Single quantum emitter Dicke enhancement

Tommaso Tufarelli,1, ∗ Daniel Friedrich,2 Heiko Groß,2 Joachim Hamm,3 Ortwin Hess,3, 4 and Bert Hecht2, † 1School of Mathematical Sciences and Centre for the Mathematics and Theoretical Physics of Quantum Non-Equilibrium Systems, University of Nottingham, Nottingham, NG7 2RD, United Kingdom 2Nano-Optics and Biophotonics Group, Experimentelle Physik 5, Physikalisches Institut, Universit¨atW¨urzburg, D-97074 W¨urzburg, 3The Blackett Laboratory, Department of Physics, , London SW7 2AZ, United Kingdom 4School of Physics and CRANN Institute, Trinity College Dublin, Dublin 2, Ireland

Coupling N identical emitters√ to the same field mode is well-established method to enhance light matter inter- action. However, the resulting N-boost of the coupling strength comes at the cost of a “linearized” (effectively semi-classical) dynamics. Here, we instead demonstrate a new approach for enhancing the coupling constant of a single quantum emitter, while retaining the nonlinear character of the light-matter interaction. We consider a single quantum emitter with N nearly degenerate transitions that are collectively coupled to the same field mode. We show that√ in such conditions an effective Jaynes-Cummings model emerges, with a boosted coupling constant of order N. The validity and consequences of our general conclusions are analytically demonstrated for the instructive case N = 2. We further observe that our system can closely match the spectral line shapes and photon autocorrelation functions typical of Jaynes-Cummings physics, hence proving that quantum optical nonlinearities are retained. Our findings match up very well with recent broadband plasmonic nanoresonator strong-coupling experiments and will therefore facilitate the control and detection of single-photon nonlineari- ties at ambient conditions.

I. INTRODUCTION In this paper we indeed show how the exploitation of a multilevel emitter, with N nearly degenerate excited states Strongly coupled light-matter systems, described by the coupled to the same field mode, increases the√ effective light- Jaynes-Cummings (JC) model [1], are of fundamental inter- matter coupling constant by a factor of oredr N. While this est in testing the quantised of coupled light and matter is the same scaling found in the Dicke [12] and TC models degrees of freedom in the context of cavity QED [2]. Im- [10] with N emitters, the crucial difference is that in our case portantly, JC physics requires that a single two-level emit- the quantum non-linearity of the interaction is preserved: we ter interacts with a single-mode electromagnetic field. Such demonstrate this based on emission spectra and second-order systems can be relevant for quantum information processing, photon autocorrelation functions. More precisely, our system for example in interfacing flying and static qubits [3] or for can closely approximate the behaviour of a JC model, in prin- quantum-state engineering [2]. Yet strong coupling of light ciple, for arbitrary values of N. These findings, in their sim- and a single emitter is not easy to achieve in practice, due plest form, are schematised in Fig 1. to the large mismatch in the spatial extension of free-space Additionally, we find that our predictions are robust against photons and typical emitters (i.e. artificial or real atoms) that typical sources of decoherence (dephasing) that are common feature electronic transitions in the optical domain [4, 5]. at elevated temperatures. Thus, while our results are fully gen- A possible route to overcome this difficulty is to engineer eral and independent of the physical implementation of the an “effective” emitter with large dipole moment. This is most model, we anticipate that our approach can be particularly commonly implemented by combining N  1 identical emit- useful in the context of room temperature strong coupling. ters coupled to the same field, resulting in the well-known We thus find it useful to provide some context on the current “Dicke enhancement” of light-matter interaction: the√ effective state-of-the-art of this research field, and to highlight how our coupling strength is boosted by a factor of order N, com- research can help it move forward. pared to the single-emitter case [6–9]. However, the relevant Recently, room-temperature strong coupling with many theory in such systems is the many-emitter limit of the Tavis- arXiv:2010.12585v2 [quant-ph] 25 Mar 2021 emitters [13–16] and even single quantum emitters [17–21] Cummings (TC) model [10], whose low-energy dynamics is have been experimentally demonstrated. These experiments effectively linear, i.e., indistinguishable from that of coupled rely on the remarkable coupling strengths achievable in plas- harmonic oscillators [11]. monic nanoresonators, thanks to the deeply sub-wavelength The key achievement of our work is to bypass this seemingly mode volumes of the latter. In order to relax the mode vol- inevitable trade-off between coupling strength and (quantum) ume requirements (which are currently pushing the limits of nonlinearity: we open an innovative route to a ‘Dicke-like’ en- nanofabrication techniques) and obtain more robust experi- hancement of the coupling constant which preserves the pre- mental realizations, it would be of great interest to find emitter cious quantum optical nonlinearities of the JC model. systems with increased dipole moments. This would result in larger coupling constants while still allowing to harness the single-emitter quantum nonlinearity of the JC model (JCM) [22, 23]. So far, only the “first rung” of the JC ladder, i.e. the ∗ [email protected] single-excitation regime of the model, has been experimen- † [email protected] tally detected at ambient conditions. Efforts are still ongoing 2

Jaynes-Cummings Model Multilevel Model Tavis-Cummings/Dicke model

single-emitter cavity mode single-emitter cavity mode multiple emitter cavity mode transition transitions transitions

2 2 2

1 1 1

0 0 0

anharmonic spectrum anharmonic spectrum harmonic spectrum with boosted coupling with boosted coupling

FIG. 1. Visual summary of the light-matter interaction models treated here. For multilevel, TC and Dicke models, only energy levels that have a dirrect correspondence in the JCM are shown. Left: Standard Jaynes-Cummings model, in which a single two-level emitter couples resonantly to a single mode field, with strength g. The resulting anharmonic spectrum is a cornerstone of modern quantum optics, but challenging to probe experimentally: the required regime g  κ is difficult to achieve, where κ is the cavity decay rate. Center: Multilevel model, the focus of this paper. A quasi-degenerate multilevel emitter, with N closely spaced excited states, interacts with a single√ mode field. When all the transitions are approximately resonant with the field, an effective coupling constant geff, enhanced by a factor ∝ N, is established. In principle, the resulting low energy spectrum can closely approximate the JC spectrum for any value of N. Right: Tavis-Cummings and√ Dicke Models, in which N  1 nearly-identical two level systems interact with the same field. A boosted effective coupling constant geff ∝ N is established also in this case, but the spectrum becomes approximately harmonic as N is increased. to detect higher rungs of the ladder to shed light on (and har- interpretation of such experiments. Importantly, our approach ness) the quantum non-linearity of the model. lives independently of the contentious issues regarding the In order to realise the JCM experimentally, great efforts calculation of mode volumes and coupling constants in quan- have been made over the years in the fabrication of artificial tum (and classical) nanophotonics, see for example [28–30]. atoms — such as semiconductor quantum dots [24] — that are These approaches will of course determine the numerical val- close to ideal two-level emitters. Indeed, so far the multi-level ues of coupling constants in different experimental situations, nature of such emitters has mainly been treated as a modelling but they do not affect the nature of our quantum optical model nuisance, i.e. as a source of errors with respect to an idealised and the general conclusions we can draw from it. two-level scenario. In this context, Madsen et al. [25] and The paper is organised as follows. In Section II we intro- our previous paper [19], have recently mentioned the possi- duce a Hamiltonian that models an emitter with N excited lev- ble effects of a more complicated energy level structure of the els coupled to the same ground state via a single mode field. emitter. To the best of our knowledge, however, our previous We will show that our system can be understood as a JC model work [19] was the first to explicitly suggest that an emitter which is weakly interacting with N − 1 “dark states” of the with multiple (> 2) energy levels can even bring about some emitter. In Section III we generalise our model to an open advantages: in the right conditions it may be used to simulate system via a master equation. This takes into account inco- a JC system with enhanced coupling constant. With reports herent processes such as photon loss, dephasing and emitter of quantum dots featuring as much as 64 quasi-degenerate ex- pumping, all of which may be required for describing realistic cited states [26, 27], a general quantum-mechanical analysis experimental scenarios. In Section IV we support our general of such a system (multilevel model for brevity) is in order. arguments with detailed analytical results for the special case The findings presented here will thus be of importance for N = 2: we find that our system’s emission spectrum features the selection and the design of suitable emitters for nanoscale approximately the same resonances as a JC model, plus addi- strong-coupling experiments at ambient conditions and for the tional features (due to the dark states) that we fully character- 3 ize. Section V presents numerical studies for N = 3, compar- ing common experimental observables in our model with the same quantities calculated in reference JC and TC/Dicke mod- els. In Section VI we draw our conclusions. For complete- ness, Appendix A contains some useful reminders and for- mulas pertaining JC, TC and coupled-oscillator models, since those are referenced throughout the paper.

