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

TKNN formula for general lattice Hamiltonian in odd dimensions

Hidenori Fukaya
Department of Physics, Osaka University, Toyonaka, Osaka 560-0043, Japan.
[email protected]
  
Department of Physics, Osaka University, Toyonaka, Osaka 560-0043, Japan.
[email protected]
   Satoshi Yamaguchi
Department of Physics, Osaka University, Toyonaka, Osaka 560-0043, Japan˙[email protected]
   Xi Wu
Physics Department, Ariel University, Ariel 40700, Israel
[email protected]
Abstract:

Topological insulators in odd dimensions are characterized by topological numbers. We prove the well-known relation between the topological number given by the Chern character of the Berry curvature and the Chern-Simons level of the low energy effective action for a general class of Hamiltonians bilinear in the fermion with general U(1) gauge interactions including non-minimal couplings by an explicit calculation. A series of Ward-Takahashi identities are crucial to relate the Chern-Simons level to a winding number, which could then be directly reduced to Chern character of Berry curvature by carrying out the integral over the temporal momenta.

1 Introduction

Topological insulators in D=2n+1D=2n+1 dimensions are characterized either by the Chern character of the Berry connection from the eigenfunctions of the Hamiltonian in the valence band [1, 2] or by the coefficient of the effective action as a functino of an external U(1) gauge field, i.e., photon [3, 4, 5, 6]. These two characterization are known to be equivalent because they both arise from the current correlation functions and there are explicit proofs for various cases [7, 8]. In Ref. [9] a proof is given for a large class of models for general odd dimensions, where they consider the most general lattice action for arbitrary free kinetic term on the lattice which is then coupled to U(1) gauge field in a minimal way, i.e. with the gauge interaction in the form of

H(A)=m,nψmhmneiAmnψn+mψmψm,\displaystyle H(A)=\sum_{m,n}\psi^{\dagger}_{m}h_{mn}e^{iA_{mn}}\psi_{n}+\sum_{m}\psi_{m}^{\dagger}\psi_{m}, (1)

where m,nm,n are the lattice sites hmnh_{mn} are the hopping parameters and AmnA_{mn} is the line integral of the gauge field along the straight line connecting the sites mm and nn. The advantage of this class of Hamiltonian is that the contact interactions such as fermion-fermion-multi-photon vertices do not contribute to the final expression so that only a set of Feynman diagram which appear also in the continuum theory gives non-vanishing contributions. Of course, this type of gauge interaction is physically motivated since it is based on the famous method of ‘Peierls substitution’ [10]. However, in more general situation, the gauge interaction may not always be described by such a single straight Wilson-line. In such cases, one has to include the contribution of contact interaction vertices. In this proceedings, we report our recent study on the equivalence for generalized lattice Hamiltonian so as to include arbitrary non-minimal gauge interactions [11].

2 Gapped fermion system on the lattice

We consider a gapped fermion system (with a gap Δ\Delta) on a lattice with the following action in Euclidean space in D=2n+1D=2n+1 dimensions.

SE=𝑑trψ(t,r)[t+iA0+H(A)]ψ(t,r),\displaystyle S_{\rm E}=\int dt\sum_{\vec{r}}\psi^{\dagger}(t,\vec{r})\left[\frac{\partial}{\partial t}+iA_{0}+H(\vec{A})\right]\psi(t,\vec{r}), (2)

where r\vec{r} runs over the 2n2n dimensional spatial lattice points. We will set x0=tx^{0}=t in the followings. The Hamiltonian H(A)H(\vec{A}) is given by a summation over all the possible hoppings on the lattice which include gauge interactions with a smooth external U(1) gauge field Aμ=(A0,A)A_{\mu}=(A_{0},\vec{A}). The fermion fields ψ(t,r)\psi^{\dagger}(t,\vec{r}) and ψ(t,r)\psi(t,\vec{r}) give creation and annihilation operators of fermions after quantization. We assume that there are NvN_{v} bands and NcN_{c} bands below and above the fermi level, respectively. Therefore the fermion fields have Nv+NcN_{v}+N_{c} components.

