We present two new developments for computing excited state energies within the GW approximation. First, calculations of the Green’s function and the screened Coulomb interaction are decomposed into two parts: one is deterministic, while the other relies on stochastic sampling. Second, this separation allows constructing a subspace self-energy, which contains dynamic correlation from only a particular (spatial or energetic) region of interest. The methodology is exemplified on large-scale simulations of nitrogen-vacancy states in a periodic hBN monolayer and hBN-graphene heterostructure. We demonstrate that the deterministic embedding of strongly localized states significantly reduces statistical errors, and the computational cost decreases by more than an order of magnitude. The computed subspace self-energy unveils how interfacial couplings affect electronic correlations and identifies contributions to excited-state lifetimes. While the embedding is necessary for the proper treatment of impurity states, the decomposition yields new physical insight into quantum phenomena in heterogeneous systems.

First-principles treatment of electron excitation energies is crucial for guiding the development of new materials with tailored optoelectronic properties. Localized quantum states became the focal point for condensed systems1–7 since the lifetime of localized excitations can be controllably modified by interfacial couplings.8–11 Besides predicting experimental observables, theoretical investigations thus help elucidate the interplay of the electron–electron interactions and the role of the environment.

The excited electrons and holes near the Fermi level are conveniently described by the quasiparticle (QP) picture: the charge carriers are characterized by renormalized interactions and a finite lifetime limited by energy dissipation, which governs the deexcitation mechanism. Quantitative predictions of QPs necessitate the inclusion of non-local many-body interactions.

The prevalent route to describe QPs in condensed systems employs the Green’s function formalism.12,13 The excitation energy and its lifetime are inferred from the QP dynamics. The many-body effects are represented by the non-local and dynamical self-energy, Σ. In practice, Σ is approximated by selected classes of interactions, forming a hierarchy of systematically improvable methods.12 The formalism also allows constructing the self-energy for distinct states with a different form of Σ.

Here, we neglect the vibrational effects, and the induced density fluctuations dominate the response of the system to the excitation. The perturbation expansion is conveniently based on the screened Coulomb interaction, W, which explicitly incorporates the system’s polarizability. The neglect of the higher-order (vertex) terms,12 canceling the “self-screening” error and capturing two-particles interactions,14,15 leads to the popular GW method.12,16–18 This approach predicts the QP gap, ionization potentials, and electron affinities, in good agreement with experiments.16,18 Within the GW approximation, Σ is a product of W and the Green’s function, G, in the time domain.

Conventional implementations of the GW self-energy scale as O(N4) with the system size, and it is usually limited to few-electron systems. Recent developments have significantly reduced the prefactor of the conventional GW implementations.19–22 Furthermore, a hybrid stochastic–deterministic approach attempted to reduce the cost by constructing the stochastic “pseudobands” when summing over quasi-degenerate states.23 In contrast, our stochastic treatment recasts the full self-energy expression as a statistical estimator. The method employs sampling of electronic wavefunctions combined with the decomposition of operators.24–26 In practice, this formulation decreases the computational time significantly and leads to a linear scaling algorithm.24–26 

The stochastic approach employs real-space random vectors that sample the Hilbert space of the electronic Hamiltonian. The statistical error in Σ is governed by the number of vectors, which are used for the decomposition of G and the evaluation of W. Besides the calculation parameters, i.e., the number of stochastic vectors, the sampling critically depends on how uniformly the Hilbert space is represented. The statistical fluctuations are often large for localized orbitals in condensed systems (e.g., defects or moiré states) since it is problematic to sample both localized and delocalized states evenly in the real-space representation.27 This translates to an increased computational cost for defects, impurities, and molecules.26,27 Furthermore, too high fluctuations potentially hinder proper convergence and may result in sampling bias.

Here, we overcome this difficulty by constructing a hybrid deterministic–stochastic approach. We show how to efficiently decompose G and W in real-time and embed the strongly localized states. In this formalism, the problematic orbitals are treated explicitly without relying on their sampling by random vectors. A similar embedding scheme was so far employed only in static ground-state calculations.28 The decomposition is general and can be used to sample arbitrary excitations. Naturally, it is especially well suited for spatially or energetically isolated electronic states. Furthermore, we employ the decomposition technique to compute Σ stochastically from an arbitrarily large subspace of interest. We show that the self-energy separation disentangles correlation contributions from different spatial regions of the system.

After deriving the formalism of the self-energy embedding and decomposition, we illustrate our methods numerically for nitrogen vacancy in large periodic cells of hBN. This system represents a realistic simulation of a prototypical single-photon emitter.2,29–32 First, we demonstrate the reduction of the stochastic error for the defect states in the hBN monolayer. Next, we study the interaction of the defect in the hBN-graphene heterostructure.

The Green’s function (G), a time-ordered correlator of creation and annihilation field operators, describes the dynamics of an individual quasiparticle. The poles of G fully determine the single-particle excitations (as well as many other properties).

Solving directly for G is often technically challenging (or outright impossible). Alternatively, the Green’s function is often sought via a perturbative expansion of the electron–electron interactions on top of a propagator of non-interacting particles (i.e., the non-interacting Green’s function, G0). The two quantities are related via the Dyson equation G1=G01Σ, where Σ is a self-energy accounting, in principle, for all the many-body effects absent in G0.

Calculations usually employ only a truncated expansion of the self-energy. Despite ongoing developments,15,33,34 the most common approach is limited to the popular GW approximation to the self-energy, which is composed of the non-local exchange (ΣX) and polarization (ΣP) terms. In the time-domain, the latter operator is expressed as

ΣP(r,r,t)=iG(r,r,t)WP(r,r,t+),
(1)

where WP is the time-ordered polarization potential due to the time-dependent induced charge density.25 The potential is conveniently expressed using the reducible polarizability χ and the Coulomb kernel ν as

WP(r,r,t)=ν(r,r)χ(rr,t)ν(r,r)drdr.
(2)

Evaluating the action of WP on individual states is the practical bottleneck of the GW approach. Hence, despite that Eq. (1) together with the Dyson equation should be self-consistent, the QP energies are commonly computed by a “one-shot” correction, denoted as G0W0, where W0 is computed from the non-interacting starting point.12,17 In practice, Eq. (1) thus contains quantities constructed from the mean-field Hamiltonian, H0, comprising one-body terms and local Hartree, ionic, and exchange-correlation potentials. For WP, it is common to employ the random phase approximation (RPA). Beyond RPA, approaches are more expensive and, in general, do not improve the QP energies unless higher-order (vertex) terms are included in Σ.15,35

Within this one-shot framework, the QP energies become12 

εQP=ε0+ϕ|ΣX+ΣP(ω=εQP)vxc|ϕ,
(3)

where ε0 are eigenvalues of H0 and vxc is the (approximate) exchange-correlation potential. In Eq. (3), ΣP is in the frequency domain.

The stochastic G0W0 method seeks the QP energy via random sampling of wavefunctions and decomposition of operators in the real-time domain.24–26 The expectation value of Σ is expressed as a statistical estimator. The result is subject to fluctuations that decrease with the number of samples as 1/N.

In this formalism, the polarization self-energy expression is separable.24–26 Specifically, for a particular H0 eigenstate ϕ, the perturbative correction becomes

ϕ|ΣP(t)|ϕ=ϕ|iG0(t)WP(t)|ϕ1Nζ¯ζ¯ϕ(r)ζ(r,t)uζ(r,t)d3r,
(4)

where uζ(r, t) is an induced charge density potential and ζ is a random vector used for sampling of G0 (discussed in detail in Sec. III and the  Appendix). The state ζ at time t is

|ζ(t)U0,tPμ(t)|ζ,
(5)

where the U0,t is time evolution operator, detailed in the  Appendix,

U0,teiH0t.
(6)

The projector Pμ(t) selects the states above or below the chemical potential, μ, depending on the sign of t. In practice, Pμ(t) is directly related to the Fermi–Dirac operator.25,36–38 Since the Green’s function is a time-ordered quantity, the vectors in the occupied and unoccupied subspace are propagated backward or forward in time and contribute selectively to the hole and particle non-interacting Green’s functions.

