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

Entanglement entropy for scale-invariant states: universal finite-size scaling

Huan-Qiang Zhou Centre for Modern Physics, Chongqing University, Chongqing 400044, The People’s Republic of China    Qian-Qian Shi Centre for Modern Physics, Chongqing University, Chongqing 400044, The People’s Republic of China    Ian P. McCulloch School of Mathematics and Physics, The University of Queensland, St. Lucia, QLD 4072, Australia    Murray T. Batchelor Centre for Modern Physics, Chongqing University, Chongqing 400044, The People’s Republic of China Mathematical Sciences Institute, The Australian National University, Canberra ACT 2601, Australia
Abstract

A universal finite system-size scaling analysis of the entanglement entropy is presented for highly degenerate ground states arising from spontaneous symmetry breaking with type-B Goldstone modes in exactly solvable one-dimensional quantum many-body systems. These states appear to be scale-invariant, but not conformally invariant. Our findings are based on a physical argument, imposing three constraints on the entanglement entropy, in addition to further confirmation from an asymptotic analysis of the entanglement entropy for the SU(2){\rm SU}(2) spin-1/21/2 ferromagnetic states. The resulting universal scaling form is demonstrated for three fundamental models – the SU(2){\rm SU}(2) spin-ss Heisenberg ferromagnetic model, the SU(N+1){\rm SU}(N+1) ferromagnetic model, and the staggered SU(3){\rm SU}(3) spin-1 ferromagnetic biquadratic model. The results point towards a classification for distinct types of scale-invariant states, relevant to a complete classification of quantum states of matter.

.

I Introduction

Symmetry offers a powerful means to understand fundamental phenomena in physics and beyond symmetry ; andersonbook ; SSBbook . A specific example concerns scale invariance, which was believed to imply conformal invariance, as first speculated by Polyakov polyakov . Historically, this speculation led to the creation of conformal field theory cft . Since then much effort has been made in an attempt to clarify a deep connection between scale and conformal invariance morecf . However, a systematic investigation of scale-invariant, but not conformally invariant, quantum states, relevant to a classification of quantum critical points, is still lacking, even for one-dimensional quantum many-body systems.

A recent conceptual development has revealed that highly degenerate ground states arising from spontaneous symmetry breaking with type-B Goldstone modes in one-dimensional quantum many-body systems are scale-invariant FMGM ; golden ; LLspin1 . Somewhat surprisingly, a paradigmatic example is the SU(2){\rm SU}(2) Heisenberg ferromagnet, with the highly degenerate ground states being the familiar ferromagnetic states. As demonstrated, the highly degenerate ground states admit an exact singular value decomposition, thus exhibiting self-similarities in the ground state subspace. As a consequence, an abstract fractal is revealed, living in a Hilbert space, which may be characterized in terms of the fractal dimension, introduced earlier by Castro-Alvaredo and Doyon doyon for the SU(2){\rm SU}(2) Heisenberg ferromagnetic states in the context of a field-theoretic approach. The fractal dimension is identified with the number of the type-B Goldstone modes FMGM . In other words, there is a fascinating connection between scale-invariant states and the counting rule of the Goldstone modes FMGM . The latter in turn is established with the introduction of the Watanabe-Brauner matrix brauner-watanabe ; watanabe , as a result of an insightful observation made by Nambu nambu , thus leading to the classification of type-A and type-B Goldstone modes.

However, a natural question arises as to whether or not it is possible to distinguish scale-invariant states from conformally invariant states, given that the entanglement entropy scales logarithmically with the block size in the thermodynamic limit, with the prefactor being half the number of the type-B Goldstone modes FMGM for scale-invariant states and c/3c/3 for conformally invariant states, where cc is the central charge. One may anticipate that a finite system-size scaling behavior of the entanglement entropy is different for scale-invariant states and conformally invariant states, thus furnishing a sensible means for distinguishing them. Physically, such a distinct finite system-size scaling behavior is deeply rooted in the fact that highly degenerate ground states arising from spontaneous symmetry breaking with type-B Goldstone modes only concerns a scale transformation, instead of being a scale transformation plus the Lorentz boost for conformally invariant states.

This question is addressed here using an heuristic physical argument, which leads to three constraints imposed on the entanglement entropy: first, it must be identical for a block and for its environment; second, it should reproduce a logarithmic scaling function with the block size in the thermodynamic limit; third, as a function of both the system size and the block size, it must be homogeneous, with the order being one, to render consistency with a scale transformation. As it turns out, the universal scaling function for such a type of scale-invariant states is the most fundamental solution satisfying the three constraints, with further confirmation coming from an asymptotic analysis of the entanglement entropy for the SU(2){\rm SU}(2) spin-1/21/2 ferromagnetic states. To illustrate our scheme, we choose the SU(2){\rm SU}(2) spin-ss Heisenberg ferromagnetic model, the SU(N+1){\rm SU}(N+1) ferromagnetic model, and the staggered SU(3){\rm SU}(3) spin-1 ferromagnetic biquadratic model.

II Generalities

Consider a quantum many-body system on a lattice, described by Hamiltonian \mathscr{H}. Suppose the system is partitioned into a block BB and its environment EE, with the block consisting of nn (contiguous) lattice sites, with the other LnL-n lattice sites constituting the environment EE. If the symmetry group GG is spontaneously broken into HH, then the counting rule is established as NA+2NB=NBGN_{A}+2N_{B}=N_{BG} watanabe , where NAN_{A} and NBN_{B} are the numbers of the type-A and type-B GMs, and NBGN_{BG} is equal to the dimension of the coset space G/HG/H. In particular, for a quantum many-body system, no type-A GM survives in one spatial dimension, as follows from the Mermin-Wagner-Coleman theorem mwc . Hence the number of the type-A GMs NAN_{A} is always zero, if one is restricted to one-dimensional quantum many-body systems.

Here we present an heuristic argument to establish a universal scaling behavior of the entanglement entropy S(L,n)S(L,n) with LL and nn for a scale-invariant quantum state, which is one of the highly degenerate ground states for a quantum many-body system undergoing spontaneous symmetry breaking with type-B Goldstone modes. For completeness, an alternative derivation is also suggested for conformally invariant quantum states.

For both scale-invariant and conformally invariant states FMGM , the entanglement entropy S(n)S(n) exhibits a logarithmic scaling behavior with the block size nn in the thermodynamic limit (up to an additive constant):

S(n)log2n.S(n)\propto\log_{2}n. (1)

Here, S(n)S(n) denotes S(L,n)S(L,n), when LL\rightarrow\infty. Note that the prefactor is NB/2N_{B}/2 for a scale-invariant quantum state and c/3c/3 for a conformally invariant quantum state, where NBN_{B} denotes the number of the type-B Goldstone modes. We remark that, physically, both central charge cc and the number of the type-B Goldstone modes NBN_{B} count the number of gapless low-lying excitations for the quantum many-body system under investigation.

