<<

Excitons in solids from periodic

equation-of-motion coupled-cluster theory

Xiao Wang† and Timothy C. Berkelbach∗,†,‡

†Center for Computational Quantum Physics, Flatiron Institute, New York, New York 10010 USA ‡Department of Chemistry, Columbia University, New York, New York 10027 USA

E-mail: [email protected]

Abstract

We present an ab initio study of electronically excited states of three-dimensional solids

using -based periodic equation-of-motion coupled-cluster theory with single and dou-

ble excitations (EOM-CCSD). The explicit use of translational symmetry, as implemented via

Brillouin zone sampling and momentum conservation, is responsible for a large reduction in

cost. Our largest system studied, which samples the Brillouin zone using 64 k-points (a 4×4×4

mesh) corresponds to a canonical EOM-CCSD calculation of 768 electrons in 640 orbitals. We

study eight simple main-group semiconductors and insulators, with direct singlet excitation en-

ergies in the range of 3 to 15 eV. Our predicted excitation energies exhibit a mean signed error

of 0.24 eV and a mean absolute error of 0.27 eV when compared to experiment. Although

this error is similar to that found for EOM-CCSD applied to , it may also reflect the

role of vibrational effects, which are neglected in the calculations. Our results support recently

proposed revisions of experimental optical gaps for AlP and cubic BN. We furthermore cal-

culate the energy of excitons with nonzero momentum and compare the exciton dispersion of

LiF with experimental data from inelastic X-ray scattering. By calculating excitation energies

under strain, we extract hydrostatic deformation potentials in order to quantify the strength of

1 interactions between excitons and acoustic phonons. Our results indicate that coupled-cluster

theory is a promising method for the accurate study of a variety of exciton phenomena in solids.

1 Introduction

Density functional theory (DFT) is the workhorse of computational materials science and com- monly predicts ground-state properties with high accuracy.1–8 Extending this success to the pre- diction of excited-state properties is an ongoing challenge. Time-dependent DFT (TDDFT)9 strug- gles to treat neutral excitations in the condensed phase and reasonable accuracy requires the use of hybrid functionals and/or frequency-dependent exchange-correlation kernels.10–20 Instead, the community typically builds upon DFT through Green’s function-based many-body .13 For weakly-correlated materials, including simple metals, semiconductors, and insula- tors, the GW approximation to the self-energy21–26 combined with the same approximation to the Bethe-Salpeter equation27–31 yields properties that are typically in good agreement with experiment. Strongly-correlated materials are more commonly treated via dynamical mean- field theory (DMFT),32,33 either in the DFT+DMFT framework34–36 or the GW+DMFT frame- work.37,38 Alternatively, methods have been adapted for the calculation of excitation energies, including variational39 and diffusion Monte Carlo40–42 and auxiliary-field quantum Monte Carlo.43 In recent years, wavefunction-based techniques from the community have been adapted for periodic boundary conditions and applied to condensed-phase systems. For ground-state properties, we mention the application of Møller-Plesset perturbation theory,44–47 coupled-cluster theory,48–52 and full configuration interaction quantum Monte Carlo.53–55 The same methods have also been used for finite-temperature properties.56–58 Charged excitation energies, as quantified through the one-particle spectral function or the band structure, have been studied by periodic equation-of-motion coupled-cluster theory for the uniform electron gas,59 for the sim-

52 60 ple semiconductors silicon and diamond, for two-dimensional MoS2, and for transition metal

2 oxides.61 Neutral excitation energies have been primarily studied by configuration interaction with sin- gle excitations (CIS),62,63 which is equivalent to the use of the Hartree-Fock self-energy in the Bethe-Salpeter equation (with the Tamm-Dancoff approximation). Due to the unscreened nature of the Coulomb interaction, periodic CIS typically predicts excitation energies which are much too large.64 Recently, periodic equation-of-motion coupled-cluster theory with single and double excitations (EOM-CCSD) was applied to one-dimensional polyethylene in a small single-particle ,50 producing results in reasonable agreement with experiment. For three-dimensional condensed-phase systems, our group has recently applied EOM-CCSD to the neutral excited-state properties of the uniform electron gas65 as well as the absorption and pump-probe spectroscopy of the naphthalene crystal at the Γ point.66 Here, we continue this endeavor and present the results of a new EOM-CCSD implemen- tation with translational symmetry and Brillouin zone sampling for three-dimensional atomistic solids. The layout of the article is as follows. In section2, we describe the theory underlying our implementation, including details about our symmetry-adapted Gaussian basis sets and periodic EOM-CCSD. In section3, we present computational details, including details about the materials studied, the basis sets used, and integral evaluation. In section4, we provide EOM-CCSD results on the excitation energies (including a comparison with CIS and a discussion of finite-size effects), the exciton binding energy, the dispersion of excitons with nonzero momentum, and the exciton- phonon interaction. In section5, we summarize our results and conclude with future directions.

2 Theory

Our periodic calculations are performed using a translational-symmetry-adapted single-particle basis,

X ik·T φµk(r) = e φ˜µ(r − T ), (1) T

3 where φ˜µ(r) is an -centered Gaussian orbital, T is a lattice translation vector, and k is a crystal momentum vector sampled from the first Brillouin zone. A Hartree-Fock calculation in this basis produces crystalline orbitals (COs)

X ψpk(r) = Cµp(k)φµk(r), (2) µ

where p is the band index and Cµp(k) are the CO coefficients. The periodic CCSD energy and cluster amplitudes are determined by the usual equations,49,67–71

E0 = hΦ0|H¯ |Φ0i, (3)

0 = hΦaka |H¯ |Φ i, (4) iki 0 0 = hΦakabkb |H¯ |Φ i, (5) iki jk j 0 where Φaka and Φakabkb are Slater determinants with one and two electron-hole pairs, indices i, j, k, l iki iki jk j denote occupied orbitals, and a, b, c, d denote virtual orbitals. The similarity transformed Hamil-

−T T tonian is given by H¯ ≡ e He , where T = T1 + T2 is a momentum-conserving cluster operator with

X X0 aka † T = t a a k , (6) 1 iki aka i i ai kaki X X0 1 akabkb † † T2 = t a a a jk aik . (7) 4 iki jk j aka bkb j i abi j kakbkik j

The primed summations indicate momentum conservation. Excited states are accessed in coupled-cluster theory using the equation-of-motion (EOM) for- malism,71–76 which amounts to diagonalizing the effective Hamiltonian H¯ in a truncated space of excitations. For neutral excitations considered in this work, we use electronic excitation (EE) EOM-CCSD, where the diagonalization is performed in the basis of 1-particle+1-hole and 2- particle+2-hole states. We study excitations with zero and nonzero momentum q by using a basis

4 of determinants with the corresponding momentum. The (right-hand) excited state is therefore given by

T |Ψ(q)i = [R1(q) + R2(q)] e |Φ0i (8) with

X X0 aka † R (q) = r a a k , (9) 1 iki aka i i ai kaki X X0 1 akabkb † † R2(q) = r a a a jk aik , (10) 4 iki jk j aka bkb j i abi j kakbkik j

where the primed summations indicate that the momenta sum to q. The use of translational sym-

2 4 4 metry in CCSD leads to a computational cost that scales like o v Nk , where o is the number of

occupied orbitals per unit cell, v is the number of virtual orbitals per unit cell, and Nk is the number of k-points sampled. Further details about periodic Gaussian-based Hartree-Fock and CCSD can be found in Ref. 52.

