q iκη z −iκη z κη iκη z −iκη z fη (z) = f √ (e ± e )ex + √ (e ∓ e )ez η κ2 +q2 κ2 +q2 L± L ± nonzero η η × [ϑ(d/2 − z) + ϑ(d/2 + z) − 1]

⊥ The boundary conditions require the continuity of n×Eη and we obtain the explicit functional form of uη for this class of n · [0Eη] across the interfaces with n being the unit vector eigenmodes as shown in the first line of TableII. This set in- perpendicular to the surfaces. The latter condition is equiva- cludes the bands labeled as TM2, RTMn=0,2,4..., SS (=TM0) lent to the continuity of the magnetic field orthogonal to the that are shown in Fig.6(a,d). In these modes, the longitudinal surface normal n × Bη. wavenumber κη in the insulator region is defined by 2 2 ωη 2 κη = 2 ξ (ωη) − q . (B24) a. TM modes c We emphasize that κη takes either a real value or a purely Without loss of generality, we here consider the TM- imaginary value. In the latter case, the eigenmodes are lo- polarized eigenmodes satisfying calized at the interfaces and represent the surface modes. We define the skin wavenumber νη, which is real and positive, as [Bη]y 6= 0, [Bη]x,z = 0, (B16) follows: r [E ] = 0, [E ] 6= 0. (B17) ω2 η y η x,z ν = q2 − η  (ω ) > 0. (B25) η c2 p η The Maxwell equation ∇ · Bˆ = 0 then leads to qy = 0, i.e., M I The coefficients uη and uη in the metals (M) and in the insu- the eigenmodes propagate along the x axis and thus we denote lator (I), respectively, satisfy the following relations imposed q = qx. The parity conditions (B4) and (B5) in the present by the boundary conditions: case can be read as M I νη d/2 uη /uη = e cos (κηd/2) [B (x, y, −z)] = −P [B (x, y, z)] , (B18) η y η y ξ (ω ) ν η η νη d/2 [E (x, y, −z)] = P [E (x, y, z)] , (B19) = − e sin (κηd/2) . (B26) η x η x p (ωη) κη [Eη(x, y, −z)]z = −P [Eη(x, y, z)]z . (B20) Aside an irrelevant phase factor, these coefficients can In terms of the mode function u⊥, the TM modes considered be fixed by further imposing the normalization condition η R ⊥ 2 here satisfy dz uη (z) = 1. In the same way as done above, we can also obtain the cor- u⊥(z) = 0, (B21) η y responding eigenmodes in the case of the odd parity P = −. Their functional forms are given in the second line in TableII. u⊥(−z) = P u⊥(z) , (B22) η x η x The corresponding coefficients satisfy the boundary condi- u⊥(−z) = −P u⊥(z) . (B23) η z η z tions M I νη d/2 We first consider the case of the even parity P = +. Solv- uη /uη = e sin (κηd/2) ing the eigenvalue equation (B7) for the mode function u⊥ ξ (ω ) ν η η η νη d/2 iq·ρ ⊥ = e cos (κηd/2) , (B27) with the transversality condition, i.e., ∇ · (e uη ) = 0, p (ωη) κη 20

FIG. 11. Spatial profiles of the TM modes in the cavity configuration at low and high in-plane momenta q. The black dashed vertical lines ⊥ indicate the metal-insulator interfaces. The blue solid curve shows the spatial profile of the out-of-plane component [iuη (z)]z of the mode function (corresponding to the z component of the electric field). The red solid curve shows the spatial profile of the in-plane component ⊥ [uη (z)]x of the mode function (corresponding to the x component of the electric field). The parameters are the same as in Fig.6 while the in-plane momentum q for low- and high-momentum results is chosen to be q = 0.1 Ω/c and q = 7 Ω/c, respectively.

R ⊥ 2 and also the normalization condition dz uη (z) = 1. The eigenvalue equation then leads to the functional form in These modes are labeled as TM1, RTMn=1,3,5..., AS1,2 that the third (fourth) line in TableII in the case of the even par- are shown in Fig.6(b,e). We show typical spatial profiles of ity P = + (the odd parity P = −). The boundary conditions ⊥ the electric fields at low and high in-plane momenta q for each require that the value of the mode function uη and its first spa- of these TM modes in Fig. 11. tial derivative must be continuous across the interfaces. More specifically, in the case of the even parity they lead to the con- M,I ditions on the coefficients uη , b. TE modes M I νη d/2 uη /uη = e cos (κηd/2) ν In the case of the TE polarization, without loss of generality η νη d/2 = − e sin (κηd/2) , (B36) we can assume the following conditions: κη while in the odd-parity case we have to impose the conditions [Eη]y 6= 0, [Eη]x,z = 0, (B28)

[Bη] = 0, [Bη] 6= 0. (B29) M I νη d/2 y x,z uη /uη = e sin (κηd/2) ν η νη d/2 From the transversality condition, = e cos (κηd/2) . (B37) κη iq·ρ ⊥ ∇ · (e uη ) = 0, (B30) R ⊥ 2 The normalization condition dz uη (z) = 1 can be also and the Maxwell equation ∇ · Dˆ = 0, we obtain the condition imposed on the coefficients. In Fig.6(c,f), the TE modes with qy = 0, i.e., the TE modes considered here also propagate the even parity are labeled as TE1, RTEn=0,2,4... while the along the x axis. The electromagnetic fields of the TE modes ones with the odd parity are labeled as TE0, RTEn=1,3,5.... transform under the parity transformation as In contrast to some of the TM modes (such as the AS1,2 and the RTM0 modes), the spatial profiles of the electric fields are [Eη(x, y, −z)]y = P [Eη(x, y, z)]y , (B31) qualitatively insensitive to the in-plane momenta for all of the TE modes; their typical results are shown in Fig. 12. [Bη(x, y, −z)]x = −P [Bη(x, y, z)]x , (B32)

