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

\pdfximage

supplement.pdf

Origin of interlayer exciton-phonon coupling in 2D heterostructures

Muralidhar Nalabothula [email protected]    Ludger Wirtz    Sven Reichardt Department of Physics and Materials Science, University of Luxembourg, 162a avenue de la Faïencerie, L-1511 Luxembourg, Luxembourg
Abstract

The coupling between excitons and phonons across adjacent layers has been experimentally observed in various heterostructures of layered materials. Yet the precise mechanism underlying this phenomenon remains elusive. Using the WSe2@hhBN heterostructure as an example, we study the origin of the interlayer exciton-phonon coupling and its signature in resonant Raman scattering through first-principles calculations. Our study emphasizes the central role of crystal symmetries in the interlayer exciton-phonon scattering processes, which are responsible for the anomalous resonant Raman intensities of the in-plane and the out-of-plane hhBN phonon modes. We find that the deformation potential induced by the hhBN phonon interacts with the hybridized hole density of WSe2 excitons near the hhBN interface, leading to interlayer exciton-phonon coupling.

preprint: APS/123-QED

In recent years, interfacing two-dimensional (2D) materials with different layers or substrates [1] has revealed fascinating properties that are difficult to achieve with individual layers. For example, interlayer electron-electron interactions, can give rise to diverse phenomena such as moiré excitons [2], superconducting phases [3], and Mott insulating states [4]. Likewise, interlayer electron-phonon interactions can affect the mobilities [5] and superconducting temperatures [6] of 2D materials.

Similarly, recent optical scattering measurements on 2D heterostructures have shed light on the coupling of excitons in one layer with phonons of the adjacent layer [7, 8, 9, 10, 11, 12]. This remarkable interlayer exciton-phonon coupling was first demonstrated in monolayer WSe2 encapsulated in hexagonal boron nitride (hhBN) using Raman spectroscopy [7]. Subsequent Raman and photoluminescence measurements have confirmed its existence in various other heterostructures [12, 9, 13, 14]. This interlayer exciton-phonon coupling can play a key role in exciton lifetimes and decoherence times [15], and the study of phonon polaritons in neighbouring layers [16].

Albeit the experimental evidence for interlayer exciton-phonon coupling, the microscopic mechanism responsible for this phenomenon still remains a mystery. In this Letter, we use WSe2@hhBN as an example to unveil the microscopic mechanism of exciton-phonon coupling across layers. Using ab initio methods, we compute resonant Raman intensities, which provide a detailed atomistic view of the interlayer exciton-phonon scattering process. We demonstrate the selection rules in interlayer exciton-phonon scattering, which are responsible for the anomalous resonant Raman intensities: when in resonance with the A exciton of WSe2, the out-of-plane hhBN phonon mode (which is Raman forbidden in pure hhBN) exhibits a much higher intensity in the heterostructure than the Raman-allowed in-plane mode [7]. Our main findings reveal that the deformation potential of the hhBN phonon scatters the hybridized part of the WSe2 exciton-hole in the vicinity of the hhBN layer, giving rise to interlayer exciton-phonon coupling. We show that this coupling is extremely sensitive to the interlayer distance and bond polarity of the phonon layer and is not required at all to observe this effect [9, 10, 11, 12].

We start our discussion by looking at the resonant Raman scattering in a monolayer WSe2@hhBN heterostructure. We follow Refs. [17, 18] and calculate the differential cross section for Stokes Raman scattering mediated by one phonon:

dσdΩωLωλωL|μνλ(ωL,ωλ)|2.\frac{\mathrm{d}\sigma}{\mathrm{d}\Omega}\propto\frac{\omega_{\mathrm{L}}-\omega_{\lambda}}{\omega_{\mathrm{L}}}|\mathcal{M}_{\mu\nu}^{\lambda}(\omega_{L},\omega_{\lambda})|^{2}. (1)

Here, μ\mu and ν\nu denote the polarization of the incoming and outgoing light, respectively, while ωL\omega_{\mathrm{L}} and ωλ\omega_{\lambda} denote the frequencies of the incoming light and the created phonon of branch λ\lambda, respectively. Within the Tamm-Dancoff approximation [19], the Raman scattering matrix element μνλ\mathcal{M}^{\lambda}_{\mu\nu} takes on the simple form