The induced potential uζ(r, t) represents the time-ordered potential of the response to the charge addition or removal,

uζ(r,t)=WP(r,r,t)ζ¯(r)ϕ(r)d3r,
(7)

where ζ¯ spans the entire Hilbert space.24–26 

In practice, we compute uζ from the retarded response potential, which is ũζ=W̃P(r,r,t)ζ¯(r)ϕ(r)d3r; the time-ordering step affects only the imaginary components of the Fourier transforms of uζ and ũζ (see the  Appendix).13,25,26

The retarded response, ũζ, is directly related to the time-evolved charge density δn(r,t)n(r,t)n(r,t=0) induced by a perturbing potential δv,

ũζ(r,t)=ν(r,r)χ(r,r,t)δv(r,r)drdrdrν(r,r)δn(r,t)dr,
(8)

where we define a perturbing potential

δv=ν(r,r)ζ¯(r)ϕ(r).
(9)

Note that δv is explicitly dependent on the state ϕ and the ζ¯ vector; the latter is part of the stochastically decomposed G0 operator.

The stochastic formalism further reduces the cost of evaluating ũζ. Instead of computing δn(r,t) by a sum over single-particle states, we use another set of random vectors η confined to the occupied subspace. Time-dependent density n(r, t) thus becomes24–26,37–39

n(r,t)=limNη1NηlNη|ηl(r,t)|2,
(10)

where ηl is propagated in time using U0,t and H0 in Eq. (6) is implicitly time-dependent. As common,12,17 we resort to the density functional theory (DFT) starting point. Since H0 is therefore a functional of n(r,t), the same holds for the time evolution operator,

η(t)=U0,t[n(t)]η.
(11)

Furthermore, we employ RPA when computing ũζ; this corresponds to evolution within the time-dependent Hartree approximation.36,40,41

Practical calculations use only a limited number of random states. Consequently, the time evolved density exhibits random fluctuations at each space–time point. To resolve the response to δv, we use a two-step propagation whose difference is the δn that typically converges fast with Nη.24–26 

The stochastic vectors ζ and η, introduced in Sec. II B, are constructed on a real-space grid and sample the occupied (or unoccupied) states. The number of these vectors (Nζ and Nη) is increased so that the statistical errors are below a predefined threshold. Furthermore, the underlying assumption is that ζ and η sample the Hilbert space uniformly.

Here, we present a stochastic approach restricted only to a subset of states, while selected orbitals, ϕ, are treated explicitly and constitute an embedded subspace. We denote this set as the {ϕ}-subspace. In the context of the G0W0 approximation, we use the hybrid approach for (i) the Green’s function, (ii) the induced potential, or (iii) both G0 and u simultaneously. In the following, we present each case separately.

The non-interacting Green’s function is decomposed into two parts (omitting for brevity the space–time coordinates),

G0G̃0+G0ϕ,
(12)

where G0ϕ is the Green’s function of the constructed explicitly from ϕ as

G0ϕ(r,r,t)=j{ϕ}(1)θ(t)iϕj(r)ϕj*(r)eiεjt,
(13)

where θ is the Heaviside step function responsible for the time-ordering (corresponding to particle and hole contributions to G0). The complementary part, G̃0, is sampled with random states as in Eq. (4).

Unlike in the fully stochastic approach (where no G0ϕ term is present), the sampling vectors are constructed as orthogonal to the {ϕ}-subspace, i.e.,

ζ¯=1Pϕζ¯0.
(14)

Here, ζ¯0 spans, in principle, the entire Hilbert space and Pϕ is

Pϕ=jϕϕjϕj.
(15)

The construction of G̃0 remains the same as in the fully stochastic case: G̃0 is decomposed by a pair of vectors ζ¯ and ζ(t), cf. Eqs. (5), (6), and (14).

Note that it is possible to generalize the projector Pϕ, Eq. (15), to an arbitrary ϕ-subspace. The particular choice of ϕ does not affect the decomposition of the Green’s function [Eq. (12)] or the time evolution of ζ(t). However, the dynamics of G0ϕ would require explicit action of U0,t on ϕ that is, in principle, not an eigenstate of H0. Furthermore, instead of an intrinsically localized ϕ-subspace, one can also employ localized orbitals constructed by a unitary transformation of orbitals.42–44 We do not pursue this route here and select {ϕ}-subspace composed from the starting point eigenstates.

The retarded induced potential ũζ is decomposed through partitioning of the time-dependent charge density, cf., Eq. (8) (omitting the space–time coordinates),

nñ+nϕ,
(16)

where the nϕ is the density constructed from occupied states ϕ (which we assume to be mutually orthogonal),

nϕ(r,t)=j{ϕ}fjϕj(r,t)2.
(17)

Here, fj is the occupation of the jth state. Note that choosing ϕ within the unoccupied subspace is meaningless in this context since only occupied states contribute to the charge density nϕ. The complementary part, ñ, is given by stochastic sampling [Eq. (10)] that employs random vectors η̃,

η̃=1P̃ϕη,
(18)

where η spans the entire occupied subspace and

P̃ϕ=jϕfjϕjϕj.
(19)

The time propagation of the charge density is similar to the fully stochastic case. The deterministic and stochastic vectors evolve as

ϕj(t)=U0,t[nϕ(t),ñ(t)]ϕj,
(20)
η̃(t)=U0,t[nϕ(t),ñ(t)]η̃,
(21)

where U0,t is explicitly expressed as a functional of the two density contributions from Eq. (16). Note that these expressions are analogous to Eq. (11).

Both partitionings are used in conjunction. While G0 contains contributions from both occupied and unoccupied states, only the former are included in ũζ. The combined partitioning may use entirely different subspaces for the Green’s function and the induced potential. In Sec. V C, we employ the decomposition in both terms because it yields the best results and significantly reduces statistical fluctuations.

In Sec. III we partition retarded induced potential and the Green’s function, intending to decrease stochastic fluctuations in the self-energy. Here, we use the partitioning to achieve the second goal of this paper—to determine the contribution to ΣP(ω).

Conceptually, we want to address quasiparticle scattering by correlations from a particular subspace. In the expression for ΣP(ω) [Eq. (4)], this corresponds to accounting for selected charge density fluctuations in ũζ. In practice, we construct the subspace polarization self-energy as

ϕ|ΣPs(t)|ϕ=1Nζ¯ζ¯ϕ(r)ζ(r,t)uζs(r,t)d3r,
(22)

where we introduced the subspace induced potential uζs, which is obtained from its retarded form [in analogy to Eq. (8)],

ũζs(r,t)=ν(r,r)δns(r,t)dr.
(23)

This potential stems from the induced charge density that includes contributions only from selected orbitals {ϕ}. Note that nsr,t is obtained either from individual single-particle states or from the stochastic sampling of the {ϕ}-subspace under consideration.

If the set of ϕ states is large, it is natural to employ the stochastic approach; the density is sampled according to Eq. (10) with vectors ηs prepared as

ηs=P̃ϕη,
(24)

where the projector is in Eq. (19) and vectors η span the entire occupied subspace.

The time evolution of ηs follows Eq. (11), i.e., it is governed by the operator U0,t that depends on the total time-dependent density,

ηs(t)=U0,t[n(t)]ηs.
(25)

Hence, despite ΣPs(t) contains only fluctuations from a particular subspace, the calculation requires knowledge of the time evolution of both ns and n.