3 Computational Details

We study eight semiconducting and insulating materials featuring the diamond/zinc-blende crystal structure and the rock salt crystal structure. The materials have a wide range of both direct and indirect band gaps and a variety of ionic and covalent bonding. The eight materials and the lattice constants used are given in Tab.1. In all calculations, the Brillouin zone was sampled with a

77 uniform Monkhorst-Pack mesh of Nk k-points that includes the Γ point. Our calculations are performed with GTH pseudopotentials,78,79 although we perform some all- electron calculations for comparison. For pseudopotential calculations, we use the corresponding polarized double- and triple-zeta basis sets DZVP and TZVP.80 For all-electron calculations, we use a modification of the cc-pVDZ basis set presented in Ref. 64, denoted AE-PVDZ. The finite-size errors of periodic calculations are influenced by the treatment of two-electron repulsion integrals (ERIs). Many of these integrals are formally divergent, due to the long-range

5 nature of the Coulomb interaction; however, these divergent ERIs enter into expressions for ob- servables as integrable divergences, producing well-defined results in the thermodynamic limit. In order to avoid divergent ERIs, we calculate all atomic orbital ERIs as

Z Z kµkν kκkλ −1 ρµν (r1)ρκλ (r2) (µkµ, νkν|κkκ, λkλ) = Nk dr1 dr2 , (11) |r1 − r2|

where the orbital-pair densities have had their net charge removed,

ρkµkν (r) = φ∗ (r)φ (r) − ρ , (12a) µν µkµ νkν µν 1 Z ρ = drφ∗ (r)φ (r), (12b) µν µkµ νkν NkΩ

and Ω is the volume of the unit cell. The ERIs are then calculated using periodic Gaussian den- sity fitting with an even-tempered auxiliary basis as described in Ref. 81. We note that the use of chargeless pair densities is equivalent to neglecting the G = 0 component of the ERIs when calculated in a plane-wave basis. At the HF level, this treatment of ERIs produces an energy that

−1/3 converges to the thermodynamic limit as Nk , due to the exchange energy; this can be corrected −1 82,83 with a Madelung constant, leading to Nk convergence. This particular finite-size behavior is also present in the occupied orbital energies εi(k). We use a closed-shell implementation of periodic EOM-CCSD and study singlet excitations in this work, which are calculated by Davidson diagonalization.84,85 The initial guess used in the iterative diagonalization is obtained from dense diagonalization of the effective Hamiltonian in the single excitation subspace. All calculations were performed with the open-source PySCF software package.86 Our imple- mentation of EOM-CCSD with k-point symmetry is not yet parallelized; for the largest systems studied here, these serial calculations take a few days on a single core.

6 4 Results and Discussion

4.1 Direct optical excitation energy

In this section, we study the performance of periodic EOM-CCSD on the lowest-lying direct sin- glet excitation energies for the eight selected solids. Such states are relevant for absorption spec- troscopy where no momentum is transferred during excitation. In periodic calculations, it is desirable to achieve convergence with respect to the single-particle basis set, the level of correlation, and Brillioun zone sampling. The first two categories (extensively characterized in molecular systems) have been studied to some extent in solids53,87,88 and the convergence behavior of the third one is well understood at the mean-field level.83,89,90 For ground-state energies, the finite-size behavior has been extensively discussed in the quantum Monte Carlo community91,92 and only more recently from the quantum chemistry perspective.51,53,93–97 The finite-size behavior of excitation energies has received considerably less attention.41,42,52 We hope our preliminary results and discussion presented below encourage future work in this direction.

4.1.1 CIS

As a warm-up to EOM-CCSD, we first present results for periodic configuration interaction with single excitations (CIS), which forms a minimal theory for electronic excited states in the con- densed phase. Importantly, the relatively low cost of CIS allows us to study the convergence with respect to Brillouin zone sampling up to relatively large k-point meshes (either 73 or 83). The finite-size error of the CIS excitation energy can be analyzed approximately by considering it as the difference between the energy of two determinants, the HF one |Φ0i and a single excitation

ak |Φik i. As discussed above, the energy of a single determinant calculated using our handling of ERIs −1/3 exhibits a finite-size error decaying like Nk , but which can be removed by a Madelung constant that depends only on the number of electrons in the unit cell. Therefore, the correction is identical for both states, and this leading-order error cancels in the energy difference. We thus posit that the

7 Figure 1: Excitation energy of LiF calculated with CIS using the DZVP basis set. The dashed line is a fit to results obtained with a number of k-points between 33 and 73. The dotted line includes only 23, 33, and 43 k-points. Both fits use the functional form given in Eq. 13.

−1 CIS energy converges at least as fast as Nk . In Fig.1, we show the excitation energy predicted by CIS for the LiF crystal as a function of

−1 −1 Nk . Clearly, the finite-size error decays at least as fast as Nk and so we use the three-parameter fitting function

−1 −2 E(Nk) = E∞ + a Nk + b Nk , (13)

in order to extrapolate to the thermodynamic limit (Nk → ∞). This fit is shown by the dashed line in Fig.1, which includes results between 3 3 and 73 k-points. Using an all-electron double-zeta basis set, we performed CIS calculations on a subset of our eight materials, in order to compare to previously published CIS results from Lorenz et al.64 By extrapolation of our results obtained at large k-point meshes using the above form, we obtain converged excitation energies (the “AE-PVDZ/Conv” column of Tab.1) that are within 0.1 eV of Ref. 64 for all materials studied. However, we emphasize that the CIS excitation energies are larger than experiment by 2 eV or more. Before moving on to our EOM-CCSD results, we use CIS to assess some of the future approxi- mations we will have to make. In particular, we will only access k-point meshes up to 4 × 4 × 4 and

8 we will use GTH pseudopotentials and corresponding DZVP basis sets. First, considering finite- size errors, we re-fit the CIS data using the same form but excluding all k-point meshes larger than 4 × 4 × 4; for LiF, the result of this fit is shown as the dotted line in Fig.1. Clearly, without larger k-point meshes, this extrapolation predicts an excitation energy which is too high by about 0.2 eV. These limited extrapolation results are shown for all materials in the “AE-PVDZ/234-Extrap” col- umn of Tab.1 and exhibit errors of about ±0.5 eV. Second, considering pseudopotential errors, in the last two columns of Tab.1, we show the excitation energies calculated with GTH pseudopoten- tials. In many cases, the pseudopotential error is less than 0.1 eV; naturally, materials containing heavier second-row such as Si or Mg exhibit the largest errors, which are about 0.3 eV.

