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

Exact spin helix eigenstates in anisotropic spin-ss Heisenberg model in arbitrary dimensions

Mingchen Zheng Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China    Chenguang Liang School of Physics, Peking University, Beijing 100871, China    Shu Chen Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China    Xin Zhang xinzhang@iphy.ac.cn Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China
Abstract

Spin-helix states are exact eigenstates of the spin-1/21/2 XXZ Heisenberg chain characterized by spatially modulated spin textures appearing under special parameter conditions. In this work, we generalize this framework and construct exact spin helix eigenstates in the fully anisotropic XYZ Heisenberg model, extending to higher spatial dimensions and spin representations, which reveals new stabilization mechanisms beyond integrability. The XXZ and XY limits are also analyzed to clarify the properties and constraints of these spin helix eigenstates. Our results broaden the class of analytically tractable exact eigenstates in quantum spin systems and deepen understanding of spatially inhomogeneous states in complex many-body systems.

I Introduction

Exact eigenstates in quantum many-body systems have attracted growing interest, as they offer rare analytic access to the structure of strongly correlated quantum matter Sutherland ; Amico et al. (2008). These states naturally arise in integrable systems Baxter (2016); Korepin et al. (1997), where an extensive set of conserved quantities enables exact solutions and leads to constrained, nonergodic dynamics describable by generalized Gibbs ensembles Ilievski et al. (2015, 2016); Essler and Fagotti (2016); Vidmar and Rigol (2016); Cecile et al. (2024) and generalized hydrodynamics Castro-Alvaredo et al. (2016); Bertini et al. (2016); Bastianello et al. (2022); Alba et al. (2021); De Nardis et al. (2022); Doyon et al. (2025). Remarkably, exact eigenstates can also emerge in nonintegrable systems, most notably as quantum many-body scar states (QMBS), in which a set of atypical, long-lived coherent states coexist with an otherwise thermalizing spectrum Serbyn et al. (2021); Moudgalya et al. (2022); Chandran et al. (2023); Turner et al. (2018a, b); Lin and Motrunich (2019); Mark et al. (2020); Moudgalya et al. (2018a, b); Schecter and Iadecola (2019); Chattopadhyay et al. (2020); Shibata et al. (2020); O’Dea et al. (2020); Pakrouski et al. (2020); Ren et al. (2021); Ivanov and Motrunich . Such states offer valuable benchmarks for understanding entanglement structure Amico et al. (2008); Eisert et al. (2010), coherence and ergodicity breaking D’Alessio et al. (2016); Nandkishore and Huse (2015); Abanin et al. (2019); Sierant et al. (2025), and serve as experimentally accessible targets for testing quantum dynamics and controlling many-body evolution in programmable quantum simulators Bernien et al. (2017); Bluvstein et al. (2021).

A particularly striking class of exact eigenstates emerges in the paradigmatic spin-1/2 XXZ Heisenberg chain with periodic boundary conditions, where the so-called spin-helix states (SHSs) arise under fine-tuned conditions (the anisotropy parameter η\eta takes root-of-unity values) Popkov et al. (2021); Jepsen et al. (2022a). These eigenstates exhibit spatially rotating spin textures and provide rare examples of highly structured, nonthermal excited eigenstates in an interacting many-body system. Their origin can be traced back to Baxter’s seminal work on the eight-vertex (XYZ) model in the 1970s Baxter (1973a, b, c), which revealed spectral degeneracies at special parameter values—later understood to stem from an enhanced quantum group symmetry Uq(𝔰𝔩2)U_{q}(\mathfrak{sl}_{2}) where qq is a root of unity Pasquier and Saleur (1990); Deguchi et al. (2001). More recently, SHSs have been reinterpreted via the Bethe ansatz as arising from “phantom” Bethe roots, in which all quasiparticles share identical momentum and do not scatter Popkov et al. (2021). Importantly, SHSs have been experimentally observed in ultracold atom systems simulating the XXZ chain, where long-lived helical spin patterns were detected Jepsen et al. (2021, 2022a). Moreover, SHSs persist beyond the integrable setting, appearing in higher-dimensional and higher-spin extensions of the XXZ model where they represent robust examples of QMBS Jepsen et al. (2022a), suggesting a unifying framework for nonthermal states across integrable and nonintegrable regimes.

Despite this progress, the status of SHSs in more general settings—such as the fully anisotropic XYZ Heisenberg model—remains poorly understood. The XYZ model features unequal couplings along all spin directions and lacks continuous symmetries, unlike the U(1)U(1)-symmetric XXZ chain. This more intricate parameter landscape raises fundamental questions about the existence and structure of exact spin helix eigenstates beyond the integrable regime. In this work, we investigate SHSs in the XYZ model and demonstrate that, under specific parameter conditions, it supports eigenstates with helical structure. We further extend the analysis to higher spatial dimensions dd and higher spin ss, and offer a unified framework for understanding spin helix eigenstates across different parameter regimes. We also examine the XXZ and XY limits in detail, clarifying the characteristics and parameter constraints of spin helix eigenstates in these cases. Our results provide a comprehensive perspective on analytically tractable, spatially inhomogeneous eigenstates in interacting quantum spin systems.

This paper is organized as follows. In Section II, we introduce the notation and conventions for elliptic theta functions, which are essential for describing the modulated structure of spin helix eigenstates in the XYZ model. Section III presents the construction and characterization of exact spin helix eigenstates in the dd-dimensional hypercubic XYZ spin model. In Section IV, we analyze the degenerate limits of the XYZ model—such as the XXZ and XY cases—and demonstrate how spin helix eigenstates persist and transform in these regimes. In Section V, we extend our discussion to other classes of spin models that support spin helix eigenstates, including systems with long-range interactions and more complex lattice geometries. Finally, Section VI is the summary and outlook.

II Theta functions

We introduce the following Jacobi theta functions ϑα(u,q)\vartheta_{\alpha}(u,q) Whittaker and Watson (1950)

ϑ1(u,q)=2n=0(1)nq(n+12)2sin[(2n+1)u],ϑ2(u,q)=2n=0q(n+12)2cos[(2n+1)u],ϑ3(u,q)=1+2n=1qn2cos(2nu),ϑ4(u,q)=1+2n=1(1)nqn2cos(2nu).\displaystyle\begin{aligned} &\vartheta_{1}(u,q)=2\sum_{n=0}^{\infty}(-1)^{n}q^{(n+\frac{1}{2})^{2}}\sin[(2n+1)u],\\ &\vartheta_{2}(u,q)=2\sum_{n=0}^{\infty}q^{(n+\frac{1}{2})^{2}}\cos[(2n+1)u],\\ &\vartheta_{3}(u,q)=1+2\sum_{n=1}^{\infty}q^{n^{2}}\cos(2nu),\\ &\vartheta_{4}(u,q)=1+2\sum_{n=1}^{\infty}(-1)^{n}q^{n^{2}}\cos(2nu).\end{aligned} (1)

For convenience, we use the following shorthand notations θα(u),θ~α(u),α=1,2,3,4\theta_{\alpha}(u),\,\tilde{\theta}_{\alpha}(u),\,\alpha=1,2,3,4

θα(u)ϑα(πu,eiπτ),θ~α(u)ϑα(πu,e2iπτ),\displaystyle\theta_{\alpha}(u)\equiv\vartheta_{\alpha}(\pi u,{\mathrm{e}}^{{\mathrm{i}}\pi\tau}),\quad\tilde{\theta}_{\alpha}(u)\equiv\vartheta_{\alpha}(\pi u,{\mathrm{e}}^{2{\mathrm{i}}\pi\tau}), (2)

where τ\tau is a complex with a positive imaginary part. Further details on theta functions can be found in Appendix A.

III XYZ Model and its exact spin helix eigenstate

The Hamiltonian of the spin-ss XYZ Heisenberg model is

H=i,jHi,j,\displaystyle H=\sum_{\langle i,j\rangle}H_{i,j},\quad (3)

where Hi,jH_{i,j} is the two-site XYZ interaction term

Hi,j=JxSixSjx+JySiySjy+JzSizSjz,H_{i,j}=J_{x}{S}_{i}^{x}S_{j}^{x}+J_{y}S_{i}^{y}S_{j}^{y}+J_{z}S_{i}^{z}S_{j}^{z}, (4)

Here, Sjα{S}_{j}^{\alpha} (α=x,y,z)(\alpha=x,\ y,\ z) denote the spin-ss operators at lattice site jj of a dd-dimensional hypercubic lattice with volume V=α=1dLαV=\prod_{\alpha=1}^{d}L_{\alpha}, where LαL_{\alpha} is the number of sites along the α\alpha-th spatial direction. The notation i,j{\langle i,j\rangle} indicates summation over nearest-neighbor site pairs, and the exchange coefficients are parameterized by theta functions as follows Wang et al. (2016)

Jx=θ4(η)θ4(0),Jy=θ3(η)θ3(0),Jz=θ2(η)θ2(0).\displaystyle J_{x}=\frac{\theta_{4}(\eta)}{\theta_{4}(0)},\quad J_{y}=\frac{\theta_{3}(\eta)}{\theta_{3}(0)},\quad J_{z}=\frac{\theta_{2}(\eta)}{\theta_{2}(0)}. (5)

Obviously, the Hamiltonian depends on two parameters: η\eta and τ\tau. From the identities

Jα|ηη+2=Jα,Jα|ηη+2τ=e4iπ(η+τ)Jα,\displaystyle J_{\alpha}|_{\eta\to\eta+2}=J_{\alpha},\quad J_{\alpha}|_{\eta\to\eta+2\tau}={\mathrm{e}}^{-4{\mathrm{i}}\pi(\eta+\tau)}J_{\alpha}, (6)

we can thus restrict parameter η\eta to the rectangle 0(η)<2, 0Im(η)<2Im(τ)0\leq(\eta)<2,\,0\leq{\rm Im}(\eta)<2{\rm Im}(\tau) without loss of generality.

When τ\tau is purely imaginary and η\eta is real or also purely imaginary, the Hamiltonian in Eq. (3) is Hermitian, as follows

  • when Im(η)=(τ)=0{\rm Im}(\eta)=(\tau)=0,   |Jx||Jy||Jz||J_{x}|\geq|J_{y}|\geq|J_{z}|,

  • when (η)=(τ)=0(\eta)=(\tau)=0,   |Jx||Jy||Jz||J_{x}|\leq|J_{y}|\leq|J_{z}|.

Let us first introduce the following local vector

ψ(s)(u)\displaystyle\psi^{(s)}(u) =1𝒩(u)n=02sκn[θ~1(u)]2sn[θ~4(u)]n|sn,\displaystyle=\frac{1}{\mathcal{N}(u)}\sum_{n=0}^{2s}\kappa_{n}\left[\tilde{\theta}_{1}(u)\right]^{2s-n}\left[\tilde{\theta}_{4}(u)\right]^{n}\ket{s-n}, (7)
κn\displaystyle\kappa_{n} =(2s)!/(n!(2sn)!),\displaystyle=\sqrt{(2s)!/(n!(2s-n)!)},
𝒩(u)\displaystyle\mathcal{N}(u) =[θ4((u))θ3(iIm(u))]s\displaystyle=\left[\theta_{4}((u))\theta_{3}({\mathrm{i}}{\rm Im}(u))\right]^{s}

where uu is a free parameter, {|m}\{\ket{m}\} are the SzS^{z} basis with Sz|m=m|mS^{z}\ket{m}=m\ket{m}. The state ψ(s)(u)\psi^{(s)}(u) in (7) is a spin coherent state which can be obtained by rotating the highest-weight state |s\ket{s} first around the yy-axis and then around the zz-axis, specifically as follows

ψ(s)(u)eiβ(u)Szeiγ(u)Sy|s,\displaystyle\psi^{(s)}(u)\propto{\mathrm{e}}^{-{\mathrm{i}}\beta(u)S_{z}}{\mathrm{e}}^{-{\mathrm{i}}\gamma(u)S_{y}}\ket{s}, (8)

where

γ(u)=2arctan|θ~4(u)θ~1(u)|,β(u)=arg(θ~4(u)θ~1(u)).\displaystyle\gamma(u)=2\arctan\left|\frac{\tilde{\theta}_{4}(u)}{\tilde{\theta}_{1}(u)}\right|,\quad\beta(u)=\arg\left(\frac{\tilde{\theta}_{4}(u)}{\tilde{\theta}_{1}(u)}\right). (9)

Thus we can also use γ(u)\gamma(u) and β(u)\beta(u) to represent the state ψ(s)(u)\psi^{(s)}(u), and the expectation value of SαS^{\alpha} in the state ψ(s)(u)\psi^{(s)}(u) reads

Sx=ssin(γ(u))cos(β(u)),Sy=ssin(γ(u))sin(β(u)),Sz=scos(γ(u)).\displaystyle\begin{aligned} \langle S_{x}\rangle&=s\sin(\gamma(u))\cos(\beta(u)),\\ \langle S_{y}\rangle&=s\sin(\gamma(u))\sin(\beta(u)),\\ \langle S_{z}\rangle&=s\cos(\gamma(u)).\end{aligned} (10)

The proof of Eqs. (8) and (10) is presented in Appendix B.

