This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Theory of polariton-electron interactions in semiconductor microcavities

Guangyao Li School of Physics and Astronomy, Monash University, Victoria 3800, Australia ARC Centre of Excellence in Future Low-Energy Electronics Technologies, Monash University, Victoria 3800, Australia    Olivier Bleu School of Physics and Astronomy, Monash University, Victoria 3800, Australia ARC Centre of Excellence in Future Low-Energy Electronics Technologies, Monash University, Victoria 3800, Australia    Jesper Levinsen School of Physics and Astronomy, Monash University, Victoria 3800, Australia ARC Centre of Excellence in Future Low-Energy Electronics Technologies, Monash University, Victoria 3800, Australia    Meera M. Parish School of Physics and Astronomy, Monash University, Victoria 3800, Australia ARC Centre of Excellence in Future Low-Energy Electronics Technologies, Monash University, Victoria 3800, Australia
Abstract

We develop a microscopic description of an electron-doped two-dimensional semiconductor embedded in a microcavity. Specifically, we investigate the interactions between exciton-polaritons and electrons for the case where the interactions between charges are strongly screened and the system is spin polarized. As a starting point, we obtain an analytic expression for the exciton-polariton wave function, and we relate the microscopic parameters of the light-matter system to experimentally measurable quantities, such as the Rabi coupling and the cavity photon frequency. We then derive the polariton-electron interaction within the standard Born approximation and compare it with the exact polariton-electron scattering TT matrix that we obtain from a diagrammatic approach that has proven highly successful in the context of nuclear physics and ultracold atomic gases. In particular, we show that the Born approximation provides an upper bound on the polariton-electron coupling strength at vanishing momentum. Using our exact microscopic calculation, we demonstrate that polariton-electron scattering can be strongly enhanced compared to the exciton-electron case, which is the opposite of that expected from the Born approximation. We furthermore expose a resonance-like peak at scattering momenta near the polariton inflection point, whose size is set by the strength of the light-matter coupling. Our results arise from the non-Galilean nature of the polariton system and should thus be applicable to a range of semiconductor microcavities such as GaAs quantum wells and atomically thin materials.

I Introduction

By embedding a two-dimensional (2D) semiconductor in a microcavity, one can achieve a strong coupling between photons and bound electron-hole pairs (i.e., excitons). This in turn gives rise to exciton-polaritons, quasiparticles which are superpositions of both light and matter [1, 2, 3]. A key advantage of polaritons over ordinary photons is that they can pairwise interact with other particles via their excitonic component, which is important for a range of applications including polariton superfluidity [4, 5], ultra-fast polariton spin switching [6], and the generation of photon correlations [7, 8]. Moreover, there is the prospect of enhancing the interactions in the polariton system by coupling to few-body bound states such as biexcitons [9] or trions [10].

Interaction processes involving polaritons are conventionally described using the Born approximation [11, 12, 13, 14, 15]. This can essentially be viewed as a two step process. First, the matter component interactions are calculated within the Born approximation, and then the result is weighted by the matter fraction of the involved particles since only this component interacts. There have been various attempts [11, 16, 17, 18, 19] to obtain perturbative corrections beyond the standard Born approximation to include higher-order effects due to the light-matter coupling. However, until now there have been no exact microscopic calculations of few-body processes that involve the constituent electron-hole-photon components of the polaritons, which is mainly due to the complexity associated with treating both the electronic and the light-matter interactions. It is therefore an open and important question whether the assumptions in the Born approximation are valid even at a qualitative level.

In this paper, which accompanies Ref. [20], we use a microscopic model that explicitly includes electrons, holes, and photons to investigate this question. The key simplification is to consider strongly screened electronic interactions. Such a simplified model has previously been used to theoretically investigate the crossover from a Bose-Einstein condensate (BEC) of polaritons to the Bardeen-Cooper-Schrieffer regime of electron-hole superfluidity [21, 22, 23]. Here we show that this approximation allows us to formulate a diagrammatic approach where we can analytically obtain the polariton wave function and propagator, as well as the associated spectrum and photon and exciton fractions. Our resulting microscopic description of exciton-polaritons complements our previous results for Coulomb electronic interactions [24], and it can act as a starting point for further few- and many-body calculations within the model of strongly screened electronic interactions.

We then apply our diagrammatic approach to perform the first exact calculation of spin-polarized polariton-electron scattering within a microscopic model. Importantly, we find that the polariton-electron scattering is strongly enhanced up to the polariton dispersion inflection point compared with exciton-electron scattering, in stark contrast to the assumptions of the Born approximation. We argue that this unexpected feature is because the light-matter coupling shifts the collision energy, which has a stronger effect than the reduction in the interaction strength due the reduced exciton fraction of the polariton. Given that this argument relies only on standard 2D scattering theory [25] which is valid also for unscreened electronic interactions, we expect the enhanced polariton-electron interactions to be a generic feature of 2D semiconductors embedded in microcavities, and we expect our results to be quite accurate even at a quantitative level in the case of more realistic electronic interactions.

Of particular interest is the scattering at low momentum, due to the potential implications for polariton BECs [26] in charge-doped semiconductors. It is well known [27] that the scattering of 2D quantum particles with short range interactions approaches zero logarithmically with momentum. In the case of exciton-electron scattering, the relevant momentum scale is the inverse exciton Bohr radius. Remarkably, in the case of polariton-electron scattering, we find that the resulting momentum scale is suppressed exponentially by the large electron-photon mass ratio, and consequently the polariton-electron scattering only approaches zero in systems that greatly exceed the size of the universe. This is a striking consequence of the polariton being formed from a superposition of particles with extremely different masses, and it allows us to define a finite coupling constant at vanishing momentum.

We furthermore find that the Born approximation severely overestimates the polariton-electron interaction constant for typical experimental parameters. We argue that this is because the Born approximation represents an upper bound on the interaction constant. While this implies that this simple approximation cannot in general be trusted, we show that it can be replaced by a similarly simple expression that calculates the interaction constant instead from the low-energy exciton-electron scattering amplitude, which represents an excellent approximation in the case of a transition metal dichalcogenide (TMD) monolayer or a single GaAs quantum well embedded in a microcavity.

The paper is organized as follows. In Sec. II, we present the microscopic description of a single polariton, taking care to appropriately renormalize the model. We obtain analytic expressions for the polariton wave function, spectrum, and excitonic and photonic fractions, and demonstrate that our results can be obtained within both an operator and a diagrammatic approach. In Sec. III we outline the Born approximation of polariton-electron scattering, which provides an upper bound for our exact calculation presented in Sec. IV. We discuss how our results depend on system parameters relevant to both semiconductor quantum wells and atomically thin semiconductors. In Sec. V we conclude. Technical details are given in the appendices.

II Theoretical description of exciton-polaritons

We consider a system consisting of spin-polarized electrons, holes, and photons in a 2D semiconductor, such as a quantum well or an atomically thin transition metal dichalcogenide material. The Hamiltonian reads

H=\displaystyle H= 𝐤(ϵ𝐤ee𝐤e𝐤+ϵ𝐤hh𝐤h𝐤)+𝐤(ω+ϵ𝐤c)c𝐤c𝐤\displaystyle\sum_{\mathbf{k}}\left(\epsilon^{e}_{\mathbf{k}}\,e^{\dagger}_{\mathbf{k}}e_{\mathbf{k}}+\epsilon^{h}_{\mathbf{k}}\,h^{\dagger}_{\mathbf{k}}h_{\mathbf{k}}\right)+\sum_{\mathbf{k}}(\omega+\epsilon^{c}_{\mathbf{k}})c^{\dagger}_{\mathbf{k}}c_{\mathbf{k}}
V0𝐤𝐤𝐪e𝐤h𝐪𝐤h𝐪𝐤e𝐤\displaystyle-V_{0}\sum_{\mathbf{k}\,\mathbf{k}^{\prime}\,\mathbf{q}}e^{\dagger}_{\mathbf{k}}h^{\dagger}_{\mathbf{q}-\mathbf{k}}h_{\mathbf{q}-\mathbf{k}^{\prime}}e_{\mathbf{k}^{\prime}}
+g𝐤𝐪(e𝐤h𝐪𝐤c𝐪+c𝐪h𝐪𝐤e𝐤).\displaystyle+g\sum_{\mathbf{k}\,\mathbf{q}}\left(e^{\dagger}_{\mathbf{k}}h^{\dagger}_{\mathbf{q}-\mathbf{k}}c_{\mathbf{q}}+c^{\dagger}_{\mathbf{q}}h_{\mathbf{q}-\mathbf{k}}e_{\mathbf{k}}\right). (1)

The first line describes the kinetic energies of the particles, where e𝐤e_{\mathbf{k}}, h𝐤h_{\mathbf{k}}, and c𝐤c_{\mathbf{k}} are electron, hole, and photon annihilation operators, respectively, with momentum 𝐤\mathbf{k} and kinetic energy ϵ𝐤e,h,c=|𝐤|2/(2me,h,c)k2/(2me,h,c)\epsilon^{e,h,c}_{\mathbf{k}}=|\mathbf{k}|^{2}/(2m_{e,h,c})\equiv k^{2}/(2m_{e,h,c}), and corresponding masses mem_{e}, mhm_{h}, and mcm_{c}. ω\omega is the cavity photon frequency measured from the bandgap. In the second line, we have the electron-hole interactions, which we take to be strongly screened contact interactions of strength V0>0V_{0}>0. Note that there are no electron-electron or hole-hole interactions in this screened case since the interactions between identical fermions formally vanish due to Pauli exclusion. The last term corresponds to the creation (annihilation) of an electron-hole pair through the absorption (emission) of a photon within the rotating wave approximation, and gg denotes the strength of the (unrenormalized) light-matter coupling. This term explicitly breaks Galilean invariance by coupling the light photon with much heavier matter particles, which has important consequences for the few-body properties of the system. Note that here and in the following we adopt units where =1\hbar=1 and the system area A=1A=1.

The Hamiltonian contains the bare parameters V0V_{0} and gg that describe the strength of matter-matter and light-matter contact interactions, respectively. We take these to be constant up to an ultraviolet momentum cutoff which is set by the detailed band structure of the 2D semiconductor. For our discussion of polariton physics, the precise value of the momentum cutoff will be irrelevant, since we aim to develop a low-energy theory that is independent of the short-distance physics [24]. Therefore, both V0V_{0} and gg need to be renormalized such that all observable quantities are independent of the momentum cutoff. For simplicity, we take the cutoffs related to V0V_{0} and gg to be the same, denoted Λ\Lambda, since they both drop out after the renormalization.

II.1 Exciton problem

Let us first discuss how the formation of an exciton is described within the contact interaction model (II). To this end, we consider the most general state of an electron-hole pair with zero center-of-mass momentum:

|Φ=𝐤ϕ𝐤e𝐤h𝐤|0,\ket{\Phi}=\sum_{\mathbf{k}}\phi_{\mathbf{k}}e^{\dagger}_{\mathbf{k}}h^{\dagger}_{-\mathbf{k}}\ket{0}, (2)

with the normalization condition Φ|Φ=𝐤|ϕ𝐤|2=1.\bra{\Phi}\ket{\Phi}=\sum_{\mathbf{k}}\absolutevalue{\phi_{\mathbf{k}}}^{2}=1. Here, |0\ket{0} is the electron-hole vacuum. In the absence of coupling to light, i.e., at g=0g=0, the wave function ϕ𝐤\phi_{\mathbf{k}} satisfies the Schrödinger equation

(Eϵ¯𝐤)ϕ𝐤\displaystyle(E-\bar{\epsilon}_{\mathbf{k}})\phi_{\mathbf{k}} =V0𝐤ϕ𝐤,\displaystyle=-V_{0}\sum_{\mathbf{k}^{\prime}}\phi_{\mathbf{k}^{\prime}}, (3)

where we define the electron-hole kinetic energy ϵ¯𝐤ϵ𝐤e+ϵ𝐤h=k2/2mr\bar{\epsilon}_{\mathbf{k}}\equiv\epsilon^{e}_{\mathbf{k}}+\epsilon^{h}_{\mathbf{k}}=k^{2}/2m_{r} with mr=(1/me+1/mh)1m_{r}=(1/m_{e}+1/m_{h})^{-1} the electron-hole reduced mass.

Equation (3) admits a single bound state with binding energy εB\varepsilon_{B}, corresponding to the 1s1s exciton state (see, e.g., Ref. [28]), which has the associated effective Bohr radius aX1/2mrεBa_{X}\equiv 1/\sqrt{2m_{r}\varepsilon_{B}}. The corresponding wave function ϕX𝐤\phi_{X\mathbf{k}} satisfies the equation

(εB+ϵ¯𝐤)ϕX𝐤\displaystyle(\varepsilon_{B}+\bar{\epsilon}_{\mathbf{k}})\phi_{X\mathbf{k}} =V0𝐤ϕX𝐤.\displaystyle=V_{0}\sum_{\mathbf{k}^{\prime}}\phi_{X\mathbf{k}^{\prime}}. (4)

Importantly, the right hand side of Eq. (4) does not depend on momentum, and therefore we immediately find that the 1s1s exciton has the wave function

ϕX𝐤=ZXεB+ϵ¯𝐤,\displaystyle\phi_{X\mathbf{k}}=\frac{\sqrt{Z_{X}}}{\varepsilon_{B}+\bar{\epsilon}_{\mathbf{k}}}, (5)

where ZXZ_{X} comes from the normalization condition

ZX=[𝐤1(εB+ϵ¯𝐤)2]1=2πεBmr.\displaystyle Z_{X}=\left[\sum_{\mathbf{k}}\frac{1}{(\varepsilon_{B}+\bar{\epsilon}_{\mathbf{k}})^{2}}\right]^{-1}=\frac{2\pi\varepsilon_{B}}{m_{r}}. (6)

Equation (4) also allows us to relate the bare coupling constant, V0V_{0}, to εB\varepsilon_{B}: Acting with the operator 𝐤1εB+ϵ¯𝐤()\sum_{\mathbf{k}}\frac{1}{\varepsilon_{B}+\bar{\epsilon}_{\mathbf{k}}}(\cdot) on Eq. (4), we find

1V0=𝐤Λ1εB+ϵ¯𝐤.\displaystyle\frac{1}{V_{0}}=\sum_{\mathbf{k}}^{\Lambda}\frac{1}{\varepsilon_{B}+\bar{\epsilon}_{\mathbf{k}}}. (7)

Here, the sum on 𝐤\mathbf{k} is logarithmically divergent, and we have therefore explicitly introduced the ultraviolet momentum cutoff Λ\Lambda. Once the coupling constant and Λ\Lambda are related to the exciton binding energy via Eq. (7), all dependence on these bare parameters is eliminated from the problem [28].

It is instructive to compare the operator formalism described above for the exciton problem with the electron-hole TT matrix [28]:

𝒯0(E)=2π/mrln[E/εB+i0]+iπ.\mathcal{T}_{0}(E)=\frac{2\pi/m_{r}}{-\ln\left[E/\varepsilon_{B}+i0\right]+i\pi}. (8)

This is discussed further below in Sec. II.3. In this formalism, the bound state emerges as a pole at the negative energy E=εBE=-\varepsilon_{B}, where the presence of the infinitesimal positive imaginary part +i0+i0 shifts the pole slightly into the lower half of the complex plane. In the vicinity of the pole, we expand the TT matrix to find 𝒯0(E)ZX/(E+εB+i0)\mathcal{T}_{0}(E)\simeq Z_{X}/(E+\varepsilon_{B}+i0). Thus, the normalization ZXZ_{X} naturally emerges in both approaches.