4.1.2 EOM-CCSD

We now move on to our results from periodic EOM-CCSD. Again, due to the high cost of these calculations, we utilized GTH pseudopotentials, sampled the Brillouin zone with meshes up to 4×4×4, and used basis set corrections. In particular, our primary calculations were based on a HF reference obtained with the full DZVP basis set; subsequent CCSD and EOM-CCSD calculations then employed frozen virtual orbitals, typically correlating 4 lowest virtual bands. The results of these calculations were used to extrapolate to the thermodynamic limit using the same empirical

Table 1: Singlet excitation energies (eV) from all-electron and pseudopotential-based periodic CIS

System a (Å) AE-PVDZ DZVP 234-Extrapa Convb Ref. 64 234-Extrapa Convb Diamond 3.567 10.99 11.76 11.72 11.07 11.81 Si 5.431 5.84 6.14 6.05 5.43 5.71 SiC 4.360 9.41 9.83 9.74 9.12 9.47 LiF 3.990 16.02 15.85 15.84 16.06 15.87 LiCl 5.130 11.04 10.89 MgO 4.213 12.00 11.91 11.94 11.69 11.66 BN 3.615 14.17 14.32 AlP 5.451 6.59 6.76 a Extrapolation based on results obtained with 23, 33, and 43 k-point meshes b Extrapolation based on results obtained with 33 up to 73 k-point meshes

9 Table 2: Singlet excitation energies (eV) from periodic EOM-CCSD

a b c System 234-Extrap ∆frz ∆TZ Final Expt. Diamond 7.70 −0.18 −0.05 7.47 7.398 Si 3.96 −0.37 −0.07 3.52 3.499 SiC 6.53 −0.19 −0.08 6.27 6.0100 LiF 14.29 −0.82 +0.01d 13.48 12.7,101 13.68102 LiCl 9.62 −0.27 −0.07 9.29 8.9103 MgO 8.55 −0.19 −0.07 8.29 7.5,104 7.6105 BN 11.48 −0.35 −0.02 11.11 11106 AlP 4.97 −0.42 −0.07 4.48 4.6107,108 MSE 0.24 MAE 0.27 a Extrapolation based on frozen-virtual DZVP results obtained with 23, 33, and 43 k-point meshes b ∆frz is the energy difference between complete DZVP and frozen- virtual DZVP on a 3×3×3 k-point grid c ∆TZ is the energy difference between TZVP and DZVP on a 2×2×2 k-point grid d For LiF, the TZVP basis set has severe linear dependencies and was modified by doubling the exponents of the two most diffuse s- and p-type primitive Gaussian functions of Li (all basis functions of F are unchanged) formula as described above (Eq. 13). These results are given in the “234-Extrap” column of Tab.2.

To these values, we then added two basis set corrections, ∆frz and ∆TZ; ∆frz is the energy difference between complete and frozen-orbital DZVP calculations with a 3 × 3 × 3 k-point mesh; ∆TZ is the energy difference between TZVP and DZVP calculations with a 2 × 2 × 2 k-point mesh. Whereas

∆frz is typically between 0.2 and 0.8 eV, ∆TZ is less than 0.1 eV.

−1 Basis-set corrected excitation energies as a function of Nk are shown in Figure2 for diamond, Si, LiF, and MgO. Our final values for all eight materials are given in the “Final” column of Tab.2 and compared to experiment. Unsurprisingly, EOM-CCSD is a massive improvement over CIS; for the eight solids studied here, EOM-CCSD predicts excitation energies with a mean signed error (MSE) of 0.24 eV and a mean absolute error (MAE) of 0.27 eV. A few results in Tab.2 are noteworthy. First, we note that the excitation energy of cubic BN is frequently reported as 6.4 eV, which is almost 5 eV lower than the EOM-CCSD prediction. How-

10 ever, a GW/BSE calculation reported in 2004109 also found a value of around 11 eV and proposed a reinterpretation of the experimental data. Indeed, a joint theory-experiment paper published in 2018106 attributed the lower energy absorption features to a combination of defects and domains of hexagonal BN, further supporting a direct excitation energy of about 11 eV. Second, the exci- tation energy of AlP has frequently been reported as 3.6 eV, about 1 eV below the EOM-CCSD prediction. A 2019 publication reporting the results of TDDFT and GW/BSE calculations108 sug- gested that the 3.6 eV feature seen in experimental spectra is due to the indirect transition of AlP. Experimental spectra show a much stronger peak at around 4.6 eV,107 which is the likely value of the first direct excitation energy. The two materials with the largest error are LiF and MgO. Whereas absorption spectra of LiF typically show a narrow peak at about 12.7 eV101 (leading to an overprediction of 0.8 eV), inelastic X-ray scattering data is consistent with a value of 13.68 eV.102 For MgO, EOM-CCSD overpredicts the excitation energy by about 0.7 eV. Dittmer et al.110 very recently applied a flavor of similarity-transformed EOM-CCSD (which we emphasize is different from EOM-CCSD111) to the calculation of band gaps and optical gaps of solids using a cluster-based approach, to be contrasted with the periodic approach taken here. Interestingly, the two solids that are shared between our works are these same two outliers, MgO and LiF, for which that work predicts optical gaps of 6.32 and 11.41 eV, respectively. Based on the experimental values identified in Tab.2, these values underestimate the experimental ones by about 1.3 eV. The tendency of vanilla EOM-CCSD to overestimate excitation energies is the same as the one typically observed in molecules and could potentially be addressed via inclusion of triple excita- tions. However, we also emphasize that our calculations include no information about exciton- phonon effects, which are expected to reduce the excitation energy from its purely electronic value.112–116 For this reason, our electronic optical gaps might actually be more accurate than sug- gested here.

11 Figure 2: Excitation energy of diamond, Si, LiF, and MgO calculated with CIS and EOM-CCSD. Results are extrapolated to Nk → ∞ using the functional form given in Eq. 13. The dashed lines are fitted based on results obtained with a number of k-points between 33 and 73. The dotted lines include only 23, 33, and 43 k-points. A variety of experimental results are indicated by solid lines.

4.2 Exciton binding energy

We now consider the exciton binding energy, defined as the difference between the fundamental band gap and the first neutral excitation energy (i.e. the optical gap). Within periodic coupled- cluster theory, the band gap is obtained from the calculation of the ionization potential (IP-EOM- CCSD) and the electron affinity (EA-EOM-CCSD), as described in Ref. 52. Here, our IP/EA- EOM-CCSD calculations are basis-set-corrected in the same way as described for our EOM-CCSD calculations. We focus on LiF, which is a direct-gap ionic insulator with a concomitantly large exciton bind- ing energy. The minimum band gap occurs at the Γ point, which is where we calculate the IP and EA. In Fig.3(a), we show the fundamental (IP /EA) gap and the optical (EE) gap, as a function of the number of k-points sampled in the Brillouin zone. Clearly the fundamental gap is larger than

