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

The Exact Susceptibility of the Spin-S Transverse Ising Chain with Next-nearest-neighbor Interactions.

Kazuhiko Minami Graduate School of MathematicsGraduate School of Mathematics Nagoya University Nagoya University
Furo-cho
Furo-cho Chikusa-ku Chikusa-ku Nagoya Nagoya Aichi Aichi 464-8602 464-8602 Japan Japan
Abstract

The zero-field susceptibility of the spin-SS transverse Ising chain with next-nearest-neighbor interactions is obtained exactly. The susceptibility is given in an explicit form for S=1/2S=1/2, and expressed in terms of the eigenvectors of the transfer matrix for general spin SS. It is found that the low-temperature limit is independent of spin SS, and is divergent at the transition point.

1 Introduction

The transverse Ising model is a simple and basic spin model with non-trivial quantum effects. Specifically the spin 1/21/2 transverse Ising chain has been investigated by many authors. Fisher[1][2] derived the exact zero-field transverse susceptibility, Katsura[3] obtained the free energy, and Pfeuty[4] also obtained the free energy through the Jordan-Wigner transformation. The one-dimensional spin 1/21/2 transverse Ising model is known to be equivalent to the two-dimensional rectangular Ising model. [5] The Jordan-Wigner transformation was generalized,[6] and an infinite number of systems which can be regarded as kind of generalizations of the transverse Ising chain were solved exactly.[6][7][8] Applications of the Jordan-Wigner transformation to higher dimensions have also been considered by many authors.[9][10][11][12][13][14]

The Ising chain with general spin SS have also been investigated. [15][16][17] The zero-field susceptibility of spin SS transverse Ising chain was exactly derived.[18][19][20] The transverse susceptibility and the specific heat were shown up to S=60S=60 and the crossover from the S=1/2S=1/2 case to the classical case was exactly observed in Ref.\citen98Minami. The transverse susceptibility with a non-zero longitudinal external field, and that of Ising models with alternate or random structures were also obtained.[21][22] The continuous spin chain which can be regarded as the infinite spin limit of the Ising chain was also investigated. [23][24][25]

The Ising chain with next-nearest-neighbor interactions have also been investigated by many authors. The free energy and correlation functions were, with the use of a dual transformation, derived by Stephenson[26] and Hornreich et al.[27] The zero-field transverse susceptibility for the spin 1/21/2 case was calculated by Harada.[28] Oguchi introduced[29] a transfer matrix for the Ising chain with nearest- and next-nearest-neighbor interactions, and the free energy was re-derived by Kassan-Ogly.[30]

In this paper, the exact zero-field susceptibility of the transverse Ising chain with the next-nearest-neighbor interactions is derived for general spin SS. In section 2.1, the transverse susceptibility is expressed as an infinite sum of correlation functions. In section 2.2, the transfer matrix VV is introduced for the Ising chain with the next-nearest-neighbor interactions. The transfer matrix for the S=1/2S=1/2 case is explicitly considered in detail as an example. In section 2.3, the susceptibility is expressed as a finite sum using the eigenvectors of VV for general spin SS, and written down explicitly specifically for S=1/2S=1/2 . In section 3, the low-temperature limit is considered, and it is derived that the T0T\to 0 limit is independent of SS, and is divergent at the point where the ground state shows phase transition.

2 Formulation

2.1 Model and Formulation

Let us consider the Hamiltonian

=(0+1)H𝒬,{\cal H}=({\cal H}_{0}+{\cal H}_{1})-H{\cal Q}, (1)

where

0=J0j=1Nsjzsj+1z,1=J1j=1Nsjzsj+2z,𝒬=gμBj=1Nsjx,\displaystyle{\cal H}_{0}=-J_{0}\sum_{j=1}^{N}s_{j}^{z}s_{j+1}^{z},\hskip 22.76228pt{\cal H}_{1}=-J_{1}\sum_{j=1}^{N}s_{j}^{z}s_{j+2}^{z},\hskip 22.76228pt{\cal Q}=g\mu_{\rm B}\sum_{j=1}^{N}s_{j}^{x}, (2)

and the periodic boundary condition is assumed. Here Is=0+1{\cal H}_{\rm Is}={\cal H}_{0}+{\cal H}_{1} is the Ising interactions, and 𝒬{\cal Q} is the transverse term which does not commute with Is{\cal H}_{\rm Is}.

The transverse susceptibility χ\chi at zero-field is defined as

χ\displaystyle\chi =\displaystyle= H𝒬|H=0\displaystyle\frac{\partial}{\partial H}\langle{\cal Q}\rangle|_{H=0} (3)
=\displaystyle= HTr𝒬exp(β)Trexp(β)|H=0.\displaystyle\frac{\partial}{\partial H}\frac{{\rm Tr\>}{\cal Q}\exp(-\beta{\cal H})}{{\rm Tr}\exp(-\beta{\cal H})}\Big{|}_{H=0}.

The expectation \langle\>\rangle is taken by {\cal H}, and β=1/kBT\beta=1/k_{\rm B}T where kBk_{\rm B} is the Boltzmann constant and TT is the temperature. For the purpose to calculate (3), let us introduce the formula

ddHexp(β)=exp(β)0βexp(λ)𝒬exp(λ)𝑑λ,\frac{d}{dH}\exp(-\beta{\cal H})=\exp(-\beta{\cal H})\int_{0}^{\beta}\exp(\lambda{\cal H}){\cal Q}\exp(-\lambda{\cal H})d\lambda, (4)

which can be derived by multiplying exp(β)\exp(\beta{\cal H}) from the left-hand side and differentiating in terms of β\beta. Using (4), 𝒬\langle{\cal Q}\rangle can be expanded in terms of HH, and we obtain

χ\displaystyle\chi =\displaystyle= H[𝒬0+H0βexp(λIs)𝒬exp(λIs)𝒬0𝑑λ1+H0β𝒬0𝑑λ+O(H2)]|H=0\displaystyle\frac{\partial}{\partial H}[\frac{\langle{\cal Q}\rangle_{0}+H\int_{0}^{\beta}\langle\exp(\lambda{\cal H}_{\rm Is}){\cal Q}\exp(-\lambda{\cal H}_{\rm Is}){\cal Q}\rangle_{0}d\lambda}{1+H\int_{0}^{\beta}\langle{\cal Q}\rangle_{0}d\lambda}+O(H^{2})]\Big{|}_{H=0} (5)
=\displaystyle= 0βexp(λIs)𝒬exp(λIs)𝒬0𝑑λ,\displaystyle\int_{0}^{\beta}\langle\exp(\lambda{\cal H}_{\rm Is}){\cal Q}\exp(-\lambda{\cal H}_{\rm Is}){\cal Q}\rangle_{0}d\lambda,

where 0\langle\>\rangle_{0} is the expectation taken at zero external field H=0H=0.

Next let us introduce inner derivatives δ0\delta_{0} and δ1\delta_{1} through the relations

δ00𝒬=𝒬,δ0n+1𝒬=[0,δ0n𝒬],\displaystyle\delta_{0}^{0}{\cal Q}={\cal Q},\hskip 22.76228pt\delta_{0}^{n+1}{\cal Q}=[{\cal H}_{0},\delta_{0}^{n}{\cal Q}],
δ10𝒬=𝒬,δ1n+1𝒬=[1,δ1n𝒬],\displaystyle\delta_{1}^{0}{\cal Q}={\cal Q},\hskip 22.76228pt\delta_{1}^{n+1}{\cal Q}=[{\cal H}_{1},\delta_{1}^{n}{\cal Q}], (6)

and introduce δ=δ0+δ1\delta=\delta_{0}+\delta_{1}. The expectation in (5) can be expanded using

exp(λIs)𝒬exp(λIs)=p=0λpp!δp𝒬,\exp(\lambda{\cal H}_{\rm Is}){\cal Q}\exp(-\lambda{\cal H}_{\rm Is})=\sum_{p=0}^{\infty}\frac{\lambda^{p}}{p!}\delta^{p}{\cal Q}, (7)

and (5) is expressed as

χ=0βp=0λpp!δp𝒬𝒬0𝑑λ.\chi=\int_{0}^{\beta}\langle\sum_{p=0}^{\infty}\frac{\lambda^{p}}{p!}\delta^{p}{\cal Q}\>{\cal Q}\rangle_{0}d\lambda. (8)

It can be seen that (8) is equal to

χ\displaystyle\chi =\displaystyle= p=01p!0βλp𝑑λδp𝒬𝒬0\displaystyle\sum_{p=0}^{\infty}\frac{1}{p!}\int_{0}^{\beta}\lambda^{p}d\lambda\langle\delta^{p}{\cal Q}\>{\cal Q}\rangle_{0} (9)
=\displaystyle= p=0βp+1(p+1)!δp𝒬𝒬0.\displaystyle\sum_{p=0}^{\infty}\frac{\beta^{p+1}}{(p+1)!}\langle\delta^{p}{\cal Q}\>{\cal Q}\rangle_{0}.