II. MODEL

Our multilevel model features a quantum emitter with elec- FIG. 2. Illustration of the transformation between the bare emitter basis and the radiation basis for N = 2 sublevels, and the resulting |Gi |e i tronic ground state and a collection of excited states k , appearance of a boosted effective coupling. The same principles hold also called sublevels, where k = 1, 2,..., N. Each internal for larger values of N. (a) In the bare emitter basis the two levels transition of the emitter, |Gi ↔ |e i, is coupled to a single- k |e1,2i are not directly coupled to each other, while ε plays the role mode field (cavity for brevity) with strength gk ∈ R. The typ- of a relative detuning between them. Both |Gi ↔ |e1,2i transitions ical physical system we have in mind is a multilevel quantum couple to the cavity field, and we are assuming for simplicity that dot embedded in a plasmonic nanoresonator, where the cavity the associated coupling constants are the same: g1 = g2 = g. (b) resonance is broadband, in the sense that it overlaps with all In the radiation basis, the bright and dark states |Bi, |Di have the the emitter transitions, and where strong field gradients may same frequency, but are coupled to each other with strength ε/2. The allow for light-matter couplings beyond the dipole approxi- transition√ |Gi ↔ |Bi is coupled to the field with enhanced strength mation (hence the possibility of coupling a large number of geff = 2g, while the dark state |Di does not directly interact with levels). Yet, it is important to note that our quantum optical the field. formalism applies to any type of emitter or resonator that sat- isfy our modelling assumptions. As usual the cavity mode is A. Bare emitter basis versus radiation basis described via the annihilation operatora ˆ, with [ˆa, aˆ†] = 1. Taking ~ = 1 for convenience, the cavity resonance has To aid understanding, the main ideas behind this subsection average frequency (hence energy) ω0, the energy of |Gi can are illustrated in Fig. 2 for the simplest case, N = 2. We begin be set to zero without loss of generality, while each state |eki our study of the general N-sublevel model by considering the has energy ω0 + ∆k. Here ∆k is the detuning between the k-th case of exact degeneracy of the excited states, ∆k = ∆ ∀k, or emitter transition (|eki → |Gi) and the cavity. For convenience equivalently ε = 0 in terms of the full range of energy detun- we assume ∆1 ≤ ∆2 ≤ ... ≤ ∆N , we indicate the full range of ings. The resulting mathematics follow in a straightforward energy detunings as ε ≡ ∆N − ∆1 and the average detuning as manner from the well-known JCM. This simple starting point 1 P ∆ = N k ∆k. we can write the total Hamiltonian as will set the stage to investigate small deviations from an ideal- ized scenario, i.e. cases with small but nonzero ε. For ε = 0, we find that any linear combination of the states |e i is itself a H = H0 + V, (1) k valid emitter excited state, or more precisely, an eigenstate of Xn † H0 with eigenvalue ω0 + ∆. Among all the states that can be H0 = ω0aˆ aˆ + (ω0 + ∆k)|ekihek|, (2) k=1 constructed in this way, of particular significance is the coher- ent superposition XN   | ih | †| ih | P V = gk aˆ ek G + aˆ G ek , (3) k gk |eki k=1 |Bi = q , (4) P 2 j g j H , V where 0 are, respectively, the free and interaction Hamil- where the denominator ensures normalization. We shall refer tonians, and we have neglected counter-rotating terms in the to |Bi as the bright state, or superradiant state (mimicking the light-matter interaction. The physics of Hamiltonian (1) can terminology of the Dicke model). State |Bi can be interpreted H in general be quite complicated, despite being number- as the only emitter excited state that couples directly to the H conserving — i.e. commutes with the number operator cavity field. Indeed it is easy to check that the light-matter Nˆ a†a P |e ihe | tot = ˆ ˆ + k k k . However, here we are interested interaction term in Eq. (1) may be rewritten in the simple form in the special case of quasi-degenerate emitters that are near-  †  resonant with the cavity, i.e. we assume the parameter regime V = geff aˆ |BihG| + aˆ |GihB| , (5) |∆k|  gk. This approximation holds e.g. for colloidal quan- tum dots coupled to plasmonic nanoresonators. Here, ∆k are where we have defined an effective coupling constant on the order of a few meV, while the coupling strengths, gk, sX 2 are on the order of 100meV [19]. In this scenario a number of geff = gk. (6) qualitative observations can be made. k 4

