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

Quenching of low-energy optical absorption in bilayer C3N polytypes

Matteo Zanfrognini [email protected] Dipartimento di Scienze Fisiche, Informatiche e Matematiche, Universita`\grave{a} di Modena e Reggio Emilia, I-41125 Modena, Italy Centro S3, CNR-Istituto Nanoscienze, I-41125 Modena, Italy    Miki Bonacci Dipartimento di Scienze Fisiche, Informatiche e Matematiche, Universita`\grave{a} di Modena e Reggio Emilia, I-41125 Modena, Italy Centro S3, CNR-Istituto Nanoscienze, I-41125 Modena, Italy    Fulvio Paleari Centro S3, CNR-Istituto Nanoscienze, I-41125 Modena, Italy    Elisa Molinari Dipartimento di Scienze Fisiche, Informatiche e Matematiche, Universita`\grave{a} di Modena e Reggio Emilia, I-41125 Modena, Italy Centro S3, CNR-Istituto Nanoscienze, I-41125 Modena, Italy    Alice Ruini Dipartimento di Scienze Fisiche, Informatiche e Matematiche, Universita`\grave{a} di Modena e Reggio Emilia, I-41125 Modena, Italy Centro S3, CNR-Istituto Nanoscienze, I-41125 Modena, Italy    Andrea Ferretti Centro S3, CNR-Istituto Nanoscienze, I-41125 Modena, Italy    Marilia J. Caldas Instituto de Física, Universidade de São Paulo, Cidade Universitária, 05508-900 São Paulo, Brazil    Daniele Varsano [email protected] Centro S3, CNR-Istituto Nanoscienze, I-41125 Modena, Italy
Abstract

In this work we provide a first principles description of the electronic and optical properties of bilayers C3N, with different stacking motifs AB, AB and AA. Starting from quasi-particle electronic band-structures, we solve the Bethe Salpeter Equation (BSE) to access the excitonic properties of these bilayers. For all stacking sequences, we see strong optical absorption at energies lower than but close to that of the monolayer. Most relevant, we predict a strong quenching of the low-energy optical absorption, with negligible oscillator strength of low-lying bound excitons. This is a unique phenomenology that does not arise in the monolayer case, nor in other common homo-bilayers. We explain these findings in terms of the small interband dipole matrix elements associated to the valence-conduction transitions involved in these excitons, and discuss them in view of the different stacking motifs.

I Introduction

Many recent experimental and theoretical works have focused on a detailed understanding of the electronic and optical properties of two-dimensional (2D) materials, due to their potential use in the design of innovative optoelectronic devices which could combine atomic-size dimensions with improved performance Mak et al. (2010); Zhang (2015); Chernikov et al. (2014); Cudazzo et al. (2016); Lin et al. (2016); Rousseau et al. (2021); Zhang et al. (2017); Galvani et al. (2016); Tran et al. (2014); Thygesen (2017); Palummo et al. (2015). Great attention has been reserved to the study of Van der Waals structures, where atomically-thin 2D materials are vertically stacked and held together by weak and long range dispersive interactions, strongly affecting both the optical Andersen et al. (2015); Paleari et al. (2018) and the electronic properties of the isolated single layers He et al. (2014). A promising material for such novel applications is graphene-like 2D polyaniline (also known as monolayer C3N), which has been recently synthesized through different approaches Mahmood et al. (2016); Yang et al. (2017): its electronic and optical properties have been extensively studied from a theoretical point of view Wu et al. (2018); Bonacci et al. (2022); Tang et al. (2022); Zanfrognini et al. (2023), revealing a quasi-particle band structure with indirect band-gap and intense optical absorption for photons in a narrow spectral range around 2 eV.

Vertical stacking of two layers of C3N is a possible approach to tune its electronic and optical properties. Few theoretical works have analysed the stability of bilayers C3N (BL-C3N) as a function of the possible stacking patterns Wu et al. (2018); Wenya et al. (2021). Following the notation of Ref. [Wenya et al., 2021], all these calculations have obtained negative formation energies for AB and AB (displaced-like), as well as AA (sandwich-like stacking) arrangements. Furthermore, BL-C3N with AA and AB stackings have also been experimentally synthesized, as described in Ref. [Wenya et al., 2021], where a detailed investigation of the electronic properties using scanning tunnelling spectroscopy (STS) has been provided. The results indicate a strong change of the electronic-transport band gap passing from monolayer to bilayer, together with relevant modifications of the electronic properties as a function of the stacking sequence.

Motivated by these experimental advances, in this work we discuss the optical properties of BL-C3N via first principles methods, properly including excitonic effects which are known to play a fundamental role in the description of optical absorption in 2D-materials Latini et al. (2015); Andersen et al. (2015); Hüser et al. (2013). As a first result, we find for all systems a strong optical absorption in an energy region around 1.7 eV which is sightly lower than the absorption peak of the isolated monolayer Bonacci et al. (2022), despite the consistent reduction of the electronic gap. This feature, combined with the energy band structure associated to these absorption peaks, is promising for photovoltaics. Moreover, our results indicate that, for all the considered stackings, the optical response does not present other bright excitonic states at lower energies. Such behaviour is not observed in other common semiconducting bilayer homo-structures, e.g. BL-hBNPaleari et al. (2018), BL-MoS2Mak et al. (2010); Palummo et al. (2015) or BL-phosphoreneTran et al. (2014). Finally, the origin of this low energy absorption quenching is explained and rationalized, focusing on the properties of the single particle states involved in the formation of the lowest-energy excitons.

The article is organized as follows. In Section II, we summarize the computational methods used within this work, while in Section III we present quasi-particle band structures for the AB and AB displaced-like stacking patterns. In Section IV we discuss the optical properties of these bilayers and in Section V we provide a rationale for the negligible oscillator strength observed for the low-lying excitons. Finally, in Section VI, we discuss the optical properties of bilayer C3N with AA sandwich-like stacking, predicting a strong optical absorption quenching starting from the symmetry properties of this polytype and confirming such interpretation with an approximate solution of the Bethe Salpeter equation.

II Computational methods

Ground state structural and electronic properties have been investigated using density functional theory (DFT), as implemented in the plane-wave-basis-set package Quantum ESPRESSO Giannozzi et al. (2009, 2017). In these calculations, we have used norm-conserving ONCV pseudopotentials Hamann (2013), within the GGA-PBE approximation Perdew et al. (1996) for the exchange-correlation potential. Van der Waals interactions between layers have been taken into account by adding the dispersion correction proposed by Grimme Grimme (2006) to the exchange-correlation energy (PBE-D2). Equilibrium structural properties have been obtained by relaxing both the in-plane unit cell and the atomic positions up to when the components of the forces acting on each atom were smaller than 5\cdot10-4 Ry/Bohr. In all ground state calculations we used a 12×\times12×\times1 Monkhorst-Pack Monkhorst and Pack (1976) k-grid to sample the Brillouin Zone (BZ) and a kinetic energy cutoff of 90 Ry for the plane wave basis set used to represent single particle wavefunctions.

In the case of AB and AB stacking motifs, Kohn-Sham wavefunctions and eigenvalues, computed from the equilibrium ground state charge density, have been used to evaluate quasi-particle (QP) corrections to DFT energies within the GW approximation Hedin (1965); Onida et al. (2002) for the electron self energy. QP corrections ϵn𝐤KS\epsilon_{n\mathbf{k}}^{\text{KS}} have been computed using the single shot G0W0 approach, evaluating the expectation value of the operator ΣvKS{\Sigma}-{v}_{\text{KS}} on the Kohn-Sham states |ψn𝐤|\psi_{n\mathbf{k}}\rangle, being Σ{\Sigma} the electron self-energy operator and vKS{v}_{KS} the exchange-correlation potential. When solding the Dyson equation we linearized the frequency dependence of the self-energy, Σ(En𝐤QP)Σ(ϵn𝐤KS)+Σϵ|ϵn𝐤KS(En𝐤QPϵn𝐤KS){\Sigma}(E^{\text{QP}}_{n\mathbf{k}})\approx{\Sigma}(\epsilon^{\text{KS}}_{n\mathbf{k}})+\frac{\partial{\Sigma}}{\partial\epsilon}\big{|}_{\epsilon^{\text{KS}}_{n\mathbf{k}}}(E^{\text{QP}}_{n\mathbf{k}}-\epsilon^{\text{KS}}_{n\mathbf{k}}), as proposed in Ref. [Hybertsen and Louie, 1986].

The electron-electron screened interaction WW has been computed using the Random Phase Approximation (RPA), as implemented in the Yambo code Marini et al. (2009); Sangalli et al. (2019). Converged QP gaps within a 10 meV threshold required the inclusion of 700 bands and a 𝐆\mathbf{G}-vector cutoff of 16 Ry in the construction of the screening matrix at the RPA level. The frequency dependence of WW has been described using the Godby-Needs plasmon-pole model Godby and Needs (1989), and 1000 bands have been included in the sum-over-states appearing in the correlation part of the self energy. To reduce spurious interactions among different cells in the stacking direction, we have used a supercell length along zz of 23.5 Å, together with a 2D cutoff Ismail-Beigi (2006); Rozzi et al. (2006) for the Coulomb potential. Finally, to speed-up the convergence of QP gaps w.r.t. the k-point mesh, we have adopted the approach recently proposed by Guandalini et al. Guandalini et al. (2022). In this work we have verified that, with this method, a 18×\times18×\times1 Monkhorst-Pack k-grid already provides converged gaps within the chosen threshold of 10 meV.

Starting from QP corrected electronic energies, we have obtained excitonic properties by solving the Bethe-Salpeter equation (BSE) Strinati (1988); Rohlfing and Louie (2000) in the resonant (Tamm-Dancoff) approximation, i.e. via diagonalization of the excitonic Hamiltonian

Hexc(vc𝐤,vc𝐤)=(Ec𝐤QPEv𝐤QP)δccδvv+Kd(vc𝐤,vc𝐤)+Kx(vc𝐤,vc𝐤)\begin{split}H_{exc}\big{(}vc\mathbf{k},v^{\prime}c^{\prime}\mathbf{k^{\prime}}\big{)}&=\big{(}E_{c\mathbf{k}}^{\text{QP}}-E_{v\mathbf{k}}^{\text{QP}}\big{)}\delta_{cc^{\prime}}\delta_{vv^{\prime}}\\ &+K^{d}\big{(}vc\mathbf{k},v^{\prime}c^{\prime}\mathbf{k}^{\prime}\big{)}+K^{x}\big{(}vc\mathbf{k},v^{\prime}c^{\prime}\mathbf{k}^{\prime}\big{)}\end{split} (1)

where KdK^{d} (KxK^{x}) is the direct (exchange) part of the BSE kernel, (v,v)(v,v^{\prime}) and (c,c)(c,c^{\prime}) are the valence and conduction bands included in the BSE and E(v,c)𝐤QPE_{(v,c)\mathbf{k}}^{QP} are the QP corrected electron energies. Converged exciton energies have been obtained including the two highest-occupied valence and the four lowest unoccupied conduction bands in the construction of Hexc(vc𝐤,vc𝐤)H_{exc}\big{(}vc\mathbf{k},v^{\prime}c^{\prime}\mathbf{k^{\prime}}\big{)}, and using a 48×\times48×\times1 Monkhorst-Pack k-grid to sample the BZ. We point out that, in the solution of the BSE, QP corrections have been approximated via a scissor-stretching operator (see Supplemental Material sup for a detailed description about the fitting procedure), while the electron-electron screened interaction has been computed at the RPA level, in the static approximation, using the same converged parameters adopted for the calculation of QP corrections.