II.2 Operator approach to exciton-polaritons

Let us now discuss how to obtain the polariton spectrum within the model (II), and how to relate this to experimentally measurable quantities. As we will show, the Hamiltonian (II) is analytically solvable in the case of the single-polariton problem. In the following, we provide a derivation of the exact wave function for a polariton at normal incidence (i.e., zero momentum), but our results can be straightforwardly extended to describe polaritons at finite momentum — see Appendix A.

To proceed, we write the most general electron-hole-photon wave function as:

|Ψ=𝐤ψ𝐤e𝐤h𝐤|0+γc𝟎|0,\ket{\Psi}=\sum_{\mathbf{k}}\psi_{\mathbf{k}}e^{\dagger}_{\mathbf{k}}h^{\dagger}_{-\mathbf{k}}\ket{0}+\gamma c^{\dagger}_{\mathbf{0}}\ket{0}, (9)

with the normalization condition:

Ψ|Ψ=𝐤|ψ𝐤|2+|γ|2=1.\bra{\Psi}\ket{\Psi}=\sum_{\mathbf{k}}\absolutevalue{\psi_{\mathbf{k}}}^{2}+\absolutevalue{\gamma}^{2}=1. (10)

The Schrödinger equation can be obtained by projecting (EH^)|Ψ=0(E-\hat{H})\ket{\Psi}=0 onto the electron-hole and photon parts of Eq. (9), which gives:

(Eϵ¯𝐤)ψ𝐤\displaystyle(E-\bar{\epsilon}_{\mathbf{k}})\psi_{\mathbf{k}} =V0𝐤ψ𝐤+gγ,\displaystyle=-V_{0}\sum_{\mathbf{k}^{\prime}}\psi_{\mathbf{k}^{\prime}}+g\gamma, (11a)
(Eω)γ\displaystyle(E-\omega)\gamma =g𝐤ψ𝐤.\displaystyle=g\sum_{\mathbf{k}}\psi_{\mathbf{k}}. (11b)

In the following, we assume that the exciton binding energy is larger than other relevant energy scales, such as the photon-exciton detuning and the light-matter Rabi coupling. This is a good approximation in the TMDs and in a single GaAs quantum well. As we show, this condition need not be strictly satisfied, and therefore our results also apply to structures containing multiple GaAs quantum wells.

Refer to caption
Figure 1: Energies (top row) and photon fractions (bottom row) of the lower and upper polaritons obtained within our electron-hole-photon model in Eq. (12) (blue solid lines) and within the two coupled oscillator model Eq. (21) (purple dashed lines). The shaded regions correspond to the electron-hole continuum. Parameters: (a) and (d) Ω/εB=0.1\Omega/\varepsilon_{B}=0.1; (b) and (e) Ω/εB=0.5\Omega/\varepsilon_{B}=0.5; (c) and (f) Ω/εB=1\Omega/\varepsilon_{B}=1.

Before delving into the technical details, we will briefly summarize our main conclusions. We emphasize that the results presented here act as a starting point for both few-body physics, as discussed in this paper, and many-body physics within the model (II). As we show in the following, the exact polariton energies E±E_{\pm} satisfy the transcendental equation

(ωE±)ln(E±εB)=Ω2εB,(\omega-E_{\pm})\ln\left(\frac{-E_{\pm}}{\varepsilon_{B}}\right)=\frac{\Omega^{2}}{\varepsilon_{B}}, (12)

where

Ωg𝐤ΛϕX𝐤=gV0ZX\displaystyle\Omega\equiv g\sum_{\mathbf{k}}^{\Lambda}\phi_{X\mathbf{k}}=\frac{g}{V_{0}}\sqrt{Z_{X}} (13)

is the effective exciton-photon Rabi coupling. The subscripts - and ++ indicate the lower and upper polaritons, respectively, and we assume that E±<0E_{\pm}<0 111If ΩεB\Omega\gtrsim\varepsilon_{B} and/or δεB\delta\gtrsim\varepsilon_{B}, the upper polariton enters the continuum of unbound electron-hole states. In that case, one must analytically continue the energy slightly into the complex plane, EE+i0E\to E+i0. Like in the exciton problem, the right hand side of Eq. (11a) is independent of momentum, and hence the electron-hole wave function has the same functional form as the exciton wave function ϕX𝐤\phi_{X\mathbf{k}} in Eq. (5222The variational approach developed by Khurgin [76] approximated the functional form of the electron-hole wave function as unchanged in the presence of light-matter coupling. While this is an approximation in the Coulomb case studied in that work [24], we see that it is exact for strongly screened electron-hole interactions.:

ψ±,𝐤=Z±1|γ±|2E±+ϵ¯𝐤.\displaystyle\psi_{\pm,\mathbf{k}}=\frac{\sqrt{Z_{\pm}}\sqrt{1-|\gamma_{\pm}|^{2}}}{-E_{\pm}+\bar{\epsilon}_{\mathbf{k}}}. (14)

Here the numerator follows from the normalization condition in Eq. (10), and

Z±=[𝐤1(E±ϵ¯𝐤)2]1=2π|E±|mr\displaystyle Z_{\pm}=\left[\sum_{\mathbf{k}}\frac{1}{(E_{\pm}-\bar{\epsilon}_{\mathbf{k}})^{2}}\right]^{-1}=\frac{2\pi|E_{\pm}|}{m_{r}} (15)

is the generalization of the exciton normalization ZXZ_{X} in Eq. (6). The corresponding photon and exciton Hopfield coefficients C±C_{\pm} and X±X_{\pm} take the form

|C±|2\displaystyle|C_{\pm}|^{2} |γ±|2=11+εB|E±|(E±ω)2Ω2,\displaystyle\equiv|\gamma_{\pm}|^{2}=\frac{1}{1+\frac{\varepsilon_{B}}{|E_{\pm}|}\frac{(E_{\pm}-\omega)^{2}}{\Omega^{2}}}, (16a)
|X±|2\displaystyle|X_{\pm}|^{2} 1|γ±|2=11+|E±|εBΩ2(E±ω)2.\displaystyle\equiv 1-|\gamma_{\pm}|^{2}=\frac{1}{1+\frac{|E_{\pm}|}{\varepsilon_{B}}\frac{\Omega^{2}}{(E_{\pm}-\omega)^{2}}}. (16b)

Finally, we note that the photon frequency relative to the exciton energy is shifted by a finite amount in the presence of the active medium, such that the effective photon-exciton detuning δ\delta is

δ=ω(εB)Ω22εB.\displaystyle\delta=\omega-(-\varepsilon_{B})-\frac{\Omega^{2}}{2\varepsilon_{B}}. (17)

Figure 1 shows the polariton energies and the corresponding photon and exciton fractions according to Eqs. (12) and (16). We find that they agree extremely well with the results of treating the excitons and photons as two coupled oscillators [31], even when the Rabi coupling is comparable to the exciton binding energy as in structures containing multiple GaAs quantum wells. In particular, our scheme provides an improved agreement between the two models than that obtained in Ref. [23], which considered a similar electron-hole-photon model. Moreover, our scheme is technically simpler to implement since Eqs. (12)-(17) are fully analytic and do not require the introduction of an additional infrared cutoff. We now proceed to derive these results.

II.2.1 Renormalization procedure

The central idea of the renormalization procedure is to relate the parameters in Eq. (11) to the experimental observables in the coupled-oscillator model of excitons and photons. At a technical level, we note that our renormalization of the electron-hole-photon model is conceptually similar to that of atoms interacting via both an open and a closed channel in three dimensions [32]. First, let us consider the case of small exciton-photon Rabi coupling such that the polariton energy is close to the exciton energy, i.e., E=εB+ΔEE=-\varepsilon_{B}+\Delta E with |ΔE|εB\absolutevalue{\Delta E}\ll\varepsilon_{B} the energy correction. In the limit EεBE\to-\varepsilon_{B}, the electron-hole part of the polariton wave function in Eq. (14) can then be approximated as proportional to the exciton wave function in Eq. (5), ψ𝐤βϕX𝐤\psi_{\mathbf{k}}\simeq\beta\phi_{X\mathbf{k}}, where β\beta is a complex number. Within this approximation, we use the exciton Schrödinger equation in Eq. (4) to find that Eq. (11) takes the form:

(E+εB)β\displaystyle(E+\varepsilon_{B})\beta =γg𝐤ϕX𝐤,\displaystyle=\gamma\,g\sum_{\mathbf{k}}\phi_{X\mathbf{k}}, (18a)
(Eω)γ\displaystyle(E-\omega)\gamma =βg𝐤ϕX𝐤.\displaystyle=\beta\,g\sum_{\mathbf{k}}\phi_{X\mathbf{k}}. (18b)

Written in matrix form we have

[εBΩΩω][βγ]=E[βγ],\displaystyle\begin{bmatrix}-\varepsilon_{B}&\Omega\\[1.99997pt] \Omega&\omega\end{bmatrix}\begin{bmatrix}\beta\\[1.99997pt] \gamma\end{bmatrix}=E\begin{bmatrix}\beta\\[1.99997pt] \gamma\end{bmatrix}, (19)

where we have identified the off-diagonal term g𝐤ϕX𝐤g\sum_{\mathbf{k}}\phi_{X\mathbf{k}} as the experimentally measurable Rabi coupling Ω\Omega introduced in Eq. (13). Evaluating the sum using Eqs. (5) and (7) allows us to relate the bare coupling gg to Ω\Omega and the exciton parameters,

g=ΩV0ZX.g=\Omega\frac{V_{0}}{\sqrt{Z_{X}}}. (20)

Since 1/V01/V_{0} diverges logarithmically with the cutoff Λ\Lambda in Eq. (7), we thus require g1/lnΛg\sim 1/\ln\Lambda to ensure that Ω\Omega is finite in our renormalization scheme.

Equation (19) yields the spectrum of two coupled oscillators:

E±osc\displaystyle E_{\pm}^{\rm osc} =εB+12(δ±δ2+4Ω2),\displaystyle=-\varepsilon_{B}+\frac{1}{2}\left(\delta\pm\sqrt{\delta^{2}+4\Omega^{2}}\right), (21)

where δ=ω(εB)\delta=\omega-(-\varepsilon_{B}) is the (bare) photon-exciton detuning. Thus we see that in the limit of small Rabi coupling we recover the usual spectrum [1, 2] of the lower ()(-) and upper (+)(+) polaritons. We also note that the photon and exciton fractions can be written as

|X±osc|2\displaystyle|X_{\pm}^{\rm osc}|^{2} =|β|2=11+Ω2(E±oscω)2,\displaystyle=|\beta|^{2}=\frac{1}{1+\frac{\Omega^{2}}{(E_{\pm}^{\rm osc}-\omega)^{2}}}, (22a)
|C±osc|2\displaystyle|C_{\pm}^{\rm osc}|^{2} =1|β|2=11+(E±oscω)2Ω2.\displaystyle=1-|\beta|^{2}=\frac{1}{1+\frac{(E_{\pm}^{\rm osc}-\omega)^{2}}{\Omega^{2}}}. (22b)

Note the similarity to our exact equation (16). In particular, the coefficients exactly match in the limit where the polariton energies approach the exciton energy.

In general, the exciton wave function will be modified by the coupling to light, unlike what we have assumed in the above analysis. We therefore now proceed to find the exact spectrum of Eq. (11), which also allows us to arrive at the modified Hopfield coefficients. To this end, we define fV0𝐤Λψ𝐤f\equiv V_{0}\sum_{\mathbf{k}}^{\Lambda}\psi_{\mathbf{k}} which is finite as Λ\Lambda\to\infty since V01/lnΛV_{0}\sim 1/\ln\Lambda and 𝐤ψ𝐤lnΛ\sum_{\mathbf{k}}\psi_{\mathbf{k}}\sim\ln\Lambda. Equation (11) then becomes

f\displaystyle f =V0𝐤1Eϵ¯𝐤(fgγ),\displaystyle=-V_{0}\sum_{\mathbf{k}}\frac{1}{E-\bar{\epsilon}_{\mathbf{k}}}(f-g\gamma), (23a)
γ\displaystyle\gamma =1EωgfV0.\displaystyle=\frac{1}{E-\omega}\frac{gf}{V_{0}}. (23b)

Rearranging Eq. (23a) and then inserting (23b), we find

1V0+𝐤1Eϵ¯𝐤\displaystyle\frac{1}{V_{0}}+\sum_{\mathbf{k}}\frac{1}{E-\bar{\epsilon}_{\mathbf{k}}} =gγf𝐤1Eϵ¯𝐤\displaystyle=\frac{g\gamma}{f}\sum_{\mathbf{k}}\frac{1}{E-\bar{\epsilon}_{\mathbf{k}}}
=1Eωg2V0𝐤1Eϵ¯𝐤\displaystyle=\frac{1}{E-\omega}\frac{g^{2}}{V_{0}}\sum_{\mathbf{k}}\frac{1}{E-\bar{\epsilon}_{\mathbf{k}}}
=1ωEΩ2ZX\displaystyle=\frac{1}{\omega-E}\frac{\Omega^{2}}{Z_{X}} (24)

where, in the last line, we have used Eq. (20) and the fact that V0𝐤1Eϵ¯𝐤1V_{0}\sum_{\mathbf{k}}\frac{1}{E-\bar{\epsilon}_{\mathbf{k}}}\to-1 as Λ\Lambda\to\infty. Replacing 1/V01/V_{0} on the left hand side using Eq. (7), we finally obtain Eq. (12) for the polariton spectrum:

(ωE)ln(EεB)=Ω2εB.(\omega-E)\ln\left(\frac{-E}{\varepsilon_{B}}\right)=\frac{\Omega^{2}}{\varepsilon_{B}}. (25)

The solutions to Eq. (25) correspond to the exact polariton energies for the Hamiltonian (II), under the assumption that E<0E<0 (which is always true for the lower polariton). Expanding the logarithm to leading order around E=εBE=-\varepsilon_{B} allows us to recover the spectrum of two coupled oscillators in Eq. (21).

To extract the photon fraction, we use Eq. (23b) and Eq. (20) to obtain

γ=fωEΩZX.\displaystyle\gamma=\frac{f}{\omega-E}\frac{\Omega}{\sqrt{Z_{X}}}. (26)

In the limit Λ\Lambda\to\infty, we have f=Z1|γ|2f=\sqrt{Z}\sqrt{1-|\gamma|^{2}} according to its definition and the form of the electron-hole wave function in Eq. (14) (dropping the ±\pm subscripts). We can thus solve for γ\gamma to obtain Eq. (16).

II.2.2 Effective photon-exciton detuning

As noted above, the coupling of the photon to the active semiconductor medium can also shift the cavity photon frequency. In experiment, the Rabi coupling and the exciton-photon detuning are parameters that are fitted from the observed polariton spectrum. Therefore, we now define an effective detuning that would result from such a fitting procedure. This allows us to directly relate the bare photon frequency ω\omega to a detuning δ\delta. In contrast to the case of Coulomb interactions [24], we find that the shift is independent of the cutoff Λ\Lambda for the contact-interaction model.