In practice, we employ a set of two independent stochastic samplings: (i) vectors η describing the entire occupied space and (ii) vectors ηs confined only to the chosen {ϕ}-subspace. The first set characterizes the total change density fluctuation and enters U0,t in Eq. (25). The stochastic–deterministic GW algorithm is in the  Appendix.

In this section, we will demonstrate the capabilities of the method introduced above. The starting-point calculations are performed with a real-space DFT implementation, employing regular grids, Troullier–Martins pseudopotentials,45 and the Generalized gradient approximation (GGA)46 functional for exchange and correlation. We investigate finite and 2D infinite systems using modified periodic boundary conditions with Coulomb interaction cutoffs.47 

The numerical verification for the SiH4 molecule is in Sec. V B. To converge the occupied H0 eigenvalues to <5 meV, we use a kinetic energy cutoff of 26 Eh and 64 × 64 × 64 real-space grid with the step of 0.3 a0.

Large calculations for the VN defect in hBN monolayer and in hBN heterostructure with graphene are in Secs. V C and V D. The heterostructure is built with an interlayer distance of 3.35 Å. In both cases, we consider relaxed rectangular 12 × 6 supercells containing 287 and 575 atoms. We performed the structural optimization in the QuantumESPRESSO code48 together with Tkatchenko–Scheffler’s total energy corrections49 and effective screening medium method.50 Our band structure calculations in low dimensional systems have been validated against the QuantumESPRESSO calculations.27 

The GW calculations were performed using a development version of the StochasticGW code.24–26 The calculations employ an additional set of 20 000 random vectors used in the sparse stochastic compression used for time-ordering of ũζ.25 The time propagation of the induced charge density is performed for a maximum propagation time of 50 a.u., with the time step of 0.05 a.u.

We verify the implementation of the methods presented in Secs. III and IV on the SiH4 molecule. In particular, we test separately (i) the construction of the embedding schemes, (ii) decomposition of the time propagation, and (iii) the evaluation of the subspace self-energy.

The reference calculation employs only one level of stochastic sampling (for the Green’s function, while the rest of the calculation is deterministic). For small systems, this approach converges fast as the stochastic fluctuations are small.25,26 Yet, we need Nζ = 1500 vectors to decrease the QP energies errors below 0.08 eV. An illustration of the self-energy for the HOMO state of SiH4 is in Fig. 1; everywhere in this figure, the stochastic error is below 0.13 eV for all frequencies.

FIG. 1.

Verification for the SiH4 molecule. (a) The comparison of the total self-energies of the highest occupied state (HOMO) obtained with reference and embedding schemes. The figure demonstrates that all five approaches yield identical Σ(ω). In the plot, G denotes the decomposition scheme where HOMO is embedded in the Green’s function, W denotes the calculation with HOMO state embedded in the screened Coulomb potential, and G+W corresponds to simultaneous decomposition of G and W. The red dashed line represents the sum over subspace contributions shown in the panel below. (b) Subspace self-energy of the HOMO state that contains only a contribution of a specific state i, denoted as Σi in the legend.

FIG. 1.

Verification for the SiH4 molecule. (a) The comparison of the total self-energies of the highest occupied state (HOMO) obtained with reference and embedding schemes. The figure demonstrates that all five approaches yield identical Σ(ω). In the plot, G denotes the decomposition scheme where HOMO is embedded in the Green’s function, W denotes the calculation with HOMO state embedded in the screened Coulomb potential, and G+W corresponds to simultaneous decomposition of G and W. The red dashed line represents the sum over subspace contributions shown in the panel below. (b) Subspace self-energy of the HOMO state that contains only a contribution of a specific state i, denoted as Σi in the legend.

Close modal

We first inspect the results for an embedded deterministic subspace. Figure 1 shows that the different schemes for the explicit treatment of the HOMO state (Sec. III) produce the same self-energy curve. The inclusion of HOMO in W is combined with the stochastic sampling of the three remaining orbitals by Nη = 16 random vectors. Note that it is not economical to sample the action of WP for small systems,25,26 and this calculation serves only as a test case. This treatment yields a statistical error of 0.08 eV, i.e., the same as the reference calculation despite the additional fluctuations due to the η vectors.

When the HOMO orbital is explicitly included in G, the resulting statistical error is below the error of the reference calculation (0.05 eV). Such a result is expected since the reference relies on the stochastic sampling of the Green’s function. The same happens when the frontier orbital is in both G and W (Nη = 16 random vectors sample the induced charge density). Tests for other states are not presented here but lead to identical conclusions.

Next, we verify that the density entering the time-evolution operator U0,t[n(t)] can be constructed from different states than it is acting on. Specifically, the induced charge and the time-dependent densities may be sampled and built by different means. To demonstrate this, we propagate each of the H0 eigenstates with U0,t[n(t)], where n(t) is stochastically sampled by Nη = 16 random vectors. Only the induced charge density, entering Eq. (8), is computed from the {ϕ} eigenvectors. The agreement with the reference self-energy curve is excellent with differences smaller than the standard deviations at each frequency point [see Fig. 1(a)].

Finally, we inspect the subspace self-energy in which U0,t[n(t)] employs the total charge density sampled by Nη = 16 random vectors. Figure 1(b) shows four different Σs(ω) curves corresponding to the contributions of individual H0 eigenstates. Since the eigenstates are orthogonal, the total self-energy is simply the sum of individual Σs(ω) components. The additivity of the subspace self-energies is demonstrated numerically in Fig. 1(a). The subspace results illustrate that HOMO and the bottom valence orbital exhibit the largest amplitudes of ΣPs; hence, these two states dominate the correlation near the ionization edge.

The deterministic subspace embedding should numerically stabilize the stochastic sampling and decreases the computational cost. To test the methodology on a realistic system, we consider the electronic structure of an infinitely periodic hBN monolayer containing a single nitrogen vacancy (VN) per a unit cell with dimensions of 3.0 × 2.6 nm2. The system comprises 1147 electrons with the defect state being singly occupied and hence positioned at the Fermi level.

In the current calculations, we enforce spin degeneracy. The reason is twofold: (i) half populated states are strongly polarizable and they exhibit stronger stochastic fluctuations in the time evolution. They are thus a more stringent test of the embedding. (ii) In Sec. V D, we compare the monolayer with a heterostructure to determine the role of substrate material on the self-energy. In the heterostructure, the magnetic splitting of spin-up/down components disappears.51 

The relaxed monolayer geometry shows only mild restructuring. The vacancy introduces three localized states with the D3h point group symmetry.51,52 The singly occupied in-gap state is labeled by its irreducible representation a2.52 The second state (e′) is doubly degenerate and pushed high in the conduction region. Due to the enforced spin-degeneracy, the electron–electron interactions are increased and e′ appears higher than in the previous calculations.29,53–55 The a2 and e′ single-particle wavefunctions are illustrated in Fig. 2.

FIG. 2.

Deterministic embedding for the nitrogen vacancy in hBN monolayer. (a) Black points represent individual states of the unfolded band structure; blue lines are a guide for the eyes. Panels (b)–(d) depict the self-energy curves (solid line) for e′, a2, and VBM state, respectively; the states are marked in panel (a) by red dashed rectangles. The statistical errors are smaller than the thickness of the line. Gray shaded areas (on each plot) correspond to the “spread” of self-energy curves for 15 distinct fully stochastic calculations; blue solid lines are the statistical averages. The light red area in panel (c) represents the spread after deterministic embedding of the a2 state; the red solid line is the self-energy curve. The third column contains the electron density of the corresponding states presented at the same isovalue.

FIG. 2.

Deterministic embedding for the nitrogen vacancy in hBN monolayer. (a) Black points represent individual states of the unfolded band structure; blue lines are a guide for the eyes. Panels (b)–(d) depict the self-energy curves (solid line) for e′, a2, and VBM state, respectively; the states are marked in panel (a) by red dashed rectangles. The statistical errors are smaller than the thickness of the line. Gray shaded areas (on each plot) correspond to the “spread” of self-energy curves for 15 distinct fully stochastic calculations; blue solid lines are the statistical averages. The light red area in panel (c) represents the spread after deterministic embedding of the a2 state; the red solid line is the self-energy curve. The third column contains the electron density of the corresponding states presented at the same isovalue.

