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

Geometric Criterion for Solvability of Lattice Spin Systems

Masahiro Ogura Yukawa Institute for Theoretical Physics, Kyoto University, Kyoto 606-8502, Japan    Yukihisa Imamura Yukawa Institute for Theoretical Physics, Kyoto University, Kyoto 606-8502, Japan    Naruhiko Kameyama Graduate School of Mathematics, Nagoya University, Nagoya 464-8602, Japan    Kazuhiko Minami Graduate School of Mathematics, Nagoya University, Nagoya 464-8602, Japan    Masatoshi Sato Yukawa Institute for Theoretical Physics, Kyoto University, Kyoto 606-8502, Japan
Abstract

We present a simple criterion for solvability of lattice spin systems on the basis of the graph theory and the simplicial homology. The lattice systems satisfy algebras with graphical representations. It is shown that the null spaces of adjacency matrices of the graphs provide conserved quantities of the systems. Furthermore, when the graphs belong to a class of simplicial complexes, the Hamiltonians are found to be mapped to bilinear forms of Majorana fermions, from which the full spectra of the systems are obtained. In the latter situation, we find a relation between conserved quantities and the first homology group of the graph, and the relation enables us to interpret the conserved quantities as flux excitations of the systems. The validity of our theory is confirmed in several known solvable spin systems including the 1d transverse-field Ising chain, the 2d Kitaev honeycomb model and the 3d diamond lattice model. We also present new solvable models on a 1d tri-junction, 2d and 3d fractal lattices, and the 3d cubic lattice.

I Introduction

Exactly solvable models have been played important roles in the understanding of physics in strongly correlated systems. In particular, exactly solvable lattice spin models have revealed many important phenomena. For instance, solving the 2d Ising model exactly, Onsager Onsager (1944) showed the presence of ferromagnetic phase transition in spin systems for the first time, which is one of milestones in statistical physics. Since Onsager’s work, other lattice spin models were solved exactly, such as the Potts model, the hard-hexagon model, and so on Baxter (2016); Wu (1971); Kadanoff and Wegner (1971). More recently, exactly solvable models also have disclosed exotic quantum phases in strongly correlated systems, such as spin liquid phases with non-abelian anyon excitations Kitaev (2006).

Quantum solvable lattice spin models are classified into three types. The first one has a Hamiltonian of which terms commute with each other, which includes the 2d Kitaev’s toric code Kitaev (2006); Kitaev and Laumann (2009), the X-cube model Castelnovo et al. (2010); Nandkishore and Hermele (2019), and so on. The second one has special symmetries such as Lie groups or quantum groups. This type includes the 1d Heisenberg model and the XXZ model Pasquier and Saleur (1990). Then, the last one can be transformed into free-fermion systems Jordan and Wigner (1928); Nambu (1995); Lieb et al. (1961); Niemeijer (1967); Katsura (1962); Perk et al. (1975); Minami (2016, 2017); Imamura and Katsura ; Prosko et al. (2017); Kaufman and Onsager (1949); Perk et al. (1975, 1984); Perk (2017); Kitaev (2006); Feng et al. (2007); Chen and Hu (2007); Chen and Nussinov (2008); Kitaev and Laumann (2009); Minami (2019). For instance, both the 1d XY model and the 1d transverse field Ising model can be converted into free-fermion systems by using the Jordan-Wigner transformation. Another example is the Kitaev’s honeycomb lattice model, which is transformed into a free fermion system by adapting a redundant representation of spins with Majorana operators.

In this paper, we present a simple criterion for the third type of solvability of lattice spin systems. Our criterion is based on the graph theory and the simplicial homology. For a lattice spin system with an algebra with a graphical representation, we show that the null space of the adjacency matrix of the graph provides conserved quantities of the system. Furthermore, when the graph belongs to a class of simplicial complexes, we reveal that the Hamiltonian is mapped to a bilinear form of Majorana fermions, from which the full spectrum of the system is obtained. We also find a relation between the conserved quantities and the first homology group of the graph. Based on the relation, we interpret the conserved quantities as flux excitations. We apply our criterion for several known solvable spin systems including the 1d transverse-field Ising chain, the 1d XY model, the 2d Kitaev honeycomb model, and the 3d diamond lattice model. We also present new solvable models on a 1d tri-junction, 2d and 3d fractal lattices, and the 3d cubic lattice.

The rest of this paper is organized as follows. In Sec. II, we present the main results. We introduce lattice models which satisfy a class of algebras. Representing the algebra in the form of a graph, we present Theorems that give the criterion of solvability in terms of the graph theory and the simplicial homology. In Sec. III, we illustrate our criterion by applying it to the 1d transverse-field Ising model, the XY model, the Kitaev honeycomb model and so on. We also provide new solvable models in Sec.IV In Sec. V, we present proofs of Theorems in Sec.II. We finally give discussions in Sec.VI.

II Main Results

First, we present our main results in this paper, which are summarized in three Theorems. The proofs of these Theorems will be given in Sec.V.

In this paper, we consider a class of Hamiltonians HH that satisfy the following properties.

  • HH has the form of H=j=1nλjhjH=\sum_{j=1}^{n}\lambda_{j}h_{j} with coefficients λj\lambda_{j}\in\mathbb{R} and operators hjh_{j} (j=1,,n)(j=1,\dots,n).

  • The operators hjh_{j} obey hj2=1h_{j}^{2}=1, hj=hjh_{j}^{\dagger}=h_{j}, and hjhk=ϵjkhkhjh_{j}h_{k}=\epsilon_{jk}h_{k}h_{j} with ϵij=±1\epsilon_{ij}=\pm 1.

The second property requires that hjh_{j}s commute or anti-commute with each other. The operators hjh_{j} generate an algebra 𝒜{\cal A} on \mathbb{C}, which we call the bond algebra (BA) Nussinov and Ortiz (2009); Cobanera et al. (2011). To represent the BA 𝒜{\cal A} visually, we introduce a graph 𝒢(𝒜){\cal G(A)} as follows.

  • Put nn vertices in general position and place hih_{i} on the ii-th vertex.

  • When hih_{i} and hjh_{j} anti-commute (commute) with each other, we draw (do not draw) a line between the vertices with hjh_{j} and hkh_{k}.

The resulting graph compactly encodes the information of the commutativity among hjh_{j}s. We call the graph 𝒢(𝒜){\cal G(A)} as commutativity graph (CG) Wang and Hazzard (2019) of 𝒜\cal{A}. The CG 𝒢(𝒜){\cal G(A)} has an algebraic representation with an adjacency matrix M(𝒜)M({\cal A}). The adjacency matrix M(𝒜)M({\cal A}) is a real symmetric n×nn\times n matrix of which elements indicate whether pairs of vertices are adjacent or not in 𝒢(𝒜){\cal G(A)}: The diagonal elements of M(𝒜)M({\cal A}) are zero and the (i,j)(i,j)-component is chosen to be 11 (0) if ii- and jj-th vertices in 𝒢(𝒜){\cal G(A)} are connected (not connected) by a line. The multiplication and the addition for M(𝒜)M({\cal A}) are defined as a matrix on the binary field 𝔽2\mathbb{F}_{2}, i.e. a matrix with entries 0 or 11, which satisfy 0+0=00+0=0, 0+1=10+1=1, 1+0=11+0=1, and 1+1=01+1=0.

Using M(𝒜)M({\cal A}), we present our first main result. A product hj1hj2hjkh_{j_{1}}h_{j_{2}}\cdots h_{j_{k}} conserves if it commutes with any hjh_{j} in HH. We find that such conserved quantities in 𝒜{\cal A} can be counted by using the adjacency matrix M(𝒜)M({\cal A}). More precisely, we have Theorem 1: {itembox}[l]Theorem 1 Let 𝒜\cal{A} be the BA of a Hamiltonian H=j=1nλjhjH=\sum_{j=1}^{n}\lambda_{j}h_{j}, 𝒢(𝒜)\cal{G(A)} be the corresponding CG of 𝒜\cal{A}, and M(𝒜)M({\cal A}) be the adjacency matrix of 𝒢(𝒜){\cal G(A)}. Then, the dimension of the kernel space of M(𝒜)M({\cal A}) coincides with the total number of conserved quantities in the form of hj1hjkh_{j_{1}}\cdots h_{j_{k}}.

Here the kernel space (or null space) of M(𝒜)M({\cal A}) is defined by

KerM(𝒜)={𝒗𝔽2n;M(𝒜)𝒗=𝟎}.\displaystyle{\rm Ker}M({\cal A})=\{{\bm{v}}\in\mathbb{F}_{2}^{n};M({\cal A}){\bm{v}}={\bm{0}}\}. (1)

As is shown in Sec.V, we can construct the conserved quantities from an element 𝒗{\bm{v}} of KerM(𝒜){\rm Ker}M({\cal A}): Let 𝒗(hj){\bm{v}}(h_{j}) be the unit vector on 𝔽2\mathbb{F}_{2} having a nonzero element only in the jj-th component,

𝒗(hj)=(00100)T.\displaystyle{\bm{v}}(h_{j})=\begin{pmatrix}0&\cdots&0&1&0\cdots&0\\ \end{pmatrix}^{T}. (2)

We can uniquely decompose 𝒗KerM(𝒜){\bm{v}}\in{\rm Ker}M({\cal A}) in the form of

𝒗=𝒗(hl1)+𝒗(hl2)++𝒗(hlm).\displaystyle{\bm{v}}={\bm{v}}(h_{l_{1}})+{\bm{v}}(h_{l_{2}})+\cdots+{\bm{v}}(h_{l_{m}}). (3)

Then, hl1hl2hlmh_{l_{1}}h_{l_{2}}\cdots h_{l_{m}} is a conserved quantity of HH.

The CG also enables us to characterize the BA geometrically. For this purpose, we adapt the notion of simplex: A dd-simplex is a dd-dimensional polyhedron having the minimal number of vertices, namely d+1d+1 vertices. For instance, a 0-simplex is a vertex, a 1-simplex is a line, a 2-simplex is a triangle, a 3-simplex is a tetrahedron, and so on. In particular, we consider a special set of simplices, which we call point-connected simplices: Let us consider a set of simplices S={s1,,sm}S=\{s_{1},\dots,s_{m}\} and let VV be a set consisting of all vertices of sαSs_{\alpha}\in S (α=1,,m\alpha=1,\dots,m). Then, we call SS as point-connected if SS is connected and any pair of sα,sβSs_{\alpha},s_{\beta}\in S (αβ\alpha\neq\beta) having a non-empty intersection shares only a single vertex vVv\in V (Namely sαsβ={v}s_{\alpha}\cap s_{\beta}=\{v\}). Furthermore, we call SS as single-point-connected if any vertex vVv\in V is shared by at most two different sαs_{\alpha}s. Adding all faces of sαSs_{\alpha}\in S (α=1,,m)(\alpha=1,\dots,m) into SS, we obtain a simplicial complex K(S)K(S), which we dub single-point-connected simplicial complex (SPSC). See Fig. 1.

Refer to caption
Figure 1: A single-point-connected simplicial complex. Two 3-simplices (dark brown tetrahedrons) , Two 2-simplices (light brown triangles), and three 1-simplices (black lines) are connected only by vertices.

Now we describe Theorem 2.