12 Figure 3: Fundamental band gap and optical gap (a) and exciton binding energy (b) of LiF from periodic EOM-CCSD. As discussed in the text, the optical gap is fit to the functional form given in Eq. 13. Due to the behavior of IP/EA-EOM-CCSD, the fundamental band gap (red squares) and exciton binding energy (brown diamonds) are extrapolated to Nk → ∞ assuming a finite-size −1/3 scaling of the form ENk = E∞ + a Nk .

the optical gap such that the exciton binding energy is positive, as expected. The exciton binding energy is plotted in Fig.3(b), and seen to be around 0.8 eV with a 4 × 4 × 4 k-point mesh. Unlike neutral excitation energies, which conserve particle number, the IP and EA are charged excitation energies corresponding to a change in particle number. In particular, the same approximation con- sidered above for CIS, i.e. the use of a single determinant, leads to Koopmans’ approximation to

the ionization potential, IP≈ −εi; as discussed above, occupied orbital energies converge slowly

with Nk when the ERIs are handled as described in section3. Therefore, as shown in Fig.3(b), we fit the exciton binding energies calculated with 3 × 3 × 3 and 4 × 4 × 4 grids to the form

−1/3 E(Nk) = E∞ + aNk , in order to extrapolate to the thermodynamic limit. Doing so gives 1.47 eV, which is in good agreement with the experimental value of 1.6 eV.105 We note that if we sepa- rately extrapolate the band gap and the optical gap and take the difference, we get a larger value of 2.74 eV.

13 4.3 Exciton dispersion

Although optical absorption spectroscopy and modern theoretical approaches have primarily fo- cused on excitons with q = 0, it is important to also consider excitons that carry a finite mo- mentum q. For example, electron-hole pairs with finite momentum are realized in many indirect semiconductors and are also important for a quantitative modeling of the exciton-phonon interac- tion, excitonic dynamics, and radiative lifetimes. The simulation of excitons with finite-momentum is straightforward in EOM-CCSD, and sim- ply requires that the involved crystal momenta sum to q in Eqs.9 and 10. The EOM Hamiltonian is block-diagonal with respect to the exciton momentum and thus all accessible momenta can be studied independentaly and calculated in parallel. Because the exciton momentum q corresponds to a momentum difference, a periodic calculation only has access to values q = k − k0, where k are momenta from the k-point mesh. Therefore, different k-point meshes can be utilized in order to access different values of the exciton momentum q, albeit with an impact on the finite-size error. Again we focus on LiF, for which we show the EOM-CCSD exciton dispersion in Fig.4. Our results are compared to inelestic X-ray scattering (IXS) spectroscopy measurements performed by Abbamonte et al.102 (open circles). We utilize 3 × 3 × 3 k-point mesh (up-pointing triangles) and 4×4×4 k-point mesh (down-pointing triangles) in order to access more momenta q in the Brillouin

Figure 4: Exciton dispersion of LiF. Periodic EOM-CCSD data are obtained with different k-point meshes in order to access more values of the momentum transfer q. Rigid shifts were applied, as described in the text. Also shown are the experimental inelastic X-ray scattering (IXS) data from Ref. 102 and the tight-binding (TB) model fitted to that data.

14 Figure 5: Ground- and excited-state potential energy surface of diamond as a function of the unit cell volume (a) and the behavior of the excitation energy as a function of hydrostatic strain (b). Results are obtained from CCSD and EOM-CCSD with the DZVP basis set and a 3 × 3 × 3 sampling of the Brillouin zone.

zone. The 3 × 3 × 3 values were rigidly shifted in order to achieve agreement at the Γ point and subsequently, both dispersion data were rigidly shifted in order to place the exciton band center at 14.2 eV, as was observed experimentally. From their experimental data along the Γ − X line, those authors paramaterized a tight-binding model (with band center 14.2 eV and nearest-neighbor transfer integral −0.065 eV), which we have extended to the entire Brillouin zone for comparison (dashed line). We can see that the EOM-CCSD results are in good agreement with the IXS data, with an error less than 0.2 eV along the Γ → X direction. Our largest discrepancy is at the L point, although we emphasize that experimental data is unavailable at that momentum and the disagreement may indicate a failure of the simple tight-binding model.

4.4 Exciton-phonon interaction

Finally, as an additional demonstration of the solid-state problems that can be studied by EOM- CCSD, we consider the behavior of excitation energies as a function of lattice strain. As an ex- ample, in Fig.5(a) we show the ground-state and first excited-state potential energy surfaces of diamond, associated with hydrostatic strain, i.e. isotropic variation of the unit cell. While the

15 ground-state energy minimum occurs at a lattice constant of 3.567 Å (which is fortuitously the exact experimental value5,117), the excited state has a minimum which is shifted to a larger lattice constant of 3.674 Å. The behavior of the excitation energy as a function of lattice strain can be used to determine properties of the exciton-phonon interaction. In particular, the interaction with acoustic phonons can be modeled within the deformation potential approximation for the change in the excitation

118,119 energy, ∆EX = 3Dε, where D is the hydrostatic deformation potential, 3ε is the trace of the

strain tensor, ε = (a − a0)/a0 is the relative strain, a is the strained lattice constant, and a0 is the unstrained lattice constant. Defined in this way, our calculations predict the excitonic hydrostatic deformation potential of diamond to be D = −2.8 eV. Repeating the same procedure for MgO, we predict a larger D = −11.7 eV, indicating a stronger coupling to hydrostatic strain than in diamond. For comparison, the use of the direct HF band gap at Γ yields D = −5.0 eV and D = −14.4 eV, indicating the importance of electron correlation. Interestingly the analogous use of the LDA band gap yields D = −2.9 eV and D = −10.0 eV, in better agreement with the EOM-CCSD result. Experimentally determined deformation potentials for exciton-phonon interactions are sporadic in the literature and we consider a direct comparison on a wider variety of materials to be a topic for future work, along with the extension to optical phonons and phonons away from the Γ point.

5 Conclusions

To summarize, we have presented the results of periodic EOM-CCSD for various neutral excited- state properties of semiconductors and insulators. The method has shown promising results for op- tical excitation energies, exciton binding energies, exciton dispersion relations, and exciton-phonon interaction energies. Collectively, these results demonstrate that EOM-CCSD is a promising and tractable approach for the study of excited-state properties of solids. While we have attempted to address finite-size errors, arising from incomplete sampling of the Brillouin zone, the high cost of EOM-CCSD precludes a definitive extrapolation. Future work

16 will be focused on both analytical and numerical exploration of finite-size errors and strategies for amelioration, such as those that have been developed for ground-state CCSD.51,94,95 In a similar direction, we are exploring the use of approximations to EOM-CCSD with reduced scaling,120 which will enable the study of simple solids with more k-points or the study of solids with more complex unit cells. Additional future work is targeted at the study of simple metals (where finite- order perturbation theory breaks down88), the study of exciton-phonon interactions beyond the deformation potential approximation, and the efficient calculation of optical spectra.

Acknowledgments

X.W. thanks Dr. Varun Rishi for helpful discussions. This work was supported in part by the National Science Foundation under Grant No. CHE-1848369. All calculations were performed using resources provided by the Flatiron Institute. The Flatiron Institute is a division of the Simons Foundation.