Close modal

We first focus on the delocalized top valence and bottom conduction states. The fully stochastic calculations converge fast for both of them. The charge density fluctuations are sampled by Nη = 8 vectors, and the Green’s function requires Nζ = 1500 to yield QP energies with statistical errors below 0.03 eV for the valence band maximum (VBM). The error for the conduction band minimum (CBM) is <0.01 eV. The computed resulting quasiparticle bandgap is 6.49 ± 0.04 eV, in excellent agreement with previous calculations and experiments (providing a range of values between 6.1 eV and 6.6 eV).56–58 

To investigate the electronic structure in greater detail, we employ the projector-based energy–momentum analysis based on supercell band unfolding.27,59–61 In practice, the individual wavefunctions within our simulation cell are projected onto the Brillouin zone of a single hBN unit cell. The resulting band structure is shown in Fig. 2(a). Since our calculations employ a rectangular supercells, the critical point K of the hexagonal Brillouin zone appears between the Γ and X points (marked on the horizontal axis by ⋆). The figure shows that the fundamental bandgap is indirect; the direct transition is 7.39 ± 0.04 eV, extremely close to the results for the pure hBN monolayer (reported to be in a range between 7.26 eV and 7.37 eV).58,62,63

The defect states appear as flat bands, labeled in Fig. 2 by their irreducible representations. As expected from the outset, the stochastic calculations for a2 exhibit large fluctuations. While each sample is numerically stable, random vectors produce a self-energy curve with a significant statistical error at each frequency point, i.e., the time evolution is strongly dependent on the initial choice of {ζ} and {η}. In this example, only the a2 state exhibits such behavior.

Figure 2 illustrates the self-energies for the band edge and the defect states. For all the cases, the plots show the spread of the ΣP(ω) curves: these correspond to the outer envelope for 15 distinct calculations, each employing 100 ζ sampling vectors (combined with Nη = 8 each). For the a2 defect state, the stochastic sampling is possibly biased. The variation is three times as big as the spread of the VBM and almost seven times larger compared to CBM. The large spread is associated with increased statistical errors (50 meV), notably larger than those of delocalized VBM (30 meV) and CBM (9 meV) states. Away from the QP energy, the fluctuations increase even further; the spread of the samples becomes two times larger near the maximum of ΣP at −20 eV. The convergence of the QP energy is poor, and the low sample standard deviation (roughly twice as big as for VBM) suggests incorrect statistics. In practice, each sampling yields a self-energy curve that lies outside of the standard deviation of the previous simulation.

The deterministic embedding remedies insufficient sampling without significantly impacting the computational cost. Naturally, we select the a2 defect state and treat it explicitly (while randomly sampling the rest of the orbitals). Hence, according to the notation of Sec. III, the {ϕ}-subspace contains only a single orbital.

The decomposition of the non-interacting Green’s function follows Eq. (12) and stabilizes the sampling. The spread of the self-energy curves decreases approximately three times for a wide range of frequencies. With the embedded a2 state, the statistical error of the QP energy decreases smoothly and uniformly with the number of samples. Each new sampling falls within the error of the calculations. Yet, the final statistical error (0.03 eV) is less than 10% larger than for the delocalized states.

The decomposition of the induced charge density alone is less promising. Fundamentally, δn(r, t) contains contributions from the entire system, and a small {ϕ}-subspace will unlikely lead to drastic improvement. Indeed, the statistical errors and convergence behavior remain the same as for the fully stochastic treatment.

Embedding of the localized state in both G0 and δn(r, t) is, however, the best strategy. If both decompositions use the same (or overlapping) {ϕ}-subspace, the induced charge density and the potential δv [Eq. (9)] share (at least some) ϕ states. Consequently, the sampling of ΣP(ω) becomes less dependent on the particular choice of ζ vectors, which sample only states orthogonal to {ϕ}. Indeed, the embedding of a single localized state in G and W results in a nearly fourfold reduction of the statistical fluctuation that translates to more than an order of magnitude savings in the computational time. The error of the QP energy of the a2 state is approximately half of the error for VBM (16 meV for Nζ = 1500).

For completeness, we also applied the three types of embedding on VBM, which does not suffer from bias or large statistical errors. Unlike for the localized state, we observe only a negligible reduction in the fluctuations. Specifically, the stochastic error for VBM decreases from 30 meV to 29 meV regardless of whether G or G + W embedding is used. For delocalized states, the fully stochastic sampling is thus sufficient as expected from our previous work.25–27 Note that the total cost of the calculation scales quadratically25 with the number of states used to evaluate the retarded potential ũζ [Eq. (8)]. An optimal computational cost is achieved when strongly localized states are treated via embedding, but the subspace of delocalized states is sampled stochastically.

Having an improved description of the localized states in hand, we now turn to the decomposition of the self-energy. The real-time approach described in Sec. IV allows inspecting the many-body interactions from a selected portion of the system. In contrast to Subsection V C, the subspace of interest contains a large number of states (irrespective of their degree of localization), and it is randomly sampled. The goal of this decomposition is to understand the role of correlation, especially at interfaces.

To test our method, we investigate a periodic hBN monolayer containing a single VN defect placed on graphene. Such a heterostructure has also been realized experimentally.64–67 The structure contains 2299 valence electrons, and it is illustrated in Fig. 3 together with selected orbital isosurfaces. The a2 state is energetically lower than e′ (as in the pristine hBN monolayer). Both defects only weakly hybridize with graphene and remain localized within the hBN sublayer. However, the presence of graphene leads to an increased delocalization of the defect states within the monolayer68 and decrease in the hBN fundamental bandgap, attributed to image-charge effects.69 

FIG. 3.

Excitations in an hBN-graphene heterostructure containing a nitrogen vacancy. (a) Calculated band structure—the states localized on graphene are in green, while blue points represent hBN.70 Lines serve as a guide for the eyes. The individual isosurfaces are shown on the right for VBM, a2 defect state, and e′ defect states. Second row contains self-energy curves with vertical lines indicating position of the QP energies: (b) VBM with QP energy −6.9 eV, (c) a2 defect state with QP energy −2.6 eV, (d) e′ defect state with QP energy −1.4 eV, and (e) CBM with QP energy −0.6 eV. The frequency axis in panels (b)–(e) is relative to the vacuum level. Black solid lines are the full heterostructure self-energies; green and blue dashed lines are the ΣPs(ω) graphene monolayer and hBN monolayer subspace self-energies, respectively. The blue line represents the self-energy of an isolated hBN monolayer with VN (discussed in Sec. V C). The third row of panels contains the ImΣPs(ω).

FIG. 3.

Excitations in an hBN-graphene heterostructure containing a nitrogen vacancy. (a) Calculated band structure—the states localized on graphene are in green, while blue points represent hBN.70 Lines serve as a guide for the eyes. The individual isosurfaces are shown on the right for VBM, a2 defect state, and e′ defect states. Second row contains self-energy curves with vertical lines indicating position of the QP energies: (b) VBM with QP energy −6.9 eV, (c) a2 defect state with QP energy −2.6 eV, (d) e′ defect state with QP energy −1.4 eV, and (e) CBM with QP energy −0.6 eV. The frequency axis in panels (b)–(e) is relative to the vacuum level. Black solid lines are the full heterostructure self-energies; green and blue dashed lines are the ΣPs(ω) graphene monolayer and hBN monolayer subspace self-energies, respectively. The blue line represents the self-energy of an isolated hBN monolayer with VN (discussed in Sec. V C). The third row of panels contains the ImΣPs(ω).