{itembox}

[l]Theorem 2 Let 𝒜\cal{A} be the BA of a Hamiltonian H=j=1nλjhjH=\sum_{j=1}^{n}\lambda_{j}h_{j} and 𝒢(𝒜)\cal{G(A)} be the corresponding CG of 𝒜\cal{A}. If 𝒢(𝒜){\cal G(A)} coincides with a SPSC K(S)K(S) with S={s1,,sm}S=\{s_{1},\dots,s_{m}\}, then HH is written by a bilinear form of mm Majorana operators. In particular, hjh_{j} is recast into

hj=iϵαβφαφβ,ϵαβ=±1,\displaystyle h_{j}=-i\epsilon_{\alpha\beta}\varphi_{\alpha}\varphi_{\beta},\quad\epsilon_{\alpha\beta}=\pm 1, (4)

where φα\varphi_{\alpha} are Majorana operators with the hermiticity φα=φα\varphi_{\alpha}^{\dagger}=\varphi_{\alpha} and the anti-commutation relation {φα,φβ}=2δα,β\{\varphi_{\alpha},\varphi_{\beta}\}=2\delta_{\alpha,\beta}.

Remarks are in order. (i) Without loss of generality, we can assume that any vertex vv of sαSs_{\alpha}\in S is shared by another sβSs_{\beta}\in S (βα\beta\neq\alpha): If not, we can add vv itself into SS as a 0-simplex to meet the assumption. (ii) Under this assumption, the Majorana operator φα\varphi_{\alpha} in Theorem 2 can be assigned to the simplex sαSs_{\alpha}\in S. Then, φα\varphi_{\alpha} and φβ\varphi_{\beta} in Eq.(4) are given by those on the simplices that share the vertex with hjh_{j}. (iii) The sign factors ϵαβ\epsilon_{\alpha\beta} in Eq.(4) are determined as follows. First, we use a sign ambiguity in Majorana operators: We can multiply φα\varphi_{\alpha} by 1-1 without changing the (anti-)commutation relations between them. Using this gauge transformation, we can change the m1m-1 relative signs between φα\varphi_{\alpha}, which enables us to erase m1m-1 ϵαβ\epsilon_{\alpha\beta}s. There still, however, remain nm+1n-m+1 ϵαβ\epsilon_{\alpha\beta}s. The following Theorem 3 tells us that these remaining sign factors are determined by conserved quantities.

{itembox}

[l]Theorem 3 Let 𝒜{\cal A} be the BA obeying the same assumption of Theorem 2. Then, K(S)K(S) has independent nm+1n-m+1 non-contractible loops as a simplicial complex on 𝔽2\mathbb{F}_{2}. Correspondingly, there exist nm+1n-m+1 conserved quantities that determine the remaining nm+1n-m+1 sign factors.

It should be noted here that for each non-contractible loop, there remains a sign factor that cannot be removed by the gauge transformation. To count the number of independent non-contractible loops in K(S)K(S), we calculate the homology group Hq(K(S))H_{q}(K(S)) of K(S)K(S). As we shall show in Sec.V, a straightforward calculation shows that Hq2(K(S))=0H_{q\geq 2}(K(S))=0 and dimH1(K(S))=nm+1{\rm dim}H_{1}(K(S))=n-m+1 when K(S)K(S) is a SPSC. The latter result implies that K(S)K(S) has nm+1n-m+1 independent non-contractible loops. We also find that each loop gives a conserved quantity: Take non-contractible loops as small as possible, then the product of all hjh_{j}s on each loop gives a conserved quantity. Furthermore, we find that the conserved quantity reduces to the sign factor on the loop by rewriting it in terms of Majorana fermions in Eq.(4).

Theorems 2 and 3 imply that HH is solvable as a free Majorana system: We can obtain the full spectrum of HH just by diagonalizing the free Majorana Hamiltonian.

We summarize the relation between the original spin model, the CG, the SPSC, and the free-fermion representation in Table 1.

Table 1: Relations between the original model, the commutativity graph (CG), the single-point-connected simplicial complex (SPSC), and the free-fermion representation.
original model \Leftrightarrow CG M(𝒜)M({\cal A}) \supset SPSC K(S)K(S) \Leftrightarrow free-fermion rep.
hjh_{j} \Leftrightarrow vertex vsαsβv\in s_{\alpha}\cap s_{\beta} \Leftrightarrow iϵαβφαφβ-i\epsilon_{\alpha\beta}\varphi_{\alpha}\varphi_{\beta}
{hi,hj}=0\{h_{i},h_{j}\}=0 \Leftrightarrow line
clique \Leftrightarrow sαK(S)s_{\alpha}\in K(S) \Leftrightarrow Majorana op. φα\varphi_{\alpha}
[h,H]=0[h,H]=0 \Leftrightarrow KerM(𝒜){\rm Ker}M({\cal A}) \supset H1(K(S))H_{1}(K(S)) \Leftrightarrow flux ϵ\epsilon

III Applications to known solvable models

In this section, we apply our theory to known solvable models, which confirms the validity of our criterion. There are also a lot of solvable lattice models by our method. For example, we have checked our method in models in Refs. Minami (2019, 2017); Nussinov and Ortiz (2009); Shi et al. (2009); Imamura and Katsura ; Prosko et al. (2017); Lee et al. (2007); Yu and Wang (2008); Chen and Kapustin (2019).

III.1 Transverse-Field Ising Model and Related Models

First, we examine a class of spin models obeying the following BA with n=2Nn=2N

hj2=1,{hj,hj+1}=0,\displaystyle h_{j}^{2}=1,\quad\{h_{j},h_{j+1}\}=0,
[hj,hk]=0,(jk±1).\displaystyle\left[h_{j},h_{k}\right]=0,\quad(j\neq k\pm 1). (5)

In the periodic boundary condition h2N+1=h1h_{2N+1}=h_{1}, the CG of this algebra is a circle in Fig.2.

Refer to caption
Figure 2: The CG of Eq.(5). The periodic boundary condition h2N+1=h1h_{2N+1}=h_{1} is imposed.

The corresponding adjacency matrix is given by

M(𝒜)=(0111010110101110).\displaystyle M({\cal A})=\begin{pmatrix}0&1&&&&1\\ 1&0&1&&\text{{\huge{0}}}&\\ &1&\ddots&\ddots&&\\ &&\ddots&\ddots&1&\\ &\text{\huge{0}}&&1&0&1\\ 1&&&&1&0\end{pmatrix}. (6)

For N2N\geq 2, the kernel space of M(𝒜)M({\cal A}) has the dimension 22, which is spanned by (1,0,1,0,)T(1,0,1,0,\dots)^{T} and (0,1,0,1,)T(0,1,0,1,\dots)^{T}. Therefore, from Theorem 1, we have two conserved quantities;

c1=h1h3h2N1,c2=h2h4h2N.\displaystyle c_{1}=h_{1}h_{3}\cdots h_{2N-1},\quad c_{2}=h_{2}h_{4}\cdots h_{2N}. (7)

Indeed, we can easily check that c1c_{1} and c2c_{2} commute with any hjh_{j}. We also find that the CG in Fig. 2 is a SPSC. Applying Theorem 2, we can rewrite hjh_{j} in the form of

hj=iϵjφj1φj,\displaystyle h_{j}=-i\epsilon_{j}\varphi_{j-1}\varphi_{j}, (8)

where φj\varphi_{j} is a Majorana operator and ϵj=±1\epsilon_{j}=\pm 1. Then, almost all ϵj\epsilon_{j}’s can be erased by redefining φj\varphi_{j} as φjϵj1φj\varphi_{j}\to\epsilon_{j}^{-1}\varphi_{j} (j=1,,2N1)j=1,\dots,2N-1), and after this, we obtain

hj=iφj1φj(j=1,,2N1),\displaystyle h_{j}=-i\varphi_{j-1}\varphi_{j}\quad(j=1,\cdots,2N-1),
h2N=iϵφ2N1φ2N.\displaystyle h_{2N}=-i\epsilon\varphi_{2N-1}\varphi_{2N}. (9)

The remaining ϵ\epsilon in Eq.(9) is determined by c1c2c_{1}c_{2},

ϵ=c1c2.\displaystyle\epsilon=-c_{1}c_{2}. (10)

The sign factor (1)Nϵ(-1)^{N}\epsilon corresponds to the π\pi-flux through the hole of the CG in Fig.2 Kaufman (1949).

In the open boundary condition, the CG is a line, and M(𝒜)M({\cal A}) becomes

M(𝒜)=(0100101001010010),\displaystyle M({\cal A})=\left(\begin{array}[]{cccccc}0&1&0&0&\cdots&\\ 1&0&1&0&\cdots&\\ 0&1&0&1&\cdots&\\ 0&0&1&0&\cdots&\\ \vdots&\vdots&\vdots&\vdots&\ddots&\\ \end{array}\right), (16)

of which kernel is dimension 0 for n=2Nn=2N. Now no conserved quantity is obtained, and thus ϵ=1\epsilon=1. In particular, in this case, our method naturally reproduces the Jordan-Wigner transformation Minami (2016). We can transform M(𝒜)M({\cal A}) into the following form

QTM(𝒜)Q=(0111101111011110),\displaystyle Q^{T}M({\cal A})Q=\left(\begin{array}[]{cccccc}0&1&1&1&\cdots&\\ 1&0&1&1&\cdots&\\ 1&1&0&1&\cdots&\\ 1&1&1&0&\cdots&\\ \vdots&\vdots&\vdots&\vdots&\ddots&\\ \end{array}\right), (22)

where QQ is given by

Q=pP[p,p+1]=(1111011100110001),\displaystyle Q=\prod_{p}P^{[p,p+1]}=\left(\begin{array}[]{cccccc}1&1&1&1&\cdots&\\ 0&1&1&1&\cdots&\\ 0&0&1&1&\cdots&\\ 0&0&0&1&\cdots&\\ \vdots&\vdots&\vdots&\vdots&\ddots&\\ \end{array}\right), (28)

where P[p,q]P^{[p,q]} is an elementary matrix with the (i,j)(i,j)-component Pij[p,q]=δij+δipδjqP_{ij}^{[p,q]}=\delta_{ij}+\delta_{ip}\delta_{jq}. As we shall show in Sec.V, P[p,q]P^{[p,q]} induces a map

{hp,,hq,}{hp,,hphq,},\displaystyle\{\ldots h_{p},\ldots,h_{q},\ldots\}\mapsto\{\ldots h_{p},\ldots,h_{p}h_{q},\ldots\}, (29)

and thus QQ gives a new bases

ej=h1h2hj.\displaystyle e_{j}=h_{1}h_{2}\cdots h_{j}. (30)

The commutation relations in QTM(𝒜)QQ^{T}M({\cal A})Q are eiej=ejeie_{i}e_{j}=-e_{j}e_{i} for all iji\neq j, those of the Clifford algebra. Introducing the initial operator h0h_{0} that obeys h02=1h_{0}^{2}=-1, {h0,h1}=0\{h_{0},h_{1}\}=0 and [h0,hj]=0[h_{0},h_{j}]=0 (j1j\neq 1), and defining φj\varphi_{j} as

φj=ij1h0h1h2hj,\displaystyle\varphi_{j}=i^{j-1}h_{0}h_{1}h_{2}\cdots h_{j}, (31)

we reproduces Eq.(9) with ϵ=1\epsilon=1. Equation (31) is an algebraic generalization of the Jordan-Wigner transformation Minami (2016). Actually, in the case of the transverse Ising chain below, by taking the initial operator as h0=iσ1xh_{0}=i\sigma^{x}_{1}, Eq.(31) reproduces the original Jordan-Wigner transformation.

For simplicity, we only consider the periodic boundary condition below.