Hence, through an appropriate change of basis for the emitter, the N degenerate transitions |Gi ↔ |eki can be replaced by a single transition |Gi ↔ |Bi, characterized by an enhanced coupling to the cavity field. Eq. (6) indeed√ shows that the ef- fective coupling scales like geff ∼ N with the number of sublevels. This boosting of the effective light-matter coupling is analogous to what happens in the Dicke [12] and TC [10] models, where N degenerate emitters are coupled to the same field. A crucial difference from the TC model is that Eq. (5) retains the structure of a JC interaction Hamiltonian, so that in the model studied here there is in principle no loss of anhar- monicity as the number of sublevels N is increased. To fully specify the mentioned change of basis, we can de- fine N −1 further linear combinations of the excited states |eki, say {|D1i, |D2i,..., |DN−1i}, which together with {|Gi, |Bi} FIG. 3. Schematic of the investigated setup. The system may be form a complete orthonormal set for the emitter internal lev- excited through an incoherent pump of intensity Λ applied to the els. When ε = 0 each |Dki is also an eigenstate of H0 with emitter, or coherently driving of the (plasmonic) cavity with field eigenvalue ω0 + ∆, due to the N-fold degeneracy. By con- amplitude Ω and frequency ωL. The bright state of the emitter is struction, these additional basis states do not directly interact coupled to the cavity with the effective strength geff , and the emitter with the cavity field, i.e. sublevels are spread in frequency over a range ε. κ is the loss rate of the cavity and γ the decay rate of all the emitter sublevels, which V|Dki = hDk|V = 0, k = 1, 2,..., N − 1, (7) are also subject to dephasing at rate γd (not shown). All the observ- ables discussed in this work can be measured by collecting the cavity and may be called subradiant states or dark states. In output field. practice, explicit expressions for the dark states may be found by standard orthogonalization procedures (details not shown). The above discussion suggests that the set sublevels. This can happen if incoherent processes couple the {|Gi, |Bi, |D1i, |D2i,..., |DN−1i}, which we will call radiation pair {|Gi, |Bi} to the dark states. We shall see in Section IV, for basis, may be a more meaningful conceptual tool to study example, that emitter dephasing provides a common mecha- light-matter interaction in our system, as compared to the bare nism for interconverting bright and dark states. Yet, our ana- emitter basis {|Gi, |e1i, |e2i,..., |eN i}. lytical and numerical results suggest that the multilevel model For the purpose of modelling the optical properties of the can provide an excellent approximation to JC physics even system, it is now tempting to simply neglect all the dark states when these imperfections are taken into account. and retain only the two emitter levels |Bi, |Gi. This is justified if there are no additional processes, coherent or incoherent, that are able to populate the dark states and/or couple them to III. MASTER EQUATION the two levels |Bi, |Gi. When all these conditions are satisfied, the system is exactly described by a JCM with coupling In order to predict relevant observables for future exper- constant geff. iments, we now introduce an open system generalisation of our model, via the well established Gorini-Kossakowski- We are now in a position to study small deviations from the Sudarshan-Lindblad (GKSL) master equation [31]. As antici- ideal case. As anticipated, the most obvious deviation from pated above, the main experimental scenario we have in mind pure JC physics occurs when the excited states |eki are not ex- is a quantum dot interacting with a plasmonic nanoresonator, actly resonant with each other (i.e. ε , 0). In the radiation yet the ideas presented here may be applicable to a wider va- basis, the (small) relative detunings between the levels |eki riety of setups. We will include cavity photon loss as the pri- will translate into (weak) couplings between the bright and mary channel for energy dissipation, while the decay rate of dark states, and between the dark states themselves. In other the emitter into free-space modes will be assumed negligible words, |Bi and |Dki cease to be eigenstates of H0 when ε , 0. (however, in simulations it will remain finite for reasons of Note however that Eqs. (5) and (7) remain valid in general, so numerical stability). Such an assumption is reasonable due to that the dark states may interact with the field only indirectly, the large β-factors achievable in plasmonic resonators. On the i.e. through the mediation of state |Bi. These additional inter- other hand, we shall assume that the emitter excited states can actions will bring about deviations from the well understood be broadened well beyond their radiative linewidth by addi- JC physics described above. While a general calculation of tional dephasing processes - typically thought of as a draw- these effects would be cumbersome and beyond the scopes of back of room temperature operation. We consider two ways this paper, the case N = 2 sublevels is simple enough to be in which energy can be supplied to the coupled system: in- treated analytically and already contains all the essential in- coherent pumping of the emitter, or a coherent driving field gredients in our problem — see Section IV below and Fig. 2. applied to the cavity. While the first is a common method to In some cases, the exact equivalence to a JC system may be feed excitations into quantum dots, the second is arguably the lost even in the case of perfect degeneracy between the excited standard avenue to test nonlinearities in a quantum optical sys- 5 √ tem. For a visual summary, the main ingredients of our open where in this case geff = 2g. As for the general case treated system model are sketched in Fig. 3. In formulas, the follow- in the previous section, a JC interaction describes the cou- ing master equation will be used to model the emitter-cavity pling between the two levels {|Gi, |Bi} and the cavity field. system, However we can also see that the relative detuning of the sub- | i X 1   levels, ε, translates into a small coupling between B and ρ˙ = −i[H + V , ρ] + 2LρL† − L†Lρ − ρL†L , (8) |Di, so that even the dark state |Di can indirectly influence drive 2 L∈J the optical behaviour of the system. The ground state of H | i where H is the Hamiltonian (1), and is easily spotted as G, 0 , with eingenvalue E = 0. Explot- ing excitation number conservation, i.e. [H, Nˆ ] = 0, where   tot iωLt † −iωLt ˆ † Vdrive = Ω aeˆ + aˆ e (9) Ntot = aˆ aˆ + |BihB| + |DihD|, one can show that the remaining eigenvalues and eigenvectors of H can be found by diagonal- is an interaction term describing coherent driving of the cavity ising the 3 × 3 matrices (n = 1, 2,..., ∞) by a classical laser at frequency ωL (Ω quantifies the driving field strength). The L’s appearing on the right hand side of  √   nω√0 geff n 0  Eq. (8) are the jump operators, used to model incoherent pro-   Hn = geff n nω0 + ∆ −ε/2  , (13) cesses. They are collected in the set J, which features the   0 −ε/2 nω0 + ∆ following elements: √ • Cavity decay (photon loss) at rate κ: L = κ aˆ which describe the Hamiltonian restricted to the subspace √ n o n o • Radiative decay of the emitter γ: L = γ |Gihek|, Nˆ = n ≡ Span |G, ni, |B, n − 1i, |D, n − 1i . (14) √ tot • Emitter dephasing γd: L = γd |ekihek|, q The behaviour of the eigenvalues of H1 is shown in Fig 4, as Λ a function of the detuning ∆. Even at the non-negligible ratio • Emitter pumping of (total) rate Λ: L = N |ekihG|, ε/geff = 0.25 used in the figure, the commonalities with a JCM where k = 1, 2,..., N. We assumed that the decay and de- spectrum are evident. Note also that the eigenvalues of Hn for phasing rates are the same for all sublevels, and that the pump n ≥ 2 will follow the same qualitative structure, as a generic power Λ is split symmetrically over the N sublevels. We also √ Hn √may be obtained from H1 by a simple rescaling geff → consider equal couplings gk = geff/ N for all k, while the de- geff n, ω0 → nω0. While a general analytical diagonalization tunings ∆k are equally spaced in the interval [∆−ε/2, ∆+ε/2]. of Eq. (13) is cumbersome, for the resonant case ∆ = 0 one All the assumed symmetries between the sublevels are of can obtain simple analytical expressions. For this case, the course an idealisation. Many generalisations of this basic eigenvalues of Hn read model can be included in principle. However, we do not ex- s pect that these can bring about significant qualitative depar- ε2 tures from the physics described here. Hence, in this work En,± = nω0 ± geff n + , (15) 4g2 we confine our analysis to the above setting for simplicity and eff computational efficiency. En,D = nω0. (16)