This formula was already derived in Ref.\citen96Minami. What we have to do next is to evaluate the correlation functions δp𝒬𝒬0\langle\delta^{p}{\cal Q}\>{\cal Q}\rangle_{0} with non-zero next-nearest-neighbor interaction, and to perform the infinite sum in (9).

Then let us consider δp𝒬𝒬0\langle\delta^{p}{\cal Q}\>{\cal Q}\rangle_{0}. Because of the commutatibity 10=01{\cal H}_{1}{\cal H}_{0}={\cal H}_{0}{\cal H}_{1}, we generally find δ1δ0=δ0δ1\delta_{1}\delta_{0}=\delta_{0}\delta_{1}, and then δp𝒬\delta^{p}{\cal Q} is written as

δp𝒬\displaystyle\delta^{p}{\cal Q} =\displaystyle= (δ0+δ1)p𝒬=gμBl=0p(pl)δ0plδ1lj=1Nsjx,\displaystyle(\delta_{0}+\delta_{1})^{p}{\cal Q}=g\mu_{\rm B}\sum_{l=0}^{p}\left(\begin{array}[]{c}p\\ l\end{array}\right)\delta_{0}^{p-l}\delta_{1}^{l}\sum_{j=1}^{N}s^{x}_{j}, (12)

where

(pl)=p!(pl)!l!.\displaystyle\left(\begin{array}[]{c}p\\ l\end{array}\right)=\frac{p!}{(p-l)!l!}. (15)

It is straightforward to show inductively that

δ0p0δ1p1j=1Nsjx\displaystyle\delta_{0}^{p_{0}}\delta_{1}^{p_{1}}\sum_{j=1}^{N}s^{x}_{j}
=j=1N((J0)p0l0=0p0(p0l0)(sj1z)p0l0(sj+1z)l0)((J1)p1l1=0p1(p1l1)(sj2z)p1l1(sj+2z)l1)τp0+p1sjk(p0+p1),\displaystyle=\sum_{j=1}^{N}\Big{(}(-J_{0})^{p_{0}}\sum_{l_{0}=0}^{p_{0}}\left(\begin{array}[]{c}p_{0}\\ l_{0}\end{array}\right)(s^{z}_{j-1})^{p_{0}-l_{0}}(s^{z}_{j+1})^{l_{0}}\Big{)}\Big{(}(-J_{1})^{p_{1}}\sum_{l_{1}=0}^{p_{1}}\left(\begin{array}[]{c}p_{1}\\ l_{1}\end{array}\right)(s^{z}_{j-2})^{p_{1}-l_{1}}(s^{z}_{j+2})^{l_{1}}\Big{)}\tau_{p_{0}+p_{1}}s^{k(p_{0}+p_{1})}_{j}, (20)
(21)

where

τp={1p=evenip=odd\displaystyle\tau_{p}=\left\{\begin{array}[]{cl}1&\hskip 8.5359ptp={\rm even}\\ i&\hskip 8.5359ptp={\rm odd}\end{array}\right. (24)

and k(p)=xk(p)=x for even pp, and k(p)=yk(p)=y for odd pp. Then from (12) and (21), we find

δp𝒬𝒬0\displaystyle\langle\delta^{p}{\cal Q}\>{\cal Q}\rangle_{0} =\displaystyle= (gμB)2l=0p(pl)δ0plδ1lj=1Nsjx,j=1Nsjx0\displaystyle(g\mu_{\rm B})^{2}\sum_{l=0}^{p}\left(\begin{array}[]{c}p\\ l\end{array}\right)\langle\delta_{0}^{p-l}\delta_{1}^{l}\sum_{j=1}^{N}s^{x}_{j},\sum_{j^{\prime}=1}^{N}s^{x}_{j^{\prime}}\rangle_{0} (27)
=\displaystyle= (gμB)2j=1Nl=0p(pl)(J0)pl(J1)l\displaystyle(g\mu_{\rm B})^{2}\sum_{j=1}^{N}\sum_{l=0}^{p}\left(\begin{array}[]{c}p\\ l\end{array}\right)(-J_{0})^{p-l}(-J_{1})^{l} (30)
×\displaystyle\times l0=0p0(p0l0)l1=0p1(p1l1)(sj2z)p1l1(sj1z)p0l0(τpsjk(p)sjx)(sj+1z)l0(sj+2z)l10,\displaystyle\sum_{l_{0}=0}^{p_{0}}\left(\begin{array}[]{c}p_{0}\\ l_{0}\end{array}\right)\sum_{l_{1}=0}^{p_{1}}\left(\begin{array}[]{c}p_{1}\\ l_{1}\end{array}\right)\langle(s^{z}_{j-2})^{p_{1}-l_{1}}(s^{z}_{j-1})^{p_{0}-l_{0}}(\tau_{p}s^{k(p)}_{j}s^{x}_{j})(s^{z}_{j+1})^{l_{0}}(s^{z}_{j+2})^{l_{1}}\rangle_{0}, (35)
(36)

where p0=plp_{0}=p-l, p1=lp_{1}=l and thus p0+p1=pp_{0}+p_{1}=p. Here we used the fact that eigenstates of Is{\cal H}_{\rm Is} are direct products of eigenstates of each sjzs_{j}^{z}, and hence (sj2z)p1l1(sj1z)p0l0τpsjk(p)(sj+1z)l0(sj+2z)l1sjx0=0\langle(s^{z}_{j-2})^{p_{1}-l_{1}}(s^{z}_{j-1})^{p_{0}-l_{0}}\tau_{p}s^{k(p)}_{j}(s^{z}_{j+1})^{l_{0}}(s^{z}_{j+2})^{l_{1}}\cdot s^{x}_{j^{\prime}}\rangle_{0}=0 when jjj^{\prime}\neq j. The correlation function in (36) can be obtained through the transfer matrix method.

2.2 Transfer Matrix

The transfer matrix is defined as the matrix composed of the Boltzmann factors between two sites as its elements:

V=sj,sj+1|sj+1eKsjsj+1sj|,\displaystyle V=\sum_{s_{j},s_{j+1}}|s_{j+1}\rangle e^{Ks_{j}s_{j+1}}\langle s_{j}|, (37)

and thus

sj+1|V|sj=eKsjsj+1.\displaystyle\langle s_{j+1}|V|s_{j}\rangle=e^{Ks_{j}s_{j+1}}. (38)

Here we consider the state |sjj|s_{j}\rangle_{j} at site jj with the eigenvalue sjs_{j} (sj=S,S1,,S)(s_{j}=S,S-1,\ldots,-S): sjz|sjj=sj|sjjs^{z}_{j}|s_{j}\rangle_{j}=s_{j}|s_{j}\rangle_{j} , and |sjj|s_{j}\rangle_{j} is written |sj|s_{j}\rangle for abbreviation. If one operates VV from the left-hand side, then one additional bond is introduced to the chain; for example V|sj=sV|s_{j}=s\rangle provides all the (j+1)(j+1)-th states |sj+1|s_{j+1}\rangle multiplied by the Boltzmann factor obtained under the condition that the site jj has fixed eigenvalue sj=ss_{j}=s. In our case, we consider the configurations of adjacent two sites to introduce the next-nearest-neighbor interactions. The transfer matrix, in our notation, is defined as

V=sj1,sj,sj,sj+1|sj|sj+1eK0sj1sjeK1sj1sj+1δsjsjsj1|sj|,\displaystyle V=\sum_{s_{j-1},s_{j},s^{\prime}_{j},s^{\prime}_{j+1}}|s^{\prime}_{j}\rangle|s^{\prime}_{j+1}\rangle e^{K_{0}s_{j-1}s^{\prime}_{j}}e^{K_{1}s_{j-1}s^{\prime}_{j+1}}\delta_{s_{j}s^{\prime}_{j}}\langle s_{j-1}|\langle s_{j}|, (39)

and thus

sj|sj+1|V|sj1|sj=eK0sj1sjeK1sj1sj+1δsjsj,\displaystyle\langle s^{\prime}_{j}|\langle s^{\prime}_{j+1}|\>V\>|s_{j-1}\rangle|s_{j}\rangle=e^{K_{0}s_{j-1}s^{\prime}_{j}}e^{K_{1}s_{j-1}s^{\prime}_{j+1}}\delta_{s_{j}s^{\prime}_{j}}, (40)

where K0=βJ0K_{0}=\beta J_{0}, K1=βJ1K_{1}=\beta J_{1}, and sj1,sj,sj,sj+1=S,S1,,Ss_{j-1},s_{j},s^{\prime}_{j},s^{\prime}_{j+1}=S,S-1,\ldots,-S.