To capture the leading order correction to the cavity frequency when ΩεB\Omega\ll\varepsilon_{B}, we take ω=δ+sεB\omega=\delta+s-\varepsilon_{B} with ss small compared to εB\varepsilon_{B}, and then insert this into Eq. (12):

(δ+sεBE±)ln(E±εB)=Ω2εB.(\delta+s-\varepsilon_{B}-E_{\pm})\ln(\frac{-E_{\pm}}{\varepsilon_{B}})=\frac{\Omega^{2}}{\varepsilon_{B}}. (27)

Using the lowest-order expressions for the polariton energies, Eq. (21), and keeping terms only up to order Ω/εB\Omega/\varepsilon_{B} and δ/εB\delta/\varepsilon_{B}, we find

s=Ω22εB.\displaystyle s=\frac{\Omega^{2}}{2\varepsilon_{B}}. (28)

This leads to the expression in Eq. (17). As seen in Fig. 1, with this definition we have an excellent agreement with the coupled oscillators model up to Rabi couplings ΩεB\Omega\sim\varepsilon_{B}, well beyond our initial assumption that ΩεB\Omega\ll\varepsilon_{B}. In particular, had we instead used the bare detuning ωεB\omega-\varepsilon_{B}, our results would be shifted by εB/2\varepsilon_{B}/2 in panels (c) and (f), and hence the definition (17) substantially improves the agreement with the coupled-oscillator model.

II.3 Diagrammatic formulation

Refer to caption
Figure 2: Feynman diagrams for the two-body scattering process. (a) Dressed photon propagator (double wavy line) in terms of the bare photon propagator (single wavy line) and the photon self-energy (shaded ellipse). The black dots represent gg. (b) Photon self-energy consisting of all possible electron (solid line) and hole (dashed line) interaction terms. The loop consisting of an electron and a hole will be referred to as a polarization bubble, and is denoted Π(E)\Pi(E). (c) Electron-hole TT matrix (shaded rectangle), where the white dots represent V0-V_{0}.

In this section, we provide an alternative description of exciton-polaritons using a diagrammatic formulation. While this is equivalent to the operator approach described above, the advantage of the diagrammatic formulation is that it allows a straightforward characterization of repeated scattering processes. As we shall see, this enables us to obtain the polariton propagator in closed form, and it provides a convenient starting point for further few-body studies such as the polariton-electron scattering discussed in Sections III and IV.

We start by considering the properties of a photon interacting with an undoped 2D semiconductor in a microcavity. The microcavity photon will be modified by repeated interactions with electron-hole pairs, which includes both bound excitons and unbound electron-hole pairs in the continuum. Referring to Fig. 2, the resulting “dressed” photon is characterized by the self-energy Σ\Sigma [33]. Following Ref. [24], we obtain the dressed photon propagator D(𝐐,E)D(\mathbf{Q},E) from the Dyson equation in Fig. 2(a):

D(𝐐,E)\displaystyle D(\mathbf{Q},E) =D0(𝐐,E)+D0(𝐐,E)Σ(𝐐,E)D(𝐐,E)\displaystyle=D_{0}(\mathbf{Q},E)+D_{0}(\mathbf{Q},E)\Sigma(\mathbf{Q},E)D(\mathbf{Q},E)
=1D0(𝐐,E)1Σ(𝐐,E),\displaystyle=\frac{1}{D_{0}(\mathbf{Q},E)^{-1}-\Sigma(\mathbf{Q},E)}, (29)

where EE and 𝐐\mathbf{Q} are, respectively, the energy and momentum of the photon. In the absence of the active medium, we have the bare photon propagator

D0(𝐐,E)D0(Eϵ𝐐c)=1Eωϵ𝐐c+i0,D_{0}(\mathbf{Q},E)\equiv D_{0}(E-\epsilon_{\mathbf{Q}}^{c})=\frac{1}{E-\omega-\epsilon^{c}_{\mathbf{Q}}+i0}, (30)

where the imaginary infinitesimal +i0+i0 shifts the poles into the lower half of the complex plane and ensures that the photon propagates forward in time. Note that we can also incorporate the decay of photons from the microcavity by setting the imaginary part equal to the decay rate. However, we neglect this decay rate since it is much smaller than the Rabi coupling in the regime of strong light-matter coupling, and it therefore only has a negligible effect on the scattering. In other words, we consider a closed quantum system. In the following, we will always implicitly assume that the energy carries an infinitesimal positive imaginary part, that is, we will be working with retarded Green’s functions.

As shown in Fig. 2(b), the photon self-energy consists of two terms: Σ(𝐐,E)=Σ(1)(Eϵ𝐐X)+Σ(2)(Eϵ𝐐X)\Sigma(\mathbf{Q},E)=\Sigma^{(1)}(E-\epsilon^{X}_{\mathbf{Q}})+\Sigma^{(2)}(E-\epsilon^{X}_{\mathbf{Q}}), with ϵ𝐐X=Q2/2mX\epsilon^{X}_{\mathbf{Q}}=Q^{2}/2m_{X} the exciton kinetic energy and mX=me+mhm_{X}=m_{e}+m_{h} the exciton mass. These contain all possible processes that involve the excitation of an electron-hole pair, and they thus only depend on the energy in the exciton center-of-mass frame. Hence, the 𝐐\mathbf{Q} dependence simply appears as a shift in the energy.

Within the model (II), we have

Σ(1)(E)\displaystyle\Sigma^{(1)}(E) =g2𝐤Λ1Eϵ¯𝐤g2Π(E),\displaystyle=g^{2}\sum_{\mathbf{k}}^{\Lambda}\frac{1}{E-\bar{\epsilon}_{\mathbf{k}}}\equiv g^{2}\,\Pi(E), (31a)
Σ(2)(E)\displaystyle\Sigma^{(2)}(E) =g2Π2(E)𝒯0(E),\displaystyle=g^{2}\,\Pi^{2}(E)\,\mathcal{T}_{0}(E), (31b)

where Π(E)\Pi(E) is the polarization bubble diagram corresponding to an electron-hole pair. Since Π(E)lnΛ\Pi(E)\sim\ln\Lambda and the bare coupling g1/lnΛg\sim 1/\ln\Lambda, the term Σ(1)(E)\Sigma^{(1)}(E) vanishes when the momentum cutoff Λ\Lambda\to\infty. On the other hand, we have g2Π2(E)=Ω2/ZXg^{2}\Pi^{2}(E)=\Omega^{2}/Z_{X} from Eq. (20) and the fact that V0Π(E)1V_{0}\Pi(E)\to-1 as Λ\Lambda\to\infty.

The electron-hole TT matrix 𝒯0(E)\mathcal{T}_{0}(E) (or exciton propagator) is shown in Fig. 2(c), and corresponds to

𝒯0(E)=1V01Π(E).\mathcal{T}_{0}(E)=\frac{1}{-V_{0}^{-1}-\Pi(E)}. (32)

We can renormalize the bare V0V_{0} using Eq. (7) to finally obtain the standard expression in Eq. (8), see Appendix B.

Refer to caption
Figure 3: Polariton propagator (double line with an arrow) given by repeated interactions between electron-hole pairs and dressed photons — see Eq. (35). The symbols are the same as in Fig. 2.

Combining all of these expressions yields the explicit form for the dressed photon propagator:

D(𝐐,E)=1Eωϵ𝐐c+Ω2εB[ln(ϵ𝐐XEεB)]1,D(\mathbf{Q},E)=\frac{1}{E-\omega-\epsilon^{c}_{\mathbf{Q}}+\frac{\Omega^{2}}{\varepsilon_{B}}\,{\left[\ln\left(\frac{\epsilon^{X}_{\mathbf{Q}}-E}{\varepsilon_{B}}\right)\right]}^{-1}}, (33)

where ω\omega can be related to the physical photon-exciton detuning δ\delta via Eq. (17). To simplify the notation, we have not explicitly written the imaginary infinitesimal on the energy. The poles of the photon propagator contain information about the entire polariton spectrum. Specifically, the denominator of Eq. (33) is zero when we have

(ω+ϵ𝐐cE)ln[ϵ𝐐XEεB]=Ω2εB.\left(\omega+\epsilon^{c}_{\mathbf{Q}}-E\right)\ln\left[\frac{\epsilon^{X}_{\mathbf{Q}}-E}{\varepsilon_{B}}\right]=\frac{\Omega^{2}}{\varepsilon_{B}}. (34)

Indeed, for 𝐐=0\mathbf{Q}=0, this exactly corresponds to Eq. (25) from the operator approach. For general 𝐐\mathbf{Q}, we denote the polariton energy solutions of Eq. (34) as E±(𝐐)E_{\pm}(\mathbf{Q}).

In addition to the dressed photon propagator, we have what we call the “polariton propagator” 𝒯(𝐐,E)\mathcal{T}(\mathbf{Q},E) shown in Fig. 3, which instead focuses on the behavior of the matter component of the polariton. Like the dressed photon, it contains all the possible scattering processes involving an electron-hole pair or photon. In the absence of light-matter coupling, it reduces to 𝒯0(Eϵ𝐐X)\mathcal{T}_{0}(E-\epsilon^{X}_{\mathbf{Q}}), the electron-hole TT matrix at center-of-mass momentum 𝐐\mathbf{Q}. Suppressing the dependence on energy and momentum, Fig. 3 reads:

𝒯\displaystyle\mathcal{T} =𝒯0+g2(𝒯0ΠD+𝒯0ΠDΠ𝒯0+D+DΠ𝒯0)\displaystyle=\mathcal{T}_{0}+g^{2}\left(\mathcal{T}_{0}\Pi D+\mathcal{T}_{0}\Pi D\Pi\mathcal{T}_{0}+D+D\Pi\mathcal{T}_{0}\right)
=𝒯0+𝒯02Dg2Π2\displaystyle=\mathcal{T}_{0}+\mathcal{T}_{0}^{2}\,D\,g^{2}\Pi^{2}
=1𝒯01D0g2Π2.\displaystyle=\frac{1}{\mathcal{T}_{0}^{-1}-D_{0}\,g^{2}\Pi^{2}}. (35)

Here, we again use g1/lnΛg\sim 1/\ln\Lambda and ΠlnΛ\Pi\sim\ln\Lambda to remove terms that vanish in the limit Λ\Lambda\to\infty, while in the last line, we use the relation D=(D01g2Π2𝒯0)1D=(D_{0}^{-1}-g^{2}\Pi^{2}\mathcal{T}_{0})^{-1}. Thus, the polariton propagator finally reads:

𝒯(𝐐,E)=2π/mrln(ϵ𝐐XEεB)Ω2εB(Eωϵ𝐐c)1.\mathcal{T}(\mathbf{Q},E)=\frac{2\pi/m_{r}}{-\ln\left(\frac{\epsilon^{X}_{\mathbf{Q}}-E}{\varepsilon_{B}}\right)-\frac{\Omega^{2}}{\varepsilon_{B}}\left(E-\omega-\epsilon^{c}_{\mathbf{Q}}\right)^{-1}}. (36)

It is also straightforward to show that the polariton and photon propagators are related via

𝒯(𝐐,E)=𝒯0(Eϵ𝐐X)D0(Eϵ𝐐c)D(𝐐,E).\displaystyle\mathcal{T}(\mathbf{Q},E)=\frac{\mathcal{T}_{0}(E-\epsilon^{X}_{\mathbf{Q}})}{D_{0}(E-\epsilon_{\mathbf{Q}}^{c})}D(\mathbf{Q},E). (37)

The polariton propagator has precisely the same poles as the dressed photon propagator in Eq. (33). However, the poles of the photon propagator have residues corresponding to the photon fractions of the polaritons:

|C±(𝐐)|2=(1+ZX/Z±(𝐐)Ω2D02(𝐐,E±(𝐐)))1,\displaystyle|C_{\pm}(\mathbf{Q})|^{2}=\left(1+\frac{Z_{X}/Z_{\pm}(\mathbf{Q})}{\Omega^{2}D_{0}^{2}(\mathbf{Q},E_{\pm}(\mathbf{Q}))}\right)^{-1}, (38)

while the residues for the poles of the polariton propagator instead correspond to the matter component,

Z±(𝐐)|X±(𝐐)|2Z±(𝐐)[1|C±(𝐐)|2],\displaystyle Z_{\pm}(\mathbf{Q})|X_{\pm}(\mathbf{Q})|^{2}\equiv Z_{\pm}(\mathbf{Q})\left[1-|C_{\pm}(\mathbf{Q})|^{2}\right], (39)

with the electron-hole normalization factors

Z±(𝐐)=2π|E±(𝐐)ϵ𝐐X|mr.\displaystyle Z_{\pm}(\mathbf{Q})=\frac{2\pi|E_{\pm}(\mathbf{Q})-\epsilon^{X}_{\mathbf{Q}}|}{m_{r}}. (40)

At 𝐐=0\mathbf{Q}=0, the Hopfield coefficients and the normalization factors reduce to the corresponding expressions in Eqs. (16) and (15).

III Polariton-electron interactions in the Born approximation

Due to the complexity of the Coulomb interaction, an exact solution of few-body scattering processes such as polariton-electron and polariton-polariton scattering in strongly coupled light matter systems remains elusive. Instead, the interactions are typically approximated by considering their corresponding excitonic counterparts multiplied by appropriate powers of the excitonic Hopfield coefficient:

geP|X|2geX,gPP|X|4gXX,\displaystyle g_{eP}\simeq|X_{-}|^{2}g_{eX},\qquad g_{PP}\simeq|X_{-}|^{4}g_{XX}, (41)

where gePg_{eP} (geX)(g_{eX}) is the polariton-electron (exciton-electron) interaction constant, while gPPg_{PP} (gXX)(g_{XX}) is the polariton-polariton (exciton-exciton) interaction constant. Note that we have specialized to the lower polariton here, but similar expressions are used for the upper polariton with the simple replacement XX+X_{-}\to X_{+}. The standard approach is to calculate the interaction strengths within the Born approximation [11, 12, 13, 14], with effects due to very strong light-matter coupling included via the calculation of additional matrix elements [11, 16, 17, 18, 19]. The polariton-polariton interaction strength has also been calculated in TT-matrix approaches that treat the exciton as an inert object, first focusing on the case of cross-circularly polarized polaritons [34] and recently for both the singlet and triplet scattering configurations in multilayer systems [35].

In this paper, we provide the first exact microscopic calculation of polariton-electron scattering, within a simplified model of contact interactions between charges. It is therefore imperative to compare our results with the standard Born approximation. Therefore, in this section we will describe the results of the Born approximation within our model. We will furthermore argue that the Born approximation provides a strict upper bound on the interaction energy shift (in the limit of zero momentum) when there are no lower-lying bound states.

III.1 Operator approach

The analytic expression for the polariton wave function obtained in the previous section (see also the finite-momentum generalization in Appendix A) allows us to straightforwardly evaluate the Born approximation for polariton-electron scattering. For simplicity, we consider the frame where the center-of-mass momentum is zero, since in that case the different partial wave components of the scattering separate. Since the electron has a relatively flat dispersion, our results are likely to be insensitive to the actual electron momentum and will thus be dominated by the polariton momentum.

