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

Solving the Bethe-Salpeter Equation in Real Frequencies at Finite Temperature

I.S. Tupitsyn Department of Physics, University of Massachusetts, Amherst, MA 01003, USA    N.V. Prokof’ev Department of Physics, University of Massachusetts, Amherst, MA 01003, USA
Abstract

Self-consistent Hartree-Fock approximation combined with solutions of the Bethe-Salpeter equation offers a powerful tool for studies of strong correlation effects arising in condensed matter models, nuclear physics, quantum field theories, and real materials. The standard finite-temperature approach would be to first solve the problem in the Matsubara representation and then apply numerical analytic continuation to the real-frequency axis to link theoretical results with experimental probes, but this ill-conditioned procedure often distorts important spectral features even for very accurate imaginary-frequency data. We demonstrate that the ladder-type finite-temperature Bethe-Salpeter equation in the Hartree-Fock basis for the 3-point vertex function and, ultimately, system’s polarization can be accurately solved directly on the real frequency axis using the diagrammatic Monte Carlo technique and series resummation. We illustrate the method by applying it to the homogeneous electron gas model and demonstrate how multiple scattering events renormalize Landau damping.

I Introduction

The Bethe-Salpeter equation (BSE) was initially introduced to study two-particle Green’s function and the corresponding bound states in nuclear physics [1]. Later on it was formulated for studies of optical spectra in solids [2, 3] and molecules [4, 5]. In material science/quantum chemistry methods one often starts by establishing the Hartree-Fock (HF) basis and proceeds with the electronic structure calculations and solutions of the BSE for two-particle Green’s function and related properties [6, 7, 8]. In strongly correlated models and quantum field theories BSE is used to calculate the effect of vertex corrections [9, 10]. This is especially important in the case of polarization when looking at charge and spin responses because neglecting vertex corrections leads to wrong results [11].

The standard finite-temperature approach is to first solve the problem in the Matsubara domain [12] and then apply a numeric analytic continuation procedure to obtain real frequency results for comparison with experimental observations. However, even if original data is known with high accuracy, the final step distorts, or even fails to resolve, important spectral features unless one imposes constrains on how the result is supposed to behave [13]. Until recently, this problem standing in the way of accurate theoretical descriptions of experimentally relevant observables was considered unavoidable.

The breakthrough development done in the context of diagrammatic Monte Carlo (diagMC) for Fermi-Hubbard model in Refs. [14, 15, 16] (see also Refs. [17, 18]) was that one can automatically express real-frequency answers for an arbitrary diagram using the so-called algorithmic Matsubara integration (AMI) protocol and, thus, completely eliminate the need for numeric analytic continuation. It works best for expansions in terms of “bare” (non-interacting) Green’s functions and frequency-independent interactions. However, for strongly correlated regimes, when diagrammatic expansions in terms of original potentials diverge—a typical example would be the fundamental homogeneous electron gas (HEG), or jellium, model—expansions should be performed in terms of “dressed” propagators and screened potentials, which include some interactions effects non-perturbatively. The corresponding generalization of the AMI protocol was described in Ref. [11] but its implementation for high-order diagrammatic simulations poses significant technical challenges.

Current applications of the AMI protocol for jellium are based on expansion in terms of the Yukawa potential, Y=4πe2/(q2+κ2)Y=4\pi e^{2}/(q^{2}+\kappa^{2}), with optimally chosen screening momentum κ\kappa and self-consistent HF single particle propagators, GG, see Refs. [19, 20, 21] and the text below Eq. (1). In brief, the AMI technique lists all diagrams of order NN for an observable of interest, eliminates all sums over internal Matsubara frequencies for every listed diagram by an exact analytic transformation, and stochastically samples the remaining momentum integrals. However, it remains biased because all poles are regularized by adding small but finite imaginary part to the external frequency, ΩΩ+iη\Omega\to\Omega+i\eta. The exact answer is recovered by taking the η+0\eta\to+0 limit which requires extremely long computation times for high-order diagrams [19].