For example in the case of S=1/2S=1/2, we find

V=(xy01/xy0x/y0y/x00y/x0x/y01/xy0xy)\displaystyle V=\left(\begin{array}[]{cccc}xy&0&1/xy&0\\ x/y&0&y/x&0\\ 0&y/x&0&x/y\\ 0&1/xy&0&xy\end{array}\right) (45)

where x=e14K0x=e^{\frac{1}{4}K_{0}} and y=e14K1y=e^{\frac{1}{4}K_{1}}. The configurations {|sj1|sj}\{|s_{j-1}\rangle|s_{j}\rangle\} and {|sj|sj+1}\{|s^{\prime}_{j}\rangle|s^{\prime}_{j+1}\rangle\} are both arranged by binary rules as |12|12,|12|12,|12|12,|12|12|\frac{1}{2}\rangle|\frac{1}{2}\rangle,|\frac{1}{2}\rangle|\frac{-1}{2}\rangle,|\frac{-1}{2}\rangle|\frac{1}{2}\rangle,|\frac{-1}{2}\rangle|\frac{-1}{2}\rangle, in which each configuration corresponds to the ρ=1,2,3,4\rho=1,2,3,4-th element of the matrix, respectively. When one operate the transfer matrix from the left-hand side, then the Boltzmann weights coming from an additional site are introduced. Configurations with the inconsistency sjsjs_{j}\neq s^{{}^{\prime}}_{j} are eliminated multiplying 0 element. Then we find the matrix elements VρρV_{\rho\rho^{\prime}} for example (V)11=12|12|V|12|12=xy(V)_{11}=\langle\frac{1}{2}|\langle\frac{1}{2}|\>V\>|\frac{1}{2}\rangle|\frac{1}{2}\rangle=xy, (V)21=12|12|V|12|12=x/y(V)_{21}=\langle\frac{1}{2}|\langle\frac{-1}{2}|\>V\>|\frac{1}{2}\rangle|\frac{1}{2}\rangle=x/y, and (V)12=12|12|V|12|12=0(V)_{12}=\langle\frac{1}{2}|\langle\frac{1}{2}|\>V\>|\frac{1}{2}\rangle|\frac{-1}{2}\rangle=0. The eigenequation is

0\displaystyle 0 =\displaystyle= det(λEV)\displaystyle{\rm det\>}(\lambda E-V) (46)
=\displaystyle= λ42xyλ3+y2(x21x2)λ2+2yx(y21y2)λ(y21y2)2\displaystyle\lambda^{4}-2xy\lambda^{3}+y^{2}(x^{2}-\frac{1}{x^{2}})\lambda^{2}+2\frac{y}{x}(y^{2}-\frac{1}{y^{2}})\lambda-(y^{2}-\frac{1}{y^{2}})^{2}
=\displaystyle= [λ2y(x+1x)λ+(y21y2)][λ2y(x1x)λ(y21y2)].\displaystyle\Big{[}\lambda^{2}-y(x+\frac{1}{x})\lambda+(y^{2}-\frac{1}{y^{2}})\Big{]}\Big{[}\lambda^{2}-y(x-\frac{1}{x})\lambda-(y^{2}-\frac{1}{y^{2}})\Big{]}.

The eigenvalues are obtained as

λ1±\displaystyle\lambda_{1}^{\pm} =\displaystyle= 12[y(x+1x)±y2(x+1x)24(y21y2)],\displaystyle\frac{1}{2}\Big{[}y(x+\frac{1}{x})\pm\sqrt{y^{2}(x+\frac{1}{x})^{2}-4(y^{2}-\frac{1}{y^{2}})}\Big{]}, (47)

and

λ2±\displaystyle\lambda_{2}^{\pm} =\displaystyle= 12[y(x1x)±y2(x1x)2+4(y21y2)].\displaystyle\frac{1}{2}\Big{[}y(x-\frac{1}{x})\pm\sqrt{y^{2}(x-\frac{1}{x})^{2}+4(y^{2}-\frac{1}{y^{2}})}\Big{]}. (48)

When J1=0J_{1}=0, then y=1y=1, and we find λ1+=x+1x\displaystyle\lambda_{1}^{+}=x+\frac{1}{x}, λ1=0\lambda_{1}^{-}=0, λ2+=x1x\displaystyle\lambda_{2}^{+}=x-\frac{1}{x}, and λ2=0\lambda_{2}^{-}=0. When J0=0J_{0}=0, then x=1x=1, and we find λ1+=y+1y\displaystyle\lambda_{1}^{+}=y+\frac{1}{y}, λ1=y1y\displaystyle\lambda_{1}^{-}=y-\frac{1}{y}, and λ2±=±y21y2\lambda_{2}^{\pm}=\pm\sqrt{\displaystyle y^{2}-\frac{1}{y^{2}}}. The eigenvector corresponding to λ1+\lambda_{1}^{+} is obtained as

𝐮1=(1/xyuu1/xy),\displaystyle{\bf u}_{1}=\left(\begin{array}[]{c}1/xy\\ u_{-}\\ u_{-}\\ 1/xy\end{array}\right), (53)

where

u±\displaystyle u_{\pm} =\displaystyle= 12(R±y(x1x)),\displaystyle\frac{1}{2}\Big{(}R_{-}\pm y\>(x-\frac{1}{x})\Big{)},
R\displaystyle R_{-} =\displaystyle= y2(x+1x)24(y21y2).\displaystyle\sqrt{y^{2}(x+\frac{1}{x})^{2}-4(y^{2}-\frac{1}{y^{2}})}. (54)

The elements of VV are all non-negative, and VV is irreducible. Hence from the Perron-Frobenius theorem, the maximum eigenvalue of VV is unique. Together with the fact that λ1+\lambda_{1}^{+} is the maximum eigenvalue for J1=0J_{1}=0 and for J0=0J_{0}=0, λ1+\lambda_{1}^{+} is also the maximum for all J0J_{0} and J1J_{1}. Let UU be a matrix whose column vectors are the eigenvectors of VV, the first column being the eigenvector 𝐮1{\bf u}_{1}. Then the first row vector 𝐮¯1{\bar{\bf u}}_{1} of the matrix U1U^{-1} is obtained as

𝐮¯1=xy2R(u+,1xy,1xy,u+).\displaystyle{\bar{\bf u}}_{1}=\frac{xy}{2R_{-}}\Big{(}u_{+},\frac{1}{xy},\frac{1}{xy},u_{+}\Big{)}. (55)

It is easy to confirm that 𝐮¯1𝐮1=1{\bar{\bf u}}_{1}\cdot{\bf u}_{1}=1.