The Born approximation consists in approximating the scattering TT matrix by the first term in the Born series, which corresponds to evaluating the matrix element of the polariton-electron interaction potential. Taking the incoming (outgoing) polariton and electron to have momenta ±𝐩1\pm\mathbf{p}_{1} (±𝐩2\pm\mathbf{p}_{2}), this matrix element takes the form

TB(𝐩1,𝐩2)\displaystyle{T}^{B}(\mathbf{p}_{1},\mathbf{p}_{2})
=0|e𝐩2P𝐩2[HE(𝐩1)ϵ𝐩1e]P𝐩1e𝐩1|0,\displaystyle\quad=\bra{0}e_{-\mathbf{p}_{2}}P_{\mathbf{p}_{2}}[H-E_{-}(\mathbf{p}_{1})-\epsilon^{e}_{\mathbf{p}_{1}}]P_{\mathbf{p}_{1}}^{\dagger}e_{-\mathbf{p}_{1}}^{\dagger}\ket{0}, (42)

where the operator P𝐩P^{\dagger}_{\mathbf{p}} creates a polariton with momentum 𝐩\mathbf{p}, see Eq. (58) in Appendix A. We assume energy and momentum conservation such that the scattering is elastic, i.e., we take |𝐩1|=|𝐩2|=p|\mathbf{p}_{1}|=|\mathbf{p}_{2}|=p. Here, the subtraction of the energy of the non-interacting particles from the Hamiltonian, Eq. (II), ensures that we only evaluate the polariton-electron interaction potential. Using the exact polariton wave function from Appendix A, we find

TB(𝐩1,𝐩2)\displaystyle{T}^{B}(\mathbf{p}_{1},\mathbf{p}_{2})
=(ψ,me𝐩1/mX+𝐩2(𝐩1))2[E(𝐩1)ϵ𝐩1+𝐩2hϵ𝐩2e]\displaystyle=-\left(\psi_{-,m_{e}\mathbf{p}_{1}/m_{X}+\mathbf{p}_{2}}^{(\mathbf{p}_{1})}\right)^{2}\left[E_{-}(\mathbf{p}_{1})-\epsilon^{h}_{\mathbf{p}_{1}+\mathbf{p}_{2}}-\epsilon^{e}_{\mathbf{p}_{2}}\right]
=Z(𝐩1)|X(𝐩1)|2E(𝐩1)+ϵ𝐩1+𝐩2h+ϵ𝐩2e,\displaystyle=\frac{Z_{-}(\mathbf{p}_{1})|X_{-}(\mathbf{p}_{1})|^{2}}{-E_{-}(\mathbf{p}_{1})+\epsilon^{h}_{\mathbf{p}_{1}+\mathbf{p}_{2}}+\epsilon^{e}_{\mathbf{p}_{2}}}, (43)

where the sign in the middle line arises from the exchange of identical fermions. As anticipated, we see that the exciton fraction |X(𝐩1)|2|X_{-}(\mathbf{p}_{1})|^{2} naturally emerges from the normalization of the polariton wave function.

Refer to caption
Figure 4: Born approximation for elastic polariton-electron scattering. (a) Momentum dependence at a fixed angle ϕ12=0\phi_{12}=0 between 𝐩1\mathbf{p}_{1} and 𝐩2\mathbf{p}_{2}, where we take |𝐩1|=|𝐩2|=p|\mathbf{p}_{1}|=|\mathbf{p}_{2}|=p. Gray vertical line indicates the selected momentum in (b). (b) Angular dependence for a fixed p=0.1aX1p=0.1a_{X}^{-1}. The results are shown for parameters relevant to a single GaAs quantum well, i.e., Ω/εB=0.2\Omega/\varepsilon_{B}=0.2, me=0.067m0m_{e}=0.067m_{0}, mh=0.45m0m_{h}=0.45m_{0} [36], and mc=104m0m_{c}=10^{-4}m_{0}, with m0m_{0} the free electron mass, at negative detuning δ/εB=0.2\delta/\varepsilon_{B}=-0.2.

In Fig. 4 we illustrate the Born approximation for elastic polariton-electron scattering. The exciton fraction quickly increases at low momentum, which dominates the behavior at momenta up to the polariton dispersion inflection point above which the exciton fraction is very close to unity. Apart from the Hopfield coefficient |X(𝐩1)|2|X_{-}(\mathbf{p}_{1})|^{2}, there is also a momentum and angular dependence through the kinetic energies and the wave function normalization; however we find that this dependence is generally quite weak. Above the inflection point, it leads to a slow decline for increasing momentum.

Specializing to zero momentum, we may define a low-energy polariton-electron interaction constant gePBg_{eP}^{B}, which takes the form

gePBTB(𝟎,𝟎)=Z|X|2E=2π|X|2mr.\displaystyle g_{eP}^{B}\equiv T^{B}(\mathbf{0},\mathbf{0})=\frac{Z_{-}|X_{-}|^{2}}{-E_{-}}=\frac{2\pi|X_{-}|^{2}}{m_{r}}. (44)

This is precisely the exciton fraction multiplied by the exciton-electron interaction, geXB=4πεBaX2=2π/mrg_{eX}^{B}=4\pi\varepsilon_{B}a_{X}^{2}=2\pi/m_{r}. Note that this is quite close to the corresponding result obtained in the case of Coulomb electronic interactions [37]: Using the formalism developed in Ref. [24], we find geXB,Coulomb=(2π/mr)(23π/8)0.8geXBg_{eX}^{B,\mathrm{Coulomb}}=(2\pi/m_{r})(2-3\pi/8)\simeq 0.8g_{eX}^{B}. We emphasize that the polariton-electron Born approximation in Eq. (44) only depends on the light-matter coupling through the exciton fraction, while it is independent of the absolute strength of the light-matter coupling, Ω/εB\Omega/\varepsilon_{B}.

III.2 Diagrammatic approach

The polariton-electron scattering process can also be calculated diagrammatically, as shown in Fig. 5. In this case, the Born approximation corresponds to the first diagram on the right hand side, which involves the exchange of a hole with momentum 𝐩1+𝐩2\mathbf{p}_{1}+\mathbf{p}_{2} [38]. This gives

TB(𝐩1,𝐩2)\displaystyle T^{B}(\mathbf{p}_{1},\mathbf{p}_{2})
=Z(𝐩1)|X(𝐩1)|2Gh(𝐩1+𝐩2,E(𝐩1)ϵ𝐩2),\displaystyle\quad=-Z_{-}(\mathbf{p}_{1})|X_{-}(\mathbf{p}_{1})|^{2}G_{h}(\mathbf{p}_{1}+\mathbf{p}_{2},E_{-}(\mathbf{p}_{1})-\epsilon_{\mathbf{p}_{2}}), (45)

where the minus sign is a consequence of fermionic statistics and

Gh(𝐩,E)=1Eϵ𝐩h+i0\displaystyle G_{h}(\mathbf{p},E)=\frac{1}{E-\epsilon^{h}_{\mathbf{p}}+i0} (46)

is the free hole Green’s function. Each external polariton line in the diagram contributes the square root of the residue of the propagator 𝒯\cal{T} at the pole, and the normalization of the incoming and outgoing lines is the same since |𝐩1|=|𝐩2||\mathbf{p}_{1}|=|\mathbf{p}_{2}|. The residue is precisely Res[𝒯(𝐩,E(𝐩)]=Z(𝐩)|X(𝐩)|2\operatorname{Res}\left[\mathcal{T}(\mathbf{p},E_{-}(\mathbf{p})\right]=Z_{-}(\mathbf{p})|X_{-}(\mathbf{p})|^{2} — see Eq. (39). Using the definition of the hole propagator, and comparing Eq. (III.1) with Eq. (III.2) we thus see that the Born approximation within the two approaches coincides, as it should.

Refer to caption
Figure 5: Diagrammatic representation of the (unnormalized) polariton-electron scattering matrix tt (shaded rectangle). Black, white, and dashed circles represent electrons, holes, and polaritons, respectively, and black and blue text are the corresponding momenta and energies of the intermediate propagators.

III.3 Upper bound on the interaction energy shift

We now argue that the Born approximation is conceptually important since it provides an upper bound on the polariton-electron interaction constant gePg_{eP}. We start by noting that the ground-state energy for an electron and a lower polariton in an area AA can be written as

EeP=ΨeP|H|ΨePΨeP|ΨeP=E+gePA,\displaystyle E_{eP}=\frac{\expectationvalue{H}{\Psi_{eP}}}{\innerproduct{\Psi_{eP}}{\Psi_{eP}}}=E_{-}+\frac{g_{eP}}{A}, (47)

where |ΨeP\ket{\Psi_{eP}} is the exact interacting polariton-electron state. Here we have assumed that there are no lower-energy polariton-electron bound states so that EePE_{eP} consists of the non-interacting kinetic energy plus a two-body interaction term. We momentarily keep the area AA explicit so that we can keep track of powers of aX2/Aa_{X}^{2}/A. In particular, note that ΨeP|ΨeP=1+O(aX2/A)\innerproduct{\Psi_{eP}}{\Psi_{eP}}=1+O(a_{X}^{2}/A), where the last term arises from the composite nature of the polariton.

Rearranging Eq. (47) and keeping only terms up to order aX2/Aa_{X}^{2}/A then gives

gePA\displaystyle\frac{g_{eP}}{A} =ΨeP|(HE)|ΨeP\displaystyle=\bra{\Psi_{eP}}(H-E_{-})\ket{\Psi_{eP}}
0|e𝟎P𝟎(HE)P𝟎e𝟎|0,\displaystyle\leq\bra{0}e_{\mathbf{0}}P_{\mathbf{0}}(H-E_{-})P_{\mathbf{0}}^{\dagger}e_{\mathbf{0}}^{\dagger}\ket{0}, (48)

where the second line follows from the fact that the non-interacting state P𝟎e𝟎|0P_{\mathbf{0}}^{\dagger}e_{-\mathbf{0}}^{\dagger}\ket{0} can be viewed as a variational approximation to the exact interacting state |ΨeP\ket{\Psi_{eP}}. Thus, we finally arrive at

gePgePB.\displaystyle g_{eP}\leq g^{B}_{eP}. (49)

Hence, the Born approximation serves as a strict upper bound on the interaction energy shift. This observation is quite general and is independent of the details of the underlying interactions, e.g., it also applies to polariton-polariton scattering and to more realistic interactions between charged particles. The only requirement is that there are no lower energy bound states (such as trions) and that the Hamiltonian faithfully reproduces the low-energy physics. Therefore, any diverging interaction strength obtained in the absence of a bound state must be an artefact of the approximation [35].

IV Polariton-electron interactions

We now present our full diagrammatic calculation of the polariton-electron interaction strength, which is exact within the Hamiltonian (II). Our approach follows similar calculations originally introduced in the context of neutron-deuteron scattering by Skorniakov and Ter-Martirosian [39], and later applied to cold atomic gases [40, 41, 42, 43]. Subsequently, the theory of the effective three-body problem at finite momentum was developed in a series of papers [44, 45, 46, 47], and this has already been successfully applied to describe the strong atom-dimer attraction observed in mass-imbalanced ultracold Fermi gases [48].

The exact scattering of an electron and a polariton can be represented as an infinite sum of terms where the two electrons involved in the process exchange a hole, and the Born approximation corresponds to keeping only the first term in this series. While it is not possible to simply calculate each term in this sum separately and then sum them up, the key observation that enables an exact solution of the polariton-electron scattering is that the sum satisfies an integral equation as illustrated in Fig. 5. Indeed, iterating the right hand side of this equation generates all possible terms where the electrons exchange a hole. Diagrammatically, the equation for the polariton-electron TT matrix takes the exact same form as the exciton-electron TT matrix, the only difference being the replacement of the exciton with the polariton propagator, and the associated change in dispersion.

To compute the TT matrix, we take the incoming and outgoing electrons to have momentum 𝐩1-\mathbf{p}_{1} and 𝐩2-\mathbf{p}_{2} and energy ϵ𝐩1e\epsilon_{\mathbf{p}_{1}}^{e} and ϵ𝐩2e\epsilon_{\mathbf{p}_{2}}^{e}, respectively, while the polaritons have momenta 𝐩1\mathbf{p}_{1} and 𝐩2\mathbf{p}_{2} and energies Eϵ𝐩1eE-\epsilon^{e}_{\mathbf{p}_{1}} and Eϵ𝐩2eE-\epsilon^{e}_{\mathbf{p}_{2}}, respectively, with EE the total collision energy. The equation for the (unnormalized) polariton-electron TT matrix illustrated in Fig. 5 takes the explicit form [44, 47]

t(𝐩1,𝐩2)=Gh(𝐩1+𝐩2,Eϵ𝐩1eϵ𝐩2e)\displaystyle t(\mathbf{p}_{1},\mathbf{p}_{2})=-G_{h}(\mathbf{p}_{1}+\mathbf{p}_{2},E-\epsilon^{e}_{\mathbf{p}_{1}}-\epsilon^{e}_{\mathbf{p}_{2}})
𝐪Gh(𝐩1+𝐪,Eϵ𝐩1eϵ𝐪e)𝒯(𝐪,Eϵ𝐪e)t(𝐪,𝐩2),\displaystyle-\sum_{\mathbf{q}}G_{h}(\mathbf{p}_{1}+\mathbf{q},E-\epsilon^{e}_{\mathbf{p}_{1}}-\epsilon^{e}_{\mathbf{q}})\mathcal{T}(\mathbf{q},E-\epsilon^{e}_{\mathbf{q}})t(\mathbf{q},\mathbf{p}_{2}), (50)

where the minus signs on the right hand side follows from the exchange of identical electrons 333Note that the electron propagator Ge(𝐪,q0)=1/(q0ϵ𝐪+i0)G_{e}(-\mathbf{q},-q_{0})=1/(-q_{0}-\epsilon_{\mathbf{q}}+i0) in the last term in Fig. 5 has a simple pole in the upper half of the complex q0q_{0} plane, while all other propagators have their non-analytic structure only in the lower half plane. This allows us to perform the integral over q0q_{0} by closing the contour in the upper half plane, effectively removing that propagator and setting the electron dispersion to its on-shell value ϵ𝐪e=𝐪2/(2me)\epsilon^{e}_{\mathbf{q}}=\mathbf{q}^{2}/(2m_{e}) in the other propagators. For more details, see, e.g., Ref. [41].. We see that this is an integral equation in the first momentum index of tt, and we can therefore only set |𝐩1|=|𝐩2||\mathbf{p}_{1}|=|\mathbf{p}_{2}| at the end of calculation. In order to have on-shell scattering, we take the total energy to be E=E(𝐩2)+ϵ𝐩2E=E_{-}(\mathbf{p}_{2})+\epsilon_{\mathbf{p}_{2}}. The normalization is the same as in Eq. (III.2), and therefore the normalized and on-shell TT matrix takes the form

T(𝐩1,𝐩2)=Z(𝐩1)|X(𝐩1)|2t(𝐩1,𝐩2).\displaystyle T(\mathbf{p}_{1},\mathbf{p}_{2})=Z_{-}(\mathbf{p}_{1})|X_{-}(\mathbf{p}_{1})|^{2}t(\mathbf{p}_{1},\mathbf{p}_{2}). (51)

This quantifies the strength of scattering between electrons and polaritons.