Keeping this fact in mind, one may assume that, when the system size LL and the block size nn are finite but large enough, the entanglement entropy S(L,n)S(L,n) scales with LL and nn as follows,

S(L,n)log2g(L,n).S(L,n)\propto\log_{2}g(L,n).\\ (2)

Here, g(L,n)g(L,n) is introduced as a universal scaling function of LL and nn, yet to be determined. Three constraints imposed on g(L,n)g(L,n) emerge from the following three physical requirements, to which the entanglement entropy S(L,n)S(L,n) is subject.

First, the fact that for any pure quantum state, the entanglement entropy for a block BB is identical to that for its environment EE nielsen implies that S(L,n)=S(L,Ln)S(L,n)=S(L,L-n). Hence, we have

g(L,n)=g(L,Ln).g(L,n)=g(L,L-n). (3)

In addition, S(L,n)S(L,n) must be monotonically increasing with nn until it reaches L/2L/2, where LL is assumed to be even.

Second, as the thermodynamic limit is approached, i.e., LL\rightarrow\infty, Eq. (2) reduces to Eq. (1). Therefore, one may expect that in this limit g(L,n)g(L,n) becomes nn:

limLg(L,n)=n.\lim_{L\rightarrow\infty}g(L,n)=n. (4)

Third, g(L,n)g(L,n) must be a first-order homogeneous function of LL and nn as a result of a scale transformation: nλnn\rightarrow\lambda\,n and LλLL\rightarrow\lambda\,L, given that it becomes nn in the thermodynamic limit [cf. Eq. (4)]. Thus

g(λL,λn)=λg(L,n).g(\lambda L,\lambda n)=\lambda\,g(L,n). (5)

As it turns out, the three constraints imposed on g(L,n)g(L,n) allow us to determine its specific forms. In fact, if one chooses λ=1/L\lambda=1/L, then we have

g(L,n)=Lg(1,n/L).g(L,n)=L\,g(1,n/L). (6)

This implies that g(L,n)g(L,n) is, up to a multiplicative factor LL, a function of n/Ln/L:

g(L,n)=Lk(n/L).g(L,n)=L\,k(n/L). (7)

Here, k(n/L)k(n/L) is a function of n/Ln/L, yet to be determined. Our task is therefore reduced to the determination of the function k(x)k(x), with x=n/Lx=n/L.

To proceed, let us first make an intriguing observation. Suppose that a function k(x)k(x) yields a solution satisfying the three constraints, then any function F(k(x))F(k(x)) of k(x)k(x) also yields a solution, subject to two conditions for F(x)F(x): (i) its constant term vanishes, F(0)=0F(0)=0 and (ii) its first-order derivative is nonzero, F(0)0F^{\prime}(0)\neq 0. This observation drastically simplifies our task, since we may restrict ourselves to searching for the most fundamental solution, corresponding to a plain scale transformation, in the sense that no other symmetry transformation is involved, if one excludes any internal symmetry.

For our purpose, it is plausible to assume that k(x)=u(x)u(1x)k(x)=u(x)u(1-x), with u(x)u(x) being a polynomial function of xx, in order to ensure that the first constraint (3) is satisfied. Then, the second constraint (4) implies that u(x)=xu(x)=x as the simplest choice. Hence, we have k(x)=x(1x)k(x)=x(1-x). This yields the simplest universal scaling function, thus constituting the most fundamental solution to the three constraints for scale-invariant states. Here, we remark that there exists another solution kc(x)k_{c}(x), which may be expressed in terms of elementary functions, if one assumes that kc(x)=v(x)+v(1x)k_{c}(x)=v(x)+v(1-x), with v(x)v(x) being a non-polynomial function of xx, in order to ensure that the first constraint (3) is satisfied. Then, the second constraint (4) implies that v(x)=1/(2iπ)exp(iπx)v(x)=1/(2i\pi)\exp(i\pi x). Hence, we have kc(x)=1/πsin(πx)k_{c}(x)=1/\pi\sin(\pi x). Indeed, kc(x)k_{c}(x) may be re-expressed as a function of k(x)k(x), namely kc(x)=1/πcos[π/214k(x)]k_{c}(x)=1/\pi\cos[\pi/2\sqrt{1-4k(x)}], as one may have anticipated from the observation above. Actually, kc(x)k_{c}(x) constitutes a universal scaling function for conformally invariant states, as predicted by Calabrese and Cardy cardy (more details on the universal scaling functions k(x)k(x) and kc(x)k_{c}(x) are given in Section A of the Supplemental Material (SM)).

In order to lend further support to our claim that k(x)k(x) is the desired universal scaling function for scale-invariant states arising from spontaneous symmetry breaking with type-B Goldstone modes, we may resort to exactly solvable quantum many-body systems. One may choose the SU(2){\rm SU}(2) spin-1/21/2 ferromagnetic Heisenberg model, with the number of the type-B Goldstone mode NBN_{B} being 1. For this model, the entanglement entropy S(L,n)S(L,n) has been derived by Popkov and Salerno popkov . As it turns out, our findings reproduce their results for the SU(2){\rm SU}(2) spin-1/21/2 ferromagnetic states (cf. Section B of the SM). For a conformally invariant state, one may choose the transverse-field spin-1/21/2 Ising chain at its critical point, with central charge c=1/2c=1/2. For this model, the entanglement entropy S(L,n)S(L,n) has been derived by Jin and Korepin korepin , thus allowing us to reproduce the prediction from conformal field theory cardy .

More precisely, for a scale-invariant state, with a given filling ff, the entanglement entropy Sf(L,n)S_{\!\!f}(L,n) takes the form

Sf(L,n)=NB2log2n(Ln)L+Sf0.S_{\!\!f}(L,n)=\frac{N_{B}}{2}\log_{2}\frac{n(L-n)}{L}+S_{\!\!f0}. (8)

On the other hand, for a conformally invariant state, the entanglement entropy S(L,n)S(L,n) takes the form

S(L,n)=c3log2(LπsinnπL)+S0.S(L,n)=\frac{c}{3}\log_{2}\left(\frac{L}{\pi}\sin{\frac{n\pi}{L}}\right)+S_{\!0}. (9)

Note that an additive nonuniversal constant Sf0S_{\!\!f0} or S0S_{\!0} has been introduced for a scale-invariant or conformally invariant state, respectively.

Three remarks are in order. First, it is necessary to introduce the filling ff to indicate distinct degenerate ground states arising from spontaneous symmetry breaking with type-B Goldstone modes, in contrast to conformally invariant states. Note that, generically, ff refers to a set of fillings f1,f2,,frf_{1},f_{2},\ldots,f_{r}, with rr being the rank of a symmetry group for a quantum many-body system. Second, the block does not necessarily consist of contiguous sites for a scale-invariant state, which appears as one of highly degenerate ground states in a quantum many-body system, as already discussed for the entanglement entropy in the thermodynamic limit FMGM . Meanwhile, the scaling relation (8) is left intact as the boundary conditions vary from periodic to open. In contrast, a conformally invariant state appears to be a non-degenerate ground state for a quantum many-body system at criticality, and the scaling relation (9) does depend on the boundary conditions adopted cardy ; korepin1 ; zhou . Third, other choices for the universal scaling function g(L,n)g(L,n) are possible for distinct types of scale-invariant states, but remain to be clarified.