The local vector ψi(s)(u)\psi^{(s)}_{i}(u) in (7) satisfies the following divergence condition:

Hi,jψi(s)(u)ψj(s)(u±η)=s2b(±u)ψi(s)(u)ψj(s)(u±η)\displaystyle H_{i,j}\,\psi^{(s)}_{i}(u)\psi^{(s)}_{j}(u\pm\eta)=s^{2}b(\pm u)\psi^{(s)}_{i}(u)\psi^{(s)}_{j}(u\pm\eta)
±s[a(u)Siza(u±η)Sjz]ψi(s)(u)ψj(s)(u±η),\displaystyle\quad\pm s\left[a(u)S_{i}^{z}-a(u\pm\eta)S_{j}^{z}\right]\,\psi^{(s)}_{i}(u)\psi^{(s)}_{j}(u\pm\eta), (11)

where

a(u)=θ1(η)θ2(u)θ2(0)θ1(u),b(u)=g(η)+g(u)g(u+η),\displaystyle a(u)=\frac{\theta_{1}(\eta)\theta_{2}(u)}{\theta_{2}(0)\theta_{1}(u)},\quad b(u)=g(\eta)+g(u)-g(u+\eta),
g(u)=θ1(η)θ1(u)θ1(0)θ1(u),θj(u)=θj(u)u.\displaystyle g(u)=\frac{\theta_{1}(\eta)\theta_{1}^{\prime}(u)}{\theta_{1}^{\prime}(0)\theta_{1}(u)},\quad\theta^{\prime}_{j}(u)=\frac{\partial\theta_{j}(u)}{\partial u}. (12)

Equation (11) implies that only two non-eigen terms arise when Hi,jH_{i,j} acts on the product state ψi(s)(u)ψj(s)(v)\psi^{(s)}_{i}(u)\psi^{(s)}_{j}(v) and the “phase” factors uu and vv in ψi(s)(u)\psi^{(s)}_{i}(u) and ψj(s)(v)\psi^{(s)}_{j}(v) has a fixed difference ±η\pm\eta. The foundational identities presented in Eq. (11) inspire the construction of specific tensor product states in Theorem 1. When s=1/2s=1/2, Eq. (11) was established in Refs. Popkov et al. (2022); Zhang et al. (2022, 2024). In this paper, we generalize it to the case of an arbitrary ss. The proof of Eq. (11) involves numerous elliptic function identities. Due to its inherent complexity, we present complete proof details in Appendix C.

We can construct the following tensor product state

|Ψ(s)(u,ϵ)=jψj(s)(u+ηϵ𝒏j),u,\displaystyle\ket{\Psi^{(s)}(u,\bm{\epsilon})}=\bigotimes_{j}\psi^{(s)}_{j}\left(u+\eta\,\bm{\epsilon}\cdot\bm{n}_{j}\right),\quad u\in\mathbb{C}, (13)

where 𝒏j=(nj,1,,nj,d)\bm{n}_{j}=(n_{j,1},\ldots,n_{j,d}) is the coordinate of site jj in the dd-dimensional hypercubic lattice, ϵ=(ϵ1,,ϵd)\bm{\epsilon}=(\epsilon_{1},\ldots,\epsilon_{d}), ϵα=±1\epsilon_{\alpha}=\pm 1 is a dd-dimensional vector, and ϵ𝒏j=ϵ1nj,1++ϵdnj,d\bm{\epsilon}\cdot\bm{n}_{j}=\epsilon_{1}n_{j,1}+\dots+\epsilon_{d}\,n_{j,d} describes a phase that exhibits linear spatial variation along each lattice direction. A visualization of the phase ϵ𝒏j\bm{\epsilon}\cdot\bm{n}_{j} is demonstrated in Fig. 1. Such an inhomogeneous state was first proposed by Baxter in the pioneering study of the one-dimensional spin-1/21/2 XYZ model Baxter (1973a, c). In Eq. (13), we generalize Baxter’s result to dd-dimensional spin-ss Heisenberg chain.

0ϵ1\epsilon_{1}2ϵ12\epsilon_{1}ϵ2\epsilon_{2}ϵ1+ϵ2\epsilon_{1}+\epsilon_{2}2ϵ1+ϵ22\epsilon_{1}+\epsilon_{2}2ϵ22\epsilon_{2}ϵ1+2ϵ2\epsilon_{1}+2\epsilon_{2}2ϵ1+2ϵ22\epsilon_{1}+2\epsilon_{2}
Figure 1: The spatial variation of the phase factor ϵ𝒏j\bm{\epsilon}\cdot\bm{n}_{j} across the 2dd lattice.
Theorem 1.

Under periodic boundary conditions, the vector |Ψ(s)(u,ϵ)\ket{\Psi^{(s)}(u,\bm{\epsilon})} in Eq. (13) is an eigenstate of HH (3) with energy

E=ds2θ1(η)θ1(0)V+4iπs2θ1(η)θ1(0)β=1dpβαβLα,\begin{split}E=&d\,s^{2}\frac{\theta_{1}^{\prime}(\eta)}{\theta_{1}^{\prime}(0)}V+4{\mathrm{i}}\pi s^{2}\frac{\theta_{1}(\eta)}{\theta_{1}^{\prime}(0)}\sum_{\beta=1}^{d}p_{\beta}\prod_{\alpha\neq\beta}L_{\alpha},\end{split} (14)

when the elliptic commensurability condition holds

Lαη=2pατ+2qα,α=1,,d,pα,qα.\displaystyle L_{\alpha}\eta=2p_{\alpha}\tau+2q_{\alpha},\,\,\alpha=1,\dots,d,\,\,p_{\alpha},q_{\alpha}\in\mathbb{Z}. (15)
Proof.

We see that the state ψ(s)(u)\psi^{(s)}(u) (7) is entirely determined by the ratio θ~4(u)θ~1(u)\frac{\tilde{\theta}_{4}(u)}{\tilde{\theta}_{1}(u)}. With the help of the following identity

θ~4(u+2k+2lτ)θ~1(u+2k+2lτ)=(A3,A4)θ~4(u)θ~1(u),k,l,\displaystyle\frac{\tilde{\theta}_{4}(u+2k+2l\tau)}{\tilde{\theta}_{1}(u+2k+2l\tau)}\overset{(\ref{periodicity;1},\ref{periodicity;2})}{=}\frac{\tilde{\theta}_{4}(u)}{\tilde{\theta}_{1}(u)},\quad k,l\in\mathbb{Z}, (16)

the elliptic commensurability condition (15) ensures

ψ(s)(u+Lαη)ψ(s)(u),\psi^{(s)}(u+L_{\alpha}\eta)\propto\psi^{(s)}(u),

so that the tensor product state (13) is compatible with periodic boundary conditions.

Due to the quasi-periodicity of theta functions (see Eqs. (A3) and (A4)), we can get these useful identities

a(u+2k+2lτ)=a(u),\displaystyle a(u+2k+2l\tau)=a(u),
g(u+2k+2lτ)=g(u)4iπlθ1(η)θ1(0),k,l,\displaystyle g(u+2k+2l\tau)=g(u)-4{\mathrm{i}}\pi l\,\frac{\theta_{1}(\eta)}{\theta_{1}^{\prime}(0)},\quad k,l\in\mathbb{Z}, (17)

where a(u)a(u) and g(u)g(u) are defined in Eq. (12).

First, we consider the one-dimensional system. Acting the Hamiltonian HH on the state |Ψ(s)(u,ϵ)k=1Lψk(s)(u+ϵkη)\ket{\Psi^{(s)}(u,\epsilon)}\equiv\bigotimes_{k=1}^{L}\psi_{k}^{(s)}(u+\epsilon k\eta), we get

H|Ψ(s)(u,ϵ)\displaystyle\quad H\ket{\Psi^{(s)}(u,\epsilon)}
=n=1LHn,n+1k=1Lψk(s)(uk(ϵ))\displaystyle=\sum_{n=1}^{L}H_{n,n+1}\bigotimes_{k=1}^{L}\psi_{k}^{(s)}(u_{k}^{(\epsilon)})
=(11)ϵsn=1L[a(un(ϵ))Snza(un+1(ϵ))Sn+1z]|Ψ(s)(u,ϵ)\displaystyle\overset{(\ref{div;eq})}{=}\epsilon s\sum_{n=1}^{L}\left[a(u_{n}^{(\epsilon)})S_{n}^{z}-a(u_{n+1}^{(\epsilon)})S_{n+1}^{z}\right]\ket{\Psi^{(s)}(u,\epsilon)}
+s2n=1Lb(ϵun(ϵ))|Ψ(s)(u,ϵ)\displaystyle\quad+s^{2}\sum_{n=1}^{L}b(\epsilon u_{n}^{(\epsilon)})\ket{\Psi^{(s)}(u,\epsilon)}
=ϵs[a(u1(ϵ))S1za(uL+1(ϵ))S1z]|Ψ(s)(u,ϵ)\displaystyle=\epsilon s\left[a(u_{1}^{(\epsilon)})S_{1}^{z}-a(u_{L+1}^{(\epsilon)})S_{1}^{z}\right]\ket{\Psi^{(s)}(u,\epsilon)}
+s2[Lg(η)+g(ϵu1(ϵ))g(ϵuL+1(ϵ))]|Ψ(s)(u,ϵ),\displaystyle\quad+s^{2}\left[Lg(\eta)+g(\epsilon u_{1}^{(\epsilon)})-g(\epsilon u_{L+1}^{(\epsilon)})\right]\ket{\Psi^{(s)}(u,\epsilon)},

where un(ϵ)=u+ϵnηu^{(\epsilon)}_{n}=u+\epsilon n\eta. Using (15) and (17), we have

a(u1(ϵ))a(uL+1(ϵ))=0,g(ϵu1(ϵ))g(ϵuL+1(ϵ))=4iπpθ1(η)θ1(0).\displaystyle\begin{aligned} a(u_{1}^{(\epsilon)})-a(u_{L+1}^{(\epsilon)})&=0,\\ g(\epsilon u_{1}^{(\epsilon)})-g(\epsilon u_{L+1}^{(\epsilon)})&=4{\mathrm{i}}\pi p\,\frac{\theta_{1}(\eta)}{\theta_{1}^{\prime}(0)}.\end{aligned} (18)

Thus,

H|Ψ(s)(u,ϵ)=s2[Lθ1(η)θ1(0)+4iπpθ1(η)θ1(0)]|Ψ(s)(u,ϵ),\displaystyle H\ket{\Psi^{(s)}(u,\epsilon)}{=}s^{2}\left[L\,\frac{\theta_{1}^{\prime}(\eta)}{\theta_{1}^{\prime}(0)}+4{\mathrm{i}}\pi p\,\frac{\theta_{1}(\eta)}{\theta_{1}^{\prime}(0)}\right]\ket{\Psi^{(s)}(u,\epsilon)}, (19)

the theorem is proved for the one-dimensional case.