The corresponding normalised eigenvectors |En,±i, |En,Di are IV. ANALYTICAL RESULTS FOR 2 SUB-LEVELS    √  1  2geff n  We here present the analytical diagonalization of Hamilto- |E i = √  |G, ni ± |B, n − 1i n,±  q  nian (1) for the case N = 2 with equal couplings g1 = g2 = g, 2  2 2  4geffn + ε as well as the associated implications for the system’s open ε dynamics. This instance of the model, illustrated in Fig. 2, − q |D, n − 1i, (17) 2 2 is particularly transparent and fully captures the essence of 8geffn + 2ε our study. We verified that analytical solutions are also avail- √ able for the cases N = 3, 4, using similar methods, but these ε 2geff n |En,Di = q |G, ni + q |D, n − 1i. (18) are significantly more complex without adding fundamen- 4g2 n + ε2 4g2 n + ε2 tally new insights. For N = 2 the detunings are given by eff eff ∆ = ∆ − ε/2 and ∆ = ∆ + ε/2, respectively. Further- 1 2 The above expressions suggest that in the regime ε  g more there are only one bright and one dark state given by eff | i | i | i−| i it is natural to identify {|En,±i, En,±} as a weakly perturbed JC |Bi = e1 √+ e2 , |Di = e1 √ e2 . In the radiation basis representa- 2 2 eigensystem (the latter is indeed recovered exactly for ε → 0). tion, the Hamiltonian may be recast as For brevity, we refer to those as the ‘JC model-type’ eigenval- H = H0 + V0, (10) ues and eigenvectors. The eigenstates |En,Di, which instead 0 feature the same energy levels of an empty cavity, may be 0 † H0 = ω0aˆ aˆ + (ω0 + ∆) (|BihB| + |DihD|), (11) understood as dressed quasi-dark states of the emitter, accom-   ε − V0 = g aˆ |BihG| + aˆ†|GihB| − (|BihD| + |DihB|) , (12) panied by approximately n 1 cavity photons. They indeed eff 2 result from a hybridization of states |G, ni and |D, n − 1i, with 6

We may immediately notice that the lowest quasi-dark state, |E1Di, is long lived, while all the remaining excited states have approximately the same lifetime of either the JCM ex- cited states or the Fock states of an empty and lossy cavity. Next, emitter dephasing will induce further broadening of the Hamiltonian energy levels, which can be estimated as

 2 2 δE j ' γd |he1|E ji| + |he2|E ji| . (22)

FIG. 4. Eigenvalues of H1 as a function of the average detuning ∆, for Some straightforward algebra yields ε = 0.25geff (blue continuous lines). The anticrossing behaviour of eigenvalues E1,± around ∆ = 0 approximates well (the maximum rel-  2  ative error in the plotted range being ∼ 1.3%) the reference JCM with  ε  δEnD ' γd 1 −  , (23) coupling geff (red circles). The energy E1,D of the quasi-dark state  4ng2  is instead approximately linear in the detuning and sits in between eff  2  the JCM-like energies. An elementary perturbation theory calcula- γd  ε    δEn± ' 1 +  , (24) ' − 2 2 2  2  tion indeed yields En,D ∆ 1 ε /4geff n , which agrees well with 4ngeff the exact value in the considered regime (relative error below 0.2%). Qualitatively similar results are obtained for higher excitation num- bers n ≥ 2, and in fact it is easy to see that the relative weight of ε where again we have expanded our results at second order and ∆ decreases with n. in ε. In this case we can see that the quasi-dark states are the most affected, since they feature a larger overlap with the bare excited states of the emitter. Note also that the δE j’s the second state being dominant in the superposition. It is in- are not associated with a decay towards states of lower ex- teresting to notice that the two states involved in such superpo- citation number (more below). The exact broadening of the sition are only coupled indirectly in the Hamiltonian, through energy levels, induced by the combined effect of photon loss the mediation of state |B, n − 1i. and emitter dephasing, can be obtained as −2Im(E˜ j), where Exploiting the analytical results obtained here, it is possible to E˜ j are the eigenvalues of the non-Hermitian operator calculate the Hamiltonian dynamics of a generic initial state via standard methods. In fact, while our work focuses on the κ γ X regime ε  g , we note that the above diagonalization pro- H˜ = H − i aˆ†aˆ − d |e ihe |. (25) eff 2 2 k k cedure is valid for a general choice of the model parameters. k