Since the fermion system is gapped with a gap size Δ>0\Delta>0 , the effective gauge action obtained by integrating out fermions can be expanded in terms of gauge invariant local actions as Seff=kakSk(A)S_{\rm eff}=\sum_{k}a_{k}S_{k}(A). Here, Seff(A)S_{\rm eff}(A) is defined as eSeff=𝒟ψ𝒟ψeSEe^{S_{\rm eff}}=\int\mathcal{D}\psi\mathcal{D}\psi^{\dagger}e^{-S_{\rm E}}, and Sk(A)S_{k}(A) are the gauge invariant actions given by the local Lagrangian Lk(A)L_{k}(A) and aka_{k} are the coefficients. By dimensional analysis, if the Lagragian Lk(A)L_{k}(A) has a mass dimension dkd_{k} the coefficient aka_{k} is suppressed by the dk(2n+1)d_{k}-(2n+1)-powers in 1Δ\frac{1}{\Delta} or lattice spacing aa. Many of the Lagrangians are given in terms of gauge invariant field FμνF_{\mu\nu} Since we do not have the Lorentz-invariance on the lattice, the structure of the coefficients aka_{k} in the effective action heavily depend the geometry of the lattice. However, there is a very special parity-violating term called Chern-Simons action Scs(A)S_{\rm cs}(A) given by

Scs(A)=d2n+1xϵα0β1α1βnαnAα0β1Aα1βnAαn.\displaystyle S_{\rm cs}(A)=\int d^{2n+1}x~{}\epsilon_{\alpha_{0}\beta_{1}\alpha_{1}\cdots\beta_{n}\alpha_{n}}A_{\alpha_{0}}\partial_{\beta_{1}}A_{\alpha_{1}}\cdots\partial_{\beta_{n}}A_{\alpha_{n}}. (3)

This action is topological and always takes this form no matter what the geometry of the lattice is. Topological information of the fermion system is contained in the effective action through the coefficient ccsc_{\rm cs} as Seff(A)=iccsScs+S_{\rm eff}(A)=ic_{\rm cs}S_{\rm cs}+\cdots where ”\cdots ” stands for other gauge invariant terms . Here the gauge invariance of the action requires that the coefficient is quantized as

ccs=k(2π)n(n+1)!\displaystyle c_{\rm cs}=\frac{k}{(2\pi)^{n}(n+1)!} ,\displaystyle,~{}~{} k.\displaystyle k\in\mathbb{Z}. (4)

Since the Chern-Simons action is of the lowest dimension in the parity-violating sector, the coefficient ccsc_{\rm cs} can be obtained by the following quantity

ccs\displaystyle c_{\rm cs} =\displaystyle= (i)n+1ϵα0β1α1βnαn(n+1)!(2n+1)!((q1)β1)((qn)βn)i=1ndDxieiqαixiδn+1Seff(A)δAα0(x0)δAα1(x1)δAαn(xn)|qi=0.\displaystyle\frac{(-i)^{n+1}\epsilon_{\alpha_{0}\beta_{1}\alpha_{1}\cdots\beta_{n}\alpha_{n}}}{(n+1)!(2n+1)!}\left(\frac{\partial}{\partial(q_{1})_{\beta_{1}}}\right)\cdots\left(\frac{\partial}{\partial(q_{n})_{\beta_{n}}}\right)\left.\prod_{i=1}^{n}\int d^{D}x_{i}e^{iq_{\alpha_{i}}x_{i}}\frac{\delta^{n+1}S_{\rm eff}(A)}{\delta A_{\alpha_{0}}(x_{0})\delta A_{\alpha_{1}}(x_{1})\cdots\delta A_{\alpha_{n}}(x_{n})}\right|_{q_{i}=0}. (5)

The effective action can be given by the log of the fermion determinant as Seff(A)=Tr[ln(D0+H(A))]S_{\rm eff}(A)=\mbox{Tr}\left[\ln\left(D_{0}+H(A)\right)\right], where D0=x0+iA0D_{0}=\frac{\partial}{\partial x^{0}}+iA_{0}. Splitting the kinetic operator D0+H(A)D_{0}+H(A) into free part and interaction part as D0+H(A)=x0+H0Γ(AD_{0}+H(A)=\frac{\partial}{\partial x_{0}}+H_{0}-\Gamma(A), where H0H_{0} is the free fermion part defined as H0H(A)|A=0H_{0}\equiv\left.H(A)\right|_{A=0} and Γ(A)\Gamma(A) is the interaction part defined as Γ(A)iA0H(A)+H0\Gamma(A)\equiv-iA_{0}-H(A)+H_{0}. Expanding in Γ(A)\Gamma(A), we obtain

