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

thanks: These authors contributed equally to this work.thanks: These authors contributed equally to this work.

Unified Investigation of Rapid Hall Coefficient Changes in Cuprates: Pseudogap and Fermi Surface Influences

Yingze Su School of Physics, Peking University, Beijing 100871, China    Hui Li School of Physics, Zhejiang University, Hangzhou 310058, China    Huaqing Huang [email protected] School of Physics, Peking University, Beijing 100871, China    Dingping Li [email protected] School of Physics, Peking University, Beijing 100871, China
Abstract

High-TcT_{c} cuprates are characterized by strong spin fluctuations, which give rise to antiferromagnetic and pseudogap phases and may be key to the high superconducting critical temperatures observed in these materials. Experimental studies have revealed significant changes in the Hall coefficient RHR_{H} across these phases, a phenomenon closely related to both spin fluctuations and changes in the Fermi surface morphology. Using the perturbation correction to Gaussian approximation (PCGA), we investigate the two-dimensional(2D) square-lattice single-band Hubbard model and obtain the self-energy with a finite imaginary part due to scattering. We calculate the density dependence of the Hall number nH=1/(qRH)n_{H}=1/(qR_{H}). For small hole (or electron) doping pp (or xx), our numerical results show that nHn_{H} transitions from pp to 1+p1+p for hole-doped systems, and from x-x to 1x1-x for electron-doped systems—both in agreement with experimental findings. Furthermore, we discuss the correlation between phase boundaries and the observed peculiar changes in the Hall number.

preprint: APS/123-QED

I Introduction

The pseudogap of cuprates is one of the most intensely debated phenomena in the studies for high-temperature superconductors [Vedeneev_2021]. Some researchers consider it a distinct phase of matter [varma_1997, kivelson_1998], while others view it as a precursor of an ordered phase [Sedrakyan_2010, Wu_2019]. The complexity of the pseudogap arises from its connection to various orders, such as charge/spin/pair density waves [Hucker_2014, Wang_2015], strip order [comin_2015], (short-range) antiferromagnetism [Baledent_2011], electronic nematicity [Cyr_2015], etc. These competing orders may provide crucial insights into the emergence of high-temperature superconductivity.

In the transition from the “normal” phase (including strange metal) to the pseudogap region, significant changes in Hall number have been observed experimentally [badoux_change_2016, greene_strange_2020]. In the electron-doped region, as doping xx decreases, the Hall number initially drops from 1x1-x to a significantly negative value, before rising back to x-x. This anomalous phenomenon is believed to be related to Fermi surface reconstruction [greene_strange_2020]. In the hole-doped region, the dependence of the Hall number on the doping pp shows a drop from 1+p1+p to pp. Utilizing the Yang-Rice-Zhang ansatz with pseudogap, Storey pointed out that this drop is associated with Ne´el\mathrm{N\acute{e}el} antiferromagnetism [storey_hall_2016]. Further investigations explained the emergence of short-range order by the Hubbard model’s spiral phase. However, it still requires numerous adjustable parameters, one of which is the scattering rate [Eberlein_2016, mitscherling_longitudinal_2018].

In this study, we employ the single-band Hubbard model to calculate the Hall coefficient across a wide range of densities. Firstly, we attribute the dominant contribution to scattering rates to Coulomb repulsive interactions, and calculate finite scattering rates in two-loop perturbation correction. Subsequently, response theory is employed to obtain the conductivity and Hall number [Voruganti_conductivity_1992]. Moreover, for both hole-doped and electron-doped scenarios, this unified description achieves behavior consistent with experimental observations under moderate doping levels. Notably, short-range correlations and Fermi surface both play important roles in the abrupt changes observed in the Hall number.

This paper is structured as follows. In section II, we introduce the antiferromagnetic phase and its perturbation correction for the Hubbard model. Subsequently, we provide an overview of the Fermi surface evolution with respect to density, and outline the calculation method for the Hall number. In section LABEL:chaps:results, we analyze the phase regions based on the Fermi surface, investigate the variation of the scattering rate with momentum points, and explore the changes in the Hall number across a broad electron filling range. Finally, section LABEL:chaps:conclusion concludes our study.