References

(1) Kohn, W.; Sham, L. J. Self-Consistent Equations Including Exchange and Correlation Ef- fects. Phys. Rev. 1965, 140, A1133.

(2) Kresse, G.; Furthmuller,¨ J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50.

(3) Staroverov, V. N.; Scuseria, G. E.; Tao, J.; Perdew, J. P. Tests of a Ladder of Density Func- tionals for Bulk Solids and Surfaces. Phys. Rev. B 2004, 69, 075102.

(4) Paier, J.; Marsman, M.; Hummer, K.; Kresse, G.; Gerber, I. C.; Angy´ an,´ J. G. Screened hybrid density functionals applied to solids. J. Chem. Phys. 2006, 124, 154709.

17 (5) Haas, P.; Tran, F.; Blaha, P. Calculation of the Lattice Constant of Solids with Semilocal Functionals. Phys. Rev. B 2009, 79, 085104.

(6) Nørskov, J. K.; Abild-Pedersen, F.; Studt, F.; Bligaard, T. Density Functional Theory in Surface Chemistry and Catalysis. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 937–943.

(7) Burke, K. Perspective on Density Functional Theory. J. Chem. Phys. 2012, 136, 150901.

(8) Freysoldt, C.; Grabowski, B.; Hickel, T.; Neugebauer, J.; Kresse, G.; Janotti, A.; Van de Walle, C. G. First-principles calculations for point defects in solids. Rev. Mod. Phys. 2014, 86, 253–305.

(9) Runge, E.; Gross, E. K. U. Density-Functional Theory for Time-Dependent Systems. Phys. Rev. Lett. 1984, 52, 997–1000.

(10) Gavrilenko, V. I.; Bechstedt, F. Optical Functions of Semiconductors beyond Density- Functional Theory and Random-Phase Approximation. Phys. Rev. B 1997, 55, 4343–4352.

(11) Tokatly, I. V.; Pankratov, O. Many-Body Diagrammatic Expansion in a Kohn-Sham Basis: Implications for Time-Dependent Density Functional Theory of Excited States. Phys. Rev. Lett. 2001, 86, 2078–2081.

(12) Reining, L.; Olevano, V.; Rubio, A.; Onida, G. Excitonic Effects in Solids Described by Time-Dependent Density-Functional Theory. Phys. Rev. Lett. 2002, 88, 066404.

(13) Onida, G.; Reining, L.; Rubio, A. Electronic Excitations: Density-Functional versus Many- Body Green’s-Function Approaches. Rev. Mod. Phys. 2002, 74, 601–659.

(14) Sottile, F.; Bruneval, F.; Marinopoulos, A. G.; Dash, L. K.; Botti, S.; Olevano, V.; Vast, N.; Rubio, A.; Reining, L. TDDFT from Molecules to Solids: The Role of Long-Range Inter- actions. Int. J. Quantum Chem. 2005, 102, 684–701.

(15) Botti, S.; Schindlmayr, A.; Sole, R. D.; Reining, L. Time-dependent density-functional the- ory for extended systems. Reports on Progress in Physics 2007, 70, 357–407.

18 (16) Paier, J.; Marsman, M.; Kresse, G. Dielectric properties and excitons for extended systems from hybrid functionals. Phys. Rev. B 2008, 78, 121201.

(17) Izmaylov, A. F.; Scuseria, G. E. Why Are Time-Dependent Density Functional Theory Exci- tations in Solids Equal to Band Structure Energy Gaps for Semilocal Functionals, and How Does Nonlocal Hartree–Fock-Type Exchange Introduce Excitonic Effects? J. Chem. Phys. 2008, 129, 034101.

(18) Sharma, S.; Dewhurst, J. K.; Sanna, A.; Gross, E. K. U. Bootstrap Approximation for the Exchange-Correlation Kernel of Time-Dependent Density-Functional Theory. Phys. Rev. Lett. 2011, 107, 186401.

(19) Rigamonti, S.; Botti, S.; Veniard, V.; Draxl, C.; Reining, L.; Sottile, F. Estimating Excitonic Effects in the Absorption Spectra of Solids: Problems and Insight from a Guided Iteration Scheme. Phys. Rev. Lett. 2015, 114, 146402.

(20) Ullrich, C. A.; Yang, Z.-h. In Density-Functional Methods for Excited States; Ferre,´ N., Filatov, M., Huix-Rotllant, M., Eds.; Springer International Publishing: Cham, 2016; pp 185–217.

(21) Hedin, L. New Method for Calculating the One-Particle Green’s Function with Application to the Electron-Gas Problem. Phys. Rev. 1965, 139, A796–A823.

(22) Strinati, G.; Mattausch, H. J.; Hanke, W. Dynamical Correlation Effects on the Quasiparticle Bloch States of a Covalent Crystal. Phys. Rev. Lett. 1980, 45, 290–294.

(23) Strinati, G.; Mattausch, H. J.; Hanke, W. Dynamical Aspects of Correlation Corrections in a Covalent Crystal. Phys. Rev. B 1982, 25, 2867–2888.

(24) Hybertsen, M. S.; Louie, S. G. First-Principles Theory of Quasiparticles: Calculation of Band Gaps in Semiconductors and Insulators. Phys. Rev. Lett. 1985, 55, 1418–1421.

19 (25) Hybertsen, M. S.; Louie, S. G. Electron Correlation in Semiconductors and Insulators: Band Gaps and Quasiparticle Energies. Phys. Rev. B 1986, 34, 5390–5413.

(26) Pulci, O.; Marsili, M.; Luppi, E.; Hogan, C.; Garbuio, V.; Sottile, F.; Magri, R.; Sole, R. D. Electronic Excitations in Solids: Density Functional and Green’s Function Theory. Phys. Status Solidi B 2005, 242, 2737–2750.

(27) Albrecht, S.; Reining, L.; Sole, R. D.; Onida, G. Excitonic Effects in the Optical Properties. Phys. Status Solidi A 1998, 170, 189–197.

(28) Hanke, W.; Sham, L. J. Many-Particle Effects in the Optical Spectrum of a Semiconductor. Phys. Rev. B 1980, 21, 4656–4673.

(29) Sham, L. J.; Rice, T. M. Many-Particle Derivation of the Effective-Mass Equation for the Wannier Exciton. Phys. Rev. 1966, 144, 708–714.

(30) Strinati, G. Effects of Dynamical Screening on Resonances at Inner-Shell Thresholds in Semiconductors. Phys. Rev. B 1984, 29, 5718–5726.

(31) Rohlfing, M.; Louie, S. G. Electron-Hole Excitations and Optical Spectra from First Princi- ples. Phys. Rev. B 2000, 62, 4927–4944.

(32) Georges, A.; Kotliar, G. Hubbard Model in Infinite Dimensions. Phys. Rev. B 1992, 45, 6479–6483.

(33) Georges, A.; Kotliar, G.; Krauth, W.; Rozenberg, M. J. Dynamical Mean-Field Theory of Strongly Correlated Fermion Systems and the Limit of Infinite Dimensions. Rev. Mod. Phys. 1996, 68, 13–125.