Seff(A)const.\displaystyle S_{\rm eff}(A)-\mbox{\rm const.} =\displaystyle= n=11nTr[(1x0+H0Γ(A))n]\displaystyle-\sum_{n=1}^{\infty}\frac{1}{n}\mbox{Tr}\left[\left(\frac{1}{\frac{\partial}{\partial x_{0}}+H_{0}}\Gamma(A)\right)^{n}\right] (6)

From Eq.(5), we find that the Chern-Simons coupling for D=2+1D=2+1 dimension is given by

ccs\displaystyle c_{\rm cs} =\displaystyle= (i)22!3!d3p(2π)3ϵα0β1α1(q1)β1\displaystyle-\frac{(-i)^{2}}{2!3!}\int\frac{d^{3}p}{(2\pi)^{3}}\epsilon_{\alpha_{0}\beta_{1}\alpha_{1}}\left(\frac{\partial}{\partial q_{1}}\right)_{\beta_{1}} (7)
{Tr[SF(p)Γ(2)[q1,α0;q1,α1;p]]+Tr[SF(pq1)Γ(1)[q1,α0;p]SF(p)Γ(1)[q1,α1;pq1]]}|q1=0,\displaystyle\left\{\mbox{Tr}\left[S_{F}(p)\Gamma^{(2)}[-q_{1},\alpha_{0};q_{1},\alpha_{1};p]\right]\right.+\left.\left.\mbox{Tr}\left[S_{F}(p-q_{1})\Gamma^{(1)}[-q_{1},\alpha_{0};p]S_{F}(p)\Gamma^{(1)}[q_{1},\alpha_{1};p-q_{1}]\right]\right\}\right|_{q_{1}=0},

where SF(p)S_{F}(p) is the fermion propagator 1ip0+H0(p)\frac{1}{ip_{0}+H_{0}(\vec{p})} and Γ(1)[q1,α1;p]\Gamma^{(1)}[q_{1},\alpha_{1};p] and Γ(2)[q1,α1;q2,α2;p]\Gamma^{(2)}[q_{1},\alpha_{1};q_{2},\alpha_{2};p] are fermion-fermion-photon and fermion-fermion-photon-photon vertices with in-coming fermion momentum pp and in-coming photon momenta qi(i=1,2)q_{i}~{}(i=1,2) with Lorentz index αi(i=1,2)\alpha_{i}~{}(i=1,2)

Γ(1)[q1,α1;p]\displaystyle\Gamma^{(1)}[q_{1},\alpha_{1};p] =\displaystyle= d2n+1x1eiq1x1d2n+1yeipyδΓ[A](x,y)δAα1(x1)|A=0,\displaystyle\int d^{2n+1}x_{1}~{}e^{iq_{1}\cdot x_{1}}\int d^{2n+1}y~{}e^{ip\cdot y}\left.\frac{\delta\Gamma[A](x,y)}{\delta A_{\alpha_{1}}(x_{1})}\right|_{A=0}, (8)
Γ(2)[q1,α1,q2;α2;p]\displaystyle\Gamma^{(2)}[q_{1},\alpha_{1},q_{2};\alpha_{2};p] =\displaystyle= i=12(d2n+1xieiqixi)d2n+1yeipyδ2Γ[A](x,y)δAα1(x1)δAα2(x2)|A=0,\displaystyle\prod_{i=1}^{2}\left(\int d^{2n+1}x_{i}~{}e^{iq_{i}\cdot x_{i}}\right)\int d^{2n+1}y~{}e^{ip\cdot y}\left.\frac{\delta^{2}\Gamma[A](x,y)}{\delta A_{\alpha_{1}}(x_{1})\delta A_{\alpha_{2}}(x_{2})}\right|_{A=0}, (9)

Note that the contributions with multi-photon vertices vanishes for the class of Hamiltonians with gauge interactions given by a single straight Wilson-line because the multi-photon vertices are symmetric under the interchange of Lorentz indices of photons. When contracted with the antisymmetric tensor, such contributions vanish. However, in general Hamiltonian we must consider these contributions.