II formalism

II.1 Model and Methodology

We start with a single-band Hubbard model on the 2D square lattice. The Hamiltonian is

H^=i,jσ(tijc^iσc^jσ+h.c.)+Uin^in^i,\hat{H}=-\sum_{\langle i,j\rangle}\sum_{\sigma}\left(t_{ij}\hat{c}^{\dagger}_{i\sigma}\hat{c}_{j\sigma}+h.c.\right)+U\sum_{i}\hat{n}_{i\uparrow}\hat{n}_{i\downarrow}, (1)

where tijt_{ij} denotes the hopping amplitude between lattice site ii and jj and UU represents the strength of the on-site Coulomb interaction. c^,c^\hat{c}^{\dagger},\hat{c} are electron creation and annihilation operators, respectively. The index σ=,\sigma=\uparrow,\downarrow describes the spin orientation. The hopping amplitudes have been investigated theoretically using density-functional-theory calculations [andersen_lda_nodate, pavarini_band-structure_2001, Imada_ab_initio_2022], and experimentally by angle-resolved photoemission spectroscopy (ARPES) [Shen_electronic_1995, Borisenko_joys_2000]. Nearest, second nearest, and third nearest neighbor hopping t,t,t′′t,t^{\prime},t^{\prime\prime} are usually taken into account and the former is adopted as the unit of energy. For La2xSrxCuO4\mathrm{La}_{2-x}\mathrm{Sr}_{x}\mathrm{CuO}_{4} (LSCO\mathrm{LSCO}), t/t0.1t^{\prime}/t\sim-0.1 while for YBa2Cu3Oy\mathrm{YBa}_{2}\mathrm{Cu}_{3}\mathrm{O}_{y} (YBCO\mathrm{YBCO}) and Bi2Sr2CaCu2O8+δ\mathrm{Bi_{2}Sr_{2}CaCu_{2}O_{8+\delta}}, t/t0.3t^{\prime}/t\sim-0.3. In our calculations, we take t=0.25t,t′′=0.10tt^{\prime}=-0.25t,t^{\prime\prime}=0.10t and a pretty strong U=6tU=6t, near the typical values presented in literature [mitscherling_longitudinal_2018, Huang_strange_2019]. Taking the lattice constant aa as the unit of length, and the energy dispersion without UU (i.e., U=0U=0) is

ϵ(k)=\displaystyle\epsilon(\vec{k})= 2t(coskx+cosky)+4|t|coskxcosky\displaystyle-2t\left(\cos k_{x}+\cos k_{y}\right)+4|t^{\prime}|\cos k_{x}\cos k_{y} (2)
2t′′(cos2kx+cos2ky).\displaystyle-2t^{\prime\prime}\left(\cos 2k_{x}+\cos 2k_{y}\right).

The Hartree-Fock approximation shows that the Hubbard model exhibits a rich phase diagram [Igoshev_Incommensurate_2010, Laughlin_Hartree-Fock_2014, Scholle_comprehensive_2023], considering both magnetic and charge fluctuations. Given the prominence of antiferromagnetic (AF) fluctuations near half-filling, one may expect the magnetic phase transition would capture its key features. For simplicity, we employ a paramagnetic(PM)-AF transition. Both phases could be described in a unified form [Kao_Unified_2023]. When adjacent sites are considered distinctly, the system exhibits a 2×2\sqrt{2}\times\sqrt{2} super-lattice with lattice vectors a1=(1,1)\vec{a}_{1}=(1,1) and a2=(1,1)\vec{a}_{2}=(-1,1). Within each super-cell, there exist two sites, denoted as rA=(0,0)\vec{r}^{A}=(0,0) and rB=(0,1)\vec{r}^{B}=(0,1). The first Brillouin zone is folded accordingly as shown in Figure 1. Kinetic term of Hamiltonian written in basis {ψA,ψB,ψA,ψB}\{\psi^{A}_{\uparrow},\psi^{B}_{\uparrow},\psi^{A}_{\downarrow},\psi^{B}_{\downarrow}\} is blocked diagonal.

H^0,(k)=[ϵ1(k)ϵ2(k)e+iφ(k)ϵ2(k)eiφ(k)ϵ1(k)],\hat{H}_{0,\uparrow}(\vec{k})=\begin{bmatrix}\epsilon_{1}(\vec{k})&\epsilon_{2}(\vec{k})e^{+\mathrm{i}\varphi(\vec{k})}\\ \epsilon_{2}(\vec{k})e^{-\mathrm{i}\varphi(\vec{k})}&\epsilon_{1}(\vec{k})\end{bmatrix}, (3)
ϵ1(k)=4|t|coskxcosky2t′′(cos2kx+cos2ky),ϵ2(k)=2t(coskx+cosky),φ(k)=krB=ky.\begin{gathered}\epsilon_{1}(\vec{k})=4|t^{\prime}|\cos k_{x}\cos k_{y}-2t^{\prime\prime}\left(\cos 2k_{x}+\cos 2k_{y}\right),\\ \epsilon_{2}(\vec{k})=-2t\left(\cos k_{x}+\cos k_{y}\right),\\ \varphi(\vec{k})=\vec{k}\cdot\vec{r}^{B}=k_{y}.\end{gathered} (4)