2.3 The Transverse Susceptibility

Let us then consider the expectation value in (36) for general spin SS. The quantum expectation of this type can be calculated using the transfer matrix method.[20] Let {|s1,,sN}\{|s_{1},\ldots,s_{N}\rangle\} be a complete set of the eigenstates of Is{\cal H}_{\rm Is}, in which each |s1,,sN=|s1|s2|sN|s_{1},\ldots,s_{N}\rangle=|s_{1}\rangle|s_{2}\rangle\cdots|s_{N}\rangle is a direct product of eigenstates of sjzs_{j}^{z} (j=1,,N)(j=1,\ldots,N). Then, the expectation value sjksjx0\langle s_{j}^{k}s_{j}^{x}\rangle_{0}, for example, can be calculated using the transfer matrix VV as

sjksjx0\displaystyle\langle s_{j}^{k}s_{j}^{x}\rangle_{0} =\displaystyle= Trsjksjxexp(βIs)Trexp(βIs)={sj}s1,,sN|sjksjxexp(βIs)|s1,,sN{sj}s1,,sN|exp(βIs)|s1,,sN\displaystyle\frac{{\rm Tr\>}s_{j}^{k}s_{j}^{x}\exp(-\beta{\cal H}_{\rm Is})}{{\rm Tr\>}\exp(-\beta{\cal H}_{\rm Is})}=\frac{\displaystyle\sum_{\{s_{j}\}}\langle s_{1},\ldots,s_{N}|s_{j}^{k}s_{j}^{x}\exp(-\beta{\cal H}_{\rm Is})|s_{1},\ldots,s_{N}\rangle}{\displaystyle\sum_{\{s_{j}\}}\langle s_{1},\ldots,s_{N}|\exp(-\beta{\cal H}_{\rm Is})|s_{1},\ldots,s_{N}\rangle} (56)
=\displaystyle= TrDj(p)VNTrVN,\displaystyle\frac{{\rm Tr\>}D_{j}^{(p)}V^{N}}{{\rm Tr\>}V^{N}},

where Dj(p)D_{j}^{(p)} is an operator defined by

Dj(p)=sj,sj+1|sj|sj+1Δρ(p)sj|sj+1|,\displaystyle D_{j}^{(p)}=\sum_{s^{\prime}_{j},s^{\prime}_{j+1}}|s^{\prime}_{j}\rangle|s^{\prime}_{j+1}\rangle\Delta_{\rho}^{(p)}\langle s^{\prime}_{j}|\langle s^{\prime}_{j+1}|, (57)

and represented by a diagonal matrix D(p)D^{(p)} with the elements

(D(p))ρρ\displaystyle(D^{(p)})_{\rho\rho^{\prime}} =\displaystyle= Δρ(p)δρρ,\displaystyle\Delta_{\rho}^{(p)}\delta_{\rho\rho^{\prime}},
Δρ(p)\displaystyle\Delta_{\rho}^{(p)} =\displaystyle= {(sxsx)ll=(S(S+1)(Sl+1)2)/2p=even(sysx)ll=(Sl+1)/2ip=odd,\displaystyle\left\{\begin{array}[]{ll}(s^{x}s^{x})_{ll}=(S(S+1)-(S-l+1)^{2})/2&p={\rm even}\\ (s^{y}s^{x})_{ll}=(S-l+1)/2i&p={\rm odd},\end{array}\right. (60)

and ρ\rho is the index which corresponds to the spin configuration (sj,sj+1)(s^{\prime}_{j},s^{\prime}_{j+1}) of the adjacent two sites jj and j+1j+1 (see for example below (45)), and ll is defined by sj=Sl+1s^{\prime}_{j}=S-l+1, i.e. ρ\rho denotes a configuration (sj,sj+1)(s^{\prime}_{j},s^{\prime}_{j+1}), and ll is determined from sjs^{\prime}_{j}. In the case of S=1/2S=1/2, for example, D(p)D^{(p)} is a diagonal matrix D(p)=diag(1/4,1/4,1/4,1/4)D^{(p)}={\rm diag\>}(1/4,1/4,1/4,1/4) when pp is even, and iD(p)=diag(1/4,1/4,1/4,1/4)iD^{(p)}={\rm diag\>}(1/4,1/4,-1/4,-1/4) when pp is odd. Let us introduce a parameter mρ=sj=Sl+1m_{\rho}=s^{\prime}_{j}=S-l+1 which is the eigenvalue satisfying sjz|sj=mρ|sjs_{j}^{z}|s^{\prime}_{j}\rangle=m_{\rho}|s^{\prime}_{j}\rangle. The parameter mρm_{\rho} will be used below.

Similarly the correlation function in (36) for general spin SS is then expressed, using the transfer matrix (39) and the operator

Mj=sj,sj+1|sj|sj+1sjsj|sj+1|,\displaystyle M_{j}=\sum_{s^{\prime}_{j},s^{\prime}_{j+1}}|s^{\prime}_{j}\rangle|s^{\prime}_{j+1}\rangle s^{\prime}_{j}\langle s^{\prime}_{j}|\langle s^{\prime}_{j+1}|, (61)

which is represented by a diagonal matrix M=diag(m1,,mn)M={\rm diag}(m_{1},\ldots,m_{n}), as

(sj2z)p1l1(sj1z)p0l0(τpsjk(p)sjx)(sj+1z)l0(sj+2z)l10=TrMp1l1VMp0l0VτpD(p)VMl0VMl1VN4TrVN\displaystyle\langle(s_{j-2}^{z})^{p_{1}-l_{1}}(s_{j-1}^{z})^{p_{0}-l_{0}}(\tau_{p}s_{j}^{k(p)}s_{j}^{x})(s_{j+1}^{z})^{l_{0}}(s_{j+2}^{z})^{l_{1}}\rangle_{0}=\frac{{\rm Tr\>}M^{p_{1}-l_{1}}VM^{p_{0}-l_{0}}V\tau_{p}D^{(p)}VM^{l_{0}}VM^{l_{1}}V^{N-4}}{{\rm Tr\>}V^{N}} (70)
=\displaystyle= TrMp1l1VMp0l0VτpD(p)VMl0VMl1U(1(λ2/λ1)N4(λn/λ1)N4)U1TrV4U(1(λ2/λ1)N4(λn/λ1)N4)U1,\displaystyle\frac{{\rm Tr\>}M^{p_{1}-l_{1}}VM^{p_{0}-l_{0}}V\tau_{p}D^{(p)}VM^{l_{0}}VM^{l_{1}}U\left(\begin{array}[]{cccc}1&&&\\ &(\lambda_{2}/\lambda_{1})^{N-4}&&\\ &&\ddots&\\ &&&(\lambda_{n}/\lambda_{1})^{N-4}\end{array}\right)U^{-1}}{{\rm Tr\>}V^{4}U\left(\begin{array}[]{cccc}1&&&\\ &(\lambda_{2}/\lambda_{1})^{N-4}&&\\ &&\ddots&\\ &&&(\lambda_{n}/\lambda_{1})^{N-4}\end{array}\right)U^{-1}},

where UU is a matrix whose column vectors are the eigenvectors of VV, the first column being the eigenvector that corresponds to the unique maximum eigenvalue λ1\lambda_{1}, and n=(2S+1)2n=(2S+1)^{2}. Matrix U1VU=ΛU^{-1}VU=\Lambda is a diagonal matrix whose diagonal elements are the eigenvalues of VV. Here let us consider the thermodynamic limit NN\rightarrow\infty, then we find (λρ/λ1)N40(\lambda_{\rho}/\lambda_{1})^{N-4}\to 0 for ρ1\rho\neq 1, and (70) is expressed as

1cι,κ=1nμ,ρ,ν=1nu¯1ιmιp1l1Vιμmμp0l0VμρτpΔρ(p)Vρνmνl0Vνκmκl1uκ1(N),\rightarrow\frac{1}{c}\sum_{\iota,\>\kappa=1}^{n}\sum_{\mu,\>\rho,\>\nu=1}^{n}{\bar{u}}_{1\iota}m_{\iota}^{p_{1}-l_{1}}V_{\iota\mu}m_{\mu}^{p_{0}-l_{0}}V_{\mu\rho}\tau_{p}\Delta_{\rho}^{(p)}V_{\rho\nu}m_{\nu}^{l_{0}}V_{\nu\kappa}m_{\kappa}^{l_{1}}u_{\kappa 1}\hskip 17.07182pt(N\to\infty), (71)

where

c=ι,κ=1nμ,ρ,ν=1nu¯1ιVιμVμρVρνVνκuκ1=λ14,\displaystyle c=\sum_{\iota,\>\kappa=1}^{n}\sum_{\mu,\>\rho,\>\nu=1}^{n}{\bar{u}}_{1\iota}V_{\iota\mu}V_{\mu\rho}V_{\rho\nu}V_{\nu\kappa}u_{\kappa 1}=\lambda_{1}^{4}, (72)

and u¯1ι=(U1)1ι{\bar{u}}_{1\iota}=(U^{-1})_{1\iota}, uκ1=(U)κ1u_{\kappa 1}=(U)_{\kappa 1}. The factor u¯1ιuκ1{\bar{u}}_{1\iota}u_{\kappa 1} is the exactly renormalized Boltzmann weight with fixed configurations ι\iota and κ\kappa. When the transfer matrix VV cannot be diagonalized, which occur for example when y2(x+1/x)/2=1y^{2}(x+1/x)/2=1 with x1x\neq 1 in the case of spin 1/21/2, we can consider the Jordan normal form, and obtain an identical result. The susceptibility in the thermodynamic limit is obtained, from (9), (36) and (71), as