In addition to the usual Ward-Takahashi identity Γ(1)[0,α;p]=\textcolorblueSF1(p)pα\Gamma^{(1)}[0,\alpha;p]=\textcolor{blue}{-}\frac{\partial S^{-1}_{F}(p)}{\partial p_{\alpha}}, we derive the Ward-Takahashi identities in Appendix  A as

2Γ(1)[k,μ;p]kνpλ|k=0=Γ(2)[k,μ;0,λ;p]kν|k=0=Γ(2)[0,λ;l,μ;p]lν|l=0.\displaystyle\left.\frac{\partial^{2}\Gamma^{(1)}[k,\mu;p]}{\partial k_{\nu}\partial p_{\lambda}}\right|_{k=0}=\left.\frac{\partial\Gamma^{(2)}[k,\mu;0,\lambda;p]}{\partial k_{\nu}}\right|_{k=0}=\left.\frac{\partial\Gamma^{(2)}[0,\lambda;l,\mu;p]}{\partial l_{\nu}}\right|_{l=0}. (10)

Using these identities, the integrand of Eq. (7) can be rewritten as

ϵα0β1α1Tr[pα0(2SF(p)Γ(1)[q,α1;p]qβ1)+SF(p)SF1(p)pα1SF(p)SF1(p)pβ1SF(p)SF1(p)pα0]|q=0.\displaystyle\epsilon_{\alpha_{0}\beta_{1}\alpha_{1}}\mbox{Tr}\left[\frac{\partial}{\partial_{p_{\alpha_{0}}}}\left(2S_{F}(p)\frac{\partial\Gamma^{(1)}[q,\alpha_{1};p]}{\partial q_{\beta_{1}}}\right)\left.+S_{F}(p)\frac{\partial S^{-1}_{F}(p)}{\partial p_{\alpha_{1}}}S_{F}(p)\frac{\partial S^{-1}_{F}(p)}{\partial p_{\beta_{1}}}S_{F}(p)\frac{\partial S^{-1}_{F}(p)}{\partial p_{\alpha_{0}}}\right]\right|_{q=0}. (11)

The first term is a total divergence which vanishes when we integrate over the momentum. Therefore, one finds that the Chern-Simons coupling is given by the winding number as

ccs\displaystyle c_{\rm cs} =\displaystyle= (i)2ϵα0β1α12!3!dp02πBZd2p(2π)2Tr[SF(p)SF1(p)pα0SF(p)SF1(p)pβ1SF(p)SF1(p)pα1].\displaystyle\frac{(-i)^{2}\epsilon_{\alpha_{0}\beta_{1}\alpha_{1}}}{2!3!}\int\frac{dp_{0}}{2\pi}\int_{\rm BZ}\frac{d^{2}p}{(2\pi)^{2}}\mbox{Tr}\left[S_{F}(p)\frac{\partial S_{F}^{-1}(p)}{\partial p_{\alpha_{0}}}S_{F}(p)\frac{\partial S^{-1}_{F}(p)}{\partial p_{\beta_{1}}}S_{F}(p)\frac{\partial S_{F}^{-1}(p)}{\partial p_{\alpha_{1}}}\right]. (12)

For the case for D=4+1D=4+1 dimension (n=2n=2), we also have the following Ward-Takahashi identitiy given in AppendixA

2Γ(3)[q,μ;r,ν;0,λ;p]qαrβ|q=r=0=3Γ(2)[q,μ;r,ν;p]qαrβpλ|q=r=0.\displaystyle\left.\frac{\partial^{2}\Gamma^{(3)}[q,\mu;r,\nu;0,\lambda;p]}{\partial q_{\alpha}\partial r_{\beta}}\right|_{q=r=0}=\left.\frac{\partial^{3}\Gamma^{(2)}[q,\mu;r,\nu;p]}{\partial q_{\alpha}\partial r_{\beta}\partial p_{\lambda}}\right|_{q=r=0}. (13)