[Bη(x, y, −z)]z = P [Bη(x, y, z)]z . (B33) These conditions can be read in terms of the mode function 2. Longitudinal mode ⊥ uη as The other type of solutions is the longitudinal mode that is u⊥(z) = 0, (B34) η x,z characterized by the longitudinal electric field u⊥(−z) = P u⊥(z) . (B35) η y η y ∇ × Eη = 0. (B38) 21

k Because of our choice of the Coulomb gauge Aˆ = 0, the lon- in-plane in-plane RTE3

gitudinal field is directly linked with the longitudinal mode of [a.u.] [a.u.] the matter field φˆk as discussed in the main text (cf. Eq. (37)). TE 1 ˆ From the Maxwell equation ∇ · D = 0, we can see that TE 0 RTE2 this mode oscillates in time at a constant frequency Ωk cor- responding to the vanishing permittivity in the insulator

field RTE field   5 RTE1 ξ Ωk = 0, (B39) p Ωk = Ω2 + g2. (B40) RTE4 RTE0 We can expand the longitudinal matter field in terms of the mode function as z [d] z [d] r ˆk ~ X  ˆk iq·ρ k  φ (r) = φqnλe fη (z) + H.c. , FIG. 12. Spatial profiles of the TE modes in the cavity configu- 2MNdΩ η ration. The black dashed vertical lines indicate the metal-insulator k (B41) interfaces. The mode functions fη of all the TE modes have only the in-plane components, whose spatial profiles are shown by the red where the longitudinal mode function satisfies solid curve (corresponding to the y component of the electric field). The parameters are the same as in Fig.6. We use q = 0.1 Ω/c for iq·ρ k ∇ × (e fη (z)) = 0. (B42) the sake of concreteness while the spatial profiles of the TE modes are qualitatively insensitive to a specific choice of the in-plane mo- ˆ ∂Bˆ mentum q. Meanwhile, the Maxwell equation ∇×E = − ∂t ensures the vanishing magnetic field Bˆ = 0. Thus, it suffices to only con- sider the longitudinal electric field and, without loss of gener- ality, we here consider the modes satisfying while similar arguments can be given for the corresponding RTE modes as well.

[Eη]y = 0, [Eη]x,z 6= 0, (B43) Inside the insulator region, the RTM modes with q = 0 B = 0. (B44) have the spatial profiles of the magnetic field either sin(κz) or η cos(κz), which we denote by the labels P = + or P = −, The corresponding in-plane momenta point to the x direction, respectively. We here recall that the parity of the eigenmodes i.e., qy = 0 and thus q = qx. The boundary conditions at the is defined by that of the transverse components of the electric interfaces lead to the vanishing electromagnetic fields in the field. In these cases, the longitudinal wavevector κ should metal mirrors. The resulting explicit functional form of the satisfy the following equations: mode function is given in the last line in TableII. Depending s ( κd  on the parity P = ± defined via the relations ξ(ω) cot 2 (P = +), = κd  (C1) |p(ω)| − tan (P = −). h k i 2 fη (z) = 0, (B45) y The Maxwell equations also lead to the additional constraint h k i h k i fη (−z) = P fη (z) , (B46) p ω x x κ = ξ(ω) . (C2) h k i h k i c fη (−z) = −P fη (z) , (B47) z z In the case of the frequency regime slightly below the bare there are two classes of solutions, which are labeled as the L± phonon frequency Ω of paraelectric materials, the dielectric modes in Fig.6(a). The corresponding longitudinal wavenum- function ξ(ω) takes a singularly large value compared with ber κη can be simply obtained by the boundary conditions |p(ω)|. Thus, the conditions (C1) and (C2) lead to the ap- h k i proximate eigenmode equation f η(±d/2) = 0. x p πc ξ(ω)ω ' n (n = 1, 2,...). (C3) d Appendix C: Origin of the resonant soft-phonon modes Owing to the singular nature of ξ(ω) in the limit of ω → Ω−0, this condition can in principle be attained for an arbitrary in- We here provide simple explanations to elucidate the origin teger n. It is this resonant structure that allows for generating of the resonant softened phonon modes that are denoted by the a large number of soft-phonon modes in the heterostructure labels RTM and RTE in Fig.6. To this end, we consider a fre- configuration. In practice, when the damping of the phonon quency regime in which the metallic permittivity is negative mode is included, there should exist an effective cutoff on pos- p(ω) < 0 while the insulator dielectric function is positive sible values of n, but still many eigensolutions should be al- ξ(ω) > 0 (cf. Eqs. (B10) and (B11)). To be specific, we fo- lowed as long as the damping rate is small enough in compar- cus on the RTM modes at the zero in-plane momentum q = 0 ison with the phonon frequency. 22

(a) (b) very different frequencies (i.e., optical vs mid-infrared), the general formalism presented in Sec. III should be applicable S’ to both cases. Either one of these modes can be introduced as Metal a nondispersive mode having the frequency Ω in Eq. (16). To ωS’ =2ω T be concrete, we focus on the case of electronic excitations, and ω argue that the proposed cavity configuration can be applied to Molecular solid improve efficiencies in molecular LEDs. S A major challenge in organic LEDs is that the lowest exci- T tonic state is a spin triplet, which is optically dark (i.e., it re- Metal S’’ combines through nonradiative channels). The reason for this is that Hund’s rule for electrons in a molecule favors the spin q alignment and thus the spin triplet excitonic states are lower in energy than the singlet ones. Having a practical method of FIG. 13. Possible applications of the cavity setting to molecular