IV.1 Scattering of slow particles

It is of particular interest to consider the scattering of polaritons at vanishing momentum due to the realization of polariton Bose-Einstein condensation [26]. First, it is important to clarify precisely what we mean by this limit. It is well-known [25, 28] that the scattering TT matrix of 2D quantum particles with short range interactions (such as the exciton-electron interaction between a charge and an induced dipole) must approach zero as the collision energy vanishes. However, in the case of polariton-electron scattering, this only becomes relevant for exceedingly small momenta pp below those where the typical scale of exciton-electron scattering 1/me\sim 1/m_{e} is comparable to 1/[mcln(εB/ϵ𝐩c)]1/[m_{c}\ln(\varepsilon_{B}/\epsilon_{\mathbf{p}}^{c})] [35]. Because of the very small photon mass, this momentum scale corresponds to a length scale much larger than the size of the universe, and this effect can therefore be discarded. Thus, we can safely take the limit p0p\to 0 in our numerical solutions, since we will never encounter such an extreme scale. We emphasize that this is a dramatic consequence of the lack of Galilean invariance in the polariton system due to the strong coupling of excitons and photons.

For scattering at small momenta, we set 𝐩2=0\mathbf{p}_{2}=0 and 𝐩1=𝐩\mathbf{p}_{1}=\mathbf{p} in Eq. (50) and perform the integration over the angle between 𝐩\mathbf{p} and 𝐪\mathbf{q} assuming E<0E<0. The remaining integral equation is then solved using Gauss-Legendre quadrature; for more details, see Appendix C. As a result, we obtain the exact polariton-electron interaction constant

geP=T(0,0)=Z|X|2t(0,0),\displaystyle g_{eP}=T(0,0)=Z_{-}|X_{-}|^{2}t(0,0), (52)

where we use the same normalization as in Eq. (51). Note the similarities between this expression and that in Eq. (44) within the Born approximation.

Refer to caption
Figure 6: Comparison between the polariton-electron interaction constant gePg_{eP} within our exact diagrammatic calculation (blue solid line), the Born approximation (gray dashed line), and the approximation Eq. (54) (yellow dotted line). Top row: GaAs quantum well microcavity systems with mc=104m0m_{c}=10^{-4}m_{0}, me=0.067m0m_{e}=0.067m_{0}, mh=0.45m0m_{h}=0.45m_{0} [36, 50], and εeX/εB1.18\varepsilon_{eX}/\varepsilon_{B}\simeq 1.18 obtained from the low-energy exciton-electron TT matrix. Bottom row: TMD monolayer microcavity systems with me=mh=m0m_{e}=m_{h}=m_{0} and εeX/εB=0.47\varepsilon_{eX}/\varepsilon_{B}=0.47 [51]; the parameters correspond approximately to the case of a MoSe2  [52], MoS2 [53], or WSe2 [54, 55] monolayer in (c) or to a WS2 [56] monolayer in (d).

Figure 6 shows a comparison of the polariton-electron interaction constant obtained from different methods for parameters relevant to both TMDs and GaAs quantum wells. The exact diagrammatic calculation, gePg_{eP}, can be used as a standard to gauge the correctness of other calculations. We see that generally the interaction strength increases as the polariton goes from being photonic at negative detuning to more excitonic at positive detuning, in agreement with the expectation from the Born approximation, Eq. (44). However, we find that the exact calculation features a peak at small positive detuning when the ratio Ω/εB\Omega/\varepsilon_{B} is not too large, which is a qualitative feature that is missed by the Born approximation. Furthermore, the Born approximation generically overestimates the interaction strength, and this effect is most dramatic for relatively small values of the ratio Ω/εB\Omega/\varepsilon_{B} such as in TMDs. For larger Ω/εB\Omega/\varepsilon_{B}, the exact result approaches gePBg_{eP}^{B} and thus we expect the Born approximation to be more accurate for GaAs quantum wells, which is consistent with recent measurements of the polariton-polariton interaction strength [57]. Note that Ω/εB0.5\Omega/\varepsilon_{B}\gtrsim 0.5 applies to the case of multiple GaAs quantum wells and thus a full description requires us to take account of the non-trivial effects of multiple layers [35].

We can understand these results by noting that at momenta above the inflection point, the polariton propagator in Eq. (36) quickly approaches the exciton propagator, 𝒯0{\cal T}_{0}, in Eq. (8). Taking the limit 𝒯𝒯0{\cal T}\to{\cal T}_{0}, we observe that Eq. (50) has the exact same functional form as in exciton-electron scattering [47], the only difference being the change in collision energy due to the light-matter coupling. This motivates the interpretation of polariton-electron scattering as an off-shell exciton-electron scattering process, similarly to recent results on polariton-polariton interactions [35]. To make this explicit, we consider the universal form of the low-energy exciton-electron scattering TT matrix

TeX(E)2πmeX1ln[εeX/(E+εB)],\displaystyle T_{eX}(E)\simeq\frac{2\pi}{m_{eX}}\frac{1}{\ln[-\varepsilon_{eX}/(E+\varepsilon_{B})]}, (53)

where we define meX=memX/(me+mX)m_{eX}=m_{e}m_{X}/(m_{e}+m_{X}) as the reduced exciton-electron mass. Unlike the light-matter coupled system, the exciton-electron scattering only depends on the collision energy, since the exciton-electron system is Galilean invariant. Equation (53) is valid at collision energies |E+εB|εeX|E+\varepsilon_{B}|\ll\varepsilon_{eX}, where the energy scale εeXεB\varepsilon_{eX}\sim\varepsilon_{B} depends on the precise details of the electronic interactions and on the electron-hole mass ratio 444In general, there are finite range corrections to Eq. (53), which add a term meXre2(E+εB)\sim m_{eX}r_{e}^{2}(E+\varepsilon_{B}) in the denominator [28]. In the absence of other scales, the effective range reaXr_{e}\sim a_{X} and therefore such terms are not important when |E+εB|εeX|E+\varepsilon_{B}|\ll\varepsilon_{eX}.. We thus estimate the polariton-electron scattering in the limit of low momentum by the off-shell exciton-electron expression

geP|X|2TeX(E),\displaystyle g_{eP}\simeq|X_{-}|^{2}T_{eX}(E_{-}), (54)

which is valid when |E+εB|εB|E_{-}+\varepsilon_{B}|\ll\varepsilon_{B}.

Refer to caption
Figure 7: Polariton-electron scattering TT matrix at finite momentum for parameters corresponding to TMD monolayers. We show our exact diagrammatic calculation as a black solid line, while the yellow dashed line is the corresponding exciton-electron scattering TT matrix. Here, me=mh=m0m_{e}=m_{h}=m_{0} and mc=104m0m_{c}=10^{-4}m_{0}, and in the top row Ω/εB=0.025\Omega/\varepsilon_{B}=0.025 corresponding approximately to the case of a MoSe2 [52], MoS2 [53], or WSe2 [54, 55] monolayer, while in the bottom row Ω/εB=0.05\Omega/\varepsilon_{B}=0.05 corresponding to a WS2 [56] monolayer.

Equation (54) yields the yellow dashed line in Fig. 6, which is seen to closely follow the exact polariton-electron scattering result for sufficiently low Rabi coupling, reproducing all qualitative features. In particular, it works exceptionally well in the case of TMD monolayers, which supports the approximation for polariton-polariton interactions introduced in Ref. [35]. Note that Eq. (54) predicts a resonance when the (negative) collision energy is comparable to εeX\varepsilon_{eX}, such as in GaAs with a very large Rabi coupling in panel (b). However, this resonance is spurious and can be cured by including higher order terms in the phase shift, and it is also absent in our exact calculation.

While Eq. (54) is conceptually similar to the Born approximation, Eq. (44), we see that the fact that the exciton-electron scattering should be considered off-shell leads to a qualitative difference in the behavior as a function of detuning. We stress that there is no a priori reason why this conclusion should not generalize beyond contact electronic interactions to Coulomb or Keldysh [59] type potentials in, e.g., GaAs or TMDs: At large electron-exciton separation rr, both of these electronic potentials will result in an effective exciton-electron potential that decays like 1/r4-1/r^{4} corresponding to the interaction between a charge and an induced dipole. Such a potential qualifies as short ranged [27], and therefore the low-energy exciton-electron scattering necessarily follows the form in Eq. (53). Furthermore, this approximation relies only on a single quantity, the low-energy scale εeX\varepsilon_{eX}, which can be fitted in experiment or calculated in exact solutions of the exciton-electron three-body problem such as in the recent work by Fey et al. [60]. Finally, we note that since εeX\varepsilon_{eX} only appears under a logarithm, our results are not very sensitive to realistic variations in this parameter, and they may therefore be expected to be quite accurate even at a quantitative level.

IV.2 Polariton-electron interactions at finite momentum

We now turn to the interaction of an electron and a polariton at finite momentum which, as we shall show, can be strongly enhanced compared with their exciton-electron counterpart. As previously noted, we will restrict our attention to the case of zero center-of-mass momentum, which allows us to perform a partial wave decomposition of the TT matrix as follows [25]:

t(𝐩1,𝐩2)\displaystyle t(\mathbf{p}_{1},\mathbf{p}_{2}) =l=0(2δl0)cos(lϕ12)tl(p1,p2),\displaystyle=\sum_{l=0}^{\infty}(2-\delta_{l0})\cos(l\phi_{12})t_{l}(p_{1},p_{2}), (55)

where ll is the angular momentum quantum number and ϕ12\phi_{12} is the angle between 𝐩1\mathbf{p}_{1} and 𝐩2\mathbf{p}_{2}. Obviously, the scattering in the limit of zero momentum discussed in the previous subsection corresponds to ss-wave, l=0l=0. This decomposition allows us to solve the integral equation as a function of a single parameter, the magnitude of the incident relative momentum. The main challenge then is that the electron-hole TT matrix in Eq. (50) has a simple pole when q=p2q=p_{2} [44], and we discuss how to deal with this in Appendix C. Once we have solved for the partial wave amplitudes tlt_{l}, we obtain the normalized scattering TT matrix in the ll’th partial wave via

Tl(p)=Z(𝐩)|X(𝐩)|2tl(p),\displaystyle T_{l}(p)=Z_{-}(\mathbf{p})|X_{-}(\mathbf{p})|^{2}t_{l}(p), (56)

similarly to Eq. (51). In this work, we focus on ss-wave scattering, which we denote by Ts(p)T_{s}(p). However, we note that higher partial waves (in particular pp-wave) can become important in the scattering of polaritons with heavy holes [45, 47] due to the presence of hole-hole-electron trion bound states which have previously been discussed in the context of quantum wells [61] and TMD monolayers [62, 63] or strongly screened electronic interactions [64, 51].

Refer to caption
Figure 8: Polariton-electron scattering TT matrix at finite momentum for parameters relevant to GaAs semiconductor microcavities. We show our exact diagrammatic calculation (black solid line) and the corresponding exciton-electron scattering TT matrix (yellow dashed line) for me=0.067m0m_{e}=0.067m_{0}, mh=0.45m0m_{h}=0.45m_{0} [36], and mc=104m0m_{c}=10^{-4}m_{0}. In the top row we have Ω/εB=0.1\Omega/\varepsilon_{B}=0.1 while in the bottom row Ω/εB=0.5\Omega/\varepsilon_{B}=0.5 [50].

Figures 7 and 8 show our results for the polariton-electron scattering at finite relative momentum for the case of a TMD monolayer and for a GaAs quantum well, respectively. A striking aspect of our results is the presence of a resonance-like peak at momenta close to the polariton inflection point. The peak is most pronounced for zero and negative detuning and it is enhanced with increased Rabi coupling. By contrast, we find that the peak is completely absent in the case of exciton-electron scattering (also shown in the figures), where we obtain the corresponding TT matrix using Eq. (53) at collision energy E+εB=p2/2meXE+\varepsilon_{B}=p^{2}/2m_{eX}. At a qualitative level, the results for GaAs and a TMD monolayer embedded in a microcavity are similar, although the generally larger ratios of Rabi coupling to exciton binding energy in GaAs means that the relative enhancement of interactions in GaAs microcavity is larger. For relative momenta above the polariton inflection point, we see that the polariton-electron TT matrix quickly reduces to the exciton-electron TT matrix, as we would expect since in that limit the exciton fraction approaches unity.

As in the case of scattering at low momenta, we can understand the resonance-like feature in the polariton-electron scattering at finite momentum in terms of off-shell exciton-electron scattering. Using a similar reasoning to that which led to Eq. (54), we approximate

Ts(p)|X(𝐩)|2TeX(E(𝐩)+ϵ𝐩e).\displaystyle T_{s}(p)\simeq|X_{-}(\mathbf{p})|^{2}T_{eX}(E_{-}(\mathbf{p})+\epsilon_{\mathbf{p}}^{e}). (57)

We illustrate this in Fig. 9 for the case of the TMDs with a smaller ratio of Rabi coupling to exciton binding energy (MoSe2, WSe2, or MoS2). We see that this approximation works extremely well for almost all momenta, the only difference being a narrow region around the momentum where the collision energy E+εBE+\varepsilon_{B} vanishes. By analyzing the two contributions in Eq. (57), we conclude that the resonance-like feature is due to a competition between the exciton fraction — which quickly approaches unity when the momentum approaches the polariton inflection point — and the fact that scattering is suppressed when the collision energy approaches zero. Figure 9 also shows that the polariton-electron TT matrix is nearly purely real at the peak, which we can also understand from Eq. (57) since exciton-electron scattering only features an imaginary part for positive collision energy.

Crucially, our results in Figures 7-9 imply that the polariton-electron TT matrix is essentially never smaller than that of an electron and an exciton at a given momentum, the only exception being at a very large negative detuning. This is in spite of the suppression due to the reduced exciton fraction in the polariton, and it demonstrates that the strong energy shift of the scattering process due to the light-matter coupling is the dominant effect. Our result can thus be understood as an effective photon-mediated enhancement of exciton-electron scattering, which is a major qualitative difference compared with the commonly applied assumption in the Born approximation, Eq. (41). As we have derived this effect in a fully microscopic model, and since it agrees with the universal low-energy behavior, Eq. (57), we expect that this conclusion carries over to more realistic interactions between charged particles.

Given the general nature of our arguments, the enhancement of interactions between polaritons and electrons is also expected to carry over to other geometries, such as systems where a 2D electron gas is in proximity to a 2D exciton-polariton condensate. Here, it would be interesting to investigate whether the enhanced interactions could lead to stronger drag effects [65], or to more robust superconductivity [66, 67].

Refer to caption
Figure 9: Comparison between the real and imaginary parts of the electron-polariton scattering TT matrix. Black and blue solid lines correspond to the real and imaginary part of Ts(p)T_{s}(p) in Eq. (56), respectively. Black and blue dashed lines are the real and imaginary parts of the off-shell exciton approximation in Eq. (57), respectively. For this comparison, we take the parameters me=mh=m0m_{e}=m_{h}=m_{0}, mc=104m0m_{c}=10^{-4}m_{0}, δ/εB=0\delta/\varepsilon_{B}=0, and Ω/εB=0.025\Omega/\varepsilon_{B}=0.025.