with which one can show extra terms add up to total derivatives so we obtain

ccs\displaystyle c_{\rm cs} =\displaystyle= (i)323!5!d5p(2π)5ϵα0β1α1β2α2\displaystyle-\frac{(-i)^{3}\cdot 2}{3!5!}\int\frac{d^{5}p}{(2\pi)^{5}}\epsilon_{\alpha_{0}\beta_{1}\alpha_{1}\beta_{2}\alpha_{2}} (14)
Tr[SF(p)SF1(p)pα0SF(p)SF1(p)pβ1SF(p)SF1(p)pα1SF(p)SF1(p)pβ2SF(p)SF1(p)pα2].\displaystyle\mbox{Tr}\left[S_{F}(p)\frac{\partial S^{-1}_{F}(p)}{\partial p_{\alpha_{0}}}S_{F}(p)\frac{\partial S^{-1}_{F}(p)}{\partial p_{\beta_{1}}}S_{F}(p)\frac{\partial S^{-1}_{F}(p)}{\partial p_{\alpha_{1}}}S_{F}(p)\frac{\partial S^{-1}_{F}(p)}{\partial p_{\beta_{2}}}S_{F}(p)\frac{\partial S^{-1}_{F}(p)}{\partial p_{\alpha_{2}}}\right].

Therefore, Chern-Simons coupling is given by the winding number with fermion propagator also for D=4+1D=4+1 case.

3 Equivalence of winding number and chern number

We now show the equivalence of the Chern-Simons coupling given by the winding number expression and the Chern character given by the Berry connection for the energy eigenstates in the valence bands. The proof of this part is already given in Ref. [9], but since the proof is simple, we give it here for completeness. We give the calculation for arbitrary odd (D=2n+1D=2n+1) dimensions, even though we have shown that the Chern-Simons coupling ccsc_{\rm cs} can be written by the winding number using SFS_{F} only for D=2+1D=2+1 and D=4+1D=4+1 dimensions.

In order to simplify the notation, hereafter we abbreviate the derivative with respect to the momentum pμp_{\mu} as μpμ\partial_{\mu}\equiv\frac{\partial}{\partial p_{\mu}}. The result of the previous section for D=2+1D=2+1 and D=4+1D=4+1 can be unified to the following results: In the expression using the fermion propagator S(p)=1ip0+HS(p)=\frac{1}{ip^{0}+H} and inserting a complete set of energy eigenstates α|αα|\displaystyle{\sum_{\alpha}|\alpha\rangle\langle\alpha|}, where α\alpha is the label of energ, the Chern-Simons coupling ccsc_{\rm cs} is given as

ccs\displaystyle c_{\rm cs} =\displaystyle= n!(i)n+2(n+1)!(2n)!d2np(2π)2nα1,,α2nϵi1i2i2ndp02πα1|i1H|α2α2|i2H|α3α2n|i2nH|α1(ip0+Eα1)2(ip0+Eα2)(ip0+Eα2n),.\displaystyle\frac{n!(-i)^{n+2}}{(n+1)!(2n)!}\int\frac{d^{2n}p}{(2\pi)^{2n}}\sum_{\alpha_{1},\cdots,\alpha_{2n}}\epsilon^{i_{1}i_{2}\cdots i_{2n}}\int\frac{dp^{0}}{2\pi}\frac{\langle\alpha_{1}|\partial_{i_{1}}H|\alpha_{2}\rangle\langle\alpha_{2}|\partial_{i_{2}}H|\alpha_{3}\rangle\cdots\langle\alpha_{2n}|\partial_{i_{2n}}H|\alpha_{1}\rangle}{(ip^{0}+E_{\alpha_{1}})^{2}(ip^{0}+E_{\alpha_{2}})\cdots(ip^{0}+E_{\alpha_{2n}})},.

where i1,,i2ni_{1},\cdots,i_{2n} stand for the spatial indices.

All we have to do is to integrate over p0p^{0} using Cauchy’s theorem. Here, we use a trick to simplify the integration. It is easy to see that the expression Eqs. (11), (14) are invariant under continuous deformation of SFS_{F} (or HH) provided that the integrand remains to have no singularities. Therefore, under a continuous change of the Hamilitonian, the winding number remains unchanged from its original value as long as the enegry spectrum is kept gapped throughout the deformation. Now, the most general Hamiltonian with NvN_{v} valence bands and NcN_{c} conduction bands is expressed as