Finally, by diagonalization of Hexc(vc𝐤,vc𝐤)H_{exc}\big{(}vc\mathbf{k},v^{\prime}c^{\prime}\mathbf{k^{\prime}}\big{)}

vc𝐤Hexc(vc𝐤,vc𝐤)Aλ(vc𝐤)=EλAλ(vc𝐤)\sum_{v^{\prime}c^{\prime}\mathbf{k}^{\prime}}H_{exc}\big{(}vc\mathbf{k},v^{\prime}c^{\prime}\mathbf{k^{\prime}}\big{)}A_{\lambda}\big{(}v^{\prime}c^{\prime}\mathbf{k}^{\prime}\big{)}=E_{\lambda}A_{\lambda}\big{(}vc\mathbf{k}\big{)} (2)

we obtained the energies EλE_{\lambda} and the envelope functions Aλ(vc𝐤)A_{\lambda}\big{(}vc\mathbf{k}\big{)} for each exciton λ\lambda. Optical absorption is finally computed as the imaginary part of the macroscopic dielectric function,

ϵM(E)=18πVλDλEEλ+iη\epsilon_{M}(E)=1-\frac{8\pi}{V}\sum_{\lambda}\frac{D_{\lambda}}{E-E_{\lambda}+i\eta} (3)

In the above expression, VV is the unit cell volume and DλD_{\lambda} the oscillator strength of exciton λ\lambda defined as

Dλ=|ϵ^vc𝐤dvc𝐤Aλ(vc𝐤)|2,D_{\lambda}=\bigg{|}\hat{\epsilon}\cdot\sum_{vc\mathbf{k}}d_{vc\mathbf{k}}A_{\lambda}(vc\mathbf{k})\bigg{|}^{2}, (4)

where ϵ^\hat{\epsilon} is the in-plane polarization direction of the incoming light, and dvc𝐤=φv𝐤|𝐫|φc𝐤d_{vc\mathbf{k}}=\langle\varphi_{v\mathbf{k}}|\mathbf{r}|\varphi_{c\mathbf{k}}\rangle the single-particle interband dipole matrix element.

III Structural and electronic properties of bilayer C3N with AB and AB stackings

Refer to caption
Figure 1: Crystal structures for bilayer C3N with AB (upper panel) and AB (lower panel) stackings. Yellow (light blue) spheres indicate Carbon (Nitrogen) atoms, while small (large) radius spheres denote atoms located on the upper (lower) layer. The red dot indicates the in-plane position of the inversion symmetry center, while dashed red lines represent mirror symmetry planes parallel to the stacking direction. Finally, the green dot in the AB bilayer denotes the in-plane position of the two three-fold rotation axes parallel to the stacking direction, while the dashed green line in the AB bilayer corresponds to an in-plane C2 rotation axis.

In Figure 1 we present the crystal structures of BL-C3N with AB (upper panel) and AB (lower panel) stackings: yellow (light blue) spheres denote Carbon (Nitrogen) atoms and small (large) atoms are located on the upper (lower) layer, denoted from now on as L1\mathrm{L}_{1} (L2\mathrm{L}_{2}). For both stacking motifs, we have obtained an in-plane lattice parameter of 4.849 Å, slightly smaller than that of the isolated monolayer (4.857 Å); the interlayer distance (evaluated as the separation along zz between Carbon atoms with the same in-plane coordinates) has been found equal to 3.22 Å for the AB and 3.21 Å  for the AB stacking. These values are in agreement with those obtained with PBE-D2 calculations in Ref. [Wenya et al., 2021], while slightly smaller than the interlayer distances computed with VdW-functionals in Ref. [Wu et al., 2018].

We now briefly discuss the crystal symmetries of the two stackings. The point group of AB-C3N is D3d, and also includes non-symmorphic operations. This stacking possesses a spatial inversion center (red dot in the upper panel of Fig. 1) together with a three-fold C3 rotation axis along zz-direction, whose in-plane position is denoted by a green dot in Fig. 1. Furthermore, this stacking motif is invariant under mirror reflections σ\sigma w.r.t. planes parallel to the stacking direction and represented by dashed red lines in Fig. 1: These planes are respectively denoted as σ^ΓM\hat{\sigma}_{\Gamma\mathrm{M}}, σ^ΓM\hat{\sigma}_{\Gamma\mathrm{M^{\prime}}}, and σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M^{\prime\prime}}} as they are aligned to these high symmetry directions in the BZ. AB-C3N has lower symmetry: Its point group is C2h, that includes the inversion symmetry, a two-fold in-plane rotation axis lying between the two C3N planes (represented by the green dashed line in the lower panel of Fig. 1 and denoted as C^2ΓK\hat{\mathrm{C}}_{2}^{\Gamma\mathrm{K}}), and a mirror symmetry plane σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M^{\prime\prime}}} parallel to the ΓM′′\Gamma\mathrm{M^{\prime\prime}} direction in the BZ.

Refer to caption
Figure 2: Electronic band structure computed at the DFT-PBE (green-dashed lines) and with G0W0 approximation (continuous black lines), for AB-C3N (a) and AB-C3N (b). The insets represent the hexagonal Brillouin zone, together with the high symmetry points defining the paths where bands are computed. The parity of the topmost valence and the lowest conduction bands along ΓM′′\Gamma\mathrm{M}^{\prime\prime} direction w.r.t. mirror symmetry σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M^{\prime\prime}}} are indicated: the notation nipn_{i}^{p} indicates that band nin_{i} has parity pp, being p=+()p=+(-) for even (odd) states and n={v,c}n=\{v,c\}. In both images, the top valence band energy is shifted to 0 eV.

In Figure 2 we present the electronic band structure of AB-C3N (panel a) and AB-C3N (panel b), computed first within DFT-PBE (dashed green lines) and then including QP corrections at the G0W0 level (solid black lines). We note that the two stackings are characterized by an indirect band gap, both at the DFT and GW level. In the case of AB stacking, the electronic band dispersion along Γ\GammaM and ΓM′′\Gamma\mathrm{M^{\prime\prime}} coincide, as a consequence of the three-fold rotational symmetry. Therefore, the highest-energy valence band has six equivalent maxima, while the bottom of the conduction band is located at the Γ\Gamma point. We also note that the presence of doubly degenerate bands at Γ\Gamma is consistent with irreducible representations of dimension 2 in the D3d\mathrm{D}_{3\mathrm{d}} point group. The calculated indirect gap at PBE level is 0.108 eV, which is increased to 0.72 eV once QP corrections are taken into account. Finally, the direct band gap is found along the ΓM\Gamma\mathrm{M} direction, at k located approximately at half the ΓM\Gamma\rightarrow\mathrm{M} path: the G0W0 gap is 1.85 eV, 0.73 eV larger than the value obtained at the PBE level (1.12 eV).

In the AB stacking, the maximum of the valence band is found at the M′′\mathrm{M}^{\prime\prime} point, while the lowest unoccupied conduction state is at Γ\Gamma, as in the case of the AB stacking motif. Our calculations give an indirect gap of 0.136 eV at the DFT-PBE level, which is increased to 0.73 eV with the inclusion of QP corrections. The obtained indirect gap for AB-C3N is slightly smaller than the one measured experimentally with STS on SiO2-Si substrates in Ref. [Wenya et al., 2021] (0.85±\pm0.03 eV), but in better agreement than other GW calculations Wu et al. (2018) where the Hybertsen-Louie plasmon-pole model was used Hybertsen and Louie (1986). Because of the lack of three-fold rotational symmetry around the zz axis, the directions ΓM\Gamma\mathrm{M} and ΓM′′\Gamma\mathrm{M}^{\prime\prime} are no longer equivalent. As a result, the minimum direct gap is found approximately at half the ΓM′′\Gamma\rightarrow\mathrm{M}^{\prime\prime} path, with a value around 1.79 eV (1.09 eV within DFT) while the indirect gap between the conduction at Γ\Gamma and the top-valence at M\mathrm{M} is slightly larger than the one between Γ\Gamma and M′′\mathrm{M}^{\prime\prime}. We obtain a ΓcMv\Gamma_{c}-\mathrm{M}_{v} gap of 0.87 eV (0.249 eV) at the G0W0 (DFT-PBE) level. Electronic gaps computed within the G0W0 approximation for both stacking motifs at high symmetry points are summarized in Table 1.

Comparing the band dispersions for AB and AB stackings, we notice that, for 𝐤\mathbf{k} along the ΓM\Gamma\mathrm{M} direction, the lowest pair of conduction bands are almost degenerate (splitting of about 1 meV) in the AB stacking, while well separated in the AB case (splitting larger than 0.2 eV). In the Supplemental Material sup , we provide a qualitative explanation of this peculiar lack of splitting among the two lowest conduction bands in AB-C3N, analysing the quasi-symmetries of the sublattice where conduction states are localized once 𝐤\mathbf{k} is taken along ΓM\Gamma\mathrm{M} direction. A similar behaviour is also observed for the two topmost valence states, where the splitting is, however, not negligible also in the AB stacking (splitting of about 50 meV). Overall, the obtained gaps for the AB and AB stacking are similar (in agreement with the hybrid-DFT results of Ref. [Wenya et al., 2021]), while the electronic dispersions differ because of the different symmetry properties of the two stackings. We point out that there are clear differences with the monolayer case, as we observe a strong reduction in both the indirect and direct electronic gaps (see Table. 1 for a comparison).