V Conclusions and outlook

We have presented a microscopic approach to exciton-polaritons in a 2D semiconductor embedded in a microcavity. By treating the semiconductor interactions as strongly screened, we have found an analytic expression for the polariton propagator, which serves as a starting point for further few- and many-body calculations. We have applied our microscopic description of polaritons to the calculation of the scattering of electrons and polaritons in a manner that is exact within our approximation. Remarkably, we have shown that polariton-electron scattering is nearly always stronger than exciton-electron scattering at a given momentum, in contrast to the prevailing belief based on the Born approximation. In particular, we have shown that the interactions can be strongly enhanced up to and beyond the inflection point of the polariton dispersion, giving rise to a resonance-like feature. While this qualitatively new and unexpected behavior may appear counter-intuitive, we have argued that the enhancement of scattering is due to the strong light-matter coupling that shifts the collision energy compared with exciton-electron scattering, and hence this is a generic effect which is independent of our approximation of screened interactions.

Our approach can be directly applied to a large class of other few-body problems in light-matter coupled systems. Of particular interest is the case of spin-polarized electrons interacting with an exciton-polariton of opposite circular polarization. In this case, the exciton-electron system features a trion bound state which leads to resonantly enhanced interactions, and this has recently enabled the observation of polaron-polaritons [68, 69]. Furthermore, using the four-body techniques developed in the context of cold atomic gases Refs. [70, 71, 41, 42], our approach can be straightforwardly generalized to the calculation of polariton-polariton and polariton-exciton interactions. This would, for instance, allow a fully microscopic description of the scattering resonances due to coupling to a biexciton bound state, as observed in Refs. [9, 72, 73] and also analyzed in Refs. [74, 75]. Finally, the microscopic theory of exciton-polaritons for the case of Coulomb interactions developed in Ref. [24] also allows a natural diagrammatic representation, which can be applied to few-body scattering problems. Hence our approach holds promise as a fully microscopic theory in which to obtain key quantities in exciton-polariton systems.

Acknowledgements.
We thank Olivier Bleu, Dmitry Efimkin, Eliezer Estrecho, Emma Laird, David Neilson, Elena Ostrovskaya, and Maciej Pieczarka for useful discussions, and we acknowledge support from the Australian Research Council Centre of Excellence in Future Low-Energy Electronics Technologies (CE170100039). JL is furthermore supported through the Australian Research Council Future Fellowship FT160100244.

Appendix A Polariton at finite momentum

In this Appendix, we apply the operator approach discussed in Sec. II.2 to a polariton at finite momentum 𝐐\mathbf{Q}. The polariton creation operator P𝐐P^{\dagger}_{\mathbf{Q}} is defined as:

P𝐐|0|Ψ𝐐=𝐤ψ𝐤(𝐐)e𝐐e+𝐤h𝐐h𝐤|0+γ(𝐐)c𝐐|0,P_{\mathbf{Q}}^{\dagger}\ket{0}\equiv\ket{\Psi_{\mathbf{Q}}}=\sum_{\mathbf{k}}\psi_{\mathbf{k}}^{(\mathbf{Q})}\,e^{\dagger}_{\mathbf{Q}_{e}+\mathbf{k}}h^{\dagger}_{\mathbf{Q}_{h}-\mathbf{k}}\ket{0}+\gamma^{(\mathbf{Q})}c^{\dagger}_{\mathbf{Q}}\ket{0}, (58)

where 𝐐e,hme,h𝐐/mX\mathbf{Q}_{e,h}\equiv m_{e,h}\,\mathbf{Q}/m_{X} with mX=me+mhm_{X}=m_{e}+m_{h} the exciton mass. The Schrödinger equation can be obtained by projecting (EH^)|Ψ𝐐=0(E-\hat{H})\ket{\Psi_{\mathbf{Q}}}=0 onto the electron-hole and photon parts of Eq. (58), which gives:

(Eϵ𝐐Xϵ¯𝐤)ψ𝐤(𝐐)\displaystyle\left(E-\epsilon^{X}_{\mathbf{Q}}-\bar{\epsilon}_{\mathbf{k}}\right)\psi^{(\mathbf{Q})}_{\mathbf{k}} =V0𝐤ψ𝐤(𝐐)+gγ(𝐐),\displaystyle=-V_{0}\sum_{\mathbf{k}^{\prime}}\psi^{(\mathbf{Q})}_{\mathbf{k}^{\prime}}+g\gamma^{(\mathbf{Q})}, (59a)
(Eωϵ𝐐c)γ(𝐐)\displaystyle\left(E-\omega-\epsilon^{c}_{\mathbf{Q}}\right)\gamma^{(\mathbf{Q})} =g𝐤φ𝐤(𝐐).\displaystyle=g\sum_{\mathbf{k}}\varphi^{(\mathbf{Q})}_{\mathbf{k}}. (59b)

The momentum 𝐐\mathbf{Q} serves as an external parameter in Eq. (59) and thus the renormalization procedure is the same as in Sec. II.2, as we now demonstrate.

We first consider the case of small exciton-photon Rabi coupling where the polariton energy is close to the exciton energy, i.e., E=εB+ϵ𝐐X+ΔEE=-\varepsilon_{B}+\epsilon^{X}_{\mathbf{Q}}+\Delta E. In the limit ΔE0\Delta E\to 0, the electron-hole part of the polariton wave function ψ𝐤(𝐐)\psi_{\mathbf{k}}^{(\mathbf{Q})} can be approximated as proportional to the exciton wave function, ψ𝐤(𝐐)β(𝐐)ϕX𝐤\psi_{\mathbf{k}}^{(\mathbf{Q})}\simeq\beta^{(\mathbf{Q})}\,\phi_{X\mathbf{k}}, where β(𝐐)\beta^{(\mathbf{Q})} is a complex number. Within this approximation, Eq. (59) takes the form:

(E+εBϵ𝐐X)β(𝐐)\displaystyle(E+\varepsilon_{B}-\epsilon^{X}_{\mathbf{Q}})\beta^{(\mathbf{Q})} =γ(𝐐)g𝐤ϕX𝐤,\displaystyle=\gamma^{(\mathbf{Q})}\,g\sum_{\mathbf{k}}\phi_{X\mathbf{k}}, (60a)
(Eωϵ𝐐c)γ(𝐐)\displaystyle(E-\omega-\epsilon^{c}_{\mathbf{Q}})\gamma^{(\mathbf{Q})} =β(𝐐)g𝐤ϕX𝐤.\displaystyle=\beta^{(\mathbf{Q})}\,g\sum_{\mathbf{k}}\phi_{X\mathbf{k}}. (60b)

Written in matrix form we have

[εB+ϵ𝐐XΩΩω+ϵ𝐐c][β(𝐐)γ(𝐐)]=E[β(𝐐)γ(𝐐)],\displaystyle\begin{bmatrix}-\varepsilon_{B}+\epsilon^{X}_{\mathbf{Q}}&\Omega\\[1.99997pt] \Omega&\omega+\epsilon^{c}_{\mathbf{Q}}\end{bmatrix}\begin{bmatrix}\beta^{(\mathbf{Q})}\\[1.99997pt] \gamma^{(\mathbf{Q})}\end{bmatrix}=E\begin{bmatrix}\beta^{(\mathbf{Q})}\\[1.99997pt] \gamma^{(\mathbf{Q})}\end{bmatrix}, (61)

where we have identified the off-diagonal term g𝐤ϕX𝐤g\sum_{\mathbf{k}}\phi_{X\mathbf{k}} as the experimentally measurable Rabi coupling Ω\Omega introduced in Eq. (13). Equation (61) yields the spectrum of two coupled oscillators as

E±osc=12[ϵ𝐐X+δ+ϵ𝐐c±(δ+ϵ𝐐cϵ𝐐X)2+4Ω2]εB,\displaystyle E^{\rm osc}_{\pm}=\frac{1}{2}\left[\epsilon^{X}_{\mathbf{Q}}+\delta+\epsilon_{\mathbf{Q}}^{c}\pm\sqrt{(\delta+\epsilon_{\mathbf{Q}}^{c}-\epsilon^{X}_{\mathbf{Q}})^{2}+4\Omega^{2}}\,\right]-\varepsilon_{B}, (62)

with the Hopfield coefficients

|C±osc(𝐐)|2\displaystyle\absolutevalue{C_{\pm}^{\rm osc}(\mathbf{Q})}^{2} =12(1ϵ𝐐Xδϵ𝐐c(ϵ𝐐Xδϵ𝐐c)2+4Ω2),\displaystyle=\frac{1}{2}\left(1\mp\frac{\epsilon^{X}_{\mathbf{Q}}-\delta-\epsilon_{\mathbf{Q}}^{c}}{\sqrt{(\epsilon^{X}_{\mathbf{Q}}-\delta-\epsilon_{\mathbf{Q}}^{c})^{2}+4\Omega^{2}}}\right), (63a)
|X±osc(𝐐)|2\displaystyle\absolutevalue{X_{\pm}^{\rm osc}(\mathbf{Q})}^{2} =12(1±ϵ𝐐Xδϵ𝐐c(ϵ𝐐Xδϵ𝐐c)2+4Ω2).\displaystyle=\frac{1}{2}\left(1\pm\frac{\epsilon^{X}_{\mathbf{Q}}-\delta-\epsilon_{\mathbf{Q}}^{c}}{\sqrt{(\epsilon^{X}_{\mathbf{Q}}-\delta-\epsilon_{\mathbf{Q}}^{c})^{2}+4\Omega^{2}}}\right). (63b)

After the renormalization, we follow a similar derivation as that carried out in Eqs. (23) and (24) to find that the exact polariton dispersion E±(𝐐)E_{\pm}(\mathbf{Q}) satisfies

[ω+ϵ𝐐cE±(𝐐)]ln[E±(𝐐)+ϵ𝐐XεB]=Ω2εB.\left[\omega+\epsilon^{c}_{\mathbf{Q}}-E_{\pm}(\mathbf{Q})\right]\ln\left[\frac{-E_{\pm}(\mathbf{Q})+\epsilon^{X}_{\mathbf{Q}}}{\varepsilon_{B}}\right]=\frac{\Omega^{2}}{\varepsilon_{B}}. (64)

With the polariton dispersion established, we can return to Eq. (59) to solve for the electron-hole wave function. Since the right hand side of Eq. (59a) is independent of 𝐤\mathbf{k}, we immediately know that the electron-hole wave function ψ±,𝐤(𝐐)\psi_{\pm,\mathbf{k}}^{(\mathbf{Q})} has the same functional form as in Eq. (14):

ψ±,𝐤(𝐐)=Z±(𝐐)1|γ±(𝐐)|2E±(𝐐)+ϵ𝐐X+ϵ¯𝐤,\displaystyle\psi_{\pm,\mathbf{k}}^{(\mathbf{Q})}=\frac{\sqrt{Z_{\pm}{(\mathbf{Q})}}\sqrt{1-|\gamma_{\pm}^{(\mathbf{Q})}|^{2}}}{-E_{\pm}(\mathbf{Q})+\epsilon^{X}_{\mathbf{Q}}+\bar{\epsilon}_{\mathbf{k}}}, (65)

with

Z±(𝐐)\displaystyle Z_{\pm}{(\mathbf{Q})} =[𝐤1(E±(𝐐)ϵ𝐐Xϵ¯𝐤)2]1\displaystyle=\left[\sum_{\mathbf{k}}\frac{1}{(E_{\pm}(\mathbf{Q})-\epsilon^{X}_{\mathbf{Q}}-\bar{\epsilon}_{\mathbf{k}})^{2}}\right]^{-1}
=2π|E±(𝐐)ϵ𝐐X|mr.\displaystyle=\frac{2\pi\left|E_{\pm}(\mathbf{Q})-\epsilon^{X}_{\mathbf{Q}}\right|}{m_{r}}. (66)

We can also extract the Hopfield coefficients in Eq. (16) in the same manner as in Sec. II.2. This straightforwardly yields the same expressions as in the diagrammatic section II.3, and allows us to obtain an analytic expression for the electron-hole wave function in terms of experimentally measurable parameters.

Appendix B Electron-hole T matrix

In this Appendix, we derive the expression for the exciton propagator, Eq. (8). Starting from the electron-hole TT matrix diagram shown in Fig. 2(c), the exciton propagator at zero momentum reads:

𝒯0(E)=1V01Π(E),\mathcal{T}_{0}(E)=\frac{1}{-V_{0}^{-1}-\Pi(E)}, (67)

where V01V_{0}^{-1} has been related to an integral by Eq. (7),

1V0=𝐤Λ1εB+ϵ¯𝐤.\displaystyle\frac{1}{V_{0}}=\sum_{\mathbf{k}}^{\Lambda}\frac{1}{\varepsilon_{B}+\bar{\epsilon}_{\mathbf{k}}}. (68)

The bubble integral Π(E)\Pi(E) in Eq. (67) can be worked out explicitly:

Π(E)=\displaystyle\Pi(E)= 𝐤Λ1Eϵ¯𝐤\displaystyle\sum_{\mathbf{k}}^{\Lambda}\frac{1}{E-\bar{\epsilon}_{\mathbf{k}}}
=\displaystyle= 𝐤Λ(1Eϵ¯𝐤+1εB+ϵ¯𝐤)𝐤Λ1εB+ϵ¯𝐤\displaystyle\sum_{\mathbf{k}}^{\Lambda\to\infty}\left(\frac{1}{E-\bar{\epsilon}_{\mathbf{k}}}+\frac{1}{\varepsilon_{B}+\bar{\epsilon}_{\mathbf{k}}}\right)-\sum_{\mathbf{k}}^{\Lambda}\frac{1}{\varepsilon_{B}+\bar{\epsilon}_{\mathbf{k}}}
=\displaystyle= mr2πln(EεB)𝐤Λ1εB+ϵ¯𝐤,\displaystyle\frac{m_{r}}{2\pi}\ln\left(\frac{-E}{\varepsilon_{B}}\right)-\sum_{\mathbf{k}}^{\Lambda}\frac{1}{\varepsilon_{B}+\bar{\epsilon}_{\mathbf{k}}}, (69)

where we can take the limit Λ\Lambda\to\infty in the difference in the second line, since this is ultraviolet convergent, while the last sum diverges as lnΛ\sim\ln\Lambda when Λ\Lambda\to\infty. Inserting Eqs. (69) and (68) into Eq. (67) we obtain the expression for 𝒯0(E)\mathcal{T}_{0}(E):

𝒯0(E)=2π/mrln[E/εB+i0]+iπ.\mathcal{T}_{0}(E)=\frac{2\pi/m_{r}}{-\ln\left[E/\varepsilon_{B}+i0\right]+i\pi}. (70)

Appendix C Partial wave decomposition and numerical solution of the polariton-electron scattering integral equation

We now outline how we numerically investigate the key equation of polariton-electron scattering, namely Eq. (50), which we rewrite here in integral form by taking the continuum limit

t(𝐩1,𝐩2)=\displaystyle t(\mathbf{p}_{1},\mathbf{p}_{2})= Gh(𝐩1+𝐩2,Eϵ𝐩1eϵ𝐩2e)\displaystyle-G_{h}(\mathbf{p}_{1}+\mathbf{p}_{2},E-\epsilon^{e}_{\mathbf{p}_{1}}-\epsilon^{e}_{\mathbf{p}_{2}})
0qdq2π02πdϕ2πGh(𝐩1+𝐪,Eϵ𝐩1eϵ𝐪e)\displaystyle-\int_{0}^{\infty}\frac{q\,dq}{2\pi}\int_{0}^{2\pi}\frac{d\phi}{2\pi}G_{h}(\mathbf{p}_{1}+\mathbf{q},E-\epsilon^{e}_{\mathbf{p}_{1}}-\epsilon^{e}_{\mathbf{q}})
×𝒯(𝐪,Eϵ𝐪e)t(𝐪,𝐩2),\displaystyle\qquad\qquad\quad\times\mathcal{T}(\mathbf{q},E-\epsilon^{e}_{\mathbf{q}})t(\mathbf{q},\mathbf{p}_{2}), (71)

where we again take the total energy E=E(𝐩2)+ϵ𝐩2E=E_{-}(\mathbf{p}_{2})+\epsilon_{\mathbf{p}_{2}}. This represents an integral equation in the first momentum of the function t(𝐩1,𝐩2)t(\mathbf{p}_{1},\mathbf{p}_{2}), while it only depends parametrically on the second. There are two main challenges to deal with when solving this equation. First, the integral equation depends on the angle between incident and outgoing relative momenta. This angular dependence increases the numerical complexity, which we resolve by performing a partial wave decomposition. Second, the polariton propagator has a simple pole, which we address by using a principal value prescription.