H(p)a=1NvEa(p)|a(p)a(p)|+b˙=1NcEb˙(p)|b˙(p)b˙(p)|,\displaystyle H(\vec{p})\equiv\sum_{a=1}^{N_{v}}E_{a}(\vec{p})|a(\vec{p})\rangle\langle a(\vec{p})|+\sum_{\dot{b}=1}^{N_{c}}E_{\dot{b}}(\vec{p})|\dot{b}(\vec{p})\rangle\langle\dot{b}(\vec{p})|, (16)

where |a(p)|a(\vec{p})\rangle labeled by aa is the energy eigenstate in the valence band with spatial momentum p\vec{p} and negative energy eigenvalue Ea(p)<0E_{a}(\vec{p})<0. The state |b˙(p)|\dot{b}(\vec{p})\rangle labeled by b˙\dot{b} is the energy eigenstate in the conduction band with spatial momentum p\vec{p} and positive energy eigenvalue Eb˙(p)>0E_{\dot{b}}(\vec{p})>0. One can continuously deform the Hamiltonian without hitting the singularity of S(p)S(p) ( i.e. keeping the system gapped ) so that all energy eigenvalues in the conduction bands and all energy eigenvalues in the valence bands are degenerate and momentum independent (i.e. flat band ) respectively. Then the deformed Hamiltonian HnewH_{\rm new} which gives the same winding number becomes

Hnew(p)=Eva=1Nv|a(p)a(p)|+Ecb˙=1Nc|b˙(p)b˙(p)|,\displaystyle H_{\rm new}(\vec{p})=E_{v}\sum_{a=1}^{N_{v}}|a(\vec{p})\rangle\langle a(\vec{p})|+E_{c}\sum_{\dot{b}=1}^{N_{c}}|\dot{b}(\vec{p})\rangle\langle\dot{b}(\vec{p})|, (17)

where Ev<0E_{v}<0, Ec>0E_{c}>0 are the momentum independent constant. Here the eigenstates are identical to those with the original Hamiltonian. Since there are only two poles, we can easily carrry our p0p^{0} integral to obtain

J=a1,,an=1Nva˙1,,a˙n=1Ncϵi1j1i2nj2n(2n)!(n!)2a1|i1a˙1a˙1|j1a2××an|ina˙na˙n|jna1.\displaystyle J=-\sum_{a_{1},\cdots,a_{n}=1}^{N_{v}}\sum_{\dot{a}_{1},\cdots,\dot{a}_{n}=1}^{N_{c}}\epsilon^{i_{1}j_{1}\cdots i_{2n}j_{2n}}\frac{(2n)!}{(n!)^{2}}\langle a_{1}|\partial_{i_{1}}\dot{a}_{1}\rangle\langle\dot{a}_{1}|\partial_{j_{1}}a_{2}\rangle\times\cdots\times\langle a_{n}|\partial_{i_{n}}\dot{a}_{n}\rangle\langle\dot{a}_{n}|\partial_{j_{n}}a_{1}\rangle.

Let us define the Berry connection using the negative energy eigenstates as

𝒜ab𝒜μabdxμ=ia|μbdxμia|db.\displaystyle\mathcal{A}^{ab}\equiv\mathcal{A}^{ab}_{\mu}dx^{\mu}=-i\langle a|\partial_{\mu}b\rangle dx^{\mu}\equiv-i\langle a|db\rangle. (18)

Then it is straightforward to show that the Berry curvature ab\mathcal{F}^{ab} is

ab\displaystyle\mathcal{F}^{ab} \displaystyle\equiv (d𝒜+i𝒜𝒜)ab=ic˙=1Nca|dc˙c˙|db\displaystyle\left(d\mathcal{A}+i\mathcal{A}\mathcal{A}\right)^{ab}=i\sum_{\dot{c}=1}^{N_{c}}\langle a|d\dot{c}\rangle\langle\dot{c}|db\rangle (19)

Using Eq. (19), the Chern-Simons coupling can be expressed as

ccs=(1)n(n+1)!(2π)nBZchn(𝒜),\displaystyle c_{\rm cs}=\frac{(-1)^{n}}{(n+1)!(2\pi)^{n}}\int_{BZ}\mbox{ch}_{n}(\mathcal{A}), (20)

where chn(𝒜)\mbox{ch}_{n}(\mathcal{A}) is the 2nd Chern character defined by chn(𝒜)=1n!1(2π)ntr(n)\mbox{ch}_{n}(\mathcal{A})=\frac{1}{n!}\frac{1}{(2\pi)^{n}}\mbox{tr}(\mathcal{F}^{n}) . Comparing this expression with Eq. (4) ccs=k(n+1)!(2π)nc_{\rm cs}=\frac{k}{(n+1)!(2\pi)^{n}},

k=(1)nBZchn(𝒜).\displaystyle k=(-1)^{n}\int_{BZ}\mbox{ch}_{n}(\mathcal{A}). (21)

We have shown the Chern-Simons level and the topological number in terms of the Berry connection is identical. On the other hand, while the Berry connection approach is limited to the free theory case, the effective theory approach can be applied also to interacting theories.