We complete our analysis by recalling that both stacking configurations are invariant under mirror reflection σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M^{\prime\prime}}}, therefore electronic states for k along this direction can be classified in terms of their parity w.r.t. such symmetry. We have numerically found that, in both stackings, the two highest occupied valence bands are even w.r.t. σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M^{\prime\prime}}}, (here denoted as vi+v_{i}^{+} in Fig. 2, with i=1,2i=1,2). On the other hand, if k has modulus in the range [|ΓM|3,|ΓM|][\frac{|\Gamma\mathrm{M}|}{3},|\Gamma\mathrm{M}|] the two lowest unoccupied conduction bands are σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M^{\prime\prime}}}-odd (cic_{i}^{-}). Instead, for k close to Γ\Gamma, the second and the third conduction bands are even w.r.t σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M^{\prime\prime}}} and exhibit a stronger dispersion with k, if compared with odd conduction states. We point out that a similar analysis can be carried out also for electronic states along this direction in monolayer C3N. In that case, the highest valence (lowest conduction) is even (odd) w.r.t. the mirror symmetry along ΓM′′\Gamma\mathrm{M}^{\prime\prime} direction, while the second conduction is even. In the case of AB stacking, the same parity analysis can be presented for the bands along the direction ΓM\Gamma\mathrm{M}, as the crystal is invariant w.r.t. σ^ΓM\hat{\sigma}_{\Gamma\mathrm{M}} mirror symmetry. Such symmetry classification will be exploited to understand bilayer optical properties in the following.

Gaps ML Ref. [Bonacci et al., 2022] AB AB
ΓcΓv\Gamma_{c}-\Gamma_{v} 2.96 2.36 2.39
McMv\mathrm{M}_{c}-\mathrm{M}_{v} 2.67 2.02 2.40
Mc′′Mv′′\mathrm{M}^{\prime\prime}_{c}-\mathrm{M}^{\prime\prime}_{v} 2.02 2.03
ΓcMv\Gamma_{c}-\mathrm{M}_{v} 1.42 0.72 0.87
ΓcMv′′\Gamma_{c}-\mathrm{M}^{\prime\prime}_{v} 0.72 0.73
Min. direct gap 2.62 1.85 1.79
Table 1: Direct and indirect gaps (in eV) at high symmetry points for AB and AB BL-C3N, obtained at the G0W0 level, compared with monolayer (ML) data (G0W0 on top of DFT-PBE) from Ref. [Bonacci et al., 2022]. The last row indicates the minimum direct electronic gap evaluated along ΓM′′\Gamma\mathrm{M}^{\prime\prime} direction.

IV Optical absorption in bilayer C3N with AB and AB stackings

Refer to caption
Figure 3: Optical absorption spectra for AB-C3N (a) and AB-C3N (b). Solid red (dashed green) lines correspond to spectra computed with light polarization along the ΓM′′{\Gamma\mathrm{M}}^{\prime\prime} (ΓK{\Gamma\mathrm{K}}) direction, while dotted lines are the independent particle spectra evaluated for polarization along ΓM′′{\Gamma\mathrm{M}}^{\prime\prime}. The polarization directions are indicated on top of the crystal structures for clarity. The insets highlight low energy quasi-dark peaks labelled as P1 and P2, while the vertical black dashed lines indicate the position of the direct QP band gap obtained within G0W0. The band structure at the G0W0 level of AB-C3N is shown in (c), with red arrows indicating the transitions mainly responsible for the C absorption peak in (a). Similarly, the single particle bands of AB-C3N are displayed in (d): green (red) arrows emphasize the transitions mostly involved in C1 (C2) peaks labelled in b.

We now turn to the discussion of the optical properties of bilayer C3N. In Figure 3 we show the absorption spectra computed for AB (panel a) and AB bilayer C3N (panel b), at the independent particle level (dotted black lines), and with the inclusion of electron-hole interaction by solving the BSE as detailed in Section II. Green dashed (solid red) lines have been obtained assuming light polarized along the ΓK{\Gamma\mathrm{K}} (ΓM′′{\Gamma\mathrm{M}^{{}^{\prime\prime}}}) direction. For clarity, polarization versors ϵ^\hat{\epsilon} are shown together with the crystal structures in the insets.

The AB spectrum is dominated by an intense peak (denoted as C) at energy E1.70E\approx 1.70 eV, whose spectral position and intensity are not dependent on the polarization direction. Such peak is due to a set of almost degenerate eigenstates of Hexc, characterized by relevant contributions from single particle transitions between the valence band v1+v^{+}_{1} and the conduction state c1c^{-}_{1} along ΓM\Gamma\mathrm{M} equivalent directions. Among these, transitions with the highest weights are denoted by arrows on the band structure shown in Fig. 3(c). We point out that transitions between v2+v^{+}_{2} and c2c^{-}_{2} for 𝐤\mathbf{k} along the same direction also contribute to this absorption peak, though with a smaller weight than v1+c1v^{+}_{1}\rightarrow c^{-}_{1}, and are omitted for clarity in Fig. 3(c). The situation is slightly different in the case of AB C3N. Indeed, also for this stacking motif the absorption spectrum is dominated by a single intense peak, but its position in energy and its strength depend on the in-plane light polarization direction. In Fig. 3(b), we have labelled as C1 the main peak at 1.71 eV obtained for light polarization along ΓK{\Gamma\mathrm{K}} and as C2 the absorption maximum at 1.73 eV found for ΓM′′{\Gamma\mathrm{M}^{{}^{\prime\prime}}}-polarized light. The C1 peak is mainly due to v1+c1v^{+}_{1}\rightarrow c^{-}_{1} for 𝐤\mathbf{k} along ΓM′′\Gamma\mathrm{M}^{{}^{\prime\prime}} direction, as shown schematically by green arrows in Fig. 3(d), again with a smaller contribution coming from v2+v^{+}_{2} to c2c^{-}_{2} transitions for the same 𝐤\mathbf{k} points (not shown in Fig. 3d). On the other hand, the C2 absorption peak comes from transitions between the two highest occupied valence states and the two lowest (quasi-degenerate) conduction bands along ΓM{\Gamma\mathrm{M}} and ΓM{\Gamma\mathrm{M}}^{\prime} directions (see red arrows in Fig. 3d).

First, we note that the energy window found for strong optical absorption in both stackings is still in a good range for solar energy conversion, not much lower than that found for monolayer C3N of 1.82 eV in Ref. [Bonacci et al., 2022]. Furthermore, the associated transitions are still in region favourable to electron-hole splitting, as in the monolayer. This can be particularly interesting, since one can expect the bilayer to have a more stable structure, compared to a monolayer, when deposited on a suitable electrode.

Second, and central to this work, the striking feature of bilayer C3N optical spectra is the apparent absence of intense absorption peaks due to strongly bound excitonic states formed by single particle transitions close to the electronic direct gap. As shown in the insets of Fig. 3(a,b), within the energy range between 1.25 and 1.5 eV, two absorption structures (labelled in both cases as P1 and P2) are present, but they exhibit optical strengths which are almost two orders of magnitude smaller than the most intense peaks.

Refer to caption
Figure 4: Contributions of single particle transitions between the highest occupied valence and the lowest unoccupied conduction in the BZ to excitons P1 and P2 in AB-C3N (a and c) and in AB-C3N (b and d), as defined by Eq. (5).

In AB-C3N, both P1 and P2 peaks are due to a pair of degenerate excitons, respectively at energies EP1E_{\mathrm{P}_{1}} = 1.35 eV and EP2E_{\mathrm{P}_{2}} = 1.47 eV, with oscillator strengths not dependent on the polarization direction. For completeness, we point out that diagonalization of the excitonic Hamiltonian also provides other excitonic resonances (the lowest with energy of 1.34 eV) which are characterized by null oscillator strength within numerical accuracy. Such excitons, dark by strict symmetry reasons, will not be considered further in the following. P1 and P2 excitons are almost totally composed by electron-hole transitions between the highest occupied valence v2+v^{+}_{2} and the lowest unoccupied conduction c1c^{-}_{1} states, with wave-vectors 𝐤\mathbf{k} along ΓM\Gamma\mathrm{M} and equivalent directions in the BZ. This is better clarified in Fig. 4(a,c), where we show plots of the quantity

Avc(𝐤)=λ|Aλ(vc𝐤)|2A^{vc}(\mathbf{k})=\sum_{\lambda}\big{|}A_{\lambda}(vc\mathbf{k})\big{|}^{2} (5)

for the P1 and P2 excitons, respectively. In Eq. (5), the index vv (cc) is fixed to the last valence (first conduction) band and the sum over λ\lambda is performed over the pair of degenerate states responsible for the P1 and P2 peaks. We notice that single particle transitions forming the excitons P1 are mainly localized in the central region of ΓM\Gamma\mathrm{M} and equivalent directions, with Avc(𝐤)A^{vc}(\mathbf{k}) having non-negligible values for |𝐤||\mathbf{k}| mainly in the interval [13,23]|ΓM|[\frac{1}{3},\frac{2}{3}]|\Gamma\mathrm{M}|. On the other hand, excitons P2 (Fig. 4c) are still localized along ΓM\Gamma\mathrm{M} directions, but the corresponding function Avc(𝐤)A^{vc}(\mathbf{k}) has intense contributions from transitions slightly closer to the M\mathrm{M} point and exhibits a node for 𝐤\mathbf{k} points along this direction.

In the case of the AB stacking, P1 and P2 peaks are related each to a single-nondegenerate exciton at energies EP1E_{\mathrm{P}_{1}} = 1.30 eV and EP2E_{\mathrm{P}_{2}} = 1.42 eV, respectively. Differently from the AB case, the oscillator strength of these excitations is polarization dependent as shown in the inset of Fig. 3(b), where we notice that both resonances are dark for light polarization along the ΓM′′{\Gamma\mathrm{M}}^{\prime\prime} direction while they exhibit a small but non-zero optical activity for incoming light polarized along ΓK{\Gamma\mathrm{K}}. As in the AB stacking, such excitons are mainly formed by transitions between the highest occupied valence and the lowest conduction band. In Fig. 4(b,d) we show the functions Avc(𝐤)A^{vc}(\mathbf{k}) computed for exciton P1 and P2, respectively. We see that, in both cases, the transitions involved in these excitations are strongly localized along the single ΓM′′\Gamma\mathrm{M}^{\prime\prime} direction of the BZ (see e.g. Fig. 2), with the P1 exciton having the main contributions coming from the middle of ΓM′′\Gamma\mathrm{M}^{\prime\prime}, where the minimum direct electronic band gap is located, and the P2 resonance characterized by a node along this BZ-line.