Close modal

By unfolding the wave functions of the bilayer onto the hBN conventional rectangular unit cell, we obtain the band structure shown in Fig. 3(a). The graphene and hBN bands are distinguished by a state projection on the densities above and below the center of the interlayer region. Some states, however, extend appreciably across this boundary and lead to the appearance of spurious low-intensity peaks in the unfolded band structure.70 Yet, individual bands are easy to distinguish, and one can clearly disentangle individual hBN and graphene states.

The graphene portion of the band structure reproduces the well-known semimetallic features with a Dirac point located at the K boundary of the hexagonal Brillouin zone. As discussed above, the K appears in between Γ and X of the rectangular cell, and it is labeled by ⋆ in Fig. 3. The typical Dirac cone dispersion of graphene is only a little affected by the hBN presence, and the Dirac point remains close to the Fermi level despite the charge transfer from the a2 defect51 This is not surprising: sparse charge defects lead to only weak doping of graphene. Furthermore, the previous DFT results employed small supercells and may suffer from significant electronic “overdelocalization”71,72 that spuriously enhances charge transfer.

The hBN part of the band structure (Fig. 3) is only weakly affected by the heterostructure formation. The delocalized states are qualitatively identical to those in the monolayer (Fig. 2). The fundamental bandgap remains indirect and reduced to 6.32 ± 0.04 eV. The screening introduced by graphene thus leads to a small change in Eg (<0.2 eV compared to the monolayer). The positions of the defect states, however, change notably. Here, both appear above the Fermi level due to the charge transfer from a2 to the graphene layer.

Due to the charge transfer and Fermi level shift, it is clear that graphene is responsible for altering the charge fluctuations, i.e., the polarization part of the self-energy. In practice, graphene acts as a dielectric background inducing a significant screening of the Coulomb interaction in hBN. It stands to reason that the localized defect states would be strongly affected by such a polarizable layer and that the corresponding self-energy should be dominated by the spectral features originating in graphene.

To investigate the degree of coupling and the contribution to the self-energy from each monolayer, we compute ΣPs(ω) using Eq. (22). Here, the {ϕ}-subspace is constructed from stochastic samples of the 576 occupied graphene states (distinguished by the green color in Fig. 3). For each sampling of the Green’s function, the subspace is described by eight random vectors in the {ϕ}-subspace. Additional eight vectors sample the complementary subspace.

Figure 3 shows the decomposition of the self-energy, indicating the contribution of graphene. For the delocalized states, the ΣP(ω) curves of the heterostructure (black solid lines) appear similar to those of an hBN monolayer (blue solid lines). Specifically, we see that ΣP(ω) for VBM [panel (b)] has practically identical frequency dependence between −15 eV and 0 eV. While in-plane contributions from hBN (blue dashed line) dominate the entire curve, screening from the graphene (green dashed line) substrate is substantial. A naïve comparison between the self-energy curves of the hBN monolayer and the heterostructure suggests that the enhanced maximum in ReΣP at −25 eV (marked by an asterisk) is caused extrinsically by graphene. However, this is not the case, and the effect is only indirect: the presence of graphene leads to shifting of the hBN spectral features. At the QP energy (marked by the dashed vertical line), graphene contributes to total ΣP of the heterostructure by 36%.

The situation is different for CBM [panel (e)]. Toward the static limit (ω → 0), the self-energy curve becomes more negative. This shift is caused directly by the induced density fluctuations in the graphene substrate. The substrate screening renormalizes the quasiparticle bandgap, as has been shown in the literature of low-dimensional materials.9,27,69,73–75

The self-energy curves of the a2 panel (c) and e′ [panel (d)] defect states are significantly different from those computed for the monolayer. In both cases, we observe a negative shift related to the QP stabilization by non-local correlations. The ΣP contributions to the defects’ QP energies are roughly three times larger in the heterostructure. Here, the intrinsic hBN interactions are the major driving force. The defect states are mostly affected by the in-plane induced density fluctuations (constituting 64% and 76% of the total polarization self-energy for a2 and e′). Like the VBM, graphene indirectly acts on the defect QP states by enhancing the in-plane fluctuations, rather than direct coupling to the defects.

In the same vein, ImΣ(ω = εQP) of the graphene subspace is small, showing that there is only a small lifetime broadening due to the substrate. In detail, the graphene monolayer contributes to the broadening of the VBM state, a2 state, and e′ around twice less than the hBN subspace. The only exception is the CBM state for which the substrate contribution dominates (by a factor of ∼3). As shown in this example, the stochastic subspace self-energy efficiently probes dynamical electron–electron interactions at interfaces. The random sampling allows selecting an arbitrary subspace to identify its contribution to the correlation energy and the excited state lifetimes, obtained from real and imaginary parts of Σ(ω = εQP), respectively.

Stochastic Green’s function approaches represent a class of low-scaling many-body methods based on stochastic decomposition and random sampling of the Hilbert space. The overall computational cost is impacted by the presence of localized states that increase the sampling errors. Here, we have introduced a practical solution for the G0W0 method. The localized states are treated explicitly (deterministically), while the rest is subject to stochastic sampling. We have further shown that the subspace-separation can be applied to decompose the dynamical self-energy into contributions from distinct states (or parts of the systems).

Using nitrogen vacancy in hBN monolayer, we demonstrate that the deterministic embedding dramatically reduces the statistical fluctuations. Consequently, the computational time decreases by more than an order of magnitude for a given target level of stochastic error. The embedding may be made within the Green’s function and the screened Coulomb interactions independently or simultaneously; the latter provides the best results. For delocalized states, embedding is not necessary and performs similar to the standard uniform sampling with real-space random vectors. The mechanism of embedding presented here is general, and it is the first step toward hybrid techniques that treat selected subspace at the distinct (higher) level of theory. The development of hybrid techniques employing various orbital localization procedures is currently underway.

We further demonstrate that the subspace self-energy contains (in principle, additive) contributions from the induced charge density. The self-energy contribution from a particular portion of the system is computed by confining sampling vectors into different (orthogonal) subspaces. We exemplify the capabilities of such calculations on defects state in the hBN-graphene heterostructure. Here, the charged excitation in one layer directly couples to density oscillations in the substrate monolayer. The electronic correlations for the defects are governed by the interactions in the host layer; the substrate only indirectly affects their strength.

This example serves as a stimulus for additional study of the defect states in heterostructures. Here, the coupling strength between individual subsystems is tunable by particular stacking order and induced strains. The subspace self-energy represents a direct route to explore such quantum many-body interfacial phenomena.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

This work was supported by the NSF through the Materials Research Science and Engineering Centers (MRSEC) Program of the NSF through Grant No. DMR-1720256 (Seed Program). We gratefully acknowledge support via the UC Santa Barbara NSF Quantum Foundry funded via the Q-AMASE-i program under Award No. DMR-1906325. The calculations were performed as part of the XSEDE76 computational Project No. TG-CHE180051. Use was made of computational facilities purchased with funds from the National Science Foundation (Grant No. CNS-1725797) and administered by the Center for Scientific Computing (CSC). The CSC is supported by the California NanoSystems Institute and the Materials Research Science and Engineering Center (MRSEC; Grant No. NSF DMR-1720256) at UC Santa Barbara.

The new implementation is derived from the fully stochastic algorithm presented in Ref. 25. The steps (1), (2), and (a) are new, i.e., a result of the embedding introduced in this article. Steps (b)–(d) and (3) are similar to the fully stochastic scheme, and they employ hybrid quantities [i.e., decomposed time-dependent density Eq. (16) and/or the decomposed Green’s function Eq. (12)]. Step 4 is identical to the original algorithm.

The steps of the algorithm are as follows:

  1. Select a set of deterministic states ϕ for embedding or subspace decomposition. Construct the projector Pϕ [Eq. (15)] for the Green’s function and/or P̃ϕ [Eq. (19)] for the retarded induced potential.