Acknowledgments.
This work is supported in part by the Japanese Grant-in-Aid for Scientific Research(Nos. 15K05054, 18H01216, 18H04484 and 18K03620).

Appendix A Ward-Takahashi identities

In this appendix, we derive various identities among vertex functions and the inverse fermion propagator obtained from gauge invariance, i.e. Ward-Takahashi identities. Finite difference operator which appears in the hopping term of the lattice fermion system can be expressed in terms of infinite series of derivatives. Therefore, we assume that the Hamiltonian can be expressed in terms of all sorts of fermion hopping terms connected by the Wilson-lines of arbitrary contours or superpositions of them. Then, the action can be formally expanded as

S=𝑑txn=0ψ(t,x)Mμ1μn(Dμ1Dμnψ)(t,x)\displaystyle S=\int dt\sum_{\vec{x}}\sum_{n=0}^{\infty}\psi^{\dagger}(t,\vec{x})M_{\mu_{1}\cdots\mu_{n}}(D_{\mu_{1}}\cdots D_{\mu_{n}}\psi)(t,\vec{x}) (22)

where summation over μ1,,μn\mu_{1},\cdots,\mu_{n} are implicit. Mμ1μnM_{\mu_{1}\cdots\mu_{n}} are some N×NN\times N matrix where N=Nc+NvN=N_{c}+N_{v} is the number of fermion degrees of freedom per site.

Expanding this action in terms of gauge fields and making Fourier transformations, one can obtain the formal expressions of the inverse propagator and the vertex functions in the momentum space. In the following, let us denote the inverse fermion propagator with momentum pp as SF1(p)S^{-1}_{F}(p) and the vertex functions with incoming fermion momentum pp and nn photons with incoming momentum kik_{i} and μi\mu_{i} components (i=1,,ni=1,\cdots,n) and outgoing fermion with momentum p+i=1nkip+\displaystyle{\sum_{i=1}^{n}k_{i}} as Γ(n)[k1,μ1;;kn,μn;p]\Gamma^{(n)}[k_{1},\mu_{1};\cdots;k_{n},\mu_{n};p]. Then the formal expression gives

SF1(p)=n=0Mμ1μni=1n(ipμi),\displaystyle S_{F}^{-1}(p)=\sum_{n=0}^{\infty}M_{\mu_{1}\cdots\mu_{n}}\prod_{i=1}^{n}\left(ip_{\mu_{i}}\right), (23)
Γ(1)[k,μ;p]=in=1a=1nMμ1μa1μμa+1μni=1a1(i(p+k)μi)i=a+1n(ipμi),\displaystyle\Gamma^{(1)}[k,\mu;p]=-i\sum_{n=1}^{\infty}\sum_{a=1}^{n}M_{\mu_{1}\cdots\mu_{a-1}\mu\mu_{a+1}\cdots\mu_{n}}\prod_{i=1}^{a-1}\left(i(p+k)_{\mu_{i}}\right)\prod_{i=a+1}^{n}\left(ip_{\mu_{i}}\right), (24)

Substituting Eqs. (23) (24), it is easy to show the usual Ward-Takahashi identity in QED holds.

Now, t is interesting to note that we could also obtain Ward-Takahashi identities for quantities involving higher order terms in photon momenta and multi-photon vertex functions. For example the two-photon vertex is given as

