Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Nonlinear spectroscopy of semiconductor moiré materials

B. Evrard Institute for Quantum Electronics, ETH Zürich, CH-8093 Zürich, Switzerland    H.S. Adlong Institute for Quantum Electronics, ETH Zürich, CH-8093 Zürich, Switzerland Institute for Theoretical Physics, ETH Zürich, Zürich, Switzerland    A.A. Ghita Institute for Quantum Electronics, ETH Zürich, CH-8093 Zürich, Switzerland    T. Uto Institute for Quantum Electronics, ETH Zürich, CH-8093 Zürich, Switzerland    L. Ciorciaro Institute for Quantum Electronics, ETH Zürich, CH-8093 Zürich, Switzerland    K. Watanabe Research Center for Electronic and Optical Materials, NIMS, 1-1 Namiki, Tsukuba 305-0044, Japan    T. Taniguchi Research Center for Electronic and Optical Materials, NIMS, 1-1 Namiki, Tsukuba 305-0044, Japan    M. Kroner Institute for Quantum Electronics, ETH Zürich, CH-8093 Zürich, Switzerland    A. İmamoğlu Institute for Quantum Electronics, ETH Zürich, CH-8093 Zürich, Switzerland
Abstract

We use time-resolved nonlinear pump–probe measurements to reveal features of semiconductor moiré materials not accessible to linear spectroscopy. With an intense, red-detuned pump pulse, we generate a high density of virtual excitons or exciton–polarons in various moiré minibands. A broadband probe pulse in turn measures the response of all optical resonances induced by the pump-generated excitations. We generically observe a coherent blue shift originating from contact-like exciton–exciton interactions. At charge neutrality, these measurements allow us to assess the spatial overlap between different optical excitations and to observe signatures of a bound biexciton state between two different moiré exciton modes. In contrast to electron doped monolayers, spatially confined moiré attractive polarons behave as an ensemble of non-interacting two-level emitters, exhibiting an electron-density-independent ac-Stark effect. Tuning the pump laser into resonance with the attractive polaron, we demonstrate the filling of the moiré lattice with localized polarons and thereby realize a nonequilibrium Bose–Fermi mixture in moiré flat bands.

I Introduction

Semiconductor moiré materials have emerged as a rich playground for exploration of strongly correlated electrons [1, 2, 3]. Recent experiments based on twisted bilayers of transition metal dichalcogenides (TMDs) have enabled the observation of a wealth of many-body states, ranging from correlated Mott–Wigner states [4, 5, 6], through kinetic magnetism [7] to fractional Chern insulators [8, 9]. These discoveries were made using various experimental techniques, such as transport, STM spectroscopy, or linear optical spectroscopy [10]. The link between the optical measurements and the electronic state stems from the dynamical dressing of tightly bound TMD excitons by itinerant charges, leading to the formation of attractive and repulsive exciton–polarons  [11, 12, 13, 14, 15, 16]. Since exciton–polarons are sensitive to both the charge and magnetic order of the electron system, they provide unequivocal signatures of correlated electrons  [17, 18, 7]. A common feature of all the measurement techniques implemented to date is that they exclusively measured linear response of electrons to external fields.

In this work, we use nonlinear optical pump–probe measurements to investigate features of semiconductor moiré materials not accessible through linear spectroscopy. In particular, we investigate their optical excitations, which emerge from a complex interplay between the moiré potential and the interaction bewteen excitons and a many-body electronic state in a flat band. Our approach relies on an intense red-detuned pump beam to generate a significant density of virtual excitations, and on a weak broadband probe pulse to monitor the subsequent modification of the optical spectrum. By detuning the pump away from any resonances, we alleviate real-absorption-induced modification of the electronic state and look at the coherent scattering response of the system. When the heterobilayer is charge neutral, we observe a blue shift of moiré exciton resonances in co-circularly-polarized pump-probe measurements [19, 20, 21, 22, 23, 24, 25, 26]: these measurements reveal the extent of spatial overlap between distinct exciton modes. We find further confirmation of this overlap under cross-circularly-polarized excitations [27, 28, 29, 30, 31] where we identify biexciton Feshbach resonances associated with bound states of both the same and different moiré exciton modes [32]. To further study the nature of these biexciton states, we develop a theoretical model for the scattering of excitons in a moiré and find good qualitative agreement with the experiment. When the moiré material is electron doped, nonlinear pump-probe measurements reveal an electron-density-independent blue shift of the attractive polaron resonance that scales linearly with the inverse pump-laser-detuning: such a single-emitter-like ac-Stark shift [33, 34, 35] demonstrates that the attractive polaron resonance should be considered as an ensemble of spatially localized and non-overlapping independent trion excitations.

Refer to caption
Figure 1: a Schematic of the device. The TMDs MoSe2 and WS2 form the moiré material that we investigate. It is encapsulated in two 35absent35\approx 35\,≈ 35nm thick hBN flakes, and dual gated with graphite electrodes to enable an independent control of the chemical potential and electric field. The latter is kept close to zero, such that only MoSe2 is doped at low densities, given the band alignment (b). The evolution of the reflection spectrum of MoSe2 as a function of the electronic density (c) reveals several moiré exciton and polaron resonances, which we investigate. We rely on a pump–probe scheme (d) where an intense red-detuned laser generates a virtual population of moiré excitations. The interaction between this background and a test excitation generated with a probe laser are then measured.

Our measurements are carried out in a 0similar-to-or-equalsabsentsuperscript0\simeq 0^{\circ}≃ 0 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT-degree stacked MoSe2/WS2 heterostructure, exhibiting a Type I band alignment where the lowest (highest) energy moiré conduction (valence) band resides in MoSe2 (Fig. 1 ab). The electron density dependent reflection contrast exhibits four bright resonances, which we identify as MX1, MX2, AP and MX3 to be consistent with the notation used in an earlier publication [7] (Fig. 1 c). A strong pump laser with a finite red detuning from a given resonance generates a large virtual population of the corresponding moiré exciton species/modes that exists only during the 0.2absent0.2\approx 0.2≈ 0.2 ps duration of the pump pulse. We estimate a maximum virtual exciton population in MX1 of 1012absentsuperscript1012\approx 10^{12}≈ 10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT cm-2 for a pump detuning of δ120subscript𝛿120\delta_{1}\approx 20italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 20 meV (for more details, see the Appendix). A weak, broadband probe pulse then measures the energy shift of all excitonic species concurrently (Fig. 1 d). Throughout this work, we set the pump laser detuning from the excitonic resonances to be much smaller than the exciton binding energy: In this limit, the dominant contribution to the light shift of the resonances originates from exciton–exciton interactions [19, 20, 21, 22, 23, 24, 25]. We determine both intra- and inter-species interaction strengths between the moiré exciton or polaron modes as function of the electron density by measuring the light shift in this small detuning limit.

II Interactions between moiré excitons

We begin by performing non-linear spectroscopy at charge neutrality in order to explore exciton–exciton interactions in the presence of a moiré. When the material is free of itinerant electrons, two bright moiré exciton resonances MX1 and MX2 are visible in the normalized reflection spectrum of MoSe2, ΔR=(RRbg)/RbgΔ𝑅𝑅subscript𝑅bgsubscript𝑅bg\Delta R=(R-R_{\mathrm{bg}})/R_{\mathrm{bg}}roman_Δ italic_R = ( italic_R - italic_R start_POSTSUBSCRIPT roman_bg end_POSTSUBSCRIPT ) / italic_R start_POSTSUBSCRIPT roman_bg end_POSTSUBSCRIPT with R𝑅Ritalic_R and Rbgsubscript𝑅bgR_{\mathrm{bg}}italic_R start_POSTSUBSCRIPT roman_bg end_POSTSUBSCRIPT the reflection spectrum from the sample and background, respectively (see Fig. 1c). To explore the interactions involving these two resonances we consider two different scenarios: co- and cross-circular polarized light of the pump and probe beams. These scenarios allow us to either probe interactions between excitons in the same valley (co-circular) or in opposite valleys (cross-circular).

II.1 Co-circularly polarized light

Refer to caption
Figure 2: Light shift of moiré excitons for co-circularly polarized pump and probe lasers. (a) Schematic of the moiré exciton energy level at charge neutrality. (b) Reflection spectrum as a function of the pump–probe delay, showing a clear blue shift of the lowest and brightest resonance MX1 at zero time delay. The differential spectrum (c), obtained by subtracting from (b) a reference spectrum (acquired at τ<2.5𝜏2.5\tau<-2.5\,italic_τ < - 2.5ps), reveals a blue shift of the upper and darker resonance MX2 as well. (d) The dependence of the blue shift of MX1 (blue circles) on the pump detuning scales as 1/δ12n1proportional-to1superscriptsubscript𝛿12subscript𝑛11/\delta_{1}^{2}\propto n_{1}1 / italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∝ italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (blue line), and stems from MX1–MX1 interactions. The shift of MX2 (red squares), is well fitted either by a 1/δ12proportional-toabsent1superscriptsubscript𝛿12\propto 1/\delta_{1}^{2}∝ 1 / italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT dependence (dashed red line) or by n1n21/(δ1δ2)proportional-tosubscript𝑛1subscript𝑛21subscript𝛿1subscript𝛿2\sqrt{n_{1}n_{2}}\propto 1/(\delta_{1}\delta_{2})square-root start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ∝ 1 / ( italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) (dotted red line). While we are not able to precisely deconvolve the role of these two potential contributions, both imply significant MX1-MX2 interaction and hence spatial overlap between these two bright moiré excitons.

Beginning with the scenario of co-circular polarization, we consider the evolution of ΔRΔ𝑅\Delta Rroman_Δ italic_R as a function of the delay τ=tprobetpump𝜏subscript𝑡probesubscript𝑡pump\tau=t_{\rm probe}-t_{\rm pump}italic_τ = italic_t start_POSTSUBSCRIPT roman_probe end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT roman_pump end_POSTSUBSCRIPT, between the pump and probe pulse, as shown in Fig. 2 b (here the pump pulse is red-detuned from the MX1 resonance by δ1=20subscript𝛿120\delta_{1}=20italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 20 meV). The dominant coupling mechanism between two same-valley excitons are electron- and hole-exchange interactions [36, 37], which leads to short-range repulsion. Correspondingly, we observe a clear blue shift of the brightest exciton MX1 for |τ|0.2𝜏0.2|\tau|\lessapprox 0.2\,| italic_τ | ⪅ 0.2ps, that is, when the two pulses overlap in time (for more detail on the experimental setup see the Appendix). To better assess the pump-induced modifications of weaker resonances, we use the differential reflection spectrum ΔR(t)ΔRrefΔ𝑅𝑡Δsubscript𝑅ref\Delta R(t)-\Delta R_{\rm ref}roman_Δ italic_R ( italic_t ) - roman_Δ italic_R start_POSTSUBSCRIPT roman_ref end_POSTSUBSCRIPT, where ΔRrefΔsubscript𝑅ref\Delta R_{\rm ref}roman_Δ italic_R start_POSTSUBSCRIPT roman_ref end_POSTSUBSCRIPT is a reference spectrum obtained when the probe pulse hits the sample significantly before the pump (τ2.5𝜏2.5\tau\lessapprox-2.5\,italic_τ ⪅ - 2.5ps): Figure 2 c shows that a smaller blue shift of MX2 is discernible in ΔR(t)ΔRrefΔ𝑅𝑡Δsubscript𝑅ref\Delta R(t)-\Delta R_{\rm ref}roman_Δ italic_R ( italic_t ) - roman_Δ italic_R start_POSTSUBSCRIPT roman_ref end_POSTSUBSCRIPT. In a mean-field picture (detailed in App. H.6), the light shift ΔisubscriptΔ𝑖\Delta_{i}roman_Δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of MXi (i=1,2𝑖12i=1,2italic_i = 1 , 2) can be expressed as

Δ1=2u11n1+2u12n2+4k1n1n2,subscriptΔ12subscript𝑢11subscript𝑛12subscript𝑢12subscript𝑛24subscript𝑘1subscript𝑛1subscript𝑛2\displaystyle\Delta_{1}=2u_{11}n_{1}+2u_{12}n_{2}+4k_{1}\sqrt{n_{1}n_{2}}\,,roman_Δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 2 italic_u start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 2 italic_u start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + 4 italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT square-root start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG , (1a)
Δ2=2u22n2+2u12n1+4k2n1n2,subscriptΔ22subscript𝑢22subscript𝑛22subscript𝑢12subscript𝑛14subscript𝑘2subscript𝑛1subscript𝑛2\displaystyle\Delta_{2}=2u_{22}n_{2}+2u_{12}n_{1}+4k_{2}\sqrt{n_{1}n_{2}}\,,roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 2 italic_u start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + 2 italic_u start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 4 italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT square-root start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG , (1b)

where nisubscript𝑛𝑖n_{i}italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the density of MXi excitons and we have introduced four different interaction terms: ui,jsubscript𝑢𝑖𝑗u_{i,j}italic_u start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT correspond to MXi+MXj{}_{j}\leftrightharpoonsstart_FLOATSUBSCRIPT italic_j end_FLOATSUBSCRIPT ⇋ MXi+MXj, while kisubscript𝑘𝑖k_{i}italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT correspond to MX1+MX2{}_{2}\leftrightharpoonsstart_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT ⇋ MXi+MXi. The density scales as ni1/δi2proportional-tosubscript𝑛𝑖1superscriptsubscript𝛿𝑖2n_{i}\propto 1/\delta_{i}^{2}italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∝ 1 / italic_δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT with the detuning δisubscript𝛿𝑖\delta_{i}italic_δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of the pump from the MXi resonance. Therefore, by tuning the pump laser frequency, one can change the density imbalance between the two excitons, and in principle deconvolve the contribution of each scattering processes to the light shift. In practice, we observe signatures of an incoherent response for a blue-detuned pump, so we focus exclusively on red detunings. We consequently always have δ1<δ2subscript𝛿1subscript𝛿2\delta_{1}<\delta_{2}italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and thus n1>n2subscript𝑛1subscript𝑛2n_{1}>n_{2}italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (see Fig. 2 a). This imbalance is further amplified by the oscillator strength difference between MX1 and MX2.

Figure 2 d shows the light shifts Δ1,2subscriptΔ12\Delta_{1,2}roman_Δ start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT as a function of δ1subscript𝛿1\delta_{1}italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT in a range where n110n2greater-than-or-equivalent-tosubscript𝑛110subscript𝑛2n_{1}\gtrsim 10\,n_{2}italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≳ 10 italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. As a result, the light shift of MX1 is dominated by MX1–MX1 interactions. From a fit we obtain the interaction strength u1,1subscript𝑢11u_{1,1}italic_u start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT, which we find to be larger by a factor u1,1/uex1.6±0.2subscript𝑢11subscript𝑢explus-or-minus1.60.2u_{1,1}/u_{\rm ex}\approx 1.6\pm 0.2italic_u start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT / italic_u start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT ≈ 1.6 ± 0.2 than the moiré-free exciton–exciton interaction strength uexsubscript𝑢exu_{\rm ex}italic_u start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT, measured on a monolayer MoSe2 region of the same device. We tentatively attribute this enhancement to the reduced spatial extent of the MX1 center-of-mass wavefunction within the moiré unit cell, which in turn increases the overlap between MX1 excitons for a given average density.

Remarkably and despite of the large imbalance n2n1much-less-thansubscript𝑛2subscript𝑛1n_{2}\ll n_{1}italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≪ italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, we observe a substantial light shift of MX2, strongly increasing as the pump wavelength approaches the MX1 resonance. This shift Δ2subscriptΔ2\Delta_{2}roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is not well reproduced by a 1/δ221superscriptsubscript𝛿221/\delta_{2}^{2}1 / italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT dependence, and attests to substantial inter-species interactions. A fit to our data does not enable us to precisely disentangle the respective contributions of the MX2+MX1{}_{1}\leftrightharpoonsstart_FLOATSUBSCRIPT 1 end_FLOATSUBSCRIPT ⇋ MX2+MX1 and MX2+MX1{}_{1}\leftrightharpoonsstart_FLOATSUBSCRIPT 1 end_FLOATSUBSCRIPT ⇋ 2MX1 processes. Nevertheless, we estimate that 0.3u1,2/u1,10.8less-than-or-similar-to0.3subscript𝑢12subscript𝑢11less-than-or-similar-to0.80.3\lesssim u_{1,2}/u_{1,1}\lesssim 0.80.3 ≲ italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT / italic_u start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT ≲ 0.8 while we are not able to reliably estimate the kisubscript𝑘𝑖k_{i}italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT parameters (see App. C for more details on the fitting procedure).

These observations suggest a significant spatial overlap of MX1 and MX2 excitons and invalidates a simplistic picture of moiré exciton modes that are tightly confined around different high-symmetry points of the moiré potential. Moreover, our experiments are in reasonable qualitative agreement with the findings of moiré exciton wavefunctions obtained for the same structure using the continuum model [38]. In particular, we find that the continuum model predicts u1,1/uex=2.2subscript𝑢11subscript𝑢ex2.2u_{1,1}/u_{\rm ex}=2.2italic_u start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT / italic_u start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT = 2.2 and the ratio of interactions to be u1,2/u1,1=0.6subscript𝑢12subscript𝑢110.6u_{1,2}/u_{1,1}=0.6italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT / italic_u start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT = 0.6 (see Appendix H for details).

We point out that Eq. (1) can be understood as arising from first order perturbation theory. This approach is valid in the limit for small enough pump power/large enough pump detuning. We find an empirical confirmation that we are indeed working in that regime by observing a linear dependence of the light shifts ΔisubscriptΔ𝑖\Delta_{i}roman_Δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT with the pump intensity (App. B). Nevertheless, we can envision interesting higher order effects arising from the mixing of the excitonic states. In particular, mixing of optically bright and dark excitons, could be observed as the emergence of new resonances in the reflection spectrum.

Refer to caption
Figure 3: Coupling to biexcitonic states. (a) Reflection spectrum as a function of the pump–probe delay when the pump is red-detuned by 222222\,22meV from MX1 and the pump and probe laser are cross-circularly polarized. The differential spectrum (b), obtained by subtracting from (a) a reference spectrum (acquired at t<2.5𝑡2.5t<-2.5\,italic_t < - 2.5ps), reveals a blue shift of both resonances MX1 and MX2. A similar measurement carried out for a detuning of 454545\,45meV from MX1 shows instead a red shift of both resonances (cd). The dependence of the light shifts of MX1 (blue dots) and MX2 (red dots) displays a sign change for a pump detuning δ136subscript𝛿136\delta_{1}\approx 36\,italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 36meV and δ128subscript𝛿128\delta_{1}\approx 28\,italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 28meV, respectively (e). (f) The corresponding theory simulation for the light-shifts of MX1 (blue) and MX2 (red) based on DFT moiré parameters. (g) The calculated spectrum which is shown in three different energy sectors with zero (defined as zero energy), one and two excitons. In the one exciton sector there are the optically bright MX1 and MX2 excitons. In the two exciton sector (focusing on zero centre-of-mass quasi-momentum) there exists a band of MX1-MX1 excitons (light blue) as well as other bands (gray), and bound biexcitons (green lines). We find that the sign change in the light shift of MX1 and MX2 in (f) can be attributed to the blue and red arrows (indicating exciton–biexciton transitions), respectively.

II.2 Cross-circularly polarized light

Since electron and hole exchange interactions are suppressed for excitons generated in opposite valleys, one may naively assume that the cross-circularly polarized scenario should yield a significantly smaller light shift. However, the bare interaction of opposite valley excitons is attractive, and supports a bound biexciton state, which has significant implications.

In order to understand how the biexciton state affects these interactions, we briefly review the simpler scenario in which there is no moiré (such as in the case of monolayer TMDs). In this case it is known that the presence of the biexciton resonance leads to an additional contribution to the light shift, scaling linearly with the inverse of the two-photon detuning δb1=(Eb+δex)1superscriptsubscript𝛿b1superscriptsubscript𝐸bsubscript𝛿ex1\delta_{\rm b}^{-1}=(-E_{\rm b}+\delta_{\rm ex})^{-1}italic_δ start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = ( - italic_E start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT + italic_δ start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, where δexsubscript𝛿ex\delta_{\rm ex}italic_δ start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT is the detuning of the pump from the exciton resonance, Ebsubscript𝐸𝑏E_{b}italic_E start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is the biexciton binding energy and the probe is assumed resonant with the exciton. A hallmark of the biexciton is the ac Stark effect, where the light shift changes in sign for detunings in the vicinity of Ebsubscript𝐸bE_{\rm b}italic_E start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT [27, 28, 29, 30, 31]. The two-photon resonance condition δb=0subscript𝛿b0\delta_{\rm b}=0italic_δ start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT = 0 can be considered as a biexciton Feshbach resonance where the effective interactions between the pump and probe generated excitons changes from being attractive (δb>0subscript𝛿b0\delta_{\rm b}>0italic_δ start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT > 0) to repulsive (δb<0subscript𝛿b0\delta_{\rm b}<0italic_δ start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT < 0).