  2. Create Nζ¯0 stochastic vectors (typically in the order of 102–103) for sampling the full Hilbert space. If including deterministic states in the Green’s function, construct ζ¯ via projection with Pϕ [Eq. (14)]. For each sampling of the Green’s function, perform the next steps:

    • Create a set of Nη̃ (typically <10) random vectors in the occupied subspace. Use P̃ϕ [Eq. (19)] and either

      • construct η̃ via Eq. (18) for the embedding scheme

      • or construct η̃s via Eq. (24) for the subspace self-energy decomposition.

    • Calculate δv(r) by Eq. (9) and perturb η̃ and/or ϕ as

η̃λ(r)=eiλδv(r)η̃(r),
(A1)
ϕλ(r)=eiλδv(r)ϕ(r),
(A2)

with λ103105Eh1.

|η̃λ(t+dt)=eiHλ(t)dt|η̃λ(t),
(A3)
|ϕλ(t+dt)=eiHλ(t)dt|ϕλ(t).
(A4)

Two propagations are needed: with λ = 0 and with a finite value of λ. We employ the random phase approximation, corresponding to the time-dependent Hartree approach. The time evolution thus depends only on the total charge density constructed using Eq. (16).

ũζ(r,t)=vHλ(r,t)vHλ=0(r,t)λ,
(A5)

and then apply a time-ordering operation13 in frequency domain,

uζ(r,ω)=ũζ(r,ω),ω0,(ũζ(r,ω))*,ω<0.
(A6)
  1. Compute ζ(r, t) for negative and positive times [for hole and particle components of the Green’s function—Eq. (5)]. If embedding in the Green’s function is applied, construct Gϕ components using Eq. (13). Accumulate the matrix element of the self-energy via Eq. (4).

  2. When steps (1)–(3) are performed, average the resulting ⟨ϕ|Σ(t)|ϕ⟩ from each sampling of the Hilbert space (which employed ζ¯ and ϕ orbitals). Perform Fourier transformation of the self-energy, and solve Eq. (3) in the frequency domain.

  • (c)

    Perform time-propagation in discrete time steps dt (typically 0.05 a.u.) for states η̃λ(r) and ϕλ(r),

  • (d)

    Compute ũζ(t) as the difference of perturbed and unperturbed Hartree potentials at each time step for λ = 0 and for a finite value of λ,