III Three illustrative examples

III.1 The SU(2){\rm SU}(2) spin-ss ferromagnetic model

Consider the SU(2){\rm SU}(2) spin-ss ferromagnetic Heisenberg model with nearest-neighbor interaction, described by the Hamiltonian

=j=1LSjSj+1,\mathscr{H}=-\sum_{j=1}^{L}\textbf{S}_{j}\cdot\textbf{S}_{j+1}, (10)

where Sj=(Sjx,Sjy,Sjz)\textbf{S}_{j}=(S^{x}_{j},S^{y}_{j},S^{z}_{j}), and SjxS^{x}_{j}, SjyS^{y}_{j}, SjzS^{z}_{j} are the spin operators for spin ss at the jj-th site. Note that for all three illustrative models the sum over jj is taken from 1 to L1L-1 for open boundary conditions (OBCs), and from 1 to LL for periodic boundary conditions (PBCs). The spin-ss Heisenberg model admits exact solutions as far as its ground state subspace is concerned, though it is only exactly solvable by means of the Bethe ansatz for s=1/2s=1/2. The degenerate ground states arise from spontaneous symmetry breaking: SU(2)U(1){\rm SU}(2)\rightarrow{\rm U}(1), thus resulting in one type-B Goldstone mode, so NB=1N_{B}=1. We remark that the degenerate ground states are identical under both OBCs and PBCs, and span an irreducible representation for the symmetry group SU(2){\rm SU}(2), with the dimension being 2sL+12sL+1.

III.2 The SU(N+1){\rm SU}(N+1) ferromagnetic model

The SU(N+1){\rm SU}(N+1) ferromagnetic model is described by the Hamiltonian

=jPjj+1,\mathscr{H}=-\sum_{j}P_{j\;j+1}, (11)

where PP is the permutation operator P=u,v=1N+1euvevuP=\sum_{u,v=1}^{N+1}e_{uv}\otimes e_{vu}, where euv=|uv|e_{uv}=|u\rangle\langle v|, with |u|u\rangle and |v|v\rangle being the uu-th and vv-th states in an orthonormal basis. Physically, the permutation operator PP may be realized in terms of the spin-ss operators S=(Sx,Sy,Sz)\textbf{S}=(S_{x},S_{y},S_{z}), with N=2sN=2s:

P=q=02s(1)2s+qmq2s2(SS)m(m+1)+2s(s+1)q(q+1)m(m+1).P=\sum_{q=0}^{2s}(-1)^{2s+q}\prod_{m\neq q}^{2s}\frac{2(\textbf{S}\otimes\textbf{S})-m(m+1)+2s(s+1)}{q(q+1)-m(m+1)}.

The SU(N+1){\rm SU}(N+1) ferromagnetic model is exactly solvable by means of the nested Bethe ansatz sutherland . The degenerate ground states arise from spontaneous symmetry breaking: SU(N+1)SU(N)×U(1){\rm SU}(N+1)\rightarrow{\rm SU}(N)\times{\rm U}(1) successively, thus resulting in NN type-B Goldstone modes, so NB=NN_{B}=N. We remark that the degenerate ground states are identical under both OBCs and PBCs, and span an irreducible representation for the symmetry group SU(N+1){\rm SU}(N+1), with the dimension being the combinatorial number CL+NNC_{L+N}^{N}.

III.3 The staggered SU(3){\rm SU}(3) spin-1 ferromagnetic biquadratic model

The staggered SU(3){\rm SU}(3) spin-1 ferromagnetic biquadratic model is described by the Hamiltonian

=j(SjSj+1)2.\mathscr{H}=\sum_{j}{\left(\textbf{S}_{j}\cdot\textbf{S}_{j+1}\right)^{2}}. (12)

Here Sj=(Sjx,Sjy,Sjz)\textbf{S}_{j}=(S^{x}_{j},S^{y}_{j},S^{z}_{j}) is the spin-1 operator at site jj. The model (12) is exactly solvable barber . Indeed, it constitutes (up to an additive constant) a representation of the Temperley-Lieb algebra tla ; baxterbook ; martin , and thus follows from a solution to the Yang-Baxter equation baxterbook ; sutherlandb ; mccoy . Note that it is peculiar, in the sense that the ground states are highly degenerate, exponential with the system size LL, thus leading to non-zero residual entropy spins , in sharp contrast to the SU(2){\rm SU}(2) spin-ss ferromagnetic model (10) and the SU(N+1){\rm SU}(N+1) ferromagnetic model (11). Remarkably, the degenerate ground states for this model arise from spontaneous symmetry breaking SU(3)U(1)×U(1){\rm SU}(3)\rightarrow{\rm U}(1)\times{\rm U}(1), with the number of the type-B Goldstone modes NB=2N_{B}=2, and the ground state degeneracies constitute two Fibonacci-Lucas sequences golden (also cf. spins ; saleur ) under OBCs and PBCs.

IV Universal finite-size scaling for the entanglement entropy Sf(L,n)S_{\!\!f}(L,n)

For the SU(2){\rm SU}(2) spin-ss ferromagnetic Heisenberg model, the entanglement entropy S(L,n,M)S(L,n,M) for the highly degenerate ground states |L,M|L,M\rangle have been derived in Ref. FMGM , with ff being the filling f=M/Lf=M/L. For convenience, the explicit expression for S(L,n,M)S(L,n,M) has been collected in Section C of the SM.

We plot the entanglement entropy S(L,n,M)S(L,n,M) vs nn against the universal finite-size scaling Sf(L,n)S_{\!\!f}(L,n) vs nn in Fig. 1 for s=1/2s=1/2, 11, 3/23/2 and 22. Our numerical data for S(L,n,M)S(L,n,M) fall on the curve Sf(L,n)S_{\!\!f}(L,n), when nn ranges from 10 to 90, with the relative errors being less than 1.5%1.5\%. Here and hereafter, we have regarded Sf(L,n)S_{\!\!f}(L,n) as a function of nn for fixed LL and ff.

Refer to caption
Figure 1: The entanglement entropy S(L,n,M)S(L,n,M) vs nn against Sf(L,n)S_{\!\!f}(L,n) vs nn for the highly degenerate ground states in the SU(2){\rm SU}(2) spin-ss ferromagnetic Heisenberg model, with the system size L=100L=100, for (a) s=1/2s=1/2, (b) s=1s=1, (c) s=3/2s=3/2 and (d) s=2s=2. The filling factor ff is chosen to be f=1/4f=1/4 with in each case L=100L=100. The best fitting yields (a) Sf0=0.834S_{\!\!f0}=0.834, (b) Sf0=0.934S_{\!\!f0}=0.934, (c) Sf0=0.963S_{\!\!f0}=0.963, and (d) Sf0=1.149S_{\!\!f0}=1.149, with the relative errors being less than 1.5%1.5\%, when nn ranges from 10 to 90.