μνλ(ωL,ωλ)=S,S(dSν)(gSSλ)dSμ(ωLES+iγ)(ωLωλES+iγ)+S,S(dSμ)gSSλdSν(ωL+ESiγ)(ωLωλ+ESiγ).\begin{split}\mathcal{M}_{\mu\nu}^{\lambda}\left(\omega_{L},\omega_{\lambda}\right)=\sum_{S,S^{\prime}}\frac{\left(d^{\nu}_{S^{\prime}}\right)^{*}(g_{SS^{\prime}}^{\lambda})^{*}d_{S}^{\mu}}{\left(\hbar\omega_{\mathrm{L}}-E_{S}+i\gamma\right)\left(\hbar\omega_{\mathrm{L}}-\hbar\omega_{\lambda}-E_{S^{\prime}}+i\gamma\right)}\\ +\sum_{S,S^{\prime}}\frac{\left(d_{S}^{\mu}\right)^{*}g_{SS^{\prime}}^{\lambda}d^{\nu}_{S^{\prime}}}{\left(\hbar\omega_{\mathrm{L}}+E_{S}-i\gamma\right)\left(\hbar\omega_{\mathrm{L}}-\hbar\omega_{\lambda}+E_{S^{\prime}}-i\gamma\right)}.\end{split} (2)

The sums run over all excitonic states SS and SS^{\prime} with energies ES()E_{S{(\prime)}} and decay constant γ\gamma. The quantity dSμd^{\mu}_{S} is the coupling matrix element between an exciton SS and a photon of polarization μ\mu, while (gSSλ)(g^{\lambda}_{SS^{\prime}})^{*} represents the exciton-phonon coupling matrix element related to the scattering of an exciton SS to an exciton SS^{\prime} via emission of one phonon of branch λ\lambda [18]. We obtain all needed quantities for an evaluation of Eq. 2 for a hetereostructure of one layer hBN and one layer of WSe2 from first principles using the GW-BSE [20, 21] formalism on top of density functional theory (see supplementary information (SM) [22] for details of the ab initio methods [23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37]).

Refer to caption
Figure 1: Raman intensities of the ZO and LO/TO modes of hhBN in a heterostructure of single layers of hhBN and WSe2 as a function of laser excitation energy. a) Computed resonant Raman intensities of the ZO mode (blue line) and LO/TO mode (orange line). The gray shaded area denotes the imaginary part of the in-plane polarizability of the heterostructure. Blue triangles represent experimental data of the ZO mode Raman intensities at 4 K, taken from Ref. [38]. (b) and (c) Computed Raman intensities for (b) the ZO mode and (c) the LO/TO mode, considering all scattering channels (red lines) or retaining only intra-exciton scattering channels (blue line).

In Fig. 1a, we show the calculated, phonon branch-dependent Raman intensities, as a function of the energy of the incoming photon. The blue line denotes the Raman intensity for the out-of-plane optical (ZO) phonon of the hBN layer. We compare the Raman intensity profiles to the imaginary part of the in-plane polarizability (absorption coefficient) of the heterostructure (shaded area) and to experimental Raman intensities from McDonnell et al. [38](blue triangles, see SM [22] for further details). The most striking features of the ZO-mode intensity profile are the two strong resonances at ωL1.72\omega_{\mathrm{L}}\approx 1.72 eV and ωL1.82\omega_{\mathrm{L}}\approx 1.82 eV, labeled “I” and “II” in Fig. 1a. The first of these two peaks coincides with a resonance in the absorption coefficient which corresponds to the well-known 1s1s state of the “A”-exciton [39] of WSe2 and can thus be interpreted as the resonant excitation of the 1s1s A-exciton in WSe2, which then couples to the hhBN ZO mode. The second peak, in contrast, has no such counterpart in the absorption spectrum. It is the quantum of one ZO mode away from the first resonance peak and corresponds to the resonant recombination of the 1s1s exciton under phonon emission (when ωL=E1s+ωZO\omega_{\mathrm{L}}=E_{1s}+\omega_{\mathrm{ZO}}), compare Eq. (2). Finally, Peak III coincides with the 2s2s-A-exciton of the absorption coefficient and is due to the resonant excitation of the 2s2s state. The results of our calculation match the the experimental data of McDonnell et al. [38] well in terms of the local peak intensities. We note that the 2s2s exciton is blue-shifted in our calculation with respect to experiment, as we consider a heterostructure of one layer WSe2 on top of one layer of hhBN vs. a system of monolayer WSe2 sandwiched in bulk hhBN in experiment (where screening is stronger and, thus, the energy difference between 1s1s and 2s2s excitons reduced) [39].