(34) Georges, A. Strongly Correlated Electron Materials: Dynamical Mean-Field Theory and Electronic Structure. AIP Conference Proceedings 2004, 715, 3–74.

20 (35) Kotliar, G.; Savrasov, S. Y.; Haule, K.; Oudovenko, V. S.; Parcollet, O.; Marianetti, C. A. Electronic structure calculations with dynamical mean-field theory. Rev. Mod. Phys. 2006, 78, 865–951.

(36) Held, K. Electronic Structure Calculations Using Dynamical Mean Field Theory. Adv. Phys. 2007, 56, 829–926.

(37) Biermann, S.; Aryasetiawan, F.; Georges, A. First-Principles Approach to the Electronic Structure of Strongly Correlated Systems: Combining the $GW$ Approximation and Dy- namical Mean-Field Theory. Phys. Rev. Lett. 2003, 90, 086402.

(38) Biermann, S. In First Principles Approaches to Spectroscopic Properties of Complex Mate- rials; Di Valentin, C., Botti, S., Cococcioni, M., Eds.; Topics in Current Chemistry; Springer Berlin Heidelberg: Berlin, Heidelberg, 2014; pp 303–345.

(39) Zhao, L.; Neuscamman, E. Variational Excitations in Real Solids: Optical Gaps and Insights into Many-Body Perturbation Theory. Phys. Rev. Lett. 2019, 123, 036402.

(40) Williamson, A. J.; Hood, R. Q.; Needs, R. J.; Rajagopal, G. Diffusion quantum Monte Carlo calculations of the excited states of silicon. Phys. Rev. B 1998, 57, 12140–12144.

(41) Hunt, R. J.; Szyniszewski, M.; Prayogo, G. I.; Maezono, R.; Drummond, N. D. Quan- tum Monte Carlo calculations of energy gaps from first principles. Phys. Rev. B 2018, 98, 075122.

(42) Yang, Y.; Gorelov, V.; Pierleoni, C.; Ceperley, D. M.; Holzmann, M. Electronic band gaps from quantum Monte Carlo methods. Phys. Rev. B 2020, 101, 085115.

(43) Ma, F.; Zhang, S.; Krakauer, H. Excited state calculations in solids by auxiliary-field quan- tum Monte Carlo. New J. Phys. 2013, 15, 093017.

(44) Sun, J.-Q.; Bartlett, R. J. Second-order Many-body Perturbation-theory Calculations in Ex- tended Systems. J. Chem. Phys. 1996, 104, 8553–8565.

21 (45) Ayala, P. Y.; Kudin, K. N.; Scuseria, G. E. Atomic Orbital Laplace-Transformed Second- Order Møller–Plesset Theory for Periodic Systems. J. Chem. Phys. 2001, 115, 9698–9707.

(46) Pisani, C.; Busso, M.; Capecchi, G.; Casassa, S.; Dovesi, R.; Maschio, L.; Zicovich- Wilson, C.; Schutz,¨ M. Local-MP2 Electron Correlation Method for Nonconducting Crys- tals. J. Chem. Phys. 2005, 122, 094113.

(47) Marsman, M.; Gruneis, A.; Paier, J.; Kresse, G. Second-Order Møller–Plesset Perturbation Theory Applied to Extended Systems. I. Within the Projector-Augmented-Wave Formalism Using a Plane Wave Basis Set. J. Chem. Phys. 2009, 130, 184103.

(48) Hirata, S.; Grabowski, I.; Tobita, M.; Bartlett, R. J. Highly Accurate Treatment of Electron Correlation in Polymers: Coupled-Cluster and Many-Body Perturbation Theories. Chem. Phys. Letters 2001, 345, 475–480.

(49) Hirata, S.; Podeszwa, R.; Tobita, M.; Bartlett, R. J. Coupled-Cluster Singles and Doubles for Extended Systems. J. Chem. Phys. 2004, 120, 2581–2592.

(50) Katagiri, H. Equation-of-Motion Coupled-Cluster Study on Exciton States of Polyethylene with Periodic Boundary Condition. J. Chem. Phys. 2005, 122, 224901.

(51) Gruneis,¨ A.; Booth, G. H.; Marsman, M.; Spencer, J.; Alavi, A.; Kresse, G. Natural Orbitals for Wave Function Based Correlated Calculations Using a Plane Wave Basis Set. J. Chem. Theory Comput. 2011, 7, 2780–2785.

(52) McClain, J.; Sun, Q.; Chan, G. K.-L.; Berkelbach, T. C. Gaussian-Based Coupled-Cluster Theory for the Ground State and Band Structure of Solids. J Chem Theory Comput 2017,

(53) Shepherd, J. J.; Gruneis,¨ A.; Booth, G. H.; Kresse, G.; Alavi, A. Convergence of Many- Body Wave-Function Expansions Using a Plane-Wave Basis: From Homogeneous Electron Gas to Solid State Systems. Phys. Rev. B 2012, 86, 035111.

22 (54) Booth, G. H.; Gruneis,¨ A.; Kresse, G.; Alavi, A. Towards an Exact Description of Electronic Wavefunctions in Real Solids. Nature 2013, 493, 365.

(55) Ruggeri, M.; R´ıos, P. L.; Alavi, A. Correlation energies of the high-density spin-polarized electron gas to meV accuracy. Phys. Rev. B 2018, 98, 161105.

(56) Malone, F. D.; Blunt, N. S.; Brown, E. W.; Lee, D. K. K.; Spencer, J. S.; Foulkes, W. M. C.; Shepherd, J. J. Accurate Exchange-Correlation Energies for the Warm Dense Electron Gas. Phys. Rev. Lett. 2016, 117, 115701.

(57) White, A. F.; Chan, G. A Time-Dependent Formulation of Coupled Cluster Theory for Many-Fermion Systems at Finite Temperature. ArXiv180709961 Cond-Mat Physicsphysics 2018,

(58) Hummel, F. Finite Temperature Coupled Cluster Theories for Extended Systems. J. Chem. Theory Comput. 2018, 14, 6505–6514, PMID: 30354112.

(59) McClain, J.; Lischner, J.; Watson, T.; Matthews, D. A.; Ronca, E.; Louie, S. G.; Berkel- bach, T. C.; Chan, G. K.-L. Spectral Functions of the Uniform Electron Gas via Coupled- Cluster Theory and Comparison to the $GW$ and Related Approximations. Phys. Rev. B 2016, 93, 235139.

(60) Pulkin, A.; Chan, G. K.-L. First Principles Coupled Cluster Theory of the Electronic Spec- trum of the Transition Metal Dichalcogenides. arXiv:1909.10886 2019,

(61) Gao, Y.; Sun, Q.; Yu, J. M.; Motta, M.; McClain, J.; White, A. F.; Minnich, A. J.; Chan, G. K.-L. Electronic Structure of Bulk Manganese Oxide and Nickel Oxide from Coupled Clus- ter Theory. arXiv:1910.02191 2019,

(62) Hirata, S.; Head-Gordon, M.; Bartlett, R. J. Configuration Interaction Singles, Time- Dependent Hartree–Fock, and Time-Dependent Density Functional Theory for the Elec- tronic Excited States of Extended Systems. J. Chem. Phys. 1999, 111, 14.