For the SU(N+1){\rm SU}(N+1) ferromagnetic model, the entanglement entropy S(L,n,M1,,MN)S(L,n,M_{1},\ldots,M_{N}) for the highly degenerate ground state |L,M1,M2,,MN|L,M_{1},M_{2},\ldots,M_{N}\rangle has been derived in Ref. FMGM , with the filling factors f1=M1/Lf_{1}=M_{1}/L, f2=M2/Lf_{2}=M_{2}/L and f3=M3/Lf_{3}=M_{3}/L. The explicit expression for S(L,n,M1,,MN)S(L,n,M_{1},\ldots,M_{N}) has been collected in Section C of the SM.

Refer to caption
Figure 2: (a) The entanglement entropy S(L,n,M1,M2)S(L,n,M_{1},M_{2}) vs nn against Sf(L,n)S_{\!\!f}(L,n) vs nn for the highly degenerate ground states in the SU(3){\rm SU}(3) spin-11 model. The filling factors are chosen to be f1=1/4f_{1}=1/4 and f2=1/4f_{2}=1/4. (b) The entanglement entropy S(L,n,M1,M2,M3)S(L,n,M_{1},M_{2},M_{3}) vs nn against Sf(L,n)S_{\!\!f}(L,n) vs nn for the highly degenerate ground states in the SU(4){\rm SU}(4) spin-3/23/2 model. The filling factors are chosen to be f1=1/4f_{1}=1/4, f2=1/4f_{2}=1/4 and f3=1/4f_{3}=1/4. For both figures the system size L=100L=100. The best fitting yields (a) Sf0=1.576S_{\!f0}=1.576 and (b) Sf0=2.102S_{\!f0}=2.102, with the relative errors being less than 1%1\%, when nn ranges from 10 to 90.

We plot the entanglement entropy S(L,n,M1,M2)S(L,n,M_{1},M_{2}) and the entanglement entropy S(L,n,M1,M2,M3)S(L,n,M_{1},M_{2},M_{3}) vs nn, with the selected fillings (a) f1=1/4f_{1}=1/4 and f2=1/4f_{2}=1/4 and (b) f1=1/4f_{1}=1/4, f2=1/4f_{2}=1/4 and f3=1/4f_{3}=1/4, respectively, against the universal finite-size scaling Sf(L,n)S_{\!\!f}(L,n) vs nn, in Fig. 2, with NB=2N_{B}=2 and NB=3N_{B}=3. Our numerical data for S(L,n,M1,M2)S(L,n,M_{1},M_{2}) and S(L,n,M1,M2,M3)S(L,n,M_{1},M_{2},M_{3}) fall on the curve Sf(L,n)S_{\!\!f}(L,n), when nn ranges from 10 to 90, with the relative errors being less than 1%1\%.

Finally for the staggered SU(3){\rm SU}(3) spin-1 ferromagnetic biquadratic model, the entanglement entropy S(L,n,M1,M2)S(L,n,M_{1},M_{2}) and the entanglement entropy S(L,n,M2,M3)S(L,n,M_{2},M_{3}) for the two degenerate ground states |L,M1,M22|L,M_{1},M_{2}\rangle_{2} and |L,M2,M34|L,M_{2},M_{3}\rangle_{4} have been derived in Ref. golden , with the filling factors f1=M1/Lf_{1}=M_{1}/L, f2=M2/Lf_{2}=M_{2}/L and f3=M3/Lf_{3}=M_{3}/L. Here the explicit expressions for S(L,n,M1,M2)S(L,n,M_{1},M_{2}) and S(L,n,M2,M3)S(L,n,M_{2},M_{3}) have been collected in Section C of the SM.

For this model we plot the entanglement entropy S(L,n,M1,M2)S(L,n,M_{1},M_{2}) and S(L,n,M2,M3)S(L,n,M_{2},M_{3}) vs nn against the universal finite-size scaling Sf(L,n)S_{\!\!f}(L,n) vs nn in Fig. 3, with the selected filling factors (a) f1=1/4f_{1}=1/4 and f2=1/4f_{2}=1/4, and (b) f2=1/4f_{2}=1/4 and f3=1/8f_{3}=1/8. Our numerical data for S(L,n,M1,M2)S(L,n,M_{1},M_{2}) and S(L,n,M2,M3)S(L,n,M_{2},M_{3}) are seen to fall on the curve Sf(L,n)S_{\!\!f}(L,n), when nn ranges from 10 to L10L-10, with the relative errors being less than 1%1\%.

Refer to caption
Figure 3: The entanglement entropy S(L,n,M1,M2)S(L,n,M_{1},M_{2}) vs nn against Sf(L,n)S_{\!\!f}(L,n) vs nn for the highly degenerate ground states from the highest weight state in the staggered SU(3){\rm SU}(3) spin-1 ferromagnetic biquadratic model, with the system size L=100L=100, when the fillings f1f_{1} and f2f_{2} are chosen to be f1=1/4f_{1}=1/4 and f2=1/4f_{2}=1/4. (b) The entanglement entropy S(L,n,M2,M3)S(L,n,M_{2},M_{3}) vs nn against Sf(L,n)S_{f}(L,n) vs nn for the highly degenerate ground states from a generalized highest weight state in the staggered SU(3){\rm SU}(3) spin-1 ferromagnetic biquadratic model, with the system size L=120L=120, when the fillings f2f_{2} and f3f_{3} are chosen to be f2=1/4f_{2}=1/4 and f3=1/8f_{3}=1/8. The best fitting yields (a) Sf0=1.308S_{\!\!f0}=1.308 and (b) Sf0=0.743S_{\!\!f0}=0.743, respectively, with the relative errors being less than 1%1\%, when nn ranges from 10 to L10L-10.

V Summary and outlook

A universal finite-size scaling behavior of the entanglement entropy has been shown for scale-invariant, but not conformally invariant, degenerate ground states, which arise from spontaneous symmetry breaking with type-B Goldstone modes in quantum many-body systems. As it turns out, scale-invariant states exhibit a universal finite system-size scaling behavior, distinct from conformally invariant states, given both are reduced to a logarithmic scaling function in the thermodynamic limit. This fact enables us to distinguish scale-invariant states from conformally invariant states. It follows that scale invariance does not necessarily imply conformal invariance, in contrast to the speculation made by Polyakov polyakov , even for the familiar SU(2){\rm SU}(2) spin-1/21/2 ferromagnetic states – a paradigmatic example for spontaneous symmetry breaking with type-B Goldstone modes. Our findings shed further light on a complete classification of quantum phase transitions and quantum states of matter entropy . In this sense, the possibility for formulating a fully-fledged theory of scale-invariant states remains largely unexplored.