Compared to the ZO mode of hhBN, its in-plane LO/TO mode displays a Raman intensity three orders of magnitude lower, yet with a qualitatively similar resonance structure as a function of ωL\omega_{\mathrm{L}}. This remarkable, anomalous behavior of two different allowed phonon modes coupling so differently when the incident light is in resonance with a neighboring layer has been reported in Ref. [7], which provided a speculative interpretation of the underlying Raman scattering process. Using our atomistic first-principles approach, we can now scrutinize the Raman scattering pathways and understand them in terms of symmetry and involved scattering events.

From Eq. (2), it is clear that any contributing scattering pathway must feature non-zero optical matrix elements dSμd^{\mu}_{S} and dSνd^{\nu}_{S^{\prime}} as well as a non-vanishing exciton-phonon coupling matrix element gS,Sλg^{\lambda}_{S,S^{\prime}}. In order to infer the corresponding selection rules, we note that the underlying C3C_{3} point group symmetry of the hhBN@WSe2 heterostructure allows the classification of zero-momentum phonons and excitons by their total angular momentum component in the out-of-plane direction, with corresponding quantum number m=+1m=+1, 0, or -1. A finite optical strength dSμd^{\mu}_{S} for light polarized parallel to the heterostructure is only possible for excitons SS with mS=±1m_{S}=\pm 1. Meanwhile, the exciton-phonon matrix elements gS,Sλg^{\lambda}_{S,S^{\prime}} are non-zero only if total angular momentum (modulo 3) is conserved (see SM [22] for derivation):

mSmSmλ0 (mod 3).m_{S}-m_{S^{\prime}}-m_{\lambda}\equiv 0\text{ (mod 3)}. (3)

For the ZO-phonon (mZO=0m_{\mathrm{ZO}}=0), this implies that active scattering pathways necessarily have mS=mSm_{S}=m_{S^{\prime}}. By contrast, for the LO/TO-phonon (mLO/TO=±1m_{\mathrm{LO/TO}}=\pm 1), we need to have mS=mS±1m_{S}=m_{S^{\prime}}\pm 1 (mod 3).

The combination of the optical and phonon-specific selection rules allows us then to understand the nature of the resonance features I, II, and III in Fig. 1a. As the 1s1s exciton of WSe2 is the only optically active low-energy exciton available, resonance I and II are associated with the scattering process 1s1s1s\to 1s, which gives rise to resonances for both the incoming (I) and the outgoing (II) light. However, while this resonance structure is seen in both the ZO and the LO/TO mode, the latter is three orders of magnitude less intense, although formerly allowed by symmetry. To understand this enormous difference between the phonon modes, we first note that the 1s1s exciton of WSe2 is in fact doubly degenerate due to time reversal symmetry [40]. The two members of the doublet are localized at different, inequivalent corners K and K’ of the first Brillouin zone and carry opposite total angular momentum m1s,K()=+()1m_{1s,K^{(\prime)}}=+(-)1. For the ZO mode, this implies that 1s1s1s\to 1s scattering is allowed within a valley (intra-valley scattering), while for the LO/TO mode with finite angular momentum, 1s1s1s\to 1s scattering is only allowed between 1s1s states in opposite valleys (inter-valley scattering). However, due to the strong localization of the 1s1s-exciton wave function in momentum space (see Fig. S4 of SM [22]), there is only a miniscule overlap between wave functions centered in different valleys and, as a result, the inter-valley scattering required for the LO/TO mode is strongly suppressed and so is consequently the Raman intensity.

In order to substantiate this, we re-calculate the Raman intensities by considering only intra-exciton scattering channels, i.e., by restricting the double sum in Eq. (2) to S=SS=S^{\prime}. As shown in Fig. 1b and c, we can indeed numerically confirm that intra-exciton scattering is absent for the LO/TO mode while being the dominant scattering process for the ZO mode. In terms of the three most prominent resonances in the ZO mode Raman intensities, we can thus conclude that resonances I and II stem from 1s1s-to-1s1s intra-valley exciton scattering and III from 2s2s-to-2s2s intra-valley scattering. We note that this finding is in contrast to previous assumptions [7] that identified peaks II and III to the two resonances associated with the inter-exciton scattering process 1s2s1s\to 2s. For the LO/TO mode, however, we can confirm that intra-valley scattering plays no role and that the suppressed finite Raman intensity around the 1s1s exciton can only stem from weak inter-exciton scattering.