Γ(2)[k,μ;l,ν;p]\displaystyle\Gamma^{(2)}[k,\mu;l,\nu;p]
=\displaystyle= i2n=1a,b=1a<bnMμ1μa1μμa+1μb1νμb+1μni=1a1(i(p+k+l)μi)i=a+1b1(i(p+l)μi)i=b+1n(ipμi)\displaystyle-i^{2}\sum_{n=1}^{\infty}\sum_{\underset{a<b}{a,b=1}}^{n}M_{\mu_{1}\cdots\mu_{a-1}\mu\mu_{a+1}\cdots\mu_{b-1}\nu\mu_{b+1}\cdots\mu_{n}}\prod_{i=1}^{a-1}\left(i(p+k+l)_{\mu_{i}}\right)\prod_{i=a+1}^{b-1}\left(i(p+l)_{\mu_{i}}\right)\prod_{i=b+1}^{n}\left(ip_{\mu_{i}}\right)
\displaystyle- i2n=1a,b=1a<bnMμ1μa1νμa+1μb1μμb+1μni=1a1(i(p+k+l)μi)i=a+1b1(i(p+k)μi)i=b+1n(ipμi),\displaystyle i^{2}\sum_{n=1}^{\infty}\sum_{\underset{a<b}{a,b=1}}^{n}M_{\mu_{1}\cdots\mu_{a-1}\nu\mu_{a+1}\cdots\mu_{b-1}\mu\mu_{b+1}\cdots\mu_{n}}\prod_{i=1}^{a-1}\left(i(p+k+l)_{\mu_{i}}\right)\prod_{i=a+1}^{b-1}\left(i(p+k)_{\mu_{i}}\right)\prod_{i=b+1}^{n}\left(ip_{\mu_{i}}\right),

Using Eqs.(24) (LABEL:eq:Gamma^(2)) , we obtain the following identities:

2Γ(1)[k,μ;p]kνpλ|k=0=Γ(2)[k,μ;l,λ;p]kν|k,l=0=Γ(2)[k,λ;l,μ;p]lν|k,l=0\displaystyle\left.\frac{\partial^{2}\Gamma^{(1)}[k,\mu;p]}{\partial k_{\nu}\partial p_{\lambda}}\right|_{k=0}=\left.\frac{\partial\Gamma^{(2)}[k,\mu;l,\lambda;p]}{\partial k_{\nu}}\right|_{k,l=0}=\left.\frac{\partial\Gamma^{(2)}[k,\lambda;l,\mu;p]}{\partial l_{\nu}}\right|_{k,l=0} (26)

Carrying out similar calculations by simply differentiaing Γ(2)\Gamma^{(2)} and Γ(3)\Gamma^{(3)} , we can see that the following identity holds:

2Γ(3)[q,μ;r,ν;s,λ;p]qαrβ|q,r,s=0=3Γ(2)[q,μ;r,ν;p]qαrβpλ|q,r=0\displaystyle\left.\frac{\partial^{2}\Gamma^{(3)}[q,\mu;r,\nu;s,\lambda;p]}{\partial q_{\alpha}\partial r_{\beta}}\right|_{q,r,s=0}=\left.\frac{\partial^{3}\Gamma^{(2)}[q,\mu;r,\nu;p]}{\partial q_{\alpha}\partial r_{\beta}\partial p_{\lambda}}\right|_{q,r=0} (27)

References

  • [1] D. J. Thouless, M. Kohmoto, M. P. Nightingale and M. den Nijs, Phys.  Rev.  Lett. 49, 405 (1982)
  • [2] Qian Niu and D J Thouless and Yong-Shi Wu, Pys.  Rev. B31 3372(1985)
  • [3] K. Ishikawa, Phys. Rev. Lett.  53, 1615 (1984). ; Phys. Rev. D 31, 1432 (1985).
  • [4] K. Ishikawa and T. Matsuyama, Z. Phys. C 33, 41 (1986). ; Nucl. Phys. B 280, 523 (1987).
  • [5] H. So, Prog. Theor. Phys.  73, 528 (1985). ; Prog. Theor. Phys.  74, 585 (1985).
  • [6] M. F. L. Golterman, K. Jansen and D. B. Kaplan, Phys. Lett. B 301, 219 (1993) [hep-lat/9209003].
  • [7] E. Witten, Riv. Nuovo Cim.  39, no. 7, 313 (2016) [arXiv:1510.07698 [cond-mat.mes-hall]].
  • [8] M. A. Zubkov, Annals Phys.  373, 298 (2016) [arXiv:1603.03665 [cond-mat.mes-hall]].
  • [9] Qi, Xiao-Liang and Hughes, Taylor and Zhang, Shou-Cheng, Phys.  Rev. B78 195424 (2008)
  • [10] Peierls, R , Z. Phys. 80 763 (1933)
  • [11] H. Fukaya, T. Onogi, S. Yamaguchi and X. Wu, arXiv:1903.11852 [hep-th].