Instead, a simple but heuristic physical argument has been proposed, which imposes three given constraints on the entanglement entropy. This in turn makes it possible to classify distinct types of scale-invariant states. Our results clearly suggest that highly degenerate ground states arising from spontaneous symmetry breaking with type-B Goldstone modes fall into the simplest category for all possible scale-invariant states, featuring the most fundamental universal scaling function. All other possible universal scaling functions might be expressed in terms of this fundamental universal scaling function, as illustrated for conformally invariant states.

VI Acknowledgements

We thank John Fjaerestad for enlightening discussions.

References

  • (1) H. Weyl, Symmetry (Princeton University Press, Princeton, NJ, 1952); E. P. Wigner, Symmetries and Reflections (Indiana University Press, Bloomington, Indiana, 1967).
  • (2) P. W. Anderson, Basic Notions in Condensed Matter Physics (CRC Press, Boca Raton, 1994).
  • (3) H. Arodz, J. Dziarmaga, and W. H. Zurek, eds. Patterns of Symmetry Breaking (NATO Science Series II: Mathematics, Physics and Chemistry, 2003).
  • (4) A. M. Polyakov, JETP Lett. 12, 381 (1970).
  • (5) A. A. Belavin, A. M. Polyakov, and A. B. Zamolodchikov, Nucl. Phys. B 2, 333 (1984); see also P. D. Francesco, P. Mathieu, and D. Sénéchal, Conformal Field Theory (Springer, Berlin, 1997).
  • (6) H. Osborn, Nucl. Phys. B 363, 486 (1991); A. LeClair, J. M. Roman and G. Sierra, Nucl. Phys. B 700, 407 (2004); A. LeClair and G. Sierra, J. Stat. Mech. P08004 (2004); Z. Komargodski and A. Schwimmer, J. High Energy Phys. 12, 099 (2011); J.-F. Fortin, B. Grinstein, and A. Stergiou, J. High Energy Phys. 01, 184 (2013); Y. Nakayama, Phys. Rep. 569, 1 (2015); A. Dymarsky, Z. Komargodski, A. Schwimmer, and S. Theisen, J. High Energy Phys. 10, 171 (2015); B. Delamotte, M. Tissier, and N. Wschebor, Phys. Rev. E 93, 012144 (2015); A. Dymarsky, K. Farnsworth, Z. Komargodski, M. A. Luty, and V. Prilepina, J. High Energy Phys. 02, 099 (2016); A. Dymarsky and A. Zhiboedov, J. Phys. A: Math. Theor. 48, 41FT01 (2015); M. Hogervorst, S. Rychkov, and B. C. van Rees, Phys. Rev. D 93, 125025 (2016).
  • (7) Q.-Q. Shi, Y.-W. Dai, H.-Q. Zhou, and I. P. McCulloch, arXiv: 2201.01071 (2022).
  • (8) H.-Q. Zhou, Q.-Q. Shi, I. P. McCulloch, and M. T. Batchelor, arXiv: 2302.13126 (2023).
  • (9) Q.-Q. Shi, Y.-W. Dai, S.-H. Li, and H.-Q. Zhou, arXiv: 2204.05692 (2022).
  • (10) O. A. Castro-Alvaredo and B. Doyon, Phys. Rev. Lett. 108, 120401 (2012).
  • (11) H. Watanabe and T. Brauner, Phys. Rev. D 84, 125013 (2011).
  • (12) H. Watanabe and H. Murayama, Phys. Rev. Lett. 108, 251602 (2012); H. Watanabe and H. Murayama, Phys. Rev. X 4, 031057 (2014).
  • (13) Y. Nambu, J. Stat. Phys. 115, 7 (2004).
  • (14) N. D. Mermin and H. Wagner, Phys. Rev. Lett. 17, 1133 (1966); S. R. Coleman, Commun. Math. Phys. 31, 259 (1973).
  • (15) M.A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information (Cambridge University Press, Cambridge, UK, 2000).
  • (16) P. Calabrese and J. Cardy, J. Stat. Mech. P06002 (2004).
  • (17) V. Popkov and M. Salerno, Phys. Rev. A 71, 012301 (2005).
  • (18) B.-Q. Jin and V. E. Korepin, J. Stat. Phys. 116, 79-95 (2004).
  • (19) V. E. Korepin, Phys. Rev. Lett. 92, 096402 (2004).
  • (20) H.-Q. Zhou, T. Barthel, J. O. Fjaerestad, and U. Schollwoeck, Phys. Rev. A 74, 050305(R) (2006).
  • (21) B. Sutherland, Phys. Rev. B 12, 3795 (1975).
  • (22) M. N. Barber and M. T. Batchelor, Phys. Rev. B40, 4621 (1989); M. T. Batchelor and M. N. Barber, J. Phys. A: Math. Gen. 23, L15 (1990); A. Klümper, Europhys. Lett. 9, 815 (1989).
  • (23) H. N. V. Temperley and E. Lieb, Proc. R. Soc. Lond. A322, 251 (1971).
  • (24) R. J. Baxter, Exactly Solved Models in Statistical Mechanics (Academic Press, London, 1982).
  • (25) P. Martin, Potts Models and Related Problems in Statistical Mechanics (World Scientific, 1991).
  • (26) B. Sutherland, Beautiful Models, 70 Years of Exactly Solved Quantum Many-Body Problems (World Scientific, Singapore, 2004).
  • (27) B. M. McCoy, Advanced Statistical Physics (Oxford University Press, Oxford, 2009).
  • (28) B. Aufgebauer and A. Klümper, J. Stat. Mech. P05018 (2010).
  • (29) N. Read and H. Saleur, Nucl. Phys. B, 777, 263-315 (2007); Y.-T. Oh, H. Katsura, H.-Y. Lee, and J. H. Han, Phys. Rev. B 96, 165126 (2017).
  • (30) H.-Q. Zhou, Q.-Q. Shi, and Y.-W. Dai, Entropy, 24, 1306 (2022).

Supplementary Material

VI.1 A: Derivation of the distinct universal scaling functions k(x)k(x) and kc(x)k_{c}(x)

As mentioned in the main text, the observation that if a function k(x)k(x) yields a solution to the three given constraints, then any function F(k(x))F(k(x)) of k(x)k(x) also yields a solution, subject to the two conditions for F(x)F(x): its constant term vanishes, F(0)=0F(0)=0, and its first-order derivative is nonzero, F(0)0F^{\prime}(0)\neq 0. Hence, one may search for the most fundamental solution, which is supposed to be a universal scaling function solely as a result of a scale transformation.

Keeping this fact in mind, we may seek for a universal scaling function k(x)k(x) of the form

k(x)=u(x)u(1x).k(x)=u(x)u(1-x). (S1)

This guarantees that the constraint (3) is satisfied. Here, u(x)u(x) is a polynomial function of xx, without a constant term, i.e., u(x)=β=1mνβxβu(x)=\sum_{\beta=1}^{m}\nu_{\beta}x^{\beta}, with mm being an integer, and ν1,ν2,,νm\nu_{1},\nu_{2},\ldots,\nu_{m} being real numbers yet to be determined. The constraint (4) yields ν1β=1mνβ=1\nu_{1}\sum_{\beta=1}^{m}\nu_{\beta}=1. As a consequence, we have to distinguish the two situations: (1) ν1=1\nu_{1}=1 and (2) ν11\nu_{1}\neq 1.