1.
H.
Yu
,
G.-B.
Liu
,
J.
Tang
,
X.
Xu
, and
W.
Yao
, “
Moiré excitons: From programmable quantum emitter arrays to spin-orbit–coupled artificial lattices
,”
Sci. Adv.
3
,
e1701696
(
2017
).
2.
G.
Grosso
,
H.
Moon
,
B.
Lienhard
,
S.
Ali
,
D. K.
Efetov
,
M. M.
Furchi
,
P.
Jarillo-Herrero
,
M. J.
Ford
,
I.
Aharonovich
, and
D.
Englund
, “
Tunable and high-purity room temperature single-photon emission from atomic defects in hexagonal boron nitride
,”
Nat. Commun.
8
,
705
(
2017
).
3.
M.
Yankowitz
,
J.
Jung
,
E.
Laksono
,
N.
Leconte
,
B. L.
Chittari
,
K.
Watanabe
,
T.
Taniguchi
,
S.
Adam
,
D.
Graf
, and
C. R.
Dean
, “
Dynamic band-structure tuning of graphene moiré superlattices with pressure
,”
Nature
557
,
404
408
(
2018
).
4.
Y.
Cao
,
V.
Fatemi
,
A.
Demir
,
S.
Fang
,
S. L.
Tomarken
,
J. Y.
Luo
,
J. D.
Sanchez-Yamagishi
,
K.
Watanabe
,
T.
Taniguchi
,
E.
Kaxiras
,
R. C.
Ashoori
, and
P.
Jarillo-Herrero
, “
Correlated insulator behaviour at half-filling in magic-angle graphene superlattices
,”
Nature
556
,
80
84
(
2018
).
5.
U.
Zondiner
,
A.
Rozen
,
D.
Rodan-Legrain
,
Y.
Cao
,
R.
Queiroz
,
T.
Taniguchi
,
K.
Watanabe
,
Y.
Oreg
,
F.
von Oppen
,
A.
Stern
,
E.
Berg
,
P.
Jarillo-Herrero
, and
S.
Ilani
, “
Cascade of phase transitions and Dirac revivals in magic-angle graphene
,”
Nature
582
,
203
208
(
2020
).
6.
M. E.
Turiansky
,
A.
Alkauskas
, and
C. G.
Van de Walle
, “
Spinning up quantum defects in 2D materials
,”
Nat. Mater.
19
,
487
489
(
2020
).
7.
A.
Gottscholl
,
M.
Kianinia
,
V.
Soltamov
,
S.
Orlinskii
,
G.
Mamin
,
C.
Bradac
,
C.
Kasper
,
K.
Krambrock
,
A.
Sperlich
,
M.
Toth
,
I.
Aharonovich
, and
V.
Dyakonov
, “
Initialization and read-out of intrinsic spin defects in a van der Waals crystal at room temperature
,”
Nat. Mater.
19
,
540
545
(
2020
).
8.
C.
Wang
,
C.
Axline
,
Y. Y.
Gao
,
T.
Brecht
,
Y.
Chu
,
L.
Frunzio
,
M. H.
Devoret
, and
R. J.
Schoelkopf
, “
Surface participation and dielectric loss in superconducting qubits
,”
Appl. Phys. Lett.
107
,
162601
(
2015
).
9.
D. Y.
Qiu
,
F. H.
da Jornada
, and
S. G.
Louie
, “
Environmental screening effects in 2D materials: Renormalization of the bandgap, electronic structure, and optical spectra of few-layer black phosphorus
,”
Nano Lett.
17
,
4706
4712
(
2017
).
10.
A.
Tartakovskii
, “
Moiré or not
,”
Nat. Mater.
19
,
581
582
(
2020
).
11.
L.
Yuan
,
B.
Zheng
,
J.
Kunstmann
,
T.
Brumme
,
A. B.
Kuc
,
C.
Ma
,
S.
Deng
,
D.
Blach
,
A.
Pan
, and
L.
Huang
, “
Twist-angle-dependent interlayer exciton diffusion in WS2–WSe2 heterobilayers
,”
Nat. Mater.
19
,
617
623
(
2020
).
12.
R. M.
Martin
,
L.
Reining
, and
D. M.
Ceperley
,
Interacting Electrons
(
Cambridge University Press
,
2016
).
13.
A. L.
Fetter
and
J. D.
Walecka
,
Quantum Theory of Many-Particle Systems
(
Dover Publications
,
2003
).
14.
C.
Mejuto-Zaera
,
G.
Weng
,
M.
Romanova
,
S. J.
Cotton
,
K. B.
Whaley
,
N. M.
Tubman
, and
V.
Vlček
, “
Are multi-quasiparticle interactions important in molecular ionization?
,” arXiv:2009.02401 [physics.chem-ph] (
2020
).
15.
V.
Vlček
, “
Stochastic vertex corrections: Linear scaling methods for accurate quasiparticle energies
,”
J. Chem. Theory Comput.
15
,
6254
6266
(
2019
).
16.
L.
Hedin
, “
New method for calculating the one-particle Green’s function with application to the electron-gas problem
,”
Phys. Rev.
139
,
A796
A823
(
1965
).
17.
M. S.
Hybertsen
and
S. G.
Louie
, “
Electron correlation in semiconductors and insulators: Band gaps and quasiparticle energies
,”
Phys. Rev. B
34
,
5390
5413
(
1986
).
18.
F.
Aryasetiawan
and
O.
Gunnarsson
, “
The GW method
,”
Rep. Prog. Phys.
61
,
237
312
(
1998
).
19.
P.
Umari
,
G.
Stenuit
, and
S.
Baroni
, “
GW quasiparticle spectra from occupied states only
,”
Phys. Rev. B
81
,
115104
(
2010
).
20.
M.
Govoni
and
G.
Galli
, “
Large scale GW calculations
,”
J. Chem. Theory Comput.
11
,
2680
2696
(
2015
).
21.
F.
Bruneval
and
X.
Gonze
, “
Accurate GW self-energies in a plane-wave basis using only a few empty states: Towards large systems
,”
Phys. Rev. B
78
,
085125
(
2008
).
22.
W.
Gao
,
W.
Xia
,
X.
Gao
, and
P.
Zhang
, “
Speeding up GW calculations to meet the challenge of large scale quasiparticle predictions
,”
Sci. Rep.
6
,
36849
(
2016
).
23.
M.
Del Ben
,
F. H.
da Jornada
,
A.
Canning
,
N.
Wichmann
,
K.
Raman
,
R.
Sasanka
,
C.
Yang
,
S. G.
Louie
, and
J.
Deslippe
, “
Large-scale GW calculations on pre-exascale HPC systems
,”
Comput. Phys. Commun.
235
,
187
195
(
2019
).
24.
D.
Neuhauser
,
Y.
Gao
,
C.
Arntsen
,
C.
Karshenas
,
E.
Rabani
, and
R.
Baer
, “
Breaking the theoretical scaling limit for predicting quasiparticle energies: The stochastic GW approach
,”
Phys. Rev. Lett.
113
,
076402
(
2014
).
25.
V.
Vlček
,
W.
Li
,
R.
Baer
,
E.
Rabani
, and
D.
Neuhauser
, “
Swift GW beyond 10 000 electrons using sparse stochastic compression
,”
Phys. Rev. B
98
,
075107
(
2018
).
26.
V.
Vlček
,
E.
Rabani
,
D.
Neuhauser
, and
R.
Baer
, “
Stochastic GW calculations for molecules
,”
J. Chem. Theory Comput.
13
,
4997
5003
(
2017
).
27.
J.
Brooks
,
G.
Weng
,
S.
Taylor
, and
V.
Vlček
, “
Stochastic many-body perturbation theory for moiré states in twisted bilayer phosphorene
,”
J. Phys.: Condens. Matter
32
,
234001
(
2020
).
28.
W.
Li
,
M.
Chen
,
E.
Rabani
,
R.
Baer
, and
D.
Neuhauser
, “
Stochastic embedding DFT: Theory and application to p-nitroaniline in water
,”
J. Chem. Phys.
151
,
174115
(
2019
).
29.
T. T.
Tran
,
K.
Bray
,
M. J.
Ford
,
M.
Toth
, and
I.
Aharonovich
, “
Quantum emission from hexagonal boron nitride monolayers
,”
Nat. Nanotechnol.
11
,
37
41
(
2016
).
30.
A. L.
Exarhos
,
D. A.
Hopper
,
R. R.
Grote
,
A.
Alkauskas
, and
L. C.
Bassett
, “
Optical signatures of quantum emitters in suspended hexagonal boron nitride
,”
ACS Nano
11
,
3328
3336
(
2017
).
31.
C.
Li
,
Z.-Q.
Xu
,
N.
Mendelson
,
M.
Kianinia
,
M.
Toth
, and
I.
Aharonovich
, “
Purification of single-photon emission from hBN using post-processing treatments
,”
Nanophotonics
8
,
2049
2055
(
2019
).
32.
N.
Mendelson
,
Z.-Q.
Xu
,
T. T.
Tran
,
M.
Kianinia
,
J.
Scott
,
C.
Bradac
,
I.
Aharonovich
, and
M.
Toth
, “
Engineering and tuning of quantum emitters in few-layer hexagonal boron nitride
,”
ACS Nano
13
,
3132
3140
(
2019
).
33.
E.
Maggio
and
G.
Kresse
, “
GW vertex corrected calculations for molecular systems
,”
J. Chem. Theory Comput.
13
,
4765
(
2017
).
34.
M.
Hellgren
,
N.
Colonna
, and
S.
de Gironcoli
, “
Beyond the random phase approximation with a local exchange vertex
,”
Phys. Rev. A
98
,
045117
(
2018
).
35.
A. M.
Lewis
and
T. C.
Berkelbach
, “
Vertex corrections to the polarizability do not improve the GW approximation for the ionization potential of molecules
,”
J. Chem. Theory Comput.
15
,
2925
2932
(
2019
).
36.
R.
Baer
and
D.
Neuhauser
, “
Real-time linear response for time-dependent density-functional theory
,”
J. Chem. Phys.
121
,
9803
9807
(
2004
).
37.
Y.
Gao
,
D.
Neuhauser
,
R.
Baer
, and
E.
Rabani
, “
Sublinear scaling for time-dependent stochastic density functional theory
,”
J. Chem. Phys.
142
,
034106
(
2015
).
38.
D.
Neuhauser
,
E.
Rabani
,
Y.
Cytter
, and
R.
Baer
, “
Stochastic optimally tuned range-separated hybrid density functional theory
,”
J. Phys. Chem. A
120
,
3071
3078
(
2016
).
39.
E.
Rabani
,
R.
Baer
, and
D.
Neuhauser
, “
Time-dependent stochastic Bethe-Salpeter approach
,”
Phys. Rev. B
91
,
235302
(
2015
).
40.
S.
Baroni
,
S.
de Gironcoli
, and
A.
Dal Corso
, “
Phonons and related crystal properties from density-functional perturbation theory
,”
Rev. Mod. Phys.
73
,
515
562
(
2001
).
41.
D.
Neuhauser
and
R.
Baer
, “
Efficient linear-response method circumventing the exchange-correlation kernel: Theory for molecular conductance under finite bias
,”
J. Chem. Phys.
123
,
204105
(
2005
).
42.
G. H.
Wannier
, “
The structure of electronic excitation levels in insulating crystals
,”
Phys. Rev.
52
,
191
197
(
1937
).
43.
N.
Marzari
,
A. A.
Mostofi
,
J. R.
Yates
,
I.
Souza
, and
D.
Vanderbilt
, “
Maximally localized Wannier functions: Theory and applications
,”
Rev. Mod. Phys.
84
,
1419
1475
(
2012
).
44.
E. Ö.
Jónsson
,
S.
Lehtola
,
M.
Puska
, and
H.
Jónsson
, “
Theory and applications of generalized Pipek–Mezey Wannier functions
,”
J. Chem. Theory Comput.
13
,
460
474
(
2017
).
45.
N.
Troullier
and
J. L.
Martins
, “
Efficient pseudopotentials for plane-wave calculations
,”
Phys. Rev. B
43
,
1993
(
1991
).
46.
J. P.
Perdew
and
Y.
Wang
, “
Accurate and simple analytic representation of the electron-gas correlation energy
,”
Phys. Rev. B
45
,
13244
13249
(
1992
).
47.
C. A.
Rozzi
,
D.
Varsano
,
A.
Marini
,
E. K. U.
Gross
, and
A.
Rubio
, “
Exact Coulomb cutoff technique for supercell calculations
,”
Phys. Rev. B
73
,
205119
(
2006
).
48.
P.
Giannozzi
,
O.
Andreussi
,
T.
Brumme
,
O.
Bunau
,
M.
Buongiorno Nardelli
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
M.
Cococcioni
,
N.
Colonna
,
I.
Carnimeo
,
A.
Dal 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.
Otero-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
, “
Advanced capabilities for materials modelling with quantum espresso
,”
J. Phys.: Condens. Matter
29
,
465901
(
2017
).
49.
A.
Tkatchenko
and
M.
Scheffler
, “
Accurate molecular van der Waals interactions from ground-state electron density and free-atom reference data
,”
Phys. Rev. Lett.
102
,
073005
(
2009
).
50.
M.
Otani
and
O.
Sugino
, “
First-principles calculations of charged surfaces and interfaces: A plane-wave nonrepeated slab approach
,”
Phys. Rev. B
73
,
115407
(
2006
).
51.
S.
Park
,
C.
Park
, and
G.
Kim
, “
Interlayer coupling enhancement in graphene/hexagonal boron nitride heterostructures by intercalated defects or vacancies
,”
J. Chem. Phys.
140
,
134706
(
2014
).
52.
A.
Sajid
,
J. R.
Reimers
, and
M. J.
Ford
, “
Defect states in hexagonal boron nitride: Assignments of observed properties and prediction of properties relevant to quantum computation
,”
Phys. Rev. B
97
,
064101
(
2018
).
53.
C.
Attaccalite
,
M.
Bockstedte
,
A.
Marini
,
A.
Rubio
, and
L.
Wirtz
, “
Coupling of excitons and defect states in boron-nitride nanostructures
,”
Phys. Rev. B
83
,
144115
(
2011
).
54.
B.
Huang
and
H.
Lee
, “
Defect and impurity properties of hexagonal boron nitride: A first-principles calculation
,”
Phys. Rev. B
86
,
245406
(
2012
).
55.
N. L.
McDougall
,
J. G.
Partridge
,
R. J.
Nicholls
,
S. P.
Russo
, and
D. G.
McCulloch
, “
Influence of point defects on the near edge structure of hexagonal boron nitride
,”
Phys. Rev. B
96
,
144106
(
2017
).
56.
M. E.
Levinshtein
,
S. L.
Rumyantsev
, and
M. S.
Shur
,
Properties of Advanced Semiconductor Materials: GaN, AIN, InN, BN, SiC, SiGe
(
Wiley & Sons
,
2001
).
57.
F.
Fuchs
,
J.
Furthmüller
,
F.
Bechstedt
,
M.
Shishkin
, and
G.
Kresse
, “
Quasiparticle band structure based on a generalized Kohn-Sham scheme
,”
Phys. Rev. B
76
,
115109
(
2007
).
58.
F.
Hüser
,
T.
Olsen
, and
K. S.
Thygesen
, “
Quasiparticle GW calculations for solids, molecules, and two-dimensional materials
,”
Phys. Rev. B
87
,
235132
(
2013
).
59.
V.
Popescu
and
A.
Zunger
, “
Extracting e versus pk effective band structure from supercell calculations on alloys and impurities
,”
Phys. Rev. B
85
,
085201
(
2012
).
60.
H.
Huang
,
F.
Zheng
,
P.
Zhang
,
J.
Wu
,
B.-L.
Gu
, and
W.
Duan
, “
A general group theoretical method to unfold band structures and its application
,”
New J. Phys.
16
,
033034
(
2014
).
61.
P. V. C.
Medeiros
,
S.
Stafström
, and
J.
Björk
, “
Effects of extrinsic and intrinsic perturbations on the electronic structure of graphene: Retaining an effective primitive cell band structure by band unfolding
,”
Phys. Rev. B
89
,
041407
(
2014
).
62.
P.
Cudazzo
,
L.
Sponza
,
C.
Giorgetti
,
L.
Reining
,
F.
Sottile
, and
M.
Gatti
, “
Exciton band structure in two-dimensional materials
,”
Phys. Rev. Lett.
116
,
066803
(
2016
).
63.
F.
Paleari
,
T.
Galvani
,
H.
Amara
,
F.
Ducastelle
,
A.
Molina-Sánchez
, and
L.
Wirtz
, “
Excitons in few-layer hexagonal boron nitride: Davydov splitting and surface localization
,”
2D Materials
5
,
045017
(
2018
).
64.
Z.-Q.
Xu
,
N.
Mendelson
,
J. A.
Scott
,
C.
Li
,
I. H.
Abidi
,
H.
Liu
,
Z.
Luo
,
I.
Aharonovich
, and
M.
Toth
, “
Charge and energy transfer of quantum emitters in 2D heterostructures
,”
2D Mater.
7
,
031001
(
2020
).
65.
O.
Salihoglu
,
N.
Kakenov
,
O.
Balci
,
S.
Balci
, and
C.
Kocabas
, “
Graphene as a reversible and spectrally selective fluorescence quencher
,”
Sci. Rep.
6
,
33911
(
2016
).
66.
J.
Lee
,
W.
Bao
,
L.
Ju
,
P. J.
Schuck
,
F.
Wang
, and
A.
Weber-Bargioni
, “
Switching individual quantum dot emission through electrically controlling resonant energy transfer to graphene
,”
Nano Lett.
14
,
7115
7119
(
2014
).
67.
C.
Bjelkevig
,
Z.
Mi
,
J.
Xiao
,
P. A.
Dowben
,
L.
Wang
,
W.-N.
Mei
, and
J. A.
Kelber
, “
Electronic structure of a graphene/hexagonal-BN heterostructure grown on Ru(0001) by chemical vapor deposition and atomic layer deposition: Extrinsically doped graphene
,”
J. Phys.: Condens. Matter
22
,
302002
(
2010
).
68.