The Hartree-Fock Green’s functions can be expressed by the local densities nA,nA,nB,nBn^{A}_{\uparrow},n^{A}_{\downarrow},n^{B}_{\uparrow},n^{B}_{\downarrow} from Dyson-Schwinger equation,

G1(k)=[iωn+μUnB00iωn+μUnA]H0,(k).G^{-1}_{\upuparrows}(k)=\begin{bmatrix}\mathrm{i}\omega_{n}+\mu-Un^{B}_{\uparrow}&0\\ 0&\mathrm{i}\omega_{n}+\mu-Un^{A}_{\uparrow}\end{bmatrix}-H_{0,\uparrow}(\vec{k}). (5)

For the AF phase, they are related to each other nA=nB,nA=nBn^{A}_{\uparrow}=n^{B}_{\downarrow},n^{A}_{\downarrow}=n^{B}_{\uparrow}; for the PM phase, they are all equal. Self-consistent equations are derived by Matsubara sum 1βNkGAA(k)=nA\frac{1}{\beta N}\sum_{k}G^{AA}_{\upuparrows}(k)=n^{A}_{\uparrow}, here k=(iωn,k)k=(\mathrm{i}\omega_{n},\vec{k}), and β,N\beta,N denotes inverse of temperature and number of sites, respectively.

Refer to caption
Figure 1: The supercell and corresponding 1st Brillouin Zone(BZ). Upper: adjacent sites are divided into AA parts(red) and BB parts(blue). The lattice vectors are a1=(1,1),a2=(1,1)\vec{a}_{1}=(1,1),\vec{a}_{2}=(-1,1). Lower: the area of Folded BZ(dashed line) is half of the original BZ(solid line). The reciprocal lattice vectors are k1=(π,π),k2=(π,π)\vec{k}_{1}=(\pi,\pi),\vec{k}_{2}=(-\pi,\pi)

Scattering due to interaction plays a pivotal role in cuprates transport property, and it leads us to compute the self-energy with a finite imaginary part at the two-loop level. Making use of standard perturbation theory, we get the perturbation correction to Gaussian approximation (PCGA) [Kao_Unified_2023]. Self-energy Σ(k)\Sigma_{\upuparrows}(k) turns out to be

Σab(k)=U2β2N2q1,q2Gba(q1+q2k)Gab(q1)Gab(q2),\Sigma^{ab}_{\upuparrows}(k)=-\frac{U^{2}}{\beta^{2}N^{2}}\sum_{q_{1},q_{2}}G^{ba}_{\downdownarrows}(q_{1}+q_{2}-k)G^{ab}_{\upuparrows}(q_{1})G^{ab}_{\downdownarrows}(q_{2}), (6)

where a,b{A,B}a,b\in\{A,B\}. We would see that the real-frequency self-energy Σ(ω,k)\Sigma_{\upuparrows}(\omega,\vec{k}) do have a spatially varying imaginary part.

II.2 Fermi Surface

Refer to caption
Figure 2: (a)Sketch for FS under Hartree-Fock Approximation, and (0,0)(0,0) at the center as Figure 1. Blue shaded area corresponds to electron pockets while orange corresponds to hole pockets. A1A6A_{1}\sim A_{6}: PM phase, from an extremely low density to pretty high density. Around A2A_{2} FS transition from convex to concave, and around A4A_{4} FS open up. B1B3B_{1}\sim B_{3}: AF phase with finite magnetic moments and electron density increases gradually. (b) shows phase boundary over density in different temperatures, which decrease from top to bottom. Appearance of B1,B3B_{1},B_{3} depends on parameters we choose.

Various methods exist for studying phase transitions in electronic systems, one of which involves detecting the evolution of the Fermi surface (FS). [storey_hall_2016, armitage_angle-resolved_2003, matsui_evolution_2007, louis_remarkable_2019]. Figure 2a shows the evolution of the FS as a function of density, which is determined using the Hartree-Fock approximation.