The observed polarization dependence in the optical absorption of AB-C3N can be rationalized via symmetry arguments, similarly to the analysis proposed in Ref. [Paleari et al., 2018] for bilayer hBN. The point group of AB-C3N is C2h, so the in-plane exciton dipole operator projected along ΓM′′{\Gamma\mathrm{M}}^{\prime\prime} (𝐃ΓM′′{\mathbf{D}}_{\Gamma\mathrm{M}^{\prime\prime}}) transforms as the Bu irreducible representation of C2h, while 𝐃ΓK{\mathbf{D}}_{\Gamma\mathrm{K}} as Au. In fact, 𝐃ΓM′′{\mathbf{D}}_{\Gamma\mathrm{M}^{\prime\prime}} is invariant under the mirror symmetry σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M}^{\prime\prime}} (as oriented along ΓM′′\Gamma\mathrm{M}^{\prime\prime}), but it changes sign under inversion and C^2ΓK\hat{C}^{\Gamma\mathrm{K}}_{2} rotation (see Fig. 1). Differently, 𝐃ΓK{\mathbf{D}}_{\Gamma\mathrm{K}} is even w.r.t. C^2ΓK\hat{C}^{\Gamma\mathrm{K}}_{2} and odd under σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M}^{\prime\prime}} and inversion, therefore behaving as the AuA_{u}(C2h) irreducible representation. Therefore, the eigenstates |λ|\lambda\rangle of HexcH_{exc} such that 0|𝐃|λ0\langle 0|\mathbf{D}|\lambda\rangle\neq 0 (|0|0\rangle being the excitonic vacuum transforming as the fully symmetric representation Ag) transform as Au or Bu if 𝐃{\mathbf{D}} is projected along ΓK{\Gamma\mathrm{K}} or ΓM′′{\Gamma\mathrm{M}}^{\prime\prime}, respectively. Considering for example Au states, these will have null optical activity by symmetry, once incoming light is polarized along ΓM′′{\Gamma\mathrm{M}}^{\prime\prime}. Therefore, we can explain the presence of absorption peaks in AB-C3N which turn on and off according to the chosen polarization direction. Furthermore, following the presented analysis, we can assign P1 and P2 excitons to the Au representation, exactly as the eigenstates of Hexc responsible for the C1 peak, while C2 is expected to be due to excitations transforming as Bu.

The situation is different for AB stacked C3N. For simplicity, we discuss the brightness of excitonic eigenstates using the subgroup C3v of D3d, formed by the mirror planes σ\sigma along ΓM\Gamma\mathrm{M} directions together with the two 3-fold rotations depicted in the upper panel of Fig. 1. The in-plane exciton dipole operator 𝐃\mathbf{D} transforms as the two-dimensional irreducible representation E of C3v. As a consequence, all bright excitons behave as E of C3v, so that they are expected to be double-degenerate and characterized by an isotropic oscillator strength, in agreement with the numerical results obtained from ab initio calculations.

We now turn our attention to the quasi-dark nature of the low-lying bound excitons P1 and P2 in both the considered stackings. We point out that such small optical activity cannot be related to an interlayer nature of these excitons, i.e. it is not due to a negligible spatial overlap between electron and hole wavefunctions. This point is better clarified by looking, for example, at the real space wavefunctions of exciton P1, as shown in Fig. 5, with AB (AB) results reported in the left (right) panel. For each stacking, the upper (lower) wavefunction has been obtained assuming the hole (represented by the red dot) fixed on layer L1 (L2) and located in the plane close to a nitrogen atom.

Refer to caption
Figure 5: Wavefunctions for exciton P1 computed in real space for AB (left) and AB (right) stackings. In each panel, the upper (lower) wavefunction has been computed assuming the hole localized on a Nitrogen atom on layer L1 (L2). The hole position along the stacking direction is indicated by the red circle.

The wavefunctions clearly indicate the intralayer nature of such excitons: In fact, looking at the excitonic wavefunction isosurfaces, we notice that the electron has a high probability to be found on the same layer on which the hole is localized.

Therefore, the negligible dipole strength of these low-lying excitons is a consequence of the small interband dipole associated to the electron-hole single particle transitions involved in these excitations. This can be understood from the independent particle (IP) absorption spectrum shown in Fig. 3(a,b) as dotted lines, where, in both cases, the optical signal is negligible for energies close to the direct QP gap, with the IP absorption onset located at higher photon energies. In the following section we discuss in greater detail the single particle states involved in these low-lying excitons and we propose a possible rationale for the observed IP absorption quenching.

V Rationale for quenching of low energy absorption in AB and AB BL-C3N

In this Section we develop a model for the electronic bands in proximity of the direct gap to rationalize the small oscillator strength of the single particle transitions close to the direct electronic gap. As the following analysis is valid for both stackings, here we focus on the case of AB-C3N (results for AB motif are presented in the Supplemental Material sup ). We will restrict our analysis to 𝐤\mathbf{k} along ΓM′′\Gamma\mathrm{M}^{\prime\prime}, in the region where valence to conduction transitions giving the highest contribution to the low-energy quasi-dark excitons are located.

As already discussed in the literature Wenya et al. (2021); Wu et al. (2018), the lowest lying conduction bands and the highest valence states have a π\pi character. therefore, we can analyse them using a tight-binding (TB) Hamiltonian, obtained considering one 2pz orbital for each atom. In the following we will denote as τα\mathbf{\tau}_{\alpha} the position of both the atom α\alpha and the 2pz orbital localized on it. In practice, we construct a TB Hamiltonian as

Hα,β2L(𝐤)=𝐑ei𝐤𝐑t(α0,β𝐑),H^{\mathrm{2L}}_{\alpha,\beta}(\mathbf{k})=\sum_{\mathbf{R}}e^{i\mathbf{k}\cdot\mathbf{R}}\,t(\alpha 0,\beta\mathbf{R}), (6)

where 𝐑\mathbf{R} is a lattice vector and t(α0,β𝐑)=α0|H|β𝐑t(\alpha 0,\beta\mathbf{R})=\langle\alpha 0|{H}|\beta\mathbf{R}\rangle are the hopping matrix elements between two 2pz orbitals, localized at sites τα\mathbf{\tau}_{\alpha} and τβ+𝐑\mathbf{\tau}_{\beta}+\mathbf{R}, respectively. These matrix elements have been computed fully ab initio by Wannierization Souza et al. (2001); Marzari et al. (2012) of DFT bands using the Wannier90 Mostofi et al. (2008, 2014); Pizzi et al. (2020) code. Details about the procedure are provided in the Supplemental Material sup .

As there are 16 atoms in the unit cell, at a general 𝐤\mathbf{k} the TB Hamiltonian can be written as a 16×\times16 Hermitian matrix, in a block-like form as

𝐇2L(𝐤)=[𝐇L1(𝐤)𝐇IL(𝐤)𝐇IL(𝐤)𝐇L2(𝐤)],\mathbf{H}^{\mathrm{2L}}(\mathbf{k})=\begin{bmatrix}\mathbf{H}^{\mathrm{L}_{1}}(\mathbf{k})&\mathbf{H}^{\mathrm{IL}}(\mathbf{k})\\ \mathbf{H}^{\mathrm{IL}}(\mathbf{k})^{\dagger}&\mathbf{H}^{\mathrm{L}_{2}}(\mathbf{k})\end{bmatrix}, (7)

where 𝐇L1\mathbf{H}^{\mathrm{L}_{1}}, 𝐇L2\mathbf{H}^{\mathrm{L}_{2}}, and 𝐇IL\mathbf{H}^{\mathrm{IL}} are 8×\times8 matrices corresponding to the different layers and their coupling. Indeed, to obtain this separation, we have grouped together the orbitals localized on L1 and on L2, associating to each orbital α1\alpha_{1} localized at τα1\mathbf{\tau}_{\alpha_{1}} on layer L1 an orbital α2\alpha_{2} at τα2\mathbf{\tau}_{\alpha_{2}} on layer L2, such that τα2=I^τα2\mathbf{\tau}_{\alpha_{2}}=\hat{\mathrm{I}}\mathbf{\tau}_{\alpha_{2}}, I^\hat{\mathrm{I}} being the inversion symmetry operator. Then 𝐇L1\mathbf{H}^{\mathrm{L}_{1}} (𝐇L2\mathbf{H}^{\mathrm{L}_{2}}) is the block of 𝐇2L\mathbf{H}^{\mathrm{2L}} which contains the intralayer hopping within layer L1 (L2), while 𝐇IL\mathbf{H}^{\mathrm{IL}} depends on the hopping integrals between orbitals on different layers. We can now write 𝐇2L(𝐤)=𝐡IN(𝐤)+𝐡IL(𝐤)\mathbf{H}^{\mathrm{2L}}(\mathbf{k})=\mathbf{h}_{\mathrm{IN}}(\mathbf{k})+\mathbf{h}_{\mathrm{IL}}(\mathbf{k}), where 𝐡IN(𝐤)\mathbf{h}_{\mathrm{IN}}(\mathbf{k}) is block-diagonal while 𝐡IL(𝐤)\mathbf{h}_{\mathrm{IL}}(\mathbf{k}) is purely off-diagonal. We note that the subscripts ”IN” and ”IL” stand for intralayer and interlayer, respectively. Using spatial (I^\hat{\mathrm{I}}) and time (T^\hat{\mathrm{T}}) inversion symmetries, in Appendix A we show that, for each 𝐤\mathbf{k}, 𝐡IN(𝐤)\mathbf{h}_{\mathrm{IN}}(\mathbf{k}) has a spectrum of eigenvalues εn𝐤0\varepsilon_{n\mathbf{k}}^{0} which are degenerate in pairs. Furthermore, we also show that the eigenspace associated to each eigenvalue εn𝐤0\varepsilon_{n\mathbf{k}}^{0} is spanned by two Bloch states |ϕn𝐤L1|\phi^{\mathrm{L}_{1}}_{n\mathbf{k}}\rangle and |ϕn𝐤L2|\phi^{\mathrm{L}_{2}}_{n\mathbf{k}}\rangle such that |ϕn𝐤L2=I^T^|ϕn𝐤L1|\phi^{\mathrm{L}_{2}}_{n\mathbf{k}}\rangle=-\hat{\mathrm{I}}\cdot\hat{\mathrm{T}}|\phi^{\mathrm{L}_{1}}_{n\mathbf{k}}\rangle and with |ϕn𝐤Li|\phi^{\mathrm{L}_{i}}_{n\mathbf{k}}\rangle localized on layer Li\mathrm{L}_{i}.