χ0\displaystyle\chi_{0} =\displaystyle= limNχN\displaystyle\lim_{N\rightarrow\infty}\frac{\chi}{N} (80)
=\displaystyle= p=0βp+1(p+1)!limN1Nδp𝒬𝒬0\displaystyle\sum_{p=0}^{\infty}\frac{\beta^{p+1}}{(p+1)!}\lim_{N\rightarrow\infty}\frac{1}{N}\langle\delta^{p}{\cal Q}{\cal Q}\rangle_{0}
=\displaystyle= p=0βp+1(p+1)!(gμB)2l=0p(pl)(J0)pl(J1)l\displaystyle\sum_{p=0}^{\infty}\frac{\beta^{p+1}}{(p+1)!}(g\mu_{\rm B})^{2}\sum_{l=0}^{p}\left(\begin{array}[]{c}p\\ l\end{array}\right)(-J_{0})^{p-l}(-J_{1})^{l}
×l0=0p0(p0l0)l1=0p1(p1l1)1cι,κ=1nμ,ρ,ν=1nu¯1ιmιp1l1Vιμmμp0l0VμρτpΔρ(p)Vρνmνl0Vνκmκl1uκ1\displaystyle\hskip 22.76228pt\times\sum_{l_{0}=0}^{p_{0}}\left(\begin{array}[]{c}p_{0}\\ l_{0}\end{array}\right)\sum_{l_{1}=0}^{p_{1}}\left(\begin{array}[]{c}p_{1}\\ l_{1}\end{array}\right)\frac{1}{c}\sum_{\iota,\>\kappa=1}^{n}\sum_{\mu,\>\rho,\>\nu=1}^{n}{\bar{u}}_{1\iota}m_{\iota}^{p_{1}-l_{1}}V_{\iota\mu}m_{\mu}^{p_{0}-l_{0}}V_{\mu\rho}\tau_{p}\Delta_{\rho}^{(p)}V_{\rho\nu}m_{\nu}^{l_{0}}V_{\nu\kappa}m_{\kappa}^{l_{1}}u_{\kappa 1}
=\displaystyle= (gμB)2cp=0βp+1(p+1)!l=0p(pl)(J0)pl(J1)l\displaystyle\frac{(g\mu_{\rm B})^{2}}{c}\sum_{p=0}^{\infty}\frac{\beta^{p+1}}{(p+1)!}\sum_{l=0}^{p}\left(\begin{array}[]{c}p\\ l\end{array}\right)(-J_{0})^{p-l}(-J_{1})^{l}
×ι,κ=1nμ,ρ,ν=1nu¯1ιVιμVμρτpΔρ(p)VρνVνκuκ1(mμ+mν)pl(mι+mκ)l\displaystyle\hskip 22.76228pt\times\sum_{\iota,\>\kappa=1}^{n}\sum_{\mu,\>\rho,\>\nu=1}^{n}{\bar{u}}_{1\iota}V_{\iota\mu}V_{\mu\rho}\tau_{p}\Delta_{\rho}^{(p)}V_{\rho\nu}V_{\nu\kappa}u_{\kappa 1}(m_{\mu}+m_{\nu})^{p-l}(m_{\iota}+m_{\kappa})^{l}
=\displaystyle= (gμB)2cι,κ=1nμ,ρ,ν=1np=0βp+1(p+1)!((J0)(mμ+mν)+(J1)(mι+mκ))pu¯1ιVιμVμρτpΔρ(p)VρνVνκuκ1.\displaystyle\frac{(g\mu_{\rm B})^{2}}{c}\sum_{\iota,\>\kappa=1}^{n}\sum_{\mu,\>\rho,\>\nu=1}^{n}\sum_{p=0}^{\infty}\frac{\beta^{p+1}}{(p+1)!}\Big{(}(-J_{0})(m_{\mu}+m_{\nu})+(-J_{1})(m_{\iota}+m_{\kappa})\Big{)}^{p}{\bar{u}}_{1\iota}V_{\iota\mu}V_{\mu\rho}\tau_{p}\Delta_{\rho}^{(p)}V_{\rho\nu}V_{\nu\kappa}u_{\kappa 1}.

The sum should be classified according to (J0)(mμ+mν)+(J1)(mι+mκ)=J(ι,μ,ν,κ)(-J_{0})(m_{\mu}+m_{\nu})+(-J_{1})(m_{\iota}+m_{\kappa})=J(\iota,\mu,\nu,\kappa) and pp. Note that J(ι,μ,ν,κ)0=1J(\iota,\mu,\nu,\kappa)^{0}=1 even if J(ι,μ,ν,κ)=0J(\iota,\mu,\nu,\kappa)=0. Then we find

χ0\displaystyle\chi_{0} =\displaystyle= (gμB)2cι,κ,μ,ν,ρ=1(J(ι,μ,ν,κ)=0)nβ11!u¯1ιVιμVμρτ0Δρ(0)VρνVνκuκ1\displaystyle\frac{(g\mu_{\rm B})^{2}}{c}\sum_{\begin{subarray}{c}\iota,\>\kappa,\>\mu,\>\nu,\>\rho=1\\ (J(\iota,\mu,\nu,\kappa)=0)\end{subarray}}^{n}\frac{\beta^{1}}{1!}{\bar{u}}_{1\iota}V_{\iota\mu}V_{\mu\rho}\tau_{0}\Delta_{\rho}^{(0)}V_{\rho\nu}V_{\nu\kappa}u_{\kappa 1} (85)
+\displaystyle+ (gμB)2cι,κ,μ,ν,ρ=1(J(ι,μ,ν,κ)0)np=0(p=even)βp+1(p+1)!J(ι,μ,ν,κ)pu¯1ιVιμVμρΔρ(even)VρνVνκuκ1\displaystyle\frac{(g\mu_{\rm B})^{2}}{c}\sum_{\begin{subarray}{c}\iota,\>\kappa,\>\mu,\>\nu,\>\rho=1\\ (J(\iota,\mu,\nu,\kappa)\neq 0)\end{subarray}}^{n}\sum_{\begin{subarray}{c}p=0\\ (p={\rm even})\end{subarray}}^{\infty}\frac{\beta^{p+1}}{(p+1)!}J(\iota,\mu,\nu,\kappa)^{p}{\bar{u}}_{1\iota}V_{\iota\mu}V_{\mu\rho}\Delta_{\rho}^{(\rm even)}V_{\rho\nu}V_{\nu\kappa}u_{\kappa 1}
+\displaystyle+ (gμB)2cι,κ,μ,ν,ρ=1(J(ι,μ,ν,κ)0)np=1(p=odd)βp+1(p+1)!J(ι,μ,ν,κ)pu¯1ιVιμVμρiΔρ(odd)VρνVνκuκ1,\displaystyle\frac{(g\mu_{\rm B})^{2}}{c}\sum_{\begin{subarray}{c}\iota,\>\kappa,\>\mu,\>\nu,\>\rho=1\\ (J(\iota,\mu,\nu,\kappa)\neq 0)\end{subarray}}^{n}\sum_{\begin{subarray}{c}p=1\\ (p={\rm odd})\end{subarray}}^{\infty}\frac{\beta^{p+1}}{(p+1)!}J(\iota,\mu,\nu,\kappa)^{p}{\bar{u}}_{1\iota}V_{\iota\mu}V_{\mu\rho}i\Delta_{\rho}^{(\rm odd)}V_{\rho\nu}V_{\nu\kappa}u_{\kappa 1},

where Δρ(even)=Δρ(0)\Delta_{\rho}^{(\rm even)}=\Delta_{\rho}^{(0)} and Δρ(odd)=Δρ(1)\Delta_{\rho}^{(\rm odd)}=\Delta_{\rho}^{(1)}. Finally we obtain