For the PM phase, in the dilute limit, the FS is nearly a circle, which is convex everywhere. As the electron density grows, the FS around nodal point (q,q)(q,q) gradually becomes concave (A1A_{1} to A3A_{3} in Fig. 2a). The point that FS transition from convex to concave is regarded as Fermi liquid starting to break down [FS_Gindikin_2024]. When anti-nodal point goes to (π,0)(\pi,0) the topological of FS changes (A4A_{4} in Fig. 2a).

The AF phase is more complicated since there are finite magnetic moments mm. Near half-filling, electron pockets and hole pockets coexist as B2B_{2} in Fig. 2a. With hole/electron doping increases, electron/hole pockets become smaller, and even disappear under some parameter regions. These topological changes also indicate phase boundaries. The boundary ends in high temperature which does not allow the AF phase in corresponding doped levels (between mid- and lower- temperature in Fig. 2b).

II.3 Current and Response functions

To calculate the Hall number, we need to evaluate longitudinal and Hall conductivity respectively. For mean-field theory with magnetic order and uniform scattering rate, there are some systematic researches [Eberlein_2016, mitscherling_longitudinal_2018]. We would derive the formula in a similar way without a uniform scattering rate assumption. Under electromagnetic field, we need to apply Peierls substitution [peierls_zur_1933, wannier_dynamics_1962, Vucice_electrical_2021] to Hamiltonian Equation (1)

tij[A]=tijexp(i(Ai+Aj)(rirj)/2),t_{ij}[\vec{A}]=t_{ij}\exp\left(\mathrm{i}\left(\vec{A}_{i}+\vec{A}_{j}\right)\cdot(\vec{r}_{i}-\vec{r}_{j})/2\right), (7)

where A\vec{A} is the vector potential of the electromagnetic field. The current operator j^α\hat{j}^{\alpha} and corresponding bare vertex γα\gamma^{\alpha} satisfy

j^α(r)=δH^[A]δAα(r)|A0=r1,r2σcσ(r1)γα(r1,r2;r)cσ(r2).\hat{j}^{\alpha}(\vec{r})=\left.\frac{\delta\hat{H}[\vec{A}]}{\delta A_{\alpha}(\vec{r})}\right|_{\vec{A}\equiv 0}=\sum_{\vec{r}_{1},\vec{r}_{2}}\sum_{\sigma}c^{\dagger}_{\sigma}(\vec{r}_{1})\gamma^{\alpha}(\vec{r}_{1},\vec{r}_{2};\vec{r})c_{\sigma}(\vec{r}_{2}). (8)

To connect (Hall) conductivity and current-correlation functions, we need to extend linear response theory up to the second order. The coefficients Π\Pis can be expressed by either correlation functions or conductivity, serving as a bridge between them.

jα(τ,r)=\displaystyle\langle j^{\alpha}(\tau,\vec{r})\rangle= 0βdτΠab(τ,r;τ)AbE(τ)\displaystyle\int_{0}^{\beta}\mathrm{d}\tau^{\prime}\ \Pi^{ab}(\tau,\vec{r};\tau^{\prime})A^{E}_{b}(\tau^{\prime}) (9)
+\displaystyle+ 0βdτrΠabc(τ,r;τ,r)AbE(τ)AcB(r)\displaystyle\int_{0}^{\beta}\mathrm{d}\tau^{\prime}\sum_{\vec{r}^{\prime}}\Pi^{abc}(\tau,\vec{r}^{\prime};\tau^{\prime},\vec{r}^{\prime})A^{E}_{b}(\tau^{\prime})A^{B}_{c}(\vec{r}^{\prime})
+\displaystyle+ higher order response,\displaystyle\text{higher order response},

where a,b,c{x,y,z}a,b,c\in\{x,y,z\} denotes spatial components, τ\tau is the imaginary time and A(τ,r)=AE(τ)+AB(r)\vec{A}(\tau,\vec{r})=\vec{A}^{E}(\tau)+\vec{A}^{B}(\vec{r}). Suppose E(τ)=E(τ)e^x,B(r)=B(r)e^z\vec{E}(\tau)=E(\tau)\hat{e}_{x},\vec{B}(\vec{r})=B(\vec{r})\hat{e}_{z}, we could select appropriate gauge to make AB\vec{A}^{B} in yy-direction.