We now extend these concepts to the case of excitons in moiré materials. By measuring the ac Stark effect of MX1 and MX2 excitons under cross-circularly-polarized pump-probe lasers as a function of δ1subscript𝛿1\delta_{1}italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, we explore the nature of biexciton resonances in moiré materials: while a monolayer hosts a single zero momentum biexciton, we find that the presence of moiré potential leads to multiple zero quasi-momentum biexcitons. We first observe that both the MX1 and MX2 excitons blue shift close to resonance (δ122subscript𝛿122\delta_{1}\approx 22italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 22 meV, Fig. 3 a and b) and red shift at larger detunings (δ145subscript𝛿145\delta_{1}\approx 45italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 45 meV, Fig. 3 c and d). Figure 3 e shows the light shift as a function of the pump detuning: from a heuristic fit with the function Δi=atan1(δ1Eb,ib)+csubscriptΔ𝑖𝑎superscript1subscript𝛿1subscript𝐸b𝑖𝑏𝑐\Delta_{i}=a\tan^{-1}{\left(\frac{\delta_{1}-E_{\mathrm{b},i}}{b}\right)}+croman_Δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_a roman_tan start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( divide start_ARG italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT roman_b , italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_b end_ARG ) + italic_c (with i=1,2𝑖12i=1,2italic_i = 1 , 2, a,b,c𝑎𝑏𝑐a,b,citalic_a , italic_b , italic_c fitting parameters), we extract the energy of two biexciton states Eb,1subscript𝐸b1E_{\mathrm{b,1}}italic_E start_POSTSUBSCRIPT roman_b , 1 end_POSTSUBSCRIPT and Eb,2subscript𝐸b2E_{\mathrm{b,2}}italic_E start_POSTSUBSCRIPT roman_b , 2 end_POSTSUBSCRIPT determining the sign change in the ac Stark shifts of MX1 and MX2. From the fit to the MX1 light shift, we extract Eb,136subscript𝐸b136E_{\rm b,1}\approx 36\,italic_E start_POSTSUBSCRIPT roman_b , 1 end_POSTSUBSCRIPT ≈ 36meV, which suggests that the ground state moiré biexciton is more strongly bound than the previously measured MoSe2 monolayer biexciton (with Eb28subscript𝐸b28E_{\rm b}\approx 28\,italic_E start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT ≈ 28meV [25]). The fit to the MX2 light shift shows the presence of a second biexciton state, with a mixed MX1/MX2 character, and a binding energy Eb,228subscript𝐸b228E_{\rm b,2}\approx 28\,italic_E start_POSTSUBSCRIPT roman_b , 2 end_POSTSUBSCRIPT ≈ 28meV measured with respect to an unbound MX1 and MX2 excitons. Importantly, Eb,2subscript𝐸b2E_{\rm b,2}italic_E start_POSTSUBSCRIPT roman_b , 2 end_POSTSUBSCRIPT is smaller than the energy splitting between MX1 and MX2 (40absent40\approx 40≈ 40 meV), and this second biexciton is higher in energy than two unbound MX1 excitons. This should be contrasted with the monolayer scenario, where the bound state by definition has lower energy than two unbound zero-momentum excitons.

II.3 Theoretical model

To further study the nature of the biexciton states, we employ a theoretical model for the scattering of excitons in a moiré potential, which draws inspiration from recent treatments of atoms interacting in low-dimensional optical lattices [39]. The Hamiltonian we consider is

H^=𝐤,σ,λE𝐤λX^𝐤λσX^𝐤λσ^𝐻subscript𝐤𝜎𝜆subscript𝐸𝐤𝜆subscriptsuperscript^𝑋𝐤𝜆𝜎subscript^𝑋𝐤𝜆𝜎\displaystyle\hat{H}=\sum_{\mathbf{k},\sigma,\lambda}E_{\mathbf{k}\lambda}\hat% {X}^{\dagger}_{\mathbf{k}\lambda\sigma}\hat{X}_{\mathbf{k}\lambda\sigma}over^ start_ARG italic_H end_ARG = ∑ start_POSTSUBSCRIPT bold_k , italic_σ , italic_λ end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT bold_k italic_λ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k italic_λ italic_σ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_k italic_λ italic_σ end_POSTSUBSCRIPT
+𝐤,𝐤,𝐪λ,λ,λ,λ𝒱λλλλ(𝐤,𝐤,𝐪)X^𝐤𝐪λX^𝐤+𝐪,λX^𝐤λX^𝐤λ,subscript𝐤superscript𝐤𝐪subscript𝜆superscriptsubscript𝜆subscript𝜆superscriptsubscript𝜆subscriptsuperscript𝒱superscriptsubscript𝜆superscriptsubscript𝜆subscript𝜆subscript𝜆𝐤superscript𝐤𝐪subscriptsuperscript^𝑋𝐤𝐪superscriptsubscript𝜆absentsubscriptsuperscript^𝑋superscript𝐤𝐪superscriptsubscript𝜆absentsubscript^𝑋superscript𝐤subscript𝜆absentsubscript^𝑋𝐤subscript𝜆absent\displaystyle+\sum_{\begin{subarray}{c}\mathbf{k},\mathbf{k}^{\prime},\mathbf{% q}\\ \lambda_{\uparrow},\lambda_{\uparrow}^{\prime},\lambda_{\downarrow},\lambda_{% \downarrow}^{\prime}\end{subarray}}\mathcal{V}^{\lambda_{\uparrow}^{\prime}% \lambda_{\downarrow}^{\prime}}_{\lambda_{\uparrow}\lambda_{\downarrow}}(% \mathbf{k},\mathbf{k}^{\prime},\mathbf{q})\hat{X}^{\dagger}_{\mathbf{k}-% \mathbf{q}\lambda_{\uparrow}^{\prime}\uparrow}\hat{X}^{\dagger}_{\mathbf{k}^{% \prime}+\mathbf{q},\lambda_{\downarrow}^{\prime}\downarrow}\hat{X}_{\mathbf{k}% ^{\prime}\lambda_{\downarrow}\downarrow}\hat{X}_{\mathbf{k}\lambda_{\uparrow}% \uparrow},+ ∑ start_POSTSUBSCRIPT start_ARG start_ROW start_CELL bold_k , bold_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , bold_q end_CELL end_ROW start_ROW start_CELL italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG end_POSTSUBSCRIPT caligraphic_V start_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( bold_k , bold_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , bold_q ) over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k - bold_q italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ↑ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT + bold_q , italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ↓ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_k start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_k italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , (2)

where X^𝐤λσsubscript^𝑋𝐤𝜆𝜎\hat{X}_{\mathbf{k}\lambda\sigma}over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_k italic_λ italic_σ end_POSTSUBSCRIPT annihilates a moiré exciton with energy E𝐤λsubscript𝐸𝐤𝜆E_{\mathbf{k}\lambda}italic_E start_POSTSUBSCRIPT bold_k italic_λ end_POSTSUBSCRIPT, quasi-momentum 𝐤𝐤\mathbf{k}bold_k and band index λ𝜆\lambdaitalic_λ. Here the valley of the exciton is indexed by the spin of the electron inside the exciton, i.e., σ=,𝜎\sigma=\uparrow,\downarrowitalic_σ = ↑ , ↓. The interaction term is written generically, and represents a short-range interaction potential in the basis of the moiré excitonic states: the potential is fixed by enforcing that it yields the correct biexciton binding in the absence of moiré. While this has been measured in monolayer as Eb28subscript𝐸b28E_{\rm b}\approx 28italic_E start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT ≈ 28 meV [25], we reduce it to 20 meV since we anticipate screening from the WSe2 layer. The parameters for the moiré potential — which uses the continuum approximation — are derived from large-scale Density Functional Theory (DFT) [7]. We find that in between MX1 (λ=0𝜆0\lambda=0italic_λ = 0) and MX2 resonances (λ=3𝜆3\lambda=3italic_λ = 3), there are two dark excitonic resonances (λ=1;2𝜆12\lambda=1;2italic_λ = 1 ; 2). While we relegate most details of our model — including a detailed description of the interaction potential and the moiré parameters — to Appendix H, we now provide an overview of the model, including its comparison to experiment.

The scattering physics between an \uparrow and \downarrow exciton is fully characterised by the T𝑇Titalic_T matrix, which corresponds to an infinite sum over exciton-exciton scattering events. We can study the experimentally relevant zero quasi-momentum biexcitonic states by calculating the matrix element of the T𝑇Titalic_T matrix given by

Tλλ;𝟎λλ;𝟎(E)0|X^𝟎λX^𝟎λT^(E)X^𝟎λX^𝟎λ|0,subscriptsuperscript𝑇subscript𝜆subscript𝜆𝟎superscriptsubscript𝜆subscriptsuperscript𝜆𝟎𝐸bra0subscript^𝑋𝟎subscript𝜆absentsubscript^𝑋𝟎subscript𝜆absent^𝑇𝐸subscriptsuperscript^𝑋𝟎superscriptsubscript𝜆absentsubscriptsuperscript^𝑋𝟎superscriptsubscript𝜆absentket0\displaystyle T^{\lambda_{\uparrow}\lambda_{\downarrow};\mathbf{0}}_{\lambda_{% \uparrow}^{\prime}\lambda^{\prime}_{\downarrow};\mathbf{0}}(E)\equiv\bra{0}% \hat{X}_{\mathbf{0}\lambda_{\downarrow}\downarrow}\hat{X}_{\mathbf{0}\lambda_{% \uparrow}\uparrow}\hat{T}(E)\hat{X}^{\dagger}_{\mathbf{0}\lambda_{\uparrow}^{% \prime}\uparrow}\hat{X}^{\dagger}_{\mathbf{0}\lambda_{\downarrow}^{\prime}% \downarrow}\ket{0},italic_T start_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ; bold_0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ; bold_0 end_POSTSUBSCRIPT ( italic_E ) ≡ ⟨ start_ARG 0 end_ARG | over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_0 italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_0 italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT over^ start_ARG italic_T end_ARG ( italic_E ) over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ↑ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ↓ end_POSTSUBSCRIPT | start_ARG 0 end_ARG ⟩ , (3)

with |0ket0\ket{0}| start_ARG 0 end_ARG ⟩ the vacuum and E𝐸Eitalic_E the energy. Importantly, the poles of Eq. (3) are the energies of the zero quasi-momentum biexcitons, which we show in Fig. 3(g). Here, the moiré leads to gaps opening in the two-body (zero quasi-momentum) scattering continuum. We find that the higher energy biexcitons can exist in these gaps, consistent with our experimental observations.

To directly compare the theoretical model to the exciton shifts measured in experiment, we employ a polaron inspired model in which a single \uparrow exciton is dressed by finite quasi-momentum excitations out of a coherent state of \downarrow excitons generated by the pump. Employing a classical description of the pump light, we find the self-energy of the \uparrow exciton within the so-called ladder approximation

Σλλ(E;ωL)=λλβλTλλ;𝟎λλ;𝟎(E+ωL+2iγ)βλ,subscriptΣsubscript𝜆superscriptsubscript𝜆𝐸subscript𝜔𝐿subscriptsubscript𝜆superscriptsubscript𝜆subscript𝛽subscript𝜆subscriptsuperscript𝑇subscript𝜆subscript𝜆𝟎superscriptsubscript𝜆subscriptsuperscript𝜆𝟎𝐸subscript𝜔𝐿2𝑖𝛾subscriptsuperscript𝛽superscriptsubscript𝜆\displaystyle\Sigma_{\lambda_{\uparrow}\lambda_{\uparrow}^{\prime}}(E;\omega_{% L})=\sum_{\lambda_{\downarrow}\lambda_{\downarrow}^{\prime}}\beta_{\lambda_{% \downarrow}}T^{\lambda_{\uparrow}\lambda_{\downarrow};\mathbf{0}}_{\lambda_{% \uparrow}^{\prime}\lambda^{\prime}_{\downarrow};\mathbf{0}}(E+\omega_{L}+2i% \gamma)\beta^{*}_{\lambda_{\downarrow}^{\prime}},roman_Σ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_E ; italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ) = ∑ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_T start_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ; bold_0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ; bold_0 end_POSTSUBSCRIPT ( italic_E + italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT + 2 italic_i italic_γ ) italic_β start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT , (4)

where the magnitude of |βλ|nλproportional-tosubscript𝛽𝜆subscript𝑛𝜆|\beta_{\lambda}|\propto\sqrt{n_{\lambda}}| italic_β start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT | ∝ square-root start_ARG italic_n start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT end_ARG (with nλsubscript𝑛𝜆n_{\lambda}italic_n start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT the density of \downarrow excitons in the λ𝜆\lambdaitalic_λ band; see Appendix H.4 for details) and γ𝛾\gammaitalic_γ is the inverse lifetime of the excitons. Here, ΩΩ\Omegaroman_Ω is the light-matter coupling and we observe that the pump light frequency ωLsubscript𝜔𝐿\omega_{L}italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT effectively leads to off-shell scattering of the excitons. The shift measured in the MX1 (λ=0𝜆0\lambda=0italic_λ = 0) and MX2 (λ=3𝜆3\lambda=3italic_λ = 3) excitons is given by Δλ(δ1)Σλλ(E𝟎,λ;ωL=E𝟎,0δ1)similar-to-or-equalssubscriptΔ𝜆subscript𝛿1subscriptΣ𝜆𝜆subscript𝐸𝟎𝜆subscript𝜔𝐿subscript𝐸𝟎0subscript𝛿1\Delta_{\lambda}(\delta_{1})\simeq\Sigma_{\lambda\lambda}(E_{\mathbf{0},% \lambda};\omega_{L}=E_{\mathbf{0},0}-\delta_{1})roman_Δ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ≃ roman_Σ start_POSTSUBSCRIPT italic_λ italic_λ end_POSTSUBSCRIPT ( italic_E start_POSTSUBSCRIPT bold_0 , italic_λ end_POSTSUBSCRIPT ; italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT bold_0 , 0 end_POSTSUBSCRIPT - italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ), where we have introduced the detuning δ1subscript𝛿1\delta_{1}italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT as defined in experiment.

Figure 3(f) shows the ac Stark shifts predicted by our model. It is necessary to normalize the experimental light shifts by the pump laser power lpumpsubscript𝑙pumpl_{\rm pump}italic_l start_POSTSUBSCRIPT roman_pump end_POSTSUBSCRIPT. Consequently, we normalize the theoretical light shifts by 2Ω2lpumpproportional-tosuperscriptPlanck-constant-over-2-pi2superscriptΩ2subscript𝑙pump\hbar^{2}\Omega^{2}\propto l_{\rm pump}roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∝ italic_l start_POSTSUBSCRIPT roman_pump end_POSTSUBSCRIPT for comparison. Remarkably, our findings are in good qualitative agreement with the experimental observations. The theoretical calculation also suggests that for the measured experimental detunings, only two biexcitons lead to a noticeable sign change in the MX1 and MX2 energies.

III Attractive polaron light shift

Refer to caption
Figure 4: Light shift of moiré polarons. (a) The peak reflection contrast as a function of gate voltage, showing that the oscillator strength of the AP linearly increases until ν=1𝜈1\nu=1italic_ν = 1 and then decreases linearly until ν=2𝜈2\nu=2italic_ν = 2. (b) Reflection spectrum ΔRAP(τ)Δsubscript𝑅AP𝜏\Delta R_{\mathrm{AP}}(\tau)roman_Δ italic_R start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT ( italic_τ ) of the AP at ν=1𝜈1\nu=1italic_ν = 1 as a function of the pump–probe delay τ𝜏\tauitalic_τ. (c) The observed AP blue shift is not driven by interactions as shown by its dependence on the pump detuning. (d) The measured light shift is remarkably constant as a function of ν𝜈\nuitalic_ν, up to a small dispersion 3%similar-toabsentpercent3\sim 3\%∼ 3 % (shown as a blue stripe) compatible with statistical fluctuations (shown as error bar, obtained for each points by repeating four time the measurement). The combination of linear 1/δAP1subscript𝛿AP1/\delta_{\mathrm{AP}}1 / italic_δ start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT dependence of the light shift and its independence of ν𝜈\nuitalic_ν demonstrate the lack of interactions between APs or trions localized on different moiré sites.

Next, we focus on the attractive polaron (AP) resonance which emerges around 35absent35\approx 35\,≈ 35meV on the red side of MX1. The corresponding trion binding energy is about 50%percent5050\%50 % larger than that observed in monolayer MoSe2, suggesting that the electron Wannier orbital is strongly localized around the minimum of the moiré potential. The picture of localized moiré trion is also supported by the doping dependence of the AP oscillator strength, which follows the number of singly occupied moiré sites: It increases linearly until ν=1𝜈1\nu=1italic_ν = 1 and then decreases linearly until ν=2𝜈2\nu=2italic_ν = 2 (Fig. 4 a).

Figure 4 b shows the pump–probe measurement as a function of τ𝜏\tauitalic_τ for δAP=12subscript𝛿AP12\delta_{\mathrm{AP}}=12italic_δ start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT = 12 meV, clearly showing a blue shift of the AP resonance with negligible alteration of the resonance for τ0.2𝜏0.2\tau\geq 0.2italic_τ ≥ 0.2 ps. While this measurement is reminiscent of the result for MX1, Fig. 4 c shows a striking difference in δAPsubscript𝛿AP\delta_{\mathrm{AP}}italic_δ start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT dependence; namely, the AP light shift is better fit using a 1/δAP1subscript𝛿AP1/\delta_{\mathrm{AP}}1 / italic_δ start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT dependence, typical of the ac Stark shift observed for an ensemble of non-interacting two-level emitters. We explain this observation by arguing that the AP resonance can be considered as stemming primarily from a collective excitation of trions localized at the M-M sites of the moiré lattice. In the absence of inter-site hopping, the excitation of a trion at a given site cannot depend on the existence of a trion on any other site, indicating that moiré trions are non-interacting excitations. The only contribution to the light shift in this limit will come from the ac-stark shift of each independent site, whose magnitude scales as 1/δAP1subscript𝛿AP1/\delta_{\mathrm{AP}}1 / italic_δ start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT. Small but non-zero hybridization of the collective trion excitation with the bare exciton could give rise to a finite interaction strength and a deviation form the pure 1/δAP1subscript𝛿AP1/\delta_{\mathrm{AP}}1 / italic_δ start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT contribution to the light shift. We note that recent experiments on the same moiré structure yielded magnetization signatures consistent with unexpectedly weak inter-site hopping [7].

We find a confirmation of this explanation when we measure the electron density dependence of the light shift: Figure 4 d shows that varying ν𝜈\nuitalic_ν from 00 to 2222 results in negligible variation in the magnitude of the light shift, despite large variation in the pump-induced virtual AP population. This behavior contrasts with the strong electron density dependence of the AP light shift previously observed in a monolayer MoSe2 [25], as well as that of MX3, which we identify as an itinerant moiré exciton-polaron existing for filling ν>1𝜈1\nu>1italic_ν > 1 (see Fig. 10 in the Appendix): in stark contrast to the AP, we observe a clear electron density dependence of MX3. Figure 10, together with Figs. 4 c–d demonstrate the power of nonlinear pump–probe measurements in uncovering the strikingly different nature of localized moiré AP from that of itinerant exciton-polaron resonances.

IV Resonant excitation of the attractive polaron

Refer to caption
Figure 5: Resonant excitation of moiré polarons. When the pump and probe lasers are cross-circularly polarized, they drive the AP transition at different moiré sites, hosting electrons with opposite spins (see the sketch in d). As a result, the pump has a negligible effect on the probe spectrum (a). A linearly polarized pump on the other hand, can generate an AP on all sites, leading to a blue shift and most prominently a reduction of the AP oscillator strength (b). For a sufficiently large intensity, all sites ends up in a statistical mixture of a single electron and a moiré trion and consequently the amplitude of the AP resonance is divided by a factor two around zero time delay (c). This phenomenon is observed at all fillings of the lattice (e). Here, the blue dots are the fitted AP amplitudes in absence of a pump pulse. The red squares show the amplitude after the pump pulse, for increasing pump power. We observe a saturation at half the reference amplitude (black dotted line).

Having identified the AP resonance as a collective excitation of independent two-level emitters where the ground and excited states correspond to an electron and a trion at a given site, respectively, we address the possibility of saturating the two-level emitters by resonantly driving the AP resonance. Since we are not interested in measuring the reflection of the strong resonant pump laser, we block it with a polarization filtering. We investigate both cross-circularly and cross-linearly polarized pump and probe configurations (Fig. 5 ab). For cross-circular polarization, the pump generates APs in sites hosting an electron in a given valley, while the probe addresses the sites where electrons are in the opposite valley. Consequently, we observe no change in the probe reflection spectrum even for τ0similar-to-or-equals𝜏0\tau\simeq 0italic_τ ≃ 0.