23 (63) Lorenz, M.; Usvyat, D.; Schutz,¨ M. Local Ab Initio Methods for Calculating Optical Band Gaps in Periodic Systems. I. Periodic Density Fitted Local Configuration Interaction Singles Method for Polymers. J. Chem. Phys. 2011, 134, 094101.

(64) Lorenz, M.; Maschio, L.; Schutz,¨ M.; Usvyat, D. Local Ab Initio Methods for Calculat- ing Optical Bandgaps in Periodic Systems. II. Periodic Density Fitted Local Configuration Interaction Singles Method for Solids. J. Chem. Phys. 2012, 137, 204119.

(65) Lewis, A. M.; Berkelbach, T. C. Ab Initio Lifetime and Concomitant Double-Excitation Character of Plasmons at Metallic Densities. Phys. Rev. Lett. 2019, 122.

(66) Lewis, A. M.; Berkelbach, T. C. Ab Initio Linear and Pump-Probe Spectroscopy of Exci- tons in Molecular Crystals. The Journal of Physical Chemistry Letters 0, 0, null, PMID: 32109074.

(67) Coester, F.; Kummel,¨ H. Short-Range Correlations in Nuclear Wave Functions. 1960, 17, 477–485.

(68) Cˇ ´ızek,ˇ J. On the Correlation Problem in Atomic and Molecular Systems. Calculation of Wavefunction Components in Ursell-Type Expansion Using Quantum-Field Theoretical Methods. J. Chem. Phys. 1966, 45, 4256–4266.

(69)K ummel,¨ H. Origins of the Coupled Cluster Method. Theoret. Chim. Acta 1991, 80, 81–89.

(70) Bartlett, R. J. Many-Body Perturbation Theory and Coupled Cluster Theory for Electron Correlation in Molecules. Ann. Rev. Phys. Chem. 1981, 32, 359–401.

(71) Bartlett, R.; Musiał, M. Coupled-Cluster Theory in Quantum Chemistry. Rev. Mod. Phys. 2007, 79, 291–352.

(72) Emrich, K. An Extension of the Coupled Cluster Formalism to Excited States: (II). Approx- imations and Tests. Nuclear Physics A 1981, 351, 397–438.

24 (73) Koch, H.; Jørgensen, P. Coupled Cluster Response Functions. J. Chem. Phys. 1990, 106, 8059.

(74) Stanton, J. F.; Bartlett, R. J. The Equation of Motion Coupled-cluster Method. A System- atic Biorthogonal Approach to Molecular Excitation Energies, Transition Probabilities, and Excited State Properties. J. Chem. Phys. 1993, 98, 7029–7039.

(75) Kobayashi, R.; Koch, H.; Jørgensen, P. Calculation of Frequency-Dependent Polarizabilities Using Coupled-Cluster Response Theory. Chem Phys Lett 1994, 219, 30–35.

(76) Krylov, A. I. Equation-of-Motion Coupled-Cluster Methods for Open-Shell and Electroni- cally Excited Species: The Hitchhiker’s Guide to Fock Space. Annu. Rev. Phys. Chem. 2008, 59, 433–462.

(77) Monkhorst, H. J.; Pack, J. D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192.

(78) Goedecker, S.; Teter, M.; Hutter, J. Separable Dual-Space Gaussian Pseudopotentials. Phys. Rev. B 1996, 54, 1703–1710.

(79) Hartwigsen, C.; Goedecker, S.; Hutter, J. Relativistic Separable Dual-Space Gaussian Pseu- dopotentials from H to Rn. Phys. Rev. B 1998, 58, 3641–3662.

(80) VandeVondele, J.; Krack, M.; Mohamed, F.; Parrinello, M.; Chassaing, T.; Hutter, J. Quick- step: Fast and Accurate Density Functional Calculations Using a Mixed Gaussian and Plane Waves Approach. Computer Physics Communications 2005, 167, 103–128.

(81) Sun, Q.; Berkelbach, T. C.; McClain, J. D.; Chan, G. K.-L. Gaussian and Plane-Wave Mixed Density Fitting for Periodic Systems. J. Chem. Phys. 2017, 147, 164119.

(82) Oba, F.; Togo, A.; Tanaka, I.; Paier, J.; Kresse, G. Defect Energetics in ZnO: A Hybrid Hartree-Fock Density Functional Study. Phys. Rev. B 2008, 77, 245202.

25 (83) Sundararaman, R.; Arias, T. A. Regularization of the Coulomb Singularity in Exact Ex- change by Wigner-Seitz Truncated Interactions: Towards Chemical Accuracy in Nontrivial Systems. Phys. Rev. B 2013, 87, 165122.

(84) Davidson, E. R. The Iterative Calculation of a Few of the Lowest Eigenvalues and Cor- responding Eigenvectors of Large Real-Symmetric Matrices. J. Comput. Phys. 1975, 17, 87–94.

(85) Hirao, K.; Nakatsuji, H. A Generalization of the Davidson’s Method to Large Nonsymmetric Eigenvalue Problems. J. Comput. Phys. 1982, 45, 246–254.

(86) Sun, Q.; Berkelbach, T. C.; Blunt, N. S.; Booth, G. H.; Guo, S.; Li, Z.; Liu, J.; Mc- Clain, J. D.; Sayfutyarova, E. R.; Sharma, S.; Wouters, S.; Chan, G. K.-L. PySCF: The Python-Based Simulations of Chemistry Framework. WIREs Comput. Mol. Sci. 2018, 8, e1340.

(87) Gruneis,¨ A.; Shepherd, J. J.; Alavi, A.; Tew, D. P.; Booth, G. H. Explicitly Correlated Plane Waves: Accelerating Convergence in Periodic Wavefunction Expansions. J. Chem. Phys. 2013, 139, 084112.

(88) Shepherd, J. J.; Gruneis,¨ A. Many-Body Quantum Chemistry for the Electron Gas: Conver- gent Perturbative Theories. Phys. Rev. Lett. 2013, 110, 226401.

(89) Carrier, P.; Rohra, S.; Gorling,¨ A. General Treatment of the Singularities in Hartree-Fock and Exact-Exchange Kohn-Sham Methods for Solids. Phys. Rev. B 2007, 75, 205126.

(90) Spencer, J.; Alavi, A. Efficient Calculation of the Exact Exchange Energy in Periodic Sys- tems Using a Truncated Coulomb Potential. Phys. Rev. B 2008, 77, 193110.

(91) Chiesa, S.; Ceperley, D. M.; Martin, R. M.; Holzmann, M. Finite-Size Error in Many-Body Simulations with Long-Range Interactions. Phys. Rev. Lett. 2006, 97, 076404.

26 (92) Drummond, N. D.; Needs, R. J.; Sorouri, A.; Foulkes, W. M. C. Finite-size errors in contin- uum quantum Monte Carlo calculations. Phys. Rev. B 2008, 78, 125106.