The resulting expressions are unwieldy, however we checked A. Qualitative analysis of the master equation for N = 2 via a numerical optimization that −2Im(E˜ j) ' Γ j +δE j, with a relative error below ' 3.1% in the range 0 ≤ ε ≤ 0.25geff, 0 ≤ The closed expressions obtained so far are particularly κ ≤ 2geff, 0 ≤ γd ≤ 2geff, which is more than enough for our helpful in analysing the open system behaviour of our multi- scopes. level model, as well as its relationship to the JC model. Since In order to gain qualitative understanding of photon emis- in this work we focus on the regime of low pumping (or low sion processes in our system, we next look at the (squared) coherent driving) and negligible emitter decay, we can gain matrix elements of the annihilation operatora ˆ, valuable intuition about the properties of the master equation (8) by setting γ = Λ = Ω = 0. First, we observe that the 2 introduction of a photon loss rate κ results in a finite lifetime Ai→ f = |hE f |aˆ|Eii| , (26) for the excited states of H, such that eigenstates with Nˆ tot = n will tend to decay towards states with Nˆ = n − 1, until the tot between an initial energy eigenstate |E i and a final one |E i. ground state |G, 0i is finally reached. For a given eigenstate i f The quantum jump approach to GKSL master equations [31] |E i, an estimate of the decay rate induced by photon loss can j provides a useful interpretation for these quantities: A is be obtained via the perturbation theory formula i→ f proportional to the probability that the transition |Eii → |E f i † occurs upon loss of a cavity photon [in the master equation (8), Γ j ' κhE j|aˆ aˆ|E ji, (19) such transitions are triggered by the term κaˆρaˆ†]. When the which at second order in ε yields (n ≥ 1) transition occurs, a photon is emitted into extra-cavity modes   with (approximately) frequency ω f i = E f − Ei and bandwidth  ε2  Γ ' κ n − 1 +  , (20) Γi + Γ f + δEi + δE f . The Ai→ f terms involving only JCM- nD  2  4ngeff type eigenstates do not present any particular surprises: they  2  approximate the well-known JCM results for ε  geff, and  1 ε  Γn± ' κ n − −  . (21) converge to them for ε → 0. We thus omit such cases for  2 2  8ngeff brevity and focus only on the matrix elements involving quasi 7 dark states, which yield the following expressions: Our analysis focuses on two observables that are ubiquitous in both theoretical and experimental investigations of the JC ε2 A → = , (27) model (proper definitions will be given below): 1D 0 2 2 4geffn + ε • The steady state spectrum S a(ω) of the cavity output 2 4geffn field, when the system is excited via incoherent pump- A → − = n − , (28) nD n 1,D 2 2 ing of the emitter; 4geffn + ε (2) AnD→n−1,± = 0, (29) • The photon autocorrelation function g (0), when the 2 2 emitter is excited via a coherent driving field with vari- 2geffε An±→n−1,D =    . (30) able detuning. 4g2 (n − 1) + ε2 4g2 n + ε2 eff eff The first observable provides valuable information about the From Eqs. (27), (28) and (29) we notice that, once the sys- energy level structure of a quantum system, in particular high- (2) tem is in one of the quasi-dark states |EnDi, the loss of a cav- lighting its optically allowed transitions. Instead, the g (0) ity photon can only produce the quasi dark state |En−1,Di for may be seen as a witness of optical√ non-linearity: we will use n ≥ 2, or |G, 0i for n = 1. This corresponds to the emission of it in our model to show√ that the N enhancement of the cou- a photon at the cavity frequency ω0, with the associated ma- pling constant, geff ∝ N, does not come at the cost of “lin- trix element of the order ∼ n − 1. Assuming for the moment earising” the light-matter interaction (something that instead γd = 0 (no dephasing), this implies that once the system has happens in the TC scenario as N is increased, as we shall see evolved into a quasi-dark state, it essentially reproduces the below). To support this last point, we will indeed compare open dynamics of an empty cavity until state |1, Di is reached. our autocorrelation functions with the g(2)(0) obtained for a From there the system will decay slowly towards the low- reference “coupled oscillator model” — i.e. the quintessen- est state |G, 0i, resulting again in the emission of a photon tial linear system in our context (see Appendix A for details). at the cavity frequency ω0, but with a much narrower band- Note that we are focusing only on observables that can be 2 2 width Γ1D ' κε /4geff. On the other hand, from Eq. (30) we measured via the cavity output field. The latter has indeed can also see a small probability that a JCM-like state decays the same definition in all models considered here, facilitating into a quasi-dark state by loss of a cavity photon. This pro- direct comparisons between them. On the other hand, emitter cess results√ in the emission of photons at average frequency observables do not have an unambiguous correspondence in ω0 ±geff n: these exactly overlap with the JCM spectral reso- the different models, as they are defined on Hilbert spaces of nances for n = 1, while they provide novel emission peaks for different dimensions. n ≥ 2. Putting back finite dephasing γd , 0 into the picture, In detail, according to standard input-output theory [32], the we see that, in addition to introducing further broadenings portion of the cavity output field that reaches our detector, say of the levels, dephasing events are capable of interconverting Eˆout (scalar for simplicity) may be expressed as a function of | i | i |h | | i|2 states B√and D . More in detail, we find f L i = γd/2 the cavity annihilation operatora ˆ as follows: for L = γ |e ihe | with k = 1, 2 and i, f ∈ {D, B}. Referring d k k ˆ ˆ again to the quantum jump approach [31] this will swap the Eout = Ecaˆ + Evac, (31) states |Bi and |Di with probability 1/2, if a dephasing event where Ec is a constant depending on the details of the cavity- occurs. plus-detector setup, while the operator Eˆvac captures all “vac- In summary, we have provided a detailed characterization of uum noise” contributions from the detector’s environment. the processes occurring in our open system for the emblem- Thanks to the assumption of a vacuum environment, which atic case N = 2, ∆ = Λ = γ = Ω = 0. While the multilevel is justified for optical frequencies at room temperature, Eˆvac model can approximate all features of a JCM for ε  geff, it does not contribute to any normally ordered observable con- displays additional phenomena due to the indirect interaction structed from Eˆout — and these are precisely the types of ob- of dark states and cavity field. In the next section we explore servables employed here. these issues quantitatively via numerical simulations. In the interest of brevity, we restrict ourselves to an experi- mentally desirable situation in which ∆ = 0 and the hierarchy ε  κ . geff holds. In other words we consider a strongly V. CLIMBING THE JC LADDER coupled system where the average transition frequency of the emitter is at resonance with the cavity, while the range of sub- In the following numerical studies, we will investigate how level detunings is within the bare cavity bandwidth. This sce- closely the physics of our system – as relevant for typical nario is indeed appropriate for the case of quantum dots em- quantum optics experiments – resembles√ a JCM with en- bedded in plasmonic nanoresonators, as anticipated in Sec- hanced coupling constant geff ∝ N. For the sake of com- tion II. We shall also take the driving strengths (Ω for the parison, we shall also display the behaviour of a TC model coherent and Λ for the incoherent case) as the smallest pa- with N atoms and the same effective coupling constant geff rameters in the problem. This ensures that the driving will not — see Appendix A for further details. Simulations are done induce significant modifications to the intrinsic resonances of for N = 3: this is one step up in complexity compared to the the coupled system, nor to the conditions for achieving strong previous section, but simple enough that the resulting calcula- coupling [33, 34]. In this setting we focus on two representa- tions can be handled by a standard computer. tive parameter regimes: 8

Single-excitation regime - κ = geff

Bi-excitation regime - κ = 0.05geff

FIG. 5. Comparison of cavity emission spectra, S a(ω), as obtained for our multilevel model, versus reference JC and TC models. The sub-level energy spread is fixed at ε = 0.05geff in all plots. Spectra are obtained under incoherent emitter pumping, i.e. Λ , 0, Ω = 0. The top plots in the “single-excitation regime” are obtained for κ = geff and Λ = 0.01κ, while the “bi-excitation regime” at the bottom is for κ = 0.05geff and Λ = 0.02κ. Dephasing increases from left to right. Solid blue lines indicate our multi-level model, dashed red lines show the quantities of interest for a reference JC model, while green dot-dashed lines are for the TC model. The spectra of the TC and JC models are rescaled by a factor (1 − pdark). Gray vertical lines indicate the resonances corresponding to the first (dot-dashed) and second rung (dotted) of the JC ladder.

(i) geff = κ, or ‘single-excitation regime’, weaker features displayed in our plots (e.g. the second-rung spectral resonances) may be challenging to detect in experi- (ii) geff = 20κ, or ‘bi-excitation regime’. ments.

Loosely speaking, case (i) gives access to the ‘first rung’ of the energy ladder, thus the resulting Physics should be simi- lar to that of a pair of coupled oscillators in all models. Case (ii) instead allows us to explore the ‘second rung’ of the lad- der, where the anharmonic nature of the investigated models becomes more prominent. In both situations we shall anal- yse the role of emitter dephasing by considering three scenar- A. Steady state spectra ios: no dephasing, low dephasing (i.e. one order of magni- tude smaller than the cavity decay rate κ) and high dephas- ing (i.e. dephasing and cavity decay rates of the same order). As anticipated, our first set of numerical examples investi- This is particularly relevant for quantum-dot implementations gates the optical emission properties of our multilevel model of our models, especially at room temperature, in which case under incoherent pumping of the emitter, as relevant in many emitter dephasing rates can be comparable to the cavity decay quantum dot experiments also at ambient conditions. Specif- rate. Finally it is important note that we shall adopt a logarith- ically, we set Ω = 0 (no coherent driving) and Λ , 0. In this mic scale in our plots, since the examined quantities display setting we are interested in the steady state spectrum of the features of very different magnitudes. This implies that the cavity output field, i.e., 9