While these symmetry considerations explain the difference between Raman intensities of the LO/TO- and the ZO-mode, a complimentary analysis is required to understand the precise mechanism for the interlayer exciton-phonon coupling. In order to elucidate the microscopic mechanism, we consider the 1s1s1s\to 1s scattering pathway for the ZO mode, which is the dominant Raman scattering channel in the region ωL\omega_{\mathrm{L}}\lesssim1.8 eV. The 1s1s exciton is mostly composed of one conduction and one valence band state [22], therefore we can approximate the 1s1s1s\to 1s exciton-phonon matrix element as (see SM [22] for details)

g1s,1sZO𝐤|A𝐤cv1s|2(g𝐤cZOg𝐤vZO),g_{1s,1s}^{\mathrm{ZO}}\approx\sum_{\mathbf{k}}\left|A^{1s}_{\mathbf{k}cv}\right|^{2}\left(g^{ZO}_{\mathbf{k}c}-g^{ZO}_{\mathbf{k}v}\right), (4)

where A𝐤cv1sA^{1s}_{\mathbf{k}cv} is the exciton envelope wave function for the single-electron transition |𝐤,v|𝐤,c|\mathbf{k},v\rangle\to|\mathbf{k},c\rangle. The diagonal electron-phonon matrix elements g𝐤nZOg^{\mathrm{ZO}}_{\mathbf{k}n} for a single-electron state |𝐤n|\mathbf{k}n\rangle within the framework of DFT are given by [41]

g𝐤nZO=d3𝐫|ψ𝐤n(𝐫)|2ZOVKS(𝐫).g^{\mathrm{ZO}}_{\mathbf{k}n}=\int\mathrm{d}^{3}\mathbf{r}\,|\psi_{\mathbf{k}n}(\mathbf{r})|^{2}\partial_{{}_{\mathrm{ZO}}}V_{\text{KS}}(\mathbf{r}). (5)

Here, ψ𝐤n(𝐫)\psi_{\mathbf{k}n}(\mathbf{r}) denotes the one-electron wave function for state |𝐤n|\mathbf{k}n\rangle, VKS(𝐫)V_{\mathrm{KS}}(\mathbf{r}) corresponds to the total self-consistent Kohn-Sham (KS) potential, and ZO\partial_{\mathrm{ZO}} represents the directional derivative along the ZO phonon mode displacement vector [41]. Combining Eqs. (4) and (5), we obtain

g1s,1sZOd3𝐫[nc1s(𝐫)nv1s(𝐫)]ZOVKS(𝐫),g_{1s,1s}^{\mathrm{ZO}}\approx\int\mathrm{d}^{3}\mathbf{r}\,\left[n^{1s}_{c}(\mathbf{r})-n^{1s}_{v}(\mathbf{r})\right]\partial_{{}_{\mathrm{ZO}}}V_{\mathrm{KS}}(\mathbf{r}), (6)

where nc(v)1s(𝐫)=𝐤,v(c)|A𝐤cv1s|2|ψ𝐤c(v)(𝐫)|2n^{1s}_{c(v)}(\mathbf{r})=\sum_{\mathbf{k},v(c)}|A^{1s}_{\mathbf{k}cv}|^{2}|\psi_{\mathbf{k}c(v)}(\mathbf{r})|^{2} denotes an “exciton-averaged” electron (cc) or hole (vv) density for the 1s1s exciton.

Previous studies [9, 10, 11, 12] have suggested that the polar nature of the ZO phonon is responsible for the origin of the interlayer exciton-phonon coupling. In order to verify this claim, we separate the total perturbed potential in Eq. (6) into macroscopic and microscopic components [42]: ZOVKS(𝐫)=ZOVmicro(𝐫)+ZOVmacro(𝐫)\partial_{{}_{\mathrm{ZO}}}V_{\mathrm{KS}}(\mathbf{r})=\partial_{{}_{\mathrm{ZO}}}V_{\mathrm{micro}}(\mathbf{r})+\partial_{{}_{\mathrm{ZO}}}V_{\mathrm{macro}}(\mathbf{r}). The macroscopic part is the dipole field resulting from the polar nature of hhBN, i.e., from its Born effective charges, and in the literature is often referred to as the Fröhlich field [42]. The microscopic part of the perturbed potential is then calculated by subtracting the macroscopic part from the total perturbed potential.

In Figure 2, we illustrate the different components of the 1s1s1s\to 1s exciton-phonon matrix element, as described by Eq. (6). We plot the in-plane average of the change in KS potential (divided into microscopic and macroscopic part) and the integrated electron and hole densities as a function of the out-of-plane spatial coordinate (zz). We distinguish between the region close to the hhBN layer, the near field (grey shading), and the remaining area, the far field.