If ν1=1\nu_{1}=1, then we have β=2mνβ=0\sum_{\beta=2}^{m}\nu_{\beta}=0. Hence, the simplest choice for u(x)u(x) is u(x)=xu(x)=x, which amounts to stating that ν1=1\nu_{1}=1 and νβ=0\nu_{\beta}=0 (β=2,,m\beta=2,\ldots,m). Accordingly, we have

k(x)=x(1x).k(x)=x(1-x). (S2)

This yields the simplest universal scaling function, which in turn constitutes the most fundamental solution to the three constraints for scale-invariant states.

Actually, this is the only solution available, as long as u(x)u(x) is a polynomial function of xx. A proof for our claim is based on mathematical induction. In fact, for m=1m=1, obviously we have ν1=1\nu_{1}=1, and for m=2m=2, it is straightforward to see that ν1=1\nu_{1}=1 and ν2=0\nu_{2}=0. Suppose our claim is valid for mm. Then, for m+1m+1, one might resort to the following lemma: if u(x)u(x) yields a solution k(x)k(x) to the three constraints, then u~(x)=(u(x)x2)/(1x)\tilde{u}(x)=(u(x)-x^{2})/(1-x) also yields a solution. Indeed, the mapping from u(x)u(x) to u~(x)\tilde{u}(x) turns a polynomial function u(x)u(x) of xx, with order m+1m+1, into a polynomial function u~(x)\tilde{u}(x) of xx, with order mm, due to the fact that u(x)x2u(x)-x^{2} may be factorized into 1x1-x times a polynomial function of xx, with order mm.

If ν1=t1\nu_{1}=t\neq 1, then we have β=2mνβ=t1t\sum_{\beta=2}^{m}\nu_{\beta}=t^{-1}-t, with tt being a nonzero real number. In this situation, we have different possible choices for a universal scaling function k(x)k(x), as long as it is monotonically increasing with xx, when xx varies from x=0x=0 to x=1/2x=1/2. In fact, it even makes sense to take the limit mm\rightarrow\infty, with properly chosen tt and νβ\nu_{\beta}. In particular, if we choose t=π/2t=\sqrt{\pi/2}, νβ=(1)δ(π/2)2δ+1/2/[(2δ+1)××1]\nu_{\beta}=(-1)^{\delta}(\pi/2)^{2\delta+1/2}/[(2\delta+1)\times\cdots\times 1] for β=2δ+1\beta=2\delta+1 and νβ=0\nu_{\beta}=0 for β=2δ\beta=2\delta, with δ=0,1,2,\delta=0,1,2,\ldots. Then we have u(x)=2/πsin(πx/2)u(x)=\sqrt{2/\pi}\sin(\pi x/2). Thus we are led to another universal scaling function kc(c)k_{c}(c), namely

kc(x)=sin(πx)π.k_{c}(x)=\frac{\sin(\pi x)}{\pi}. (S3)

This is precisely the universal scaling function, as predicted by Calabrese and Cardy cardy-S , for conformally invariant states.

Two remarks are in order. First, kc(x)k_{c}(x) not only admits a factorization in a multiplicative way, but also splits in an additive way. This might be attributed to the fact that the Lorentz invariance is part of the symmetry group, in addition to a scale transformation, for conformally invariant states. In fact, kc(x)k_{c}(x) may be constructed as follows. In order to satisfy the constraint (3), we assume that kc(x)=v(x)+v(1x)k_{c}(x)=v(x)+v(1-x), with v(x)v(x) being a non-polynomial function of xx. As a possible trial solution, we choose v(x)=bexp(ax)v(x)=b\exp(ax), with aa and bb being real numbers yet to be determined. Hence, we have

kc(x)=2bexp(a/2)cosh[a(x12)].k_{c}(x)=2b\exp(a/2)\cosh[a(x-\frac{1}{2})]. (S4)

When x=0x=0, kc(x)k_{c}(x) must vanish. Thus, we have a=iπa=i\pi. Substituting a=iπa=i\pi back into (S4), we have kc(x)=2ibsin(πx)k_{c}(x)=2ib\sin(\pi x). To ensure that the constraint (4) is satisfied, we have kc(x)xk_{c}(x)\rightarrow x for x0x\rightarrow 0. This implies that b=1/(2iπ)b=1/(2i\pi). Hence, we are led to the same universal scaling function as that in Eq. (S3). Second, as follows from Eq. (S2), xx may be re-expressed in terms of k(x)k(x) as an inverse function, i.e., x=[1±14k(x)]/2x=[1\pm\sqrt{1-4k(x)}]/2. Hence, kc(x)k_{c}(x) turns out to be a function of k(x)k(x), with kc(x)=1/πcos[π/214k(x)]k_{c}(x)=1/\pi\cos[\pi/2\sqrt{1-4k(x)}]. This provides an illustrative example for the observation discussed in the main text.

The justification for k(x)k(x) as the universal scaling function for highly degenerate ground states arising from spontaneous symmetry breaking with type-B Goldstone modes lies in the fact that only a scale transformation is involved. Further confirmation comes from an asymptotic analysis of the entanglement entropy for the SU(2){\rm SU}(2) spin-1/21/2 ferromagnetic states, as discussed in the next Section.

VI.2 B: Entanglement entropy for the SU(2){\rm SU}(2) spin-1/21/2 ferromagnetic states

Alternatively, the scaling relations of the entanglement entropy Sf(L,n)S_{\!\!f}(L,n) for both scale-invariant and conformally invariant states, as presented in Eq.(8) and Eq.(9) in the main text, respectively, may be justified by extending a heuristic argument in Ref. FMGM-S in the thermodynamic limit.

Given that the number of the type-B Goldstone modes and central charge measure the ability for a system to react when a length scale is present and count the number of low-lying gapless excitations, we anticipate that, for scale-invariant states, Sf(L,n)S_{f}(L,n) takes the form

Sf(L,n)=NB2log2g(L,n)+Sf0,S_{\!\!f}(L,n)=\frac{N_{B}}{2}\log_{2}g(L,n)+S_{\!\!f0}, (S5)

where g(L,n)=Lk(n/L)g(L,n)=L\,k({n}/{L}), with k(n/L)k(n/L) being a function of n/Ln/L, yet to be determined. This may be determined from an exactly solvable quantum many-body model – the SU(2){\rm SU}(2) spin-1/21/2 ferromagnetic model, with NB=1N_{B}=1.