In this work, we apply the AMI and series resummation techniques for solving the ladder-type finite-temperature BSE equation for the 33-point vertex function Γ(3)\Gamma^{(3)} convoluted with two Green’s functions to obtain system’s polarization directly on the real-frequency axis. We take the crucial new step of implementing the limit of η+0\eta\to+0 explicitly; the same procedure may be used for any model with instantaneous interaction. The resulting scheme is not only numerically exact, it is also very efficient and allows one to reach expansion orders high-enough for performing reliable resummations of divergent series. The same expansion orders cannot be reached if simulations are performed with appropriately small finite-η\eta regularization.

For illustration, we compute the polarization of the HEG in the HF basis, and compare it to the random-phase-approximation (RPA) predictions (also in the HF basis) to see the effects of multiple particle-hole scattering processes. In particular, we find that the Landau damping coefficient is significantly increased relative to the RPA prediction.

II Model and simulation setup

Hamiltonian of the HEG is defined by

H=iki22m+i<je2|𝐫i𝐫j|μN,H=\sum_{i}\frac{k_{i}^{2}}{2m}+\sum_{i<j}\frac{e^{2}}{|\mathbf{r}_{i}-\mathbf{r}_{j}|}-\mu N, (1)

with mm the electron mass, μ\mu the chemical potential, and ee the electron charge. We use the inverse Fermi momentum, 1/kF1/k_{F}, and Fermi energy, εF=kF2/2m\varepsilon_{F}=k_{F}^{2}/2m, as units of length and energy, respectively; the definition of the Coulomb parameter rsr_{s} in terms of the particle number density, ρ=kF3/3π2\rho=k_{F}^{3}/3\pi^{2}, and Bohr radius, aB=1/me2a_{B}=1/me^{2}, is standard: 4πrs3/3=1/ρaB34\pi r_{s}^{3}/3=1/\rho a_{B}^{3}.

Refer to caption


Figure 1: Flowchart of the technique. Calculation starts from the self-consistent HF solution for the proper self-energy, ΣHF\Sigma_{HF} (a), with the chemical potential, μ\mu, constantly adjusted to keep the number density, nn, fixed. Converged solution, GG, is subsequently used for computing the system’s polarization, Π\Pi (b), from the series of ladder diagrams on the real frequency axis. The series can be written in terms of the BSE (c) for the dressed vertex function, Γ(3)\Gamma^{(3)}, or as Taylor series expansion in the number of interaction lines. Monte Carlo simulations in the η0\eta\to 0 limit are used to compute partial ii-th order contributions (d). In all diagrams solid and dashed lines correspond to GG and Yukawa potential, respectively.

For a meaningful diagrammatic expansion one has to work with screened interactions and counter-terms. In this work we follow the setup formulated in Ref. [20] and expand in the Yukawa potential on top of the self-consistent HF solution for the Green’s function: G1=G01ΣHF[G]G^{-1}=G^{-1}_{0}-\Sigma_{HF}[G], where G0G_{0} is the bare Green’s function and ΣHF\Sigma_{HF} is the Fock proper self-energy shown in Fig. 1(a) (Hartree diagram is absent by charge neutrality). In the converged HF solution, G=(iωsk2/2mΣHF(𝐤)+μ)1(iωsϵk+μ)1G=(i\omega_{s}-k^{2}/2m-\Sigma_{HF}(\mathbf{k})+\mu)^{-1}\equiv(i\omega_{s}-\epsilon_{k}+\mu)^{-1}, the chemical potential is also self-consistently adjusted to keep the electron density fixed, see upper part of Fig. 1. By incorporating Fock diagrams into GG one simplifies the expansion by omitting all diagrams with Fock-type self-energy insertions.

Formally, not only the screening momentum κ\kappa in the Yukawa potential, but the entire potential itself is arbitrary as long as it does not feature divergent behavior at q=0q=0. Any difference between the Coulomb, V(q)=4πe2/q2V(q)=4\pi e^{2}/q^{2}, and expansion potentials is taken care of by a counter-term C(q)=1/V(q)1/Y(q)C(q)=1/V(q)-1/Y(q) which keeps the field theoretical representation of the model exact [22, 23, 20]. For example, the physical screened interaction is given by 1/W=1/Y[ΠC(q)]1/VΠ1/W=1/Y-[\Pi-C(q)]\equiv 1/V-\Pi where Π\Pi is the system polarization. Ultimately, final results are supposed to show little dependence on the choice of Y(q)Y(q). Flexibility in reformulating the diagrammatic expansion can, and should be, used for optimization of series convergence properties. This is precisely how recent work reported in Refs. [20, 21] achieved unprecedented accuracy for single-particle and static linear response properties of HEG.