Refer to caption
Figure 2: Illustration of the different components of 1s1s1s\rightarrow 1s exciton-phonon matrix element as given in Eq. (6), plotted against the out-of-plane spatial coordinate. The horizontal orange, grey, and green lines correspond to the spatial positions of the Se, W, and hhBN atomic layers, respectively. The grey/white shading represents the near-field/far-field regime of the hhBN layer. (a) In-plane averaged macroscopic (magenta) and microscopic components (blue) of the total field. The arrows on the B and N atoms represent the ZO phonon mode displacement pattern. (b) In-plane integrated electron and hole densities of the 1s1s exciton.

Figure 2a shows the in-plane average of the microscopic (magenta line) and macroscopic (blue line) components of the total field (refer to SM [22] for computational details). The macroscopic component attains a constant value on both sides of the hhBN layer and extends entirely along the out-of-plane direction, resembling the field generated by a uniformly charged plate. Conversely, the microscopic component vanishes asymptotically due to the charge neutrality of the system. Interestingly, the microscopic part exhibits two prominent features: (i) a large and rapidly varying part in the near field, and (ii) a tiny part in the far field that is localized within the WSe2 layer, as depicted in the inset of Fig. 2b. The former arises from the change in lattice potential due to the displacement of nuclei and the induced field of the hhBN electron density; the latter is due to the induced field generated by the WSe2 electron density [41].

The different components of the total field are felt by the 1s1s exciton charge density of the WSe2 layer. In Fig. 2b, we depict the in-plane integrated electron and hole densities (nc/v1s(𝐫)n^{1s}_{c/v}(\mathbf{r})) of the 1s1s exciton. Both the electron and hole densities of the 1s1s exciton are mostly made up of tungsten dd orbitals [43]. However, while the electron density of the 1s1s exciton is completely localized within the WSe2 layer, the hole density also has a small, but finite component in the hhBN layer (see inset of Fig. 2c). This component corresponds to the hybridization of the pzp_{z} orbitals of the hhBN layer with the dd orbitals of the WSe2 layer, i.e., to the weak chemical bond between the two layers. It is noteworthy that the hybridization only occurs for the valence, i.e., bonding orbitals, and not for the conduction band orbitals, owing to the favourable or unfavorable band alignment between the hhBN and WSe2 valence and conduction bands, respectively. This hybridization has been experimentally observed and reported in Ref. [44].

Refer to caption
Figure 3: Cumulative work done by (a) the total field, (b) its macroscopic component, and (c) its microscopic component as defined in Eq. (7). The vertical dashed lines indicate the asymptotic values.

Finally, we consider the interaction of the electron and hole densities of the 1s1s exciton with the total field to gain atomistic insight on the exciton-phonon interaction across layers. The 1s1s1s\rightarrow 1s exciton-phonon matrix element can a priori receive contributions from three different channels: (i) the constant, macroscopic part interacting with the electron and hole densities over all space (ii) the large near field interacting with the small hybridized hole density in the near-field, and (iii) The small far field in the WSe2 layer interacting with the large electron and hole densities.

To quantify the contribution of each channel to the overall exciton-phonon interaction, we define the cumulative work W(z)W(z) along the z axis for a field V(𝐫)\partial V(\mathbf{r}) as

W(z)=zd3𝐫(nc1s(𝐫)nv1s(𝐫))(V(𝐫)V(𝐫)).W(z)=\int_{-\infty}^{z}\mathrm{d}^{3}\mathbf{r}\,\big{(}n^{1s}_{c}(\mathbf{r})-n^{1s}_{v}(\mathbf{r})\big{)}(\partial V(\mathbf{r})-\partial V(\mathbf{r}\rightarrow\infty)). (7)

The cumulative work for the total field asymptotically reaches the diagonal exciton-phonon matrix of the 1s1s exciton as given in Eq. 6.

In Fig. 3, we depict the cumulative work done by the total field and its constituents, denoting their asymptotic values with dashed, vertical lines. The cumulative work for all component fields starts from zero and reaches the asymptotic value by the end of the near-field volume. This implies that the coupling of the 1s1s exciton to the ZO phonon arises almost entirely from the electric fields in the near-field region and their interaction with the hybridized part of the charge density of the 1s1s exciton. Given that only the hole of the 1s1s exciton hybridizes with the hhBN orbitals, the 1s1s1s\rightarrow 1s exciton-phonon scattering can be interpreted as being primarily due to the scattering of the hole by the ZO phonon.