reversing the order of excitonic states by making the singlet and excitonic systems. (a) Schematic figure illustrating the proposed state of an exciton lower in energy will have important impli- setup, in which molecular solids or assembled molecules interact with light modes confined between metal mirrors. (b) Schematic fig- cations for improving the efficiency of organic LEDs [114]. In ure illustrating applications of the elementary excitations in the cav- the cavity configuration, since triplet excitonic states are dark, ity setting to an organic light-emitting device and a photon-exciton they do not couple to light and are not strongly modified. In conversion. The black horizontal solid (dotted) line denoted as T (S) contrast, the singlet excitonic state hybridizes with confined indicates a bare nondispersive triplet (singlet) energy of a molecular light modes and forms the softened collective modes such as state or an electron-hole pair. Only the singlet state S couples to con- the RTM0 mode shown in Fig.6(a). This should make it pos- fined light modes, resulting in the split into the multiple branches of sible to have the lowest excitonic state as a spin singlet, thus the in-plane dispersions. A lower dispersion (red solid curve denoted 00 realizing efficient organic LEDs (see the red solid curve in as S ) can potentially be applied to enhance the efficiency of organic Fig. 13(b)). While this type of hybridizations have been dis- light-emitting devices. Tuning the in-plane momentum q or the thick- cussed in the literature before [114], the novelty of the present ness of the layer, a higher dispersion (blue solid curve denoted as S0) consideration is the inclusion of the continuum of both col- can be used to realize the resonance condition ωS0 = 2ωT, leading to a conversion of a photon into a pair of triplet excitons as denoted lective matter excitations and cavity light modes in a slab of by two dashed black arrows. a finite thickness as well as taking into account the dispersive nature of the hybridized excitations there. In general, the dis- persive nature could be utilized to further tune the selective Appendix D: Applications of cavity-based control of polariton control of chemical reactivity landscapes in order to enhance dispersions or suppress certain types of photochemical reactions. We next discuss a complementary idea of using a cavity to Substantial modifications of the polariton dispersions in the facilitate the conversion of photons into pairs of triplet exci- cavity geometry (see e.g., Fig.6 in Sec. III) are generic fea- tons. This process could be useful for generating entangled tures for collective dipolar modes coupled to confined light pairs of excitations. At ambient conditions, such conversion modes. In this respect, the cavity-based control of energy- process is challenging to realize because the resonant absorp- level structures can potentially be applied to a wide range tion of light occurs at the energy of singlet excitons and this of systems such as ionic crystals, molecular solids, assem- energy is only slightly above the energy of the triplet excitons; bled molecules, and excitonic systems. Organic molecules are typical energy spacing is characterized by the Hund’s splitting particularly promising for such applications since their exci- and is an order of a few meV. tations have large dipole moments and crystals can achieve This limitation can be overcome by utilizing the proposed high-molecular densities [177]. Recent experiments have cavity setup. As discussed before, when a molecular solid demonstrated the strong coupling between a single molecule is confined in the cavity, the singlet exciton hybridizes with and a cavity mode even for the modest cavity quality factors light modes while the triplet one does not. This can lead to [118, 178]. This means that the strength of the coupling be- the formation of a hybridized singlet mode above the origi- tween the molecular optical excitation and the cavity mode ex- nal exciton energy (see e.g., the blue solid curve S’ above the ceeds the decoherence rate of the excitations [61, 82, 98, 179]. black dashed horizontal line S in Fig. 13(b)). One may, for Besides a case study of the cavity-enhanced ferroelectricity in example, use the SS(=TM0) and AS2 modes for this purpose. Sec.IV, we discuss below that the cavity-induced changes in Owing to the nontrivial dispersion of such a hybridized mode, the hybridized dispersion can be utilized to improve the effi- one can tune the in-plane momentum q (and the thickness d or ciency of organic light-emitting devices (LEDs) and to realize the plasma frequency ωp if necessary) such that a hybridized exciton-photon converters. singlet energy ωS0 becomes equal to twice of the triplet energy We consider a crystal of polar molecules surrounded by 2ωT and thus the conversion process of a single photon into metallic electrodes as shown in Fig. 13. The relevant infrared two triplet excitons becomes resonant (cf. the dashed black ar- active mode can be either an electronic excitation, such as rows in Fig. 13(b)). One can create singlet exciton-polaritons the E1 electronic transition, or a vibrational excitation of the at finite in-plane momenta by, e.g., having incident light at molecular crystal. While these two types of excitations have nonzero angle or constructing a grating on the surface of the 23