Before proceeding, we clarify the physical meaning of the splitting of 𝐇2L\mathbf{H}^{\mathrm{2L}} into 𝐡IN\mathbf{h}_{\mathrm{IN}} and 𝐡IL\mathbf{h}_{\mathrm{IL}}. In particular, 𝐡IN\mathbf{h}_{\mathrm{IN}} can be thought as an intralayer Hamiltonian, describing the two layers as not interacting with each other. Note however that the presence of the other layer is implicitly considered as the intralayer matrix elements in 𝐇L1\mathbf{H}^{\mathrm{L}_{1}} (𝐇L2\mathbf{H}^{\mathrm{L}_{2}}) are affected by the presence of opposite layer L2 (L1). On the other hand, 𝐡IL\mathbf{h}_{\mathrm{IL}} describes the interlayer interaction and acts as a perturbation of 𝐡IN\mathbf{h}_{\mathrm{IN}} since interlayer hopping integrals are typically smaller than the intralayer ones. With this interpretation, the states |ϕn𝐤L1|\phi^{\mathrm{L}_{1}}_{n\mathbf{k}}\rangle and |ϕn𝐤L2|\phi^{\mathrm{L}_{2}}_{n\mathbf{k}}\rangle can be thought as Bloch states with the same energy εn𝐤0\varepsilon_{n\mathbf{k}}^{0} localized on one of the two monolayers, if the coupling 𝐡IL\mathbf{h}_{IL} is set to zero. We now define εv𝐤0\varepsilon_{v\mathbf{k}}^{0} (εc𝐤0\varepsilon_{c\mathbf{k}}^{0}) as the energy of the highest occupied valence (lowest unoccupied conduction) band on these two non-interacting layers. The effect of the interlayer coupling will be to mix these layer-localized wavefunctions, in order to give the electronic states of the bilayer.

While the discussion presented so far is general and valid for each 𝐤\mathbf{k}-point of the BZ, we now specialize to 𝐤\mathbf{k}-vectors along the ΓM′′\Gamma\mathrm{M}^{\prime\prime} direction. For these wavevectors, 𝐇2L\mathbf{H}^{\mathrm{2L}} commutes with σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M}^{\prime\prime}} and the same is valid for 𝐡IN\mathbf{h}_{\mathrm{IN}} and 𝐡IL\mathbf{h}_{\mathrm{IL}} separately. Therefore,

σ^ΓM′′|ϕv𝐤Li=|ϕv𝐤Li,σ^ΓM′′|ϕc𝐤Li=|ϕc𝐤Li,\begin{split}\hat{\sigma}_{\Gamma\mathrm{M}^{\prime\prime}}|\phi^{\mathrm{L}_{i}}_{v\mathbf{k}}\rangle&=|\phi^{\mathrm{L}_{i}}_{v\mathbf{k}}\rangle,\\ \hat{\sigma}_{\Gamma\mathrm{M}^{\prime\prime}}|\phi^{\mathrm{L}_{i}}_{c\mathbf{k}}\rangle&=-|\phi^{\mathrm{L}_{i}}_{c\mathbf{k}}\rangle,\end{split} (8)

with i=1,2i=1,2 (we have also numerically verified these relations by computing the eigenstates of 𝐡IN\mathbf{h}_{\mathrm{IN}}). We now include the effect of 𝐡IL\mathbf{h}_{\mathrm{IL}} using first order degenerate perturbation theory, separately diagonalizing the matrix representation of hIL{h}_{\mathrm{IL}} on the two subspaces {|ϕv𝐤L1,|ϕv𝐤L2}\{|\phi^{\mathrm{L}_{1}}_{v\mathbf{k}}\rangle,|\phi^{\mathrm{L}_{2}}_{v\mathbf{k}}\rangle\} and {|ϕc𝐤L1,|ϕc𝐤L2}\{|\phi^{\mathrm{L}_{1}}_{c\mathbf{k}}\rangle,|\phi^{\mathrm{L}_{2}}_{c\mathbf{k}}\rangle\}. With this procedure, we obtain the two highest-energy (σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M}^{\prime\prime}}-even) valence bands {|φv1𝐤,|φv2𝐤}\{|\varphi_{v_{1}\mathbf{k}}\rangle,|\varphi_{v_{2}\mathbf{k}}\rangle\} and the two lowest-energy (σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M}^{\prime\prime}}-odd) conduction states {|φc1𝐤,|φc2𝐤}\{|\varphi_{c_{1}\mathbf{k}}\rangle,|\varphi_{c_{2}\mathbf{k}}\rangle\} in the bilayer. Such states have been labelled as (v1+v^{+}_{1},v2+v^{+}_{2}) and (c1c^{-}_{1},c2c^{-}_{2}), respectively, in Fig. 2(b). They can be compactly written as

|φnj𝐤\displaystyle|\varphi_{n_{j}\mathbf{k}}\rangle =\displaystyle= 12[|ϕ~n𝐤L1+sj|ϕ~n𝐤L2]\displaystyle\frac{1}{\sqrt{2}}\bigg{[}|\tilde{\phi}^{\mathrm{L}_{1}}_{n\mathbf{k}}\rangle+s_{j}|\tilde{\phi}^{\mathrm{L}_{2}}_{n\mathbf{k}}\rangle\bigg{]} (9)
Enj𝐤\displaystyle E_{n_{j}\mathbf{k}} =\displaystyle= εn,𝐤0+sj|Δn𝐤|.\displaystyle\varepsilon_{n,\mathbf{k}}^{0}+s_{j}|\Delta_{n\mathbf{k}}|. (10)

In Eqs. (9)-(10), n=v,cn=v,c, j=1,2j=1,2, sj=1s_{j}=-1 (+1+1) for j=1j=1 (j=2j=2), Δn𝐤=ϕn𝐤L1|hIL|ϕn𝐤L2\Delta_{n\mathbf{k}}=\langle\phi^{\mathrm{L}_{1}}_{n\mathbf{k}}|{h}_{\mathrm{IL}}|\phi^{\mathrm{L}_{2}}_{n\mathbf{k}}\rangle, and

|ϕ~n𝐤L1=e+iγn𝐤2|ϕn𝐤L1,|ϕ~n𝐤L2=eiγn𝐤2|ϕn𝐤L2,\begin{split}|\tilde{\phi}^{\mathrm{L}_{1}}_{n\mathbf{k}}\rangle&=e^{+i\frac{\gamma_{n\mathbf{k}}}{2}}|\phi^{\mathrm{L}_{1}}_{n\mathbf{k}}\rangle,\\[5.0pt] |\tilde{\phi}^{\mathrm{L}_{2}}_{n\mathbf{k}}\rangle&=e^{-i\frac{\gamma_{n\mathbf{k}}}{2}}|\phi^{\mathrm{L}_{2}}_{n\mathbf{k}}\rangle,\end{split} (11)

where γn𝐤=Arg[Δn𝐤]\gamma_{n\mathbf{k}}=\mathrm{Arg}[\Delta_{n\mathbf{k}}] guarantees that the relative phase between the projections cα(nj𝐤)=α𝐤|φnj𝐤c_{\alpha}(n_{j}\mathbf{k})=\langle\alpha\mathbf{k}|\varphi_{n_{j}\mathbf{k}}\rangle of a Bloch state |φnj𝐤|\varphi_{n_{j}\mathbf{k}}\rangle on 2pz orbitals localized on different layers is gauge invariant, i.e. it does not change under the transformation cα1(n𝐤)eiηcα1(n𝐤)c_{\alpha_{1}}(n\mathbf{k})\rightarrow e^{i\eta}c_{\alpha_{1}}(n\mathbf{k}), η\eta being an arbitrary phase. See App. A for further details.

We point out that the zero-order expression given by Eq. (9) is a good approximation for the two lowest σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M}^{\prime\prime}}-odd conduction bands and the two highest-energy σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M}^{\prime\prime}}-even valence bands, for the considered 𝐤\mathbf{k} vectors along ΓM′′\Gamma\mathrm{M}^{\prime\prime} direction. In principle one should also consider other terms in the expression of the eigenstatesSakurai and Napolitano (2017), coming from higher orders of the perturbative series, describing the coupling between |ϕn𝐤Li|\phi^{\mathrm{L}_{i}}_{n\mathbf{k}}\rangle and the eigenstates of 𝐡IN\mathbf{h}_{\mathrm{IN}} with different eigenvalues. For 𝐤\mathbf{k} points around the middle of the ΓM′′\Gamma\mathrm{M}^{\prime\prime} direction, these terms can be neglected in first approximation, as the other eigenstates of 𝐡IN\mathbf{h}_{\mathrm{IN}} with the same σ^ΓM′′\hat{\sigma}_{\Gamma\mathrm{M}^{\prime\prime}}-parity of |ϕn𝐤Li|\phi^{\mathrm{L}_{i}}_{n\mathbf{k}}\rangle have energies far from εn𝐤0\varepsilon_{n\mathbf{k}}^{0} (w.r.t. to the interlayer coupling strength), so that the hybridization is negligible. Our numerical results (not shown) indicate that these neglected terms become more relevant for 𝐤\mathbf{k} along the same direction, but closer to the Γ\Gamma point. Therefore, looking at Eq. (9), we can understand that the lowest odd-conduction band and the highest even-valence can be seen, respectively, as antibonding and bonding combinations of the conduction and the valence states localized on the two monolayers. We recall that the states defined in Eq. (11) are still eigenstates of the intralayer Hamiltonian h^IN\hat{h}_{\mathrm{IN}}, with |ϕ~n𝐤L2=I^T^|ϕ~n𝐤L1|\tilde{\phi}^{\mathrm{L}_{2}}_{n\mathbf{k}}\rangle=-\hat{\mathrm{I}}\cdot\hat{\mathrm{T}}|\tilde{\phi}^{\mathrm{L}_{1}}_{n\mathbf{k}}\rangle.

Refer to caption
Figure 6: Interband dipole matrix elements for 𝐤\mathbf{k} along ΓM′′\Gamma\mathrm{M}^{\prime\prime} direction, in the interval [|ΓM′′|3,|ΓM′′|][\frac{|\Gamma\mathrm{M}^{\prime\prime}|}{3},|\Gamma\mathrm{M}^{\prime\prime}|]. In the upper panel of (a) we show the moduli of dLiϵd^{\epsilon}_{L_{i}} for light polarization along ΓK\Gamma\mathrm{K} for the two layers i=1,2i=1,2, while in the lower panel the phase difference among dL1ϵd^{\epsilon}_{L_{1}} and dL2ϵd^{\epsilon}_{L_{2}} is presented. In (b) the red continuous (blue dashed) line corresponds to the interband dipole between |φv2𝐤|\varphi_{v_{2}\mathbf{k}}\rangle (|φv1𝐤|\varphi_{v_{1}\mathbf{k}}\rangle) and |φc1𝐤|\varphi_{c_{1}\mathbf{k}}\rangle computed using the TB model. Dots and triangles represent the same quantities computed fully ab initio using Yambo code. Light polarization versor is assumed aligned along ΓK\Gamma\mathrm{K} direction.