III.1.1 Transverse-Field Ising Chain

The Hamiltonian of the transverse-field Ising chain is given by

H=Jj=1Nσjxσj+1xhj=1Nσjz,\displaystyle H=-J\sum_{j=1}^{N}\sigma_{j}^{x}\sigma_{j+1}^{x}-h\sum_{j=1}^{N}\sigma_{j}^{z}, (32)

where JJ is the exchange constant and hh is a transverse magnetic filed. From Eq.(32), the generator of the BA reads

h2j1=σjz,h2j=σjxσj+1x,\displaystyle h_{2j-1}=\sigma_{j}^{z},\quad h_{2j}=\sigma_{j}^{x}\sigma_{j+1}^{x}, (33)

which satisfies Eq.(5). The conserved quantities in Eq.(7) are given by

c1=j=1Nσjz.c2=1,\displaystyle c_{1}=\prod_{j=1}^{N}\sigma_{j}^{z}.\quad c_{2}=1, (34)

and thus the sign factor in Eq.(10) is

ϵ=j=1Nσjz.\displaystyle\epsilon=-\prod_{j=1}^{N}\sigma_{j}^{z}. (35)

From Eq.(9), the Hamiltonian is recast into

H\displaystyle H =hj=1Niφ2j2φ2j1+Jj=1N1iφ2j1φ2j\displaystyle=h\sum_{j=1}^{N}i\varphi_{2j-2}\varphi_{2j-1}+J\sum_{j=1}^{N-1}i\varphi_{2j-1}\varphi_{2j}
+Jiϵφ2N1φ2N,\displaystyle+Ji\epsilon\varphi_{2N-1}\varphi_{2N}, (36)

which reproduces the result in Ref.Minami (2016).

III.1.2 Orbital Compass Chain

Another model obeying Eq.(5) is the orbital compass chain,

H=Jxj=1Nσ2j1xσ2jxJyj=1Nσ2jyσ2j+1y,\displaystyle H=-J_{x}\sum_{j=1}^{N}\sigma_{2j-1}^{x}\sigma_{2j}^{x}-J_{y}\sum_{j=1}^{N}\sigma_{2j}^{y}\sigma_{2j+1}^{y}, (37)

where Eq.(5) is obtained by the following identification,

h2j1=σ2j1xσ2jx,h2j=σ2jyσ2j+1y.\displaystyle h_{2j-1}=\sigma_{2j-1}^{x}\sigma_{2j}^{x},\quad h_{2j}=\sigma_{2j}^{y}\sigma_{2j+1}^{y}. (38)

The conserved quantities c1c_{1} and c2c_{2} in Eq.(7) become

c1=j=12Nσjx,c2=j=12Nσjy,\displaystyle c_{1}=\prod_{j=1}^{2N}\sigma_{j}^{x},\quad c_{2}=\prod_{j=1}^{2N}\sigma_{j}^{y}, (39)

and thus ϵ\epsilon in Eq.(10) is

ϵ=(1)N+1j=12Nσjz.\displaystyle\epsilon=(-1)^{N+1}\prod_{j=1}^{2N}\sigma_{j}^{z}. (40)

In terms of Majorana operators, HH in Eq.(37) is given by

H=\displaystyle H= Jxj=1Niφ2j2φ2j1+Jyj=1N1iφ2j1φ2j\displaystyle J_{x}\sum_{j=1}^{N}i\varphi_{2j-2}\varphi_{2j-1}+J_{y}\sum_{j=1}^{N-1}i\varphi_{2j-1}\varphi_{2j}
+Jyiϵφ2N1φ2N,\displaystyle+J_{y}i\epsilon\varphi_{2N-1}\varphi_{2N}, (41)

which coincides with Eq.(36) if we identify JxJ_{x} and JyJ_{y} with hh and JJ. Therefore, there is a one-to-one correspondence between the spectrum of the orbital compass chain and that of the transverse-field Ising chain.

On the other hand, there exist additional degeneracies in the orbital compass chain. First, c2c_{2} in Eq.(39) can be ±1\pm 1, which gives two-fold degeneracy of each state. Moreover, we also have additional 2N2^{N}-fold degeneracy. This originates from the mismatch between the original spin degrees of freedom and the transformed Majorana degrees of freedom: The original spin space is 22N2^{2N}-dimensional, while the space of Majorana fermions is 2N2^{N}-dimensional. Correspondingly, there are additional conserved quantities djd_{j} (j=1,,2Nj=1,\dots,2N) which cannot be written by hjh_{j},

d2j1=σ2j1yσ2jy,d2j=σ2jxσ2j+1x.\displaystyle d_{2j-1}=\sigma_{2j-1}^{y}\sigma_{2j}^{y},\quad d_{2j}=\sigma_{2j}^{x}\sigma_{2j+1}^{x}. (42)

They satisfy the same BA as hjh_{j};

dj2=1,{dj,dj+1}=0,\displaystyle d_{j}^{2}=1,\quad\{d_{j},d_{j+1}\}=0,
[dj,dk]=0,(jk±1),\displaystyle\left[d_{j},d_{k}\right]=0,\quad(j\neq k\pm 1), (43)

and thus these operators are equivalent to 2N2N Majorana fermions. As a result, they generate additional 2N2^{N}-fold degeneracy.

III.2 XY Model and Related Models

Let hjh_{j}, hjh_{j}^{\prime}, and gjg_{j} (j=1,,2Nj=1,\dots,2N) be operators obeying

hj2=(hj)2=gj2=1,{hj,hj+1}={hj,hj+1}=0,\displaystyle h_{j}^{2}=(h_{j}^{\prime})^{2}=g_{j}^{2}=1,\quad\{h_{j},h_{j+1}\}=\{h_{j}^{\prime},h_{j+1}^{\prime}\}=0,
{hj,gj}={hj,gj}={hj+1,gj}={hj+1,gj}=0,\displaystyle\{h_{j},g_{j}\}=\{h_{j}^{\prime},g_{j}\}=\{h_{j+1},g_{j}\}=\{h_{j+1}^{\prime},g_{j}\}=0, (44)

where the other relations are commutative and the periodic boundary condition is assumed,

hi+2N=hi,hi+2N=hi,gi+2N=gi.\displaystyle h_{i+2N}=h_{i},\quad h_{i+2N}^{\prime}=h_{i}^{\prime},\quad g_{i+2N}=g_{i}. (45)

This algebra defines a class of models with the CG in Fig. 3.

Refer to caption
Figure 3: The CG of Eq.(44)

The dimension of the kernel space of the adjacency matrix is 2N+22N+2, and we have 2N+22N+2 conservative quantities:

ch=h1h2N,ch=h1h2N,cg=g1g2N,\displaystyle c_{h}=h_{1}\cdots h_{2N},\quad c_{h^{\prime}}=h_{1}^{\prime}\cdots h_{2N}^{\prime},\quad c_{g}=g_{1}\cdots g_{2N},
cj=gj1hjgjhj(j=1,,2N),\displaystyle c_{j}=g_{j-1}h_{j}^{\prime}g_{j}h_{j}\quad(j=1,\dots,2N), (46)

which satisfy

chchc1c2N=1.\displaystyle c_{h}c_{h^{\prime}}c_{1}\cdots c_{2N}=1. (47)

Since the CG in Fig.3 is a SPSC, the operators in Eq.(44) can be written by Majorana operators. Using the sign ambiguity (gauge degrees of freedom) of Majorana operators, we have

hj=iφj1φj,hj=iφj1φj\displaystyle h_{j}=-i\varphi_{j-1}\varphi_{j},\quad h_{j}^{\prime}=-i\varphi_{j-1}^{\prime}\varphi_{j}^{\prime}
gj=iϵjφjφj(j=1,,2N1),\displaystyle g_{j}=-i\epsilon_{j}\varphi_{j}\varphi_{j}^{\prime}\quad(j=1,\dots,2N-1),
h2N=iϵφ2N1φ2N,h2N=iϵφ2N1φ2N,\displaystyle h_{2N}=-i\epsilon\varphi_{2N-1}\varphi_{2N},\quad h_{2N}^{\prime}=-i\epsilon^{\prime}\varphi_{2N-1}^{\prime}\varphi_{2N}^{\prime},
g2N=iφ2Nφ2N,\displaystyle g_{2N}=-i\varphi_{2N}\varphi_{2N}^{\prime}, (48)

where φi\varphi_{i} and φi\varphi_{i}^{\prime} are Majorana operators. The sign factors ϵj\epsilon_{j}, ϵ\epsilon and ϵ\epsilon^{\prime} are determined by the conserved quantities in Eq.(46),

ϵj=k=1jck,ϵ=(1)Nch,ϵ=(1)Nch.\displaystyle\epsilon_{j}=\prod_{k=1}^{j}c_{k},\quad\epsilon=(-1)^{N}c_{h},\quad\epsilon^{\prime}=(-1)^{N}c_{h^{\prime}}. (49)

III.2.1 XY Model

As a prime example of models with the CG in Fig.3, we consider the XY model,

H=Ji=12N{(1+γ)σixσi+1x+(1γ)σiyσi+1y}hi=12Nσiz,H=-J\sum_{i=1}^{2N}\left\{(1+\gamma)\sigma_{i}^{x}\sigma_{i+1}^{x}+(1-\gamma)\sigma_{i}^{y}\sigma_{i+1}^{y}\right\}-h\sum_{i=1}^{2N}\sigma_{i}^{z}, (50)

where JJ is the exchange constant, γ\gamma is the asymmetric parameter, and hh is a magnetic field. Actually, with the following identification

h2j1=σ2j1xσ2jx,h2j=σ2jyσ2j+1y,\displaystyle h_{2j-1}=\sigma_{2j-1}^{x}\sigma_{2j}^{x},\quad h_{2j}=\sigma_{2j}^{y}\sigma_{2j+1}^{y},
h2j1=σ2j1yσ2jy,h2j=σ2jxσ2j+1x,\displaystyle h_{2j-1}^{\prime}=\sigma_{2j-1}^{y}\sigma_{2j}^{y},\quad h_{2j}^{\prime}=\sigma_{2j}^{x}\sigma_{2j+1}^{x},
gj=σj+1z,\displaystyle g_{j}=\sigma_{j+1}^{z}, (51)

we reproduce the BA in Eq.(44). In this model, the conserved quantities obey

c1==c2N=1,ch=ch=cg=j=12Nσjz,\displaystyle c_{1}=\cdots=c_{2N}=1,\quad c_{h}=c_{h}^{\prime}=-c_{g}=-\prod_{j=1}^{2N}\sigma_{j}^{z}, (52)

and thus we have

ϵj=1,ϵ=ϵ=(1)N+1j=12Nσjz.\displaystyle\epsilon_{j}=1,\quad\epsilon=\epsilon^{\prime}=(-1)^{N+1}\prod_{j=1}^{2N}\sigma_{j}^{z}. (53)

Therefore, Eq.(48) leads to

H\displaystyle H =iJj=1N{(1+γ)(φ2j2φ2j1+φ2j1φ2j)}\displaystyle=iJ\sum_{j=1}^{N}\left\{(1+\gamma)(\varphi_{2j-2}\varphi_{2j-1}+\varphi_{2j-1}^{\prime}\varphi_{2j}^{\prime})\right\}
+iJj=1N{(1γ)(φ2j1φ2j+φ2j2φ2j1)}\displaystyle+iJ\sum_{j=1}^{N}\left\{(1-\gamma)(\varphi_{2j-1}\varphi_{2j}+\varphi_{2j-2}^{\prime}\varphi_{2j-1}^{\prime})\right\}
+ihj=12Nφjφj\displaystyle+ih\sum_{j=1}^{2N}\varphi_{j}\varphi_{j}^{\prime}
iJ(1ϵ){(1+γ)φ2N1φ2N+(1γ)φ2N1φ2N}.\displaystyle-iJ(1-\epsilon)\left\{(1+\gamma)\varphi_{2N-1}\varphi_{2N}+(1-\gamma)\varphi_{2N-1}^{\prime}\varphi_{2N}^{\prime}\right\}. (54)