Analogously, our proof can be generalized to higher-dimensional systems. By acting the Hamiltonian HH on the state |Ψ(s)(u,ϵ)\ket{\Psi^{(s)}(u,\bm{\epsilon})} and applying Eq. (11) to each nearest-neighbor bond, we find that the only non-eigen contributions arise from SzS^{z} terms. However, these terms cancel out across the lattice due to the alternating phase structure and the periodicity condition (15). This confirms that |Ψ(s)(u,ϵ)\ket{\Psi^{(s)}(u,\bm{\epsilon})} is indeed an eigenstate of HH and the corresponding eigenvalue is given by the sum of all coefficients preceding the eigenterms, which equals Eq. (14).

In the state |Ψ(s)(u,ϵ)\ket{\Psi^{(s)}(u,\bm{\epsilon})}, the polar and azimuthal angles γj(u)\gamma_{j}(u) and βj(u)\beta_{j}(u) defined in Eq. (9) vary systematically with site index jj, forming a spatially modulated spin texture characteristic of a spin helix. The winding direction is set by the vector ϵ\bm{\epsilon}, while the pitch and phase of the helix depend sensitively on the parameters uu, η\eta, and the modular parameter τ\tau, as illustrated in Figs. 2 and 3. Owing to the factorized structure of the state, its spin configuration admits a direct classical interpretation in the large-ss limit, where quantum fluctuations are suppressed and the local spin orientation approaches a classical vector field.

Refer to caption
Refer to caption
Figure 2: Visualization of the spin helix structure in n=1Lψn(s)(uη+nη)\bigotimes_{n=1}^{L}\psi_{n}^{(s)}(u-\eta+n\eta) for the 1D spin-ss XYZ chain with L=11,η=211,τ=0.8iL=11,\eta=\frac{2}{11},\,\tau=0.8{\mathrm{i}} (Jx=1.1128,Jy=0.9184,Jz=0.8348J_{x}=1.1128,\,J_{y}=0.9184,\,J_{z}=0.8348). Upper panel: u=0.28\,u=0.28. Lower panel: u=0.28+τ2u=0.28+\frac{\tau}{2}. The arrows indicate local spin directions, while the numbers denote lattice site coordinates. Upper panel: All spins are confined to the xzxz-plane, exhibiting periodic oscillations with lattice site variation. Lower panel: All spins are restricted to the xyxy-plane, forming a winding structure. In generic case, both γ\gamma and β\beta vary with the lattice coordinate. In this figure, we constrain all spins to a fixed plane to demonstrate the helix structure more clearly.

The eigenstate |Ψ(s)(u,ϵ)\ket{\Psi^{(s)}(u,\bm{\epsilon})} depends on a free phase parameter uu, whereas the energy EE remains independent of this parameter. This implies that the spin helix eigenstates form a degenerate manifold. |Ψ(s)(u,ϵ)\ket{\Psi^{(s)}(u,\bm{\epsilon})} is an exact nonthermal eigenstate that exists in higher dimension XYZ models, which are generally believed to be non-integrable Shiraishi and Tasaki . As such, |Ψ(s)(u,ϵ)\ket{\Psi^{(s)}(u,\bm{\epsilon})} constitutes a QMBS.

For finite systems, Eq. (15) implies the parameter η\eta can only take several discrete values

η=2pατ+2qαLα,pα,qα.\eta=\frac{2p_{\alpha}\tau+2q_{\alpha}}{L_{\alpha}},\,\,p_{\alpha},q_{\alpha}\in\mathbb{Z}. (20)

In the thermodynamic limit, these discrete points become densely distributed, allowing η\eta to approach any value within the complex plane. This demonstrates the universality of spin helix eigenstates in Heisenberg models.

Refer to caption
Refer to caption
Figure 3: (a) Local spin configuration of a spin helical state in the two-dimensional XYZ model, shown for a subset of a L×LL\times L lattice. The arrows represent spin orientations at selected lattice sites. (b) Local spin expectation values SjαS^{\alpha}_{j} (α=x,y,z\alpha=x,y,z) are plotted along the xx-axis of the lattice for site indices j=0,,L1j=0,\ldots,L-1. Local spin expectation values L=27L=27. Common Parameters: s=1s=1, u=0.4u=0.4, τ=0.7i\tau=0.7{\mathrm{i}} and η=10τ/L\eta=10\tau/L (Jx=0.5353,Jy=1.3020,Jz=1.4046J_{x}=0.5353,\ J_{y}=1.3020,\ J_{z}=1.4046).

IV Spin helix eigenstate in XXZ model and XY model

The XYZ model constitutes the most general form of the Heisenberg model class. In specific parameter regimes, it reduces to partially anisotropic models, such as the XXZ and XY models.

IV.1 Spin-ss XXZ model

In the trigonometric limit τ+i\tau\to+{\mathrm{i}}\infty, the XYZ Hamiltonian reduces to the XXZ model with

Hi,jXXZ=SixSjx+SiySjy+cos(πη)SizSjz.\displaystyle H_{i,j}^{\rm XXZ}={S}_{i}^{x}S_{j}^{x}+S_{i}^{y}S_{j}^{y}+\cos(\pi\eta)S_{i}^{z}S_{j}^{z}. (21)

In this limit, some functions in the previous section degenerate to Zhang et al. (2024):

limτ+iθ~4(u)θ~1(u)=eiπv,limτ+ib(u)=cos(πη),limτ+ia(u)=isin(πη),\displaystyle\begin{aligned} &\lim_{\tau\to+{\mathrm{i}}\infty}\frac{\tilde{\theta}_{4}(u)}{\tilde{\theta}_{1}(u)}={\mathrm{e}}^{{\mathrm{i}}\pi v},\\ &\lim_{\tau\to+{\mathrm{i}}\infty}b(u)=\cos(\pi\eta),\\ &\lim_{\tau\to+{\mathrm{i}}\infty}a(u)=-{\mathrm{i}}\sin(\pi\eta),\end{aligned} (22)

where we introduce the shifted variable u=v+1+τ2u=v+\frac{1+\tau}{2} for convenience.

Correspondingly, the elliptic local spinor ψ(s)(u)\psi^{(s)}(u) degenerates into the following trigonometric spin-coherent state

ψ¯(s)(u)=1𝒩¯(u)n=02sκneiπnu|sn,\displaystyle\bar{\psi}^{(s)}(u)=\frac{1}{\bar{\mathcal{N}}(u)}\sum_{n=0}^{2s}\kappa_{n}\,{\mathrm{e}}^{{\mathrm{i}}\pi nu}\ket{s-n}, (23)

where the normalization factor is

𝒩¯(u)=[1+e2πIm(u)]s.\bar{\mathcal{N}}(u)=\left[1+{\mathrm{e}}^{-2\pi{\rm Im}(u)}\right]^{s}. (24)

From Eqs. (22) and (23), we can get the divergence condition for the XXZ model

Hi,jXXZ\displaystyle H_{i,j}^{\rm XXZ}\, ψ¯i(s)(u)ψ¯j(s)(u±η)=s2cos(πη)ψ¯i(s)(u)ψ¯j(s)(u±η)\displaystyle\bar{\psi}^{(s)}_{i}(u)\bar{\psi}^{(s)}_{j}(u\pm\eta)\!=\!s^{2}\cos(\pi\eta)\,\bar{\psi}^{(s)}_{i}(u)\bar{\psi}^{(s)}_{j}(u\pm\eta)
issin(πη)(SizSjz)ψ¯i(s)(u)ψ¯j(s)(u±η).\displaystyle\quad\mp{\mathrm{i}}s\sin(\pi\eta)(S_{i}^{z}-S_{j}^{z})\bar{\psi}^{(s)}_{i}(u)\bar{\psi}^{(s)}_{j}(u\pm\eta). (25)

For the XXZ model, we can construct the following spin helix state

|Ψ¯(s)(u,ϵ)=jψ¯j(s)(u+ηϵ𝒏j),u.\displaystyle\ket{\bar{\Psi}^{(s)}(u,\bm{\epsilon})}=\bigotimes_{j}\bar{\psi}^{(s)}_{j}\left(u+\eta\,\bm{\epsilon}\cdot\bm{n}_{j}\right),\qquad u\in\mathbb{C}. (26)

By taking the limit τ+i\tau\to+{\mathrm{i}}\infty in the context of Theorem 1, we directly derive Theorem 2.

Theorem 2.

Under the root of unity condition

Lαη=2qα,qα,α=1,,d.L_{\alpha}\eta=2q_{\alpha},\quad q_{\alpha}\in\mathbb{Z},\quad\alpha=1,\dots,d. (27)

The state |Ψ¯(s)(u,ϵ)\ket{\bar{\Psi}^{(s)}(u,\bm{\epsilon})} proposed in Eq. (26) is an exact eigenstate of the XXZ Hamiltonian. The corresponding energy is given by

E=ds2cos(πη)V,E=ds^{2}\cos(\pi\eta)\,V, (28)

It should be noted that the spin helix eigenstate of the Hamiltonian only exists in the easy plane regime (η\eta is real). The state |Ψ¯(s)(u,ϵ)\ket{\bar{\Psi}^{(s)}(u,\bm{\epsilon})} now exhibits a “perfect” transverse helix structure with

γ(u±η)=γ(u)=2arctan(eπIm(u)),\displaystyle\gamma(u\pm\eta)=\gamma(u)=2\arctan({\mathrm{e}}^{-\pi{\rm Im}(u)}),
β(u±η)=β(u)±πη,β(u)=πu.\displaystyle\beta(u\pm\eta)=\beta(u)\pm\pi\eta,\quad\beta(u)=\pi u. (29)

In the XXZ model, all local spins are confined to the same plane (with fixed polar angle γ\gamma), while the azimuthal angles β\beta vary linearly with the lattice site. The spin helix eigenstate of the spin-ss XXZ model has been discussed in Ref. Jepsen et al. (2022b), which is a remarkable generalization of the phantom Bethe state discovered in the one-dimensional integrable spin-12\frac{1}{2} XXZ model Popkov et al. (2021).

It should be remarked that the spin helix eigenstate |Ψ(s)(u,ϵ)\ket{{\Psi}^{(s)}(u,\bm{\epsilon})} has another trivial degenerations in the limit τ+i\tau\to+{\mathrm{i}}\infty

limτ+iu/τ1|Ψ(s)(u,ϵ)\displaystyle\lim_{\begin{subarray}{c}\tau\to+{\mathrm{i}}\infty\\ u/\tau\to 1\end{subarray}}\ket{{\Psi}^{(s)}(u,\bm{\epsilon})} |Ω=|s|s,\displaystyle\propto\ket{\Omega}=\ket{s}\otimes\cdots\otimes\ket{s}, (30)
limτ+iu/τ0|Ψ(s)(u,ϵ)\displaystyle\lim_{\begin{subarray}{c}\tau\to+{\mathrm{i}}\infty\\ u/\tau\to 0\end{subarray}}\ket{{\Psi}^{(s)}(u,\bm{\epsilon})} |Ω¯=|s|s,\displaystyle\propto\ket{\bar{\Omega}}=\ket{-s}\otimes\cdots\otimes\ket{-s}, (31)

which is consistent with the fact that the highest-weight state |Ω\ket{\Omega} and the lowest-weight state |Ω¯\ket{\bar{\Omega}} remain exact eigenstates of the XXZ Hamiltonian.

Unlike the XYZ model, the XXZ Hamiltonian possesses a U(1)U(1) symmetry and commutes with the total zz-component magnetization operator jSjz\sum_{j}S_{j}^{z}. Given the high degeneracy of the energy level E=ds2cos(πη)VE=ds^{2}\cos(\pi\eta)\,V, a natural question arises: can we derive common eigenstates of this subspace from spin helix eigenstates?