metals.

[email protected] D. Jaksch, and A. Cavalleri, Possible light-induced super- [1] C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg, Pho- conductivity in K3C60 at high temperature, Nature 530, 461 tons and Atoms (Wiley, New York, 1989). (2016). [2] T. Kampfrath, K. Tanaka, and K. A. Nelson, Resonant and [19] E. Pomarico, M. Mitrano, H. Bromberger, M. A. Sentef, A. Al- nonresonant control over matter and light by intense terahertz Temimy, C. Coletti, A. Stohr,¨ S. Link, U. Starke, C. Cacho, transients, Nat. Photon. 7, 680 (2013). R. Chapman, E. Springate, A. Cavalleri, and I. Gierz, En- [3] P. W. Anderson and A. H. Dayem, Radio-frequency effects in hanced electron-phonon coupling in graphene with periodi- superconducting thin film bridges, Phys. Rev. Lett. 13, 195 cally distorted lattice, Phys. Rev. B 95, 024304 (2017). (1964). [20] A. A. Patel and A. Eberlein, Light-induced enhancement [4] A. F. G. Wyatt, V. M. Dmitriev, W. S. Moore, and F. W. of superconductivity via melting of competing bond-density Sheard, Microwave-enhanced critical supercurrents in con- wave order in underdoped cuprates, Phys. Rev. B 93, 195139 stricted tin films, Phys. Rev. Lett. 16, 1166 (1966). (2016). [5] G. M. Eliashberg, Film superconductivity stimulated by a [21] M. Knap, M. Babadi, G. Refael, I. Martin, and E. Demler, Dy- high- frequency field, JETP LETT. 11, 114 (1970). namical Cooper pairing in nonequilibrium electron-phonon [6] B. I. Ivlev, S. G. Lisitsyn, and G. M. Eliashberg, Nonequilib- systems, Phys. Rev. B 94, 214504 (2016). rium excitations in superconductors in high-frequency fields, [22] A. Komnik and M. Thorwart, BCS theory of driven supercon- J. Low Temp. Phys. 10, 449 (1973). ductivity, Eur. Phys. J. B 89, 244 (2016). [7] A. Schmid, Stability of radiation-stimulated superconductiv- [23] M. A. Sentef, A. Tokuno, A. Georges, and C. Kollath, The- ity, Phys. Rev. Lett. 38, 922 (1977). ory of laser-controlled competing superconducting and charge [8] T. Oka and H. Aoki, Photovoltaic hall effect in graphene, orders, Phys. Rev. Lett. 118, 087002 (2017). Phys. Rev. B 79, 081406 (2009). [24] M. A. Sentef, Light-enhanced electron-phonon coupling from [9] T. Kitagawa, T. Oka, A. Brataas, L. Fu, and E. Demler, Trans- nonlinear electron-phonon coupling, Phys. Rev. B 95, 205111 port properties of nonequilibrium systems under the applica- (2017). tion of light: Photoinduced quantum hall insulators without [25] D. M. Kennes, E. Y. Wilner, D. R. Reichman, and A. J. Mil- landau levels, Phys. Rev. B 84, 235108 (2011). lis, Transient superconductivity from electronic squeezing of [10] N. H. Lindner, G. Refael, and V. Galitski, Floquet topological optically pumped phonons, Nat. Phys. 13, 479 (2017). insulator in semiconductor quantum wells, Nat. Phys. 7, 490 [26] N. Tancogne-Dejean, M. A. Sentef, and A. Rubio, Ultrafast (2011). modification of Hubbard U in a strongly correlated material: [11] Y. H. Wang, H. Steinberg, P. Jarillo-Herrero, and N. Gedik, Ab initio high-harmonic generation in NiO, Phys. Rev. Lett. Observation of floquet-bloch states on the surface of a topo- 121, 097402 (2018). logical insulator, Science 342, 453 (2013). [27] J. Zhang and R. Averitt, Dynamics and control in complex [12] J. W. McIver, B. Schulte, F.-U. Stein, T. Matsuyama, G. Jotzu, transition metal oxides, Annu. Rev. Mat. Res. 44, 19 (2014). G. Meier, and A. Cavalleri, Light-induced anomalous hall ef- [28] A. Cavalleri, Photo-induced superconductivity, Contemp. fect in graphene, Nat. Phys. 16, 38 (2020). Phys. 59, 31 (2018). [13] M. Rini, N. Dean, J. Itatani, Y. Tomioka, Y. Tokura, R. W. [29] H. Patel, L. Huang, C.-J. Kim, J. Park, and M. W. Graham, Schoenlein, and A. Cavalleri, Control of the electronic phase Stacking angle-tunable photoluminescence from interlayer ex- of a manganite by mode-selective vibrational excitation, Na- citon states in twisted bilayer graphene, Nat. Commun. 10, 1 ture 449, 72 (2007). (2019). [14]M.F orst,¨ C. Manzoni, S. Kaiser, Y. Tomioka, Y. Tokura, [30] A. Cartella, T. F. Nova, M. Fechner, R. Merlin, and A. Caval- R. Merlin, and A. Cavalleri, Nonlinear phononics as an ul- leri, Parametric amplification of optical phonons, Proc. Natl. trafast route to lattice control, Nat. Phys. 7, 854 (2011). Acad. Sci. U.S.A. 115, 12148 (2018). [15] R. Mankowsky, A. Subedi, M. Forst,¨ S. O. Mariager, M. Chol- [31] Y. K. Srivastava, M. Manjappa, H. N. S. Krishnamoorthy, and let, H. Lemke, J. S. Robinson, J. M. Glownia, M. P. Minitti, R. Singh, Accessing the high-Q dark plasmonic Fano reso- A. Frano, M. Fechner, N. A. Spaldin, T. Loew, B. Keimer, nances in superconductor metasurfaces, Adv. Opt. Mater. 4, A. Georges, and A. Cavalleri, Nonlinear lattice dynamics as 1875 (2016). a basis for enhanced superconductivity in YBa2Cu3O6.5, Na- [32] M. Fleischhauer, A. Imamoglu, and J. P. Marangos, Electro- ture 516, 71 (2014). magnetically induced transparency: Optics in coherent media, [16] R. Matsunaga, N. Tsuji, H. Fujita, A. Sugioka, K. Makise, Rev. Mod. Phys. 77, 633 (2005). Y. Uzawa, H. Terai, Z. Wang, H. Aoki, and R. Shimano, [33] B. Y. Zeldovich, N. F. Pilipetsky, and V. V. Shkunov, Princi- Light-induced collective pseudospin precession resonating ples of Phase Conjugation (Springer, Berlin, 1985). with Higgs mode in a superconductor, Science 345, 1145 [34] S. Sugiura, E. A. Demler, M. Lukin, and D. Podolsky, Reso- (2014). nantly enhanced polariton wave mixing and Floquet paramet- [17] A. F. Kemper, M. A. Sentef, B. Moritz, J. K. Freericks, and ric instability, arXiv:1910.03582 (2019). T. P. Devereaux, Direct observation of Higgs mode oscillations [35] R. H. Dicke, Coherence in spontaneous radiation processes, in the pump-probe photoemission spectra of electron-phonon Phys. Rev. 93, 99 (1954). mediated superconductors, Phys. Rev. B 92, 224517 (2015). [36] M. Tavis and F. W. Cummings, Exact solution for an N- [18] M. Mitrano, A. Cantaluppi, D. Nicoletti, S. Kaiser, A. Peruc- molecule—radiation-field hamiltonian, Phys. Rev. 170, 379 chi, S. Lupi, P. Di Pietro, D. Pontiroli, M. Ricco,` S. R. Clark, (1968). 24