Equation (54) reproduces the known fermion representation of the XY model: Introducing the fermion operators aja_{j} as

φ2j1\displaystyle\varphi_{2j-1} =u2j1(a2j1+a2j1),\displaystyle=u_{2j-1}(a_{2j-1}+a_{2j-1}^{\dagger}),
φ2j1\displaystyle\varphi_{2j-1}^{\prime} =iu2j1(a2j1a2j1),\displaystyle=iu_{2j-1}(a_{2j-1}-a_{2j-1}^{\dagger}),
φ2j\displaystyle\varphi_{2j} =iu2j(a2ja2j),\displaystyle=-iu_{2j}(a_{2j}-a_{2j}^{\dagger}),
φ2j\displaystyle\varphi_{2j}^{\prime} =u2j(a2j+a2j),\displaystyle=u_{2j}(a_{2j}+a_{2j}^{\dagger}), (55)

with uj=(1)j(j1)/2u_{j}=(-1)^{j(j-1)/2}, we obtain

H=\displaystyle H= 2Jj=12N1[(ajaj+1+aj+1aj)+γ(ajaj+1+aj+1aj)]\displaystyle-2J\sum_{j=1}^{2N-1}\left[(a_{j}^{\dagger}a_{j+1}+a_{j+1}^{\dagger}a_{j})+\gamma(a_{j}^{\dagger}a_{j+1}^{\dagger}+a_{j+1}a_{j})\right]
2hj=12N(ajaj12)\displaystyle-2h\sum_{j=1}^{2N}\left(a_{j}^{\dagger}a_{j}-\frac{1}{2}\right)
+2Jcg[(ajaj+1+aj+1aj)+γ(ajaj+1+aj+1aj)],\displaystyle+2Jc_{g}\left[(a_{j}^{\dagger}a_{j+1}+a_{j+1}^{\dagger}a_{j})+\gamma(a_{j}^{\dagger}a_{j+1}^{\dagger}+a_{j+1}a_{j})\right], (56)

which is the same fermion reprentation in Ref. Niemeijer (1967).

III.2.2 Ladder Model

The second example is the ladder model DeGottardi et al. (2011),

H=\displaystyle H= Jtj=1N(σ2j1xσ2jx+σ2jyσ2j+1y)\displaystyle-J_{\rm t}\sum_{j=1}^{N}\left(\sigma_{2j-1}^{x}\sigma_{2j}^{x}+\sigma_{2j}^{y}\sigma_{2j+1}^{y}\right)
Jbj=1N(τ2j1xτ2jx+τ2jyτ2j+1y)\displaystyle-J_{\rm b}\sum_{j=1}^{N}\left(\tau_{2j-1}^{x}\tau_{2j}^{x}+\tau_{2j}^{y}\tau_{2j+1}^{y}\right)
Jj=12N(σjzτjz),\displaystyle-J_{\perp}\sum_{j=1}^{2N}\left(\sigma_{j}^{z}\tau_{j}^{z}\right), (57)

where JtJ_{\rm t} (JbJ_{\rm b}) is the intra exchange constant between top (bottom) spin chains, and JJ_{\perp} is the inter exchange constant between top and bottom chains. This model gives

h2j1=σ2j1xσ2jx,h2j=σ2jyσ2j+1y,\displaystyle h_{2j-1}=\sigma_{2j-1}^{x}\sigma_{2j}^{x},\quad h_{2j}=\sigma_{2j}^{y}\sigma_{2j+1}^{y},
h2j1=τ2j1xτ2jx,h2j=τ2jyτ2j+1y,\displaystyle h_{2j-1}^{\prime}=\tau_{2j-1}^{x}\tau_{2j}^{x},\quad h_{2j}^{\prime}=\tau_{2j}^{y}\tau_{2j+1}^{y},
gj=σjzτjz,\displaystyle g_{j}=\sigma_{j}^{z}\tau_{j}^{z}, (58)

which satisfy Eq.(44). In this model, we have

ch=j=12Nσjz,ch=j=12Nτjz,cg=chch,\displaystyle c_{h}=-\prod_{j=1}^{2N}\sigma_{j}^{z},\quad c_{h^{\prime}}=-\prod_{j=1}^{2N}\tau_{j}^{z},\quad c_{g}=c_{h}c_{h^{\prime}},
c2j1=σ2j1yσ2jyτ2j1yτ2jy,\displaystyle c_{2j-1}=-\sigma_{2j-1}^{y}\sigma_{2j}^{y}\tau_{2j-1}^{y}\tau_{2j}^{y},
c2j=σ2jxσ2j+1xτ2jxτ2j+1x,\displaystyle c_{2j}=-\sigma_{2j}^{x}\sigma_{2j+1}^{x}\tau_{2j}^{x}\tau_{2j+1}^{x}, (59)

which lead to

ϵ2j1=σ1yτ1y(k=22j1σkzτkz)σ2jyτ2jy,\displaystyle\epsilon_{2j-1}=-\sigma_{1}^{y}\tau_{1}^{y}\left(\prod_{k=2}^{2j-1}\sigma_{k}^{z}\tau_{k}^{z}\right)\sigma_{2j}^{y}\tau_{2j}^{y},\quad
ϵ2j=σ1yτ1y(k=22jσkzτkz)σ2j+1xτ2j+1x\displaystyle\epsilon_{2j}=-\sigma_{1}^{y}\tau_{1}^{y}\left(\prod_{k=2}^{2j}\sigma_{k}^{z}\tau_{k}^{z}\right)\sigma_{2j+1}^{x}\tau_{2j+1}^{x}
ϵ=(1)N+1j=12Nσjz,ϵ=(1)N+1j=12Nτjz,\displaystyle\epsilon^{\prime}=(-1)^{N+1}\prod_{j=1}^{2N}\sigma_{j}^{z},\quad\epsilon=(-1)^{N+1}\prod_{j=1}^{2N}\tau_{j}^{z}, (60)

where k=21σkzτkz1\prod_{k=2}^{1}\sigma_{k}^{z}\tau_{k}^{z}\equiv 1. The Hamiltonian is equivalent to

H\displaystyle H =iJtj=12N1φj1φj+iJtϵφ2N1φ2N\displaystyle=iJ_{\rm t}\sum_{j=1}^{2N-1}\varphi_{j-1}\varphi_{j}+iJ_{\rm t}\epsilon\varphi_{2N-1}\varphi_{2N}
+iJbj=12N1φj1φj+iJbϵφ2N1φ2N\displaystyle+iJ_{\rm b}\sum_{j=1}^{2N-1}\varphi^{\prime}_{j-1}\varphi^{\prime}_{j}+iJ_{\rm b}\epsilon^{\prime}\varphi_{2N-1}^{\prime}\varphi_{2N}^{\prime}
+iJj=12N1ϵjφjφj+iJφ2Nφ2N.\displaystyle+iJ_{\perp}\sum_{j=1}^{2N-1}\epsilon_{j}\varphi_{j}\varphi_{j}^{\prime}+iJ_{\perp}\varphi_{2N}\varphi_{2N}^{\prime}. (61)

III.2.3 Double Spin-Majorana Model

The third example is the double spin-Majorana model,

H=\displaystyle H= igj=12N(γjσjxγj+1+γjτjxγj+1)\displaystyle-ig\sum_{j=1}^{2N}\left(\gamma_{j}\sigma_{j}^{x}\gamma_{j+1}+\gamma_{j}^{\prime}\tau_{j}^{x}\gamma_{j+1}^{\prime}\right)
Jj=12Nσjzσj+1zτjzτj+1z,\displaystyle-J\sum_{j=1}^{2N}\sigma_{j}^{z}\sigma_{j+1}^{z}\tau_{j}^{z}\tau_{j+1}^{z}, (62)

where gg and JJ are real parameters, and γj\gamma_{j}’s are Majorana operators.  The BA of this model reads

hj=iγjσjxγj+1,hj=iγjτjxγj+1,\displaystyle h_{j}=i\gamma_{j}\sigma_{j}^{x}\gamma_{j+1},\quad h_{j}^{\prime}=i\gamma_{j}^{\prime}\tau_{j}^{x}\gamma_{j+1}^{\prime},
gj=σjzσj+1zτjzτj+1z,\displaystyle g_{j}=\sigma_{j}^{z}\sigma_{j+1}^{z}\tau_{j}^{z}\tau_{j+1}^{z}, (63)

which reproduces Eq.(44), and we obtain

ch=(1)Nj=12Nσjx,ch=(1)Nj=12Nτjx,cg=1,\displaystyle c_{h}=(-1)^{N}\prod_{j=1}^{2N}\sigma_{j}^{x},\quad c_{h^{\prime}}=(-1)^{N}\prod_{j=1}^{2N}\tau_{j}^{x},\quad c_{g}=1,
cj=σj1zτj1zσjxτjxσj+1zτj+1zγjγjγj+1γj+1.\displaystyle c_{j}=-\sigma_{j-1}^{z}\tau_{j-1}^{z}\sigma_{j}^{x}\tau_{j}^{x}\sigma_{j+1}^{z}\tau_{j+1}^{z}\gamma_{j}\gamma_{j}^{\prime}\gamma_{j+1}\gamma_{j+1}^{\prime}. (64)

Therefore,

ϵ1=σ2Nzτ2Nzσ1xτ1xσ2zτ2zγ1γ1γ2γ2,\displaystyle\epsilon_{1}=-\sigma_{2N}^{z}\tau_{2N}^{z}\sigma_{1}^{x}\tau_{1}^{x}\sigma_{2}^{z}\tau_{2}^{z}\gamma_{1}\gamma_{1}^{\prime}\gamma_{2}\gamma_{2}^{\prime},
ϵj=σ2Nzτ2Nzσ1yτ1y(k=2j1σkzτkz)\displaystyle\epsilon_{j}=-\sigma_{2N}^{z}\tau_{2N}^{z}\sigma_{1}^{y}\tau_{1}^{y}\left(\prod_{k=2}^{j-1}\sigma_{k}^{z}\tau_{k}^{z}\right)
×σjyτjyσj+1zτj+1zγ1γ1γj+1γj+1(j=2,,2N1),\displaystyle\times\sigma_{j}^{y}\tau_{j}^{y}\sigma_{j+1}^{z}\tau_{j+1}^{z}\gamma_{1}\gamma_{1}^{\prime}\gamma_{j+1}\gamma_{j+1}^{\prime}\quad(j=2,\dots,2N-1),
ϵ=j=12Nσjz,ϵ=j=12Nτjz,\displaystyle\epsilon=\prod_{j=1}^{2N}\sigma_{j}^{z},\quad\epsilon^{\prime}=\prod_{j=1}^{2N}\tau_{j}^{z}, (65)

where k=21σkzτkz1\prod_{k=2}^{1}\sigma_{k}^{z}\tau_{k}^{z}\equiv 1. The Hamiltonian is recast into

H\displaystyle H =igj=12N1(φj1φj+φj1φj)\displaystyle=ig\sum_{j=1}^{2N-1}\left(\varphi_{j-1}\varphi_{j}+\varphi_{j-1}^{\prime}\varphi_{j}^{\prime}\right)
+ig(ϵφ2N1φ2N+ϵφ2N1φ2N)\displaystyle+ig\left(\epsilon\varphi_{2N-1}\varphi_{2N}+\epsilon^{\prime}\varphi_{2N-1}^{\prime}\varphi_{2N}^{\prime}\right)
+iJj=12N1ϵjφjφj+iJφ2Nφ2N.\displaystyle+iJ\sum_{j=1}^{2N-1}\epsilon_{j}\varphi_{j}\varphi_{j}^{\prime}+iJ\varphi_{2N}\varphi^{\prime}_{2N}. (66)

In a manner similar to the orbital compass chain in Sec.III.1.2, this model hosts additional degeneracies originating from the mismatch between the original degrees of freedom and the transformed Majorana ones: It is found that the following operators djd_{j} and djd^{\prime}_{j} (j=1,,2Nj=1,\dots,2N) commute with hj,hj,gjh_{j},h_{j}^{\prime},g_{j},

dj=σj1zγjσjz,dj=τj1zγjτjz,\displaystyle d_{j}=\sigma_{j-1}^{z}\gamma_{j}\sigma_{j}^{z},\quad d_{j}^{\prime}=\tau_{j-1}^{z}\gamma_{j}^{\prime}\tau_{j}^{z}, (67)

which satisfies

{dj,dk}={dj,dk}=2δj,k,{dj,dk}=0.\displaystyle\{d_{j},d_{k}\}=\{d_{j}^{\prime},d_{k}^{\prime}\}=2\delta_{j,k},\quad\{d_{j},d_{k}^{\prime}\}=0. (68)

Thus, each state of this model has 22N2^{2N}-fold degeneracy.