{jx(τ,r)=0βdτΠxx(τ,r;τ)AE(τ)jy(τ,r)=0βdτrΠyxy(τ,r;τ,r)AE(τ)AB(r)E(τ)=τAE(τ),B(r)=xAB(r)\left\{\begin{aligned} \langle j^{x}(\tau,\vec{r})\rangle=&\int_{0}^{\beta}\mathrm{d}\tau\ \Pi^{xx}(\tau,\vec{r};\tau^{\prime})A^{E}(\tau^{\prime})\\ \langle j^{y}(\tau,\vec{r})\rangle=&\int_{0}^{\beta}\mathrm{d}\tau\sum_{\vec{r}^{\prime}}\Pi^{yxy}(\tau,\vec{r};\tau^{\prime},\vec{r}^{\prime})A^{E}(\tau^{\prime})A^{B}(\vec{r}^{\prime})\\ E(\tau)=&-\partial_{\tau}A^{E}(\tau),\quad B(\vec{r})=\partial_{x}A^{B}(\vec{r})\end{aligned}\right. (10)

The longitudinal and Hall conductivity σ,σHall\sigma,\sigma_{\mathrm{Hall}} are described by the current response to homogeneous static electric and magnetic fields,

jx=σE,jy=σHallBE.j^{x}=\sigma E,\quad j^{y}=\sigma_{\mathrm{Hall}}BE. (11)

And we get Π\Pis expressed by frequency-dependent conductivity [Voruganti_conductivity_1992]

σ(ω)=1iω(Πxx(ω,k=0)(ω0)),σHall(ω)=1ωkx(Πyxy(ω,k=0)(ω0)).\begin{gathered}\sigma(\omega)=\frac{1}{\mathrm{i}\omega}\left(\Pi^{xx}(\omega,\vec{k}=\vec{0})-(\omega\to 0)\right),\\ \sigma_{\mathrm{Hall}}(\omega)=\frac{1}{\omega}\frac{\partial}{\partial k_{x}}\left(\Pi^{yxy}(\omega,\vec{k}=\vec{0})-(\omega\to 0)\right).\end{gathered} (12)

By definition, static limit σ=σ(ω=0),σHall=σHall(ω=0)\sigma=\sigma(\omega=0),\sigma_{\mathrm{Hall}}=\sigma_{\mathrm{Hall}}(\omega=0). Our supercells with A,BA,B sites break translation invariance, so Π\Pis in momentum space should be symmetrized as

Πxx(k)=1NreikrΠxx(r),\displaystyle\Pi^{xx}(\vec{k})=\frac{1}{N}\sum_{\vec{r}}e^{-\mathrm{i}\vec{k}\cdot\vec{r}}\Pi^{xx}(\vec{r}), (13)
Πyxy(k)=1N2r1,r2eik(rr2)Πxx(r1,r2).\displaystyle\Pi^{yxy}(\vec{k})=\frac{1}{N^{2}}\sum_{\vec{r}_{1},\vec{r}_{2}}e^{-\mathrm{i}\vec{k}\cdot(\vec{r}-\vec{r}_{2})}\Pi^{xx}(\vec{r}_{1},\vec{r}_{2}). (14)

On the other hand, in the language of path-integral, the expectation value of current can be expressed by action including electromagnetic field S[ψ,ψ;A]S[\psi^{*},\psi;\vec{A}]

jα(τ,r)=δδAα(τ,r)lnD[ψ,ψ]eS[ψ,ψ;A].\langle j^{\alpha}(\tau,\vec{r})\rangle=-\frac{\delta}{\delta A_{\alpha}(\tau,\vec{r})}\ln\int D[\psi^{*},\psi]\ e^{-S[\psi^{*},\psi;\vec{A}]}. (15)

Furthermore, take the derivative of A\vec{A} by Eq. (9). Since conductivity only relate to derivative of Πxx\Pi^{xx} and Πyxy\Pi^{yxy} as Eq. (12), any terms with δ(τ,τ)\delta(\tau,\tau^{\prime}) or δ(rx,rx)\delta(r_{x},r^{\prime}_{x}) could be dropped.