We observe that the spin helix eigenstate |Ψ¯(s)(u,ϵ)\ket{\bar{\Psi}^{(s)}(u,\bm{\epsilon})} in Eq. (26) can be expanded as a linear combination of a set of uu-independent vectors, as demonstrated below

|Ψ¯(s)(u,ϵ)=1j𝒩¯(u+ηϵ𝒏j)n=02sVeiπnu|Ψn,ϵ(s),\displaystyle\ket{\bar{\Psi}^{(s)}(u,\bm{\epsilon})}=\frac{1}{\prod_{j}\bar{\mathcal{N}}(u+\eta\,\bm{\epsilon}\cdot\bm{n}_{j})}\sum_{n=0}^{2sV}{\mathrm{e}}^{{\mathrm{i}}\pi nu}\ket{\Psi_{n,\bm{\epsilon}}^{(s)}}, (32)

where {|Ψn,ϵ(s)}\left\{\ket{\Psi^{(s)}_{n,\bm{\epsilon}}}\right\} are the eigenstates of jSjz\sum_{j}S_{j}^{z}

jSjz|Ψn,ϵ(s)=(sVn)|Ψn,ϵ(s),n=0,,2sV.\displaystyle\sum_{j}S_{j}^{z}\ket{\Psi^{(s)}_{n,\bm{\epsilon}}}=(sV-n)\ket{\Psi^{(s)}_{n,\bm{\epsilon}}},\,\,n=0,\ldots,2sV. (33)

The state |Ψn,ϵ(s)\ket{\Psi^{(s)}_{n,\bm{\epsilon}}} is generated by acting the nonlocal chiral ladder operator on the vacuum state |Ω\ket{\Omega}, as follows

|Ψn,ϵ(s)=1n!(Jϵ)n|Ω,\displaystyle\ket{\Psi_{n,\bm{\epsilon}}^{(s)}}=\frac{1}{n!}\left(J_{\bm{\epsilon}}^{-}\right)^{n}\ket{\Omega}, (34)
Jϵ±=jexp(iπηϵ𝒏j)Sj±.\displaystyle J_{\bm{\epsilon}}^{\pm}=\sum_{j}\exp\left({\mathrm{i}}\pi\eta\,\bm{\epsilon}\cdot\bm{n}_{j}\right)S^{\pm}_{j}. (35)

Here ϵ\bm{\epsilon} specifies distinct chiral sectors.

Theorem 3.

The state |Ψ¯(s)(u,ϵ)\ket{\bar{\Psi}^{(s)}(u,\bm{\epsilon})} remains an eigenstate of the Hamiltonian for arbitrary uu under condition (27). It follows directly that all vectors in the set {|Ψn,ϵ(s)}\left\{\ket{\Psi^{(s)}_{n,\bm{\epsilon}}}\right\} are common eigenstates of HH and jSjz\sum_{j}S_{j}^{z}, and share the same energy eigenvalue E=ds2cos(πη)V.E=ds^{2}\cos(\pi\eta)\,V.

The state |Ψn,ϵ(s)\ket{\Psi_{n,\bm{\epsilon}}^{(s)}} can be considered as a collective magnon excitation, where all spin flips are coherently modulated across the lattice with a momentum πηϵ\pi\eta\,\bm{\epsilon}. The states |Ψn,ϵ(s)\ket{\Psi^{(s)}_{n,\bm{\epsilon}}} with different values of nn are mutually orthogonal. For fixed nn, the states with different ϵ\bm{\epsilon} are generally non-orthogonal but can be shown to be linearly independent. Consequently, the set of states {|Ψ¯(s)(u,ϵ)}\left\{\ket{\bar{\Psi}^{(s)}(u,\bm{\epsilon})}\right\} spans a degenerate invariant subspace of dimension

2d(2sV+1)2(2d1)=2d(2sV1)+2.2^{d}(2sV+1)-2(2^{d}-1)=2^{d}(2sV-1)+2.

For the one-dimensional spin-12\frac{1}{2} XXZ model, the observed degeneracy is known to arise from the enhanced quantum group symmetry Uq(𝔰𝔩2)U_{q}(\mathfrak{sl}_{2}) when the deformation parameter qq is a root of unity Pasquier and Saleur (1990); Deguchi et al. (2001). Given the structural similarity of the exact states constructed here, it is natural to speculate that the invariant subspace identified in our higher-dimensional spin-ss model may originate from a similar algebraic mechanism. However, a precise understanding of the underlying symmetry structure in this generalized setting remains an open problem in mathematical physics.

In the limit η0\eta\to 0, the XXZ model reduces to the isotropic XXX model. Consequently, all local states in |Ψ¯(s)(u,ϵ)\ket{\bar{\Psi}^{(s)}(u,\bm{\epsilon})} align with the same spin direction, resulting in the absence of chirality. In the isotropic case, the energy level E=ds2VE=ds^{2}V is 2sV+12sV+1-fold degenerate, with the degeneracy originating from the SU(2)SU(2) symmetry of the system.

The states defined in Eq. (34) provide a tractable framework for computing the exact bipartite entanglement entropy, defined as

SA=Tr(ρAlnρA),S_{A}=-\mathrm{Tr}(\rho_{A}\ln\rho_{A}), (36)

where ρA\rho_{A} is the reduced density matrix of a subsystem AA with volume VAV_{A}. The bipartite entanglement entropy of the tower state |Ψn,ϵ(s)\ket{\Psi_{n,\bm{\epsilon}}^{(s)}} takes the form Schecter and Iadecola (2019)

SA=j=12sVA(2sVAj)(2sV2sVAnj)(2sVn)ln((2sVAj)(2sV2sVAnj)(2sVn)),S_{A}=-\sum_{j=1}^{2sV_{A}}\frac{\binom{2sV_{A}}{j}\binom{2sV-2sV_{A}}{n-j}}{\binom{2sV}{n}}\ln\left(\frac{\binom{2sV_{A}}{j}\binom{2sV-2sV_{A}}{n-j}}{\binom{2sV}{n}}\right), (37)

where we restrict to the case VAn/2V_{A}\leq n/2.

In the special case n=VA=sVn=V_{A}=sV, one can analytically evaluate the asymptotic behavior of SAS_{A} in the thermodynamic limit. This yields

SA12ln(sπV4)+1,S_{A}\simeq\frac{1}{2}\ln\left(\frac{s\pi V}{4}\right)+1, (38)

demonstrating that the entanglement entropy grows logarithmically with the system size. This sub-volume scaling of entanglement entropy is markedly different from the volume-law behavior typical of thermal states and thus provides strong evidence that these exact states are QMBSs.

IV.2 Spin-ss XY model

When η=12\eta=\frac{1}{2}, the exchange coefficients are given by

Jx=θ3(0)θ4(0),Jy=θ4(0)θ3(0),Jz=0.\displaystyle J_{x}=\frac{\theta_{3}(0)}{\theta_{4}(0)},\,\,J_{y}=\frac{\theta_{4}(0)}{\theta_{3}(0)},\,\,J_{z}=0. (39)

This corresponds to an anisotropic XY model with zero longitudinal coupling. In this case, Eq. (15) becomes

Lα=4pατ+4qα.α=1,,d,pα,qα,L_{\alpha}=4p_{\alpha}\tau+4q_{\alpha}.\quad\alpha=1,\dots,d,\quad p_{\alpha},q_{\alpha}\in\mathbb{Z},

Since τ\tau has non-zero imaginary part, this equation holds only when pα=0p_{\alpha}=0, which implies Lα=4qαL_{\alpha}=4q_{\alpha}. In other words, the spin helix eigenstates exist in the XY model (39) for arbitrary spin ss, when the site number in each direction is a multiple of 4. The spin helix eigenstate now reads

|Ψ~(s)(u,ϵ)=jψj(s)(u+12ϵ𝒏j),u.\displaystyle\ket{\tilde{\Psi}^{(s)}(u,\bm{\epsilon})}=\bigotimes_{j}\psi^{(s)}_{j}\left(u+\tfrac{1}{2}\,\bm{\epsilon}\cdot\bm{n}_{j}\right),\quad u\in\mathbb{C}. (40)

The state (40) is a tensor product state composed of four local states: ψj(s)(u)\psi^{(s)}_{j}(u), ψj(s)(u+12)\psi^{(s)}_{j}(u+\tfrac{1}{2}), ψj(s)(u+1)\psi^{(s)}_{j}(u+1) and ψj(s)(u+32)\psi^{(s)}_{j}(u+\tfrac{3}{2}). From Eq. (14), we can prove that the state (40) is a zero-energy eigenstate.

It should be noted that the SHS in the XY model is still modulated by the theta function, and is significantly different from its XXZ counterpart. To the best of our knowledge, no simpler expression of Eq. (40) is available in generic case. However, for specific values of uu, the SHS can be considerably simplified. Let u=k+τ2u=\frac{k+\tau}{2}, one gets

θ~4(τ+k+12)θ~1(τ+k+12)=ik.\displaystyle\frac{\tilde{\theta}_{4}(\frac{\tau+k+1}{2})}{\tilde{\theta}_{1}(\frac{\tau+k+1}{2})}={\mathrm{i}}^{k}. (41)

It follows that

ψ(s)(1+τ2)\displaystyle\psi^{(s)}\!\left(\tfrac{1+\tau}{2}\right) |sx,\displaystyle\propto\ket{s}_{x}, ψ(s)(2+τ2)\displaystyle\psi^{(s)}\!\left(\tfrac{2+\tau}{2}\right) |sy,\displaystyle\propto\ket{s}_{y},
ψ(s)(3+τ2)\displaystyle\psi^{(s)}\!\left(\tfrac{3+\tau}{2}\right) |sx,\displaystyle\propto\ket{-s}_{x}, ψ(s)(4+τ2)\displaystyle\psi^{(s)}\!\left(\tfrac{4+\tau}{2}\right) |sy,\displaystyle\propto\ket{-s}_{y}, (42)

where |mx\ket{m}_{x} and |my\ket{m}_{y} are eigenstates of SxS^{x} and SyS^{y}, respectively:

Sx|mx=m|mx,Sy|my=m|my.S^{x}\ket{m}_{x}=m\ket{m}_{x},\quad S^{y}\ket{m}_{y}=m\ket{m}_{y}.

Therefore, a family of spin helix eigenstates exists, constructed from the four fundamental local states |±sx\ket{\pm s}_{x}, |±sy\ket{\pm s}_{y} with the following properties: (1) the first local state is chosen among the four fundamental states. (2) for each spatial direction, the spin sequence evolves through either a counter-clockwise cycle in the xyxy-plane |sx|sy|sx|sy|sx\ket{s}_{x}\to\ket{s}_{y}\to\ket{-s}_{x}\to\ket{-s}_{y}\to\ket{s}_{x} or a clockwise cycle |sx|sy|sx|sy|sx\ket{s}_{x}\to\ket{-s}_{y}\to\ket{-s}_{x}\to\ket{s}_{y}\to\ket{s}_{x}. An example of these SHSs is visualized in Fig. 4.

|sx\ket{s}_{x}|sy\ket{s}_{y}|sx\ket{-s}_{x}|sy\ket{-s}_{y}|sx\ket{s}_{x}|sy\ket{-s}_{y}|sx\ket{s}_{x}|sy\ket{s}_{y}|sx\ket{-s}_{x}|sy\ket{-s}_{y}|sx\ket{-s}_{x}|sy\ket{-s}_{y}|sx\ket{s}_{x}|sy\ket{s}_{y}|sx\ket{-s}_{x}|sy\ket{s}_{y}|sx\ket{-s}_{x}|sy\ket{-s}_{y}|sx\ket{s}_{x}|sy\ket{s}_{y}|sx\ket{s}_{x}|sy\ket{s}_{y}|sx\ket{-s}_{x}|sy\ket{-s}_{y}|sx\ket{s}_{x}
Figure 4: A 2dd spin helix structure of the spin-ss XY model, where each local spin is polarized along either the ±x\pm x or ±y\pm y directions, forming a chiral winding configuration.