[37] K. Hepp and E. H. Lieb, On the superradiant phase transition 075101 (2009). for molecules in a quantized radiation field: the Dicke maser [59] E. M. Purcell, H. C. Torrey, and R. V. Pound, Resonance ab- model, Ann. Phys. 76, 360 (1973). sorption by nuclear magnetic moments in a solid, Phys. Rev. [38] F. T. Hioe, Phase transitions in some generalized Dicke mod- 69, 37 (1946). els of superradiance, Phys. Rev. A 8, 1440 (1973). [60] E. T. Jaynes and F. W. Cummings, Comparison of quantum [39] C. Emary and T. Brandes, Quantum chaos triggered by precur- and semiclassical radiation theories with application to the sors of a quantum phase transition: The Dicke model, Phys. beam maser, Proc. IEEE 51, 89 (1963). Rev. Lett. 90, 044101 (2003). [61] H. J. Kimble, Strong interactions of single atoms and photons [40] C. Emary and T. Brandes, Chaos and the quantum phase tran- in cavity QED, Phys. Scr. 76, 127 (1998). sition in the Dicke model, Phys. Rev. E 67, 066203 (2003). [62] J. M. Raimond, M. Brune, and S. Haroche, Manipulating [41] M. Gaudin, Diagonalisation d’une classe d’hamiltoniens de quantum entanglement with atoms and photons in a cavity, spin, J. Phys. (France) 37, 1087 (1976). Rev. Mod. Phys. 73, 565 (2001). [42] R. Richardson, A restricted class of exact eigenstates of the [63] R. Miller, T. E. Northup, K. M. Birnbaum, A. Boca, A. D. pairing-force Hamiltonian, Phys. Lett. 3, 277 (1963). Boozer, and H. J. Kimble, Trapped atoms in cavity QED: cou- [43] O. Tsyplyatyev, J. von Delft, and D. Loss, Simplified deriva- pling quantized light and matter, J. Phys. B 38, S551 (2005). tion of the Bethe-ansatz equations for the Dicke model, Phys. [64] H. Walther, B. T. H. Varcoe, B.-G. Englert, and T. Becker, Rev. B 82, 092203 (2010). Cavity quantum electrodynamics, Rep. Prog. Phys. 69, 1325 [44] K. Rzazewski,˙ K. Wodkiewicz,´ and W. Zakowicz,˙ Phase tran- (2006). sitions, two-level atoms, and the A2 term, Phys. Rev. Lett. 35, [65] A. Boca, R. Miller, K. M. Birnbaum, A. D. Boozer, J. Mc- 432 (1975). Keever, and H. J. Kimble, Observation of the vacuum Rabi [45] P. Nataf and C. Ciuti, No-go theorem for superradiant quan- spectrum for one trapped atom, Phys. Rev. Lett. 93, 233603 tum phase transitions in cavity QED and counter-example in (2004). circuit QED, Nat. Commun. 1, 72 (2010). [66] K. M. Birnbaum, A. Boca, R. Miller, A. D. Boozer, T. E. [46] G. Chen, Z. Chen, and J. Liang, Simulation of the superra- Northup, and H. J. Kimble, Photon blockade in an optical diant quantum phase transition in the superconducting charge cavity with one trapped atom, Nature 436, 87 (2005). qubits inside a cavity, Phys. Rev. A 76, 055803 (2007). [67] C. Hamsen, K. N. Tolazzi, T. Wilk, and G. Rempe, Two- [47] J. Koch and K. Le Hur, Superfluid–Mott-insulator transition of photon blockade in an atom-driven cavity QED system, Phys. light in the Jaynes-Cummings lattice, Phys. Rev. A 80, 023811 Rev. Lett. 118, 133604 (2017). (2009). [68] G. Khitrova, H. Gibbs, M. Kira, S. W. Koch, and A. Scherer, [48] T. Grießer, A. Vukics, and P. Domokos, Depolarization shift Vacuum Rabi splitting in semiconductors, Nat. Phys. 2, 81 of the superradiant phase transition, Phys. Rev. A 94, 033815 (2006). (2016). [69] J. P. Reithmaier, G. Sek, A. Loffler,¨ C. Hofmann, S. Kuhn, [49] G. Mazza and A. Georges, Superradiant quantum materials, S. Reitzenstein, L. Keldysh, V. Kulakovskii, T. Reinecke, Phys. Rev. Lett. 122, 017401 (2019). and A. Forchel, Strong coupling in a single quantum dot– [50] L. Chirolli, M. Polini, V. Giovannetti, and A. H. MacDonald, semiconductor microcavity system, Nature 432, 197 (2004). Drude weight, cyclotron resonance, and the Dicke model of [70] T. Yoshie, A. Scherer, J. Hendrickson, G. Khitrova, H. Gibbs, graphene cavity QED, Phys. Rev. Lett. 109, 267404 (2012). G. Rupper, C. Ell, O. Shchekin, and D. Deppe, Vacuum [51] F. Dimer, B. Estienne, A. S. Parkins, and H. J. Carmichael, Rabi splitting with a single quantum dot in a photonic crys- Proposed realization of the Dicke-model quantum phase tran- tal nanocavity, Nature 432, 200 (2004). sition in an optical cavity QED system, Phys. Rev. A 75, [71] K. Hennessy, A. Badolato, M. Winger, D. Gerace, M. Atature,¨ 013804 (2007). S. Gulde, S. Falt,¨ E. L. Hu, and A. Imamoglu,˘ Quantum na- [52] E. G. D. Torre, S. Diehl, M. D. Lukin, S. Sachdev, and ture of a strongly coupled single quantum dot–cavity system, P. Strack, Keldysh approach for nonequilibrium phase tran- Nature 445, 896 (2007). sitions in quantum optics: Beyond the Dicke model in optical [72] M. T. Rakher, N. G. Stoltz, L. A. Coldren, P. M. Petroff, cavities, Phys. Rev. A 87, 023831 (2013). and D. Bouwmeester, Externally mode-matched cavity quan- [53] Z. Zhiqiang, C. H. Lee, R. Kumar, K. J. Arnold, S. J. Mas- tum electrodynamics with charge-tunable quantum dots, Phys. son, A. S. Parkins, and M. D. Barrett, Nonequilibrium phase Rev. Lett. 102, 097403 (2009). transition in a spin-1 Dicke model, Optica 4, 424 (2017). [73] L. Greuter, S. Starosielec, A. V. Kuhlmann, and R. J. Warbur- [54] Z. Zhang, C. H. Lee, R. Kumar, K. J. Arnold, S. J. Masson, ton, Towards high-cooperativity strong coupling of a quantum A. L. Grimsmo, A. S. Parkins, and M. D. Barrett, Dicke-model dot in a tunable microcavity, Phys. Rev. B 92, 045302 (2015). simulation via cavity-assisted Raman transitions, Phys. Rev. A [74] C.-H. Su, A. D. Greentree, and L. C. L. Hollenberg, Towards 97, 043858 (2018). a picosecond transform-limited nitrogen-vacancy based single [55] C. Maschler, I. B. Mekhov, and H. Ritsch, Ultracold atoms in photon source, Opt. Express 16, 6240 (2008). optical lattices generated by quantized light fields, Eur. Phys. [75] A. Faraon, C. Santori, Z. Huang, V. M. Acosta, and R. G. J. D 46, 545 (2008). Beausoleil, Coupling of nitrogen-vacancy centers to photonic [56] K. Baumann, C. Guerlin, F. Brennecke, and T. Esslinger, crystal cavities in monocrystalline diamond, Phys. Rev. Lett. Dicke quantum phase transition with a superfluid gas in an 109, 033604 (2012). optical cavity, Nature 464, 1301 (2010). [76] R. Albrecht, A. Bommer, C. Deutsch, J. Reichel, and [57] C. Hamner, C. Qu, Y. Zhang, J. Chang, M. Gong, C. Zhang, C. Becher, Coupling of a single nitrogen-vacancy center in and P. Engels, Dicke-type phase transition in a spin-orbit- diamond to a fiber-based microcavity, Phys. Rev. Lett. 110, coupled Bose–Einstein condensate, Nat. Commun. 5, 1 243602 (2013). (2014). [77] L. Li, T. Schroder,¨ E. H. Chen, M. Walsh, I. Bayn, J. Gold- [58]L.P alov´ a,´ P. Chandra, and P. Coleman, Quantum critical stein, O. Gaathon, M. E. Trusheim, M. Lu, J. Mower, M. Cot- paraelectrics and the casimir effect in time, Phys. Rev. B 79, let, M. L. Markham, D. J. Twitchen, and D. Englund, Coher- 25