Let us first discuss the partial wave decomposition. As in Eq. (55), we define

t(𝐩1,𝐩2)\displaystyle t(\mathbf{p}_{1},\mathbf{p}_{2}) =l=0(2δl0)cos(lϕ12)tl(p1,p2),\displaystyle=\sum_{l=0}^{\infty}(2-\delta_{l0})\cos(l\phi_{12})t_{l}(p_{1},p_{2}), (72)

with ϕ12\phi_{12} the angle between 𝐩1\mathbf{p}_{1} and 𝐩2\mathbf{p}_{2}. This can be inverted to find

tl(p1,p2)\displaystyle t_{l}(p_{1},p_{2}) =02πdϕ122πcos(lϕ12)t(𝐩1,𝐩2).\displaystyle=\int_{0}^{2\pi}\frac{d\phi_{12}}{2\pi}\cos(l\phi_{12})t(\mathbf{p}_{1},\mathbf{p}_{2}). (73)

Similarly, we define the partial wave decomposition of the hole Green’s function

Gh(𝐩1+𝐩2,Eϵ𝐩1eϵ𝐩2e)\displaystyle G_{h}(\mathbf{p}_{1}+\mathbf{p}_{2},E-\epsilon^{e}_{\mathbf{p}_{1}}-\epsilon^{e}_{\mathbf{p}_{2}})
=l=0(2δl0)cos(lϕ12)gl(p1,p2),\displaystyle\qquad=\sum_{l=0}^{\infty}(2-\delta_{l0})\cos(l\phi_{12})g_{l}(p_{1},p_{2}), (74)

and

gl(p1,p2)\displaystyle g_{l}(p_{1},p_{2})
=02πdϕ122πcos(lϕ12)Gh(𝐩1+𝐩2,Eϵ𝐩1eϵ𝐩2e).\displaystyle\qquad=\int_{0}^{2\pi}\frac{d\phi_{12}}{2\pi}\cos(l\phi_{12})G_{h}(\mathbf{p}_{1}+\mathbf{p}_{2},E-\epsilon^{e}_{\mathbf{p}_{1}}-\epsilon^{e}_{\mathbf{p}_{2}}). (75)

Now let us define the angles of the vectors 𝐩1\mathbf{p}_{1}, 𝐩2\mathbf{p}_{2}, and 𝐪\mathbf{q} in Eq. (71) as ϕ1\phi_{1}, ϕ2\phi_{2}, ϕq\phi_{q}, respectively. Then we can take ϕϕq2=ϕqϕ2\phi\equiv\phi_{q2}=\phi_{q}-\phi_{2} in the integral, while the angle between 𝐩1\mathbf{p}_{1} and 𝐪\mathbf{q} is ϕ1qϕ1ϕq=ϕ12ϕq2\phi_{1q}\equiv\phi_{1}-\phi_{q}=\phi_{12}-\phi_{q2} with ϕ12=ϕ1ϕ2\phi_{12}=\phi_{1}-\phi_{2}. Inserting the expansions (72) and (74) in Eq. (71) and focussing only on the angular integral, we have

02πdϕq22π(2δlh0)(2δlt0)cos(lhϕ1q)cos(ltϕq2)\displaystyle\int_{0}^{2\pi}\frac{d\phi_{q2}}{2\pi}(2-\delta_{l_{h}0})(2-\delta_{l_{t}0})\cos(l_{h}\phi_{1q})\cos(l_{t}\phi_{q2})
=δlhlt(2δlt0)cos(ltϕ12),\displaystyle\quad=\delta_{l_{h}l_{t}}(2-\delta_{l_{t}0})\cos(l_{t}\phi_{12}), (76)

where lhl_{h} and ltl_{t} are the partial wave quantum numbers of the hole propagator and the polariton-electron TT matrix inside the integral, respectively. We see that this only depends on the angle ϕ12\phi_{12} which is the same that we have on the left hand side of Eq. (71). Therefore, applying the operator 02πdϕ122πcos(lϕ12)[]\int_{0}^{2\pi}\frac{d\phi_{12}}{2\pi}\cos(l\phi_{12})[\cdot] to the equation, we find that we must have lh=lt=ll_{h}=l_{t}=l, and we arrive at the equation

tl(p1,p2)=gl(p1,p2)\displaystyle t_{l}(p_{1},p_{2})=-g_{l}(p_{1},p_{2})
0qdq2πgl(p1,q)𝒯(q,Eϵqe)tl(q,p2).\displaystyle\qquad-\int_{0}^{\infty}\frac{q\,dq}{2\pi}g_{l}(p_{1},q)\mathcal{T}(q,E-\epsilon^{e}_{q})t_{l}(q,p_{2}). (77)

The polariton propagator in Eq. (77) has a simple pole at q=p2+i0q=p_{2}+i0, since 𝒯(q,Eϵqe)=𝒯(q,E(p2)+ϵp2eϵqe)\mathcal{T}(q,E-\epsilon^{e}_{q})=\mathcal{T}(q,E_{-}(p_{2})+\epsilon^{e}_{p_{2}}-\epsilon^{e}_{q}) by definition diverges as qp2q\to p_{2}. To circumvent this pole, we use the fact that

1qki0=𝒫qk+iπδ(qk),\displaystyle\frac{1}{q-k-i0}=\frac{\cal P}{q-k}+i\pi\delta(q-k), (78)

where 𝒫{\cal P} denotes the principal part. Using this in Eq. (77), we find

tl(p1,p2)=[1iπα(p2)tl(p2,p2)]gl(p1,p2)\displaystyle t_{l}(p_{1},p_{2})=-[1-i\pi\alpha(p_{2})t_{l}(p_{2},p_{2})]g_{l}(p_{1},p_{2})
𝒫0qdq2πgl(p1,q)𝒯(q,Eϵqe)tl(q,p2),\displaystyle\qquad-{\cal P}\int_{0}^{\infty}\frac{q\,dq}{2\pi}g_{l}(p_{1},q)\mathcal{T}(q,E-\epsilon^{e}_{q})t_{l}(q,p_{2}), (79)

where

α(p)\displaystyle\alpha(p) =limqpq(qp)2π𝒯(q,E(p)+ϵpeϵqe)\displaystyle=-\lim_{q\to p}\frac{q(q-p)}{2\pi}{\cal T}(q,E_{-}(p)+\epsilon_{p}^{e}-\epsilon_{q}^{e}) (80)
=p2π|X(p)|2Z(p)(E(p)+ϵpe)/p.\displaystyle=\frac{p}{2\pi}\frac{|X_{-}(p)|^{2}Z_{(}p)}{\partial(E_{-}(p)+\epsilon^{e}_{p})/\partial p}. (81)

The prefactor in the first term of Eq. (79) is independent of p1p_{1} and depends only on p2p_{2}. Since the multiplication by a constant prefactor in the inhomogenous term of an integral equation does not change its structure, we can instead solve the auxilliary equation

Kl(p1,p2)=gl(p1,p2)\displaystyle K_{l}(p_{1},p_{2})=-g_{l}(p_{1},p_{2})
𝒫0qdq2πgl(p1,q)𝒯(q,Eϵqe)Kl(q,p2),\displaystyle\qquad-{\cal P}\int_{0}^{\infty}\frac{q\,dq}{2\pi}g_{l}(p_{1},q)\mathcal{T}(q,E-\epsilon^{e}_{q})K_{l}(q,p_{2}), (82)

where the pole is explicitly excluded using the principal value prescription [44]. We then relate the on-shell unnormalized tt matrix to KK via

tl(k,k)=[1iπα(k)tl(k,k)]Kl(k,k),\displaystyle t_{l}(k,k)=[1-i\pi\alpha(k)t_{l}(k,k)]K_{l}(k,k), (83)

or, in other words,

tl(k,k)=1Kl1(k,k)+iπα(k).\displaystyle t_{l}(k,k)=\frac{1}{K_{l}^{-1}(k,k)+i\pi\alpha(k)}. (84)

We solve Eq. (82) using Gauss-Legendre quadrature.

In the case of scattering at low momentum, we can further simplify the problem, since the simple pole in this case disappears. We then set p2=0p_{2}=0 and p1=pp_{1}=p in Eq. (77), which can then be written as

t0(p,0)=1Ep2/2mr\displaystyle t_{0}(p,0)=-\frac{1}{E-p^{2}/2m_{r}}
+0qdq2π𝒯(q,Eq2/2me)t0(q,0)(Ep2/2mrq2/2mr)2(pq/mh)2.\displaystyle\quad+\int_{0}^{\infty}\frac{q\,dq}{2\pi}\frac{\mathcal{T}(q,E-q^{2}/2m_{e})t_{0}(q,0)}{\sqrt{(E-p^{2}/2m_{r}-q^{2}/2m_{r})^{2}-(p\,q/m_{h})^{2}}}. (85)