When η=12\eta=\frac{1}{2}, the coupling constants in Eq. (39) satisfy Jx/Jy>0J_{x}/J_{y}>0. However, if we set η=12τ\eta=\frac{1}{2}-\tau, the XYZ model reduces to a different XY model with

Jx=eiπτθ3(0)θ4(0),Jy=eiπτθ4(0)θ3(0),Jz=0.\displaystyle J_{x}={\mathrm{e}}^{-{\mathrm{i}}\pi\tau}\frac{\theta_{3}(0)}{\theta_{4}(0)},\,\,J_{y}=-{\mathrm{e}}^{-{\mathrm{i}}\pi\tau}\frac{\theta_{4}(0)}{\theta_{3}(0)},\,\,J_{z}=0. (43)

In this regime, the spin helix eigenstate (40) remains a zero-energy eigenstate of the Hamiltonian (43) under the replacement 1212τ\frac{1}{2}\to\frac{1}{2}-\tau, provided the number of lattice sites in each spatial direction is a multiple of 4.

It is necessary to clarify that the states presented in Eq. (40) do not belong to the following SU(2)SU(2)-protected exact eigenstates previously identified in the spin-1 XY model Schecter and Iadecola (2019).

|Φ(u)=jϕj(u+𝟏𝒏j2),u,\displaystyle\ket{\Phi(u)}=\bigotimes_{j}\phi_{j}(u+\tfrac{\bm{1}\cdot\bm{n}_{j}}{2}),\qquad u\in\mathbb{C}, (44)
ϕ(u)=θ~12(u)|1θ~42(u)|1,  1={1,1,,1}.\displaystyle\phi(u)=\tilde{\theta}_{1}^{2}(u)\ket{1}-\tilde{\theta}_{4}^{2}(u)\ket{-1},\,\,\bm{1}=\{1,1,\ldots,1\}.

The tensor product state in Eq. (44) exhibits similar helical properties and each local basis vector ϕj(u)\phi_{j}(u) composing the eigenstate is a superposition of |1\ket{1} and |1\ket{-1}, with {L1,,Ld}\{L_{1},\ldots,L_{d}\} restricted to even integers to match the periodic boundary condition. By contrast, our local vectors {ψj(1)(u)}\{\psi_{j}^{(1)}(u)\} are linear combinations of |1,|1\ket{1},\ket{-1} and |0\ket{0}, and {L1,,Ld}\{L_{1},\ldots,L_{d}\} are constrained to multiples of 4. Most importantly, the eigenstate introduced in Eq. (44) is valid solely for the spin-11 XY system, whereas we demonstrate a universal result applicable to arbitrary spin-ss systems. See Appendix D for more details.

The exact spin helix state exhibits hallmark features of QMBSs. However, the structure of the degenerate subspace they occupy is much more complex, and it remains difficult to express it in a clear form as is available in the spin-s XXZ model. While the explicit algebraic structure of towel states has not yet been identified, we conjecture that this subspace contains a set of nonthermal eigenstates with sub-volume entanglement entropy. This suggests the presence of a broader class of scar eigenstates beyond the current framework.

V spin helix eigenstate in the Heisenberg models with other settings

In the previous sections, we study the Heisenberg model with only nearest-neighbor interactions in a hypercubic lattice. In this section, we will demonstrate that the Heisenberg model with other settings can still have spin helix eigenstates, provided the exchange coefficients are appropriately modulated.

Let us consider the following periodic spin-ss Heisenberg model

H=kFki,jkHi,j(k),\displaystyle H=\sum_{k}F_{k}\sum_{\langle i,j\rangle_{k}}H_{i,j}^{(k)}, (45)
Hi,j(k)=Jx(k)SixSjx+Jy(k)SiySjy+Jz(k)SizSjz,\displaystyle H_{i,j}^{(k)}=J_{x}^{(k)}{S}_{i}^{x}S_{j}^{x}+J_{y}^{(k)}S_{i}^{y}S_{j}^{y}+J_{z}^{(k)}S_{i}^{z}S_{j}^{z}, (46)

where i,jk\langle i,j\rangle_{k} represents the kk-th nearest-neighbor pairs along the axial lattice directions, {Fk}\{F_{k}\} are free parameters and the exchange coefficients {Jα(k)}\{J_{\alpha}^{(k)}\} are

{Jx(k),Jy(k),Jz(k)}={θ4(kη)θ4(0),θ3(kη)θ3(0),θ2(kη)θ2(0)}.\displaystyle\left\{J_{x}^{(k)},J_{y}^{(k)},J_{z}^{(k)}\right\}=\left\{\frac{\theta_{4}(k\eta)}{\theta_{4}(0)},\frac{\theta_{3}(k\eta)}{\theta_{3}(0)},\frac{\theta_{2}(k\eta)}{\theta_{2}(0)}\right\}. (47)

Using the divergence condition (11), we can prove that the state |Ψ¯(s)(u,ϵ)\ket{\bar{\Psi}^{(s)}(u,\bm{\epsilon})} defined in Eq. (13) is an eigenstate of the Hamiltonian in Eq. (45) under the condition in Eq. (15).

We focus on the hypercubic lattice in the previous sections. However, the spin helix eigenstates can also exist in triangular and kagome lattices when the anisotropic parameter η\eta takes the following specific values Jepsen et al. (2022b)

η=2pατ3+2qα3,Lα=3nα,pα,qα,nα+.\eta=\frac{2p_{\alpha}\tau}{3}+\frac{2q_{\alpha}}{3},\quad L_{\alpha}=3n_{\alpha},\quad p_{\alpha},q_{\alpha}\in\mathbb{Z},\quad n_{\alpha}\in\mathbb{N}^{+}.

VI Summary and outlook

In this paper, we study the dd-dimensional spin-ss Heisenberg model under periodic boundary conditions. We examine both the fully anisotropic XYZ model and partially anisotropic cases, including the XXZ and XY models. It is demonstrated that the anisotropic Heisenberg model has the spin helix eigenstate

jψj(s)(u+ηϵ𝒏j).\bigotimes_{j}\psi^{(s)}_{j}\left(u+\eta\,\bm{\epsilon}\cdot\bm{n}_{j}\right).

This type of tensor-product state exhibits the following characteristics: (1) The phase factor uu in ψj(s)(u)\psi^{(s)}_{j}(u) varies linearly with the lattice coordinate. (2) The phase difference between neighboring sites is ±η\pm\eta, which is compatible with the anisotropic parameters. (3) The parameter η\eta and the system length {L1,,Ld}\{L_{1},\ldots,L_{d}\} must satisfy specific constraints (Eqs. (15) and (27)) to comply with the periodic boundary conditions.

An interesting follow-up study concerns the spin helix eigenstate of the Heisenberg chain under other boundary conditions. For instance, one can verify that the following one-dimensional open Heisenberg chain Hamiltonian

H=j=1L1Hj,j+1a(u0+η)S1z+a(u0+Lη)SLz,\displaystyle H=\sum_{j=1}^{L-1}H_{j,j+1}-a(u_{0}+\eta)S_{1}^{z}+a(u_{0}+L\eta)S_{L}^{z}, (48)

admits a SHS of the form n=1Lψn(s)(u0+nη)\bigotimes_{n=1}^{L}\psi_{n}^{(s)}(u_{0}+n\eta). Unlike the periodic system, the phase factor is uniquely determined by the boundary magnetic fields.

Another open question is the degeneracy of the spin helix eigenstates. In the context of the XXZ model, we constructed a set of mutually independent vectors to explain its degeneracy. Although numerical results on small-size systems suggest a similar degeneracy structure in the XYZ case, this problem remains unsolved for high-dimensional and high-spin systems. In Appendix E, we outline a feasible method to expand the spin helix eigenstate and derive some independent vectors.

Note added: After completing this work, we became aware of a recent preprint with some overlapping content Bhowmick et al. . They note that the identification of spin helix eigenstates in the one-dimensional high-spin XYZ model dating back to 1985 Granovskii and Zhedanov (1985a, b). Our study was conducted independently, employs a different set of notations, and provides a more detailed analysis of spin helix eigenstates in the XXZ and XY limits, as well as extensions to higher spatial dimensions.

Acknowledgements.
We thank Y. Wan for valuable discussions. M.Z. acknowledges financial support from the National Natural Science Foundation of China (No. 12447130). X.Z. acknowledges financial support from the National Natural Science Foundation of China (No. 12204519).

References

Appendix A Jacobi theta functions

Some useful identities for the elliptic functions used in this paper are

θ2(u)=θ1(u+12),θ3(u)=eiπ(u+τ4)θ1(u+1+τ2),θ4(u)=eiπ(u+τ4+12)θ1(u+τ2),\displaystyle\theta_{2}(u)=\theta_{1}(u+\tfrac{1}{2}),\quad\theta_{3}(u)={\mathrm{e}}^{{\mathrm{i}}\pi(u+\frac{\tau}{4})}\theta_{1}(u+\tfrac{1+\tau}{2}),\quad\theta_{4}(u)=-{\mathrm{e}}^{{\mathrm{i}}\pi(u+\frac{\tau}{4}+\frac{1}{2})}\theta_{1}(u+\tfrac{\tau}{2}), (A1)
θ1(u)=θ1(u),θα(u)=θα(u),α=2,3,4,\displaystyle\theta_{1}(-u)=-\theta_{1}(u),\quad\theta_{\alpha}(-u)=\theta_{\alpha}(u),\quad\alpha=2,3,4, (A2)
θα(u+1)=θα(u),θα(u+1)=θα(u),α=1,2,α=3,4,\displaystyle\theta_{\alpha}(u+1)=-\theta_{\alpha}(u),\quad\theta_{\alpha^{\prime}}(u+1)=\theta_{\alpha^{\prime}}(u),\quad\alpha=1,2,\quad\alpha^{\prime}=3,4, (A3)
θα(u+τ)=eiπ(2u+τ)θα(u),θα(u+τ)=eiπ(2u+τ)θα(u),α=1,4,α=2,3,\displaystyle\theta_{\alpha}(u+\tau)=-{\mathrm{e}}^{-{\mathrm{i}}\pi(2u+\tau)}\theta_{\alpha}(u),\quad\theta_{\alpha^{\prime}}(u+\tau)={\mathrm{e}}^{-{\mathrm{i}}\pi(2u+\tau)}\theta_{\alpha^{\prime}}(u),\quad\alpha=1,4,\quad\alpha^{\prime}=2,3, (A4)
θ~1(2u)θ~4(0)=θ1(u)θ2(u)θ3(0)θ4(0),θ~4(2u)θ~4(0)=θ3(u)θ4(u)θ3(0)θ4(0),θ1(u)θ2(0)=θ~1(u)θ~4(u)θ~2(0)θ~3(0),\displaystyle\frac{\tilde{\theta}_{1}(2u)}{\tilde{\theta}_{4}(0)}=\frac{\theta_{1}(u)\theta_{2}(u)}{\theta_{3}(0)\theta_{4}(0)},\quad\frac{\tilde{\theta}_{4}(2u)}{\tilde{\theta}_{4}(0)}=\frac{\theta_{3}(u)\theta_{4}(u)}{\theta_{3}(0)\theta_{4}(0)},\quad\frac{\theta_{1}(u)}{\theta_{2}(0)}=\frac{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u)}{\tilde{\theta}_{2}(0)\tilde{\theta}_{3}(0)}, (A5)
θ~1(u+v)θ~1(uv)θ~42(0)=θ~12(u)θ~42(v)θ~12(v)θ~42(u),\displaystyle\tilde{\theta}_{1}(u+v)\tilde{\theta}_{1}(u-v)\tilde{\theta}_{4}^{2}(0)=\tilde{\theta}_{1}^{2}(u)\tilde{\theta}_{4}^{2}(v)-\tilde{\theta}_{1}^{2}(v)\tilde{\theta}_{4}^{2}(u), (A6)
θ~4(u+v)θ~4(uv)θ~42(0)=θ~42(u)θ~42(v)θ~12(v)θ~12(u),\displaystyle\tilde{\theta}_{4}(u+v)\tilde{\theta}_{4}(u-v)\tilde{\theta}_{4}^{2}(0)=\tilde{\theta}_{4}^{2}(u)\tilde{\theta}_{4}^{2}(v)-\tilde{\theta}_{1}^{2}(v)\tilde{\theta}_{1}^{2}(u), (A7)
2θ~1(u+v)θ~1(uv)=θ4(u)θ3(v)θ4(v)θ3(u),\displaystyle 2\tilde{\theta}_{1}(u+v)\tilde{\theta}_{1}(u-v)=\theta_{4}(u)\theta_{3}(v)-\theta_{4}(v)\theta_{3}(u), (A8)
2θ~4(u+v)θ~4(uv)=θ4(u)θ3(v)+θ4(v)θ3(u),\displaystyle 2\tilde{\theta}_{4}(u+v)\tilde{\theta}_{4}(u-v)=\theta_{4}(u)\theta_{3}(v)+\theta_{4}(v)\theta_{3}(u), (A9)
2θ~4(u+v)θ~1(uv)=θ1(u)θ2(v)θ1(v)θ2(u).\displaystyle 2\tilde{\theta}_{4}(u+v)\tilde{\theta}_{1}(u-v)=\theta_{1}(u)\theta_{2}(v)-\theta_{1}(v)\theta_{2}(u). (A10)