Refer to caption Refer to caption

Figure 2: Real and imaginary parts of the polarization Π0\Pi_{0} in the HF basis (see Fig.1(b) with Γ(3)=1\Gamma^{(3)}=1) at rs=2r_{s}=2 for Q/kF=0.09844Q/k_{F}=0.09844 and Ω/εF=0.2\Omega/\varepsilon_{F}=0.2 as a function of η\eta (the smallest value of η/εF\eta/\varepsilon_{F} here is 10510^{-5}).

In this work we apply the above field-theoretical setup to explore how the polarization is modified by multiple scattering of the particle-hole pair at finite temperature. Since the calculation is performed directly on the real frequency axis, our results can be immediately used to obtain the Landau damping parameter, see Ref. [24], or describe the so-call charge loss function, ImW\mathrm{Im}W. The set of diagrams involved is presented in Figs. 1 (b), (c), and (d) where the ladder-diagram series for polarization are first expressed in terms of the BSE for the dressed three-point vertex function Γ(3)\Gamma^{(3)}, see 1 (c), which is then convoluted with two Green’s functions, see 1 (b). As noticed in Ref. [11], if the calculation is restricted to have no more than one ladder rung, the result for Π=Π0+Π1\Pi=\Pi_{0}+\Pi_{1}, where Πi\Pi_{i} is the contribution of the diagram with ii rungs Fig. 1(d), may violate causality. Also, an observation that Π1\Pi_{1} is not small compared to Π0\Pi_{0} raises the question of importance of contributions from high-order ladder diagrams. Summation of the entire ladder series is supposed to solve both problems.

Before proceeding, let us elaborate on the importance of considering very small values of η\eta when using it for regularization of poles. Even the lowest order diagram demonstrates significant finite-η\eta correction at low temperature, see Fig. 2. At T/εF=0.02T/\varepsilon_{F}=0.02 the finite-η\eta effect in Π0\Pi_{0} is as large as 20%20\% for η/εF=0.005\eta/\varepsilon_{F}=0.005. One may attempt to extrapolate results towards η=0\eta=0, but this procedure is not guaranteed to work for all higher-order contributions because the η\eta-dependence is not necessarily monotonic.

BSE in the η+0\eta\to+0 limit. As discussed above, physical results may not depend on the choice of Y(q)Y(q) for setting up the expansion. This, in particular, implies that good approximations should produce results nearly independent of the choice of the screening momentum in Y(q)Y(q); any significant dependence would be a clear sign of the approximation bias. This is precisely why Π=Π0+Π1\Pi=\Pi_{0}+\Pi_{1} needs to be supplemented with additional diagrams which are supposed to restore causality and change the function shape. In what follows we demonstrate that summation of the entire ladder series produces results with relatively weak dependence on κ2\kappa^{2} when the latter is changed by more than a factor of two in the vicinity of the Thomas-Fermi momentum κTF=6πρe2/εF\kappa_{TF}=6\pi\rho e^{2}/\varepsilon_{F}. In this work we perform all calculations for rs=2r_{s}=2 when κTF1.15kF\kappa_{TF}\approx 1.15k_{F}.