χ0(gμB)2\displaystyle\frac{\chi_{0}}{(g\mu_{\rm B})^{2}} =\displaystyle= βcι,κ,μ,ν,ρ=1(J(ι,μ,ν,κ)=0)nu¯1ιVιμVμρΔρ(0)VρνVνκuκ1\displaystyle\frac{\beta}{c}\sum_{\begin{subarray}{c}\iota,\>\kappa,\>\mu,\>\nu,\>\rho=1\\ (J(\iota,\mu,\nu,\kappa)=0)\end{subarray}}^{n}{\bar{u}}_{1\iota}V_{\iota\mu}V_{\mu\rho}\Delta_{\rho}^{(0)}V_{\rho\nu}V_{\nu\kappa}u_{\kappa 1} (86)
+\displaystyle+ 1cι,κ,μ,ν,ρ=1(J(ι,μ,ν,κ)0)nsinhβJ(ι,μ,ν,κ)J(ι,μ,ν,κ)u¯1ιVιμVμρΔρ(even)VρνVνκuκ1\displaystyle\frac{1}{c}\sum_{\begin{subarray}{c}\iota,\>\kappa,\>\mu,\>\nu,\>\rho=1\\ (J(\iota,\mu,\nu,\kappa)\neq 0)\end{subarray}}^{n}\frac{\sinh\beta J(\iota,\mu,\nu,\kappa)}{J(\iota,\mu,\nu,\kappa)}{\bar{u}}_{1\iota}V_{\iota\mu}V_{\mu\rho}\Delta_{\rho}^{(\rm even)}V_{\rho\nu}V_{\nu\kappa}u_{\kappa 1}
+\displaystyle+ 1cι,κ,μ,ν,ρ=1(J(ι,μ,ν,κ)0)ncoshβJ(ι,μ,ν,κ)1J(ι,μ,ν,κ)u¯1ιVιμVμρiΔρ(odd)VρνVνκuκ1.\displaystyle\frac{1}{c}\sum_{\begin{subarray}{c}\iota,\>\kappa,\>\mu,\>\nu,\>\rho=1\\ (J(\iota,\mu,\nu,\kappa)\neq 0)\end{subarray}}^{n}\frac{\cosh\beta J(\iota,\mu,\nu,\kappa)-1}{J(\iota,\mu,\nu,\kappa)}{\bar{u}}_{1\iota}V_{\iota\mu}V_{\mu\rho}i\Delta_{\rho}^{(\rm odd)}V_{\rho\nu}V_{\nu\kappa}u_{\kappa 1}.

Equation (86) provides the exact zero-field transverse susceptibility of the spin-SS transverse Ising model (1). The result is written in terms of finite summations.

In the case of S=1/2S=1/2, long and straightforward calculations yield that the susceptibility is

χ0(gμB)2\displaystyle\frac{\chi_{0}}{(g\mu_{\rm B})^{2}} =\displaystyle= β4c[(x2+1x2)+yR(x+1x)((x1x)2+2y4)]\displaystyle\frac{\beta}{4c}\Big{[}(x^{2}+\frac{1}{x^{2}})+\frac{y}{R_{-}}(x+\frac{1}{x})\Big{(}(x-\frac{1}{x})^{2}+\frac{2}{y^{4}}\Big{)}\Big{]} (87)
+\displaystyle+ 1J012cxyR(x21x2)[Rxy+1x(x+1x)]\displaystyle\frac{1}{J_{0}}\frac{1}{2c}\frac{xy}{R_{-}}(x^{2}-\frac{1}{x^{2}})\Big{[}\frac{R_{-}}{xy}+\frac{1}{x}(x+\frac{1}{x})\Big{]}
+\displaystyle+ 1J112cxyR1y2(y21y2)[Rxy+1x(x+1x)]\displaystyle\frac{1}{J_{1}}\frac{1}{2c}\frac{xy}{R_{-}}\frac{1}{y^{2}}(y^{2}-\frac{1}{y^{2}})\Big{[}\frac{R_{-}}{xy}+\frac{1}{x}(x+\frac{1}{x})\Big{]}
+\displaystyle+ 1J0J112cxyR1x2(x21y21x2y2)[yxu+1y2]\displaystyle\frac{1}{J_{0}-J_{1}}\frac{1}{2c}\frac{xy}{R_{-}}\frac{1}{x^{2}}(x^{2}\frac{1}{y^{2}}-\frac{1}{x^{2}}y^{2})\Big{[}\frac{y}{x}u_{-}+\frac{1}{y^{2}}\Big{]}
+\displaystyle+ 1J0+J112cxyR(x2y21x2y2)[xyu++1y2],\displaystyle\frac{1}{J_{0}+J_{1}}\frac{1}{2c}\frac{xy}{R_{-}}(x^{2}y^{2}-\frac{1}{x^{2}y^{2}})\Big{[}xyu_{+}+\frac{1}{y^{2}}\Big{]},

where

c\displaystyle c =\displaystyle= 12(y4(x4+1x4)+4(x2+1x2)+4+2y4)\displaystyle\frac{1}{2}\Big{(}y^{4}(x^{4}+\frac{1}{x^{4}})+4(x^{2}+\frac{1}{x^{2}})+4+\frac{2}{y^{4}}\Big{)} (88)
+\displaystyle+ 12yR(x+1x)((x2+1x2)+2y4)((x1x)2y4+4)\displaystyle\frac{1}{2}\frac{y}{R_{-}}(x+\frac{1}{x})\Big{(}(x^{2}+\frac{1}{x^{2}})+\frac{2}{y^{4}}\Big{)}\Big{(}(x-\frac{1}{x})^{2}y^{4}+4\Big{)}
=\displaystyle= (λ1+)4,\displaystyle(\lambda_{1}^{+})^{4},

and x=eK0/4x=e^{K_{0}/4}, y=eK1/4y=e^{K_{1}/4}, and u±u_{\pm} and RR_{-} are defined in (54). Generally in the case of spin SS, we have to find the eigenvectors of n×n=(2S+1)2×(2S+1)2n\times n=(2S+1)^{2}\times(2S+1)^{2} matrix VV.

The first term in (87) corresponds to the case mμ+mν=0m_{\mu}+m_{\nu}=0 and mι+mκ=0m_{\iota}+m_{\kappa}=0 which yields J(ι,μ,ν,κ)=0J(\iota,\mu,\nu,\kappa)=0 for all J0J_{0} and J1J_{1}. Other mρm_{\rho}’s sometimes yield J(ι,μ,ν,κ)=0J(\iota,\mu,\nu,\kappa)=0 when J0J_{0} and J1J_{1} take some special values (for example, J1=J0J_{1}=-J_{0}, mμ+mν=2Sm_{\mu}+m_{\nu}=2S, mι+mκ=2Sm_{\iota}+m_{\kappa}=2S result in J(ι,μ,ν,κ)=0J(\iota,\mu,\nu,\kappa)=0). Contributions from these latter cases, if exist, can be obtained by taking limit J(ι,μ,ν,κ)0J(\iota,\mu,\nu,\kappa)\to 0 in the remaining four terms (thus the susceptibility for J1=J0J_{1}=-J_{0}, for example, is simply obtained by taking limit J1J0J_{1}\to-J_{0} in (87)).

When J1=0J_{1}=0 or J0=0J_{0}=0, the system reduces to the nearest-neighbor transverse Ising chain. For example in (87), let J10J_{1}\to 0 then y1y\to 1 and we obtain

J0χ0(gμB)2=K0/2(x+1x)2+12x1xx+1x=12(K0/4cosh214K0+tanh14K0),\displaystyle\frac{J_{0}\chi_{0}}{(g\mu_{\rm B})^{2}}=\frac{K_{0}/2}{\displaystyle(x+\frac{1}{x})^{2}}+\frac{1}{2}\frac{\displaystyle x-\frac{1}{x}}{\displaystyle x+\frac{1}{x}}=\frac{1}{2}\Big{(}\frac{K_{0}/4}{\cosh^{2}\frac{1}{4}K_{0}}+\tanh\frac{1}{4}K_{0}\Big{)}, (89)

which is the transverse susceptibility of the nearest-neighbor Ising chain.[1][2] When J00J_{0}\to 0 then x1x\to 1 and we again obtain (89) in which xx is replaced by yy, and J0J_{0} is replaced by J1J_{1}.

The susceptibility is invariant under the change of sign J0J0J_{0}\mapsto-J_{0}. For example (87) and (88) is invariant under x1/xx\mapsto 1/x. This invariance comes from the symmetry of the system under transformations J0J0J_{0}\mapsto-J_{0} and (σjx,σjy,σjz)(σjx,σjy,σjz)(\sigma^{x}_{j},\sigma^{y}_{j},\sigma^{z}_{j})\mapsto(\sigma^{x}_{j},-\sigma^{y}_{j},-\sigma^{z}_{j}) for odd jj.

The transverse susceptibility of the spin SS Ising chain is shown in Fig.1-Fig.3 for various values of the nearest-neighbor interaction J0J_{0} and the next-nearest-neighbor interaction J1J_{1}. The susceptibility is calculated using the analytic result (87) for S=1/2S=1/2, and calculated from numerically obtained u¯1ι{\bar{u}}_{1\iota} and uκ1u_{\kappa 1} for S1S\geq 1.