Πxx(τ,r;τ)=\displaystyle\Pi^{xx}(\tau,\vec{r};\tau^{\prime})= rjx(τ,r)jx(τ,r)c+δjx(τ,r)δAE(τ)rjx(τ,r)jx(τ,r)c,\displaystyle\sum_{\vec{r}^{\prime}}\left\langle j^{x}(\tau,\vec{r})j^{x}(\tau^{\prime},\vec{r}^{\prime})\right\rangle_{c}+\left\langle\frac{\delta j^{x}(\tau,\vec{r})}{\delta A^{E}(\tau^{\prime})}\right\rangle\to\sum_{\vec{r}^{\prime}}\left\langle j^{x}(\tau,\vec{r})j^{x}(\tau^{\prime},\vec{r}^{\prime})\right\rangle_{c}, (16)
Πyxy(τ,r;τ,r)=\displaystyle\Pi^{yxy}(\tau,\vec{r};\tau^{\prime},\vec{r}^{\prime})= dτ′′r′′jy(τ,r)jx(τ,r′′)jy(τ′′,r)c+dτ′′δjy(τ,r)δAE(τ)jy(τ′′,r)c\displaystyle\int\mathrm{d}\tau^{\prime\prime}\sum_{\vec{r}^{\prime\prime}}\ \left\langle j^{y}(\tau,\vec{r})j^{x}(\tau^{\prime},\vec{r}^{\prime\prime})j^{y}(\tau^{\prime\prime},\vec{r}^{\prime})\right\rangle_{c}+\int\mathrm{d}\tau^{\prime\prime}\ \left\langle\frac{\delta j^{y}(\tau,\vec{r})}{\delta A^{E}(\tau^{\prime})}j^{y}(\tau^{\prime\prime},\vec{r}^{\prime})\right\rangle_{c}
+\displaystyle+ rδjy(τ,r)δAB(r)jx(τ,r′′)c+r′′jy(τ,r)δjx(τ,r′′)δAB(r)c+δ2jy(τ,r)δAB(r)δAE(τ)\displaystyle\sum_{\vec{r}^{\prime}}\left\langle\frac{\delta j^{y}(\tau,\vec{r})}{\delta A^{B}(\vec{r}^{\prime})}j^{x}(\tau^{\prime},\vec{r}^{\prime\prime})\right\rangle_{c}+\sum_{\vec{r}^{\prime\prime}}\ \left\langle j^{y}(\tau,\vec{r})\frac{\delta j^{x}(\tau^{\prime},\vec{r}^{\prime\prime})}{\delta A^{B}(\vec{r}^{\prime})}\right\rangle_{c}+\left\langle\frac{\delta^{2}j^{y}(\tau,\vec{r})}{\delta A^{B}(\vec{r}^{\prime})\delta A^{E}(\tau^{\prime})}\right\rangle
\displaystyle\to dτ′′r′′jy(τ,r)jx(τ,r′′)jy(τ′′,r)c+r′′jy(τ,r)δjx(τ,r′′)δAB(r)c.\displaystyle\int\mathrm{d}\tau^{\prime\prime}\sum_{\vec{r}^{\prime\prime}}\ \left\langle j^{y}(\tau,\vec{r})j^{x}(\tau^{\prime},\vec{r}^{\prime\prime})j^{y}(\tau^{\prime\prime},\vec{r}^{\prime})\right\rangle_{c}+\sum_{\vec{r}^{\prime\prime}}\ \left\langle j^{y}(\tau,\vec{r})\frac{\delta j^{x}(\tau^{\prime},\vec{r}^{\prime\prime})}{\delta A^{B}(\vec{r}^{\prime})}\right\rangle_{c}.

Here c\langle\cdot\rangle_{c} denotes connected correlation functions. The lowest order of Πyxy\Pi^{yxy} consists 33 Feynman diagrams as Fig. LABEL:fig:3. Bare vertices are derivatives of Hamiltonian like Eq. (8)

δH^[A]δAα(r)|A0=r1,r2σcσ(r1)γα(r1,r2;r)cσ(r2),δ2H^[A]δAα(r)δAβ(r)|A0=r1,r2σcσ(r1)γαβ(r1,r2;r)cσ(r2).\begin{gathered}\left.\frac{\delta\hat{H}[\vec{A}]}{\delta A_{\alpha}(\vec{r})}\right|_{\vec{A}\equiv 0}=\sum_{\vec{r}_{1},\vec{r}_{2}}\sum_{\sigma}c^{\dagger}_{\sigma}(\vec{r}_{1})\gamma^{\alpha}(\vec{r}_{1},\vec{r}_{2};\vec{r})c_{\sigma}(\vec{r}_{2}),\\ \left.\frac{\delta^{2}\hat{H}[\vec{A}]}{\delta A_{\alpha}(\vec{r})\delta A_{\beta}(\vec{r})}\right|_{\vec{A}\equiv 0}=\sum_{\vec{r}_{1},\vec{r}_{2}}\sum_{\sigma}c^{\dagger}_{\sigma}(\vec{r}_{1})\gamma^{\alpha\beta}(\vec{r}_{1},\vec{r}_{2};\vec{r})c_{\sigma}(\vec{r}_{2}).\end{gathered} (17)