In stark contrast, a linearly polarized pump pulse addresses all sites, leading to a blue shift of the AP resonance in conjunction with a drastic reduction of its oscillator strength. These two effects relax on a timescale of 5absent5\approx 5≈ 5 ps – much longer than the pump pulse duration τpump0.1similar-to-or-equalssubscript𝜏pump0.1\tau_{\mathrm{pump}}\simeq 0.1italic_τ start_POSTSUBSCRIPT roman_pump end_POSTSUBSCRIPT ≃ 0.1 ps. Such an unexpectedly long relaxation time may stem partly from a long radiative lifetime of trions and partly from generation of dark collective trion states [40, 41, 42, 43]. For short timescales ττpumpless-than-or-similar-to𝜏subscript𝜏pump\tau\lesssim\tau_{\mathrm{pump}}italic_τ ≲ italic_τ start_POSTSUBSCRIPT roman_pump end_POSTSUBSCRIPT, we observe a strong saturation of the AP amplitude, so that the resonant probe reflection contrast reduces to 50%percent5050\%50 % of its value in the absence of the pump pulse (c). This observation indicates that the pump intensity is strong enough to ensure that each moiré site hosting a single electron is driven into a balanced mixture of its ground (single electron) and excited (trion) states. We observe these signatures for all electronic densities in the range 0ν20𝜈20\leq\nu\leq 20 ≤ italic_ν ≤ 2 (Fig. 5 e).

The experiments detailed in Fig. 5 present a realization of a nonequilibrium Bose–Fermi mixture consisting of electrons in a flat moiré band and optically injected excitons [44, 45]. The choice of resonant excitation of the AP transition forces the mixture into a state that can be described as a high density moiré trion gas. Using a circularly polarized resonant pump laser and an external magnetic field to valley polarize electrons, it may be possible to create a trion at each moiré site, thereby realizing a solid-state analog of the Dicke model.

V Discussion

Our findings shed new light on the nature of moiré excitons and polarons, which remains a topic of active research [46, 47, 48, 49, 50, 5, 51, 52, 53, 54, 55, 56, 57, 58, 59, 60, 32]. In particular, we show how non-linear spectroscopy unveil the itinerant or localized character of moiré optical excitations. This insight is crucial for the interpretation of experiments aimed at optical sensing of correlated electronic states [17, 18, 7] and the realization of degenerate Bose-Fermi mixtures in van der Waals heterostructures [44, 45]. Our experiments allow us to assess the extent of spatial overlap between different moiré excitons or attractive polarons. It is somewhat remarkable that this information is accessible to far-field optics given that the moiré length scale is about two orders of magnitude bellow the optical resolution.

The dressing of a quantum material with virtual optical excitations could be a promising route to engineer new phases of matter. The idea of using an intense laser pulse to modify material properties has already been demonstrated, but it often suffers from incoherent pumping and heating effects when an electronic polarization mode of the system is resonantly driven. Similar, albeit much less severe, problems in driven atomic systems can be alleviated using Rydberg dressing, where an off-resonant laser effects coherent hybridization of a ground (or long-lived low-energy) state and a Rydberg state: The atoms remain mostly in the ground state and spontaneous emission is strongly suppressed, but virtual excitations to the Rydberg state still ensure long-range interactions. The strategy we followed is somewhat analogous, albeit employed for a different purpose: The pump laser realizes an excitonic dressing of the moiré system, while keeping light absorption negligible thanks to a large enough detuning from electronic resonances. In turn, the probe pulse monitors the coherent response of the dressed moiré material, yieding information about the nature of the elementary excitations and their interactions.

VI Acknowledgement

We thank M. Hafezi, A. Srivastava, A. Christianen, A. G. Salvador and A. Müller for inspiring discussions. This work was supported by the Swiss National Science Foundation (SNSF) under Grant Number 200020__\__207520. B. E. acknowledges funding from an ETH postdoc fellowship. H. S. A. acknowledges support from the Swiss Government Excellence Scholarship. T. U. acknowledges support from the Funai Overseas Scholarship.

The data are available at the ETH Research Collection [61].

VII Appendix

Appendix A Experimental setup and sample fabrication

The sample was measured in a dry cryostat (Attodry800, Attocube) at cryogenic temperaures of 5absent5\approx 5≈ 5K with free-space optical access and equipped with nanopositioners allowing displacement along the three axes. For the pump and probe, we used a mode-locked Ti:sapphire laser (Tsunami, Spectra-Physics), with a repetition rate of 76 MHz and pulse duration 100absent100\approx 100≈ 100 fs. The pulse is split along two paths for the pump and the probe. The bandwidth of the pump is reduced using a pulse shaper and its power is controlled using a motorized optical attenuator. We achieve a larger spectral width for the probe using a nonlinear fiber (femtowhite 800, NKT Photonics) which produces a quasi-continuum around the investigated resonances. The length of the probe optical path can be varied using a retroreflector on a motorized translation stage, enabling a fine tuning of the time delay between the two pulses. Both beams were focused on a diffraction-limited spot on the sample using an apochromatic microscope objective with NA =0.8absent0.8=0.8= 0.8 (LT-APO/VISIR/0.82, Attocube). The typical probe power is on the order of a few microwatts. The reflected light spectra were recorded using a Peltier-cooled CCD camera.

For the sample fabrication, few-layer graphite, 35absent35\approx 35≈ 35 nm hBN, monolayer MoSe2 and WS2 were mechanically exfoliated. The layers were assembled using the dry-transfer technique with a poly(bisphenol A carbonate) film on a polydimethylsiloxane (PDMS) stamp and deposited on a 285 nm Si/SiO2 substrate [62]. The crystal alignement of the TMDs was determined prior to the stacking measuring the generation of second-harmonic light as a function of the polarization of an incoming infrared laser pulse. They were then stacked with a negligible twist. The graphene top and bottom gates and TMDs were contacted using gold electrodes deposited using optical lithography and electron beam deposition.

Appendix B Data analysis

In order to extract the light shift amplitude, we first fit the reflection spectrum for various time delays and extract the resonance position(s). The result is shown in Fig. 6a  in the case of the MX1 resonance. We then perform a Gaussian fit of the measured line shift as a function of the pump–probe delay in order to exctract the light shift amplitude at zero time delay. To determine the dependence of the light shift on the pump detuning, we perform for each pump wavelength a measurement at various pump powers as shown in Fig. 6b. In this way, we can use relatively low power for a near resonant excitation and larger power at larger detunings, always making sure that we are in a regime of linear scaling with power. In the figures of the main text, we plot the slope Δ/IpumpΔsubscript𝐼pump\Delta/I_{\rm pump}roman_Δ / italic_I start_POSTSUBSCRIPT roman_pump end_POSTSUBSCRIPT computed from a linear fit at each pump detuning δ𝛿\deltaitalic_δ.

Refer to caption
Figure 6: Data analysis. (a) Light shift as a function of the pump–probe delay for MX1 for a pump power of 250 µWtimes250microwatt250\text{\,}\mathrm{\SIUnitSymbolMicro W}start_ARG 250 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_W end_ARG and detuning of 11 meV, showing a clear blue shift at zero time delay. The fit corresponds to a Gaussian envelope. (b) The dependence of the light shift of MX1 on pump power at different pump detunings δ1subscript𝛿1\delta_{1}italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT exhibits a linear dependence.

Appendix C Fit of the wavelength dependence

We discuss here our analysis of the wavelenght dependance of the light shift. Taking into account a single resonance, the light shift ΔΔ\Deltaroman_Δ can be expanded in a series of 1/δ1𝛿1/\delta1 / italic_δ using perturbation theory [21]. The first order 1/δ1𝛿1/\delta1 / italic_δ term comes from light-matter interaction, while exciton-exciton interactions contribute to higher order term, in 1/δ21superscript𝛿21/\delta^{2}1 / italic_δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT or 1/(δiδj)1subscript𝛿𝑖subscript𝛿𝑗1/(\delta_{i}\delta_{j})1 / ( italic_δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) when several modes contributes. In principle, a fine analysis of the detuning dependence of the light shift would enable to deconvolve the various contributions. In practice, this analysis can be challenging due to the finite range of detuning in which we can take reliable data. Indeed, for all resonances, we had to restrict the data to a window of [20,80]absent2080\approx[20,80]≈ [ 20 , 80 ] meV. Going closer to resonance we face two issues: First, incoherent effects become more prominent as pump photons carry enough energy to generate a real population of excitons and hence are more likely to be absorbed. Second, the perturbative expansion of the light shift becomes inconsistent when Δδsimilar-toΔ𝛿\Delta\sim\deltaroman_Δ ∼ italic_δ. Conversely, going further away from resonance the signal reduces (given the available laser power) and the extraction of the light shift becomes unreliable. Furthermore, in that regime, the light–matter terms 1/δproportional-toabsent1𝛿\propto 1/\delta∝ 1 / italic_δ becomes stronger compared to the interaction term 1/δ2proportional-toabsent1superscript𝛿2\propto 1/\delta^{2}∝ 1 / italic_δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT that we are interested in.

C.1 Charge neutrality

Refer to caption
Figure 7: Pump wavelength dependence. From a fit we infer the origin of the light shift of MX1 (a), MX2 exciton (b) and of the attractive polaron at ν=1𝜈1\nu=1italic_ν = 1 c. At charge neutrality (a and b), we are not able to deconvolve the contribution of the u1,2superscriptsubscript𝑢12u_{1,2}^{\prime}italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT and k2superscriptsubscript𝑘2k_{2}^{\prime}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT fitting parameters (see Eq. [8]). We thus perform two fits, setting k2superscriptsubscript𝑘2k_{2}^{\prime}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT to its upper and lower bounds, shown as solid blue and red line, respectively. For the MX1 exciton (a), the light shift is dominated by the 2u1,1n12subscript𝑢11subscript𝑛12u_{1,1}n_{1}2 italic_u start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT term (shown as dashed line) and barely sensitive to the MX2-dependent terms, so both fit are nearly identical and we obtain a good estimate of u1,1subscript𝑢11u_{1,1}italic_u start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT. For MX2 (b), both fits are in fair agreement with the data, but the contribution of the 2u1,2n12subscript𝑢12subscript𝑛12u_{1,2}n_{1}2 italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT term (dotted line) is significantly different, and we are thus only able to provide a rather broad confidence interval for u1,2subscript𝑢12u_{1,2}italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT. For the AP, a fit AAP/δ1+BAP/δAP2subscript𝐴APsubscript𝛿1subscript𝐵APsuperscriptsubscript𝛿AP2A_{\mathrm{AP}}/\delta_{1}+B_{\mathrm{AP}}/\delta_{\mathrm{AP}}^{2}italic_A start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT / italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_B start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT / italic_δ start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (solid black line, c) suggests that the usual AC-Stark shift scaling as AAP/δ1subscript𝐴APsubscript𝛿1A_{\mathrm{AP}}/\delta_{1}italic_A start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT / italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (dashed-dotted line) is occurring, instead of an interaction-driven shift scaling as BAP/δAP2subscript𝐵APsuperscriptsubscript𝛿AP2B_{\mathrm{AP}}/\delta_{\mathrm{AP}}^{2}italic_B start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT / italic_δ start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (dotted line).

With the relatively small detuning we use (compared to the exciton Rydberg energy), the exciton-exciton interaction is expected to be the dominant contribution to the light shift [25]. This is indeed confirmed by our data which are not compatible with a 1/δ1𝛿1/\delta1 / italic_δ law. Let us thus focus on the interaction induced light shift, which we generalize to our moiré system with two bright exciton modes. As shown in App. H.6, the light shifts of MX1 and MX2 read

Δ1subscriptΔ1\displaystyle\Delta_{1}roman_Δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =2u1,1n1+2u1,2n2+4k1n1n2,absent2subscript𝑢11subscript𝑛12subscript𝑢12subscript𝑛24subscript𝑘1subscript𝑛1subscript𝑛2\displaystyle=2u_{1,1}n_{1}+2u_{1,2}n_{2}+4k_{1}\sqrt{n_{1}n_{2}}\,,= 2 italic_u start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 2 italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + 4 italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT square-root start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG , (5)
Δ2subscriptΔ2\displaystyle\Delta_{2}roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT =2u2,2n2+2u1,2n1+4k2n1n2.absent2subscript𝑢22subscript𝑛22subscript𝑢12subscript𝑛14subscript𝑘2subscript𝑛1subscript𝑛2\displaystyle=2u_{2,2}n_{2}+2u_{1,2}n_{1}+4k_{2}\sqrt{n_{1}n_{2}}\,.= 2 italic_u start_POSTSUBSCRIPT 2 , 2 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + 2 italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + 4 italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT square-root start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG . (6)

with the densities scaling as ni|Ωiϕi(0)|2/δi2proportional-tosubscript𝑛𝑖superscriptsubscriptΩ𝑖subscriptitalic-ϕ𝑖02superscriptsubscript𝛿𝑖2n_{i}\propto|\Omega_{i}\phi_{i}(0)|^{2}/\delta_{i}^{2}italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∝ | roman_Ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_δ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The ratio r=|Ω1ϕ1(0)|2/|Ω2ϕ2(0)|2=γrad,1/γrad,20.48𝑟superscriptsubscriptΩ1subscriptitalic-ϕ102superscriptsubscriptΩ2subscriptitalic-ϕ202subscript𝛾rad1subscript𝛾rad20.48r=|\Omega_{1}\phi_{1}(0)|^{2}/|\Omega_{2}\phi_{2}(0)|^{2}=\gamma_{\rm rad,1}/% \gamma_{\rm rad,2}\approx 0.48italic_r = | roman_Ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( 0 ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / | roman_Ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( 0 ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_γ start_POSTSUBSCRIPT roman_rad , 1 end_POSTSUBSCRIPT / italic_γ start_POSTSUBSCRIPT roman_rad , 2 end_POSTSUBSCRIPT ≈ 0.48 can be obtained from a fit of the reflection contrast (see App. D). Using this independent estimate and performing a joint fit of Δ1,2subscriptΔ12\Delta_{1,2}roman_Δ start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT we reduce the number of fitting parameters to five, A𝐴Aitalic_A, u2,2superscriptsubscript𝑢22u_{2,2}^{\prime}italic_u start_POSTSUBSCRIPT 2 , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, u1,2superscriptsubscript𝑢12u_{1,2}^{\prime}italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, k1superscriptsubscript𝑘1k_{1}^{\prime}italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT and k2superscriptsubscript𝑘2k_{2}^{\prime}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, where ui,j=uij/u1,1superscriptsubscript𝑢𝑖𝑗subscript𝑢𝑖𝑗subscript𝑢11u_{i,j}^{\prime}=u_{ij}/u_{1,1}italic_u start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_u start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT / italic_u start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT, ki=ki/u1,1superscriptsubscript𝑘𝑖subscript𝑘𝑖subscript𝑢11k_{i}^{\prime}=k_{i}/u_{1,1}italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / italic_u start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT and

Δ1subscriptΔ1\displaystyle\Delta_{1}roman_Δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =A(1δ12+r2u1,2δ22+2rk1δ1δ2),absent𝐴1superscriptsubscript𝛿12superscript𝑟2superscriptsubscript𝑢12superscriptsubscript𝛿222𝑟superscriptsubscript𝑘1subscript𝛿1subscript𝛿2\displaystyle=A\left(\frac{1}{\delta_{1}^{2}}+\frac{r^{2}u_{1,2}^{\prime}}{% \delta_{2}^{2}}+\frac{2rk_{1}^{\prime}}{\delta_{1}\delta_{2}}\right)\,,= italic_A ( divide start_ARG 1 end_ARG start_ARG italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG start_ARG italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG 2 italic_r italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG start_ARG italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) , (7)
Δ2subscriptΔ2\displaystyle\Delta_{2}roman_Δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT =A(r2u2,2δ22+u1,2δ12+2rk22,2δ1δ2).absent𝐴superscript𝑟2superscriptsubscript𝑢22superscriptsubscript𝛿22superscriptsubscript𝑢12superscriptsubscript𝛿122𝑟superscriptsubscript𝑘222subscript𝛿1subscript𝛿2\displaystyle=A\left(\frac{r^{2}u_{2,2}^{\prime}}{\delta_{2}^{2}}+\frac{u_{1,2% }^{\prime}}{\delta_{1}^{2}}+\frac{2rk_{2}^{\prime 2,2}}{\delta_{1}\delta_{2}}% \right)\,.= italic_A ( divide start_ARG italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_u start_POSTSUBSCRIPT 2 , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG start_ARG italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG start_ARG italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG 2 italic_r italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ 2 , 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_δ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) . (8)

Furthermore, within a Born approximation Uijkl𝑑rϕiϕjϕkϕlproportional-tosuperscriptsubscript𝑈𝑖𝑗𝑘𝑙differential-d𝑟subscriptitalic-ϕ𝑖subscriptitalic-ϕ𝑗superscriptsubscriptitalic-ϕ𝑘superscriptsubscriptitalic-ϕ𝑙U_{ij}^{kl}\propto\int dr\phi_{i}\phi_{j}\phi_{k}^{*}\phi_{l}^{*}italic_U start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k italic_l end_POSTSUPERSCRIPT ∝ ∫ italic_d italic_r italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT (see App H.6) and using the Cauchy-Schwarz inequality, we obtain the following bounds 0<u1,2<[u2,2]1/20superscriptsubscript𝑢12superscriptdelimited-[]superscriptsubscript𝑢22120<u_{1,2}^{\prime}<[u_{2,2}^{\prime}]^{1/2}0 < italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT < [ italic_u start_POSTSUBSCRIPT 2 , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT , |k1|<[u1,2]1/2superscriptsubscript𝑘1superscriptdelimited-[]subscript𝑢1212|k_{1}^{\prime}|<[u_{1,2}]^{1/2}| italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | < [ italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT , |k2|<[u2,2u1,2]1/2superscriptsubscript𝑘2superscriptdelimited-[]superscriptsubscript𝑢22superscriptsubscript𝑢1212|k_{2}^{\prime}|<[u_{2,2}^{\prime}u_{1,2}^{\prime}]^{1/2}| italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | < [ italic_u start_POSTSUBSCRIPT 2 , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT, which we enforce to improve the convergence of the fit.

For MX1, see fig. 7a, the full fit reveals the dominant contribution to be the MX1-MX1 interaction. We obtain the fitting parameter A𝐴Aitalic_A, which we can compare to the value obtain from fitting the light shift of a monolayer exciton (on a monolayer MoSe2 region of the same device). We estimate u1,1/uex1.6±0.2subscript𝑢11subscript𝑢explus-or-minus1.60.2u_{1,1}/u_{\rm ex}\approx 1.6\pm 0.2italic_u start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT / italic_u start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT ≈ 1.6 ± 0.2, where uexsubscript𝑢exu_{\rm ex}italic_u start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT is the monolayer exciton-exciton interaction (see App. E for an absolute calibration).

For MX2, see fig. 7b, the fit is unable to reliably disentangle the contribution of the second and third terms of 8, corresponding the sattering of an MX1 exciton, without (1st term) or with (2nd term) a change of moiré band. Indeed, performing a fit with each of these two terms independently, we obtain in both case a reasonable agreement with the data over the full detuning range. We thus perform two fits, where the coupling k2=±[u2,2u1,2]1/2superscriptsubscript𝑘2plus-or-minussuperscriptdelimited-[]superscriptsubscript𝑢22superscriptsubscript𝑢1212k_{2}^{\prime}=\pm[u_{2,2}^{\prime}u_{1,2}^{\prime}]^{1/2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = ± [ italic_u start_POSTSUBSCRIPT 2 , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_u start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT staturates the Cauchy-Schwartz inequality. In this way we obtain the bounds 0.3u2,20.8less-than-or-approximately-equals0.3subscript𝑢22less-than-or-approximately-equals0.80.3\lessapprox u_{2,2}\lessapprox 0.80.3 ⪅ italic_u start_POSTSUBSCRIPT 2 , 2 end_POSTSUBSCRIPT ⪅ 0.8.

C.2 Moiré attractive polaron

For the AP (c), the light shift is typically smaller and more noisy due to the weakness of the transition. As a result, it is more difficult to discriminate a potential 1/δ1𝛿1/\delta1 / italic_δ and 1/δ21superscript𝛿21/\delta^{2}1 / italic_δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT dependence. Nevertheless, the fit does suggest that the former is here the leading contribution, consistent with the picture of an ensemble of distinguishable and non-interacting two-level systems, as argued in the main text.

Appendix D Transfer matrix simulation

Refer to caption
 Resonance  ν𝜈\nuitalic_ν  γradPlanck-constant-over-2-pisubscript𝛾rad\hbar\gamma_{\rm rad}roman_ℏ italic_γ start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT [meV]  γnon-radPlanck-constant-over-2-pisubscript𝛾non-rad\hbar\gamma_{\text{non-rad}}roman_ℏ italic_γ start_POSTSUBSCRIPT non-rad end_POSTSUBSCRIPT [meV]
MX1 0 1.0±0.2plus-or-minus1.00.21.0\pm 0.21.0 ± 0.2 1.3±0.3plus-or-minus1.30.31.3\pm 0.31.3 ± 0.3
MX2 0 0.48±0.1plus-or-minus0.480.10.48\pm 0.10.48 ± 0.1 2.8±0.4plus-or-minus2.80.42.8\pm 0.42.8 ± 0.4
AP 1 0.27±0.1plus-or-minus0.270.10.27\pm 0.10.27 ± 0.1 2.7±0.4plus-or-minus2.70.42.7\pm 0.42.7 ± 0.4
MX2superscriptsubscriptabsent2{}_{2}^{\prime}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT 1 1.1±0.3plus-or-minus1.10.31.1\pm 0.31.1 ± 0.3 2.9±0.4plus-or-minus2.90.42.9\pm 0.42.9 ± 0.4
MX3 2 0.52±0.1plus-or-minus0.520.10.52\pm 0.10.52 ± 0.1 2.2±0.3plus-or-minus2.20.32.2\pm 0.32.2 ± 0.3
Figure 8: Fit of the reflection spectrum using transfer matrix simulation. The blue dots are the data, the black solid line is a fit (grey area span by varying the fitting parameters within the confidence interval).