(93) Gruneis,¨ A.; Marsman, M.; Kresse, G. Second-Order Møller–Plesset Perturbation Theory Applied to Extended Systems. II. Structural and Energetic Properties. J. Chem. Phys. 2010, 133, 074107.

(94) Booth, G. H.; Tsatsoulis, T.; Chan, G. K.-L.; Gruneis,¨ A. From Plane Waves to Local Gaus- sians for the Simulation of Correlated Periodic Systems. J. Chem. Phys. 2016, 145, 084111.

(95) Liao, K.; Gruneis,¨ A. Communication: Finite Size Correction in Periodic Coupled Cluster Theory Calculations of Solids. J. Chem. Phys. 2016, 145, 141102.

(96) Gruber, T.; Liao, K.; Tsatsoulis, T.; Hummel, F.; Gruneis,¨ A. Applying the Coupled-Cluster Ansatz to Solids and Surfaces in the Thermodynamic Limit. Phys. Rev. X 2018, 8, 021043.

(97) Mihm, T. N.; McIsaac, A. R.; Shepherd, J. J. An optimized twist angle to find the twist- averaged correlation energy applied to the uniform electron gas. J. Chem. Phys. 2019, 150, 191101.

(98) Phillip, H. R.; Taft, E. A. Kramers-Kronig Analysis of Reflectance Data for Diamond. Phys. Rev. 1964, 136, A1445–A1448.

(99) Lautenschlager, P.; Garriga, M.; Vina, L.; Cardona, M. Temperature Dependence of the Dielectric Function and Interband Critical Points in Silicon. Phys. Rev. B 1987, 36, 4821– 4830.

(100) Logothetidis, S.; Petalas, J. Dielectric Function and Reflectivity of 3C–Silicon Carbide and the Component Perpendicular to the c Axis of 6H–Silicon Carbide in the Energy Region 1.5–9.5 eV. J. Appl. Phys. 1996, 80, 1768–1772.

(101) Rao, K. K.; Moravec, T. J.; Rife, J. C.; Dexter, R. N. Vacuum Ultraviolet Reflectivities of LiF, NaF, and KF. Phys. Rev. B 1975, 12, 5937–5950.

27 (102) Abbamonte, P.; Graber, T.; Reed, J. P.; Smadici, S.; Yeh, C.-L.; Shukla, A.; Rueff, J.-P.; Ku, W. Dynamical Reconstruction of the Exciton in LiF with Inelastic X-Ray Scattering. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 12159–12163.

(103) Eby, J. E.; Teegarden, K. J.; Dutton, D. B. Ultraviolet Absorption of Alkali Halides. Phys. Rev. 1959, 116, 1099–1105.

(104) Bortz, M. L.; French, R. H.; Jones, D. J.; Kasowski, R. V.; Ohuchi, F. S. Temperature de- pendence of the electronic structure of oxides: MgO, MgAl2O4and Al2O3. Physica Scripta 1990, 41, 537–541.

(105) Roessler, D. M.; Walker, W. C. Optical Constants of Magnesium Oxide and Lithium Fluo- ride in the Far Ultraviolet*. J. Opt. Soc. Am. 1967, 57, 835.

(106) Tararan, A.; di Sabatino, S.; Gatti, M.; Taniguchi, T.; Watanabe, K.; Reining, L.; Tizei, L. H. G.; Kociak, M.; Zobelli, A. Optical Gap and Optically Active Intragap Defects in Cubic BN. Phys. Rev. B 2018, 98, 094106.

(107) Hwang, S.; Kim, T.; Jung, Y.; Barange, N.; Park, H.; Kim, J.; Kang, Y.; Kim, Y.; Shin, S.; Song, J.; Liang, C.-T.; Chang, Y.-C. Dielectric function and critical points of AlP determined by spectroscopic ellipsometry. Journal of Alloys and Compounds 2014, 587, 361 – 364.

(108) Wing, D.; Haber, J. B.; Noff, R.; Barker, B.; Egger, D. A.; Ramasubramaniam, A.; Louie, S. G.; Neaton, J. B.; Kronik, L. Comparing Time-Dependent Density Functional The- ory with Many-Body Perturbation Theory for Semiconductors: Screened Range-Separated Hybrids and the $GW$ plus Bethe-Salpeter Approach. Phys. Rev. Materials 2019, 3, 064603.

(109) Satta, G.; Cappellini, G.; Olevano, V.; Reining, L. Many-body effects in the electronic spec- tra of cubic boron nitride. Phys. Rev. B 2004, 70, 195212.

28 (110) Dittmer, A.; Izsak,´ R.; Neese, F.; Maganas, D. Accurate Band Gap Predictions of Semi- conductors in the Framework of the Similarity Transformed Equation of Motion Coupled Cluster Theory. Inorg. Chem. 2019, 58, 9303–9315, PMID: 31240911.

(111) Sous, J.; Goel, P.; Nooijen, M. Similarity transformed equation of motion coupled cluster theory revisited: a benchmark study of valence excited states. Mol. Phys. 2014, 112, 616– 638.

(112) Marini, A. Ab Initio Finite-Temperature Excitons. Phys. Rev. Lett. 2008, 101, 106405.

(113) Giustino, F.; Louie, S. G.; Cohen, M. L. Electron-Phonon Renormalization of the Direct Band Gap of Diamond. Phys. Rev. Lett. 2010, 105, 265501.

(114) Antonius, G.; Ponce,´ S.; Boulanger, P.; Cotˆ e,´ M.; Gonze, X. Many-Body Effects on the Zero-Point Renormalization of the Band Structure. Phys. Rev. Lett. 2014, 112, 215501.

(115) Ponce,´ S.; Antonius, G.; Gillet, Y.; Boulanger, P.; Laflamme Janssen, J.; Marini, A.; Cotˆ e,´ M.; Gonze, X. Temperature dependence of electronic eigenenergies in the adiabatic harmonic approximation. Phys. Rev. B 2014, 90, 214304.

(116) Zacharias, M.; Patrick, C. E.; Giustino, F. Stochastic Approach to Phonon-Assisted Optical Absorption. Phys. Rev. Lett. 2015, 115, 177401.

(117) Kittel, C.; McEuen, P.; McEuen, P. Introduction to Solid State Physics; Wiley New York, 1996; Vol. 8.

(118) Bardeen, J.; Shockley, W. Deformation Potentials and Mobilities in Non-Polar Crystals. Phys. Rev. 1950, 80, 72–80.

(119) Herring, C.; Vogt, E. Transport and Deformation-Potential Theory for Many-Valley Semi- conductors with Anisotropic Scattering. Phys. Rev. 1956, 101, 944–961.

29 (120) Goings, J. J.; Caricato, M.; Frisch, M. J.; Li, X. Assessment of Low-Scaling Approximations to the Equation of Motion Coupled-Cluster Singles and Doubles Equations. J. Chem. Phys. 2014, 141, 164116.

30 Graphical TOC Entry

31