We further examine the contributions of each individual field to the total exciton-phonon coupling by considering the cumulative work of the macroscopic and microscopic components separately. Although the macroscopic part interacts with the entire electron and hole densities, the cumulative work reaches its asymptotic value already at the interface of the hhBN layer and the region around the WSe2 layer does not yield a net contribution. Given that the macroscopic component is nearly constant, the work done by the electron and by the hole compensate each other almost perfectly, as a consequence of the exciton being a charge neutral excitation. However, the macroscopic part does contribute a finite amount of work to the exciton-phonon coupling through its interaction with the hybridized part of the hole density. Due to the local horizontal mirror symmetry of the ZO phonon, the macroscopic part flips its sign at the hhBN layer and consequently does a finite amount of work to the hole (Fig. 3a and b).

At the same time, due to lowering of symmetry in the heterostructure, the hybridized hole density around the hhBN layer is not mirror symmetric with respect to the hhBN plane. In combination with the large microscopic field, this asymmetry leads to a finite amount of work in the near-field region, see Fig. 3c. By contrast, in the far-field regime, the electron and hole distribution are mirror symmetric with respect to the W-layer and in consequence, the net work done by the ZO mode to the exciton cancels out.

It is worthwhile to stress that, unlike the interlayer electron-phonon coupling, which is very strong and is predominantly caused by the macroscopic component in the far-field region [45], the interlayer exciton-phonon coupling is much weaker and originates in the near-field region, with nearly equal contributions coming from both macroscopic and microscopic components. Remarkably, exciton-phonon scattering takes place even in the absence of the macroscopic component. This implies that bond polarity is not a prerequisite for the occurrence of the interlayer exciton-phonon scattering, contrary to current speculations [9, 10, 11, 12], which hypothesized a macroscopic dipole-dipole coupling mechanism. Furthermore, in Fig. S5 of the SM [22], we show that the Raman intensities decreases rapidly with increasing inter-layer distance, as observed experimentally in Ref. [8], confirming the near-field effect of this phenomenon.

In summary, we theoretically and computationally investigated the coupling of excitons and phonons across layers and its signature in resonant Raman scattering. Using symmetries, we identified the main exciton-phonon scattering channel in the Raman process of a prototypical WSe2@hhBN heterostructures. We find that the exciton-phonon coupling is due to the scattering of the hybridized part of the WSe2 exciton hole with the deformation potential of the hhBN phonon. This finding sheds light on the nature of the exciton-phonon coupling, which previously was hypothesized to be due to a macroscopic, polar dipole-dipole interaction, but which we find to actually be a near-field effect. This understanding of the interlayer exciton-phonon coupling in two-dimensional heterostructures can be a valuable guide for tailoring inter-layer interactions in such systems.

Acknowledgements– We acknowledge the use of the HPC facilities of the University of Luxembourg [46]. S.R. acknowledges funding from the National Research Fund (FNR) Luxembourg, project C20/MS/14802965/RESRAMAN. L.W. acknowledges funding by the FNR through project C22/MS/17415967/ExcPhon. We thank Aseem Rajan Kshirsagar and Henry Fried for helpful discussions.