ent spin control of a nanocavity-enhanced qubit in diamond, and material science perspective, Acc. Chem. Res. 49, 2403 Nat. Commun. 6, 6173 (2015). (2016). [78] J. Riedrich-Moller,¨ S. Pezzagna, J. Meijer, C. Pauly, [97] J. Flick, M. Ruggenthaler, H. Appel, and A. Rubio, F. Mucklich,¨ M. Markham, A. M. Edmonds, and Kohn–sham approach to quantum electrodynamical density- C. Becher, Nanoimplantation and purcell enhancement of sin- functional theory: Exact time-dependent effective potentials in gle nitrogen-vacancy centers in photonic crystal cavities in di- real space, Proc. Natl. Acad. Sci. U.S.A. 112, 15285 (2015). amond, Appl. Phys. Lett. 106, 221103 (2015). [98] J. Flick, M. Ruggenthaler, H. Appel, and A. Rubio, Atoms [79] A. Sipahigil, R. E. Evans, D. D. Sukachev, M. J. Burek, J. Bor- and molecules in cavities, from weak to strong coupling in regaard, M. K. Bhaskar, C. T. Nguyen, J. L. Pacheco, H. A. quantum-electrodynamics (QED) chemistry, Proc. Natl. Acad. Atikian, C. Meuwly, R. M. Camacho, F. Jelezko, E. Biele- Sci. U.S.A. 114, 3026 (2017). jec, H. Park, M. Loncar,ˇ and M. D. Lukin, An integrated dia- [99] J. Feist, J. Galego, and F. J. Garcia-Vidal, Polaritonic chem- mond nanophotonics platform for quantum-optical networks, istry with organic molecules, ACS Photonics 5, 205 (2017). Science 354, 847 (2016). [100] M. A. Sentef, M. Ruggenthaler, and A. Rubio, Cavity [80] D. Riedel, I. Sollner,¨ B. J. Shields, S. Starosielec, P. Appel, quantum-electrodynamical polaritonically enhanced electron- E. Neu, P. Maletinsky, and R. J. Warburton, Deterministic phonon coupling and its influence on superconductivity, Sci. enhancement of coherent photon generation from a nitrogen- Adv. 4 (2018). vacancy center in ultrapure diamond, Phys. Rev. X 7, 031040 [101] F. Schlawin, A. Cavalleri, and D. Jaksch, Cavity-mediated (2017). electron-photon superconductivity, Phys. Rev. Lett. 122, [81] P. Lodahl, S. Mahmoodian, and S. Stobbe, Interfacing sin- 133602 (2019). gle photons and single quantum dots with photonic nanostruc- [102] J. B. Curtis, Z. M. Raines, A. A. Allocca, M. Hafezi, and tures, Rev. Mod. Phys. 87, 347 (2015). V. M. Galitski, Cavity quantum eliashberg enhancement of su- [82] B. Hacker, S. Welte, G. Rempe, and S. Ritter, A photon– perconductivity, Phys. Rev. Lett. 122, 167002 (2019). photon quantum gate based on a single atom in an optical [103] E. Orgiu, J. George, J. Hutchison, E. Devaux, J. Dayen, resonator, Nature 536, 193 (2016). B. Doudin, F. Stellacci, C. Genet, J. Schachenmayer, [83] S. Sun, H. Kim, Z. Luo, G. S. Solomon, and E. Waks, A C. Genes, G. Pupillo, P. Samori, and T. W. Ebbesen, Conduc- single-photon switch and transistor enabled by a solid-state tivity in organic semiconductors hybridized with the vacuum quantum memory, Science 361, 57 (2018). field, Nat. Mater. 14, 1123 (2015). [84] I. Aharonovich, D. Englund, and M. Toth, Solid-state single- [104] J. Feist and F. J. Garcia-Vidal, Extraordinary exciton con- photon emitters, Nat. Photon. 10, 631 (2016). ductance induced by strong coupling, Phys. Rev. Lett. 114, [85] X. Ding, Y. He, Z.-C. Duan, N. Gregersen, M.-C. Chen, S. Un- 196402 (2015). sleber, S. Maier, C. Schneider, M. Kamp, S. Hofling,¨ C.-Y. Lu, [105] J. Schachenmayer, C. Genes, E. Tignone, and G. Pupillo, and J.-W. Pan, On-demand single photons with high extrac- Cavity-enhanced transport of excitons, Phys. Rev. Lett. 114, tion efficiency and near-unity indistinguishability from a res- 196403 (2015). onantly driven quantum dot in a micropillar, Phys. Rev. Lett. [106] D. Hagenmuller,¨ J. Schachenmayer, S. Schutz,¨ C. Genes, and 116, 020401 (2016). G. Pupillo, Cavity-enhanced transport of charge, Phys. Rev. [86] H. Snijders, J. A. Frey, J. Norman, V. P. Post, A. C. Gos- Lett. 119, 223601 (2017). sard, J. E. Bowers, M. P. van Exter, W. Loffler,¨ and [107] X. Zhong, T. Chervy, L. Zhang, A. Thomas, J. George, D. Bouwmeester, Fiber-coupled cavity-QED source of iden- C. Genet, J. A. Hutchison, and T. W. Ebbesen, Energy trans- tical single photons, Phys. Rev. Applied 9, 031002 (2018). fer between spatially separated entangled molecules, Angew. [87] S. Daiss, S. Welte, B. Hacker, L. Li, and G. Rempe, Single- Chem. Int. Ed. 56, 9034 (2017). photon distillation via a photonic parity measurement using [108] D. Hagenmuller,¨ S. Schutz,¨ J. Schachenmayer, C. Genes, and cavity QED, Phys. Rev. Lett. 122, 133603 (2019). G. Pupillo, Cavity-assisted mesoscopic transport of fermions: [88] S.-B. Zheng and G.-C. Guo, Efficient scheme for two-atom Coherent and dissipative dynamics, Phys. Rev. B 97, 205303 entanglement and quantum information processing in cavity (2018). QED, Phys. Rev. Lett. 85, 2392 (2000). [109] G. L. Paravicini-Bagliani, F. Appugliese, E. Richter, F. Val- [89] L.-M. Duan and H. J. Kimble, Scalable photonic quantum morra, J. Keller, M. Beck, N. Bartolo, C. Rossler,¨ T. Ihn, computation through cavity-assisted interactions, Phys. Rev. K. Ensslin, C. Ciuti, G. Scalari, and J. Faist, Magneto- Lett. 92, 127902 (2004). transport controlled by Landau polariton states, Nat. Phys. 15, [90] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Cavity 186 (2019). optomechanics, Rev. Mod. Phys. 86, 1391 (2014). [110] J. Li, D. Golez, G. Mazza, A. Millis, A. Georges, and [91] W. L. Barnes, A. Dereux, and T. W. Ebbesen, Surface plasmon M. Eckstein, Electromagnetic coupling in tight-binding mod- subwavelength optics, Nature 424, 824 (2003). els for strongly correlated light and matter, arXiv:2001.09726 [92] J. M. Pitarke, V. M. Silkin, E. V. Chulkov, and P. M. (2020). Echenique, Theory of surface plasmons and surface-plasmon [111] R. J. Holmes and S. R. Forrest, Strong exciton-photon coupling polaritons, Rep. Prog. Phys. 70, 1 (2006). and exciton hybridization in a thermally evaporated polycrys- [93] M. S. Tame, K. McEnery, S¸. Ozdemir,¨ J. Lee, S. Maier, and talline film of an organic small molecule, Phys. Rev. Lett. 93, M. Kim, Quantum plasmonics, Nat. Phys. 9, 329 (2013). 186404 (2004). [94] J. A. Hutchison, T. Schwartz, C. Genet, E. Devaux, and T. W. [112]S.K ena-Cohen´ and S. Forrest, Room-temperature polariton Ebbesen, Modifying chemical landscapes by coupling to vac- lasing in an organic single-crystal microcavity, Nat. Photon. uum fields, Angew. Chem. Int. Ed. 51, 1592 (2012). 4, 371 (2010). [95] J. Galego, F. J. Garcia-Vidal, and J. Feist, Cavity-induced [113] T. Chervy, J. Xu, Y. Duan, C. Wang, L. Mager, M. Frerejean, modifications of molecular structure in the strong-coupling J. A. Mnninghoff, P. Tinnemans, J. A. Hutchison, C. Genet, regime, Phys. Rev. X 5, 041022 (2015). A. E. Rowan, T. Tasing, and T. W. Ebbesen, High-efficiency [96] T. W. Ebbesen, Hybrid light–matter states in a molecular second-harmonic generation from hybrid light-matter states, 26