In order to infer the interaction strength of the various moiré excitons from their light shift, we need to estimate the exciton density that we generate, and hence the exciton oscillator strength. The latter can be obtained from a fit of the reflection spectrum. Such a fit needs to include the reflection of the electromagnetic field on the interfaces between the various dielectrics that make our van der Waals heterostructure. We do this using the transfer matrix method [63]. Two fitting parameters for the background are the hBN thickness 31absent31\approx 31\,≈ 31nm and 37absent37\approx 37\,≈ 37nm for the top and bottom layer and the hBN refractive index nhBN2.15subscript𝑛hBN2.15n_{\rm hBN}\approx 2.15italic_n start_POSTSUBSCRIPT roman_hBN end_POSTSUBSCRIPT ≈ 2.15 [64, 65]. Then, for each resonance, we have three additional fitting parameters, namely its energy, radiative and non-radiative decay rates. The results of this fit are shown in Fig. 8. We show here the bare reflection spectrum, obtained using a light source that is to a very good approximation spectrally flat in the energy range shown in the figure. We are unable to obtain a perfect fit of the background, using the hBN thicknesses and refractive index as free parameters. The discrepancy that we observe, especially on the edge of the spectrum could be due to chromatic aberrations (although we are using a microscope apochromatic objective to limit those). Nevertheless, in the center of the spectrum, we are able to reproduce our spectra very well at all fillings.

Appendix E Interaction strength

The radiative (γrsubscript𝛾r\gamma_{\rm r}italic_γ start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT) and non-radiative (γnrsubscript𝛾nr\gamma_{\rm nr}italic_γ start_POSTSUBSCRIPT roman_nr end_POSTSUBSCRIPT) decay rates can be used to extract the density of excitons nexsubscript𝑛exn_{\rm ex}italic_n start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT induced by the pump. Using the optical Bloch equations within the adiabatic approximation we have [63]

nex=2Iiγr(γr+γnr)2+δex2,subscript𝑛ex2subscript𝐼isubscript𝛾rsuperscriptsubscript𝛾rsubscript𝛾nr2superscriptsubscript𝛿ex2\displaystyle n_{\rm ex}=\frac{2I_{\rm i}\gamma_{\rm r}}{(\gamma_{\rm r}+% \gamma_{\rm nr})^{2}+\delta_{\rm ex}^{2}}\,,italic_n start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT = divide start_ARG 2 italic_I start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT italic_γ start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT end_ARG start_ARG ( italic_γ start_POSTSUBSCRIPT roman_r end_POSTSUBSCRIPT + italic_γ start_POSTSUBSCRIPT roman_nr end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_δ start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (9)

where Iisubscript𝐼iI_{\rm i}italic_I start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT is the photon flux. For reference, we first look at the exciton light shift ΔexsubscriptΔex\Delta_{\rm ex}roman_Δ start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT measured on a monolayer region of the sample, as a function of the exciton density, see Fig. 9. We observe a linear dependence Δex=UexnexsubscriptΔexsubscript𝑈exsubscript𝑛ex\Delta_{\rm ex}=U_{\rm ex}n_{\rm ex}roman_Δ start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT = italic_U start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT, from which we extract the exciton–exciton interaction strength Uex0.06(3) µeV µm2subscript𝑈extimesuncertain0.063timesmicroelectronvoltmicrometer2U_{\rm ex}\approx$0.06(3)\text{\,}\mathrm{\SIUnitSymbolMicro eV}\text{\,}{% \mathrm{\SIUnitSymbolMicro m}}^{2}$italic_U start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT ≈ start_ARG start_ARG 0.06 end_ARG start_ARG ( 3 ) end_ARG end_ARG start_ARG times end_ARG start_ARG start_ARG roman_µ roman_eV end_ARG start_ARG times end_ARG start_ARG power start_ARG roman_µ roman_m end_ARG start_ARG 2 end_ARG end_ARG end_ARG. The latter is compatible with previous measurements [66, 63, 40, 25], although significantly smaller than theoretical estimates, 3Eexaex21 µeV µm2similar-toabsent3subscript𝐸exsuperscriptsubscript𝑎ex2similar-totimes1timesmicroelectronvoltmicrometer2\sim 3\,E_{\mathrm{ex}}a_{\mathrm{ex}}^{2}\sim$1\text{\,}\mathrm{% \SIUnitSymbolMicro eV}\text{\,}{\mathrm{\SIUnitSymbolMicro m}}^{2}$∼ 3 italic_E start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∼ start_ARG 1 end_ARG start_ARG times end_ARG start_ARG start_ARG roman_µ roman_eV end_ARG start_ARG times end_ARG start_ARG power start_ARG roman_µ roman_m end_ARG start_ARG 2 end_ARG end_ARG end_ARG, where Eexsubscript𝐸exE_{\mathrm{ex}}italic_E start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT is the exciton binding energy and aexsubscript𝑎exa_{\mathrm{ex}}italic_a start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT is the exciton Bohr radius [36, 37].

From the measurement of the light shift of the moiré excitons and polaron shown in Fig. 7 we obtain UMX1–MX11.6Uexsubscript𝑈MX1–MX11.6subscript𝑈exU_{\text{MX1--MX1}}\approx 1.6\,U_{\rm ex}italic_U start_POSTSUBSCRIPT MX1–MX1 end_POSTSUBSCRIPT ≈ 1.6 italic_U start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT , UMX1–MX20.7Uexsubscript𝑈MX1–MX20.7subscript𝑈exU_{\text{MX1--MX2}}\approx 0.7\,U_{\rm ex}italic_U start_POSTSUBSCRIPT MX1–MX2 end_POSTSUBSCRIPT ≈ 0.7 italic_U start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT and UMX3–MX35.3Uexsubscript𝑈MX3–MX35.3subscript𝑈exU_{\text{MX3--MX3}}\approx 5.3\,U_{\rm ex}italic_U start_POSTSUBSCRIPT MX3–MX3 end_POSTSUBSCRIPT ≈ 5.3 italic_U start_POSTSUBSCRIPT roman_ex end_POSTSUBSCRIPT. The enhancement of the interaction strength of MX1 (charge neutrality) could stem from the partial confinement induced by the moiré potential. The larger enhancement for MX3 reflects the polaronic nature of this resonance, as previously observed in a monolayer sample [25].

Refer to caption
Figure 9: Light shift as a function of the exciton density. The blue dots are the data, obtained for various powers and detunings. The error bars come from the uncertainty of the exciton decay rates and on the power on the sample. The black solid line is a fit with the grey area span obtained by varying the fitting parameters within the confidence interval.

Appendix F Light shift of MX3 in the regime of large doping

Refer to caption
Figure 10: Light shift of MX3. The resonance MX3 shows distinctive kinks in its amplitude (a) and energy (b) at integer fillings of the moiré lattice with electrons, following the changes in the 2DES compressibility. The light shift of the moiré exciton MX3 at various fillings (c), is well fitted by a line on top of a Gaussian function (black line). The latter captures the coherent response of the system, which corresponds to an interaction-induced blue shift for co-circularly polarized pump and probe, and possibly to a red shift stemming from the coupling to the biexciton states in a cross-circular configuration. The amplitudes of these two shifts are sensitive to the polaron dressing of the exciton, and consequently are extremal at integer filling of the moiré lattice (d).

Even though MX3 is the dominant excitonic resonance for electron filling factors ν1.5𝜈1.5\nu\geq 1.5italic_ν ≥ 1.5, its identification has remained unclear. To gain insight, we investigate the nonlinear response of the MX3 resonance. Figure 10a, b show the reflection amplitude and the energy of MX3 for 1.5ν3.41.5𝜈3.41.5\leq\nu\leq 3.41.5 ≤ italic_ν ≤ 3.4 in the absence of a pump laser: Consistent with earlier observations, we find that the resonance energy as well as the reflection strength, or equivalently the oscillator strength, of MX3 exhibit local maxima at integer fillings ν=2𝜈2\nu=2italic_ν = 2 and ν=3𝜈3\nu=3italic_ν = 3. These features could be explained by partial suppression of dynamical dressing of excitons by electrons, when the two dimensional electron systems (2DES) is in an incompressible state [67, 17]. In contrast, when the electrons form a Fermi liquid (ν2,3𝜈23\nu\neq 2,3italic_ν ≠ 2 , 3), the dynamical dressing of MX3 is more effective and results in a red shift together with a reduction of the oscillator strength.

Figure 10c shows the light shift as a function of pump–probe delay τ𝜏\tauitalic_τ for four representative filling factors obtained for δ3=80subscript𝛿380\delta_{3}=80italic_δ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 80 meV. We observe that the the light shift for τ0similar-to-or-equals𝜏0\tau\simeq 0italic_τ ≃ 0 indicates repulsive (attractive) interactions between same (opposite) valley MX3 excitons generated by co- (cross-) circularly polarized pump–probe fields. While attractive interactions between opposite valley excitons has been reported before, it is surprising that the magnitude of the light shift is comparable in the two cases. The attractive interactions for the cross-polarized configuration may be explained through a near-resonant two-photon (pump+probe) excitation of the biexciton resonance at ωXXsubscript𝜔XX\omega_{\mathrm{XX}}italic_ω start_POSTSUBSCRIPT roman_XX end_POSTSUBSCRIPT. Verification of this hypothesis could be achieved by changing the pump detuning δ3subscript𝛿3\delta_{3}italic_δ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT so as to probe both δ3ωXXsubscript𝛿3subscript𝜔XX\delta_{3}\leq\omega_{\mathrm{XX}}italic_δ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ≤ italic_ω start_POSTSUBSCRIPT roman_XX end_POSTSUBSCRIPT and δ3ωXXsubscript𝛿3subscript𝜔XX\delta_{3}\geq\omega_{\mathrm{XX}}italic_δ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ≥ italic_ω start_POSTSUBSCRIPT roman_XX end_POSTSUBSCRIPT, since for the latter case, the biexciton-mediated interactions would become repulsive. As we had to choose δ3<ωXXsubscript𝛿3subscript𝜔XX\delta_{3}<\omega_{\mathrm{XX}}italic_δ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT < italic_ω start_POSTSUBSCRIPT roman_XX end_POSTSUBSCRIPT to avoid strong background absorption, we could not verify the role of biexciton in the measured light shift.

In contrast to the light shift measurements in the charge-neutFral regime, we find that the pump pulse results in a MX3 line shift that increases linearly with τ𝜏\tauitalic_τ for 0.2t3.00.2𝑡3.00.2\leq t\leq 3.00.2 ≤ italic_t ≤ 3.0 ps. Moreover, the linear shift at a given ν𝜈\nuitalic_ν is identical for co- and cross-circularly polarized pump–probe configurations, but has a different sign for compressible (ν2,3𝜈23\nu\neq 2,3italic_ν ≠ 2 , 3) and incompressible (ν=2,3𝜈23\nu=2,3italic_ν = 2 , 3) electron states. We tentatively explain this feature by generation of free carriers by non-resonant absorption of pump-photons that change the electron density for timescales well exceeding the pump duration.

Since the MX3 resonance energy has maxima (minima) for ν=2,3𝜈23\nu=2,3italic_ν = 2 , 3 (ν=3/2,5/2𝜈3252\nu=3/2,5/2italic_ν = 3 / 2 , 5 / 2), any pump-induced change in electron density will result in a red (blue) shift of the resonance energy. While we do not understand why the red (blue) shift depends linearly on τ𝜏\tauitalic_τ for τ>τpump𝜏subscript𝜏pump\tau>\tau_{\mathrm{pump}}italic_τ > italic_τ start_POSTSUBSCRIPT roman_pump end_POSTSUBSCRIPT, we speculate that pump-induced charges are initially generated in high-energy bands and that they influence the nonlinear response only as they relax to the lowest-energy available moiré band.

We also observe in Fig. 10c that the magnitude of the light shift for the co-circularly polarized pump–probe configuration is smaller for incompressible states. Plotting Δ3subscriptΔ3\Delta_{3}roman_Δ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT as a function of ν𝜈\nuitalic_ν (Fig. 10d) shows that the blue shift is indeed minimal for ν=2,3𝜈23\nu=2,3italic_ν = 2 , 3. This is at a first glance surprising given that the oscillator strength, and consequently the generated MX3 population, is maximal for these incompressible states. However, it was recently shown that interactions between attractive exciton–polarons [25, 68, 40] mediated by their dressing cloud, are dramatically enhanced compared to those of bare excitons. Such an enhancement of interaction strength may overcome the reduction of the oscillator strength of MX3 for ν2,3𝜈23\nu\neq 2,3italic_ν ≠ 2 , 3. This tentative explanation suggests that the MX3 mode may be identified as a second attractive polaron mode where the exciton is dressed by electrons in the upper moiré band. Last but not least, we find that the red shift of MX3 in the cross-circularly polarized configuration is maximal when the electronic state is incompressible; we currently do not have an explanation for this observation.

Appendix G Light shift of MX2subscriptsuperscriptabsent2{}^{\prime}_{2}start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT at ν=1𝜈1\nu=1italic_ν = 1

At a unity filling ν=1𝜈1\nu=1italic_ν = 1 of the moiré potential, we observe two bright resonances, the AP which we discussed in detail in the main text, and MX1superscriptsubscriptabsent1{}_{1}^{\prime}start_FLOATSUBSCRIPT 1 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, emerging from MX2, and which we now investigate. Depending on the pump and probe polarizations, we observed different behaviors. In co-circular polarization, we obtain the usual blue shift which we attribute to MX2superscriptsubscriptabsent2{}_{2}^{\prime}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT–MX2superscriptsubscriptabsent2{}_{2}^{\prime}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT interactions. Contrary to other resonances, we cannot confirm this claim by an analysis of the detuning dependence. Indeed, we observed strong incoherent behavior for a pump blue detuned from the AP, and we therefore only explored the red detuned situation. Specifically, we explore the range δMX2[80,110]subscript𝛿superscriptMX280110\delta_{\rm MX2^{\prime}}\approx[80,110]italic_δ start_POSTSUBSCRIPT MX2 start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ≈ [ 80 , 110 ] meV in which we observe no significant evolution of the light shift, as expected from a scaling as 1/δMX221superscriptsubscript𝛿superscriptsubscriptMX221/\delta_{\rm MX_{2}^{\prime}}^{2}1 / italic_δ start_POSTSUBSCRIPT roman_MX start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, see Fig. 11a. By contrast, in cross-circular polarization we observe a distinct redshift which diverges close to the AP resonance, in excellent agreement with a 1/δAP21superscriptsubscript𝛿AP21/\delta_{\rm AP}^{2}1 / italic_δ start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT scaling and suggesting an attractive interaction between AP and MX2superscriptsubscriptabsent2{}_{2}^{\prime}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT in opposite valleys. We point out that a similar behavior was observed in a monolayer system, and tentatively attributed to the reduction of the phase space filling upon the generation of an AP, for an opposite valley exciton [25]. The pump power dependence (b) of the light shift shows an interesting behavior for a near-resonant excitation of the AP δAP8subscript𝛿AP8\delta_{\rm AP}\approx 8italic_δ start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT ≈ 8 meV. In that case, at high intensity, we expect a saturation of the AP density as the moiré potential is filled up, as described in the main text (although here, keeping a finite δAPsubscript𝛿AP\delta_{\mathrm{AP}}italic_δ start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT we are unable to fully saturate the transition). Indeed, we observe a sub-linear increase of the cross-polarized light shift attributed to AP–MX2superscriptsubscriptabsent2{}_{2}^{\prime}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT interactions. On the contrary, the co-polarized light shift which we attribute to MX2superscriptsubscriptabsent2{}_{2}^{\prime}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT–MX2superscriptsubscriptabsent2{}_{2}^{\prime}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT interactions is linear in the pump power, as the density of MX2superscriptsubscriptabsent2{}_{2}^{\prime}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT remains far from saturation of the moiré lattice (δMX280subscript𝛿superscriptsubscriptMX280\delta_{{\rm MX}_{2}^{\prime}}\approx 80\,italic_δ start_POSTSUBSCRIPT roman_MX start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ≈ 80meV).

Refer to caption
Figure 11: Light shift of MX2superscriptsubscriptabsent2{}_{2}^{\prime}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT at ν=1𝜈1\nu=1italic_ν = 1. (a) Detuning dependence of the light shift of the resonance MX2subscriptsuperscriptabsent2{}^{\prime}_{2}start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT at a filling ν=1𝜈1\nu=1italic_ν = 1. Red dots correspond to cross-polarized pump and probe and show a red shift of the MX2subscriptsuperscriptabsent2{}^{\prime}_{2}start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT resonance which can be well fitted (red line) by a dependence 1/δAP2proportional-toabsent1superscriptsubscript𝛿AP2\propto 1/\delta_{\rm AP}^{2}∝ 1 / italic_δ start_POSTSUBSCRIPT roman_AP end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT stemming from MX2subscriptsuperscriptabsent2{}^{\prime}_{2}start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT–AP interactions. Blue dots correspond to co-polarized pump and probe and show a blue shift of the MX2subscriptsuperscriptabsent2{}^{\prime}_{2}start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT resonance which can be well fitted (blue line) assuming MX2superscriptsubscriptabsent2{}_{2}^{\prime}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT–MX2subscriptsuperscriptabsent2{}^{\prime}_{2}start_FLOATSUPERSCRIPT ′ end_FLOATSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT interactions and a scaling 1/δMX22proportional-toabsent1superscriptsubscript𝛿superscriptsubscriptMX22\propto 1/\delta_{\rm MX_{2}^{\prime}}^{2}∝ 1 / italic_δ start_POSTSUBSCRIPT roman_MX start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. (b) Lightshift of MX2superscriptsubscriptabsent2{}_{2}^{\prime}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT as a function of pump power in different polarizations. Cross-circular polarization shows a distinct red shift; the deviation from a linear dependence is due to saturation of the AP population at large powers. Co-linear polarization shows a smaller blue shift which is well in the linear regime. (c) MX2superscriptsubscriptabsent2{}_{2}^{\prime}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT shift vs. pump–probe time delay in co-circular (blue) for a pump power of 500 µWtimes500microwatt500\text{\,}\mathrm{\SIUnitSymbolMicro W}start_ARG 500 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_W end_ARG and cross-circular (red) polarizations for 50 µWtimes50microwatt50\text{\,}\mathrm{\SIUnitSymbolMicro W}start_ARG 50 end_ARG start_ARG times end_ARG start_ARG roman_µ roman_W end_ARG.

Appendix H Theoretical model

In this section, we primarily provide a detailed description of the theoretical model used to analyze the charge-neutral, cross-circular polarization data. The foremost goal of this model is to qualitatively capture the energy shifts observed in the MX1 and MX2 exciton resonances when subjected to the influence of a pump laser.

As discussed in the main text, the energy shifts can be interpreted as the dressing of the probe exciton (denoted by \uparrow) by virtual excitons (\downarrow) created by the pump. To elucidate this mechanism, we begin by examining the scattering processes involving two distinguishable excitons. Specifically, we calculate the exciton-exciton T𝑇Titalic_T matrix in the presence of a moiré potential, which is inspired by a recent treatment of two interacting atoms in a 2D square optical lattice [39]. This approach allows us to capture the essential physics of the interaction under the influence of the pump. The dressing effect is then incorporated by applying non-self-consistent T𝑇Titalic_T matrix theory.

We conclude the section with a brief discussion on a model for the co-circular polarization data. For ease of notation we set the reduced Planck’s constant and the system area and to unity (=𝒜=1Planck-constant-over-2-pi𝒜1\hbar=\mathcal{A}=1roman_ℏ = caligraphic_A = 1).