References

  • Geim and Grigorieva [2013] A. K. Geim and I. V. Grigorieva, Nature 499, 419 (2013).
  • Rivera et al. [2016] P. Rivera, K. L. Seyler, H. Yu, J. R. Schaibley, J. Yan, D. G. Mandrus, W. Yao, and X. Xu, Science 351, 688 (2016).
  • Cao et al. [2018] Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras, and P. Jarillo-Herrero, Nature 556, 43 (2018).
  • 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, M. Blei, J. D. Carlström, K. Watanabe, T. Taniguchi, S. Tongay, M. Crommie, A. Zettl, and F. Wang, Nature 579, 359 (2020).
  • Chen et al. [2008] J.-H. Chen, C. Jang, S. Xiao, M. Ishigami, and M. S. Fuhrer, Nature Nanotechnology 3, 206 (2008).
  • Wang et al. [2012] Q.-Y. Wang, Z. Li, W.-H. Zhang, Z.-C. Zhang, J.-S. Zhang, W. Li, H. Ding, Y.-B. Ou, P. Deng, K. Chang, et al., Chinese Physics Letters 29, 037402 (2012).
  • Jin et al. [2017] C. Jin, J. Kim, J. Suh, Z. Shi, B. Chen, X. Fan, M. Kam, K. Watanabe, T. Taniguchi, S. Tongay, A. Zettl, J. Wu, and F. Wang, Nature Physics 13, 127 (2017).
  • Li et al. [2022] Y. Li, X. Zhang, J. Wang, X. Ma, J.-A. Shi, X. Guo, Y. Zuo, R. Li, H. Hong, N. Li, K. Xu, X. Huang, H. Tian, Y. Yang, Z. Yao, P. Liao, X. Li, J. Guo, Y. Huang, P. Gao, L. Wang, X. Yang, Q. Dai, E. Wang, K. Liu, W. Zhou, X. Yu, L. Liang, Y. Jiang, X.-Z. Li, and L. Liu, Nano Letters 22, 2725 (2022).
  • Jadczak et al. [2021] J. Jadczak, J. Kutrowska-Girzycka, J. J. Schindler, J. Debus, K. Watanabe, T. Taniguchi, C.-H. Ho, and L. Bryja, Materials 14 (2021).
  • Du et al. [2019] L. Du, Y. Zhao, Z. Jia, M. Liao, Q. Wang, X. Guo, Z. Shi, R. Yang, K. Watanabe, T. Taniguchi, J. Xiang, D. Shi, Q. Dai, Z. Sun, and G. Zhang, Phys. Rev. B 99, 205410 (2019).
  • Du et al. [2018] L. Du, M. Liao, J. Tang, Q. Zhang, H. Yu, R. Yang, K. Watanabe, T. Taniguchi, D. Shi, Q. Zhang, and G. Zhang, Phys. Rev. B 97, 235145 (2018).
  • Chow et al. [2017] C. M. Chow, H. Yu, A. M. Jones, J. Yan, D. G. Mandrus, T. Taniguchi, K. Watanabe, W. Yao, and X. Xu, Nano Letters 17, 1194 (2017).
  • Chen et al. [2019] C. Chen, X. Chen, H. Yu, Y. Shao, Q. Guo, B. Deng, S. Lee, C. Ma, K. Watanabe, T. Taniguchi, J.-G. Park, S. Huang, W. Yao, and F. Xia, ACS Nano 13, 552 (2019).
  • Hennighausen et al. [2023] Z. Hennighausen, J. Moon, K. M. McCreary, C. H. Li, O. M. J. van ’t Erve, and B. T. Jonker, ACS Nano 17, 2529 (2023).
  • Chan et al. [2023] Y.-h. Chan, J. B. Haber, M. H. Naik, J. B. Neaton, D. Y. Qiu, F. H. da Jornada, and S. G. Louie, Nano Letters 23, 3971 (2023).
  • Viner et al. [2021] J. J. S. Viner, L. P. McDonnell, P. Rivera, X. Xu, and D. C. Smith, Phys. Rev. B 104, 165404 (2021).
  • Reichardt and Wirtz [2019] S. Reichardt and L. Wirtz, Phys. Rev. B 99, 174312 (2019).
  • Reichardt and Wirtz [2020] S. Reichardt and L. Wirtz, Science Advances 6, eabb5915 (2020).
  • Dancoff [1950] S. M. Dancoff, Phys. Rev. 78, 382 (1950).
  • Hybertsen and Louie [1986] M. S. Hybertsen and S. G. Louie, Phys. Rev. B 34, 5390 (1986).
  • Rohlfing and Louie [2000] M. Rohlfing and S. G. Louie, Phys. Rev. B 62, 4927 (2000).
  • [22] See Supplemental Material at [URL] for the ab-initio calculation details, convergence studies, derivation of selection rules, GW band structures, distance dependence of Raman intenties, and reciprocal space plots of 1s1s and 2s2s excitons.
  • Perdew et al. [1996] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