References

  • Deng et al. [2010] H. Deng, H. Haug,  and Y. Yamamoto, Exciton-polariton Bose-Einstein condensationRev. Mod. Phys. 82, 1489 (2010).
  • Carusotto and Ciuti [2013] I. Carusotto and C. Ciuti, Quantum fluids of lightRev. Mod. Phys. 85, 299 (2013).
  • Byrnes et al. [2014] T. Byrnes, N. Y. Kim,  and Y. Yamamoto, Exciton-polariton condensatesNat. Phys. 10, 803 (2014).
  • Amo et al. [2009] A. Amo, J. Lefrère, S. Pigeon, C. Adrados, C. Ciuti, I. Carusotto, R. Houdré, E. Giacobino,  and A. Bramati, Superfluidity of polaritons in semiconductor microcavitiesNature Physics 5, 805 (2009).
  • Sanvitto et al. [2010] D. Sanvitto, F. M. Marchetti, M. H. Szymańska, G. Tosi, M. Baudisch, F. P. Laussy, D. N. Krizhanovskii, M. S. Skolnick, L. Marrucci, A. Lemaître, J. Bloch, C. Tejedor,  and L. Viña, Persistent currents and quantized vortices in a polariton superfluidNature Physics 6, 527 (2010).
  • Amo et al. [2010] A. Amo, T. C. H. Liew, C. Adrados, R. Houdré, E. Giacobino, A. V. Kavokin,  and A. Bramati, Exciton-polariton spin switchesNature Photonics 4, 361 (2010).
  • Muñoz-Matutano et al. [2019] G. Muñoz-Matutano, A. Wood, M. Johnsson, X. Vidal, B. Q. Baragiola, A. Reinhard, A. Lemaître, J. Bloch, A. Amo, G. Nogues, B. Besga, M. Richard,  and T. Volz, Emergence of quantum correlations from interacting fibre-cavity polaritonsNature Materials 18, 213 (2019).
  • Delteil et al. [2019] A. Delteil, T. Fink, A. Schade, S. Höfling, C. Schneider,  and A. İmamoğlu, Towards polariton blockade of confined exciton–polaritonsNature Materials 18, 219 (2019).
  • Takemura et al. [2014] N. Takemura, S. Trebaol, M. Wouters, M. T. Portella-Oberli,  and B. Deveaud, Polaritonic Feshbach resonanceNature Physics 10, 500 (2014).
  • Tan et al. [2020] L. B. Tan, O. Cotlet, A. Bergschneider, R. Schmidt, P. Back, Y. Shimazaki, M. Kroner,  and A. İmamoğlu, Interacting Polaron-PolaritonsPhys. Rev. X 10, 021011 (2020).
  • Tassone and Yamamoto [1999] F. Tassone and Y. Yamamoto, Exciton-exciton scattering dynamics in a semiconductor microcavity and stimulated scattering into polaritonsPhys. Rev. B 59, 10830 (1999).
  • Ramon et al. [2002] G. Ramon, R. Rapaport, A. Qarry, E. Cohen, A. Mann, A. Ron,  and L. N. Pfeiffer, Scattering of polaritons by a two-dimensional electron gas in a semiconductor microcavityPhys. Rev. B 65, 085323 (2002).
  • Malpuech et al. [2002] G. Malpuech, A. Kavokin, A. Di Carlo,  and J. J. Baumberg, Polariton lasing by exciton-electron scattering in semiconductor microcavitiesPhys. Rev. B 65, 153310 (2002).
  • Hartwell and Snoke [2010] V. E. Hartwell and D. W. Snoke, Numerical simulations of the polariton kinetic energy distribution in GaAs quantum-well microcavity structuresPhys. Rev. B 82, 075307 (2010).
  • Shahnazaryan et al. [2017] V. Shahnazaryan, I. Iorsh, I. A. Shelykh,  and O. Kyriienko, Exciton-exciton interaction in transition-metal dichalcogenide monolayersPhys. Rev. B 96, 115409 (2017).
  • Rochat et al. [2000] G. Rochat, C. Ciuti, V. Savona, C. Piermarocchi, A. Quattropani,  and P. Schwendimann, Excitonic Bloch equations for a two-dimensional system of interacting excitonsPhys. Rev. B 61, 13856 (2000).
  • Combescot et al. [2007] M. Combescot, M. A. Dupertuis,  and O. Betbeder-Matibet, Polariton-polariton scattering: Exact results through a novel approachEurophysics Letters (EPL) 79, 17001 (2007).
  • Combescot et al. [2008] M. Combescot, O. Betbeder-Matibet,  and F. Dubin, The many-body physics of composite bosonsPhysics Reports 463, 215 (2008).
  • Glazov et al. [2009] M. M. Glazov, H. Ouerdane, L. Pilozzi, G. Malpuech, A. V. Kavokin,  and A. D’Andrea, Polariton-polariton scattering in microcavities: A microscopic theoryPhys. Rev. B 80, 155306 (2009).
  • Li et al. [2020] G. Li, O. Bleu, M. M. Parish,  and J. Levinsen, Enhanced scattering between electrons and exciton-polaritons in a microcavity (2020), arXiv:2008.09281 .
  • Yamaguchi et al. [2012] M. Yamaguchi, K. Kamide, T. Ogawa,  and Y. Yamamoto, BEC–BCS-laser crossover in Coulomb-correlated electron–hole–photon systemsNew Journal of Physics 14, 065001 (2012).
  • Hanai et al. [2017] R. Hanai, P. B. Littlewood,  and Y. Ohashi, Dynamical instability of a driven-dissipative electron-hole condensate in the BCS-BEC crossover regionPhys. Rev. B 96, 125206 (2017).
  • Hu and Liu [2020] H. Hu and X.-J. Liu, Quantum fluctuations in a strongly interacting Bardeen-Cooper-Schrieffer polariton condensate at thermal equilibriumPhys. Rev. A 101, 011602 (2020).
  • Levinsen et al. [2019a] J. Levinsen, G. Li,  and M. M. Parish, Microscopic description of exciton-polaritons in microcavitiesPhys. Rev. Research 1, 033120 (2019a).
  • Adhikari [1986] S. K. Adhikari, Quantum scattering in two dimensionsAmerican Journal of Physics 54, 362 (1986).
  • Kasprzak et al. [2006] J. Kasprzak, M. Richard, S. Kundermann, A. Baas, P. Jeambrun, J. M. J. Keeling, F. M. Marchetti, M. H. Szymańska, R. André, J. L. Staehli, V. Savona, P. B. Littlewood, B. Deveaud,  and L. S. Dang, Bose–Einstein condensation of exciton polaritonsNature 443, 409 (2006).
  • Landau and Lifshitz [2013] L. D. Landau and E. M. Lifshitz, Quantum mechanics: non-relativistic theory, Vol. 3 (Elsevier, 2013).
  • Levinsen and Parish [2015] J. Levinsen and M. M. Parish, Strongly interacting two-dimensional Fermi gasesAnnu. Rev. Cold Atoms Mol. 3, 1 (2015).
  • Note [1] If ΩεB\Omega\gtrsim\varepsilon_{B} and/or δεB\delta\gtrsim\varepsilon_{B}, the upper polariton enters the continuum of unbound electron-hole states. In that case, one must analytically continue the energy slightly into the complex plane, EE+i0E\to E+i0.
  • Note [2] The variational approach developed by Khurgin [76] approximated the functional form of the electron-hole wave function as unchanged in the presence of light-matter coupling. While this is an approximation in the Coulomb case studied in that work [24], we see that it is exact for strongly screened electron-hole interactions.
  • Hopfield [1958] J. J. Hopfield, Theory of the Contribution of Excitons to the Complex Dielectric Constant of CrystalsPhys. Rev. 112, 1555 (1958).
  • Werner et al. [2009] F. Werner, L. Tarruell,  and Y. Castin, Number of closed-channel molecules in the BEC-BCS crossoverThe European Physical Journal B 68, 401 (2009).
  • Fetter and Walecka [2003] A. L. Fetter and J. D. Walecka, Quantum Theory of Many-Particle Systems (Dover, New York, USA, 2003).
  • Wouters [2007] M. Wouters, Resonant polariton-polariton scattering in semiconductor microcavitiesPhys. Rev. B 76, 045319 (2007).
  • Bleu et al. [2020] O. Bleu, G. Li, J. Levinsen,  and M. M. Parish, Polariton interactions in microcavities with atomically thin semiconductor layersPhys. Rev. Research 2, 043185 (2020).
  • Sze [2007] S. M. Sze, Physics of semiconductor devices, 3rd ed. (Hoboken, NJ : Wiley-Interscience, Hoboken, NJ Hoboken, N.J., 2007).
  • Ramon et al. [2003] G. Ramon, A. Mann,  and E. Cohen, Theory of neutral and charged exciton scattering with electrons in semiconductor quantum wellsPhys. Rev. B 67, 045323 (2003).
  • Ngampruetikorn et al. [2013a] V. Ngampruetikorn, M. M. Parish,  and J. Levinsen, Three-body problem in a two-dimensional Fermi gasEPL (Europhysics Letters) 102, 13001 (2013a).
  • Skorniakov and Ter-Martirosian [1957] G. V. Skorniakov and K. A. Ter-Martirosian, Three Body Problem for Short Range Forces. I. Scattering of Low Energy Neutrons by DeuteronsSov. Phys. JETP 4, 648 (1957).
  • Petrov [2003] D. S. Petrov, Three-body problem in Fermi gases with short-range interparticle interactionPhys. Rev. A 67, 010703 (2003).
  • Brodsky et al. [2006] I. V. Brodsky, M. Y. Kagan, A. V. Klaptsov, R. Combescot,  and X. Leyronas, Exact diagrammatic approach for dimer-dimer scattering and bound states of three and four resonantly interacting particlesPhys. Rev. A 73, 032724 (2006).
  • Levinsen and Gurarie [2006] J. Levinsen and V. Gurarie, Properties of strongly paired fermionic condensatesPhys. Rev. A 73, 053607 (2006).
  • [43] D. S. Petrov, in Proceedings of the Les Houches Summer Schools, Session 94, edited by C. Salomon, G. V. Shlyapnikov, and L. F. Cugliandolo (Oxford University Press, Oxford, England, 2013), arXiv:1206.5752 .
  • Bedaque et al. [1998] P. F. Bedaque, H.-W. Hammer,  and U. van Kolck, Effective theory for neutron-deuteron scattering: Energy dependencePhys. Rev. C 58, R641 (1998).
  • Levinsen et al. [2009] J. Levinsen, T. G. Tiecke, J. T. M. Walraven,  and D. S. Petrov, Atom-Dimer Scattering and Long-Lived Trimers in Fermionic MixturesPhys. Rev. Lett. 103, 153202 (2009).
  • Levinsen and Petrov [2011] J. Levinsen and D. S. Petrov, Atom-dimer and dimer-dimer scattering in fermionic mixtures near a narrow Feshbach resonanceThe European Physical Journal D 65, 67 (2011).
  • Ngampruetikorn et al. [2013b] V. Ngampruetikorn, M. M. Parish,  and J. Levinsen, Three-body problem in a two-dimensional Fermi gasEPL (Europhysics Letters) 102, 13001 (2013b).
  • Jag et al. [2014] M. Jag, M. Zaccanti, M. Cetina, R. S. Lous, F. Schreck, R. Grimm, D. S. Petrov,  and J. Levinsen, Observation of a Strong Atom-Dimer Attraction in a Mass-Imbalanced Fermi-Fermi MixturePhys. Rev. Lett. 112, 075302 (2014).
  • Note [3] Note that the electron propagator Ge(𝐪,q0)=1/(q0ϵ𝐪+i0)G_{e}(-\mathbf{q},-q_{0})=1/(-q_{0}-\epsilon_{\mathbf{q}}+i0) in the last term in Fig. 5 has a simple pole in the upper half of the complex q0q_{0} plane, while all other propagators have their non-analytic structure only in the lower half plane. This allows us to perform the integral over q0q_{0} by closing the contour in the upper half plane, effectively removing that propagator and setting the electron dispersion to its on-shell value ϵ𝐪e=𝐪2/(2me)\epsilon^{e}_{\mathbf{q}}=\mathbf{q}^{2}/(2m_{e}) in the other propagators. For more details, see, e.g., Ref. [41].
  • Brodbeck et al. [2017] S. Brodbeck, S. De Liberato, M. Amthor, M. Klaas, M. Kamp, L. Worschech, C. Schneider,  and S. Höfling, Experimental Verification of the Very Strong Coupling Regime in a GaAs Quantum Well MicrocavityPhys. Rev. Lett. 119, 027401 (2017).
  • [51] Reference [47] determined the scattering phase shift of atoms and diatomic molecules in a two-component Fermi gas with equal mass fermions, and in particular identified the low-energy energy scale εeX1/[2meX(1.27aX)2]\varepsilon_{eX}\simeq 1/[2m_{eX}(1.27a_{X})^{2}]. In the absence of coupling to light, this scenario is equivalent to exciton-electron scattering given our approximation of contact interactions. Likewise, Ref. [64] determined the bound three-body states in the same problem.
  • Dufferwiel et al. [2015] S. Dufferwiel, S. Schwarz, F. Withers, A. A. P. Trichet, F. Li, M. Sich, O. Del Pozo-Zamudio, C. Clark, A. Nalitov, D. D. Solnyshkov, G. Malpuech, K. S. Novoselov, J. M. Smith, M. S. Skolnick, D. N. Krizhanovskii,  and A. I. Tartakovskii, Exciton-polaritons in van der Waals heterostructures embedded in tunable microcavitiesNature Communications 6, 8579 (2015).
  • Liu et al. [2014] X. Liu, T. Galfsky, Z. Sun, F. Xia, E.-c. Lin, Y.-H. Lee, S. Kéna-Cohen,  and V. M. Menon, Strong light–matter coupling in two-dimensional atomic crystalsNature Photonics 9, 30 (2014).
  • Lundt et al. [2016] N. Lundt, S. Klembt, E. Cherotchenko, S. Betzold, O. Iff, A. V. Nalitov, M. Klaas, C. P. Dietrich, A. V. Kavokin, S. Höfling,  and C. Schneider, Room-temperature Tamm-plasmon exciton-polaritons with a WSe2 monolayerNature Communications 7, 13328 (2016).
  • He et al. [2014] K. He, N. Kumar, L. Zhao, Z. Wang, K. F. Mak, H. Zhao,  and J. Shan, Tightly Bound Excitons in Monolayer WSe2{\mathrm{WSe}}_{2}Phys. Rev. Lett. 113, 026803 (2014).
  • Flatten et al. [2016] L. C. Flatten, Z. He, D. M. Coles, A. A. P. Trichet, A. W. Powell, R. A. Taylor, J. H. Warner,  and J. M. Smith, Room-temperature exciton-polaritons with two-dimensional WS2Scientific Reports 6, 33134 (2016).
  • Estrecho et al. [2019] E. Estrecho, T. Gao, N. Bobrovska, D. Comber-Todd, M. D. Fraser, M. Steger, K. West, L. N. Pfeiffer, J. Levinsen, M. M. Parish, T. C. H. Liew, M. Matuszewski, D. W. Snoke, A. G. Truscott,  and E. A. Ostrovskaya, Direct measurement of polariton-polariton interaction strength in the Thomas-Fermi regime of exciton-polariton condensationPhys. Rev. B 100, 035306 (2019).
  • Note [4] In general, there are finite range corrections to Eq. (53\@@italiccorr), which add a term meXre2(E+εB)\sim m_{eX}r_{e}^{2}(E+\varepsilon_{B}) in the denominator [28]. In the absence of other scales, the effective range reaXr_{e}\sim a_{X} and therefore such terms are not important when |E+εB|εeX|E+\varepsilon_{B}|\ll\varepsilon_{eX}.
  • Keldysh [1979] L. V. Keldysh, Coulomb interaction in thin semiconductor and semimetal films, Soviet Journal of Experimental and Theoretical Physics Letters 29, 658 (1979).
  • Fey et al. [2020] C. Fey, P. Schmelcher, A. Imamoglu,  and R. Schmidt, Theory of exciton-electron scattering in atomically thin semiconductorsPhys. Rev. B 101, 195417 (2020).
  • Sergeev and Suris [2001] R. Sergeev and R. Suris, The Triplet State of X+ Trion in 2D Quantum Wellsphysica status solidi (b) 227, 387 (2001).
  • Ganchev et al. [2015] B. Ganchev, N. Drummond, I. Aleiner,  and V. Fal’ko, Three-Particle Complexes in Two-Dimensional SemiconductorsPhys. Rev. Lett. 114, 107401 (2015).
  • Courtade et al. [2017] E. Courtade, M. Semina, M. Manca, M. M. Glazov, C. Robert, F. Cadiz, G. Wang, T. Taniguchi, K. Watanabe, M. Pierre, W. Escoffier, E. L. Ivchenko, P. Renucci, X. Marie, T. Amand,  and B. Urbaszek, Charged excitons in monolayer WSe2{\mathrm{WSe}}_{2}: Experiment and theoryPhys. Rev. B 96, 085302 (2017).
  • Pricoupenko and Pedri [2010] L. Pricoupenko and P. Pedri, Universal (1+21+2)-body bound states in planar atomic waveguidesPhys. Rev. A 82, 033625 (2010).
  • Berman et al. [2010] O. L. Berman, R. Y. Kezerashvili,  and Y. E. Lozovik, Drag effects in a system of electrons and microcavity polaritonsPhys. Rev. B 82, 125307 (2010).
  • Laussy et al. [2010] F. P. Laussy, A. V. Kavokin,  and I. A. Shelykh, Exciton-Polariton Mediated SuperconductivityPhys. Rev. Lett. 104, 106402 (2010).
  • Cotleţ et al. [2016] O. Cotleţ, S. Zeytinoǧlu, M. Sigrist, E. Demler,  and A. Imamoǧlu, Superconductivity and other collective phenomena in a hybrid Bose-Fermi mixture formed by a polariton condensate and an electron system in two dimensionsPhys. Rev. B 93, 054510 (2016).
  • Sidler et al. [2017] M. Sidler, P. Back, O. Cotlet, A. Srivastava, T. Fink, M. Kroner, E. Demler,  and A. Imamoglu, Fermi polaron-polaritons in charge-tunable atomically thin semiconductorsNature Physics 13, 255 (2017).
  • Chervy et al. [2020] T. Chervy, P. Knüppel, H. Abbaspour, M. Lupatini, S. Fält, W. Wegscheider, M. Kroner,  and A. Imamoǧlu, Accelerating Polaritons with External Electric and Magnetic FieldsPhys. Rev. X 10, 011040 (2020).
  • Petrov et al. [2003] D. S. Petrov, M. A. Baranov,  and G. V. Shlyapnikov, Superfluid transition in quasi-two-dimensional Fermi gasesPhys. Rev. A 67, 031601 (2003).
  • Petrov et al. [2005] D. S. Petrov, C. Salomon,  and G. V. Shlyapnikov, Diatomic molecules in ultracold Fermi gases—novel composite bosonsJournal of Physics B: Atomic, Molecular and Optical Physics 38, S645 (2005).
  • Takemura et al. [2017] N. Takemura, M. D. Anderson, M. Navadeh-Toupchi, D. Y. Oberli, M. T. Portella-Oberli,  and B. Deveaud, Spin anisotropic interactions of lower polaritons in the vicinity of polaritonic Feshbach resonancePhys. Rev. B 95, 205303 (2017).
  • Navadeh-Toupchi et al. [2019] M. Navadeh-Toupchi, N. Takemura, M. D. Anderson, D. Y. Oberli,  and M. T. Portella-Oberli, Polaritonic Cross Feshbach ResonancePhys. Rev. Lett. 122, 047402 (2019).
  • Levinsen et al. [2019b] J. Levinsen, F. M. Marchetti, J. Keeling,  and M. M. Parish, Spectroscopic Signatures of Quantum Many-Body Correlations in Polariton MicrocavitiesPhys. Rev. Lett. 123, 266401 (2019b).
  • Bastarrachea-Magnani et al. [2019] M. A. Bastarrachea-Magnani, A. Camacho-Guardian, M. Wouters,  and G. M. Bruun, Strong interactions and biexcitons in a polariton mixturePhys. Rev. B 100, 195301 (2019).
  • Khurgin [2001] J. Khurgin, Excitonic radius in the cavity polariton in the regime of very strong couplingSolid State Communications 117, 307 (2001).