H.1 Model for interactions of two distinguishable excitons

We begin by modelling the interactions between two distinguishable rigid (1s) excitons in the presence of moiré. The Hamiltonian we consider consists of four terms

H=HX+HX+Hd+V.𝐻subscript𝐻𝑋absentsubscript𝐻𝑋absentsubscript𝐻𝑑𝑉\displaystyle H=H_{X\uparrow}+H_{X\downarrow}+H_{d}+V.italic_H = italic_H start_POSTSUBSCRIPT italic_X ↑ end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_X ↓ end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT + italic_V . (10)

The first two terms represent the exciton Hamiltonians for each valley (σ=,𝜎\sigma={\uparrow,\downarrow}italic_σ = ↑ , ↓):

HXσ=𝐊ϵ𝐊,XX^𝐊σX^𝐊σ+𝐊𝐐V~X(𝐐)X^𝐊+𝐐σX^𝐊σ.subscript𝐻𝑋𝜎subscript𝐊subscriptitalic-ϵ𝐊𝑋subscriptsuperscript^𝑋𝐊𝜎subscript^𝑋𝐊𝜎subscript𝐊𝐐subscript~𝑉𝑋𝐐subscriptsuperscript^𝑋𝐊𝐐𝜎subscript^𝑋𝐊𝜎\displaystyle H_{X\sigma}=\sum_{\mathbf{K}}\epsilon_{\mathbf{K},X}\hat{X}^{% \dagger}_{\mathbf{K}\sigma}\hat{X}_{\mathbf{K}\sigma}+\sum_{\mathbf{K}\mathbf{% Q}}\tilde{V}_{X}(\mathbf{Q})\hat{X}^{\dagger}_{\mathbf{K}+\mathbf{Q}\sigma}% \hat{X}_{\mathbf{K}\sigma}.italic_H start_POSTSUBSCRIPT italic_X italic_σ end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_K end_POSTSUBSCRIPT italic_ϵ start_POSTSUBSCRIPT bold_K , italic_X end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_K italic_σ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_K italic_σ end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT bold_K bold_Q end_POSTSUBSCRIPT over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT ( bold_Q ) over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_K + bold_Q italic_σ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_K italic_σ end_POSTSUBSCRIPT . (11)

Here, X^𝐊σsubscript^𝑋𝐊𝜎\hat{X}_{\mathbf{K}\sigma}over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_K italic_σ end_POSTSUBSCRIPT annihilates an exciton of type σ𝜎\sigmaitalic_σ with energy ϵ𝐊,X=ϵX+|𝐊|2/2mXsubscriptitalic-ϵ𝐊𝑋subscriptitalic-ϵ𝑋superscript𝐊22subscript𝑚𝑋\epsilon_{\mathbf{K},X}=\epsilon_{X}+|\mathbf{K}|^{2}/2m_{X}italic_ϵ start_POSTSUBSCRIPT bold_K , italic_X end_POSTSUBSCRIPT = italic_ϵ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT + | bold_K | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 italic_m start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT, where mXsubscript𝑚𝑋m_{X}italic_m start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT is the exciton mass and ϵXsubscriptitalic-ϵ𝑋\epsilon_{X}italic_ϵ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT is the 1s exciton energy. The term V~Xsubscript~𝑉𝑋\tilde{V}_{X}over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT is the Fourier transform of the exciton moiré potential, which is approximated in real space by [56]

V(𝐱)=j=16Vjei𝐆j𝐱,𝑉𝐱superscriptsubscript𝑗16subscript𝑉𝑗superscript𝑒𝑖subscript𝐆𝑗𝐱\displaystyle V(\mathbf{x})=\sum_{j=1}^{6}V_{j}e^{i\mathbf{G}_{j}\cdot\mathbf{% x}},italic_V ( bold_x ) = ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_V start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i bold_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⋅ bold_x end_POSTSUPERSCRIPT , (12)

where 𝐆jsubscript𝐆𝑗\mathbf{G}_{j}bold_G start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT are the first six reciprocal lattice vectors. The three-fold symmetry and the realness of the periodic potential imply that V1=V3=V5subscript𝑉1subscript𝑉3subscript𝑉5V_{1}=V_{3}=V_{5}italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_V start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = italic_V start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT, V2=V4=V6subscript𝑉2subscript𝑉4subscript𝑉6V_{2}=V_{4}=V_{6}italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_V start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = italic_V start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT, and V1=V4subscript𝑉1superscriptsubscript𝑉4V_{1}=V_{4}^{*}italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_V start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, parametrized by V1=Veiψsubscript𝑉1𝑉superscript𝑒𝑖𝜓V_{1}=Ve^{i\psi}italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_V italic_e start_POSTSUPERSCRIPT italic_i italic_ψ end_POSTSUPERSCRIPT, with V𝑉Vitalic_V determining the potential depth and ψ𝜓\psiitalic_ψ its shape.

The third term in the Hamiltonian describes the closed-channel molecule, which mediates interactions between excitons:

Hd=𝐊(ϵ𝐊,d+δcc)d^𝐊d^𝐊+𝐊𝐐2V~X(𝐐)d^𝐊+𝐐d^𝐊,subscript𝐻𝑑subscript𝐊subscriptitalic-ϵ𝐊𝑑subscript𝛿ccsubscriptsuperscript^𝑑𝐊subscript^𝑑𝐊subscript𝐊𝐐2subscript~𝑉𝑋𝐐subscriptsuperscript^𝑑𝐊𝐐subscript^𝑑𝐊\displaystyle H_{d}=\sum_{\mathbf{K}}(\epsilon_{\mathbf{K},d}+\delta_{\rm cc})% \hat{d}^{\dagger}_{\mathbf{K}}\hat{d}_{\mathbf{K}}+\sum_{\mathbf{K}\mathbf{Q}}% 2\tilde{V}_{X}(\mathbf{Q})\hat{d}^{\dagger}_{\mathbf{K}+\mathbf{Q}}\hat{d}_{% \mathbf{K}},italic_H start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT bold_K end_POSTSUBSCRIPT ( italic_ϵ start_POSTSUBSCRIPT bold_K , italic_d end_POSTSUBSCRIPT + italic_δ start_POSTSUBSCRIPT roman_cc end_POSTSUBSCRIPT ) over^ start_ARG italic_d end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_K end_POSTSUBSCRIPT over^ start_ARG italic_d end_ARG start_POSTSUBSCRIPT bold_K end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT bold_K bold_Q end_POSTSUBSCRIPT 2 over~ start_ARG italic_V end_ARG start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT ( bold_Q ) over^ start_ARG italic_d end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_K + bold_Q end_POSTSUBSCRIPT over^ start_ARG italic_d end_ARG start_POSTSUBSCRIPT bold_K end_POSTSUBSCRIPT , (13)

where d^𝐊subscript^𝑑𝐊\hat{d}_{\mathbf{K}}over^ start_ARG italic_d end_ARG start_POSTSUBSCRIPT bold_K end_POSTSUBSCRIPT annihilates a closed-channel molecule with energy ϵ𝐊,d=2ϵX+|𝐊|2/2Msubscriptitalic-ϵ𝐊𝑑2subscriptitalic-ϵ𝑋superscript𝐊22𝑀\epsilon_{\mathbf{K},d}=2\epsilon_{X}+|\mathbf{K}|^{2}/2Mitalic_ϵ start_POSTSUBSCRIPT bold_K , italic_d end_POSTSUBSCRIPT = 2 italic_ϵ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT + | bold_K | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 italic_M (M=2mX𝑀2subscript𝑚𝑋M=2m_{X}italic_M = 2 italic_m start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT) and detuning δccsubscript𝛿cc\delta_{\rm cc}italic_δ start_POSTSUBSCRIPT roman_cc end_POSTSUBSCRIPT. The closed-channel experiences both exciton potentials, hence the factor of two in the moiré potential.

The interactions are mediated according to

V=g𝐊𝐐χ(𝐊)(d^𝐐X^𝐐𝐊,X^𝐊,+h.c.),\displaystyle V=g\sum_{\mathbf{K}\mathbf{Q}}\chi(\mathbf{K})(\hat{d}^{\dagger}% _{\mathbf{Q}}\hat{X}_{\mathbf{Q}-\mathbf{K},\uparrow}\hat{X}_{\mathbf{K},% \downarrow}+\rm{h.c.}),italic_V = italic_g ∑ start_POSTSUBSCRIPT bold_K bold_Q end_POSTSUBSCRIPT italic_χ ( bold_K ) ( over^ start_ARG italic_d end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_Q end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_Q - bold_K , ↑ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_K , ↓ end_POSTSUBSCRIPT + roman_h . roman_c . ) , (14)

where χ(𝐊)𝜒𝐊\chi(\mathbf{K})italic_χ ( bold_K ) regularizes the ultraviolet (UV) divergence. Throughout this section we will use χ(𝐊)=Θ(Λ|𝐊|)𝜒𝐊ΘΛ𝐊\chi(\mathbf{K})=\Theta(\Lambda-|\mathbf{K}|)italic_χ ( bold_K ) = roman_Θ ( roman_Λ - | bold_K | ), where ΛΛ\Lambdaroman_Λ is the UV cutoff. The bare coupling g𝑔gitalic_g and detuning δccsubscript𝛿cc\delta_{\rm cc}italic_δ start_POSTSUBSCRIPT roman_cc end_POSTSUBSCRIPT are renormalized as [69]

δccg2=𝐊χ(𝐊)1EBX+2ϵ𝐊,X,subscript𝛿ccsuperscript𝑔2subscript𝐊𝜒𝐊1subscript𝐸BX2subscriptitalic-ϵ𝐊𝑋\displaystyle\frac{\delta_{\rm cc}}{g^{2}}=\sum_{\mathbf{K}}\chi(\mathbf{K})% \frac{1}{E_{\rm BX}+2\epsilon_{\mathbf{K},X}},divide start_ARG italic_δ start_POSTSUBSCRIPT roman_cc end_POSTSUBSCRIPT end_ARG start_ARG italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = ∑ start_POSTSUBSCRIPT bold_K end_POSTSUBSCRIPT italic_χ ( bold_K ) divide start_ARG 1 end_ARG start_ARG italic_E start_POSTSUBSCRIPT roman_BX end_POSTSUBSCRIPT + 2 italic_ϵ start_POSTSUBSCRIPT bold_K , italic_X end_POSTSUBSCRIPT end_ARG , (15)

with EBXsubscript𝐸BXE_{\rm BX}italic_E start_POSTSUBSCRIPT roman_BX end_POSTSUBSCRIPT being the energy of the biexciton without moiré.

The two-channel model introduced above is often employed in the study of ultra-cold atoms. There it is understood as being inspired by the underlying microscopics, whereby the interactions between atoms are mediated by a molecular state (i.e., a Feshbach resonance). However, when taking the so-called single-channel model limit (δcc,gsubscript𝛿cc𝑔\delta_{\rm cc},g\to\inftyitalic_δ start_POSTSUBSCRIPT roman_cc end_POSTSUBSCRIPT , italic_g → ∞), the two-channel model is entirely equivalent to contact interactions with coupling constant U=g2/δcc𝑈superscript𝑔2subscript𝛿ccU=-g^{2}/\delta_{\rm cc}italic_U = - italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_δ start_POSTSUBSCRIPT roman_cc end_POSTSUBSCRIPT, which connects the Hamiltonian presented here to the Hamiltonian in Eq. (II.3). Throughout this work, we will exclusively work in the single-channel model limit since, for our purposes, the two-channel model is only used as a tool to simplify the calculation of the exciton-exciton T𝑇Titalic_T matrix.

H.2 Single-particle Hamiltonians

By Bloch’s theorem we can diagonalize the single-particle Hamiltonians in Eq. (10). For an exciton of type σ𝜎\sigmaitalic_σ, we have

HX,σ|𝐤,λ,σ=E𝐤,λσ|𝐤,λ,σsubscript𝐻𝑋𝜎ket𝐤𝜆𝜎subscriptsuperscript𝐸𝜎𝐤𝜆ket𝐤𝜆𝜎\displaystyle H_{X,\sigma}\ket{\mathbf{k},\lambda,\sigma}=E^{\sigma}_{\mathbf{% k},\lambda}\ket{\mathbf{k},\lambda,\sigma}italic_H start_POSTSUBSCRIPT italic_X , italic_σ end_POSTSUBSCRIPT | start_ARG bold_k , italic_λ , italic_σ end_ARG ⟩ = italic_E start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ end_POSTSUBSCRIPT | start_ARG bold_k , italic_λ , italic_σ end_ARG ⟩ (16)

and for the closed-channel molecule,

Hd|𝐤,λ,d=(E𝐤,λd+δcc)|𝐤,λ,d,subscript𝐻𝑑ket𝐤𝜆𝑑subscriptsuperscript𝐸𝑑𝐤𝜆subscript𝛿ccket𝐤𝜆𝑑\displaystyle H_{d}\ket{\mathbf{k},\lambda,d}=(E^{d}_{\mathbf{k},\lambda}+% \delta_{\rm cc})\ket{\mathbf{k},\lambda,d},italic_H start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT | start_ARG bold_k , italic_λ , italic_d end_ARG ⟩ = ( italic_E start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ end_POSTSUBSCRIPT + italic_δ start_POSTSUBSCRIPT roman_cc end_POSTSUBSCRIPT ) | start_ARG bold_k , italic_λ , italic_d end_ARG ⟩ , (17)

where λ𝜆\lambdaitalic_λ is the band index and 𝐤𝐤\mathbf{k}bold_k is the quasi-momentum. Here and throughout, quasi-momentum is denoted with lowercase letters, while real momentum is denoted with capital letters.

The Bloch states |𝐤,λ,σket𝐤𝜆𝜎\ket{\mathbf{k},\lambda,\sigma}| start_ARG bold_k , italic_λ , italic_σ end_ARG ⟩ and |𝐤,λ,dket𝐤𝜆𝑑\ket{\mathbf{k},\lambda,d}| start_ARG bold_k , italic_λ , italic_d end_ARG ⟩ can be expanded according to

|𝐤,λ,σket𝐤𝜆𝜎\displaystyle\ket{\mathbf{k},\lambda,\sigma}| start_ARG bold_k , italic_λ , italic_σ end_ARG ⟩ =m,nϕ(m,n)(𝐤,λ,σ)X^𝐤+𝐆mn,σ|0absentsubscript𝑚𝑛subscriptsuperscriptitalic-ϕ𝐤𝜆𝜎𝑚𝑛subscriptsuperscript^𝑋𝐤subscript𝐆𝑚𝑛𝜎ket0\displaystyle=\sum_{m,n}\phi^{(\mathbf{k},\lambda,\sigma)}_{(m,n)}\hat{X}^{% \dagger}_{\mathbf{k}+\mathbf{G}_{mn},\sigma}\ket{0}= ∑ start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUPERSCRIPT ( bold_k , italic_λ , italic_σ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_m , italic_n ) end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k + bold_G start_POSTSUBSCRIPT italic_m italic_n end_POSTSUBSCRIPT , italic_σ end_POSTSUBSCRIPT | start_ARG 0 end_ARG ⟩ (18)
|𝐤,λ,dket𝐤𝜆𝑑\displaystyle\ket{\mathbf{k},\lambda,d}| start_ARG bold_k , italic_λ , italic_d end_ARG ⟩ =m,nϕ(m,n)(𝐤,λ,d)d^𝐤+𝐆mn|0absentsubscript𝑚𝑛subscriptsuperscriptitalic-ϕ𝐤𝜆𝑑𝑚𝑛subscriptsuperscript^𝑑𝐤subscript𝐆𝑚𝑛ket0\displaystyle=\sum_{m,n}\phi^{(\mathbf{k},\lambda,d)}_{(m,n)}\hat{d}^{\dagger}% _{\mathbf{k}+\mathbf{G}_{mn}}\ket{0}= ∑ start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUPERSCRIPT ( bold_k , italic_λ , italic_d ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ( italic_m , italic_n ) end_POSTSUBSCRIPT over^ start_ARG italic_d end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k + bold_G start_POSTSUBSCRIPT italic_m italic_n end_POSTSUBSCRIPT end_POSTSUBSCRIPT | start_ARG 0 end_ARG ⟩ (19)

where 𝐆mn=mG^0+nG^1subscript𝐆𝑚𝑛𝑚subscript^𝐺0𝑛subscript^𝐺1\mathbf{G}_{mn}=m\hat{G}_{0}+n\hat{G}_{1}bold_G start_POSTSUBSCRIPT italic_m italic_n end_POSTSUBSCRIPT = italic_m over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_n over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, with G^0subscript^𝐺0\hat{G}_{0}over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and G^1subscript^𝐺1\hat{G}_{1}over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT the moiré reciprocal lattice basis vectors and m𝑚mitalic_m, n𝑛nitalic_n integers. Here the expansion coefficients satisfy

E𝐤,λσϕ𝐦(𝐤,λ,σ)subscriptsuperscript𝐸𝜎𝐤𝜆subscriptsuperscriptitalic-ϕ𝐤𝜆𝜎𝐦\displaystyle E^{\sigma}_{\mathbf{k},\lambda}\phi^{(\mathbf{k},\lambda,\sigma)% }_{\mathbf{m}}italic_E start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ end_POSTSUBSCRIPT italic_ϕ start_POSTSUPERSCRIPT ( bold_k , italic_λ , italic_σ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT =|𝐤+𝐆𝐦|22mXϕ𝐦(𝐤,λ,σ)+𝐦V~(𝐆𝐦)ϕ𝐦+𝐦(𝐤,λ,σ).absentsuperscript𝐤subscript𝐆𝐦22subscript𝑚𝑋subscriptsuperscriptitalic-ϕ𝐤𝜆𝜎𝐦subscriptsuperscript𝐦~𝑉subscript𝐆superscript𝐦subscriptsuperscriptitalic-ϕ𝐤𝜆𝜎𝐦superscript𝐦\displaystyle=\frac{|\mathbf{k}+\mathbf{G}_{\mathbf{m}}|^{2}}{2m_{X}}\phi^{(% \mathbf{k},\lambda,\sigma)}_{\mathbf{m}}+\sum_{\mathbf{m}^{\prime}}\tilde{V}(% \mathbf{G}_{\mathbf{m}^{\prime}})\phi^{(\mathbf{k},\lambda,\sigma)}_{\mathbf{m% }+\mathbf{m}^{\prime}}.= divide start_ARG | bold_k + bold_G start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT end_ARG italic_ϕ start_POSTSUPERSCRIPT ( bold_k , italic_λ , italic_σ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_m end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT bold_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT over~ start_ARG italic_V end_ARG ( bold_G start_POSTSUBSCRIPT bold_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ) italic_ϕ start_POSTSUPERSCRIPT ( bold_k , italic_λ , italic_σ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_m + bold_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT . (20)

where we have set 𝐦=(m,n)𝐦𝑚𝑛\mathbf{m}=(m,n)bold_m = ( italic_m , italic_n ). The equations for the dressed dimmer are almost identical, with σd𝜎𝑑\sigma\to ditalic_σ → italic_d, mXMsubscript𝑚𝑋𝑀m_{X}\to Mitalic_m start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT → italic_M and V~2V~~𝑉2~𝑉\tilde{V}\to 2\tilde{V}over~ start_ARG italic_V end_ARG → 2 over~ start_ARG italic_V end_ARG.

Table 1 summarises the relevant parameters for our system as determined by large-scale Density Functional Theory (DFT) calculations [7]. Here, we also include the exciton moiré parameters which are determined by assuming a tightly bound electron-hole pair such that VXeiψXVeeiψe+Vheiψhsubscript𝑉𝑋superscript𝑒𝑖subscript𝜓𝑋subscript𝑉𝑒superscript𝑒𝑖subscript𝜓𝑒subscript𝑉superscript𝑒𝑖subscript𝜓V_{X}e^{i\psi_{X}}\equiv V_{e}e^{i\psi_{e}}+V_{h}e^{i\psi_{h}}italic_V start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_ψ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ≡ italic_V start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_ψ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_POSTSUPERSCRIPT + italic_V start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_ψ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, where Visubscript𝑉𝑖V_{i}italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and ψisubscript𝜓𝑖\psi_{i}italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are the moiré parameters for the exciton (i=X𝑖𝑋i=Xitalic_i = italic_X), electron (i=e𝑖𝑒i=eitalic_i = italic_e) and hole (i=h𝑖i=hitalic_i = italic_h).

Particle V𝑉Vitalic_V [meVmeV\rm meVroman_meV] ψ𝜓\psiitalic_ψ m𝑚mitalic_m [mesubscript𝑚𝑒m_{e}italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT]
Electron 6.36.3-6.3- 6.3 0superscript00^{\circ}0 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT 0.450.450.450.45
Hole 1.91.91.91.9 59superscript5959^{\circ}59 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT 0.550.550.550.55
Exciton 5.65.65.65.6 163superscript163163^{\circ}163 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT 1111
Table 1: Parameters for electron, hole, and exciton from density functional theory calculations [7]. The mass of the particles is given in units of the bare electron mass mesubscript𝑚𝑒m_{e}italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT. The assumed moiré length is aM=8.2subscript𝑎𝑀8.2a_{M}=8.2italic_a start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT = 8.2 nm.