For the SU(2){\rm SU}(2) spin-1/21/2 ferromagnetic states, the entanglement entropy S(L,n,M)S(L,n,M) for a degenerate ground state |L,M|L,M\rangle has been derived in Refs. popkov-S ; FMGM-S . It takes the form

S(L,n,M)=k=1nΛ(L,n,k,M)log2Λ(L,n,k,M),S(L,n,M)=-\sum_{k=1}^{n}\Lambda(L,n,k,M)\log_{2}\Lambda(L,n,k,M), (S6)

where the eigenvalues Λ(L,n,k,M)\Lambda(L,n,k,M) of the reduced density matrix ρL(n,M)\rho_{L}(n,M) are popkov-S ; FMGM-S

Λ(L,n,k,M)=CnkCLnMkCLM.\Lambda(L,n,k,M)=\frac{C_{n}^{k}C_{L-n}^{M-k}}{C_{L}^{M}}. (S7)

The eigenvalues of the reduced density matrix, for large enough LL, may be approximated as a result of the Stirling approximation: m!=mmexp(m)2πmm!=m^{m}\exp(-m)\sqrt{2\pi m},

Λ(L,n,k,M)=Cnkpkqnk,\Lambda(L,n,k,M)=C_{n}^{k}p^{k}q^{n-k}, (S8)

where p=M/Lp=M/L and q=1pq=1-p.

For large enough n1n\gg 1, the dominant contribution to the sum (S6) comes from the eigenvalues Λ(L,n,k,M)\Lambda(L,n,k,M) with large kk’s. Indeed, the binomial coefficients in (S8) with 0<p<10<p<1 may be replaced by the normal distribution (cf. Ref. Allen-S ):

Cnmpmqnm12πnpqexp((mnp)22npq), npq1,C_{n}^{m}p^{m}q^{n-m}\approx\frac{1}{\sqrt{2\pi npq}}\exp\left(-\frac{(m-np)^{2}}{2npq}\right),\text{ \ \ }npq\gg 1, (S9)

Hence, the eigenvalues in Eq. (S8) become

Λ(L,n,k,M)\displaystyle\Lambda(L,n,k,M) =CknpkqnkCMkLnpMkqLnM+kCMLpMqLM\displaystyle=\frac{C_{k}^{n}p^{k}q^{n-k}C_{M-k}^{L-n}p^{M-k}q^{L-n-M+k}}{C_{M}^{L}p^{M}q^{L-M}}
1n12παexp((knp)22α),\displaystyle\approx\frac{1}{n}\frac{1}{\sqrt{2\pi\alpha}}\exp\left(-\frac{(\frac{k}{n}-p)^{2}}{2\alpha}\right), (S10)

where α=pq(Ln)/nL\alpha=pq(L-n)/nL.

Substituting (S10) into (S6) and replacing the sum with an integral, S(L,n,M)S(L,n,M) takes the form

S(L,n,M)01R(log2Rn)𝑑x,\displaystyle S(L,n,M)\approx\int_{0}^{1}R\left(\log_{2}{\frac{R}{n}}\right)dx,
R=12παexp((xp)22α).\displaystyle\qquad R=\frac{1}{\sqrt{2\pi\alpha}}\exp\left(-\frac{(x-p)^{2}}{2\alpha}\right).

For large enough nn, it yields

S(L,n,M)12log2(2πepq)+12log2n(Ln)L.S(L,n,M)\approx\frac{1}{2}\log_{2}(2\pi epq)+\frac{1}{2}\log_{2}{\frac{n(L-n)}{L}}. (S11)

We remark that (S11) is asymptotically valid for npq1npq\gg 1 and becomes exact in the limit npqnpq\rightarrow\infty.

Comparing Eq. (S11) with Eq. (S5), we are led to conclude that k(x)=x(1x)k(x)=x(1-x). Hence the scaling relation of the entanglement entropy Sf(L,n)S_{\!\!f}(L,n) with LL and nn in Eq. (8) for scale-invariant states is established.

This argument also works for the scaling relation of the entanglement entropy S(L,n)S(L,n) with LL and nn in Eq. (9) for conformally invariant states, since the exact results are available for the transverse-field spin-1/21/2 Ising model at its critical point, with central charge c=1/2c=1/2 korepin-S .

VI.3 C: Eigenvalues of the reduced density matrices for the highly degenerate ground states

Here we collect the explicit expressions for the eigenvalues of the reduced density matrices, which have been derived in Refs. FMGM-S ; golden-S for the highly degenerate ground states of the three models under consideration.

For the SU(2){\rm SU}(2) spin-ss ferromagnetic Heisenberg model, the entanglement entropy S(L,n,M)S(L,n,M) for a highly degenerate ground state |L,M|L,M\rangle has been derived in Ref. FMGM-S . It takes the form

S(L,n,M)=k=02snΛ(L,n,k,M)log2Λ(L,n,k,M),S(L,n,M)=-\sum_{k=0}^{2sn}\Lambda(L,n,k,M)\log_{2}\Lambda(L,n,k,M), (S12)

where the eigenvalues Λ(L,n,k,M)\Lambda(L,n,k,M) of the reduced density matrix ρL(n,M)\rho_{L}(n,M) are

Λ(L,n,k,M)=μ(L,k,M)ν(L,k,M),\Lambda(L,n,k,M)=\frac{\mu(L,k,M)}{\nu(L,k,M)}, (S13)

with

μ(L,k,M)=ns,,ns,ls,,lsr,t=ss1ε(s,r)nrCnm=sr1nmnrε(s,t)ltCLnm=st1lmlt,\mu(L,k,M)\!=\!{\sum}^{\prime}_{n_{-\!s},...,\;n_{s},\atop l_{-\!s},...,\;l_{s}}\prod_{r,t=-s}^{s-1}\!\varepsilon(s,r)^{n_{r}}{C_{n\!-\!\sum_{m=-s}^{r-1}\!n_{m}}^{n_{r}}}\!\varepsilon(s,t)^{l_{t}}{C_{L\!-\!n-\!\sum_{m=-s}^{t-1}l_{m}}^{l_{t}}},

and

ν(L,k,M)=Ns,,Nsr=ss1ε(s,r)NrCLm=sr1NmNr.\nu(L,k,M)={\sum}^{\prime}_{N_{-s},...,N_{s}}\prod_{r=-s}^{s-1}\varepsilon(s,r)^{N_{r}}{C_{L-\sum_{m=-s}^{r-1}N_{m}}^{N_{r}}}.

Here, the sum ns,,ns\sum^{\prime}_{n_{-s},\ldots,\;n_{s}} is taken over all possible values of nsn_{-s}, …, nsn_{s}, subject to the constraints m=ssnm=n\sum_{m=-s}^{s}n_{m}=n and m=ss(sm)nm=k\sum_{m=-s}^{s}(s-m)n_{m}=k, ls,,ls\sum^{\prime}_{l_{-s},\ldots,\;l_{s}} is taken over all the possible values of lsl_{-s}, …, lsl_{s}, subject to the constraints: m=sslm=Ln\sum_{m=-s}^{s}l_{m}=L-n. The sum m=ss(sm)lm=Mk\sum_{m=-s}^{s}(s-m)l_{m}=M-k, Ns,,Ns\sum^{\prime}_{N_{-s},...,\;N_{s}} is taken over all possible values of NsN_{-s}, …, NsN_{s}, subject to the constraints m=ssNm=L\sum_{m=-s}^{s}N_{m}=L and m=ss(sm)Nm=M\sum_{m=-s}^{s}(s-m)N_{m}=M. The factor ε(s,r)\varepsilon(s,r) takes the form

ε(s,r)=m=r+1s(s+m)(sm+1)m=rs1(sm)2.\varepsilon(s,r)=\frac{\prod_{m=r+1}^{s}{(s+m)(s-m+1)}}{\prod_{m=r}^{s-1}(s-m)^{2}}.