Starting from Eq. (9), we can evaluate the interband matrix element between the last occupied valence |φv2𝐤|\varphi_{v_{2}\mathbf{k}}\rangle and the lowest unoccupied conduction |φc1𝐤|\varphi_{c_{1}\mathbf{k}}\rangle as

dv2c1𝐤ϵ=12\displaystyle d^{\epsilon}_{v_{2}c_{1}\mathbf{k}}=\frac{1}{2}\!\!\!\!\! [ϕ~v𝐤L1|ϵ^𝐫|ϕ~c𝐤L1ϕ~v𝐤L2|ϵ^𝐫|ϕ~c𝐤L2\displaystyle\bigg{[}\langle\tilde{\phi}^{\mathrm{L}_{1}}_{v\mathbf{k}}|\hat{\epsilon}\cdot\mathbf{r}|\tilde{\phi}^{\mathrm{L}_{1}}_{c\mathbf{k}}\rangle-\langle\tilde{\phi}^{\mathrm{L}_{2}}_{v\mathbf{k}}|\hat{\epsilon}\cdot\mathbf{r}|\tilde{\phi}^{\mathrm{L}_{2}}_{c\mathbf{k}}\rangle (12)
+ϕ~v𝐤L2|ϵ^𝐫|ϕ~c𝐤L1ϕ~v𝐤L1|ϵ^𝐫|ϕ~c𝐤L2],\displaystyle+\langle\tilde{\phi}^{\mathrm{L}_{2}}_{v\mathbf{k}}|\hat{\epsilon}\cdot\mathbf{r}|\tilde{\phi}^{\mathrm{L}_{1}}_{c\mathbf{k}}\rangle-\langle\tilde{\phi}^{\mathrm{L}_{1}}_{v\mathbf{k}}|\hat{\epsilon}\cdot\mathbf{r}|\tilde{\phi}^{\mathrm{L}_{2}}_{c\mathbf{k}}\rangle\bigg{]},

where ϵ^\hat{\epsilon} is the light polarization. To make the treatment simpler, in the following we neglect the last two terms in Eq. (12), as they involve states localized on different layers, so that their value is generally small as a result of the reduced overlap among the wavefunctions. In this way, we obtain that the interband dipole is the difference between intralayer-interband dipoles:

dLiϵ(𝐤)=ϕ~v,𝐤Li|ϵ^𝐫|ϕ~c,𝐤Li.d_{\mathrm{L}_{i}}^{\epsilon}(\mathbf{k})=\langle\tilde{\phi}^{\mathrm{L}_{i}}_{v,\mathbf{k}}|\hat{\epsilon}\cdot\mathbf{r}|\tilde{\phi}^{\mathrm{L}_{i}}_{c,\mathbf{k}}\rangle. (13)

The approach adopted to compute these matrix elements starting from the proposed tight binding model is discussed in the Supplemental Material sup .

If ϵ^\hat{\epsilon} is chosen along the ΓM′′\Gamma\mathrm{M}^{\prime\prime} direction, we immediately find that dL1ϵ(𝐤)=dL2ϵ(𝐤)=0d_{\mathrm{L}_{1}}^{\epsilon}(\mathbf{k})=d_{\mathrm{L}_{2}}^{\epsilon}(\mathbf{k})=0, because of the parity of layer-resolved states, given by Eq. (8). However, if ϵ^\hat{\epsilon} is taken along the direction ΓK\Gamma\mathrm{K}, orthogonal to ΓM′′\Gamma\mathrm{M}^{\prime\prime}, we cannot explain the quenching by straightforward symmetry arguments. To clarify this point, in Fig. 6(a) we show the modulus (upper panel) of the intralayer-interband dipoles dL1ΓK(𝐤)d_{\mathrm{L}_{1}}^{\Gamma\mathrm{K}}(\mathbf{k}) and dL2ΓK(𝐤)d_{\mathrm{L}_{2}}^{\Gamma\mathrm{K}}(\mathbf{k}), together with their relative phase (lower panel) Δφd(𝐤)=Arg[dL1ΓK(𝐤)]Arg[dL2ΓK(𝐤)]\Delta\varphi_{d}(\mathbf{k})=\mathrm{Arg}[d_{\mathrm{L}_{1}}^{\Gamma\mathrm{K}}(\mathbf{k})]-\mathrm{Arg}[d_{\mathrm{L}_{2}}^{\Gamma\mathrm{K}}(\mathbf{k})]. Our results indicate that the obtained intralayer dipole matrix elements are equal in modulus and they exhibit a relative phase close to zero in the range of 𝐤\mathbf{k} points here considered, i.e. |𝐤||\mathbf{k}| in [|ΓM′′|3,|ΓM′′|][\frac{|\Gamma\mathrm{M}^{\prime\prime}|}{3},|\Gamma\mathrm{M}^{\prime\prime}|]. As the total interband dipole is the difference among intralayer contributions – see Eq. (12) –, it will be almost zero, as a consequence of the destructive interference of the two layer-resolved components. In other words, the transition probability from |φv2𝐤|\varphi_{v_{2}\mathbf{k}}\rangle to |φc1𝐤|\varphi_{c_{1}\mathbf{k}}\rangle due to ΓK\Gamma\mathrm{K}-polarized light can be interpreted as the quantum superposition of the interband scattering processes occurring on the two layers separately, whose probability amplitudes are out-of-phase, giving an overall negligible interband oscillator strength.

In Fig. 6(b), the continuous red line indicates the interband dipole between |φv2𝐤|\varphi_{v_{2}\mathbf{k}}\rangle and |φc1𝐤|\varphi_{c_{1}\mathbf{k}}\rangle, computed using the perturbative solution of the TB model (see Supplemental Material sup for details), while the red dots are the same quantities obtained ab initio using Yambo, to check the validity of our approximate treatment. We notice that, as this cancellation is not symmetry-constrained, the interband dipole is small, but not exactly zero. Such cancellation is exact at M′′\mathrm{M}^{\prime\prime} point, because of symmetry reasons. In fact, as M′′\mathrm{M}^{\prime\prime} is invariant under spatial inversion I^\hat{\mathrm{I}}, we can assign inversion-parity labels to the states at this point. Ab initio results indicate that both |φv2𝐤|\varphi_{v_{2}\mathbf{k}}\rangle and |φc1𝐤|\varphi_{c_{1}\mathbf{k}}\rangle are odd under I^\hat{\mathrm{I}} exactly as 𝐝ΓK\mathbf{d}_{\Gamma\mathrm{K}}, therefore the overall matrix element is zero.

We point out that the interband dipole is non-zero once the transition between a pair of bonding or antibonding combinations is considered. For example, taking into account the scattering |φv1𝐤|φc1𝐤|\varphi_{v_{1}\mathbf{k}}\rangle\rightarrow|\varphi_{c_{1}\mathbf{k}}\rangle, the intralayer-interband transition amplitudes sum constructively giving an intense overall interband dipole. The intense optical activity of these transitions is responsible for the main absorption peak denoted as C1 in Fig. 3. Such prediction is confirmed by data in Fig. 6(b), where this quantity is shown as computed using the model (dashed blue line) and fully ab initio using Yambo (blue triangles). We notice that the reasonable agreement between the model and the ab initio results is an a posteriori confirmation of the validity of Eq. (9) to describe single particle states along ΓM′′\Gamma\mathrm{M}^{\prime\prime}.

VI The case of AA stacking

As discussed in the Introduction, together with AB and AB stackings, another stable bilayer C3N motif is AA. Previous DFT calculations have effectively shown that these three stackings exhibit similar energies and coexistence of these motifs is expected at room temperatureWu et al. (2018). The crystal structure of AA-C3N is shown in Fig. 7(a). This stacking has an inversion symmetry center (shown by the red dot in the figure), two mirror symmetry planes, represented by red dashed lines, and a two-fold rotation axis parallel to the stacking direction (indicated by the green dot). Interestingly, this stacking-motif is also invariant w.r.t. the non-symmorphic symmetry operation {σxy|𝝉}\{\sigma_{xy}|\bm{\tau}\}, corresponding to zzz\rightarrow-z mirror symmetry followed by fractional translation of the vector 𝝉\bm{\tau}, represented by the red arrow.

Refer to caption
Figure 7: Crystal structure for bilayer C3N with AA stacking is shown in (a). The red dot indicates the in-plane position of the inversion symmetry center, the dashed red lines represent mirror symmetry planes parallel to the stacking direction and the green dot denotes the in-plane position of the two-fold rotation axis parallel to the stacking direction. Finally, the red arrow represents the fractional translation τ\vec{\tau} discussed in the main test. The DFT-PBE electronic structure is shown in (b) along the direction ΓM\Gamma\mathrm{M}^{\prime} the highest valence and the lowest conduction bands are labelled according to their parity w.r.t. {σxy|𝝉}\{\sigma_{xy}|\bm{\tau}\} symmetry operation.

Structural optimization performed within PBE-D2 provides an in-plane lattice parameter aa = 4.849 Å, while the interlayer distance is equal to 3.22 Å, similarly to the other two stackings. For this relaxed atomic structure, the electronic structure at the DFT level is shown in Fig. 7(b). We notice that the highest valence state (found at the M\mathrm{M}^{\prime} point) has a higher energy than the lowest conduction state, occurring at the Γ\Gamma point, in agreement with the DFT results of Ref. [Wu et al., 2018] (our results give a negative ”gap” Γc\Gamma_{c} - Mv{}^{\prime}_{v} = -0.31 eV). Such metallicity is a problem related to the use of the PBE functional within the Kohn-Sham DFT scheme, since, experimentally, AA-C3N has been shown to have a finite gap of about 0.4 eV Wenya et al. (2021) and hybrid DFT calculations Wenya et al. (2021) provide a semiconducting ground state.

The use of such metallic ground state to compute both QP corrections and optical properties is problematic, as it would induce a fictitious over-screening effect once the electron-electron screened interaction is evaluated using RPA approximation, providing inaccurate values for electronic gaps and exciton binding energies. Nevertheless, in the following, we will assume that the Kohn-Sham states computed at the PBE level are anyhow a good approximation for electronic wavefunctions, despite the problems of the associated Kohn-Sham energies.