Using the DFT moiré parameters, we compare the theoretical exciton spectrum to experimental data. In particular, Fig. 12 shows the spectrum at charge neutrality, which qualitatively matches the experimental observations. The lowest two peaks in the spectrum correspond to MX1 and MX2. The exciton spectrum is given by

A(E)=1πλIm[|ϕ𝟎(𝟎,λ,)|2EE𝟎,λ+iγ],𝐴𝐸1𝜋subscript𝜆superscriptsuperscriptsubscriptitalic-ϕ𝟎𝟎𝜆2𝐸subscript𝐸𝟎𝜆𝑖𝛾\displaystyle A(E)=-\frac{1}{\pi}\sum_{\lambda}\imaginary\left[\frac{|\phi_{% \mathbf{0}}^{(\mathbf{0},\lambda,\uparrow)}|^{2}}{E-E_{\mathbf{0},\lambda}+i% \gamma}\right],italic_A ( italic_E ) = - divide start_ARG 1 end_ARG start_ARG italic_π end_ARG ∑ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_OPERATOR roman_Im end_OPERATOR [ divide start_ARG | italic_ϕ start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( bold_0 , italic_λ , ↑ ) end_POSTSUPERSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_E - italic_E start_POSTSUBSCRIPT bold_0 , italic_λ end_POSTSUBSCRIPT + italic_i italic_γ end_ARG ] , (21)

which is proportional to the reflectance measured in experiment. Here, γ=1𝛾1\gamma=1italic_γ = 1 meV is the inverse lifetime of the exciton, which we set to roughly match experiment. It is noteworthy that the DFT moiré parameters underestimate the separation between MX1 and MX2 by approximately 10 meV. We remind the reader, as shown in Fig. 3(g), there exists dark excitonic states between MX1 and MX2 (and above), which can play a significant role in the biexciton spectrum.

Refer to caption
Figure 12: Exciton spectral function at zero doping using the DFT parameters in Table 1. In agreement with experiment, the theory predicts two dominant excitonic resonances, which we identify as MX1 and MX2. We have included a broadening from the finite exciton lifetime of γ=1𝛾1\gamma=1italic_γ = 1 meV.

To conclude this section, we use the Bloch states introduced above to recast the two-channel interactions. Noting that the interactions will conserve quasi-momentum, the relevant matrix elements of the interactions are

𝒱𝐤λλ𝐪λd𝐪,λd,d|V|𝐪𝐤,λ,;𝐤,λ,/g2subscriptsuperscript𝒱𝐪subscript𝜆𝑑𝐤subscript𝜆subscript𝜆bra𝐪subscript𝜆𝑑𝑑𝑉ket𝐪𝐤subscript𝜆𝐤subscript𝜆superscript𝑔2\displaystyle\mathcal{V}^{\mathbf{q}\lambda_{d}}_{\mathbf{k}\lambda_{\uparrow}% \lambda_{\downarrow}}\equiv\bra{\mathbf{q},\lambda_{d},d}V\ket{\mathbf{q}-% \mathbf{k},\lambda_{\uparrow},\uparrow;\mathbf{k},\lambda_{\downarrow},% \downarrow}/g^{2}caligraphic_V start_POSTSUPERSCRIPT bold_q italic_λ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT ≡ ⟨ start_ARG bold_q , italic_λ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT , italic_d end_ARG | italic_V | start_ARG bold_q - bold_k , italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , ↑ ; bold_k , italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT , ↓ end_ARG ⟩ / italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
=𝐍𝐜𝐍𝐫χ(𝐤+𝐍𝐜/2𝐍𝐫)ϕ𝐍𝐜(𝐪,λd,d)ϕ𝐍𝐜/2+𝐍𝐫(𝐪𝐤,λ,)ϕ𝐍𝐜/2𝐍𝐫(𝐤,λ,).absentsubscript𝐍𝐜𝐍𝐫𝜒𝐤𝐍𝐜2𝐍𝐫subscriptsuperscriptitalic-ϕ𝐪subscript𝜆𝑑𝑑𝐍𝐜subscriptsuperscriptitalic-ϕ𝐪𝐤subscript𝜆𝐍𝐜2𝐍𝐫subscriptsuperscriptitalic-ϕ𝐤subscript𝜆𝐍𝐜2𝐍𝐫\displaystyle{}\quad=\sum_{\mathbf{N_{c}}\mathbf{N_{r}}}\chi(\mathbf{k}+% \mathbf{N_{c}}/2-\mathbf{N_{r}})\phi^{(\mathbf{q},\lambda_{d},d)*}_{\mathbf{N_% {c}}}\phi^{(\mathbf{q}-\mathbf{k},\lambda_{\uparrow},\uparrow)}_{\mathbf{N_{c}% }/2+\mathbf{N_{r}}}\phi^{(\mathbf{k},\lambda_{\downarrow},\downarrow)}_{% \mathbf{N_{c}}/2-\mathbf{N_{r}}}.= ∑ start_POSTSUBSCRIPT start_ID bold_N start_POSTSUBSCRIPT bold_c end_POSTSUBSCRIPT end_ID start_ID bold_N start_POSTSUBSCRIPT bold_r end_POSTSUBSCRIPT end_ID end_POSTSUBSCRIPT italic_χ ( bold_k + start_ID bold_N start_POSTSUBSCRIPT bold_c end_POSTSUBSCRIPT end_ID / 2 - start_ID bold_N start_POSTSUBSCRIPT bold_r end_POSTSUBSCRIPT end_ID ) italic_ϕ start_POSTSUPERSCRIPT ( bold_q , italic_λ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT , italic_d ) ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_N start_POSTSUBSCRIPT bold_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ϕ start_POSTSUPERSCRIPT ( bold_q - bold_k , italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , ↑ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT start_ID bold_N start_POSTSUBSCRIPT bold_c end_POSTSUBSCRIPT end_ID / 2 + start_ID bold_N start_POSTSUBSCRIPT bold_r end_POSTSUBSCRIPT end_ID end_POSTSUBSCRIPT italic_ϕ start_POSTSUPERSCRIPT ( bold_k , italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT , ↓ ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT start_ID bold_N start_POSTSUBSCRIPT bold_c end_POSTSUBSCRIPT end_ID / 2 - start_ID bold_N start_POSTSUBSCRIPT bold_r end_POSTSUBSCRIPT end_ID end_POSTSUBSCRIPT . (22)

We point out that since the UV cutoff is large, we can take χ(𝐤+𝐍𝐜/2𝐍𝐫)χ(𝐍𝐫)𝜒𝐤𝐍𝐜2𝐍𝐫𝜒𝐍𝐫\chi(\mathbf{k}+\mathbf{N_{c}}/2-\mathbf{N_{r}})\to\chi(\mathbf{N_{r}})italic_χ ( bold_k + start_ID bold_N start_POSTSUBSCRIPT bold_c end_POSTSUBSCRIPT end_ID / 2 - start_ID bold_N start_POSTSUBSCRIPT bold_r end_POSTSUBSCRIPT end_ID ) → italic_χ ( start_ID bold_N start_POSTSUBSCRIPT bold_r end_POSTSUBSCRIPT end_ID ), where 𝐍𝐫𝐍𝐫\mathbf{N_{r}}bold_N start_POSTSUBSCRIPT bold_r end_POSTSUBSCRIPT can be understood as the relative reciprocal lattice vector of the two excitons.

H.3 T Matrix

To study the interactions we calculate the T𝑇Titalic_T matrix, which provides an exact solution to the full two-body problem. To begin we introduce the free exciton and closed-channel Green’s functions:

G^(0)(E)superscript^𝐺0𝐸\displaystyle\hat{G}^{(0)}(E)over^ start_ARG italic_G end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT ( italic_E ) =1EH^XH^X,absent1𝐸subscript^𝐻𝑋absentsubscript^𝐻𝑋\displaystyle=\frac{1}{E-\hat{H}_{X\uparrow}-\hat{H}_{X,\downarrow}}= divide start_ARG 1 end_ARG start_ARG italic_E - over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_X ↑ end_POSTSUBSCRIPT - over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_X , ↓ end_POSTSUBSCRIPT end_ARG (23)
D^(0)(E)superscript^𝐷0𝐸\displaystyle\hat{D}^{(0)}(E)over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT ( italic_E ) =1EH^d.absent1𝐸subscript^𝐻𝑑\displaystyle=\frac{1}{E-\hat{H}_{d}}.= divide start_ARG 1 end_ARG start_ARG italic_E - over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_ARG . (24)

The two-body T𝑇Titalic_T matrix is given by the infinite series

T𝑇\displaystyle Titalic_T =V^D^(0)V^+V^D^(0)V^G^(0)V^D^(0)V^+absent^𝑉superscript^𝐷0^𝑉^𝑉superscript^𝐷0^𝑉superscript^𝐺0^𝑉superscript^𝐷0^𝑉\displaystyle=\hat{V}\hat{D}^{(0)}\hat{V}+\hat{V}\hat{D}^{(0)}\hat{V}\hat{G}^{% (0)}\hat{V}\hat{D}^{(0)}\hat{V}+\dots= over^ start_ARG italic_V end_ARG over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT over^ start_ARG italic_V end_ARG + over^ start_ARG italic_V end_ARG over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT over^ start_ARG italic_V end_ARG over^ start_ARG italic_G end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT over^ start_ARG italic_V end_ARG over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT over^ start_ARG italic_V end_ARG + …
=V^(D^(0)+D^(0)V^G^(0)V^D^(0)+)V^absent^𝑉superscript^𝐷0superscript^𝐷0^𝑉superscript^𝐺0^𝑉superscript^𝐷0^𝑉\displaystyle=\hat{V}\left(\hat{D}^{(0)}+\hat{D}^{(0)}\hat{V}\hat{G}^{(0)}\hat% {V}\hat{D}^{(0)}+\dots\right)\hat{V}= over^ start_ARG italic_V end_ARG ( over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT + over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT over^ start_ARG italic_V end_ARG over^ start_ARG italic_G end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT over^ start_ARG italic_V end_ARG over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT + … ) over^ start_ARG italic_V end_ARG
=V^D^V^absent^𝑉^𝐷^𝑉\displaystyle=\hat{V}\hat{D}\hat{V}= over^ start_ARG italic_V end_ARG over^ start_ARG italic_D end_ARG over^ start_ARG italic_V end_ARG (25)

where we have suppressed the energy dependencies for brevity and D^(E)^𝐷𝐸\hat{D}(E)over^ start_ARG italic_D end_ARG ( italic_E ) is the closed-channel Green’s function. Thus, by finding the closed-channel Green’s function, we can immediately calculate the T𝑇Titalic_T matrix. This approach simplifies the calculation and provides easier access to the biexciton energies, which are both the poles of the T𝑇Titalic_T matrix and D^(E)^𝐷𝐸\hat{D}(E)over^ start_ARG italic_D end_ARG ( italic_E ).

The closed-channel Green’s function is given by

D^(E)^𝐷𝐸\displaystyle\hat{D}(E)over^ start_ARG italic_D end_ARG ( italic_E ) =D^(0)+D^(0)V^G^(0)V^D^(0)+absentsuperscript^𝐷0superscript^𝐷0^𝑉superscript^𝐺0^𝑉superscript^𝐷0\displaystyle=\hat{D}^{(0)}+\hat{D}^{(0)}\hat{V}\hat{G}^{(0)}\hat{V}\hat{D}^{(% 0)}+\dots= over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT + over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT over^ start_ARG italic_V end_ARG over^ start_ARG italic_G end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT over^ start_ARG italic_V end_ARG over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT + … (26a)
=1[D^(0)(E)]1g2Π^(E),absent1superscriptdelimited-[]superscript^𝐷0𝐸1superscript𝑔2^Π𝐸\displaystyle=\frac{1}{[\hat{D}^{(0)}(E)]^{-1}-g^{2}\hat{\Pi}(E)},= divide start_ARG 1 end_ARG start_ARG [ over^ start_ARG italic_D end_ARG start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT ( italic_E ) ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT - italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over^ start_ARG roman_Π end_ARG ( italic_E ) end_ARG , (26b)

where we have introduce the polarization bubble, which has matrix elements

Πλλ𝐪=subscriptsuperscriptΠ𝐪𝜆superscript𝜆absent\displaystyle\Pi^{\mathbf{q}}_{\lambda\lambda^{\prime}}=roman_Π start_POSTSUPERSCRIPT bold_q end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT = 𝐤,λ,λ𝒱𝐤λλ𝐪,λ1EE𝐤,𝐪;λ,λ𝒱𝐤λλ𝐪,λ,subscript𝐤subscript𝜆subscript𝜆subscriptsuperscript𝒱𝐪𝜆𝐤subscript𝜆subscript𝜆1𝐸subscript𝐸𝐤𝐪subscript𝜆subscript𝜆subscriptsuperscript𝒱𝐪superscript𝜆𝐤subscript𝜆subscript𝜆\displaystyle\sum_{\mathbf{k},\lambda_{\uparrow},\lambda_{\downarrow}}\mathcal% {V}^{\mathbf{q},\lambda}_{\mathbf{k}\lambda_{\uparrow}\lambda_{\downarrow}}% \frac{1}{E-E_{\mathbf{k},\mathbf{q};\lambda_{\uparrow},\lambda_{\downarrow}}}% \mathcal{V}^{\mathbf{q},\lambda^{\prime}*}_{\mathbf{k}\lambda_{\uparrow}% \lambda_{\downarrow}},∑ start_POSTSUBSCRIPT bold_k , italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT caligraphic_V start_POSTSUPERSCRIPT bold_q , italic_λ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG italic_E - italic_E start_POSTSUBSCRIPT bold_k , bold_q ; italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT end_ARG caligraphic_V start_POSTSUPERSCRIPT bold_q , italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT , (27)

where E𝐤,𝐪;λ,λE𝐪𝐤,λ+E𝐤,λsubscript𝐸𝐤𝐪subscript𝜆subscript𝜆subscriptsuperscript𝐸𝐪𝐤subscript𝜆subscriptsuperscript𝐸𝐤subscript𝜆E_{\mathbf{k},\mathbf{q};\lambda_{\uparrow},\lambda_{\downarrow}}\equiv E^{% \uparrow}_{\mathbf{q}-\mathbf{k},\lambda_{\uparrow}}+E^{\downarrow}_{\mathbf{k% },\lambda_{\downarrow}}italic_E start_POSTSUBSCRIPT bold_k , bold_q ; italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT ≡ italic_E start_POSTSUPERSCRIPT ↑ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_q - bold_k , italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT end_POSTSUBSCRIPT + italic_E start_POSTSUPERSCRIPT ↓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT.

We note that the polarization bubble calculation is numerically demanding due to the need to contract large tensors. In order to optimize the calculation, we reduce the size of interaction tensor 𝒱𝐤λλ𝐪,λsubscriptsuperscript𝒱𝐪𝜆𝐤subscript𝜆subscript𝜆\mathcal{V}^{\mathbf{q},\lambda}_{\mathbf{k}\lambda_{\uparrow}\lambda_{% \downarrow}}caligraphic_V start_POSTSUPERSCRIPT bold_q , italic_λ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT by eliminating all elements that are smaller than a specified tolerance. We have checked that this tolerance does not affect the final calculation.

H.4 Effect of the pump laser

To incorporate the effects of the pump laser, which generates \downarrow excitons in our notation, we consider the following modification to the Hamiltonian H^Xsubscript^𝐻𝑋absent\hat{H}_{X\downarrow}over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_X ↓ end_POSTSUBSCRIPT[21]

HXsubscript𝐻𝑋absent\displaystyle H_{X\downarrow}italic_H start_POSTSUBSCRIPT italic_X ↓ end_POSTSUBSCRIPT =𝐤,λE𝐤,λX^𝐤,λ,X^𝐤,λ,+ΩeiωLtλϕ𝟎(𝟎,λ,)X^𝟎,λ,absentsubscript𝐤𝜆subscriptsuperscript𝐸𝐤𝜆subscriptsuperscript^𝑋𝐤𝜆subscript^𝑋𝐤𝜆Ωsuperscriptsuperscript𝑒𝑖subscript𝜔𝐿𝑡subscript𝜆superscriptsubscriptitalic-ϕ𝟎𝟎𝜆subscriptsuperscript^𝑋𝟎𝜆\displaystyle=\sum_{\mathbf{k},\lambda}E^{\downarrow}_{\mathbf{k},\lambda}\hat% {X}^{\dagger}_{\mathbf{k},\lambda,\downarrow}\hat{X}_{\mathbf{k},\lambda,% \downarrow}+\Omega{}^{*}e^{i\omega_{L}t}\sum_{\lambda}\phi_{\mathbf{0}}^{(% \mathbf{0},\lambda,\downarrow)*}\hat{X}^{\dagger}_{\mathbf{0},\lambda,\downarrow}= ∑ start_POSTSUBSCRIPT bold_k , italic_λ end_POSTSUBSCRIPT italic_E start_POSTSUPERSCRIPT ↓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ , ↓ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_k , italic_λ , ↓ end_POSTSUBSCRIPT + roman_Ω start_FLOATSUPERSCRIPT ∗ end_FLOATSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT italic_t end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( bold_0 , italic_λ , ↓ ) ∗ end_POSTSUPERSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 , italic_λ , ↓ end_POSTSUBSCRIPT
+ΩeiωLtλϕ𝟎(𝟎,λ,)X^𝟎,λ,.Ωsuperscript𝑒𝑖subscript𝜔𝐿𝑡subscript𝜆superscriptsubscriptitalic-ϕ𝟎𝟎𝜆subscript^𝑋𝟎𝜆\displaystyle{}\qquad+\Omega e^{-i\omega_{L}t}\sum_{\lambda}\phi_{\mathbf{0}}^% {(\mathbf{0},\lambda,\downarrow)}\hat{X}_{\mathbf{0},\lambda,\downarrow}.+ roman_Ω italic_e start_POSTSUPERSCRIPT - italic_i italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT italic_t end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( bold_0 , italic_λ , ↓ ) end_POSTSUPERSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_0 , italic_λ , ↓ end_POSTSUBSCRIPT . (28)

Here, ΩΩ\Omegaroman_Ω represents the light-matter interaction strength; we assume that the light couples only to the zero-momentum exciton and use the rotating wave approximation (|ϵXωL|ϵX+ωLmuch-less-thansubscriptitalic-ϵ𝑋subscript𝜔𝐿subscriptitalic-ϵ𝑋subscript𝜔𝐿|\epsilon_{X}-\omega_{L}|\ll\epsilon_{X}+\omega_{L}| italic_ϵ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT | ≪ italic_ϵ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT). To remove the time-dependence, we apply the unitary transformation

U^(t)=exp(iωL𝐤,λX^𝐤,λ,X^𝐤,λ,)^𝑈𝑡𝑖subscript𝜔𝐿subscript𝐤𝜆subscriptsuperscript^𝑋𝐤𝜆subscript^𝑋𝐤𝜆\displaystyle\hat{U}(t)=\exp(-i\omega_{L}\sum_{\mathbf{k},\lambda}\hat{X}^{% \dagger}_{\mathbf{k},\lambda,\downarrow}\hat{X}_{\mathbf{k},\lambda,\downarrow})over^ start_ARG italic_U end_ARG ( italic_t ) = roman_exp ( start_ARG - italic_i italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT bold_k , italic_λ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ , ↓ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_k , italic_λ , ↓ end_POSTSUBSCRIPT end_ARG ) (29)

to move into the rotating frame. This gives us the Hamiltonian

H~Xsubscript~𝐻𝑋absent\displaystyle\tilde{H}_{X\downarrow}over~ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_X ↓ end_POSTSUBSCRIPT =𝐤,λ(E𝐤,λωL)X^𝐤,λ,X^𝐤,λ,absentsubscript𝐤𝜆subscriptsuperscript𝐸𝐤𝜆subscript𝜔𝐿subscriptsuperscript^𝑋𝐤𝜆subscript^𝑋𝐤𝜆\displaystyle=\sum_{\mathbf{k},\lambda}\left(E^{\downarrow}_{\mathbf{k},% \lambda}-\omega_{L}\right)\hat{X}^{\dagger}_{\mathbf{k},\lambda,\downarrow}% \hat{X}_{\mathbf{k},\lambda,\downarrow}= ∑ start_POSTSUBSCRIPT bold_k , italic_λ end_POSTSUBSCRIPT ( italic_E start_POSTSUPERSCRIPT ↓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ) over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ , ↓ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_k , italic_λ , ↓ end_POSTSUBSCRIPT
+Ωλϕ𝟎(𝟎,λ,)X^𝟎,λ,+Ωλϕ𝟎(𝟎,λ,)X^𝟎,λ,.Ωsuperscriptsubscript𝜆superscriptsubscriptitalic-ϕ𝟎𝟎𝜆subscriptsuperscript^𝑋𝟎𝜆Ωsubscript𝜆superscriptsubscriptitalic-ϕ𝟎𝟎𝜆subscript^𝑋𝟎𝜆\displaystyle+\Omega{}^{*}\sum_{\lambda}\phi_{\mathbf{0}}^{(\mathbf{0},\lambda% ,\downarrow)*}\hat{X}^{\dagger}_{\mathbf{0},\lambda,\downarrow}+\Omega\sum_{% \lambda}\phi_{\mathbf{0}}^{(\mathbf{0},\lambda,\downarrow)}\hat{X}_{\mathbf{0}% ,\lambda,\downarrow}.+ roman_Ω start_FLOATSUPERSCRIPT ∗ end_FLOATSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( bold_0 , italic_λ , ↓ ) ∗ end_POSTSUPERSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 , italic_λ , ↓ end_POSTSUBSCRIPT + roman_Ω ∑ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( bold_0 , italic_λ , ↓ ) end_POSTSUPERSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_0 , italic_λ , ↓ end_POSTSUBSCRIPT .