The ii-th order contribution to the ladder diagram series for Π\Pi, see Figs. 1(b) and 1(c), is given by (in what follows we treat Ω\Omega and 𝐐{\bf Q} as external parameters and do not mention them explicitly as function variables)

Πi=2(j=1i+1d𝐩j(2π)3)j=1i+1𝐩jj=1iY(𝐩j+1𝐩j),\Pi_{i}=-2\left(\prod_{j=1}^{i+1}\int\frac{d{\bf p}_{j}}{(2\pi)^{3}}\right)\prod_{j=1}^{i+1}{\cal F}_{{\bf p}_{j}}\prod_{j=1}^{i}Y({\bf p}_{j+1}-{\bf p}_{j})\,, (2)

where

𝐩=n𝐩+𝐐n𝐩Ωϵ𝐩+𝐐+ϵ𝐩+iη,{\cal F}_{{\bf p}}=\frac{n_{{\bf p}+{\bf Q}}-n_{\mathbf{p}}}{\Omega-\epsilon_{{\bf p}+{\bf Q}}+\epsilon_{\mathbf{p}}+i\eta}, (3)

with n𝐩=(eϵpμ+1)1n_{\mathbf{p}}=(e^{\epsilon_{p}-\mu}+1)^{-1}. While this setup is immediately suitable for Monte Carlo simulations with small finite η\eta, it quickly becomes too expensive as ii increases. The reason is the severe sign problem coming from Monte Carlo sampling of the principle part integrals around poles of FF functions.

This problem can be eliminated by taking the limit of η+0\eta\to+0 explicitly with the help of the limη+01x+iη=iπδ(x)+𝒫1x\lim_{\eta\to+0}\frac{1}{x+i\eta}=-i\pi\delta(x)+{\cal P}\frac{1}{x} formula were 𝒫{\cal P} stands for the principal part value. In practice it means that integrals featuring simple poles can be transformed identically as

ab𝑑xh(x)Ωg(x)+iη=ab𝑑x[α(x,x0)+iγ(x,x0)],\int_{a}^{b}dx\frac{h(x)}{\Omega-g(x)+i\eta}=\int_{a}^{b}dx\left[\alpha(x,x_{0})+i\gamma(x,x_{0})\right]\,, (4)

where x0x_{0} is the solution of the g(x0)=Ωg(x_{0})=\Omega equation and

γ(x,x0)\displaystyle\gamma(x,x_{0}) =\displaystyle= πh(x0)|g(x0)|N(x,x0)ifx0(a,b),\displaystyle-\pi\frac{h(x_{0})}{|g^{\prime}(x_{0})|}N(x,x_{0})\;\;\mbox{if}\;x_{0}\in(a,b),
γ(x,x0)\displaystyle\gamma(x,x_{0}) =\displaystyle= 0otherwise,\displaystyle 0\qquad\qquad\qquad\qquad\;\;\;\mbox{otherwise}, (5)
α(x,x0)\displaystyle\alpha(x,x_{0}) =\displaystyle= 12[h(x)Ωg(x)+h(xr)Ωg(xr)]if(x0,xr)(a,b)\displaystyle\frac{1}{2}\left[\frac{h(x)}{\Omega-g(x)}+\frac{h(x_{r})}{\Omega-g(x_{r})}\right]\;\;\mbox{if}\;(x_{0},x_{r})\in(a,b)
α(x,x0)\displaystyle\alpha(x,x_{0}) =\displaystyle= h(x)Ωg(x)otherwise,\displaystyle\frac{h(x)}{\Omega-g(x)}\qquad\qquad\qquad\qquad\;\;\mbox{otherwise}, (6)

with xr=2x0xx_{r}=2x_{0}-x. Here N(x,x0)N(x,x_{0}) is an arbitrary function normalized to unity on the (a,b)(a,b) interval–it may explicitly depend on x0x_{0}. By construction, the α+iγ\alpha+i\gamma function is not singular at the pole location. In our case, we choose variables x=cos𝐩𝐐^x=\cos\widehat{{\bf p}{\bf Q}} to regularize poles. For a given spherically symmetric dispersion relation ϵ𝐩\epsilon_{\mathbf{p}} the location of the pole x0(p,Q,Ω)x_{0}(p,Q,\Omega) as a function of pp is tabulated on a fine grid at the start of the simulation and subsequently is treated as a known function.