Define the following functions

θj(u)=θj(u)u,θ~j(u)=θ~j(u)u,ζ(u)=θ1(u)θ1(u),ζ~(u)=θ~1(u)θ~1(u).\displaystyle\theta^{\prime}_{j}(u)=\frac{\partial\theta_{j}(u)}{\partial u},\quad\tilde{\theta}_{j}^{\prime}(u)=\frac{\partial\tilde{\theta}_{j}(u)}{\partial u},\quad\zeta(u)=\frac{\theta_{1}^{\prime}(u)}{\theta_{1}(u)},\quad\tilde{\zeta}(u)=\frac{\tilde{\theta}_{1}^{\prime}(u)}{\tilde{\theta}_{1}(u)}. (A11)

which possess the following properties

ζ(u)=ζ(u),ζ(u+1)=ζ(u),ζ(u+τ)=ζ(u)2iπ,\displaystyle\zeta(u)=-\zeta(-u),\quad\zeta(u+1)=\zeta(u),\quad\zeta(u+\tau)=\zeta(u)-2{\mathrm{i}}\pi, (A12)
ζ~(u)=ζ~(u),ζ~(u+1)=ζ~(u),ζ~(u+2τ)=ζ~(u)2iπ,\displaystyle\tilde{\zeta}(u)=-\tilde{\zeta}(-u),\quad\tilde{\zeta}(u+1)=\tilde{\zeta}(u),\quad\tilde{\zeta}(u+2\tau)=\tilde{\zeta}(u)-2{\mathrm{i}}\pi, (A13)
2ζ~(2u)=ζ(u)+ζ(u+12),ζ(u)=iπ+ζ~(u)+ζ~(u+τ),\displaystyle 2\tilde{\zeta}(2u)=\zeta(u)+\zeta(u+\tfrac{1}{2}),\quad\zeta(u)={\mathrm{i}}\pi+\tilde{\zeta}(u)+\tilde{\zeta}(u+\tau), (A14)
2ζ(u)=2iπ+ζ(u2)+ζ(u+12)+ζ(u+τ2)+ζ(u+τ+12).\displaystyle 2\zeta(u)=2{\mathrm{i}}\pi+\zeta(\tfrac{u}{2})+\zeta(\tfrac{u+1}{2})+\zeta(\tfrac{u+\tau}{2})+\zeta(\tfrac{u+\tau+1}{2}). (A15)

The functions θα(u),θ~α(u),ζ(u),ζ¯(u)\theta_{\alpha}(u),\,\tilde{\theta}_{\alpha}(u),\,\zeta(u),\,\bar{\zeta}(u) satisfy the following identities

θ2(u)θ1(u)=θ2(0)θ1(0)[ζ(u2)+ζ(u+12)ζ(u)],\displaystyle\frac{\theta_{2}(u)}{\theta_{1}(u)}=\frac{\theta_{2}(0)}{\theta_{1}^{\prime}(0)}\left[\zeta(\tfrac{u}{2})+\zeta(\tfrac{u+1}{2})-\zeta(u)\right], (A16)
θ~4(η)θ~1(x1+x2)θ~1(x1+η)θ~1(x2+η)θ~4(0)θ~1(x1)θ~1(x2)θ~1(x1+x2+η)=θ1(η)θ1(0)[ζ~(x1)+ζ~(x2)+ζ~(η)ζ~(x1+x2+η)].\displaystyle\frac{\tilde{\theta}_{4}(\eta)\tilde{\theta}_{1}(x_{1}+x_{2})\tilde{\theta}_{1}(x_{1}+\eta)\tilde{\theta}_{1}(x_{2}+\eta)}{\tilde{\theta}_{4}(0)\tilde{\theta}_{1}(x_{1})\tilde{\theta}_{1}(x_{2})\tilde{\theta}_{1}(x_{1}+x_{2}+\eta)}=\frac{\theta_{1}(\eta)}{\theta_{1}^{\prime}(0)}\left[\tilde{\zeta}(x_{1})+\tilde{\zeta}(x_{2})+\tilde{\zeta}(\eta)-\tilde{\zeta}(x_{1}+x_{2}+\eta)\right]. (A17)

Appendix B Proof of Eqs. (8) and (10)

Define the following operator

𝕊(x,y)=sin(x)cos(y)Sx+sin(x)sin(y)Sy+cos(x)Sz.\displaystyle\mathbb{S}(x,y)=\sin(x)\cos(y)S_{x}+\sin(x)\sin(y)S_{y}+\cos(x)S_{z}. (B18)

Suppose

θ~4(u)θ~1(u)=eα(u)+iβ(u),α(u),β(u),\displaystyle\frac{\tilde{\theta}_{4}(u)}{\tilde{\theta}_{1}(u)}={\mathrm{e}}^{\alpha(u)+{\mathrm{i}}\beta(u)},\quad\alpha(u),\beta(u)\in\mathbb{R}, (B19)
or equivalentlyβ(u)=argθ~4(u)θ~1(u),eα(u)=|θ~4(u)θ~1(u)|.\displaystyle\mbox{or equivalently}\,\,\beta(u)=\arg\frac{\tilde{\theta}_{4}(u)}{\tilde{\theta}_{1}(u)},\quad{\mathrm{e}}^{\alpha(u)}=\left|\frac{\tilde{\theta}_{4}(u)}{\tilde{\theta}_{1}(u)}\right|. (B20)

Then, the operator 𝕊(γ(u),β(u))\mathbb{S}(\gamma(u),\beta(u)) reads

𝕊(γ,β)=eiβeα+eαS+eiβeα+eαS+eαeαeα+eαSz,tan(γ2)=eα.\displaystyle\mathbb{S}(\gamma,\beta)=\frac{{\mathrm{e}}^{{\mathrm{i}}\beta}}{{\mathrm{e}}^{\alpha}+{\mathrm{e}}^{-\alpha}}S_{-}+\frac{{\mathrm{e}}^{-{\mathrm{i}}\beta}}{{\mathrm{e}}^{\alpha}+{\mathrm{e}}^{-\alpha}}S_{+}-\frac{{\mathrm{e}}^{\alpha}-{\mathrm{e}}^{-\alpha}}{{\mathrm{e}}^{\alpha}+{\mathrm{e}}^{-\alpha}}S_{z},\quad\tan\left(\frac{\gamma}{2}\right)={\mathrm{e}}^{\alpha}. (B21)

Here and below, we omit the parameter uu for convenience. One can derive

𝕊(γ,β)ψ(s)(u)\displaystyle\quad\mathbb{S}(\gamma,\beta)\psi^{(s)}(u)
=𝒳𝕊(γ,β){mκsm[θ~1(u)θ~4(u)]m|m}\displaystyle=\mathcal{X}\mathbb{S}(\gamma,\beta)\left\{\sum_{m}\kappa_{s-m}\left[\frac{\tilde{\theta}_{1}(u)}{\tilde{\theta}_{4}(u)}\right]^{m}\ket{m}\right\}
=𝒳𝕊(γ,β){mκsmemαimβ|m}\displaystyle=\mathcal{X}\mathbb{S}(\gamma,\beta)\left\{\sum_{m}\kappa_{s-m}{\mathrm{e}}^{-m\alpha-{\mathrm{i}}m\beta}\ket{m}\right\}
=(B21)𝒳{mκsmeαeαeα+eαemαimβm|m+mκsm+1emα+αimβeα+eαλm1+|m\displaystyle\overset{(\ref{OperatorS-2})}{=}\mathcal{X}\left\{-\sum_{m}\kappa_{s-m}\frac{{\mathrm{e}}^{\alpha}-{\mathrm{e}}^{-\alpha}}{{\mathrm{e}}^{\alpha}+{\mathrm{e}}^{-\alpha}}{\mathrm{e}}^{-m\alpha-{\mathrm{i}}m\beta}m\ket{m}+\sum_{m}\kappa_{s-m+1}\frac{{\mathrm{e}}^{-m\alpha+\alpha-{\mathrm{i}}m\beta}}{{\mathrm{e}}^{\alpha}+{\mathrm{e}}^{-\alpha}}\lambda_{m-1}^{+}\ket{m}\right.
+mκsm1emααimβeα+eαλm+1|m}\displaystyle\quad+\left.\sum_{m}\kappa_{s-m-1}\frac{{\mathrm{e}}^{-m\alpha-\alpha-{\mathrm{i}}m\beta}}{{\mathrm{e}}^{\alpha}+{\mathrm{e}}^{-\alpha}}\lambda_{m+1}^{-}\ket{m}\right\}
=(C31)𝒳{smemαimβ|m}=sψ(s)(u),\displaystyle\overset{(\ref{lambda;kappa})}{=}\mathcal{X}\left\{s\sum_{m}{\mathrm{e}}^{-m\alpha-{\mathrm{i}}m\beta}\ket{m}\right\}=s\psi^{(s)}(u), (B22)

where 𝒳=[θ~1(u)θ~4(u)]s/𝒩(u)\mathcal{X}=[\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u)]^{s}/\mathcal{N}(u). Equation (B22) demonstrates that ψ(s)(u)\psi^{(s)}(u) is the highest-weight state of 𝕊(γ(u),β(u))\mathbb{S}(\gamma(u),\beta(u)), and this yields Eqs. (8) and (10).

Appendix C Proof of Eq. (11)

The SzS^{z} basis {|m}\{\ket{m}\} satisfy

Sz|m=m|m,m=s,s1,,s,\displaystyle S^{z}\ket{m}=m\ket{m},\quad m=s,s-1,\ldots,-s, (C23)
S+|m=λm+|m+1,λm+=(sm)(s+m+1),\displaystyle S^{+}\ket{m}=\lambda^{+}_{m}\ket{m+1},\quad\lambda^{+}_{m}=\sqrt{(s-m)(s+m+1)}, (C24)
S|m=λm|m1,λm=(s+m)(sm+1),\displaystyle S^{-}\ket{m}=\lambda^{-}_{m}\ket{m-1},\quad\lambda^{-}_{m}=\sqrt{(s+m)(s-m+1)}, (C25)

where S±=Sx±iSyS^{\pm}=S^{x}\pm{\mathrm{i}}S^{y}. Let us recall the vector ψ(s)(u)\psi^{(s)}(u)

ψ(s)(u)=m=ssκsmwm(u)|m,wm(u)=[θ~1(u)]s+m[θ~4(u)]sm\displaystyle\psi^{(s)}(u)=\sum_{m=-s}^{s}\kappa_{s-m}w_{m}(u)\ket{m},\quad w_{m}(u)=\left[\tilde{\theta}_{1}(u)\right]^{s+m}\left[\tilde{\theta}_{4}(u)\right]^{s-m} (C26)

Here we omit the normalization factor for convenience. Rewrite the local Hamiltonian Hi,jH_{i,j} as