This Hamiltonian can be diagonalized using the multi-mode shift operator,

D¯^(β)=exp[λ(βλX^𝟎,λ,βλX^𝟎,λ,)],^¯𝐷𝛽subscript𝜆subscript𝛽𝜆subscriptsuperscript^𝑋𝟎𝜆superscriptsubscript𝛽𝜆subscript^𝑋𝟎𝜆\displaystyle\hat{\underline{D}}(\beta)=\exp[\sum_{\lambda}\left(\beta_{% \lambda}\hat{X}^{\dagger}_{\mathbf{0},\lambda,\downarrow}-\beta_{\lambda}^{*}% \hat{X}_{\mathbf{0},\lambda,\downarrow}\right)],over^ start_ARG under¯ start_ARG italic_D end_ARG end_ARG ( italic_β ) = roman_exp [ ∑ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_β start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 , italic_λ , ↓ end_POSTSUBSCRIPT - italic_β start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_0 , italic_λ , ↓ end_POSTSUBSCRIPT ) ] , (30)

which has the property that D¯^(β)X^𝟎,λ,D¯^(β)=X^𝟎,λ,βλ^¯𝐷𝛽subscript^𝑋𝟎𝜆superscript^¯𝐷𝛽subscript^𝑋𝟎𝜆subscript𝛽𝜆\hat{\underline{D}}(\beta)\hat{X}_{\mathbf{0},\lambda,\downarrow}\hat{% \underline{D}}^{\dagger}(\beta)=\hat{X}_{\mathbf{0},\lambda,\downarrow}-\beta_% {\lambda}over^ start_ARG under¯ start_ARG italic_D end_ARG end_ARG ( italic_β ) over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_0 , italic_λ , ↓ end_POSTSUBSCRIPT over^ start_ARG under¯ start_ARG italic_D end_ARG end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ( italic_β ) = over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_0 , italic_λ , ↓ end_POSTSUBSCRIPT - italic_β start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT. The diagonalized Hamiltonian is

H~Xsubscript~𝐻𝑋absent\displaystyle\tilde{H}_{X\downarrow}over~ start_ARG italic_H end_ARG start_POSTSUBSCRIPT italic_X ↓ end_POSTSUBSCRIPT =λ(E𝟎,λωL)(X^𝟎,λ,βλ)(X^𝟎,λ,βλ)absentsubscript𝜆subscriptsuperscript𝐸𝟎𝜆subscript𝜔𝐿subscriptsuperscript^𝑋𝟎𝜆subscriptsuperscript𝛽𝜆subscript^𝑋𝟎𝜆subscript𝛽𝜆\displaystyle=\sum_{\lambda}\left(E^{\downarrow}_{\mathbf{0},\lambda}-\omega_{% L}\right)\left(\hat{X}^{\dagger}_{\mathbf{0},\lambda,\downarrow}-\beta^{*}_{% \lambda}\right)\left(\hat{X}_{\mathbf{0},\lambda,\downarrow}-\beta_{\lambda}\right)= ∑ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_E start_POSTSUPERSCRIPT ↓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 , italic_λ end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ) ( over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 , italic_λ , ↓ end_POSTSUBSCRIPT - italic_β start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ) ( over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_0 , italic_λ , ↓ end_POSTSUBSCRIPT - italic_β start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT )
+𝐤𝟎,λ(E𝐤,λωL)X^𝐤,λ,X^𝐤,λ,+E0,subscript𝐤𝟎𝜆subscriptsuperscript𝐸𝐤𝜆subscript𝜔𝐿subscriptsuperscript^𝑋𝐤𝜆subscript^𝑋𝐤𝜆subscript𝐸0\displaystyle{}\quad+\sum_{\mathbf{k}\neq\mathbf{0},\lambda}\left(E^{% \downarrow}_{\mathbf{k},\lambda}-\omega_{L}\right)\hat{X}^{\dagger}_{\mathbf{k% },\lambda,\downarrow}\hat{X}_{\mathbf{k},\lambda,\downarrow}+E_{0},+ ∑ start_POSTSUBSCRIPT bold_k ≠ bold_0 , italic_λ end_POSTSUBSCRIPT ( italic_E start_POSTSUPERSCRIPT ↓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ) over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ , ↓ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_k , italic_λ , ↓ end_POSTSUBSCRIPT + italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , (31)

where βλ=ϕ𝟎(𝟎,λ,)Ω/(E𝟎,λωL)subscript𝛽𝜆superscriptsubscriptitalic-ϕ𝟎𝟎𝜆superscriptΩsubscriptsuperscript𝐸𝟎𝜆subscript𝜔𝐿\beta_{\lambda}=-\phi_{\mathbf{0}}^{(\mathbf{0},\lambda,\downarrow)*}\Omega^{*% }/(E^{\downarrow}_{\mathbf{0},\lambda}-\omega_{L})italic_β start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = - italic_ϕ start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( bold_0 , italic_λ , ↓ ) ∗ end_POSTSUPERSCRIPT roman_Ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT / ( italic_E start_POSTSUPERSCRIPT ↓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 , italic_λ end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ) and

E0=λ[(E𝟎,λωL)|βλ|2+2Re(Ωϕ𝟎(𝟎,λ,))]subscript𝐸0subscript𝜆delimited-[]subscriptsuperscript𝐸𝟎𝜆subscript𝜔𝐿superscriptsubscript𝛽𝜆22Ωsuperscriptsubscriptitalic-ϕ𝟎𝟎𝜆\displaystyle E_{0}=\sum_{\lambda}\left[(E^{\downarrow}_{\mathbf{0},\lambda}-% \omega_{L})|\beta_{\lambda}|^{2}+2\real\left(\Omega\,\phi_{\mathbf{0}}^{(% \mathbf{0},\lambda,\downarrow)}\right)\right]italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT [ ( italic_E start_POSTSUPERSCRIPT ↓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 , italic_λ end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ) | italic_β start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 start_OPERATOR roman_Re end_OPERATOR ( roman_Ω italic_ϕ start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( bold_0 , italic_λ , ↓ ) end_POSTSUPERSCRIPT ) ] (32)

is the ground state energy. Enforcing that all (X^𝐤,λ,δ𝐤,𝟎βλ)subscript^𝑋𝐤𝜆subscript𝛿𝐤𝟎subscript𝛽𝜆\left(\hat{X}_{\mathbf{k},\lambda,\downarrow}-\delta_{\mathbf{k},\mathbf{0}}% \beta_{\lambda}\right)( over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_k , italic_λ , ↓ end_POSTSUBSCRIPT - italic_δ start_POSTSUBSCRIPT bold_k , bold_0 end_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ) annihilate the ground state (|ΦketΦ\ket{\Phi}| start_ARG roman_Φ end_ARG ⟩), we find |Φ=D¯^(β)|0ketΦ^¯𝐷𝛽ket0\ket{\Phi}=\hat{\underline{D}}(\beta)\ket{0}| start_ARG roman_Φ end_ARG ⟩ = over^ start_ARG under¯ start_ARG italic_D end_ARG end_ARG ( italic_β ) | start_ARG 0 end_ARG ⟩.

Through this analysis, we see that a classical treatment of the laser can be accounted for by substituting X^𝟎,λ,X^𝟎,λ,βλsubscript^𝑋𝟎𝜆subscript^𝑋𝟎𝜆subscript𝛽𝜆\hat{X}_{\mathbf{0},\lambda,\downarrow}\to\hat{X}_{\mathbf{0},\lambda,% \downarrow}-\beta_{\lambda}over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_0 , italic_λ , ↓ end_POSTSUBSCRIPT → over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_0 , italic_λ , ↓ end_POSTSUBSCRIPT - italic_β start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT and E𝐤,λE𝐤,λωLsubscriptsuperscript𝐸𝐤𝜆subscriptsuperscript𝐸𝐤𝜆subscript𝜔𝐿E^{\downarrow}_{\mathbf{k},\lambda}\to E^{\downarrow}_{\mathbf{k},\lambda}-% \omega_{L}italic_E start_POSTSUPERSCRIPT ↓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ end_POSTSUBSCRIPT → italic_E start_POSTSUPERSCRIPT ↓ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT. In principle, moving into the rotating frame causes the interaction in Eq. (14) to develop a time-dependence. However, we are free to cancel this time-dependence with a counter-rotating frame for the \uparrow exciton terms, (i.e., \downarrow\to\uparrow↓ → ↑ and ωLωLsubscript𝜔𝐿subscript𝜔𝐿\omega_{L}\to-\omega_{L}italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT → - italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT in Eq. (29)). This will of course also shift the \uparrow exciton spectrum by the constant ωLsubscript𝜔𝐿\omega_{L}italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT, but since we are only interested in physics measured relative to \uparrow spectrum, this is not necessary to include.

H.5 Dressing probe exciton by pump excitons

We now consider how excitons generated by the pump (denoted by \downarrow) dress a single probe exciton (\uparrow). This is analogous to the theory of the Bose polaron, where we now effectively have multiple non-interacting Bose-Einstein condensates of \downarrow excitons in various zero quasi-momentum bands. Within the framework of non-self-consistent T-matrix theory, the self energy of the probe exciton is given by

Σλλ(E;ωL)=λλβλTλλ;𝟎λλ;𝟎(E+ωL+2iγ)βλ,subscriptΣsubscript𝜆superscriptsubscript𝜆𝐸subscript𝜔𝐿subscriptsubscript𝜆superscriptsubscript𝜆subscript𝛽subscript𝜆subscriptsuperscript𝑇subscript𝜆subscript𝜆𝟎superscriptsubscript𝜆subscriptsuperscript𝜆𝟎𝐸subscript𝜔𝐿2𝑖𝛾subscriptsuperscript𝛽superscriptsubscript𝜆\displaystyle\Sigma_{\lambda_{\uparrow}\lambda_{\uparrow}^{\prime}}(E;\omega_{% L})=\sum_{\lambda_{\downarrow}\lambda_{\downarrow}^{\prime}}\beta_{\lambda_{% \downarrow}}T^{\lambda_{\uparrow}\lambda_{\downarrow};\mathbf{0}}_{\lambda_{% \uparrow}^{\prime}\lambda^{\prime}_{\downarrow};\mathbf{0}}(E+\omega_{L}+2i% \gamma)\beta^{*}_{\lambda_{\downarrow}^{\prime}},roman_Σ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_E ; italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ) = ∑ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_β start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_T start_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ; bold_0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ; bold_0 end_POSTSUBSCRIPT ( italic_E + italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT + 2 italic_i italic_γ ) italic_β start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT , (33)

where Tλλ;𝟎λλ;𝟎(E)𝟎,λ,;𝟎,λ,|T^(E)|𝟎,λ,;𝟎,λ,subscriptsuperscript𝑇subscript𝜆subscript𝜆𝟎superscriptsubscript𝜆subscriptsuperscript𝜆𝟎𝐸bra𝟎subscript𝜆𝟎subscript𝜆^𝑇𝐸ket𝟎superscriptsubscript𝜆𝟎superscriptsubscript𝜆T^{\lambda_{\uparrow}\lambda_{\downarrow};\mathbf{0}}_{\lambda_{\uparrow}^{% \prime}\lambda^{\prime}_{\downarrow};\mathbf{0}}(E)\equiv\bra{\mathbf{0},% \lambda_{\uparrow},\uparrow;\mathbf{0},\lambda_{\downarrow},\downarrow}\hat{T}% (E)\ket{\mathbf{0},\lambda_{\uparrow}^{\prime},\uparrow;\mathbf{0},\lambda_{% \downarrow}^{\prime},\downarrow}italic_T start_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ; bold_0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ; bold_0 end_POSTSUBSCRIPT ( italic_E ) ≡ ⟨ start_ARG bold_0 , italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , ↑ ; bold_0 , italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT , ↓ end_ARG | over^ start_ARG italic_T end_ARG ( italic_E ) | start_ARG bold_0 , italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , ↑ ; bold_0 , italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , ↓ end_ARG ⟩ and both excitons contribute a term iγ𝑖𝛾i\gammaitalic_i italic_γ due to their finite lifetimes. Here, we explicitly include the laser frequency ωLsubscript𝜔𝐿\omega_{L}italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT as a parameter in the self energy to emphasize its significance: we observe that the laser light shifts the scattering off-shell, similar to the effect discussed in the context of polariton-electron scattering [70].

Non-self-consistent T𝑇Titalic_T matrix theory is a non-perturbative approach in that it sums over all ladder diagrams, capturing an infinite series of scattering events to account for strong interactions. Alternatively, the non-self-consistent T𝑇Titalic_T matrix theory can be understood as entirely equivalent to a variational approach based on a generalization of the celebrated Chevy ansatz [71]:

|ΨketΨ\displaystyle\ket{\Psi}| start_ARG roman_Ψ end_ARG ⟩ =λψλ(0)X^𝟎,λ,|Φ+λψλ(0;d)d^𝟎,λ|Φabsentsubscriptsubscript𝜆subscriptsuperscript𝜓0subscript𝜆subscriptsuperscript^𝑋𝟎subscript𝜆ketΦsubscript𝜆superscriptsubscript𝜓𝜆0𝑑subscriptsuperscript^𝑑𝟎𝜆ketΦ\displaystyle=\sum_{\lambda_{\uparrow}}\psi^{(0)}_{\lambda_{\uparrow}}\hat{X}^% {\dagger}_{\mathbf{0},\lambda_{\uparrow},\uparrow}\ket{\Phi}+\sum_{\lambda}% \psi_{\lambda}^{(0;d)}\hat{d}^{\dagger}_{\mathbf{0},\lambda}\ket{\Phi}= ∑ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ψ start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 , italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , ↑ end_POSTSUBSCRIPT | start_ARG roman_Φ end_ARG ⟩ + ∑ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT italic_ψ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 0 ; italic_d ) end_POSTSUPERSCRIPT over^ start_ARG italic_d end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 , italic_λ end_POSTSUBSCRIPT | start_ARG roman_Φ end_ARG ⟩
𝐤λλψ𝐤λλ(1)X^𝐤,λ,X^𝐤,λ,|Φ.subscript𝐤subscript𝜆subscript𝜆subscriptsuperscript𝜓1𝐤subscript𝜆subscript𝜆subscriptsuperscript^𝑋𝐤subscript𝜆subscriptsuperscript^𝑋𝐤subscript𝜆ketΦ\displaystyle{}\qquad\sum_{\mathbf{k}\lambda_{\uparrow}\lambda_{\downarrow}}% \psi^{(1)}_{\mathbf{k}\lambda_{\uparrow}\lambda_{\downarrow}}\hat{X}^{\dagger}% _{\mathbf{k},\lambda_{\uparrow},\uparrow}\hat{X}^{\dagger}_{-\mathbf{k},% \lambda_{\downarrow},\downarrow}\ket{\Phi}.∑ start_POSTSUBSCRIPT bold_k italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ψ start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_k , italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , ↑ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - bold_k , italic_λ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT , ↓ end_POSTSUBSCRIPT | start_ARG roman_Φ end_ARG ⟩ . (34)

By deriving the variational equations, one can show that they can be rearranged into the form

Eψλ(0)=E𝟎,λψλ(0)+λΣλ,λ(E;ωL)ψλ(0),𝐸subscriptsuperscript𝜓0subscript𝜆subscript𝐸𝟎subscript𝜆subscriptsuperscript𝜓0subscript𝜆subscriptsuperscriptsubscript𝜆subscriptΣsubscript𝜆superscriptsubscript𝜆𝐸subscript𝜔𝐿subscriptsuperscript𝜓0superscriptsubscript𝜆\displaystyle E\psi^{(0)}_{\lambda_{\uparrow}}=E_{\mathbf{0},\lambda_{\uparrow% }}\psi^{(0)}_{\lambda_{\uparrow}}+\sum_{\lambda_{\uparrow}^{\prime}}\Sigma_{% \lambda_{\uparrow},\lambda_{\uparrow}^{\prime}}(E;\omega_{L})\psi^{(0)}_{% \lambda_{\uparrow}^{\prime}},italic_E italic_ψ start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT bold_0 , italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ψ start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT roman_Σ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT , italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_E ; italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ) italic_ψ start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT , (35)

which clearly demonstrates the equivalence of the two methods.

Since we only aim for qualitative accuracy, we approximate the shift in the exciton resonance by the self-energy,

Δλ(ωL)Σλλ(E𝟎,λ;ωL).similar-to-or-equalssubscriptΔ𝜆subscript𝜔𝐿subscriptΣ𝜆𝜆subscript𝐸𝟎𝜆subscript𝜔𝐿\displaystyle\Delta_{\lambda}(\omega_{L})\simeq\Sigma_{\lambda\lambda}(E_{% \mathbf{0},\lambda};\omega_{L}).roman_Δ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ) ≃ roman_Σ start_POSTSUBSCRIPT italic_λ italic_λ end_POSTSUBSCRIPT ( italic_E start_POSTSUBSCRIPT bold_0 , italic_λ end_POSTSUBSCRIPT ; italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ) . (36)

It is important to note that Eq. (36) captures the ac Stark effect, whereby the shift can change sign when the laser frequency ωLsubscript𝜔𝐿\omega_{L}italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT is tuned into resonance with the biexciton bound state, i.e., E𝟎,λ=ωLEBX;αsubscript𝐸𝟎𝜆subscript𝜔𝐿subscript𝐸BX𝛼E_{\mathbf{0},\lambda}=\omega_{L}-E_{\rm{BX};\alpha}italic_E start_POSTSUBSCRIPT bold_0 , italic_λ end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT roman_BX ; italic_α end_POSTSUBSCRIPT, where EBX;αsubscript𝐸BX𝛼E_{\rm{BX};\alpha}italic_E start_POSTSUBSCRIPT roman_BX ; italic_α end_POSTSUBSCRIPT denotes the energy of a zero quasi-momentum biexciton state. However, unlike the simpler case of a monolayer TMD, this sign change does not occur at all biexciton energies, as not all biexciton states will couple to the λ𝜆\lambdaitalic_λ-th exciton. We can therefore expect in the experiment, which only observes shifts in the optically active MX1 and MX2, to not detect all biexciton states. This is discussed further in the main text.

H.6 Same valley interactions

To conclude our work we briefly consider interactions from excitons in the same valley, which do not support a bound state. Owing to this, we simply take the Born approximation of the T𝑇Titalic_T matrix assuming repulsive contact interactions. Focusing on zero quasi-momentum, the interactions between the excitons are given by

𝒱^=g2Uλ3λ4λ1λ2X^𝟎,λ1,X^𝟎,λ2,X^𝟎,λ3,X^𝟎,λ4,.^𝒱𝑔2subscriptsuperscript𝑈subscript𝜆1subscript𝜆2subscript𝜆3subscript𝜆4subscriptsuperscript^𝑋𝟎subscript𝜆1subscriptsuperscript^𝑋𝟎subscript𝜆2subscript^𝑋𝟎subscript𝜆3subscript^𝑋𝟎subscript𝜆4\displaystyle\hat{\mathscr{V}}=\frac{g}{2}\sum U^{\lambda_{1}\lambda_{2}}_{% \lambda_{3}\lambda_{4}}\hat{X}^{\dagger}_{\mathbf{0},\lambda_{1},\downarrow}% \hat{X}^{\dagger}_{\mathbf{0},\lambda_{2},\downarrow}\hat{X}_{\mathbf{0},% \lambda_{3},\downarrow}\hat{X}_{\mathbf{0},\lambda_{4},\downarrow}.over^ start_ARG script_V end_ARG = divide start_ARG italic_g end_ARG start_ARG 2 end_ARG ∑ italic_U start_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 , italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , ↓ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 , italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , ↓ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_0 , italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , ↓ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_0 , italic_λ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , ↓ end_POSTSUBSCRIPT . (37)

Here g𝑔gitalic_g is the strength of the repulsion, and we have introduce the overlap integral