However, taking the limit η+0\eta\to+0 also involves “splitting” the contribution of the diagram with variable xx into up to three contributions from variables xx, x0x_{0}, and xrx_{r} if (x0,xr)(a,b)(x_{0},x_{r})\in(a,b). As the ladder diagram order increases, one now faces the problem of computing up to 3i+13^{i+1} contributions for some multi-dimensional momentum points {𝐩j}\{{\bf p}_{j}\} because momentum variables are linked by potentials YY, see (2). This remaining problem is solved by “decoupling” momentum integrals using Fourier transforms of Yukawa potentials

Y(𝐩j+1𝐩j)=𝑑𝐫jY(rj)ei(𝐩j+1𝐩j)𝐫j,Y({\bf p}_{j+1}-{\bf p}_{j})=\int d{\bf r}_{j}Y(r_{j})\,e^{i({\bf p}_{j+1}-{\bf p}_{j})\cdot{\bf r}_{j}}\,, (7)

where Y(r)=(e2/r)eκrY(r)=(e^{2}/r)\,e^{-\kappa r}. Now

Πi=2(j=1i𝑑𝐫jY(rj))j=1i+1d𝐩j(2π)3~𝐩j,𝐫j1𝐫j\Pi_{i}=-2\left(\prod_{j=1}^{i}\int d{\bf r}_{j}Y(r_{j})\right)\prod_{j=1}^{i+1}\int\frac{d{\bf p}_{j}}{(2\pi)^{3}}\tilde{\cal F}_{{\bf p}_{j},{\bf r}_{j-1}-{\bf r}_{j}} (8)

with

~𝐩,𝐑=ei𝐑𝐩n𝐩+𝐐n𝐩Ωϵ𝐩+𝐐+ϵ𝐩+iη.\tilde{\cal F}_{{\bf p},{\bf R}}=e^{i{\bf R}\cdot{\bf p}}\frac{n_{{\bf p}+{\bf Q}}-n_{\mathbf{p}}}{\Omega-\epsilon_{{\bf p}+{\bf Q}}+\epsilon_{\mathbf{p}}+i\eta}. (9)

Here we assume that 𝐫0=𝐫i+1=0{\bf r}_{0}={\bf r}_{i+1}=0 for convenience of compact notations. According to (8), for a given set of integration variables the integrand is a product of functions which depend only on one momentum variable. When the η+0\eta\to+0 limit is taken, each ~\tilde{\cal F} function may be split into up to three contributions, but the computational cost of evaluating the total contribution is now reduced to 3(i+1)3(i+1) instead of 3i+13^{i+1}. Additional real space integrals do not pose any technical or efficiency problems for Monte Carlo simulations, which all can be performed on a single laptop.

III Series Convergence

The convergence of the ladder series strongly depends on R=Ω/QvFR=\Omega/Qv_{F} ratio with vFv_{F} the Fermi velocity, see panels (a) and (b) in Fig. 3 where we plot partial contributions to polarization Π\Pi from ladder diagrams in Γ(3)\Gamma^{(3)}. For large and small RR we observe excellent convergence for all values of κ\kappa considered in this work. However, in the R1R\sim 1 region the series start diverging for κ1\kappa\leq 1, and the problem is getting progressively more severe as κ\kappa is reduced. At Q/kF0.1Q/k_{F}\approx 0.1, the series for κ=0.8\kappa=0.8 are already visibly beyond the convergence radius. The dimensionless parameter controlling series convergence can be estimated from

gρF4πe2κ2κTF22κ2g\sim\rho_{F}\frac{4\pi e^{2}}{\kappa^{2}}\sim\frac{\kappa_{TF}^{2}}{2\kappa^{2}} (10)

where ρF\rho_{F} is the Fermi-energy density of states per spin. When the screening momentum is increased to κ=1.2\kappa=1.2, the series start to converge for any value of RR (similarly to what is observed for larger momentum and higher temperature in panels (a) and (b) in Fig. 4).

Refer to caption Refer to caption

Refer to caption Refer to caption

Figure 3: Partial ImΠi\textrm{Im}\Pi_{i} (a) and ReΠi\textrm{Re}\Pi_{i} (b) contributions to polarization from ladder diagrams with ii rungs for Q/kF=0.09844Q/k_{F}=0.09844, rs=2r_{s}=2, T/εF=0.02T/\varepsilon_{F}=0.02, and κ/kF=0.8\kappa/k_{F}=0.8. Results for the final answer after series resummation (see text) are shown in panels (c) and (d) for ImΠ\textrm{Im}\Pi and ReΠ\textrm{Re}\Pi, respectively. The thin dashed lines in panels (c) and (d) are the zeroth-order Π0\Pi_{0} results for κ/kF=0.8\kappa/k_{F}=0.8. Error bars are smaller than symbol sizes.