III.3 Kitaev Honeycomb Lattice Model

The Kitaev honeycomb lattice is described by the following Hamiltonian with the nearest neighbour spin couplings,

H=\displaystyle H= Jxx-linksσjxσkxJyy-linksσjyσky\displaystyle-J_{x}\sum_{\text{$x$-links}}\sigma_{j}^{x}\sigma_{k}^{x}-J_{y}\sum_{\text{$y$-links}}\sigma_{j}^{y}\sigma_{k}^{y}
Jzz-linksσjzσkz,\displaystyle-J_{z}\sum_{\text{$z$-links}}\sigma_{j}^{z}\sigma_{k}^{z}, (69)

where the orientation of the xx, yy, and zz-links are indicated in Fig.4.

Refer to caption
Figure 4: xx-, yy- and zz-links in honeycomb lattice.
Refer to caption
Figure 5: The CG of the Kitaev honeycomb lattice model

Each term of Eq.(69) anti-commutes or commutes with each other, and thus it defines the BA. The CG of this model is the Kagome lattice in Fig. 5. The Kagome lattice is dual to the original honeycomb lattice, and each vertex in the Kagome lattice corresponds to a link in the honeycomb lattice. We assign an operator

hj,k=σjμ(j,k)σkμ(j,k)\displaystyle h_{j,k}=\sigma_{j}^{\mu(j,k)}\sigma_{k}^{\mu(j,k)} (70)

in the BA to each vertex of the Kagome lattice, where μ(j,k)=x,y,z\mu(j,k)=x,y,z is the spin-orientation at the corresponding (j,k)(j,k)-link in the honeycomb lattice. The conservative quantities are

cp=(j,k)phj,k,cz=(j,k):zlinkhj,k,\displaystyle c_{p}=\prod_{(j,k)\in\partial p}h_{j,k},\quad c_{z}=\prod_{(j,k):z-\text{link}}h_{j,k}, (71)

where pp is a hexagon in Fig. 5

Regarding triangles in Fig. 5 as 2-simplices, the CG can be identified with a SPSC. Therefore, we can apply Theorems 2 and 3 to the Kitaev honeycomb lattice model. The operator hj,kh_{j,k} is converted into a Majorana bi-linear form

hj,k=iϵjkφjφk,\displaystyle h_{j,k}=-i\epsilon_{jk}\varphi_{j}\varphi_{k}, (72)

so the Hamiltonian is equivalent to

H=j,kiJμ(j,k)ϵjkφjφk,\displaystyle H=\sum_{\langle j,k\rangle}iJ_{\mu(j,k)}\epsilon_{jk}\varphi_{j}\varphi_{k}, (73)

where ϵij\epsilon_{ij}’s are determined by the conserved quantities in Eq.(71). This result reproduces that in Ref.Kitaev and Laumann (2009), although our derivation is much simpler than the original one.

III.4 Diamond lattice model

The diamond lattice is a three-dimensional analog of the honeycomb lattice Ryu (2009); Wu et al. (2009). We can generalize the Kitaev honeycomb lattice model in three-dimensions. The Hamiltonian is given by

H=j,kJjk(αjμ(j,k)αkμ(j,k)+ζjμ(j,k)ζkμ(j,k)),\displaystyle H=-\sum_{\langle j,k\rangle}J_{jk}\left(\alpha_{j}^{\mu(j,k)}\alpha_{k}^{\mu(j,k)}+\zeta_{j}^{\mu(j,k)}\zeta_{k}^{\mu(j,k)}\right), (74)

where αjμ\alpha_{j}^{\mu} and ζjμ\zeta_{j}^{\mu} (μ=1,2,3,4)(\mu=1,2,3,4) are two sets of Dirac matrices,

αja=σjaτjx,αj4=σj0τjz,\displaystyle\alpha_{j}^{a}=\sigma_{j}^{a}\otimes\tau_{j}^{x},\quad\alpha_{j}^{4}=\sigma_{j}^{0}\otimes\tau_{j}^{z},
ζja=σjaτjz,ζj4=σj0τjx,\displaystyle\zeta_{j}^{a}=-\sigma_{j}^{a}\otimes\tau_{j}^{z},\quad\zeta_{j}^{4}=\sigma_{j}^{0}\otimes\tau_{j}^{x}, (75)

with a=1,2,3a=1,2,3, jj is the site index, and μ(j,k)=1,2,3,4\mu(j,k)=1,2,3,4 indicates the orientation of the gamma matrix at (j,k)(j,k)-link, as illustrated in Fig.6.

Refer to caption
Figure 6: Diamond lattice. The number at the link indicates the orientation μ\mu of the gamma matrix in the diamond lattice model.

We assign the operators hj,kh_{j,k} and hj,kh^{\prime}_{j,k} as

hj,k=αjμ(j,k)αkμ(j,k),hj,k=ζjμ(j,k)ζkμ(j,k),\displaystyle h_{j,k}=\alpha_{j}^{\mu(j,k)}\alpha_{k}^{\mu(j,k)},\quad h_{j,k}^{\prime}=\zeta_{j}^{\mu(j,k)}\zeta_{k}^{\mu(j,k)}, (76)

which satisfy

[hj,k,hl,m]=0.\displaystyle\left[h_{j,k},h_{l,m}^{\prime}\right]=0. (77)

The CGs of hj,kh_{j,k} and hj,kh_{j,k}^{\prime} are two identical pyrochlore lattices in Fig.7. By regarding tetrahedrons as 3-simplices, the pyrochlore lattice is identified with a SPSC. From straightforward calculation, we also find that conserved quantities in two CGs are the same. Therefore, we can transform hj,kh_{j,k}s and hj,kh_{j,k}^{\prime}s into Majorana bi-linear forms,

hj,k=iϵj,kφjφk,hj,k=iϵj,kφjφk.\displaystyle h_{j,k}=-i\epsilon_{j,k}\varphi_{j}\varphi_{k},\quad h_{j,k}^{\prime}=-i\epsilon_{j,k}\varphi_{j}^{\prime}\varphi_{k}^{\prime}. (78)

Consequently, the Hamiltonian is converted into

H=ij,kJjkϵj,k(φjφk+φjφk),\displaystyle H=i\sum_{\langle j,k\rangle}J_{jk}\epsilon_{j,k}\left(\varphi_{j}\varphi_{k}+\varphi_{j}^{\prime}\varphi_{k}^{\prime}\right), (79)

which reproduces that in Ryu (2009).

Refer to caption
Figure 7: The CG of the diamond lattice model with χ=α,ζ\chi=\alpha,\zeta.

IV New solvable models

So far, we have applied our method to known solvable models. Our approach also provides a powerful method to construct new solvable models in variety of lattices. In this section, we present such new solvable models.

IV.1 Tri-Junction Model

We first consider the transverse-field Ising chains with the tri-junction Giuliano et al. (2016); Backens et al. (2019); Giuliano et al. (2020). The Hamiltonian is given by

H=\displaystyle H= \displaystyle- a=13[Jaj=1N1σa,jzσa,j+1z+haj=2Nσa,jx]\displaystyle\sum_{a=1}^{3}\left[J_{a}\sum_{j=1}^{N-1}\sigma_{a,j}^{z}\sigma_{a,j+1}^{z}+h_{a}\sum_{j=2}^{N}\sigma_{a,j}^{x}\right] (80)
\displaystyle- t12σ1,1xσ2,1zt23σ2,1xσ3,1zt31σ3,1xσ1,1z,\displaystyle t_{12}\sigma_{1,1}^{x}\sigma_{2,1}^{z}-t_{23}\sigma_{2,1}^{x}\sigma_{3,1}^{z}-t_{31}\sigma_{3,1}^{x}\sigma_{1,1}^{z},

where JaJ_{a} and hah_{a} are the exchange constant and a magnetic field of aa-th chain, and tabt_{ab} are the coupling between aa-th and bb-th chains. The CG of this model is Fig. 8, where ha,jh_{a,j} (j=1,,2N1)(j=1,\dots,2N-1) is defined by

ha,1=σa,1xσa+1,1z,\displaystyle h_{a,1}=\sigma_{a,1}^{x}\sigma_{a+1,1}^{z},
ha,2l=σa,lzσa,l+1z,ha,2l+1=σa,l+1x.\displaystyle h_{a,2l}=\sigma_{a,l}^{z}\sigma_{a,l+1}^{z},\quad h_{a,2l+1}=\sigma_{a,l+1}^{x}. (81)

From the adjacency matrix of the CG, we find a conserved quantity

c\displaystyle c =ia=13j=1Nha,2j1\displaystyle=-i\prod_{a=1}^{3}\prod_{j=1}^{N}h_{a,2j-1}
=(a=13j=2Nσa,jx)(a=13σa,1y).\displaystyle=\left(\prod_{a=1}^{3}\prod_{j=2}^{N}\sigma_{a,j}^{x}\right)\left(\prod_{a=1}^{3}\sigma_{a,1}^{y}\right). (82)
Refer to caption
Figure 8: The CG of the tri-junction model

The CG in Fig.8 can be identified with a SPSC consisting of lines and a triangle. Therefore, applying Theorem 2 to this model, we have

ha,1=iφa,1φ,\displaystyle h_{a,1}=-i\varphi_{a,1}\varphi,
ha,j=iφa,j1φa,j,(j=2,,N).\displaystyle h_{a,j}=-i\varphi_{a,j-1}\varphi_{a,j},\quad(j=2,\dots,N). (83)

By using this, the Hamiltonian is recast into the bilinear form of Majorana operators,

H\displaystyle H =ia=13[Jaj=1Nφa,2j1φa,2j+ihaj=1Nφa,2jφa,2j+1]\displaystyle=i\sum_{a=1}^{3}\left[J_{a}\sum_{j=1}^{N}\varphi_{a,2j-1}\varphi_{a,2j}+ih_{a}\sum_{j=1}^{N}\varphi_{a,2j}\varphi_{a,2j+1}\right]
+(t12φ1,1+t23φ2,1+t31φ3,1)φ.\displaystyle+\left(t_{12}\varphi_{1,1}+t_{23}\varphi_{2,1}+t_{31}\varphi_{3,1}\right)\varphi. (84)

This model hosts implicit conserved quantities that is not obtained by hjh_{j},

ca=σa1,1zj=1Nσa,jx(a=1,2,3),\displaystyle c_{a}=\sigma_{a-1,1}^{z}\prod_{j=1}^{N}\sigma_{a,j}^{x}\quad(a=1,2,3), (85)

which satisfies

[ca,hb,j]=0,{ca,cb}=2δa,b,ic1c2c3=c.\displaystyle\left[c_{a},h_{b,j}\right]=0,\quad\{c_{a},c_{b}\}=2\delta_{a,b},\quad ic_{1}c_{2}c_{3}=c. (86)

These operators induce additional 22-fold degeneracy.