the bottom of Fig. 5 that multilevel and JC models are in- Z ∞ ! deed in agreement on the location of both single-excitation ˆ† ˆ −iωτ ω ' ω ± g ω ' S a(ω) = 2Re hEout(τ)Eout(0)i∞e dτ , ( √0 eff, gray dot-dashed lines) and bi-excitation [ 0 ω0 ± ( 2 ± 1)geff], gray dotted lines) resonances, while the Z ∞ ! † −iωτ bi-excitation peaks of the TC model are visibly shifted. Also ∝ 2Re haˆ (τ)ˆa(0)i∞e dτ , (32) 0 in this case the multilevel model presents an additional reso- nance (or perhaps a collection of closely spaced resonances) where Eq. (31) has been used, the two-point correlation func- ' † at ω ω0. Referring again to the notation of Section IV A, tion haˆ (τ)ˆa(0)i∞ is calculated via the quantum regression the- these can be ascribed to two processes: the photons emitted orem [31], and the expectation value is calculated on ρ∞, the in the |E1Di → |G, 0i transition, which is associated with a steady state of the master equation. The latter is found by small decay rate, and those emitted in the “second rung” pro- solving the linear system Lρ∞ = 0, Tr[ρ∞] = 1, where the cess |E2Di → |E1Di, which we recall is very similar to photon superoperator L is implicitly defined by rewriting the mas- emission by an empty cavity, and hence has a singnificantly ter equation (8) asρ ˙ = Lρ. We perform such calculations higher decay rate ' κ. We again notice that the ω ' ω0 reso- numerically with Python, by truncating the Hilbert space of nance can be weakened in the presence of dephasing, however the cavity to a sufficiently high dimension to obtain numerical it remains prominent even in the “high dephasing” scenario. convergence. In addition to the cavity spectrum, we shall also Our interpretation for this is the following: while dephasing be interested in the quantity of order κ can suppress the |E1Di → |G, 0i process, as dis- X cussed above, the |E i → |E i decay channel remains a | ih | 2D 1D pdark = Tr[ρ∞ Dk Dk ] probable process since it is also associated with a decay rate k of order κ. = 1 − Tr[ρ∞(|BihB| + |GihG|)], (33) It is worth pointing out that if one goes beyond our symmet- ric pump hypothesis (populating all the emitter excited states i.e. the steady-state probability that the emitter has settled in with equal probability - see Section III) it may of course be a dark state (or a mixture thereof). The resulting spectra, in possible to reduce the value of p and hence the prominence logarithmic scale, are plotted in Fig. 5 together with those dark of dark-state features in the cavity emission spectrum. of the reference TC and JC models (again, see Appendix A for details). For a fairer visual comparison between the To summarize, the results of this subsection tell us that the three models we have also rescaled the spectrum of JC and multilevel√ model is a viable system to test the characteristic ∝ ngeff splitting of energy levels that is still the holy grail TC models by a factor 1 − pdark. We indeed recall from Section IV A that the multilevel model cannot display JC of experimental JC research. In this light, the main strength physics once it gets “stuck” into a dark state. of the multilevel model is that the effective coupling geff may be significantly boosted by considering emitters with many In the single excitation regime, we find that all three models closely spaced sublevels. display the expected resonances at ω ' ω0 ± geff, but the mul- tilevel model presents additional sharp features at ω ' ω0. In the case of no dephasing, these resonances account for a sig- B. Photon autocorrelation function nificant portion of the total emitted energy, and correspond- ingly pdark is close to one. We argue that the ω ' ω0 features As a second step, we will look at the steady state behaviour can be ascribed to the slow decay of the long-lived quasi-dark of the zero-delay photon autocorrelation function, under weak states of the single-excitation sector: referring to our discus- coherent driving of the cavity. In all three models this quantity sion and notation in Section IV A, this involves transitions of is defined as follows: the form |E Di → |G, 0i (where |E Di can be any of the two 1 1 ˆ† ˆ† ˆ ˆ quasi-dark states occurring for N = 3). As dephasing is in- hEoutEoutEoutEouti g(2)(0) = lim creased, going left to right in Fig. 5, we find that the three t→∞ ˆ† ˆ 2 hEoutEouti models show increasing agreement, and at the same time p dark haˆ†aˆ†aˆaˆi is significantly reduced. Referring again to our analytical dis- = lim † , (34) cussion in Section IV A, we recall that dephasing is indeed t→∞ haˆ aˆi2 able to swap bright and dark states. This opens a new de- where again we made use of Eq. (31) in order to obtain an cay channel for the quasi-dark states |E i: as dephasing is 1D expression that only features system operators. In detail, in increased, we may reach a parameter regime where the quasi- this subsection we consider the master equation (8), setting dark states |E i are more likely to decay by a process of the 1D Λ = 0, Ω 0, and vary the driving field frequency ω . form |E i → |E i → |G, 0i, where the first step is due to de- , L 1D 1± While the resulting master equation is explicitly time depen- phasing, the second to photon emission. Hence, in such condi- dent, it is still possible to define a steady state by consid- tions JC-like features can become more prominent in the mea- ering an interaction picture with respect to the Hamiltonian sured spectrum. At high dephasing (γ = κ), the dark-state d H = ω aˆ†aˆ +ω Pn |e ihe |. This is a standard procedure in features are effectively invisible in the emission spectrum. L L L k=1 k k → (int) The bi-excitation scenario is where the multilevel model truly quantum optics and it amounts to replacing Vdrive Vdrive =  † (int) shines, compared to the reference TC model: we can see in Ω aˆ + aˆ and H → H = H − HL in the master equa- 10

Single-excitation regime - κ = g

Bi-excitation regime - κ = 0.05g

(2) FIG. 6. steady-state photon autocorrelation g (0) for N = 3 emitter sublevels and a sub-level energy spread fixed at ε = 0.05geff . We set the incoherent pump to zero and introduce a coherent cavity drive, corresponding to model parameters Λ = 0, Ω = 0.01geff . Dephasing increases from left to right as before. Solid blue lines = multi-level model; dashed red lines = reference JC model; dot-dashed green lines = reference TC model; dotted black lines = reference coupled oscillators model. For details on all the auxiliary models used here see Appendix A

tion (8). In turn, this procedure implicitly defines a time inde- even farther off from the JC predictions. As expected, the lat- pendent superoperator L(int) which does admit a steady state ter indeed displays features that are several orders of magni- (int) ρ∞ . Such a steady state is again found numerically with the tude smaller than those in all the other models, plus it satisfies same process described in the previous section, and it can be g(2)(0) > 1 at all times. As anticipated, however, the multi- easily shown that it provides the correct expectation values in- level model displays additional features in the region ω ' ω0: volved in Eq. (34). The obtained autocorrelation functions for at low or no dephasing a double-peak structure can be clearly multilevel, JC and TC models are displayed in Fig. 6 as func- seen, likely due to processes involving the quasi-dark states tions of the driving field detuning. We note that in this case no (we recall that for N = 3 there are two quasi-dark states for rescaling of the TC and JC plots is required, since the g2(0) is each value of the excitation number n ≥ 1). As dephasing is normalised by definition. increased the prominence of these peaks is reduced, presum- ably due to the introduction of additional decay channels for ' We observe that, excluding the region ω ω0, the multi- the quasi-dark states (see also previous Subsection). However, level model does a good job at matching the JC model pre- dephasing also results in the appearance of a central peak in dictions, capturing both the position and magnitude of res- the g(2)(0), which may be due to an increased likelihood that a onances and dips. These observations are true both in the (2) pair of driving photons excite a quasi-dark state of the system, single- and bi-excitation regimes. In contrast, the g (0) of and are later emitted at the frequency of an empty cavity by a the TC model displays features that are significantly less pro- process of the form |E2Di → |E1Di → |G, 0i. While a more nounced than their JC counterparts (recall that we are using rigorous analysis of these effects would be certainly worth- a logscale), as well as being shifted in frequency in the bi- while to confirm our intuitions, we feel that it goes beyond excitation regime. This is compatible with the intuition that the scopes of this work. the TC model is “more linear” than both the JC and multi- level ones. Finally, the coupled oscillator model is of course As anticipated, the g(2)(0) is one of the most commonly 11 adopted witnesses of nonlinearity in emitter-cavity systems emission spectra and photon autocorrelation functions. More- such as the ones we are investigating. The results of this sub- over, we have been able to characterise the main effects asso- section thus confirm that the multilevel model closely matches ciated with the additional levels of the emitter: the appearance the nonlinear signatures of JC physics, and in this respect it of very interesting sharp resonances in the middle of the emis- clearly outperforms the reference TC model with the same sion spectrum, associated with the slow decay of “quasi dark values of N and geff. states”. While it may be possible to observe such resonances in an experiment, our analysis shows that these are quickly washed out in the presence of dephasing. Observing dark- state signatures would thus require cryogenic temperatures. VI. CONCLUSIONS Our findings are particularly relevant in solid-state and plas- monic implementations of Cavity QED, where the emitters of We explored a novel route to realise the strong coupling choice are often multi-level quantum dots, and may pave the regime of light-matter interaction, exploiting a multi-level way towards the observation and exploitation of light-matter emitter coupled to a single-mode electromagnetic field. When strong coupling at room temperature. the N excited sublevels of the emitter are nearly degenerate, we showed both analytically and numerically that our pro- posed system can closely approximate many aspects of JC VII. ACKNOWLEDGMENTS physics. The√ associated light-matter coupling constant scales as geff ∝ N and, crucially, the quantum nonlinearities of our We thank M. G. Genoni for the useful discussion on mod- model are not suppressed with increasing N. Indeed, we are elling dephasing. Support by the Science Foundation Ireland able to√ observe the characteristic energy level splitting of or- (SFI) under grant 18/RP/6236 is gratefully acknowledged. TT der ∼ n, where n is the number of combined (emitter + field) acknowledges support from the University of Nottingham via excitations, as well as the associated nonlinear signatures in a Nottingham Research Fellowship.