If we formally consider ladder diagrams as Taylor series expansion in powers of parameter ξ\xi with the physical result corresponding to ξ0=1\xi_{0}=1, see the lower-right block in Fig. 1, then series resummation can be performed with the help of the conformal map transformation, z=ξf(ξ)ξ=zs(z)m=1cmzmz=\xi f(\xi)\rightleftarrows\xi=zs(z)\equiv\sum_{m=1}^{\infty}c_{m}z^{m}. By substituting ξ(z)\xi(z) into the expression for Π\Pi in Fig. 1(e), one generates Taylor series in powers of zz with the physical result corresponding to z0=ξ0f(ξ0)z_{0}=\xi_{0}f(\xi_{0}). The map is designed to have z0z_{0} within the convergence radius so that the new Taylor series in powers of z0z_{0} converge. We find that all our data can be successfully resummed by using a simple-pole map, z=ξ/(ξ+ξp(R))z=\xi/(\xi+\xi_{p}(R)), with ξp(R=1)1\xi_{p}(R=1)\sim 1 and increasing away from the R=1R=1 point. The final results for three values of the Yukawa screening are shown in panels (c) and (d) in Figs. 3 and 4. Remarkably, despite substantial differences from the zeroth-order RPA-type results, there is extremely close agreement between the final curves for different values of κ\kappa (note that κ2\kappa^{2} was changed by more than a factor of two). Resummation of Taylor series, as opposed to self-consistent and skeleton sequences, does not lead to multi-valued solutions of the type observed in Ref. [25], not to mention that resummed and converges ”as is” curves end up on top of each other.

Refer to caption Refer to caption

Refer to caption Refer to caption

Figure 4: Partial ImΠi\textrm{Im}\Pi_{i} (a) and ReΠi\textrm{Re}\Pi_{i} (b) contributions to polarization from ladder diagrams with ii rungs for Q/kF=1.0Q/k_{F}=1.0, rs=2r_{s}=2, T/εF=0.1T/\varepsilon_{F}=0.1, and κ/kF=1.2\kappa/k_{F}=1.2. Results for the final answer after summing up the series are shown in panels (c) and (d) for ImΠ\textrm{Im}\Pi and ReΠ\textrm{Re}\Pi, respectively. The thin dashed lines in panels (c) and (d) are the zeroth-order Π0\Pi_{0} results for κ/kF=0.8\kappa/k_{F}=0.8. Error bars are smaller than symbol sizes.

At large momentum Q/kF=1Q/k_{F}=1 and higher temperature T/εFT/\varepsilon_{F}, the series converge well for all values of κ/kF\kappa/k_{F} considered and can be summed up “as is”. The final converged results are shown in panels (c) and (d) in Fig. 4. The dependence on the Yukawa screening is barely detectable in this parameter regime. While Figs. 3 and 4 present results for rs=2r_{s}=2, we did not see any notable differences in efficiency of calculations neither for rs=1r_{s}=1 nor for rs=4r_{s}=4, in agreement with Eq. (10) which predicts that for κκTF\kappa\sim\kappa_{TF} one has g1g\sim 1.

Within the RPA, the slope of ImΠΩ/Q\textrm{Im}\Pi\propto\Omega/Q dependence is determined by the electron dispersion near the Fermi surface. Multiple scattering changes this result by nearly 100%100\%, which has direct implications for accuracy of Landau damping (and other energy looses in metals) estimates based on the RPA-type calculations (for more details on Landau damping see Ref. [24]).

IV Conclusions

We presented an efficient technique for computing ladder-diagram series for polarization on the real frequency axis at finite temperature, or, equivalently, for solving the Bethe-Salpeter equation for the 33-point vertex function in the ladder approximation and convoluting it with two Green’s functions in the Hartree-Fock basis. Our method is numerically exact once all approximations at the level of diagrams involved are specified, i.e. it does not involve the ill-conditioned numerical analytic continuation, finite regularization of the poles, or momentum grids for evaluating multi-dimensional integrals.