J+4(SiSj++Si+Sj)+J4(Si+Sj++SiSj)+JzSizSjz,\displaystyle\frac{J_{+}}{4}(S_{i}^{-}S_{j}^{+}+S_{i}^{+}S_{j}^{-})+\frac{J_{-}}{4}(S_{i}^{+}S_{j}^{+}+S_{i}^{-}S_{j}^{-})+J_{z}S_{i}^{z}S_{j}^{z},\quad (C27)

where

J+=Jx+Jy=(A9)2θ~42(η)θ~42(0),J=JxJy=(A8)2θ~12(η)θ~42(0).\displaystyle J_{+}=J_{x}+J_{y}\overset{(\ref{theta;44})}{=}\frac{2\tilde{\theta}_{4}^{2}(\eta)}{\tilde{\theta}_{4}^{2}(0)},\quad J_{-}=J_{x}-J_{y}\overset{(\ref{theta;11})}{=}\frac{2\tilde{\theta}_{1}^{2}(\eta)}{\tilde{\theta}_{4}^{2}(0)}. (C28)

Let Lm,n(u)L_{m,n}(u) and Rm,n(u)R_{m,n}(u) represent the overlap of m|in|j\bra{m}_{i}\otimes\bra{n}_{j} with the left and right sides of Eq. (11) (here we choose the ++ sign), respectively. One can easily obtain

Rm,n(u)κsmκsnwm(u)wn(u+η)=sma(u)sna(u+η)+s2b(u).\displaystyle\quad\frac{R_{m,n}(u)}{\kappa_{s-m}\kappa_{s-n}w_{m}(u)w_{n}(u+\eta)}=sma(u)-sna(u+\eta)+s^{2}b(u). (C29)

After some calculations, we derive

Lm,n(u)κsmκsnwm(u)wn(u+η)\displaystyle\quad\frac{L_{m,n}(u)}{\kappa_{s-m}\kappa_{s-n}w_{m}(u)w_{n}(u+\eta)}
=Jzmn+J+4λm+1λn1+κsm1κsn+1wm+1(u)wn1(u+η)κsmκsnwm(u)wn(u+η)\displaystyle=J_{z}mn+\frac{J_{+}}{4}\frac{\lambda^{-}_{m+1}\lambda^{+}_{n-1}\kappa_{s-m-1}\kappa_{s-n+1}w_{m+1}(u)w_{n-1}(u+\eta)}{\kappa_{s-m}\kappa_{s-n}w_{m}(u)w_{n}(u+\eta)}
+J+4λm1+λn+1κsm1κsn+1wm1(u)wn+1(u+η)κsmκsnwm(u)wn(u+η)\displaystyle\quad+\frac{J_{+}}{4}\frac{\lambda^{+}_{m-1}\lambda^{-}_{n+1}\kappa_{s-m-1}\kappa_{s-n+1}w_{m-1}(u)w_{n+1}(u+\eta)}{\kappa_{s-m}\kappa_{s-n}w_{m}(u)w_{n}(u+\eta)}
+J4λm+1λn+1κsm1κsn1wm+1(u)wn+1(u+η)κsmκsnwm(u)wn(u+η)\displaystyle\quad+\frac{J_{-}}{4}\frac{\lambda^{-}_{m+1}\lambda^{-}_{n+1}\kappa_{s-m-1}\kappa_{s-n-1}w_{m+1}(u)w_{n+1}(u+\eta)}{\kappa_{s-m}\kappa_{s-n}w_{m}(u)w_{n}(u+\eta)}
+J4λm1+λn1+κsm+1κsn+1wm1(u)wn1(u+η)κsmκsnwm(u)wn(u+η)\displaystyle\quad+\frac{J_{-}}{4}\frac{\lambda^{+}_{m-1}\lambda^{+}_{n-1}\kappa_{s-m+1}\kappa_{s-n+1}w_{m-1}(u)w_{n-1}(u+\eta)}{\kappa_{s-m}\kappa_{s-n}w_{m}(u)w_{n}(u+\eta)}
=Jzmn+J+4(sm)(s+n)wm+1(u)wn1(u+η)wm(u)wn(u+η)+J+4(s+m)(sn)wm1(u)wn+1(u+η)wm(u)wn(u+η)\displaystyle=J_{z}mn+\frac{J_{+}}{4}(s-m)(s+n)\frac{w_{m+1}(u)w_{n-1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}+\frac{J_{+}}{4}(s+m)(s-n)\frac{w_{m-1}(u)w_{n+1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}
+J4(sm)(sn)wm+1(u)wn+1(u+η)wm(u)wn(u+η)+J4(s+m)(s+n)wm1(u)wn1(u+η)wm(u)wn(u+η)\displaystyle\quad+\frac{J_{-}}{4}(s-m)(s-n)\frac{w_{m+1}(u)w_{n+1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}+\frac{J_{-}}{4}(s+m)(s+n)\frac{w_{m-1}(u)w_{n-1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}
=s2s,s(u)+sms,m(u)+sns,n(u)+mnm,n(u).\displaystyle=s^{2}{\cal F}_{s,s}(u)+sm{\cal F}_{s,m}(u)+sn{\cal F}_{s,n}(u)+mn{\cal F}_{m,n}(u). (C30)

In Eq. (C30), we use the identities

λk+1κsk1=(sk)κsk,λk1+κsk+1=(s+k)κsk.\displaystyle\lambda^{-}_{k+1}\kappa_{s-k-1}=(s-k)\kappa_{s-k},\quad\lambda^{+}_{k-1}\kappa_{s-k+1}=(s+k)\kappa_{s-k}. (C31)

The function m,n(u){\cal F}_{m,n}(u) in Eq. (C30) is

JzJ+4wm+1(u)wn1(u+η)wm(u)wn(u+η)J+4wm1(u)wn+1(u+η)wm(u)wn(u+η)\displaystyle J_{z}-\frac{J_{+}}{4}\frac{w_{m+1}(u)w_{n-1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}-\frac{J_{+}}{4}\frac{w_{m-1}(u)w_{n+1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}
+J4wm+1(u)wn+1(u+η)wm(u)wn(u+η)+J4wm1(u)wn1(u+η)wm(u)wn(u+η)\displaystyle+\frac{J_{-}}{4}\frac{w_{m+1}(u)w_{n+1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}+\frac{J_{-}}{4}\frac{w_{m-1}(u)w_{n-1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}
=Jzθ~42(η)2θ~42(0)θ~1(u)θ~4(u+η)θ~4(u)θ~1(u+η)+θ~12(η)2θ~42(0)θ~4(u)θ~4(u+η)θ~1(u)θ~1(u+η)\displaystyle=J_{z}-\frac{\tilde{\theta}_{4}^{2}(\eta)}{2\tilde{\theta}_{4}^{2}(0)}\frac{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u+\eta)}{\tilde{\theta}_{4}(u)\tilde{\theta}_{1}(u+\eta)}+\frac{\tilde{\theta}_{1}^{2}(\eta)}{2\tilde{\theta}_{4}^{2}(0)}\frac{\tilde{\theta}_{4}(u)\tilde{\theta}_{4}(u+\eta)}{\tilde{\theta}_{1}(u)\tilde{\theta}_{1}(u+\eta)}
θ~42(η)2θ~42(0)θ~4(u)θ~1(u+η)θ~1(u)θ~4(u+η)+θ~12(η)2θ~42(0)θ~1(u)θ~1(u+η)θ~4(u)θ~4(u+η)\displaystyle\quad-\frac{\tilde{\theta}_{4}^{2}(\eta)}{2\tilde{\theta}_{4}^{2}(0)}\frac{\tilde{\theta}_{4}(u)\tilde{\theta}_{1}(u+\eta)}{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u+\eta)}+\frac{\tilde{\theta}_{1}^{2}(\eta)}{2\tilde{\theta}_{4}^{2}(0)}\frac{\tilde{\theta}_{1}(u)\tilde{\theta}_{1}(u+\eta)}{\tilde{\theta}_{4}(u)\tilde{\theta}_{4}(u+\eta)}
=(A6),(A7)Jz12θ~1(uη)θ~4(u+η)θ~1(u)θ~4(u)12θ~4(uη)θ~1(u+η)θ~1(u)θ~4(u)\displaystyle\overset{(\ref{th_th;1}),(\ref{th_th;2})}{=}J_{z}-\frac{1}{2}\frac{\tilde{\theta}_{1}(u-\eta)\tilde{\theta}_{4}(u+\eta)}{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u)}-\frac{1}{2}\frac{\tilde{\theta}_{4}(u-\eta)\tilde{\theta}_{1}(u+\eta)}{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u)}
=(A10)Jzθ1(u)θ2(η)θ1(u)θ2(0)=0.\displaystyle\overset{(\ref{theta;14})}{=}J_{z}-\frac{\theta_{1}(u)\theta_{2}(\eta)}{\theta_{1}(u)\theta_{2}(0)}=0. (C32)

The function s,m(u){\cal F}_{s,m}(u) in Eq. (C30) is

J+4wm+1(u)wn1(u+η)wm(u)wn(u+η)+J+4wm1(u)wn+1(u+η)wm(u)wn(u+η)\displaystyle-\frac{J_{+}}{4}\frac{w_{m+1}(u)w_{n-1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}+\frac{J_{+}}{4}\frac{w_{m-1}(u)w_{n+1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}
J4wm+1(u)wn+1(u+η)wm(u)wn(u+η)+J4wm1(u)wn1(u+η)wm(u)wn(u+η)\displaystyle\quad-\frac{J_{-}}{4}\frac{w_{m+1}(u)w_{n+1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}+\frac{J_{-}}{4}\frac{w_{m-1}(u)w_{n-1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}
=θ~42(η)2θ~42(0)θ~1(u)θ~4(u+η)θ~4(u)θ~1(u+η)+θ~12(η)2θ~42(0)θ~4(u)θ~4(u+η)θ~1(u)θ~1(u+η)\displaystyle=-\frac{\tilde{\theta}_{4}^{2}(\eta)}{2\tilde{\theta}_{4}^{2}(0)}\frac{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u+\eta)}{\tilde{\theta}_{4}(u)\tilde{\theta}_{1}(u+\eta)}+\frac{\tilde{\theta}_{1}^{2}(\eta)}{2\tilde{\theta}_{4}^{2}(0)}\frac{\tilde{\theta}_{4}(u)\tilde{\theta}_{4}(u+\eta)}{\tilde{\theta}_{1}(u)\tilde{\theta}_{1}(u+\eta)}
+θ~42(η)2θ~42(0)θ~4(u)θ~1(u+η)θ~1(u)θ~4(u+η)θ~12(η)2θ~42(0)θ~1(u)θ~1(u+η)θ~4(u)θ~4(u+η)\displaystyle\quad+\frac{\tilde{\theta}_{4}^{2}(\eta)}{2\tilde{\theta}_{4}^{2}(0)}\frac{\tilde{\theta}_{4}(u)\tilde{\theta}_{1}(u+\eta)}{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u+\eta)}-\frac{\tilde{\theta}_{1}^{2}(\eta)}{2\tilde{\theta}_{4}^{2}(0)}\frac{\tilde{\theta}_{1}(u)\tilde{\theta}_{1}(u+\eta)}{\tilde{\theta}_{4}(u)\tilde{\theta}_{4}(u+\eta)}
=(A6),(A7)12θ~1(uη)θ~4(u+η)θ~1(u)θ~4(u)+12θ~4(uη)θ~1(u+η)θ~1(u)θ~4(u)\displaystyle\overset{(\ref{th_th;1}),(\ref{th_th;2})}{=}-\frac{1}{2}\frac{\tilde{\theta}_{1}(u-\eta)\tilde{\theta}_{4}(u+\eta)}{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u)}+\frac{1}{2}\frac{\tilde{\theta}_{4}(u-\eta)\tilde{\theta}_{1}(u+\eta)}{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u)}
=(A10)a(u).\displaystyle\overset{(\ref{theta;14})}{=}a(u). (C33)