By same method, we can construct n-junction model whose junction is a n1n-1-simplex. We can also design tree-like models by junctions.

IV.2 Hanoi graph model

We can construct solvable models in 2d and 3d fractal lattices. Let us consider the Hanoi graph in Fig.9, and place a spin operator on each site of the Hanoi graph.

Refer to caption
Figure 9: Hanoi graph. xx, yy, and zz on each site denote the spin-orientation of the exchange interaction.

Then, we consider the Hamiltonian

H=\displaystyle H= J1σ1z\displaystyle-J_{1}\sigma_{1}^{z}
J12σ1xσ2zJ13σ1yσ3z\displaystyle-J_{12}\sigma_{1}^{x}\sigma_{2}^{z}-J_{13}\sigma_{1}^{y}\sigma_{3}^{z}
J23σ2yσ3xJ24σ2xσ4zJ35σ3yσ5z\displaystyle-J_{23}\sigma_{2}^{y}\sigma_{3}^{x}-J_{24}\sigma_{2}^{x}\sigma_{4}^{z}-J_{35}\sigma_{3}^{y}\sigma_{5}^{z}
,\displaystyle-\cdots, (87)

where σiμ\sigma^{\mu}_{i} is the μ\mu-th Pauli matrix at the ii-th site in Fig.9, and JijJ_{ij} is the exchange constant. The spin-orientation of the exchange interaction is determined as illustrated in Fig.9: In the case of the (1,2) link, for instance, we take σx\sigma^{x} and σz\sigma^{z} from site 1 and site 2, respectively.

The CG of this model is the Sierpinski gasket in Fig 10, where the operators at vertices are given by

h1=σ1z,\displaystyle h_{1}=\sigma_{1}^{z},
h1,2=σ1xσ2z,h1,3=σ1yσ3z,\displaystyle h_{1,2}=\sigma_{1}^{x}\sigma_{2}^{z},\quad h_{1,3}=\sigma_{1}^{y}\sigma_{3}^{z},
h2,3=σ2yσ3x,h2,4=σ2xσ4z,h3,5=σ3yσ5z,\displaystyle h_{2,3}=\sigma_{2}^{y}\sigma_{3}^{x},\quad h_{2,4}=\sigma_{2}^{x}\sigma_{4}^{z},\quad h_{3,5}=\sigma_{3}^{y}\sigma_{5}^{z},
.\displaystyle\cdots. (88)

Since the Sierpinski gasket is a SPSC generated by 2-simplices, the Hamiltonian (87) can be transformed into a Majorana-bilinear form. Note that the Sierpinski gasket is dual to the Hanoi graph.

Refer to caption
Figure 10: Sierpinski gasket
Refer to caption
Figure 11: Sierpinski tetrahedron

This model has 3d generalization. Instead of the Hanoi graph, we use the dual lattice of the Sierpinski tetrahedron in Fig.11. Placing a Spin(4) generator at each site, we can construct the Hamiltonian of which the CG is the Sierpinski tetrahedron. In the same way as the Hanoi graph, this model can be transformed into a Majorana-bilinear form.

IV.3 Octahedron model

The dimension of simplices in a SPSC can be higher than the space dimension. To illustrate this, we consider a spin model in the cubic lattice. We place an SO(6) spin (i.e. a Spin(6) generator) on each site of the cubic lattice, and consider the nearest neighbor interaction:

H=12𝒋μ=13Jμγ𝒋μγ𝒋+𝒆μμ+3g𝒋γ𝒋7,\displaystyle H=-\frac{1}{2}\sum_{\bm{j}}\sum_{\mu=1}^{3}J_{\mu}\gamma^{\mu}_{\bm{j}}\gamma^{\mu+3}_{\bm{j}+\bm{e}_{\mu}}-g\sum_{\bm{j}}\gamma_{\bm{j}}^{7}, (89)

where JμJ_{\mu} is the exchange constant, γ𝒋μ\gamma_{\bm{j}}^{\mu} is the SO(6) gamma matrix at the site 𝒋{\bm{j}}, and 𝒆μ{\bm{e}}_{\mu} is the unit vector in the μ\mu-th direction. We assign operators

h𝒋μ=γ𝒋μγ𝒋+𝒆μμ+3,h𝒋=γ𝒋7.\displaystyle h_{\bm{j}}^{\mu}=\gamma^{\mu}_{\bm{j}}\gamma^{\mu+3}_{\bm{j}+\bm{e}_{\mu}},\quad h_{\bm{j}}^{\prime}=\gamma_{\bm{j}}^{7}. (90)

The conserved quantities are

c𝒋μ,ν\displaystyle c_{\bm{j}}^{\mu,\nu} =h𝒋νh𝒋+𝒆νμh𝒋+𝒆μνh𝒋μ.\displaystyle=h_{\bm{j}}^{\nu}h_{\bm{j}+\bm{e}_{\nu}}^{\mu}h_{\bm{j}+\bm{e}_{\mu}}^{\nu}h_{\bm{j}}^{\mu}. (91)

The CG of this model is vertex-sharing octahedra with central vertex in Fig.12. It is a SPSC since an octahedron with central vertex is a 6-simplex. Thus, we can transform these operators into

h𝒋μ=iϵ𝒋μφ𝒋φ𝒋+𝒆μ,h𝒋=iφ𝒋φ𝒋,\displaystyle h_{\bm{j}}^{\mu}=-i\epsilon^{\mu}_{\bm{j}}\varphi_{\bm{j}}\varphi_{\bm{j}+\bm{e}_{\mu}},\quad h_{\bm{j}}^{\prime}=-i\varphi_{\bm{j}}\varphi_{\bm{j}}^{\prime}, (92)

and coserved quantities into

c𝒋μ,ν=ϵ𝒋νϵ𝒋+𝒆νμϵ𝒋+𝒆μνϵ𝒋μ.\displaystyle c_{\bm{j}}^{\mu,\nu}=\epsilon^{\nu}_{\bm{j}}\epsilon^{\mu}_{\bm{j}+\bm{e}_{\nu}}\epsilon^{\nu}_{\bm{j}+\bm{e}_{\mu}}\epsilon^{\mu}_{\bm{j}}. (93)

Therefore, the Hamiltonian is recast into

H=i2𝒋μ=13Jμϵ𝒋μφ𝒋φ𝒋+𝒆μ+ig𝒋φ𝒋φ𝒋.\displaystyle H=\frac{i}{2}\sum_{\bm{j}}\sum_{\mu=1}^{3}J_{\mu}\epsilon^{\mu}_{\bm{j}}\varphi_{\bm{j}}\varphi_{\bm{j}+\bm{e}_{\mu}}+ig\sum_{\bm{j}}\varphi_{\bm{j}}\varphi_{\bm{j}}^{\prime}. (94)
Refer to caption
Figure 12: Octahedron

In the following discussion, we take g=0g=0 for simplicity. In this case, iφ𝒋φ𝒋+𝒆μi\varphi^{\prime}_{\bm{j}}\varphi^{\prime}_{{\bm{j}}+{\bm{e}}_{\mu}} conserves, which induces additional 2N/22^{N/2}-fold degeneracy with the number of vertices NN. From Lieb’s theorem Lieb (2004), the ground state is realized when ϵ𝒋μ=1\epsilon^{\mu}_{\bm{j}}=1. In this case, the Hamiltonian becomes

H=i2𝒋μ=13Jμφ𝒋φ𝒋+𝒆μ.\displaystyle H=\frac{i}{2}\sum_{\bm{j}}\sum_{\mu=1}^{3}J_{\mu}\varphi_{\bm{j}}\varphi_{\bm{j}+\bm{e}_{\mu}}. (95)

By the Fourier transformation,

φ𝒋=d3p(2π)3(ei𝒑𝒋a𝒑+ei𝒑𝒋a𝒑),\displaystyle\varphi_{\bm{j}}=\int\frac{d^{3}p}{(2\pi)^{3}}\left(e^{i\bm{p}\cdot\bm{j}}a_{\bm{p}}+e^{-i\bm{p}\cdot\bm{j}}a_{\bm{p}}^{\dagger}\right), (96)

we have

H=i2μ=13Jdd3p(2π)3[eipμa𝒑a𝒑+eipμa𝒑a𝒑\displaystyle H=\frac{i}{2}\sum_{\mu=1}^{3}J_{d}\int\frac{d^{3}p}{(2\pi)^{3}}\left[e^{ip_{\mu}}a_{\bm{p}}a_{-\bm{p}}+e^{-ip_{\mu}}a_{\bm{p}}^{\dagger}a_{-\bm{p}}^{\dagger}\right.
+eipμa𝒑a𝒑+eipμa𝒑a𝒑].\displaystyle\left.+e^{ip_{\mu}}a_{\bm{p}}^{\dagger}a_{\bm{p}}+e^{-ip_{\mu}}a_{\bm{p}}a_{\bm{p}}^{\dagger}\right]. (97)

By diagonalizing this, the quasi-particle spectrum ε𝒑\varepsilon_{\bm{p}} is obtained as

ε𝒑=μ=13Jμsinpμ,\displaystyle\varepsilon_{\bm{p}}=\sum_{\mu=1}^{3}J_{\mu}\sin p_{\mu}, (98)

where the negative energy states are occupied in the ground state.

V Proofs

Now we prove our main results, Theorems 1-3, in Sec.II. To prove Theorem 1, we examine the basic properties of the CG. Let us consider a transformation of the operators

{hp,,hq,}{hp,,hphq,}.\displaystyle\{\ldots h_{p},\ldots,h_{q},\ldots\}\mapsto\{\ldots h_{p},\ldots,h_{p}h_{q},\ldots\}. (99)

Corresponding to this transformation, the CG is modified as follows:

  • i)

    Draw new lines from hphqh_{p}h_{q} to all the hkh_{k}’s that satisfy hphk=hkhph_{p}h_{k}=-h_{k}h_{p}.

  • ii)

    If there exist two lines from hphqh_{p}h_{q} to hkh_{k}, these lines should be eliminated and there remains no line between hphqh_{p}h_{q} and hkh_{k}.

Here the rule ii) corresponds to the fact that when hph_{p} and hqh_{q} anti-commutate with hkh_{k}, then the product hphqh_{p}h_{q} commutes with hkh_{k}.

We represent the modification i) ii) in terms of the adjacency matrix on 𝔽2\mathbb{F}_{2}: Let M(𝒜)M({\cal A}) be the adjacency matrix of the CG 𝒢(𝒜){\cal G(A)}, i.e.