  • Giannozzi et al. [2017] P. Giannozzi, O. Andreussi, T. Brumme, O. Bunau, M. B. Nardelli, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, M. Cococcioni, N. Colonna, I. Carnimeo, A. D. Corso, S. de Gironcoli, P. Delugas, R. A. DiStasio, A. Ferretti, A. Floris, G. Fratesi, G. Fugallo, R. Gebauer, U. Gerstmann, F. Giustino, T. Gorni, J. Jia, M. Kawamura, H.-Y. Ko, A. Kokalj, E. Küçükbenli, M. Lazzeri, M. Marsili, N. Marzari, F. Mauri, N. L. Nguyen, H.-V. Nguyen, A. O. de-la Roza, L. Paulatto, S. Poncé, D. Rocca, R. Sabatini, B. Santra, M. Schlipf, A. P. Seitsonen, A. Smogunov, I. Timrov, T. Thonhauser, P. Umari, N. Vast, X. Wu, and S. Baroni, Journal of Physics: Condensed Matter 29, 465901 (2017).
  • Marini et al. [2009] A. Marini, C. Hogan, M. Grüning, and D. Varsano, Computer Physics Communications 180, 1392 (2009).
  • Sangalli et al. [2019] D. Sangalli, A. Ferretti, H. Miranda, C. Attaccalite, I. Marri, E. Cannuccia, P. Melo, M. Marsili, F. Paleari, A. Marrazzo, G. Prandini, P. Bonfà, M. O. Atambo, F. Affinito, M. Palummo, A. Molina-Sánchez, C. Hogan, M. Grüning, D. Varsano, and A. Marini, Journal of Physics: Condensed Matter 31, 325902 (2019).
  • Hamann [2013] D. R. Hamann, Phys. Rev. B 88, 085117 (2013).
  • Scherpelz et al. [2016] P. Scherpelz, M. Govoni, I. Hamada, and G. Galli, Journal of Chemical Theory and Computation 12, 3523 (2016).
  • Pizzi et al. [2020] G. Pizzi, V. Vitale, R. Arita, S. Blügel, F. Freimuth, G. Géranton, M. Gibertini, D. Gresch, C. Johnson, T. Koretsune, J. Ibañez-Azpiroz, H. Lee, J.-M. Lihm, D. Marchand, A. Marrazzo, Y. Mokrousov, J. I. Mustafa, Y. Nohara, Y. Nomura, L. Paulatto, S. Poncé, T. Ponweiser, J. Qiao, F. Thöle, S. S. Tsirkin, M. Wierzbowska, N. Marzari, D. Vanderbilt, I. Souza, A. A. Mostofi, and J. R. Yates, Journal of Physics: Condensed Matter 32, 165902 (2020).
  • Sohier et al. [2017] T. Sohier, M. Calandra, and F. Mauri, Phys. Rev. B 96, 075448 (2017).
  • Lazić [2015] P. Lazić, Computer Physics Communications 197, 324 (2015).
  • Guandalini et al. [2023] A. Guandalini, P. D’Amico, A. Ferretti, and D. Varsano, npj Computational Materials 9, 44 (2023).
  • Bruneval and Gonze [2008] F. Bruneval and X. Gonze, Phys. Rev. B 78, 085125 (2008).
  • Gerber and Marie [2018] I. C. Gerber and X. Marie, Phys. Rev. B 98, 245126 (2018).
  • Poulson et al. [2013] J. Poulson, B. Marker, R. A. van de Geijn, J. R. Hammond, and N. A. Romero, ACM Trans. Math. Softw. 3910.1145/2427023.2427030 (2013).
  • Godby and Needs [1989] R. W. Godby and R. J. Needs, Phys. Rev. Lett. 62, 1169 (1989).
  • Momma and Izumi [2008] K. Momma and F. Izumi, Journal of Applied Crystallography 41, 653 (2008).
  • McDonnell et al. [2020] L. P. McDonnell, J. J. S. Viner, P. Rivera, X. Xu, and D. C. Smith, 2D Materials 7, 045008 (2020).
  • Aslan et al. [2021] B. Aslan, C. Yule, Y. Yu, Y. J. Lee, T. F. Heinz, L. Cao, and M. L. Brongersma, 2D Materials 9, 015002 (2021).
  • He et al. [2014] K. He, N. Kumar, L. Zhao, Z. Wang, K. F. Mak, H. Zhao, and J. Shan, Phys. Rev. Lett. 113, 026803 (2014).
  • Giustino [2017] F. Giustino, Rev. Mod. Phys. 89, 015003 (2017).
  • Verdi and Giustino [2015] C. Verdi and F. Giustino, Phys. Rev. Lett. 115, 176401 (2015).
  • Wang et al. [2018] G. Wang, A. Chernikov, M. M. Glazov, T. F. Heinz, X. Marie, T. Amand, and B. Urbaszek, Rev. Mod. Phys. 90, 021001 (2018).
  • Magorrian et al. [2022] S. J. Magorrian, A. J. Graham, N. Yeung, F. Ferreira, P. V. Nguyen, A. Barinov, V. I. Fal’ko, N. R. Wilson, and N. D. M. Hine, 2D Materials 9, 045036 (2022).
  • Sohier et al. [2021] T. Sohier, M. Gibertini, and M. J. Verstraete, Phys. Rev. Mater. 5, 024004 (2021).
  • Varrette et al. [2022] S. Varrette, H. Cartiaux, S. Peter, E. Kieffer, T. Valette, and A. Olloh, in Proc. of the 6th ACM High Performance Computing and Cluster Technologies Conf. (HPCCT 2022) (Association for Computing Machinery (ACM), Fuzhou, China, 2022).
\foreach\x

in 1,…,0 See pages \x, of supplement.pdf