Nano Lett. 16, 7352 (2016). [133] S. Hayashi and T. Okamoto, Plasmonics: visit the past to know [114] K. Stranius, M. Hertzog, and K. Borjesson,¨ Selective manip- the future, J. Phys. D 45, 433001 (2012). ulation of electronically excited states through strong light- [134] E. N. Economou, Surface plasmons in thin films, Phys. Rev. matter interactions, Nat. Commun. 9, 2273 (2018). 182, 539 (1969). [115] D. M. Coles, Y. Yang, Y. Wang, R. T. Grant, R. A. Taylor, [135] L. Novotny and B. Hecht, Principles of Nano-Optics (Cam- S. K. Saikin, A. Aspuru-Guzik, D. G. Lidzey, J. K.-H. Tang, bridge University Press, Cambridge, 2006). and J. M. Smith, Strong coupling between chlorosomes of pho- [136] C. Maissen, G. Scalari, F. Valmorra, M. Beck, J. Faist, tosynthetic bacteria and a confined optical cavity mode, Nat. S. Cibella, R. Leoni, C. Reichl, C. Charpentier, and Commun. 5, 5561 (2014). W. Wegscheider, Ultrastrong coupling in the near field of com- [116] A. Thomas, J. George, A. Shalabney, M. Dryzhakov, S. J. plementary split-ring resonators, Phys. Rev. B 90, 205309 Varma, J. Moran, T. Chervy, X. Zhong, E. Devaux, C. Genet, (2014). J. A. Hutchison, and T. W. Ebbesen, Ground-state chemical [137] P. Pretre,ˆ L.-M. Wu, R. A. Hill, and A. Knoesen, Characteri- reactivity under vibrational coupling to the vacuum electro- zation of electro-optic films by use of decal-deposited magnetic field, Angew. Chem. Int. Ed. 55, 11462 (2016). reflection fabry–perot microcavities, J. Opt. Soc. Am. B 15, [117] A. Thomas, L. Lethuillier-Karl, K. Nagarajan, R. M. A. 379 (1998). Vergauwe, J. George, T. Chervy, A. Shalabney, E. Devaux, [138] T. Sagawa, Second law-like inequalities with quantum rela- C. Genet, J. Moran, and T. W. Ebbesen, Tilting a ground-state tive entropy: An introduction, edited by M. Nakahara and reactivity landscape by vibrational strong coupling, Science S. Tanaka (World Scientific, Singapore, 2012). 363, 615 (2019). [139] R. P. Feynman, Statistical Mechanics (CRC Press, Florida, [118] R. Chikkaraddy, B. De Nijs, F. Benz, S. J. Barrow, O. A. 1998). Scherman, E. Rosta, A. Demetriadou, P. Fox, O. Hess, and [140] J. M. Edge, Y. Kedem, U. Aschauer, N. A. Spaldin, and J. J. Baumberg, Single-molecule strong coupling at room tem- A. V. Balatsky, Quantum critical origin of the superconduct- perature in plasmonic nanocavities, Nature 535, 127 (2016). ing dome in SrTiO3, Phys. Rev. Lett. 115, 247002 (2015). [119] A. Thomas, E. Devaux, K. Nagarajan, T. Chervy, M. Sei- [141] W. Jantsch, A. Bussmann-Holder, H. Bilz, and P. Vogl, Dy- del, D. Hagenmuller,¨ S. Schutz,¨ J. Schachenmayer, C. Genet, namical Properties of IV-VI Compounds, edited by G. Holer G. Pupillo, and T. W. Ebbesen, Exploring Superconductiv- (Springer, Berlin, 2001). ity under Strong Coupling with the Vacuum Electromagnetic [142] M. Kozina, M. Fechner, P. Marsik, T. van Driel, J. M. Glow- Field, arXiv:1911.01459 (2019). nia, C. Bernhard, M. Radovic, D. Zhu, S. Bonetti, U. Staub, [120] H. Yu, Y. Peng, Y. Yang, and Z.-Y. Li, Plasmon-enhanced and M. C. Hoffmann, Terahertz-driven phonon upconversion light–matter interactions and applications, npj Comput. in SrTiO3, Nat. Phys. 15, 387 (2019). Mater. 5, 1 (2019). [143] We note that the temperature is specified to be a concrete value [121] J. J. Baumberg, J. Aizpurua, M. H. Mikkelsen, and D. R. in order to fix a length scale in the setup of interest. Smith, Extreme nanophotonics from ultrathin metallic gaps, [144] J. R. Tischler, M. S. Bradley, V. Bulovic, J. H. Song, and Nat. Mater. 18, 668 (2019). A. Nurmikko, Strong coupling in a microcavity LED, Phys. [122] S. Rowley, L. Spalek, R. Smith, M. Dean, M. Itoh, J. Scott, Rev. Lett. 95, 036401 (2005). G. Lonzarich, and S. Saxena, Ferroelectric quantum critical- [145] T. Schwartz, J. A. Hutchison, C. Genet, and T. W. Ebbesen, ity, Nat. Phys. 10, 367 (2014). Reversible switching of ultrastrong light-molecule coupling, [123] D. E. Chang, A. S. Sørensen, E. A. Demler, and M. D. Lukin, Phys. Rev. Lett. 106, 196405 (2011). A single-photon transistor using nanoscale surface plasmons, [146] D. I. Schuster, L. S. Bishop, I. L. Chuang, D. DeMille, and Nat. Phys. 3, 807 (2007). R. J. Schoelkopf, Cavity qed in a molecular ion trap, Phys. [124] R. Ruppin and R. Englman, Optical phonons of small crystals, Rev. A 83, 012311 (2011). Rep. Prog. Phys. 33, 149 (1970). [147] V. Buzek,ˇ G. Drobny,´ M. S. Kim, G. Adam, and P. L. Knight, [125] G. D. Mahan, in Ionic Crystals and Polar Semicon- Cavity qed with cold trapped ions, Phys. Rev. A 56, 2352 ductors, edited by J. T. Devreese (North-Holland, Amsterdam, (1997). 1972). [148] G. S. Agarwal, Vacuum-field rabi splittings in microwave ab- [126] X. Li, M. Bamba, Q. Zhang, S. Fallahi, G. C. Gardner, W. Gao, sorption by rydberg atoms in a cavity, Phys. Rev. Lett. 53, M. Lou, K. Yoshioka, M. J. Manfra, and J. Kono, Vacuum 1732 (1984). Bloch–Siegert shift in Landau polaritons with ultra-high co- [149] C. R. Dean, A. F. Young, I. Meric, C. Lee, L. Wang, S. Sorgen- operativity, Nat. Photon. 12, 324 (2018). frei, K. Watanabe, T. Taniguchi, P. Kim, K. L. Shepard, and [127] T. G. Philbin, Canonical quantization of macroscopic electro- J. Hone, Boron nitride substrates for high-quality graphene magnetism, New J. Phys. 12, 123008 (2010). electronics, Nat. Nanotech. 5, 722 (2010). [128] P. Y. Yu and M. Cardona, Fundamentals of Semiconductors [150] G. Constantinescu, A. Kuc, and T. Heine, Stacking in bulk and (Springer, New York, 2010). bilayer hexagonal boron nitride, Phys. Rev. Lett. 111, 036104 [129] J. Vuckovic, M. Loncar, and A. Scherer, Surface plasmon (2013). enhanced light-emitting diode, IEEE J. Quantum Electron. 36, [151] A. J. Giles, S. Dai, I. Vurgaftman, T. Hoffman, S. Liu, L. Lind- 1131 (2000). say, C. T. Ellis, N. Assefa, I. Chatzakis, T. L. Reinecke, J. G. [130] D. Koller, A. Hohenau, H. Ditlbacher, N. Galler, F. Reil, Tischler, M. M. Fogler, J. H. Edgar, D. N. Basov, and J. D. F. Aussenegg, A. Leitner, E. List, and J. Krenn, Organic Caldwell, Ultralow-loss polaritons in isotopically pure boron plasmon-emitting diode, Nat. Photon. 2, 684 (2008). nitride, Nat. Mater. 17, 134 (2018). [131] J. Lambe and S. L. McCarthy, Light emission from inelastic [152] Y. Yamamoto, F. Tassone, and H. Cao, Semiconductor Cavity electron tunneling, Phys. Rev. Lett. 37, 923 (1976). Quantum Electrodynamics (Springer, New York, 2002). [132] S. Astilean, P. Lalanne, and M. Palamaru, Light transmission [153] A. V. Kavokin, J. J. Baumberg, G. Malpuech, and F. P. Laussy, through metallic channels much smaller than the wavelength, Microcavities (Oxford Science Publications, Oxford, 2007). Opt. Commun. 175, 265 (2000). [154] J. Kasprzak, M. Richard, S. Kundermann, A. Baas, P. Jeam- 27