M(𝒜)ij={0(hihj=hjhi)1(hihj=hjhi).\displaystyle M({\cal A})_{ij}=\left\{\begin{array}[]{cl}0&(h_{i}h_{j}=h_{j}h_{i})\\ 1&(h_{i}h_{j}=-h_{j}h_{i}).\end{array}\right. (102)

M(𝒜)M({\cal A}) is symmetric and its diagonal elements are all 0. The multiplication of hph_{p} to hqh_{q} corresponds to the row and column additions of M(𝒜)M({\cal A}), i.e. the qq-th row is replaced by the sum of qq-th and pp-th row, and the qq-th column is replaced by the sum of qq-th and pp-th column. The row and column additions are given by

M(𝒜)P[p,q]TM(𝒜)P[p,q],\displaystyle M({\cal A})\mapsto P^{[p,q]T}M({\cal A})P^{[p,q]}, (103)

where P[p,q]P^{[p,q]} is an elementary matrix with the (i,j)(i,j)-component Pij[p,q]=δij+δipδjqP^{[p,q]}_{ij}=\delta_{ij}+\delta_{ip}\delta_{jq}. Here the rule 1+1=01+1=0 in the matrix corresponds to the rule ii) above.

We can also represent Eq.(99) using the same elementary matrix P[p,q]P^{[p,q]}: Let 𝒗(hj){\bm{v}}(h_{j}) be the unit vector on 𝔽2\mathbb{F}_{2} having a nonzero element only in the jj-th component,

𝒗(hj)=(00100)T.\displaystyle{\bm{v}}(h_{j})=\begin{pmatrix}0&\cdots&0&1&0\cdots&0\\ \end{pmatrix}^{T}. (104)

Then, we have

P[p,q]𝒗(hj)={𝒗(hp)+𝒗(hq)for j=q,𝒗(hj)for jq,\displaystyle P^{[p,q]}{\bm{v}}(h_{j})=\left\{\begin{array}[]{ll}{\bm{v}}(h_{p})+{\bm{v}}(h_{q})&\mbox{for $j=q$},\\ {\bm{v}}(h_{j})&\mbox{for $j\neq q$},\end{array}\right. (107)

which reproduces Eq.(99) by regarding the addition 𝒗(hp)+𝒗(hq){\bm{v}}(h_{p})+{\bm{v}}(h_{q}) as the product hphqh_{p}h_{q}.

Now consider the following operations on the CG: If there are vertices hih_{i} and hjh_{j} that are connected to each other with a line, then multiply hih_{i} to all the vertices hkh_{k} that satisfy hkhj=hjhkh_{k}h_{j}=-h_{j}h_{k}, and multiply hjh_{j} to all the vertices hkh_{k} that satisfy hkhi=hihkh_{k}h_{i}=-h_{i}h_{k}. Then there remains no line beginning from hih_{i} and hjh_{j}, except a line between hih_{i} and hjh_{j}. As a result, we obtain a graph consisting only of hih_{i} and hjh_{j}, and a graph with other vertices. Repeating the same procedure for the latter graph, we inductively obtain graphs composed of only pairs and those with isolated vertices.

This modification leads to Theorem 1: After the modification of the CG, M(𝒜)M({\cal A}) is block diagonalized with r/2r/2 number of blocks with the form (0110)\displaystyle\left(\begin{array}[]{cc}0&1\\ 1&0\end{array}\right) and nrn-r number of blocks with 0 111This fact itself is already known in the context of the matrix theory (see for example Theorem 8.10.1 in Ref.Godsil and Royle (2001)).. (rr is even.) Here r/2r/2 is the number of the pairs and nrn-r is the number of the isolated vertices in the above. Since rr coincides with rankM(𝒜){\rm rank\>}M({\cal A}), the number of the pairs is unique. When hh belongs to the kernel of M(𝒜)M({\cal A}), it is evident that hh commutes with all the hih_{i}’s, and hence [H,h]=0[H,h]=0. Conversely, assume that h=hj1hj2hjkh=h_{j_{1}}h_{j_{2}}\cdots h_{j_{k}} satisfies [H,h]=0[H,h]=0. Then, we find hhi=ϵihihhh_{i}=\epsilon_{i}h_{i}h, ϵi=+1\epsilon_{i}=+1 or 1-1, for all hih_{i}. If hh is a constant, hh generates an isolated vertex, and belongs to the kernel. Otherwise from the condition [H,h]=0[H,h]=0 and the independence of hjh_{j}s, it is easy to derive that hh commutes with all h1h_{1},…, hnh_{n}, and hence hh belongs to the kernel of M(𝒜)M({\cal A}). Therefore, Theorem 1 holds.

By nothing that the (0110)\displaystyle\left(\begin{array}[]{cc}0&1\\ 1&0\end{array}\right) block and the 0 block correspond to the Clifford algebras Cl2Cl_{2} and Cl1Cl_{1}, respectively, the above modification process also implies Proposition: {itembox}[l]Proposition Let 𝒜(X){\cal A}(X) be the BA generated from the set of independent operators XX, and M(𝒜(X))M({\cal A}(X)) be its adjacency matrix. Then, we find 𝒜(X)(Cl2)r/2(Cl1)nr{\cal A}(X)\simeq(Cl_{2})^{r/2}\otimes(Cl_{1})^{n-r}, and 𝒜(X)𝒜(X){\cal A}(X)\simeq{\cal A}(X^{\prime}) if and only if rankM(𝒜(X))=rankM(𝒜(X)){\rm rank\>}M({\cal A}(X))={\rm rank\>}M({\cal A}(X^{\prime})).

In particular, when 𝒜{\cal A} gives the complete graph with nn vertices, i.e. a graph in which all vertices are connected to each other, and when we separate a pair of operators in a manner similar to the above, it is easy to convince that the remaining graph with n2n-2 vertices becomes again a complete graph. Iterating this procedure, we finally obtain n/2n/2 pairs when nn is even, and obtain (n1)/2(n-1)/2 pairs and an isolated vertex when nn is odd. The inverse of this modification is always possible. Since the complete graph with nn vertices represents the Clifford algebra with nn operators ClnCl_{n}, the rank of the adjacency matrix of the Clifford algebra with nn operators is nn when nn is even, and n1n-1 when nn is odd. This corresponds to the known fact Cl2nCl2nCl_{2n}\simeq Cl_{2}^{\otimes n} and Cl2n+1Cl2nCl1Cl_{2n+1}\simeq Cl_{2}^{\otimes n}\otimes Cl_{1}. Therefore, Proposition implies that a BA 𝒜{\cal A} with nn operators coincides with the Clifford algebra if rankM(𝒜)=n{\rm rank}M({\cal A})=n (rankM(𝒜)=n1{\rm rank}M({\cal A})=n-1) for even (odd) n.

Theorem 2 follows from the fact that hjh_{j} in Eq.(4) reproduces the BA of the CG that coincides with a SPSC: Let K(S)K(S) with S={s1,,sm}S=\{s_{1},\dots,s_{m}\} be the SPSC for the BA, and assign a Majorana operator φα\varphi_{\alpha} on each simplex sαSs_{\alpha}\in S. As we mentioned in Remark (i) in Sec.II, without loss of generality, we can assume that any vertex vv of sαSs_{\alpha}\in S is shared by another sβSs_{\beta}\in S (βα\beta\neq\alpha). Moreover, only the two simplices share vv since SS is single-point-connected. Under this assumption, we consider hj0iϵαβφαφβh_{j}^{0}\equiv-i\epsilon_{\alpha\beta}\varphi_{\alpha}\varphi_{\beta} for the vertex vjv_{j} with hjh_{j}, where φα\varphi_{\alpha} and φβ\varphi_{\beta} are located on the simplices that share vjv_{j}. Then, we find that {hi0,hj0}=0\{h^{0}_{i},h^{0}_{j}\}=0 ([hi0,hj0]=0[h^{0}_{i},h_{j}^{0}]=0) if viv_{i} and vjv_{j} are (not) vertices of the same simplex. These relations reproduce the BA of the SPSC, and thus, we can identify hj0h_{j}^{0} with hjh_{j}.

Finally, we prove Theorem 3. For preparation, we first show the following Lemma: {itembox}[l]Lemma Let K(S)K(S) with S={s1,,sm}S=\{s_{1},\dots,s_{m}\} be a SPSC. Then we have

Cq(K(S))=Cq(K(s1))Cq(K(sm))(q1),\displaystyle C_{q}(K(S))=C_{q}(K(s_{1}))\oplus\cdots\oplus C_{q}(K(s_{m}))\quad(q\geq 1), (108)

where CqC_{q} is the qq-chain on 𝔽2\mathbb{F}_{2}, and \oplus is the direct sum (i.e. Cq(K(sα))Cq(K(sβ))={0}C_{q}(K(s_{\alpha}))\cap C_{q}(K(s_{\beta}))=\{0\} for αβ\alpha\neq\beta). We also have

Hq(K(S))=0(q2).\displaystyle H_{q}(K(S))=0\quad(q\geq 2). (109)

The proof is as follow: Since K(S)K(S) consists of all faces of s1,,sms_{1},\dots,s_{m}, we have

Cq(K(S))=Cq(K(s1))++Cq(K(sm))(q1).\displaystyle C_{q}(K(S))=C_{q}(K(s_{1}))+\cdots+C_{q}(K(s_{m}))\quad(q\geq 1). (110)

Furthermore, it holds that Cq(K(sα))Cq(K(sβ))={0}C_{q}(K(s_{\alpha}))\cap C_{q}(K(s_{\beta}))=\{0\} for αβ\alpha\neq\beta and q1q\geq 1 since K(S)K(S) is a SPSC. Thus, Eq.(108) holds. Equation (109) immediately follows from Eq.(108): Since the boundary operator \partial maps a qq-chain to (q1)(q-1)-chain as,

:Cq(K(sα))Cq1(K(sα)),\displaystyle\partial:C_{q}(K(s_{\alpha}))\to C_{q-1}(K(s_{\alpha})), (111)

we obtain

Hq(K(S))=Hq(K(s1))Hq(K(sm))(q2),\displaystyle H_{q}(K(S))=H_{q}(K(s_{1}))\oplus\cdots\oplus H_{q}(K(s_{m}))\quad(q\geq 2), (112)

which turns to be zero because Hq(K(sα))=0H_{q}(K(s_{\alpha}))=0 (q1q\geq 1).