Uλ3λ4λ1λ2=d2rφλ1,(𝐫)φλ2,(𝐫)φλ3,(𝐫)φλ4,(𝐫)subscriptsuperscript𝑈subscript𝜆1subscript𝜆2subscript𝜆3subscript𝜆4superscript𝑑2𝑟superscriptsubscript𝜑subscript𝜆1𝐫superscriptsubscript𝜑subscript𝜆2𝐫subscript𝜑subscript𝜆3𝐫subscript𝜑subscript𝜆4𝐫\displaystyle U^{\lambda_{1}\lambda_{2}}_{\lambda_{3}\lambda_{4}}=\int d^{2}r% \,\varphi_{\lambda_{1},\downarrow}^{*}(\mathbf{r})\varphi_{\lambda_{2},% \downarrow}^{*}(\mathbf{r})\varphi_{\lambda_{3},\downarrow}(\mathbf{r})\varphi% _{\lambda_{4},\downarrow}(\mathbf{r})italic_U start_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = ∫ italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_r italic_φ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_r ) italic_φ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_r ) italic_φ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , ↓ end_POSTSUBSCRIPT ( bold_r ) italic_φ start_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT , ↓ end_POSTSUBSCRIPT ( bold_r ) (38)

where φλ,subscript𝜑𝜆\varphi_{\lambda,\downarrow}italic_φ start_POSTSUBSCRIPT italic_λ , ↓ end_POSTSUBSCRIPT is the Bloch wavefunction in real space at zero quasi-momentum with band index λ𝜆\lambdaitalic_λ and valley index \downarrow. In order to derive the mean-field Hamiltonian given in the main text, we restrict ourselves to the optically bright excitonic states MX1 (λ=0𝜆0\lambda=0italic_λ = 0) and MX2 (λ=3𝜆3\lambda=3italic_λ = 3); for further details on these states see the discussion below Eq. (II.3). We furthermore use the fact that βλ=κλnλsubscript𝛽𝜆subscript𝜅𝜆subscript𝑛𝜆\beta_{\lambda}=\kappa_{\lambda}\sqrt{n_{\lambda}}italic_β start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = italic_κ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT square-root start_ARG italic_n start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT end_ARG (with κλ=eiarg(ϕ𝟎(𝟎,λ,)Ω)subscript𝜅𝜆superscript𝑒𝑖argsuperscriptsubscriptitalic-ϕ𝟎𝟎𝜆superscriptsuperscriptΩ\kappa_{\lambda}=e^{i\rm{arg}(-\phi_{\mathbf{0}}^{(\mathbf{0},\lambda,% \downarrow)}{}^{*}\Omega^{*})}italic_κ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT italic_i roman_arg ( - italic_ϕ start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( bold_0 , italic_λ , ↓ ) end_POSTSUPERSCRIPT start_FLOATSUPERSCRIPT ∗ end_FLOATSUPERSCRIPT roman_Ω start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) end_POSTSUPERSCRIPT) and Φ|X^𝟎,λ,X^𝟎,λ,|Φ=βλβλbraΦsubscriptsuperscript^𝑋𝟎𝜆subscript^𝑋𝟎superscript𝜆ketΦsuperscriptsubscript𝛽𝜆subscript𝛽superscript𝜆\bra{\Phi}\hat{X}^{\dagger}_{\mathbf{0},\lambda,\downarrow}\hat{X}_{\mathbf{0}% ,\lambda^{\prime},\downarrow}\ket{\Phi}=\beta_{\lambda}^{*}\beta_{\lambda^{% \prime}}⟨ start_ARG roman_Φ end_ARG | over^ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_0 , italic_λ , ↓ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT bold_0 , italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , ↓ end_POSTSUBSCRIPT | start_ARG roman_Φ end_ARG ⟩ = italic_β start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT italic_λ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT. Decomposing the interaction in Eq. (37) using mean-field Hartree-Fock theory then yields Eq. (1). We point out that for contact interactions with bosons, the Hartree and Fock terms are identical in magnitude and sign. The relevant overlap integrals are

u11/gsubscript𝑢11𝑔\displaystyle u_{11}/gitalic_u start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT / italic_g =U0,00,0=2.2absentsubscriptsuperscript𝑈00002.2\displaystyle=U^{0,0}_{0,0}=2.2= italic_U start_POSTSUPERSCRIPT 0 , 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , 0 end_POSTSUBSCRIPT = 2.2 (39a)
u22/gsubscript𝑢22𝑔\displaystyle u_{22}/gitalic_u start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT / italic_g =U3,33,3=1.9absentsubscriptsuperscript𝑈33331.9\displaystyle=U^{3,3}_{3,3}=1.9= italic_U start_POSTSUPERSCRIPT 3 , 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 3 , 3 end_POSTSUBSCRIPT = 1.9 (39b)
u12/gsubscript𝑢12𝑔\displaystyle u_{12}/gitalic_u start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT / italic_g =U0,30,3=1.3absentsubscriptsuperscript𝑈03031.3\displaystyle=U^{0,3}_{0,3}=1.3= italic_U start_POSTSUPERSCRIPT 0 , 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , 3 end_POSTSUBSCRIPT = 1.3 (39c)
k1/gsubscript𝑘1𝑔\displaystyle k_{1}/gitalic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_g =κ03κ3U0,30,0=1.2absentsuperscriptsubscript𝜅03subscript𝜅3subscriptsuperscript𝑈00031.2\displaystyle=\kappa_{0}^{3}\kappa_{3}U^{0,0}_{0,3}=-1.2= italic_κ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_κ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT italic_U start_POSTSUPERSCRIPT 0 , 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , 3 end_POSTSUBSCRIPT = - 1.2 (39d)
k2/gsubscript𝑘2𝑔\displaystyle k_{2}/gitalic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_g =κ0κ33U3,03,3=0.4,absentsubscript𝜅0superscriptsubscript𝜅33subscriptsuperscript𝑈33300.4\displaystyle=\kappa_{0}\kappa_{3}^{3}U^{3,3}_{3,0}=-0.4,= italic_κ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_U start_POSTSUPERSCRIPT 3 , 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 3 , 0 end_POSTSUBSCRIPT = - 0.4 , (39e)

where we have assumed that κ𝜅\kappaitalic_κ are real. All other interaction terms can be derived from these five, since the overlap integral only depends only on the number of each index (e.g., U1,01,1=U0,11,1=U1,11,0=U1,10,1subscriptsuperscript𝑈1110subscriptsuperscript𝑈1101subscriptsuperscript𝑈1011subscriptsuperscript𝑈0111U^{1,1}_{1,0}=U^{1,1}_{0,1}=U^{1,0}_{1,1}=U^{0,1}_{1,1}italic_U start_POSTSUPERSCRIPT 1 , 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 , 0 end_POSTSUBSCRIPT = italic_U start_POSTSUPERSCRIPT 1 , 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , 1 end_POSTSUBSCRIPT = italic_U start_POSTSUPERSCRIPT 1 , 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT = italic_U start_POSTSUPERSCRIPT 0 , 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT).

References

  • Mak and Shan [2022] K. F. Mak and J. Shan, Nature Nanotechnology 17, 686 (2022).
  • Andrei et al. [2021] E. Y. Andrei, D. K. Efetov, P. Jarillo-Herrero, A. H. MacDonald, K. F. Mak, T. Senthil, E. Tutuc, A. Yazdani, and A. F. Young, Nature Reviews Materials 6, 201 (2021).
  • Kennes et al. [2021] D. M. Kennes, M. Claassen, L. Xian, A. Georges, A. J. Millis, J. Hone, C. R. Dean, D. N. Basov, A. N. Pasupathy, and A. Rubio, Nat. Phys. 17, 155 (2021).
  • Shimazaki et al. [2020] Y. Shimazaki, I. Schwartz, K. Watanabe, T. Taniguchi, M. Kroner, and A. Imamoğlu, Nature 580, 472 (2020).
  • Tang et al. [2020] Y. Tang, L. Li, T. Li, Y. Xu, S. Liu, K. Barmak, K. Watanabe, T. Taniguchi, A. H. MacDonald, J. Shan, and K. F. Mak, Nature 579, 353 (2020).
  • Regan et al. [2020] E. C. Regan, D. Wang, C. Jin, M. I. Bakti Utama, B. Gao, X. Wei, S. Zhao, W. Zhao, Z. Zhang, K. Yumigeta, et al., Nature 579, 359 (2020).
  • Ciorciaro et al. [2023] L. Ciorciaro, T. Smoleński, I. Morera, N. Kiper, S. Hiestand, M. Kroner, Y. Zhang, K. Watanabe, T. Taniguchi, E. Demler, and A. İmamoğlu, Nature 623, 509–513 (2023).
  • Park et al. [2023] H. Park, J. Cai, E. Anderson, Y. Zhang, J. Zhu, X. Liu, C. Wang, W. Holtzmann, C. Hu, Z. Liu, T. Taniguchi, K. Watanabe, J.-H. Chu, T. Cao, L. Fu, W. Yao, C.-Z. Chang, D. Cobden, D. Xiao, and X. Xu, Nature 622, 74 (2023).
  • Zeng et al. [2023] Y. Zeng, Z. Xia, K. Kang, J. Zhu, P. Knüppel, C. Vaswani, K. Watanabe, T. Taniguchi, K. F. Mak, and J. Shan, Nature 622, 69 (2023).
  • Shree et al. [2021] S. Shree, I. Paradisanos, X. Marie, C. Robert, and B. Urbaszek, Nature Reviews Physics 3, 39 (2021).
  • Rapaport et al. [2000] R. Rapaport, R. Harel, E. Cohen, A. Ron, E. Linder, and L. N. Pfeiffer, Physical Review Letters 84, 1607 (2000).
  • Suris [2003] R. A. Suris, in Optical Properties of 2D Systems with Interacting Electrons (Springer, 2003) pp. 111–124.
  • Sidler et al. [2017] M. Sidler, P. Back, O. Cotlet, A. Srivastava, T. Fink, M. Kroner, E. Demler, and A. Imamoglu, Nature Physics 13, 255 (2017).
  • Efimkin and MacDonald [2017] D. K. Efimkin and A. H. MacDonald, Physical Review B 95, 035417 (2017).
  • Glazov [2020] M. M. Glazov, The Journal of Chemical Physics 153 (2020).
  • Wang et al. [2018] G. Wang, A. Chernikov, M. M. Glazov, T. F. Heinz, X. Marie, T. Amand, and B. Urbaszek, Reviews of Modern Physics 90, 021001 (2018).
  • Xu et al. [2020] Y. Xu, S. Liu, D. A. Rhodes, K. Watanabe, T. Taniguchi, J. Hone, V. Elser, K. F. Mak, and J. Shan, Nature 587, 214–218 (2020).
  • Smoleński et al. [2021] T. Smoleński, P. E. Dolgirev, C. Kuhlenkamp, A. Popert, Y. Shimazaki, P. Back, X. Lu, M. Kroner, K. Watanabe, T. Taniguchi, et al., Nature 595, 53 (2021).
  • Cunningham et al. [2019a] P. D. Cunningham, A. T. Hanbicki, T. L. Reinecke, K. M. McCreary, and B. T. Jonker, Nature communications 10, 5539 (2019a).
  • Slobodeniuk et al. [2023] A. O. Slobodeniuk, P. Koutenský, M. Bartoš, F. Trojánek, P. Malý, T. Novotný, and M. Kozák, npj 2D Materials and Applications 710.1038/s41699-023-00385-1 (2023).
  • Combescot [1992] M. Combescot, Physics reports 221, 167 (1992).
  • Schmitt-Rink and Chemla [1986] S. Schmitt-Rink and D. Chemla, Physical Review Letters 57, 2752 (1986).
  • Zimmermann [1990] R. Zimmermann, Festkörperprobleme 30: Plenary Lectures of the Divisions Semiconductor Physics Thin Films Dynamics and Statistical Physics Magnetism Metal Physics Surface Physics Low Temperature Physics Molecular Physics of the German Physical Society (DPG), Regensburg, March 26 to 30, 1990 , 295 (1990).
  • Haug and Koch [2009] H. Haug and S. W. Koch, Quantum theory of the optical and electronic properties of semiconductors (World Scientific Publishing Company, 2009).
  • Uto et al. [2024] T. Uto, B. Evrard, K. Watanabe, T. Taniguchi, M. Kroner, and A. İmamoğlu, Phys. Rev. Lett. 132, 056901 (2024).
  • Sarkar et al. [2024] S. Sarkar, M. J. Mehrabad, D. G. Suárez-Forero, L. Gu, C. J. Flower, L. Xu, K. Watanabe, T. Taniguchi, S. Park, H. Jang, Y. Zhou, and M. Hafezi, Sub-wavelength optical lattice in 2d materials (2024), arXiv:2406.00464 [cond-mat.mes-hall] .
  • Yong et al. [2018] C.-K. Yong, J. Horng, Y. Shen, H. Cai, A. Wang, C.-S. Yang, C.-K. Lin, S. Zhao, K. Watanabe, T. Taniguchi, et al., Nature Physics 14, 1092 (2018).
  • Sie et al. [2016] E. J. Sie, C. H. Lui, Y.-H. Lee, J. Kong, and N. Gedik, Nano letters 16, 7421 (2016).
  • Hao et al. [2017] K. Hao, J. F. Specht, P. Nagler, L. Xu, K. Tran, A. Singh, C. K. Dass, C. Schüller, T. Korn, M. Richter, et al., Nature communications 8, 15552 (2017).
  • Cunningham et al. [2019b] P. D. Cunningham, A. T. Hanbicki, T. L. Reinecke, K. M. McCreary, and B. T. Jonker, Nature communications 10, 5539 (2019b).
  • Cam et al. [2022] H. N. Cam, N. T. Phuc, and V. A. Osipov, npj 2D Materials and Applications 6, 22 (2022).
  • Brem and Malic [2024] S. Brem and E. Malic, 2D Materials 11, 025030 (2024).
  • Kim et al. [2014] J. Kim, X. Hong, C. Jin, S.-F. Shi, C.-Y. S. Chang, M.-H. Chiu, L.-J. Li, and F. Wang, Science 346, 1205 (2014).
  • Sie et al. [2014] E. J. Sie, J. W. McIver, Y.-H. Lee, L. Fu, J. Kong, and N. Gedik, Nature Materials 14, 290 (2014).
  • Sie et al. [2017] E. J. Sie, C. H. Lui, Y.-H. Lee, L. Fu, J. Kong, and N. Gedik, Science 355, 1066 (2017).
  • Ciuti et al. [1998] C. Ciuti, V. Savona, C. Piermarocchi, A. Quattropani, and P. Schwendimann, Physical Review B 58, 7926 (1998).
  • Shahnazaryan et al. [2017] V. Shahnazaryan, I. Iorsh, I. A. Shelykh, and O. Kyriienko, Physical Review B 96, 115409 (2017).
  • Polovnikov et al. [2022] B. Polovnikov, J. Scherzer, S. Misra, X. Huang, C. Mohl, Z. Li, J. Göser, J. Förste, I. Bilgin, K. Watanabe, et al., arXiv preprint arXiv:2208.04056  (2022).
  • Adlong et al. [2024] H. S. Adlong, J. Levinsen, and M. M. Parish (2024), unpublished manuscript.
  • Tan et al. [2020] L. B. Tan, O. Cotlet, A. Bergschneider, R. Schmidt, P. Back, Y. Shimazaki, M. Kroner, and A. İmamoğlu, Physical Review X 10, 021011 (2020).
  • Han et al. [2018] B. Han, C. Robert, E. Courtade, M. Manca, S. Shree, T. Amand, P. Renucci, T. Taniguchi, K. Watanabe, X. Marie, et al., Physical Review X 8, 031073 (2018).
  • Robert et al. [2016] C. Robert, D. Lagarde, F. Cadiz, G. Wang, B. Lassagne, T. Amand, A. Balocchi, P. Renucci, S. Tongay, B. Urbaszek, et al., Physical review B 93, 205423 (2016).
  • Yang et al. [2022] M. Yang, L. Ren, C. Robert, D. Van Tuan, L. Lombez, B. Urbaszek, X. Marie, and H. Dery, Physical Review B 105, 085302 (2022).
  • Xiong et al. [2023] R. Xiong, J. H. Nie, S. L. Brantly, P. Hays, R. Sailus, K. Watanabe, T. Taniguchi, S. Tongay, and C. Jin, Science 380, 860 (2023)https://www.science.org/doi/pdf/10.1126/science.add5574 .
  • Gao et al. [2023] B. Gao, D. G. Suárez-Forero, S. Sarkar, T.-S. Huang, D. Session, M. J. Mehrabad, R. Ni, M. Xie, J. Vannucci, S. Mittal, K. Watanabe, T. Taniguchi, A. Imamoglu, Y. Zhou, and M. Hafezi, Excitonic mott insulator in a bose-fermi-hubbard system of moiré WS2subscriptWS2\rm{WS}_{2}roman_WS start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT/WSe2subscriptWSe2\rm{WSe}_{2}roman_WSe start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT heterobilayer (2023), arXiv:2304.09731 [cond-mat.mes-hall] .
  • Huang et al. [2022] D. Huang, J. Choi, C.-K. Shih, and X. Li, Nature Nanotechnology 17, 227 (2022).
  • Du et al. [2023] L. Du, M. R. Molas, Z. Huang, G. Zhang, F. Wang, and Z. Sun, Science 379, eadg0014 (2023).
  • Zhang et al. [2018] N. Zhang, A. Surrente, M. Baranowski, D. K. Maude, P. Gant, A. Castellanos-Gomez, and P. Plochocka, Nano letters 18, 7651 (2018).
  • Jin et al. [2019] C. Jin, E. C. Regan, A. Yan, M. Iqbal Bakti Utama, D. Wang, S. Zhao, Y. Qin, S. Yang, Z. Zheng, S. Shi, et al., Nature 567, 76 (2019).
  • Liu et al. [2021] E. Liu, E. Barré, J. van Baren, M. Wilson, T. Taniguchi, K. Watanabe, Y.-T. Cui, N. M. Gabor, T. F. Heinz, Y.-C. Chang, et al., Nature 594, 46 (2021).
  • Wu et al. [2022] B. Wu, H. Zheng, S. Li, J. Ding, J. He, Y. Zeng, K. Chen, Z. Liu, S. Chen, A. Pan, et al., Light: Science & Applications 11, 166 (2022).
  • Tran et al. [2019] K. Tran, G. Moody, F. Wu, X. Lu, J. Choi, K. Kim, A. Rai, D. A. Sanchez, J. Quan, A. Singh, J. Embley, A. Zepeda, M. Campbell, T. Autry, T. Taniguchi, K. Watanabe, N. Lu, S. K. Banerjee, K. L. Silverman, S. Kim, E. Tutuc, L. Yang, A. H. MacDonald, and X. Li, Nature 567, 71–75 (2019).
  • Alexeev et al. [2019] E. M. Alexeev, D. A. Ruiz-Tijerina, M. Danovich, M. J. Hamer, D. J. Terry, P. K. Nayak, S. Ahn, S. Pak, J. Lee, J. I. Sohn, et al., Nature 567, 81 (2019).
  • Seyler et al. [2019] K. L. Seyler, P. Rivera, H. Yu, N. P. Wilson, E. L. Ray, D. G. Mandrus, J. Yan, W. Yao, and X. Xu, Nature 567, 66–70 (2019).
  • Lin et al. [2023] B.-H. Lin, Y.-C. Chao, I.-T. Hsieh, C.-P. Chuu, C.-J. Lee, F.-H. Chu, L.-S. Lu, W.-T. Hsu, C.-W. Pao, C.-K. Shih, et al., Nano letters 23, 1306 (2023).
  • Wu et al. [2017] F. Wu, T. Lovorn, and A. H. MacDonald, Physical review letters 118, 147401 (2017).
  • Brem et al. [2020] S. Brem, C. Linderälv, P. Erhart, and E. Malic, Nano letters 20, 8534 (2020).
  • Naik et al. [2022] M. H. Naik, E. C. Regan, Z. Zhang, Y.-H. Chan, Z. Li, D. Wang, Y. Yoon, C. S. Ong, W. Zhao, S. Zhao, et al., Nature 609, 52 (2022).
  • Huang et al. [2023] T.-S. Huang, P. Lunts, and M. Hafezi, Non-bosonic moiré excitons (2023).
  • Guo et al. [2020] H. Guo, X. Zhang, and G. Lu, Science Advances 610.1126/sciadv.abc5638 (2020).
  • [61] See https://www.research-collection.ethz.ch/handle/20.500.11850/661389.
  • Zomer et al. [2014] P. Zomer, M. Guimarães, J. Brant, N. Tombros, and B. Van Wees, Applied Physics Letters 105, 013101 (2014).
  • Scuri et al. [2018] G. Scuri, Y. Zhou, A. A. High, D. S. Wild, C. Shu, K. De Greve, L. A. Jauregui, T. Taniguchi, K. Watanabe, P. Kim, et al., Physical review letters 120, 037402 (2018).
  • Laturia et al. [2018] A. Laturia, M. L. Van de Put, and W. G. Vandenberghe, npj 2D Materials and Applications 2, 6 (2018).
  • Kim et al. [2012] K. K. Kim, A. Hsu, X. Jia, S. M. Kim, Y. Shi, M. Dresselhaus, T. Palacios, and J. Kong, ACS nano 6, 8583 (2012).
  • Barachati et al. [2018] F. Barachati, A. Fieramosca, S. Hafezian, J. Gu, B. Chakraborty, D. Ballarini, L. Martinu, V. Menon, D. Sanvitto, and S. Kéna-Cohen, Nature nanotechnology 13, 906 (2018).
  • Smoleński et al. [2019] T. Smoleński, O. Cotlet, A. Popert, P. Back, Y. Shimazaki, P. Knüppel, N. Dietler, T. Taniguchi, K. Watanabe, M. Kroner, and A. Imamoglu, Physical Review Letters 123, 097403 (2019).
  • Muir et al. [2022] J. B. Muir, J. Levinsen, S. K. Earl, M. A. Conway, J. H. Cole, M. Wurdack, R. Mishra, D. J. Ing, E. Estrecho, Y. Lu, et al., Nature Communications 13, 6164 (2022).
  • Levinsen and Parish [2015] J. Levinsen and M. M. Parish, Annu. Rev. Cold Atoms Mol. 3, 1 (2015).
  • Kumar et al. [2023] S. S. Kumar, B. C. Mulkerin, M. M. Parish, and J. Levinsen, Phys. Rev. B 108, 125416 (2023).
  • Chevy [2006] F. Chevy, Physical Review A 74, 063628 (2006).