3 Low-temperature Limit

Next we physically consider the low-temperature behavior of the susceptibility. Let us follow the ground-state property of the system. Because (86) is a finite sum, the leading terms in (86) dominate the susceptibility at T0T\to 0. The phase diagram was investigated in Refs.\citen72KatsuraOhminami-\citen96Muraoka, and we know the ground state configurations in each phase. Let us consider the case J0>0J_{0}>0, because the susceptibility is invariant under the change J0J0J_{0}\mapsto-J_{0} and the result for J0<0J_{0}<0 and J1J_{1} is identical with that for |J0||J_{0}| and J1J_{1}.

When J1>J0/2J_{1}>-J_{0}/2, the ground state is ferromagnetic: sj=Ss_{j}=S for all jj, or sj=Ss_{j}=-S for all jj. The leading terms in (72) are

u¯11(eS2K0+S2K1)4u11+u¯1n(e(S)2K0+(S)2K1)4un1.\displaystyle{\bar{u}_{11}}(e^{S^{2}K_{0}+S^{2}K_{1}})^{4}u_{11}+{\bar{u}_{1n}}(e^{(-S)^{2}K_{0}+(-S)^{2}K_{1}})^{4}u_{n1}. (90)

Note that u¯1ιuρ1{\bar{u}_{1\iota}}u_{\rho 1} is the exactly renormalized Boltzmann weight with fixed spin configurations ι\iota and κ\kappa, and satisfies u¯1ιuρ1>0{\bar{u}_{1\iota}}u_{\rho 1}>0. Because U1U=EU^{-1}U=E, we find that ι=1nu¯1ιuι1=1\sum_{\iota=1}^{n}{\bar{u}_{1\iota}}u_{\iota 1}=1, and thus all the u¯1ιuκ1{\bar{u}_{1\iota}}u_{\kappa 1} are finite when ι=κ\iota=\kappa. The leading terms in (86) are found in (1,1)(1,1) and (n,n)(n,n)-elements of V2τpD(p)V2V^{2}\tau_{p}D^{(p)}V^{2}, with the condition mι+mκ=mμ+mν=2Sm_{\iota}+m_{\kappa}=m_{\mu}+m_{\nu}=2S and mι+mκ=mμ+mν=2Sm_{\iota}+m_{\kappa}=m_{\mu}+m_{\nu}=-2S, respectively. They come from the ground-state configurations together with the first excitations:

|Sj2|Sj1|Sj|Sj+1|Sj+2|Sj+3,\displaystyle|S\rangle_{j-2}|S\rangle_{j-1}|S\rangle_{j}|S\rangle_{j+1}|S\rangle_{j+2}|S\rangle_{j+3},
|Sj2|Sj1|S1j|Sj+1|Sj+2|Sj+3\displaystyle|S\rangle_{j-2}|S\rangle_{j-1}|S-1\rangle_{j}|S\rangle_{j+1}|S\rangle_{j+2}|S\rangle_{j+3} (91)

and

|Sj2|Sj1|Sj|Sj+1|Sj+2|Sj+3,\displaystyle|-S\rangle_{j-2}|-S\rangle_{j-1}|-S\rangle_{j}|-S\rangle_{j+1}|-S\rangle_{j+2}|-S\rangle_{j+3},
|Sj2|Sj1|S+1j|Sj+1|Sj+2|Sj+3,\displaystyle|-S\rangle_{j-2}|-S\rangle_{j-1}|-S+1\rangle_{j}|-S\rangle_{j+1}|-S\rangle_{j+2}|-S\rangle_{j+3}, (92)

i.e. obtained from the terms

1c\displaystyle\frac{1}{c} [\displaystyle\Big{[} sinh(2Sβ(J0+J1))2S(J0+J1)\displaystyle\frac{\sinh(-2S\beta(J_{0}+J_{1}))}{-2S(J_{0}+J_{1})}
×(Δ1(even)(eS2K0+S2K1)4+Δ2(even)(eS2K0+S2K1)2(eS(S1)K0+S(S1)K1)2)u¯11u11\displaystyle\hskip 34.14322pt\times\Big{(}\Delta_{1}^{({\rm even})}(e^{S^{2}K_{0}+S^{2}K_{1}})^{4}+\Delta_{2}^{({\rm even})}(e^{S^{2}K_{0}+S^{2}K_{1}})^{2}(e^{S(S-1)K_{0}+S(S-1)K_{1}})^{2}\Big{)}{\bar{u}_{11}}u_{11}
+sinh(2Sβ(J0+J1))2S(J0+J1)\displaystyle+\frac{\sinh(2S\beta(J_{0}+J_{1}))}{2S(J_{0}+J_{1})}
×(Δn(even)(e(S)2K0+(S)2K1)4+Δn1(even)(e(S)2K0+(S)2K1)2(e(S)(S+1)K0+(S)(S+1)K1)2)u¯1nun1\displaystyle\hskip 34.14322pt\times\Big{(}\Delta_{n}^{({\rm even})}(e^{(-S)^{2}K_{0}+(-S)^{2}K_{1}})^{4}+\Delta_{n-1}^{({\rm even})}(e^{(-S)^{2}K_{0}+(-S)^{2}K_{1}})^{2}(e^{(-S)(-S+1)K_{0}+(-S)(-S+1)K_{1}})^{2}\Big{)}{\bar{u}_{1n}}u_{n1}
+cosh(2Sβ(J0+J1))12S(J0+J1)\displaystyle+\frac{\cosh(-2S\beta(J_{0}+J_{1}))-1}{-2S(J_{0}+J_{1})}
×(iΔ1(odd)(eS2K0+S2K1)4+iΔ2(odd)(eS2K0+S2K1)2(eS(S1)K0+S(S1)K1)2)u¯11u11\displaystyle\hskip 34.14322pt\times\Big{(}i\Delta_{1}^{({\rm odd})}(e^{S^{2}K_{0}+S^{2}K_{1}})^{4}+i\Delta_{2}^{({\rm odd})}(e^{S^{2}K_{0}+S^{2}K_{1}})^{2}(e^{S(S-1)K_{0}+S(S-1)K_{1}})^{2}\Big{)}{\bar{u}_{11}}u_{11}
+cosh(2Sβ(J0+J1))12S(J0+J1)\displaystyle+\frac{\cosh(2S\beta(J_{0}+J_{1}))-1}{2S(J_{0}+J_{1})}
×(iΔn(odd)(e(S)2K0+(S)2K1)4+iΔn1(odd)(e(S)2K0+(S)2K1)2(e(S)(S+1)K0+(S)(S+1)K1)2)u¯1nun1],\displaystyle\hskip 34.14322pt\times\Big{(}i\Delta_{n}^{({\rm odd})}(e^{(-S)^{2}K_{0}+(-S)^{2}K_{1}})^{4}+i\Delta_{n-1}^{({\rm odd})}(e^{(-S)^{2}K_{0}+(-S)^{2}K_{1}})^{2}(e^{(-S)(-S+1)K_{0}+(-S)(-S+1)K_{1}})^{2}\Big{)}{\bar{u}_{1n}}u_{n1}\Big{]},

where Δ1(even)=Δn(even)=12(S(S+1)S2)=12S\Delta_{1}^{({\rm even})}=\Delta_{n}^{({\rm even})}=\frac{1}{2}(S(S+1)-S^{2})=\frac{1}{2}S, Δ2(even)=Δn1(even)=12(S(S+1)(S1)2)=12(3S1)\Delta_{2}^{({\rm even})}=\Delta_{n-1}^{({\rm even})}=\frac{1}{2}(S(S+1)-(S-1)^{2})=\frac{1}{2}(3S-1), iΔ1(odd)=12Si\Delta_{1}^{({\rm odd})}=\frac{1}{2}S, iΔ2(odd)=12(S1)i\Delta_{2}^{({\rm odd})}=\frac{1}{2}(S-1), iΔn1(odd)=12(S1)i\Delta_{n-1}^{({\rm odd})}=-\frac{1}{2}(S-1), and iΔn(odd)=12Si\Delta_{n}^{({\rm odd})}=-\frac{1}{2}S. Contributions from the first sum in (86) vanish at T0T\to 0. Short calculations yield that the susceptibility at low temperatures behaves

χ0(gμB)2121J0+J1(T0),\displaystyle\frac{\chi_{0}}{(g\mu_{\rm B})^{2}}\to\frac{1}{2}\frac{1}{J_{0}+J_{1}}\hskip 17.07182pt(T\to 0), (94)

which is independent of SS.