Figure 7(b) indicates that the lowest direct band gap occurs almost in the middle of the ΓM\Gamma\mathrm{M}^{\prime} direction. As a consequence, it is reasonable to expect that the low energy transitions which contribute to the lowest energy excitons also come from this portion of the BZ. Notably, 𝐤\mathbf{k} points along this direction are invariant both under the σ^ΓM\hat{\sigma}_{\Gamma\mathrm{M}^{\prime}} mirror symmetry and {σxy|𝝉}\{\sigma_{xy}|\bm{\tau}\}, therefore the electronic states can be properly labelled according to how they transform under these operations. Our DFT results indicate that the highest valence band is even under {σxy|𝝉}\{\sigma_{xy}|\bm{\tau}\}, while the lowest conduction is odd. In Fig. 7(b) these two states are indicated as v2+v^{+}_{2} and c1c^{-}_{1}, respectively. As the dipole operator 𝐝ϵ\mathbf{d}_{\epsilon} is invariant under {σxy|𝝉}\{\sigma_{xy}|\bm{\tau}\} (assuming, as usual, incoming light with polarization direction ϵ^\hat{\epsilon} orthogonal to the stacking direction zz), the matrix element φv2𝐤|ϵ^𝐫|φc1𝐤\langle\varphi_{v_{2}\mathbf{k}}|\hat{\epsilon}\cdot\mathbf{r}|\varphi_{c_{1}\mathbf{k}}\rangle is zero, independently of the direction of the polarization versor ϵ^\hat{\epsilon}. Therefore, we expect light-induced scattering between these bands to be forbidden by symmetry and consequently the low-energy excitons composed by these transitions to be optically dark.

To confirm this symmetry-based analysis, we solve BSE computing the static electron-electron interaction in the direct kernel using PBE single particle wavefunctions and applying a rigid scissor s0Ws^{W}_{0} to all the unoccupied bands, to manually remove DFT-spurious metallicity. Such scissor parameter has been chosen so that the minimum gap Γc\Gamma_{c} - Mv{}^{\prime}_{v} of the resulting band-structure was positive but smaller than the one found experimentally, to avoid under-screening effects. Furthermore, in the independent-particle part of the excitonic Hamiltonian, Eq. (1), we mimic quasi-particle corrections via a scissor operator applied to the DFT bands, manually chosen to obtain a minimum indirect band gap Γc\Gamma_{c} - Mv{}^{\prime}_{v} equal to the experimental one (0.4 eV). We underline that such scissor is therefore larger than the one introduced in the calculation of electronic screening.

The results of these calculations are shown in Fig. 8, where we have chosen s0W=0.35s^{W}_{0}=0.35 eV , which fulfills the above-mentioned requirement to induce a small, positive band gap. The obtained spectra confirm the symmetry-based discussion just outlined. In fact, in both cases, low energy excitons (whose positions are indicated by black vertical bars in the inset) are optically dark independently of the polarization direction, and no absorption structure is observed in the low-energy region between 0.75 eV and 1.25 eV, where low-lying discrete excitons are found. The dark bound excitons are due to single particle transitions between bands v2+v^{+}_{2} and c1c^{-}_{1} along the Γ\Gamma-M direction. This point is further clarified in the Supplemental Material sup , where, for completeness, we report the exciton wavefunctions of the low-lying excitons in reciprocal space. We point out that the observed optical quenching is stable with respect to small variations of the s0Ws^{W}_{0} scissor parameter. In the Supplemental Material sup this is further confirmed by solving BSE using a RPA screening obtained with s0W=s^{W}_{0}=0.45 eV.

We emphasize that a more complete analysis of the optical properties of AA stacking would require a better starting point for G0W0 and BSE calculations (i.e. a semiconducting electronic ground state). This is presently beyond the scope of this work and is left for future investigations.

Refer to caption
Figure 8: Absorption spectra for AA-C3N, computed using rigid shift parameter s0Ws^{W}_{0} equal to 0.35 eV. Polarization directions are shown on top of the crystal structure, while the vertical dashed line indicates the energy of the minimum direct gap. Finally, the black dotted line represents the independent particle absorption spectra. The inset underlines the dark nature of low-lying excitons, whose spectral positions are represented by vertical black bars.

VII Conclusions

In this work we discuss electronic and optical properties of bilayer C3N using state-of-the-art ab initio calculations. As a first point, we find that optical absorption around 1.7 eV is favoured for all stacking structures studied. Furthermore, our results suggest a particular behaviour of BL-C3N, i.e. the absence of low energy absorption peaks due to strongly bound excitons composed by electron-hole transitions with energies close to the minimum electronic direct gap.

These findings are explained in terms of independent particle effects as due to the negligible interband dipole between the lowest conduction and the topmost valence bands involved in the formation of the involved excitons. In the case of AB and AB stackings, we develop a model for the single particle states of interest, to demonstrate that the overall interband dipole assumes negligible values because of the destructive interference of the contributions coming from the two layers. Furthermore, we also justify the quenching of low-energy absorption in AA-C3N stacking, exploiting the symmetry properties of the crystal.

This work paves the way to future theoretical and experimental investigations on multilayer C3N. On the one hand, it could be fascinating to investigate how optical properties of C3N can be tuned varying the number of stacked monolayers or changing the twist angle between them. On the other hand, the abundant presence of dark or quasi-dark low-energy excitons in BL-C3N could have important effects on exciton lifetimes and dynamics.

ACKNOWLEDGEMENTS

M.Z. thanks Marco Gibertini and Alberto Guandalini for stimulating discussions during this work and Nicola Spallanzani and Claudia Cardoso for support in the compilation of ab initio codes on different clusters. Lattice structures have been displayed using the XCrySDen Kokalj (1999) program. This work was partially funded by MaX – MAterials design at the eXascale – a European Centre of Excellence funded by the European Union’s program HORIZON-EUROHPC-JU-2021-COE-01 (Grant No. 101093374). It was also supported by the Italian national program PRIN2017 No. 2017BZPKSZ ”Excitonic insulator in two-dimensional long-range interacting systems” (EXC-INS).

We acknowledge ICSC – Centro Nazionale di Ricerca in High Performance Computing, Big Data and Quantum Computing and ECOSISTER - Ecosystem for Sustainable Transition in Emilia-Romagna, funded under the National Recovery and Resilience Plan (NRRP) - NextGenerationEU - Mission 04, Component 2, Investments 1.4 and 1.5 [Call for tender n. 3277 dated 30/12/2021. Award Number:0001052 dated 23/06/2022].

M.J.C. acknowledges Brazilian agency CNPq and INEO (Nat. Inst. Sci. Technol. Organic Electronics - CNPq, FAPESP). Computational time on the Marconi100 machine at CINECA was provided by the Italian ISCRA program. We acknowledge EuroHPC Joint Undertaking for awarding us access to MeluXina at LuxProvide, Luxembourg. The authors gratefully acknowledge the LuxProvide teams for their expert support.

Appendix A Properties of intralayer Hamiltonian

We consider the block diagonal Hamiltonian 𝐡IN\mathbf{h}_{\mathrm{IN}} defined in Section V. For each 2pz orbital on layer L1 at position τα1\tau_{\alpha_{1}} there is an analogous state localized at τα2=I^τα1\tau_{\alpha_{2}}=\hat{I}\tau_{\alpha_{1}} on layer L2, because of the inversion symmetry of the bilayer (with both AB and AB stacking). Thus, the diagonal blocks can be related to each other, i.e.

Hα1,β1L1(𝐤)=𝐑ei𝐤𝐑t(α10,β1𝐑)=𝐑ei𝐤𝐑t(α20,β2𝐑)=Hα2,β2L2(𝐤)\begin{split}H^{\mathrm{L}_{1}}_{\alpha_{1},\beta_{1}}(\mathbf{k})&=\sum_{\mathbf{R}}e^{i\mathbf{k}\cdot\mathbf{R}}\,t(\alpha_{1}0,\beta_{1}\mathbf{R})\\ &=\sum_{\mathbf{R}}e^{-i\mathbf{k}\cdot\mathbf{R}}\,t(\alpha_{2}0,\beta_{2}\mathbf{R})\\ &=H^{\mathrm{L}_{2}}_{\alpha_{2},\beta_{2}}(\mathbf{k})^{*}\end{split} (14)

where we have used the fact that the Hamiltonian of the system H{H} is invariant under spatial inversion symmetry and the hopping integrals are real because of time reversal and I^|β1𝐑=|β2I^𝐑\hat{\mathrm{I}}|\beta_{1}\mathbf{R}\rangle=-|\beta_{2}\hat{\mathrm{I}}\mathbf{R}\rangle. Equation (14) indicates that the matrix 𝐇L1\mathbf{H}^{\mathrm{L}_{1}} is the complex conjugate of 𝐇L2\mathbf{H}^{\mathrm{L}_{2}}, therefore these matrices have the same eigenvalues. As the spectrum of 𝐡IN\mathbf{h}_{\mathrm{IN}} is the union of the spectra of 𝐇L1\mathbf{H}^{\mathrm{L}_{1}} and 𝐇L2\mathbf{H}^{\mathrm{L}_{2}}, each eigenvalue εn𝐤0\varepsilon_{n\mathbf{k}}^{0} of 𝐡IN\mathbf{h}_{\mathrm{IN}} will be twofold degenerate, for each 𝐤\mathbf{k}.

Furthermore, as 𝐇L1=𝐇L2\mathbf{H}^{\mathrm{L}_{1}}=\mathbf{H}^{\mathrm{L}_{2}*}, we can associate to each eigenvalue εn𝐤0\varepsilon_{n\mathbf{k}}^{0} the pair of eigenstates

|ϕn𝐤L1=α1cα1(n𝐤)|α1𝐤,|ϕn𝐤L2=α2cα2(n𝐤)|α2𝐤,\begin{split}|\phi^{\mathrm{L}_{1}}_{n\mathbf{k}}\rangle&=\sum_{\alpha_{1}}c_{\alpha_{1}}(n\mathbf{k})|\alpha_{1}\mathbf{k}\rangle,\\ |\phi^{\mathrm{L}_{2}}_{n\mathbf{k}}\rangle&=\sum_{\alpha_{2}}c_{\alpha_{2}}(n\mathbf{k})|\alpha_{2}\mathbf{k}\rangle,\end{split} (15)

where |α𝐤=1N𝐑ei𝐤𝐑|α𝐑|\alpha\mathbf{k}\rangle=\frac{1}{\sqrt{N}}\sum_{\mathbf{R}}e^{i\mathbf{k}\cdot\mathbf{R}}|\alpha\mathbf{R}\rangle and cα2(n𝐤)=cα1(n𝐤)c_{\alpha_{2}}(n\mathbf{k})=c_{\alpha_{1}}(n\mathbf{k})^{*}, being

β1Hα1β1L1(𝐤)cβ1(n𝐤)=εn𝐤0cα1(n𝐤).\sum_{\beta_{1}}H^{\mathrm{L}_{1}}_{\alpha_{1}\beta_{1}}(\mathbf{k})c_{\beta_{1}}(n\mathbf{k})=\varepsilon_{n\mathbf{k}}^{0}c_{\alpha_{1}}(n\mathbf{k}). (16)