[1] B. W. Shore and P. L. Knight, The jaynes-cummings model, [11] N. Quesada, Strong coupling of two quantum emitters to a sin- Journal of Modern Optics 40, 1195 (1993). gle light mode: The dissipative Tavis-Cummings ladder, Physi- [2] S. Haroche and J.-M. Raimond, Exploring the quantum: atoms, cal Review A 86, 013836 (2012), publisher: American Physical cavities, and photons (Oxford university press, 2006). Society. [3] A. Reiserer, N. Kalb, G. Rempe, and S. Ritter, A quantum gate [12] R. H. Dicke, Coherence in spontaneous radiation processes, between a flying optical photon and a single trapped atom, Na- Phys. Rev. 93, 99 (1954). ture 508, 237 (2014), number: 7495 Publisher: Nature Publish- [13] M. Wersall,¨ J. Cuadra, T. J. Antosiewicz, S. Balci, and T. She- ing Group. gai, Observation of Mode Splitting in Photoluminescence of In- [4] P. Biagioni, J.-S. Huang, and B. Hecht, Nanoantennas for vis- dividual Plasmonic Nanoparticles Strongly Coupled to Molecu- ible and infrared radiation, Reports on Progress in Physics 75, lar Excitons, Nano Letters 17, 551 (2017), publisher: American 024402 (2012). Chemical Society. [5] H. Walther, B. T. Varcoe, B.-G. Englert, and T. Becker, Cavity [14] D. G. Baranov, B. Munkhbat, E. Zhukova, A. Bisht, A. Canales, quantum electrodynamics, Reports on Progress in Physics 69, B. Rousseaux, G. Johansson, T. J. Antosiewicz, and T. Shegai, 1325 (2006). Ultrastrong coupling between nanoparticle and cav- [6] P. Torm¨ a¨ and W. L. Barnes, Strong coupling between surface ity photons at ambient conditions, Nature Communications 11, polaritons and emitters: a review, Reports on Progress 2715 (2020). in Physics 78, 013901 (2015). [15] M. Geisler, X. Cui, J. Wang, T. Rindzevicius, L. Gammel- [7] D. S. Dovzhenko, S. V. Ryabchuk, Y. P. Rakovich, and I. R. gaard, B. S. Jessen, P. A. D. Gonc¸alves, F. Todisco, P. Bø ggild, Nabiev, Light–matter interaction in the strong coupling regime: A. Boisen, M. Wubs, N. A. Mortensen, S. Xiao, and N. Stenger, configurations, conditions, and applications, Nanoscale 10, Single-Crystalline Gold Nanodisks on WS2 Mono- and Multi- 3589 (2018). layers for Strong Coupling at Room Temperature, ACS Photon- [8] X. Yu, Y. Yuan, J. Xu, K.-T. Yong, J. Qu, and J. Song, Strong ics 6, 994 (2019), publisher: American Chemical Society. Coupling in Microcavity Structures: Principle, Design, and [16] A. J. Moilanen, T. K. Hakala, and P. Torm¨ a,¨ Active Control of Practical Application, Laser & Photonics Reviews 13, 1800219 Surface Plasmon–Emitter Strong Coupling, ACS Photonics 5, (2019), zSCC: 0000013. 54 (2018). [9] G. Zengin, G. Johansson, P. Johansson, T. J. Antosiewicz, [17] K. Santhosh, O. Bitton, L. Chuntonov, and G. Haran, Vac- M. Kall,¨ and T. Shegai, Approaching the strong coupling limit uum Rabi splitting in a plasmonic cavity at the single quan- in single plasmonic nanorods interacting with J-aggregates, Sci- tum emitter limit, Nature Communications 7, ncomms11823 entific Reports 3, 3074 (2013), number: 1 Publisher: Nature (2016), number: 1 Publisher: Nature Publishing Group. Publishing Group. [18] R. Chikkaraddy, B. de Nijs, F. Benz, S. J. Barrow, O. A. Scher- [10] M. Tavis and F. W. Cummings, Exact solution for an n- man, E. Rosta, A. Demetriadou, P. Fox, O. Hess, and J. J. molecule—radiation-field hamiltonian, Physical Review 170, Baumberg, Single-molecule strong coupling at room temper- 379 (1968). ature in plasmonic nanocavities, Nature 535, 127 (2016), num- ber: 7610 Publisher: Nature Publishing Group. 12