For the SU(N+1){\rm SU}(N+1) ferromagnetic model, the entanglement entropy S(L,n,M1,,MN)S(L,n,M_{1},\ldots,M_{N}) for a highly degenerate ground state |L,M1,M2,,MN|L,M_{1},M_{2},\ldots,M_{N}\rangle has been derived in Ref. FMGM-S . It takes the form

S(L,n,M1,,MN)=k1,,kN=0nΛ(L,n,k1,,kN,M1,,MN)log2Λ(L,n,k1,,kN,M1,,MN),S(L,n,M_{1},\ldots,M_{N})=-\sum_{k_{1},\ldots,k_{N}=0}^{n}\Lambda(L,n,k_{1},\ldots,k_{N},M_{1},\ldots,M_{N})\log_{2}\Lambda(L,n,k_{1},\ldots,k_{N},M_{1},\ldots,M_{N}), (S14)

where the eigenvalues Λ(L,n,k1,,kN,M1,,MN)\Lambda(L,n,k_{1},\ldots,k_{N},M_{1},\ldots,M_{N}) of the reduced density matrix ρL(n,M1,,MN)\rho_{L}(n,M_{1},\ldots,M_{N}) are

Λ(L,n,k1,,kN,M1,,MN)=α=1NCnβ=1α1kβkαγ=1NCLnβ=1γ1(Mβkβ)Mγkγα=1NCLβ=1α1MβMα.\displaystyle\Lambda(L,n,k_{1},\ldots,k_{N},M_{1},\ldots,M_{N})=\frac{\prod_{\alpha=1}^{N}{C_{n-\sum_{\beta=1}^{\alpha-1}{k_{\beta}}}^{k_{\alpha}}\prod_{\gamma=1}^{N}C_{L-n-\sum_{\beta=1}^{\gamma-1}{(M_{\beta}-k_{\beta})}}^{M_{\gamma}-k_{\gamma}}}}{\prod_{\alpha=1}^{N}{C_{L-\sum_{\beta=1}^{\alpha-1}{M_{\beta}}}^{M_{\alpha}}}}\;. (S15)

For the staggered SU(3){\rm SU}(3) spin-1 ferromagnetic biquadratic model, S(L,n,M1,M2)S(L,n,M_{1},M_{2}) and S(L,n,M2,M3)S(L,n,M_{2},M_{3}) for two highly degenerate ground states |L,M1,M22|L,M_{1},M_{2}\rangle_{2} and |L,M2,M34|L,M_{2},M_{3}\rangle_{4}, with the period being 2 and 4, respectively, have been derived in Ref. golden-S . The entanglement entropy S(L,n,M1,M2)S(L,n,M_{1},M_{2}) is

S(L,n,M1,M2)=k1=0min(M1,n/2)k2=0min(M2,nk1)Λ(L,n,k1,k2,M1,M2)log2Λ(L,n,k1,k2,M1,M2),S(L,n,M_{1},M_{2})=-\sum\limits_{k_{1}=0}^{\min(M_{1},n/2)}\sum\limits_{k_{2}=0}^{\min(M_{2},n-k_{1})}\Lambda(L,n,k_{1},k_{2},M_{1},M_{2})\log_{2}\Lambda(L,n,k_{1},k_{2},M_{1},M_{2}), (S16)

where nn is a multiple of two and the eigenvalues Λ(L,n,k1,k2,M1,M2)\Lambda(L,n,k_{1},k_{2},M_{1},M_{2}) of the reduced density matrix ρL(n,M1,M2)\rho_{L}(n,M_{1},M_{2}) are

Λ(L,n,k1,k2,M1,M2)=Cn/2k1Cnk1k2C(Ln)/2M1k1CLnM1+k1M2k2CL/2M1CLM1M2.\Lambda(L,n,k_{1},k_{2},M_{1},M_{2})=\frac{C_{n/2}^{k_{1}}C_{n-k_{1}}^{k_{2}}C_{(L-n)/2}^{M_{1}-k_{1}}C_{L-n-M_{1}+k_{1}}^{M_{2}-k_{2}}}{C_{L/2}^{M_{1}}C_{L-M_{1}}^{M_{2}}}. (S17)

The entanglement entropy S(L,n,M2,M3)S(L,n,M_{2},M_{3}) is

S(L,n,M2,M3)=k2=03n/4k3k3=0n/4Λ(L,k2,k3,M2,M3)log2Λ(L,k2,k3,M2,M3),S(L,n,M_{2},M_{3})=-\sum\limits_{k_{2}=0}^{3n/4-k_{3}}\sum\limits_{k_{3}=0}^{n/4}\Lambda(L,k_{2},k_{3},M_{2},M_{3})\log_{2}\Lambda(L,k_{2},k_{3},M_{2},M_{3}), (S18)

where nn is a multiple of four, and the eigenvalues Λ(L,n,k2,k3,M2,M3)\Lambda(L,n,k_{2},k_{3},M_{2},M_{3}) are

Λ(L,n,k2,k3,M2,M3)=Cn/4k3C3n/4k3k2C(Ln)/4M3k3C3(Ln)/4M3+k3M2k2CL/4M3C3L/4M3M2.\Lambda(L,n,k_{2},k_{3},M_{2},M_{3})=\frac{C_{n/4}^{k_{3}}C_{3n/4-k_{3}}^{k_{2}}C_{(L-n)/4}^{M_{3}-k_{3}}C_{3(L-n)/4-M_{3}+k_{3}}^{M_{2}-k_{2}}}{C_{L/4}^{M_{3}}C_{3L/4-M_{3}}^{M_{2}}}. (S19)

References

  • (1) P. Calabrese and J. Cardy, J. Stat. Mech. P06002 (2004).
  • (2) Q.-Q. Shi, Y.-W. Dai, H.-Q. Zhou, and I. P. McCulloch, arXiv: 2201.01071 (2022).
  • (3) V. Popkov and M. Salerno, Phys. Rev. A 71, 012301 (2005).
  • (4) A. O. Allen, Probability, statistics and queueing theory, Academic Press, Inc. (1990), Chap.3, p.160.
  • (5) B.-Q. Jin and V. E. Korepin, J. Stat. Phys. 116, 79-95 (2004).
  • (6) H.-Q. Zhou, Q.-Q. Shi, I. P. McCulloch, and M. T. Batchelor, arXiv: 2302.13126 (2023).