We notice that |ϕn𝐤L1|\phi^{\mathrm{L}_{1}}_{n\mathbf{k}}\rangle (|ϕn𝐤L2|\phi^{\mathrm{L}_{2}}_{n\mathbf{k}}\rangle) is a Bloch function localized on layer L1 (L2) as it only involves 2pz orbitals localized on that layer. Further, defining the time inversion operator T^=K^\hat{\mathrm{T}}=\hat{\mathrm{K}}, i.e. equal to the complex conjugate operator, one can show that |ϕn𝐤L2=I^T^|ϕn𝐤L1|\phi^{\mathrm{L}_{2}}_{n\mathbf{k}}\rangle=-\hat{\mathrm{I}}\cdot\hat{\mathrm{T}}|\phi^{\mathrm{L}_{1}}_{n\mathbf{k}}\rangle. In fact,

I^T^|ϕn𝐤L1=α1cα1(n𝐤)I^T^|α1𝐤-\hat{I}\cdot\hat{T}|\phi^{L_{1}}_{n\mathbf{k}}\rangle=-\sum_{\alpha_{1}}c_{\alpha_{1}}(n\mathbf{k})^{*}\hat{I}\cdot\hat{T}|\alpha_{1}\mathbf{k}\rangle (17)

By using the reality of 2pz orbitals together with the relation: I^|α1𝐑=|α2I^𝐑\hat{I}|\alpha_{1}\mathbf{R}\rangle=-|\alpha_{2}\hat{I}\mathbf{R}\rangle we find

I^T^|α1𝐤=1N𝐑ei𝐤𝐑|α2𝐑.\hat{I}\cdot\hat{T}|\alpha_{1}\mathbf{k}\rangle=-\frac{1}{\sqrt{N}}\sum_{\mathbf{R}}e^{i\mathbf{k}\cdot\mathbf{R}}|\alpha_{2}\mathbf{R}\rangle. (18)

Combining Eq. (17) and Eq. (18) and reminding cα2(n𝐤)=cα1(n𝐤)c_{\alpha_{2}}(n\mathbf{k})=c_{\alpha_{1}}(n\mathbf{k})^{*} we finally obtain |ϕn𝐤L2=I^T^|ϕn𝐤L1|\phi^{\mathrm{L}_{2}}_{n\mathbf{k}}\rangle=-\hat{\mathrm{I}}\cdot\hat{\mathrm{T}}|\phi^{\mathrm{L}_{1}}_{n\mathbf{k}}\rangle.

References

  • Mak et al. (2010) K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, Phys. Rev. Lett. 105, 136805 (2010).
  • Zhang (2015) H. Zhang, ACS Nano 9, 9451 (2015).
  • Chernikov et al. (2014) A. Chernikov, T. C. Berkelbach, H. M. Hill, A. Rigosi, Y. Li, B. Aslan, D. R. Reichman, M. S. Hybertsen, and T. F. Heinz, Phys. Rev. Lett. 113, 076802 (2014).
  • Cudazzo et al. (2016) P. Cudazzo, L. Sponza, C. Giorgetti, L. Reining, F. Sottile, and M. Gatti, Phys. Rev. Lett. 116, 066803 (2016).
  • Lin et al. (2016) Z. Lin, A. McCreary, N. Briggs, S. Subramanian, K. Zhang, Y. Sun, X. Li, N. J. Borys, H. Yuan, S. K. Fullerton-Shirey, A. Chernikov, H. Zhao, S. McDonnell, A. M. Lindenberg, K. Xiao, B. J. LeRoy, M. Drndić, J. C. M. Hwang, J. Park, M. Chhowalla, R. E. Schaak, A. Javey, M. C. Hersam, J. Robinson, and M. Terrones, 2D Materials 3, 042001 (2016).
  • Rousseau et al. (2021) A. Rousseau, L. Ren, A. Durand, P. Valvin, B. Gil, K. Watanabe, T. Taniguchi, B. Urbaszek, X. Marie, C. Robert, and G. Cassabois, Nano Letters 21, 10133 (2021).
  • Zhang et al. (2017) Zhang, H. Guowei, C. Shenyang, S. Andrey, O. Chaoyu, L. V. Ongun, Y. Tony, and Hugen, Nature Commun. 8, 14071 (2017).
  • Galvani et al. (2016) T. Galvani, F. Paleari, H. P. C. Miranda, A. Molina-Sánchez, L. Wirtz, S. Latil, H. Amara, and F. m. c. Ducastelle, Phys. Rev. B 94, 125303 (2016).
  • Tran et al. (2014) V. Tran, R. Soklaski, Y. Liang, and L. Yang, Phys. Rev. B 89, 235319 (2014).
  • Thygesen (2017) K. S. Thygesen, 2D Materials 4, 022004 (2017).
  • Palummo et al. (2015) M. Palummo, M. Bernardi, and J. C. Grossman, Nano Letters 15, 2794 (2015).
  • Andersen et al. (2015) K. Andersen, S. Latini, and K. S. Thygesen, Nano Letters 15, 4616 (2015).
  • Paleari et al. (2018) F. Paleari, T. Galvani, H. Amara, F. Ducastelle, A. Molina-Sánchez, and L. Wirtz, 2D Materials 5, 045017 (2018).
  • He et al. (2014) J. He, K. Hummer, and C. Franchini, Phys. Rev. B 89, 075409 (2014).
  • Mahmood et al. (2016) J. Mahmood, E. K. Lee, M. Jung, D. Shin, H.-J. Choi, J.-M. Seo, S.-M. Jung, D. Kim, F. Li, M. S. Lah, N. Park, H.-J. Shin, J. H. Oh, and J.-B. Baek, Proc. Natl. Ac. Sci. 113, 7414 (2016).
  • Yang et al. (2017) S. Yang, W. Li, C. Ye, G. Wang, H. Tian, C. Zhu, P. He, G. Ding, X. Xie, Y. Liu, Y. Lifshitz, S.-T. Lee, Z. Kang, and M. Jiang, Adv. Mater. 29, 1605625 (2017).
  • Wu et al. (2018) Y. Wu, W. Xia, W. Gao, F. Jia, P. Zhang, and W. Ren, 2D Materials 6, 015018 (2018).
  • Bonacci et al. (2022) M. Bonacci, M. Zanfrognini, E. Molinari, A. Ruini, M. J. Caldas, A. Ferretti, and D. Varsano, Phys. Rev. Materials 6, 034009 (2022).
  • Tang et al. (2022) Z. Tang, G. J. Cruz, Y. Wu, W. Xia, F. Jia, W. Zhang, and P. Zhang, Phys. Rev. Applied 17, 034068 (2022).
  • Zanfrognini et al. (2023) M. Zanfrognini, N. Spallanzani, M. Bonacci, E. Molinari, A. Ruini, M. J. Caldas, A. Ferretti, and D. Varsano, Phys. Rev. B 107, 045430 (2023).
  • Wenya et al. (2021) W. Wenya, Y. Siwei, W. Gang, Z. Teng, P. Wei, C. Zenghua, Y. Yucheng, Z. Li, H. Peng, W. Lei, B. Ardeshir, Z. Quanzhen, L. Liwei, W. Yeliang, D. Guqiao, K. Zhenhui, Y. B. I., S. D. J., and Y. Qinghong, Nature Electronics 4, 486–494 (2021).
  • Latini et al. (2015) S. Latini, T. Olsen, and K. S. Thygesen, Phys. Rev. B 92, 245123 (2015).
  • Hüser et al. (2013) F. Hüser, T. Olsen, and K. S. Thygesen, Phys. Rev. B 88, 245309 (2013).
  • Giannozzi et al. (2009) P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. D. Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari, and R. M. Wentzcovitch, J. Phys.: Condens. Matter 21, 395502 (2009).
  • 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, J. Phys.: Condens. Matter 29, 465901 (2017).
  • Hamann (2013) D. R. Hamann, Phys. Rev. B 88, 085117 (2013).
  • Perdew et al. (1996) J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
  • Grimme (2006) S. Grimme, J. Comput. Chem. 27, 1787 (2006).
  • Monkhorst and Pack (1976) H. J. Monkhorst and J. D. Pack, Physical review B 13, 5188 (1976).
  • Hedin (1965) L. Hedin, Phys. Rev. 139, A796 (1965).
  • Onida et al. (2002) G. Onida, L. Reining, and A. Rubio, Rev. Mod. Phys. 74, 601 (2002).
  • Hybertsen and Louie (1986) M. S. Hybertsen and S. G. Louie, Phys. Rev. B 34, 5390 (1986).
  • Marini et al. (2009) A. Marini, C. Hogan, M. Grüning, and D. Varsano, Comput. Phys. Commun. 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, J. Phys.: Condens. Matter 31, 325902 (2019).
  • Godby and Needs (1989) R. W. Godby and R. J. Needs, Phys. Rev. Lett. 62, 1169 (1989).
  • Ismail-Beigi (2006) S. Ismail-Beigi, Phys. Rev. B 73, 233103 (2006).
  • Rozzi et al. (2006) C. A. Rozzi, D. Varsano, A. Marini, E. K. U. Gross, and A. Rubio, Phys. Rev. B 73, 205119 (2006).
  • Guandalini et al. (2022) A. Guandalini, P. D’Amico, A. Ferretti, and D. Varsano, arXiv  (2022)arXiv:2205.11946v1 [cond-mat.mtrl-sci] .
  • Strinati (1988) G. Strinati, La Rivista del Nuovo Cimento (1978-1999) 11, 1 (1988).
  • Rohlfing and Louie (2000) M. Rohlfing and S. G. Louie, Phys. Rev. B 62, 4927 (2000).
  • (41) See Supplemental Material for details on the ab initio calculations and discussion of low-energy absorption quenching in AB-C3N.
  • Souza et al. (2001) I. Souza, N. Marzari, and D. Vanderbilt, Phys. Rev. B 65, 035109 (2001).
  • Marzari et al. (2012) N. Marzari, A. A. Mostofi, J. R. Yates, I. Souza, and D. Vanderbilt, Rev. Mod. Phys. 84, 1419 (2012).
  • Mostofi et al. (2008) A. A. Mostofi, J. R. Yates, Y.-S. Lee, I. Souza, D. Vanderbilt, and N. Marzari, Comput. Phys. Commun. 178, 685 (2008).
  • Mostofi et al. (2014) A. A. Mostofi, J. R. Yates, G. Pizzi, Y.-S. Lee, I. Souza, D. Vanderbilt, and N. Marzari, Comput. Phys. Commun. 185, 2309 (2014).
  • 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, J. Phys.: Condens. Matter 32, 165902 (2020).
  • Sakurai and Napolitano (2017) J. J. Sakurai and J. Napolitano, Modern Quantum Mechanics, 2nd ed. (Cambridge University Press, 2017).
  • Kokalj (1999) A. Kokalj, J. Mol. Graphics Modell. 17, 176 (1999).