Remarkably, final results are nearly independent of the Yukawa screening used to setup the expansion, indicating small bias coming from this auxiliary parameter. We find that multiple scattering of the particle-hole pair results in significant changes of the polarization dependence on the Ω/Q\Omega/Q ratio and, correspondingly, of the Landau damping. We expect that our scheme can be adapted for analogous simulations of realistic materials, and further advanced to include retardation effects for effective interaction potentials.

V Acknowledgements

This work was supported by the U.S. Department of Energy, Office of Science, Basic Energy Sciences, under Award DE-SC0023141.

References

  • [1] E.E. Salpeter and H.A. Bethe, Phys. Rev. 84, 1232 (1951).
  • [2] W. Hanke and L. J. Sham, Phys. Rev. Lett. 43, 387 (1979).
  • [3] L.X. Benedict, E.L. Shirley, and R. B. Bohn, Phys. Rev. Lett. 80, 4514 (1998).
  • [4] M. Rohlfing and S.G. Louie, Phys. Rev. Lett. 81, 2312 (1998).
  • [5] B. Baumeier, D. Andrienko, and M. Rohlfing, J. Chem. Theory Comput. 8, 2790 (2012).
  • [6] J. Li, M. Holzmann, I. Duchemin, X. Blase, and V. Olevano, Phys. Rev. Lett. 118, 163001 (2017).
  • [7] X. Leng, F. Jin, M. Wei, and Y. Ma, WIREs Comput. Mol. Sci. 6, 532, (2016).
  • [8] A. Förster and L. Fissher, J. Chem. Theory Comput. 18, 6779, (2022).
  • [9] G.D. Mahan, Many-particle physics, 3rd ed., Springer New York, NY (2013).
  • [10] L. Hedin, Phys. Rev. 139, A796 (1965).
  • [11] I.S. Tupitsyn, A.M. Tsvelik, R.M. Konik, and N.N. Prokof’ev, Phys. Rev. Lett. 127, 026403 (2021).
  • [12] T. Matsubara, Prog. Theor. Phys. 14, 351 (1955).
  • [13] O. Goulko, A.S. Mishchenko, L. Pollet, N. Prokof’ev, and B. Svistunov, Phys. Rev. B 95, 014102 (2017).
  • [14] A. Taheridehkordi, S.H. Curnoe, and J.P.F. LeBlanc, Phys. Rev. B 99, 035120 (2019).
  • [15] A. Taheridehkordi, S.H. Curnoe, and J.P.F. LeBlanc, Phys. Rev. B 101, 125109 (2020);
  • [16] A. Taheridehkordi, S.H. Curnoe, and J.P.F. LeBlanc, Phys. Rev. B 102, 045115 (2020).
  • [17] J. Vučičević and M. Ferrero, Phys. Rev. B 101, 075113 (2020).
  • [18] J. Vučičević, P. Stipsić, and M. Ferrero, Phys. Rev. Research 3, 023082 (2021).
  • [19] J.P.F. LeBlanc, K. Chen, K. Haule, N.V. Prokof’ev, and I.S. Tupitsyn, Phys. Rev. Lett. 129, 246401 (2022).
  • [20] K. Chen, and K. Haule, Nat. Commun. 10, 3725 (2019).
  • [21] K. Haule and K. Chen, Scientific Reports 12, 2294 (2022).
  • [22] R. Rossi, F. Werner, N. Prokof’ev, and B. Svistunov, Phys. Rev. B 93, 161102(R) (2016).
  • [23] A.J. Kim, N.V. Prokof’ev, B.V. Svistunov, and E. Kozik, Phys. Rev. Lett. 126, 257001 (2021).
  • [24] I.S. Tupitsyn and N.V. Prokof’ev, arXiv:2311.05611v1 (2023).
  • [25] E. Kozik, M. Ferrero, and A. Georges, Phys. Rev. Lett. 114, 156402 (2015).