When J1<J0/2J_{1}<-J_{0}/2, the ground state configurations are obtained by iterations of |S|S|S|S|S\rangle|S\rangle|-S\rangle|-S\rangle. [33][32] Specifically from the configurations

|Sj2|Sj1|Sj|Sj+1|Sj+2|Sj+3,|Sj2|Sj1|Sj|Sj+1|Sj+2|Sj+3,\displaystyle|S\rangle_{j-2}|S\rangle_{j-1}|-S\rangle_{j}|-S\rangle_{j+1}|S\rangle_{j+2}|S\rangle_{j+3},\hskip 25.6073pt|S\rangle_{j-2}|-S\rangle_{j-1}|-S\rangle_{j}|S\rangle_{j+1}|S\rangle_{j+2}|-S\rangle_{j+3},
|Sj2|Sj1|Sj|Sj+1|Sj+2|Sj+3,|Sj2|Sj1|Sj|Sj+1|Sj+2|Sj+3,\displaystyle|-S\rangle_{j-2}|-S\rangle_{j-1}|S\rangle_{j}|S\rangle_{j+1}|-S\rangle_{j+2}|-S\rangle_{j+3},\hskip 14.22636pt|-S\rangle_{j-2}|S\rangle_{j-1}|S\rangle_{j}|-S\rangle_{j+1}|-S\rangle_{j+2}|S\rangle_{j+3},

and their excitations, the leading terms of (86) are generated in the diagonal elements of V2τpD(p)V2V^{2}\tau_{p}D^{(p)}V^{2}, and again straightforwrd calculations yield that the low-temperature limit of the susceptibility is

χ0(gμB)2121J1=121|J1|(T0).\displaystyle\frac{\chi_{0}}{(g\mu_{\rm B})^{2}}\to\frac{-1}{2}\frac{1}{J_{1}}=\frac{1}{2}\frac{1}{|J_{1}|}\hskip 17.07182pt(T\to 0). (96)

Note that the nearest-neighbor interactions cancel with each other and J0J_{0} does not appear in the ground state configurations.

When J1=J0/2J_{1}=-J_{0}/2, configurations obtained by arbitrary iterations of |S|S|S\rangle|S\rangle and |S|S|-S\rangle|-S\rangle belong to the ground state, e.g. configurations |S|S|S|S|S|S|S\rangle|S\rangle|S\rangle|S\rangle|S\rangle|S\rangle\cdots, and |S|S|S|S|S|S|S\rangle|S\rangle|-S\rangle|-S\rangle|S\rangle|S\rangle\cdots, and |S|S|S|S|S|S|S\rangle|S\rangle|S\rangle|S\rangle|-S\rangle|-S\rangle\cdots provide the ground state energy. The ground state is therefore highly degenerate, and for example the configurations

|Sj2|Sj1|Sj|Sj+1|Sj+2|Sj+3,|Sj2|Sj1|Sj|Sj+1|Sj+2|Sj+3,\displaystyle|S\rangle_{j-2}|S\rangle_{j-1}|-S\rangle_{j}|-S\rangle_{j+1}|-S\rangle_{j+2}|-S\rangle_{j+3},\hskip 17.07182pt|S\rangle_{j-2}|S\rangle_{j-1}|S\rangle_{j}|-S\rangle_{j+1}|-S\rangle_{j+2}|-S\rangle_{j+3}, (97)

generate some of the leading terms. They satisfy mι+mκ=0m_{\iota}+m_{\kappa}=0 and mμ+mν=0m_{\mu}+m_{\nu}=0, and hence the first term in (86) remains even when T0T\to 0, and we find the divergence coming from β\beta at low temperatures.

{acknowledgment}

This work was supported by JSPS KAKENHI Grant No. JP19K03668.

[Uncaptioned image]
Refer to caption
Figure 1: The transverse susceptibility of the spin SS Ising chain with the nearest-neighbor-interaction J0=1.0J_{0}=1.0, and with various values of the next-nearest-neighbor interaction J1J_{1}, where (a) S=1/2S=1/2, (b) S=1S=1, (c) S=3/2S=3/2, and (d) S=2S=2.
Refer to caption
Figure 2: The transverse susceptibility of the spin SS Ising chain with the next-nearest-neighbor-interaction J1=1.0J_{1}=1.0, and with various values of the nearest-neighbor interaction J0>0J_{0}>0, where (a) S=1/2S=1/2, (b) S=1S=1, (c) S=3/2S=3/2, and (d) S=2S=2.
Refer to caption
Figure 3: The transverse susceptibility of the spin SS Ising chain with the next-nearest-neighbor-interaction J1=1.0J_{1}=-1.0, and with various values of the nearest-neighbor interaction J0>0J_{0}>0, where (a) S=1/2S=1/2, (b) S=1S=1, (c) S=3/2S=3/2, and (d) S=2S=2. The curves for J0=0.0J_{0}=0.0 and for J0=0.2J_{0}=0.2 are very close to each other. The curve for J0=1.0J_{0}=1.0 is larger than the other four at low temperatures, and is smaller than the other four at high temperatures.

References

  • [1] M. E. Fisher, Physica 26, 618 (1960).
  • [2] M. E. Fisher, J. Math. Phys. 4, 124 (1963).
  • [3] S. Katsura, Phys. Rev. 127, 1508 (1962).
  • [4] P. Pfeuty, Ann. Phys. 57, 79 (1970).
  • [5] M.Suzuki, Prog. Theor. Phys. 46, 1337 (1971), see also K. Minami, EPL (Europhysics Letters) 108.3, 30001 (2014).
  • [6] K. Minami, J. Phys. Soc. Jpn. 85, 024003 (2016).
  • [7] K. Minami, Nucl. Phys. B 925, 144 (2017).
  • [8] Y. Yanagihara and K. Minami, Prog. Theor. Exp. Phys. 2020.11, 113A01 (2020).
  • [9] X-Y. Feng, G-M. Zhang, and T. Xiang, Phys. Rev. Lett. 98, 087204 (2007).
  • [10] H-D. Chen, and J. Hu, Phys. Rev. B 76, 193101 (2007).
  • [11] H-D. Chen, and Z. Nussinov, J. Phys. A: Math. Theor. 41, 075001 (2008).
  • [12] K. Minami, Nucl. Phys. B 939, 465 (2019).
  • [13] W. Cao, M. Yamazaki, and Y. Zheng Phys. Rev. B 106, 075150 (2022).
  • [14] K. Li and H. C. Po Phys. Rev. B 106, 115109 (2022).
  • [15] M. Suzuki, B. Tsujiyama and S. Katsura J. Math. Phys. 8, 124 (1967)
  • [16] T. Obokata and T. Oguchi J. Phys. Soc. Jpn 25, 322 (1968)
  • [17] J. F. Dobson J. Math. Phys. 10, 40 (1969)
  • [18] M. F. Thorpe and M.Thomsen, J. Phys. C: Solid State Phys. 16, L237 (1983)
  • [19] I. Chatterjee, J. Phys. C: Solid State Phys. 18, L1097 (1985)
  • [20] K. Minami, J. Phys. A: Math. Gen. 29, 6395 (1996).
  • [21] K. Minami, J. Phys. Soc. Jpn. 67, 2255 (1998).
  • [22] K. Minami, J. Phys. A: Math. Theor. 46, 505005 (2013).
  • [23] G. S. Joyce: Phys. Rev. Lett. 19, 581 (1967)
  • [24] C. J. Thompson: J. Math. Phys. 9, 241 (1968)
  • [25] T. Horiguchi: J. Phys. Soc. Jpn. 59, 3142 (1990)
  • [26] J. Stephenson: Can. J. Phys. 48, 1724 (1970)
  • [27] R. M. Hornreich, R. Liebmann, H. G. Schuster and W. Selke, Z. Physik B35, 91 (1979)
  • [28] I. Harada: J. Phys. Soc. Jpn. 52, 4099 (1983)
  • [29] T. Oguchi, J. Phys. Soc. Jpn. 20, 2236 (1965)
  • [30] F. A. Kassan-Ogly, Phase Transitions 74, 353 (2001)
  • [31] S. Katsura and M. Ohminami, J. Phys. A: Math. Gen. 5, 95 (1972)
  • [32] T. Morita and T. Horiguchi, Phys. Lett. 38A, 223 (1972)
  • [33] Y. Muraoka, K. Oda, J. W. Tucker and T. Idogaki, J. Phys. A: Math. Gen. 29, 949 (1996)