The distributions of the wavefunctions are governed by the local external and Hartree potentials (since the orbitals correspond to eigenstates of the mean-field Hamiltonian H0).

69.
J. B.
Neaton
,
M. S.
Hybertsen
, and
S. G.
Louie
, “
Renormalization of molecular electronic levels at metal-molecule interfaces
,”
Phys. Rev. Lett.
97
,
216405
(
2006
).
70.

Low-intensity points of the graphene bands around −3 eV correspond to artifacts of the projection technique discussed in the main text. These points correspond to the hBN states that exceed the middle of the interlayer distance and spuriously appear in the graphene projection. The energy position of the low-intensity points is further smeared out by applying the QP correction.

71.
P.
Mori-Sánchez
,
A. J.
Cohen
, and
W.
Yang
, “
Localization and delocalization errors in density functional theory and implications for band-gap prediction
,”
Phys. Rev. Lett.
100
,
146401
(
2008
).
72.
A. J.
Cohen
,
P.
Mori-Sánchez
, and
W.
Yang
, “
Insights into current limitations of density functional theory
,”
Science
321
,
792
794
(
2008
).
73.
Y.
Liang
and
L.
Yang
, “
Carrier plasmon induced nonlinear band gap renormalization in two-dimensional semiconductors
,”
Phys. Rev. Lett.
114
,
063001
(
2015
).
74.
M. M.
Ugeda
,
A. J.
Bradley
,
S.-F.
Shi
,
F. H.
da Jornada
,
Y.
Zhang
,
D. Y.
Qiu
,
W.
Ruan
,
S.-K.
Mo
,
Z.
Hussain
,
Z.-X.
Shen
,
F.
Wang
,
S. G.
Louie
, and
M. F.
Crommie
, “
Giant bandgap renormalization and excitonic effects in a monolayer transition metal dichalcogenide semiconductor
,”
Nat. Mater.
13
,
1091
1095
(
2014
).
75.
Y.
Cho
and
T. C.
Berkelbach
, “
Environmentally sensitive theory of electronic and optical transitions in atomically thin semiconductors
,”
Phys. Rev. B
97
,
041409
(
2018
).
76.
J.
Towns
,
T.
Cockerill
,
M.
Dahan
,
I.
Foster
,
K.
Gaither
,
A.
Grimshaw
,
V.
Hazlewood
,
S.
Lathrop
,
D.
Lifka
,
G. D.
Peterson
,
R.
Roskies
,
J. R.
Scott
, and
N.
Wilkins-Diehr
, “
XSEDE: Accelerating scientific discovery
,”
Comput. Sci. Eng.
16
,
62
74
(
2014
).