Now we can show that K(S)K(S) has nm+1n-m+1 independent non-contractible loops. Let hjh_{j} (j=1,,nj=1,\dots,n) be the generators of a BA and S={s1,,sm}S=\{s_{1},\dots,s_{m}\} be a set of simplices of which K(S)K(S) is a SPSC of the BA. Consider the Euler characteristic of χ(K(S))\chi(K(S)),

χ(K(S))\displaystyle\chi(K(S))
=q=0dimK(S)(1)q(the number of q-faces in K(S)),\displaystyle=\sum_{q=0}^{{\rm dim}K(S)}(-1)^{q}(\mbox{the number of $q$-faces in $K(S)$}), (113)

where a qq-face is a qq-simplex included in K(S)K(S) (namely a 0-face is a vertex of K(S)K(S), a 11-face is a hinge of K(S)K(S), and so on.) In terms of homology groups, χ(K(S))\chi(K(S)) is also written as Nakahara (2003)

χ(K(S))=q=0dimK(S)(1)qdimHq(K(S)).\displaystyle\chi(K(S))=\sum_{q=0}^{{\rm dim}K(S)}(-1)^{q}{\rm dim}H_{q}(K(S)). (114)

Since K(S)K(S) is connected, we have

dimH0(K(S))=1,\displaystyle{\rm dim}H_{0}(K(S))=1, (115)

and from Lemma, it holds that

dimHq2(K(S))=0.\displaystyle{\rm dim}H_{q\geq 2}(K(S))=0. (116)

Thus, dimH1(K(S)){\rm dim}H_{1}(K(S)) is evaluated as

dimH1(K(S))\displaystyle{\rm dim}H_{1}(K(S))
=1χ(K(S))\displaystyle=1-\chi(K(S))
=1q=0dimK(S)(1)q(the number of q-faces in K(S)).\displaystyle=1-\sum_{q=0}^{{\rm dim}K(S)}(-1)^{q}(\mbox{the number of $q$-faces in $K(S)$}). (117)

We compare this with the Euler characteristic of K(sα)K(s_{\alpha}) defined by

χ(K(sα))\displaystyle\chi(K(s_{\alpha}))
=q=0dimsα(1)q(the number of q-faces in sα).\displaystyle=\sum_{q=0}^{{\rm dim}s_{\alpha}}(-1)^{q}(\mbox{the number of $q$-faces in $s_{\alpha}$}). (118)

As sαs_{\alpha} is a simplex, we have

χ(K(sα))=1,\displaystyle\chi(K(s_{\alpha}))=1, (119)

and thus, summing the both sides of Eq.(118) for all sαSs_{\alpha}\in S, we obtain

m=α=1mq=0dimsα(1)q(the number of q-faces in sα).\displaystyle m=\sum_{\alpha=1}^{m}\sum_{q=0}^{{\rm dim}s_{\alpha}}(-1)^{q}(\mbox{the number of $q$-faces in $s_{\alpha}$}). (120)

On the other hand, as K(S)K(S) is a SPSC, we have

q=0dimK(S)(1)q(the number of q-faces in K(S))\displaystyle\sum_{q=0}^{{\rm dim}K(S)}(-1)^{q}(\mbox{the number of $q$-faces in $K(S)$})
=α=1mq=0dimsα(1)q(the number of q-faces in sα)n\displaystyle=\sum_{\alpha=1}^{m}\sum_{q=0}^{{\rm dim}s_{\alpha}}(-1)^{q}(\mbox{the number of $q$-faces in $s_{\alpha}$})-n (121)

Combing Eqs.(120) and (121) with Eq.(117), we get

dimH1(K(S))=nm+1,\displaystyle{\rm dim}H_{1}(K(S))=n-m+1, (122)

which implies that there exist nm+1n-m+1 non-contractible loops in K(S)K(S).

The nm+1n-m+1 non-contractible loops give nm+1n-m+1 conserved quantities: For each non-contractible loop, consider a product of hjh_{j} on all vertices in the loop. Obviously, the product reduces to a constant if we rewrite it in terms of Majorana fermions in Theorem 2. Thus, it conserves and Theorem 3 holds.

VI Discussion

In this paper, we present a simple criterion for solvability of lattice spin systems on the basis of the graph theory and the simplicial homology. When the lattice systems obey a class of algebras with the graphical representations, the spin systems can be converted into free Majorana fermion systems. We illustrate the validity of our criterion in a variery of spin systems.

Our method may reveal interesting aspects of lattice spin systems. After the conversion to Majorana bilinear forms, the lattice spin systems exhibit particle-hole symmetry, in a manner similar to superconductors, because of the self-conjugate property of Majorana fermions. Hence, they can be a kind of topological superconductors Sato and Ando (2017), although the origin of particle-hole symmetry is completely different. The Kitaev honeycomb lattice, for instance, exhibits a 2d non-abelian topological phase analogue to chiral superconductors, in the presence of time-reversal breaking perturbation Kitaev (2006). Our approach provides a systematic way to explore other interesting topological superconducting phases in spin systems; 3d non-abelian topological phase Sato (2003); Teo and Kane (2010), gapless topological phases Sato and Fujimoto (2010); Baum et al. (2015); Kobayashi et al. (2014); Agterberg et al. (2017), and topological crystalline superconductors Shiozaki and Sato (2014); Shiozaki et al. (2016). Searching such interesting phases is left for future work.

Acknowledgement

This work was supported by a Grant-in-Aid for Scientific Research on Innovative Areas “Topological Materials Science” (KAKENHI Grant No. JP15H05855) from the Japan Society for the Promotion of Science (JSPS). This work was also supported by JST CREST Grant No. JPMJCR19T2, Japan. M.S. was supported by KAKENHI Grant Nos. JP17H02922 and JP20H00131 from the JSPS. K.M. was supported by JSPS KAKENHI Grant No. JP19K03668.

Note added. — After completion of this work, we became aware of a recent related work Chapman and Flammia (2020).

References

  • Onsager (1944) L. Onsager, Physical Review 65, 117 (1944).
  • Baxter (2016) R. J. Baxter, Exactly solved models in statistical mechanics (Elsevier, 2016).
  • Wu (1971) F. Wu, Physical Review B 4, 2312 (1971).
  • Kadanoff and Wegner (1971) L. P. Kadanoff and F. J. Wegner, Physical Review B 4, 3989 (1971).
  • Kitaev (2006) A. Kitaev, Annals of Physics 321, 2 (2006).
  • Kitaev and Laumann (2009) A. Kitaev and C. Laumann, Exact methods in low-dimensional statistical physics and quantum computing,” Lecture Notes of the Les Houches Summer School , 101 (2009).
  • Castelnovo et al. (2010) C. Castelnovo, C. Chamon,  and D. Sherrington, Physical Review B 81, 184303 (2010).
  • Nandkishore and Hermele (2019) R. M. Nandkishore and M. Hermele, Annual Review of Condensed Matter Physics 10, 295 (2019).
  • Pasquier and Saleur (1990) V. Pasquier and H. Saleur, Nuclear Physics B 330, 523 (1990).
  • Jordan and Wigner (1928) P. Jordan and E. Wigner, Z. Physik 47, 631 (1928).
  • Nambu (1995) Y. Nambu, Broken Symmetry: Selected Papers of Y Nambu 13, 1 (1995).
  • Lieb et al. (1961) E. Lieb, T. Schultz,  and D. Mattis, Annals of Physics 16, 407 (1961).
  • Niemeijer (1967) T. Niemeijer, Physica 36, 377 (1967).
  • Katsura (1962) S. Katsura, Physical Review 127, 1508 (1962).
  • Perk et al. (1975) J. Perk, H. Capel, M. Zuilhof,  and T. J. Siskens, Physica A: Statistical Mechanics and its Applications 81, 319 (1975).
  • Minami (2016) K. Minami, Journal of the Physical Society of Japan 85, 024003 (2016).
  • Minami (2017) K. Minami, Nuclear Physics B 925, 144 (2017).
  • (18) Y. Imamura and H. Katsura, In preparation.
  • Prosko et al. (2017) C. Prosko, S.-P. Lee,  and J. Maciejko, Physical Review B 96, 205104 (2017).
  • Kaufman and Onsager (1949) B. Kaufman and L. Onsager, Physical Review 76, 1244 (1949).
  • Perk et al. (1984) J. Perk, H. Capel, G. Quispel,  and F. Nijhoff, Physica A: Statistical Mechanics and its Applications 123, 1 (1984).
  • Perk (2017) J. H. Perk, arXiv preprint arXiv:1710.03384  (2017).
  • Feng et al. (2007) X.-Y. Feng, G.-M. Zhang,  and T. Xiang, Physical review letters 98, 087204 (2007).
  • Chen and Hu (2007) H.-D. Chen and J. Hu, Physical Review B 76, 193101 (2007).
  • Chen and Nussinov (2008) H.-D. Chen and Z. Nussinov, Journal of Physics A: Mathematical and Theoretical 41, 075001 (2008).
  • Minami (2019) K. Minami, Nuclear Physics B 939, 465 (2019).
  • Nussinov and Ortiz (2009) Z. Nussinov and G. Ortiz, Physical Review B 79, 214440 (2009).
  • Cobanera et al. (2011) E. Cobanera, G. Ortiz,  and Z. Nussinov, Advances in physics 60, 679 (2011).
  • Wang and Hazzard (2019) Z. Wang and K. R. Hazzard, arXiv preprint arXiv:1908.03997  (2019).
  • Shi et al. (2009) X.-F. Shi, Y. Yu, J. You,  and F. Nori, Physical Review B 79, 134431 (2009).
  • Lee et al. (2007) D.-H. Lee, G.-M. Zhang,  and T. Xiang, Physical review letters 99, 196805 (2007).
  • Yu and Wang (2008) Y. Yu and Z. Wang, EPL (Europhysics Letters) 84, 57002 (2008).
  • Chen and Kapustin (2019) Y.-A. Chen and A. Kapustin, Physical Review B 100, 245127 (2019).
  • Kaufman (1949) B. Kaufman, Physical Review 76, 1232 (1949).
  • DeGottardi et al. (2011) W. DeGottardi, D. Sen,  and S. Vishveshwara, New Journal of Physics 13, 065028 (2011).
  • Ryu (2009) S. Ryu, Physical Review B 79, 075124 (2009).
  • Wu et al. (2009) C. Wu, D. Arovas,  and H.-H. Hung, Physical Review B 79, 134427 (2009).
  • Giuliano et al. (2016) D. Giuliano, G. Campagnano,  and A. Tagliacozzo, The European Physical Journal B 89, 251 (2016).
  • Backens et al. (2019) S. Backens, A. Shnirman,  and Y. Makhlin, Scientific reports 9, 1 (2019).
  • Giuliano et al. (2020) D. Giuliano, A. Trombettoni,  and P. Sodano, arXiv preprint arXiv:2002.06677  (2020).
  • Lieb (2004) E. H. Lieb, in Condensed Matter Physics and Exactly Soluble Models (Springer, 2004) pp. 79–82.
  • Note (1) This fact itself is already known in the context of the matrix theory (see for example Theorem 8.10.1 in Ref.Godsil and Royle (2001)).
  • Nakahara (2003) M. Nakahara, Geometry, Topology, and Physics (Taylor and Francis Group, LLC, 2003).
  • Sato and Ando (2017) M. Sato and Y. Ando, Reports on Progress in Physics 80, 076501 (2017).
  • Sato (2003) M. Sato, Physics Letters B 575, 126 (2003).
  • Teo and Kane (2010) J. C. Y. Teo and C. L. Kane, Phys. Rev. Lett. 104, 046401 (2010).
  • Sato and Fujimoto (2010) M. Sato and S. Fujimoto, Phys. Rev. Lett. 105, 217001 (2010).
  • Baum et al. (2015) Y. Baum, T. Posske, I. C. Fulga, B. Trauzettel,  and A. Stern, Phys. Rev. B 92, 045128 (2015).
  • Kobayashi et al. (2014) S. Kobayashi, K. Shiozaki, Y. Tanaka,  and M. Sato, Phys. Rev. B 90, 024516 (2014).
  • Agterberg et al. (2017) D. F. Agterberg, P. M. R. Brydon,  and C. Timm, Phys. Rev. Lett. 118, 127001 (2017).
  • Shiozaki and Sato (2014) K. Shiozaki and M. Sato, Phys. Rev. B 90, 165114 (2014).
  • Shiozaki et al. (2016) K. Shiozaki, M. Sato,  and K. Gomi, Phys. Rev. B 93, 195413 (2016).
  • Chapman and Flammia (2020) A. Chapman and S. T. Flammia, arXiv preprint arXiv:2003.05465  (2020).
  • Godsil and Royle (2001) C. Godsil and G. Royle, Algebraic Graph Theory (Springer, 2001).