brun, J. Keeling, F. Marchetti, M. Szymanska,´ R. Andre,´ action in ultracold gases, Phys. Rev. A 94, 053615 (2016). J. Staehli, V. Savona, P. Littlewood, B. Deveaud, and L. Dang, [166] L. M. Sieberer, S. D. Huber, E. Altman, and S. Diehl, Dynam- Bose–Einstein condensation of exciton polaritons, Nature 443, ical critical phenomena in driven-dissipative systems, Phys. 409 (2006). Rev. Lett. 110, 195301 (2013). [155] T. Byrnes, N. Y. Kim, and Y. Yamamoto, Exciton–polariton [167] J. Marino and S. Diehl, Driven markovian quantum criticality, condensates, Nat. Phys. 10, 803 (2014). Phys. Rev. Lett. 116, 070407 (2016). [156] D. Sanvitto and S. Kena-Cohen,´ The road towards polaritonic [168] C.-E. Bardyn, M. A. Baranov, C. V. Kraus, E. Rico, devices, Nat. Mater. 15, 1061 (2016). A. Imamo˙ glu,˘ P. Zoller, and S. Diehl, Topology by dissipa- [157] I. D. Leroux, M. H. Schleier-Smith, and V. Vuletic,´ tion, New J. Phys. 15, 085001 (2013). Orientation-dependent entanglement lifetime in a squeezed [169] H. Dehghani, T. Oka, and A. Mitra, Dissipative floquet topo- atomic clock, Phys. Rev. Lett. 104, 250801 (2010). logical systems, Phys. Rev. B 90, 195429 (2014). [158] J. Borregaard, E. J. Davis, G. S. Bentsen, M. H. Schleier- [170] L. Lu, J. D. Joannopoulos, and M. Soljaciˇ c,´ Topological pho- Smith, and A. S. Sørensen, One- and two-axis squeezing of tonics, Nat. Photon. 8, 821 (2014). atomic ensembles in optical cavities, New J. Phys. 19, 093021 [171] T. Karzig, C.-E. Bardyn, N. H. Lindner, and G. Refael, Topo- (2017). logical polaritons, Phys. Rev. X 5, 031001 (2015). [159] J. Hu, W. Chen, Z. Vendeiro, A. Urvoy, B. Braverman, and [172] Z. Gong, Y. Ashida, K. Kawabata, K. Takasan, S. Hi- V. Vuletic,´ Vacuum spin squeezing, Phys. Rev. A 96, 050301 gashikawa, and M. Ueda, Topological phases of non- (2017). hermitian systems, Phys. Rev. X 8, 031079 (2018). [160] M. A. Norcia, R. J. Lewis-Swan, J. R. K. Cline, B. Zhu, A. M. [173] C.-E. Bardyn, L. Wawer, A. Altland, M. Fleischhauer, and Rey, and J. K. Thompson, Cavity-mediated collective spin- S. Diehl, Probing the topology of density matrices, Phys. Rev. exchange interactions in a strontium superradiant laser, Sci- X 8, 011035 (2018). ence 361, 259 (2018). [174] R. W. Damon and J. R. Eshbach, Magnetostatic modes of a [161] J. A. Hertz, Quantum critical phenomena, Phys. Rev. B 14, ferromagnet slab, J. Phys. Chem. Solids 19, 308 (1961). 1165 (1976); A. J. Millis, Effect of a nonzero temperature [175] W. Eerenstein, N. Mathur, and J. F. Scott, Multiferroic and on quantum critical points in itinerant fermion systems, Phys. magnetoelectric materials, Nature 442, 759 (2006). Rev. B 48, 7183 (1993); T. Moriya, Spin Fluctuations in Itin- [176] J. Ruhman and P. A. Lee, Superconductivity at very low den- erant Electron Magnetism (Springer-Verlag, New York, 1985). sity: The case of strontium titanate, Phys. Rev. B 94, 224515 [162] A. Mitra, S. Takei, Y. B. Kim, and A. J. Millis, Nonequi- (2016). librium quantum criticality in open electronic systems, Phys. [177] D. G. Lidzey, D. Bradley, M. Skolnick, T. Virgili, S. Walker, Rev. Lett. 97, 236808 (2006). and D. Whittaker, Strong exciton–photon coupling in an or- [163] T. Prosen and I. Pizorn,ˇ Quantum phase transition in a far- ganic semiconductor microcavity, Nature 395, 53 (1998). from-equilibrium steady state of an XY spin chain, Phys. Rev. [178] H. Bahsoun, T. Chervy, A. Thomas, K. Brjesson, M. Hertzog, Lett. 101, 105701 (2008). J. George, E. Devaux, C. Genet, J. A. Hutchison, and T. W. [164] S. Diehl, M. Baranov, A. J. Daley, and P. Zoller, Observability Ebbesen, Electronic light–matter strong coupling in nanoflu- of quantum criticality and a continuous in atomic idic fabry–perot´ cavities, ACS Photonics 5, 225 (2017). gases, Phys. Rev. Lett. 104, 165301 (2010). [179] Y. Kaluzny, P. Goy, M. Gross, J. M. Raimond, and S. Haroche, [165] Y. Ashida, S. Furukawa, and M. Ueda, Parity-time-symmetric Observation of self-induced Rabi oscillations in two-level quantum critical phenomena, Nat. Commun. 8, 15791 (2017); atoms excited inside a resonant cavity: The ringing regime Quantum critical behavior influenced by measurement back- of superradiance, Phys. Rev. Lett. 51, 1175 (1983).