[19] H. Groß, J. M. Hamm, T. Tufarelli, O. Hess, and B. Hecht, tween Single-Nanoparticle Plasmons and Molecular Excitons Near-field strong coupling of single quantum dots, Science Ad- at Ambient Conditions, Physical Review Letters 114, 157401 vances 4, eaar4906 (2018). (2015), publisher: American Physical Society. [20] H. Leng, B. Szychowski, M.-C. Daniel, and M. Pelton, Strong [36] Y. Sugawara, T. A. Kelf, J. J. Baumberg, M. E. Abdelsalam, coupling and induced transparency at room temperature with and P. N. Bartlett, Strong Coupling between Localized Plas- single quantum dots and gap plasmons, Nature Communica- mons and Organic Excitons in Metal Nanovoids, Physical Re- tions 9, 10.1038/s41467-018-06450-4 (2018). view Letters 97, 266808 (2006), publisher: American Physical [21] K.-D. Park, M. A. May, H. Leng, J. Wang, J. A. Kropp, Society. T. Gougousi, M. Pelton, and M. B. Raschke, Tip-enhanced [37] A. E. Schlather, N. Large, A. S. Urban, P. Nordlander, and strong coupling spectroscopy, imaging, and control of a single N. J. Halas, Near-Field Mediated Plexcitonic Coupling and Gi- quantum emitter, Science Advances 5, eaav5931 (2019). ant Rabi Splitting in Individual Metallic Dimers, Nano Letters [22] J. Kasprzak, S. Reitzenstein, E. A. Muljarov, C. Kistner, 13, 3281 (2013), publisher: American Chemical Society. C. Schneider, M. Strauss, S. Hofling,¨ A. Forchel, and W. Lang- [38] M. Brune, S. Haroche, J. M. Raimond, L. Davidovich, and bein, Up on the Jaynes-Cummings ladder of a quantum- N. Zagury, Manipulation of photons in a cavity by dispersive dot/microcavity system, Nature Materials 9, 304 (2010), num- atom-field coupling: Quantum-nondemolition measurements ber: 4 Publisher: Nature Publishing Group. and generation of “schrodinger¨ cat” states, Physical Review A [23] J. M. Fink, M. Goppl,¨ M. Baur, R. Bianchetti, P. J. Leek, 45, 5193 (1992), publisher: American Physical Society. A. Blais, and A. Wallraff, Climbing the Jaynes–Cummings lad- [39] C. K. Law and J. H. Eberly, Arbitrary Control of a Quan- der and observing its nonlinearity in a cavity QED system, Na- tum Electromagnetic Field, Physical Review Letters 76, 1055 ture 454, 315 (2008), number: 7202 Publisher: Nature Publish- (1996), publisher: American Physical Society. ing Group. [40] F. P. Laussy, E. del Valle, and C. Tejedor, Luminescence spectra [24] J. P. Reithmaier, G. Se¸k, A. Lo¨ffler, C. Hofmann, S. Kuhn, of quantum dots in microcavities. I. Bosons, Physical Review B S. Reitzenstein, L. V. Keldysh, V. D. Kulakovskii, T. L. Rei- 79, 235325 (2009), publisher: American Physical Society. necke, and A. Forchel, Strong coupling in a single quantum dot- [41] D. J. Wineland, Nobel lecture: Superposition, entanglement, semiconductor microcavity system, Nature 432, 197 (2004), and raising schrodinger’s¨ cat, Rev. Mod. Phys. 85, 1103 (2013). number: 7014 Publisher: Nature Publishing Group. [42] D. G. Baranov, M. Wersall,¨ J. Cuadra, T. J. Antosiewicz, [25] K. H. Madsen, T. B. Lehmann, and P. Lodahl, Role of multi- and T. Shegai, Novel Nanostructures and Materials for Strong level states on quantum-dot emission in photonic-crystal cavi- Light–Matter Interactions, ACS Photonics 5, 24 (2017). ties, Physical Review B 94, 235301 (2016), publisher: Ameri- [43] Y. Huang, F. Wu, and L. Yu, Rabi oscillation study of strong can Physical Society. coupling in a plasmonic nanocavity, New Journal of Physics [26] H. Liu and P. Guyot-Sionnest, Photoluminescence Lifetime of 22, 10.1088/1367-2630/ab9222 (2020). Lead Selenide Colloidal Quantum Dots, The Journal of Phys- ical Chemistry C 114, 14860 (2010), publisher: American Chemical Society. [27] Z. Hens and I. Moreels, Light absorption by colloidal semicon- Appendix A: Auxiliary models ductor quantum dots, Journal of Materials Chemistry 22, 10406 (2012), publisher: The Royal Society of Chemistry. Here we briefly list the Hamiltonians and Lindblad opera- [28] C. Tserkezis, A. I. Fernandez-Dom´ ´ınguez, P. A. D. Gonc¸alves, tors for the auxiliary models used throughout the paper. In all F. Todisco, J. D. Cox, K. Busch, N. Stenger, S. I. Bozhevol- cases the field mode is described by the same annihilation op- nyi, N. A. Mortensen, and C. Wolff, On the applicability of eratora ˆ (and corresponding Hilbert space) as in the main text. quantum-optical concepts in strong-coupling nanophotonics, In turn, photon loss and coherent cavity driving are described Reports on Progress in Physics 83, 082401 (2020). [29] N. Kongsuwan, A. Demetriadou, M. Horton, R. Chikkaraddy, by the same operators shown in the main text (Section III). J. J. Baumberg, and O. Hess, Plasmonic nanocavity modes: What changes between the different models is the structure of From near-field to far-field radiation, ACS Photonics 7, 463 the emitter, and the associated Lindblad operators. (2020), https://doi.org/10.1021/acsphotonics.9b01445. [30] S. Franke, S. Hughes, M. K. Dezfouli, P. T. Kristensen, K. Busch, A. Knorr, and M. Richter, Quantization of quasi- 1. Jaynes-Cummings Model normal modes for open cavities and plasmonic cavity quantum electrodynamics, Phys. Rev. Lett. 122, 213901 (2019). [31] H. P. Breuer and F. Petruccione, The theory of open quan- The Hamiltonian of the JC model adopted in the main text tum systems (Oxford University Press, Great Clarendon Street, reads 2002). [32] D. F. Walls and G. J. Milburn, Quantum optics (Springer Sci-  †  † HJC = ω0 aˆ aˆ + |eihe| + geff(ˆa |gihe| + aˆ|eihg|), (A1) ence & Business Media, 2007). [33] F. P. Laussy, E. del Valle, and C. Tejedor, Luminescence spectra | i of quantum dots in microcavities. I. Bosons, Physical Review B where the emitter is a two-level system with excited state e 79, 235325 (2009), publisher: American Physical Society. and ground state |gi. The Lindblad operators for emitter de- [34] E. del Valle, F. P. Laussy, and C. Tejedor, Luminescence spec- cay, dephasing and incoherent pumping are respectively tra of quantum dots in microcavities. II. Fermions, Physical Re- √ view B 79, 235326 (2009), publisher: American Physical Soci- L = γ|gihe|, (A2) ety. √ L = γd|eihe|, (A3) [35] G. Zengin, M. Wersall,¨ S. Nilsson, T. J. Antosiewicz, M. Kall,¨ √ and T. Shegai, Realizing Strong Light-Matter Interactions be- L = Λ|eihg| (A4) 13

2. Tavis-Cummings model with N atoms quency) reads

The TC model used in the main text, with N atoms and effective coupling geff, reads  † †  † † Hosc = ω0 aˆ aˆ + bˆ bˆ + geff(ˆa bˆ + bˆ aˆ), (A9) † Jˆz geff + † − HTC = ω0aˆ aˆ + ω0 + √ (Jˆ aˆ + aˆ Jˆ ), (A5) 2 N where Jˆ , Jˆ , Jˆ are spin-N/2 operators, and Jˆ± = Jˆ ± iJˆ are x y z x y a, bˆ the spin-ladder operators. The Lindblad operators for emitter where ˆ are annihilation operators of distinct Bosonic decay, dephasing and incoherent pumping are respectively modes. The Lindblad operator for emitter decay, dephasing and incoherent pumping are respectively √ L = γJˆ−, (A6) √ L = γd Jˆz/2, (A7) √ √ L = ΛJˆ+ (A8) L = γbˆ, (A10) √ † L = γdbˆ bˆ, (A11) √ † 3. Coupled oscillators model L = Λbˆ . (A12)

In the rotating wave approximation, the Hamiltonian of two coupled harmonic oscillators (both having the same fre-