The function s,n(u){\cal F}_{s,n}(u) in Eq. (C30) is

J+4wm+1(u)wn1(u+η)wm(u)wn(u+η)J+4wm1(u)wn+1(u+η)wm(u)wn(u+η)\displaystyle\frac{J_{+}}{4}\frac{w_{m+1}(u)w_{n-1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}-\frac{J_{+}}{4}\frac{w_{m-1}(u)w_{n+1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}
J4wm+1(u)wn+1(u+η)wm(u)wn(u+η)+J4wm1(u)wn1(u+η)wm(u)wn(u+η)\displaystyle\quad-\frac{J_{-}}{4}\frac{w_{m+1}(u)w_{n+1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}+\frac{J_{-}}{4}\frac{w_{m-1}(u)w_{n-1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}
=θ~42(η)2θ~42(0)θ~1(u)θ~4(u+η)θ~4(u)θ~1(u+η)θ~12(η)2θ~42(0)θ~1(u)θ~1(u+η)θ~4(u)θ~4(u+η)\displaystyle=\frac{\tilde{\theta}_{4}^{2}(\eta)}{2\tilde{\theta}_{4}^{2}(0)}\frac{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u+\eta)}{\tilde{\theta}_{4}(u)\tilde{\theta}_{1}(u+\eta)}-\frac{\tilde{\theta}_{1}^{2}(\eta)}{2\tilde{\theta}_{4}^{2}(0)}\frac{\tilde{\theta}_{1}(u)\tilde{\theta}_{1}(u+\eta)}{\tilde{\theta}_{4}(u)\tilde{\theta}_{4}(u+\eta)}
θ~42(η)2θ~42(0)θ~4(u)θ~1(u+η)θ~1(u)θ~4(u+η)+θ~12(η)2θ~42(0)θ~4(u)θ~4(u+η)θ~1(u)θ~1(u+η)\displaystyle\quad-\frac{\tilde{\theta}_{4}^{2}(\eta)}{2\tilde{\theta}_{4}^{2}(0)}\frac{\tilde{\theta}_{4}(u)\tilde{\theta}_{1}(u+\eta)}{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u+\eta)}+\frac{\tilde{\theta}_{1}^{2}(\eta)}{2\tilde{\theta}_{4}^{2}(0)}\frac{\tilde{\theta}_{4}(u)\tilde{\theta}_{4}(u+\eta)}{\tilde{\theta}_{1}(u)\tilde{\theta}_{1}(u+\eta)}
=(A6),(A7)12θ~1(u)θ~4(u+2η)θ~1(u+η)θ~4(u+η)12θ~4(u)θ~1(u+2η)θ~1(u+η)θ~4(u+η)\displaystyle\overset{(\ref{th_th;1}),(\ref{th_th;2})}{=}\frac{1}{2}\frac{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u+2\eta)}{\tilde{\theta}_{1}(u+\eta)\tilde{\theta}_{4}(u+\eta)}-\frac{1}{2}\frac{\tilde{\theta}_{4}(u)\tilde{\theta}_{1}(u+2\eta)}{\tilde{\theta}_{1}(u+\eta)\tilde{\theta}_{4}(u+\eta)}
=(A10)a(u+η).\displaystyle\overset{(\ref{theta;14})}{=}-a(u+\eta). (C34)

The function s,s(u){\cal F}_{s,s}(u) in Eq. (C30) is

J+4wm+1(u)wn1(u+η)wm(u)wn(u+η)+J+4wm1(u)wn+1(u+η)wm(u)wn(u+η)\displaystyle\frac{J_{+}}{4}\frac{w_{m+1}(u)w_{n-1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}+\frac{J_{+}}{4}\frac{w_{m-1}(u)w_{n+1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}
+J4wm+1(u)wn+1(u+η)wm(u)wn(u+η)+J4wm1(u)wn1(u+η)wm(u)wn(u+η)\displaystyle+\frac{J_{-}}{4}\frac{w_{m+1}(u)w_{n+1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}+\frac{J_{-}}{4}\frac{w_{m-1}(u)w_{n-1}(u+\eta)}{w_{m}(u)w_{n}(u+\eta)}
=(C32)Jz+J+2[θ~4(u)θ~1(u+η)θ~1(u)θ~4(u+η)+θ~1(u)θ~4(u+η)θ~4(u)θ~1(u+η)]\displaystyle\overset{(\ref{Fmn})}{=}-J_{z}+\frac{J_{+}}{2}\left[\frac{\tilde{\theta}_{4}(u)\tilde{\theta}_{1}(u+\eta)}{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u+\eta)}+\frac{\tilde{\theta}_{1}(u)\tilde{\theta}_{4}(u+\eta)}{\tilde{\theta}_{4}(u)\tilde{\theta}_{1}(u+\eta)}\right]
=(A16)θ1(η)θ1(0)[ζ(η)2ζ~(η)]+θ~4(η)θ~1(η+τ)θ~4(0)θ~1(τ)[θ~1(u+τ)θ~1(u+η)θ~1(u)θ~1(u+η+τ)+θ~1(u+2τ)θ~1(u+η+τ)θ~1(u+τ)θ~1(u+η+2τ)]\displaystyle\overset{(\ref{zeta;sigma;1})}{=}\frac{\theta_{1}(\eta)}{\theta_{1}^{\prime}(0)}[\zeta(\eta)-2\tilde{\zeta}(\eta)]+\frac{\tilde{\theta}_{4}(\eta)\tilde{\theta}_{1}(\eta+\tau)}{\tilde{\theta}_{4}(0)\tilde{\theta}_{1}(\tau)}\left[\frac{\tilde{\theta}_{1}(u+\tau)\tilde{\theta}_{1}(u+\eta)}{\tilde{\theta}_{1}(u)\tilde{\theta}_{1}(u+\eta+\tau)}+\frac{\tilde{\theta}_{1}(u+2\tau)\tilde{\theta}_{1}(u+\eta+\tau)}{\tilde{\theta}_{1}(u+\tau)\tilde{\theta}_{1}(u+\eta+2\tau)}\right]
=(A17)θ1(η)θ1(0)[ζ(η)+ζ~(u)+ζ~(u+τ)ζ~(u+η+τ)ζ~(u+η)]\displaystyle\overset{(\ref{zeta;sigma;5})}{=}\frac{\theta_{1}(\eta)}{\theta_{1}^{\prime}(0)}[\zeta(\eta)+\tilde{\zeta}(u)+\tilde{\zeta}(u+\tau)-\tilde{\zeta}(u+\eta+\tau)-\tilde{\zeta}(u+\eta)]
=(A14)θ1(η)θ1(0)[ζ(η)+ζ(u)ζ(u+η)]=g(η)+g(u)g(u+η).\displaystyle\overset{(\ref{zeta;3})}{=}\frac{\theta_{1}(\eta)}{\theta_{1}^{\prime}(0)}[\zeta(\eta)+\zeta(u)-\zeta(u+\eta)]=g(\eta)+g(u)-g(u+\eta). (C35)

Substituting Eqs. (C32) -(C35) into Eq. (C30), we conclude that

Lm,n(u)=Rm,n(u),\displaystyle L_{m,n}(u)=R_{m,n}(u), (C36)

which gives the ++ sign version of Eq. (11). The corresponding - sign result is obtained by noting the Hamiltonian is an even function of η\eta.

Appendix D Another spin helix eigenstates of the spin-1 XY model

There exists another spin helix eigenstates in the spin-1 XY model, as addressed in Ref. Schecter and Iadecola (2019). Introduce two vectors

φ(u)=|1+ρ(u)|1,φ¯(u)=|1+ρ¯(u)|1.\displaystyle\varphi(u)=\ket{1}+\rho(u)\ket{-1},\quad\bar{\varphi}(u)=\ket{1}+\bar{\rho}(u)\ket{-1}. (D37)

For the spin-1 XY model, we can prove

Hi,jφi(u)φ¯j(u)=Hi,jφ¯i(u)φj(u)=0,whenρ¯(u)=J+ρ(u)+JJρ(u)+J+.\displaystyle H_{i,j}\varphi_{i}(u)\,\bar{\varphi}_{j}(u)=H_{i,j}\bar{\varphi}_{i}(u)\,\varphi_{j}(u)=0,\quad\mbox{when}\,\,\bar{\rho}(u)=-\frac{J_{+}\rho(u)+J_{-}}{J_{-}\rho(u)+J_{+}}. (D38)

Suppose

ρ(u)=θ~42(u)θ~12(u).\displaystyle\rho(u)=-\frac{\tilde{\theta}_{4}^{2}(u)}{\tilde{\theta}_{1}^{2}(u)}. (D39)

Then, we can get

ρ¯(u)\displaystyle\bar{\rho}(u) =J+ρ(u)+JJρ(u)+J+=J+θ~42(u)+Jθ~12(u)Jθ~42(u)+J+θ~12(u)\displaystyle=-\frac{J_{+}\rho(u)+J_{-}}{J_{-}\rho(u)+J_{+}}=-\frac{-J_{+}\tilde{\theta}_{4}^{2}(u)+J_{-}\tilde{\theta}_{1}^{2}(u)}{-J_{-}\tilde{\theta}_{4}^{2}(u)+J_{+}\tilde{\theta}_{1}^{2}(u)}
=(C28)θ~42(12)θ~42(u)+θ~12(12)θ~12(u)θ~12(12)θ~42(u)+θ~42(12)θ~12(u)=(A6),(A7)θ~32(u)θ~22(u).\displaystyle\overset{(\ref{Jpm})}{=}-\frac{-\tilde{\theta}_{4}^{2}(\frac{1}{2})\tilde{\theta}_{4}^{2}(u)+\tilde{\theta}_{1}^{2}(\frac{1}{2})\tilde{\theta}_{1}^{2}(u)}{-\tilde{\theta}_{1}^{2}(\frac{1}{2})\tilde{\theta}_{4}^{2}(u)+\tilde{\theta}_{4}^{2}(\frac{1}{2})\tilde{\theta}_{1}^{2}(u)}\overset{(\ref{th_th;1}),(\ref{th_th;2})}{=}-\frac{\tilde{\theta}_{3}^{2}(u)}{\tilde{\theta}_{2}^{2}(u)}. (D40)

Therefore, for the spin-1 XY model with even {L1,,Ld}\{L_{1},\ldots,L_{d}\}, the Hamiltonian has the following eigenstate (E=0E=0)

|Φ(u)=j[θ~12(u+𝟏𝒏j2)|1jθ~42(u+𝟏𝒏j2)|1j],u.\displaystyle\ket{\Phi(u)}=\bigotimes_{j}\left[\,\tilde{\theta}_{1}^{2}(u+\tfrac{\bm{1}\cdot\bm{n}_{j}}{2})\ket{1}_{j}-\tilde{\theta}_{4}^{2}(u+\tfrac{\bm{1}\cdot\bm{n}_{j}}{2})\ket{-1}_{j}\right],\quad u\in\mathbb{C}. (D41)

In the XX case, we get J=0J_{-}=0, ρ¯(u)=ρ(u)\bar{\rho}(u)=-\rho(u). Consequently, the eigenstate degenerates into Schecter and Iadecola (2019)

|Φ(u)=j[|1j+uexp(π𝟏𝒏j)|1j],u.\displaystyle\ket{\Phi(u)}=\bigotimes_{j}\left[\ket{1}_{j}+u\exp(\pi\bm{1}\cdot\bm{n}_{j})\ket{-1}_{j}\right],\quad u\in\mathbb{C}. (D42)

Appendix E Expansion of the SHS

Define the following functions

Q(u)=θ~1(u)θ~4(u),P(u)=1πθ~2(0)θ~3(0)Q(u)u.\displaystyle Q(u)=\frac{\tilde{\theta}_{1}(u)}{\tilde{\theta}_{4}(u)},\quad P(u)=\frac{1}{\pi\tilde{\theta}_{2}(0)\tilde{\theta}_{3}(0)}\frac{\partial Q(u)}{\partial u}. (E43)

which satisfy the following identities

Q(u+v)=Q(u)P(v)+Q(v)P(u)1Q2(u)Q2(v),\displaystyle Q(u+v)=\frac{Q(u)P(v)+Q(v)P(u)}{1-Q^{2}(u)Q^{2}(v)}, (E44)
P2(u)=[1Q2(u)Q2(12)][1Q2(12)Q2(u)].\displaystyle P^{2}(u)=\left[1-\frac{Q^{2}(u)}{Q^{2}(\frac{1}{2})}\right]\left[1-Q^{2}(\tfrac{1}{2})Q^{2}(u)\right]. (E45)

Now the state ψ(s)(u+kη)\psi^{(s)}(u+k\eta) can be rewritten as

ψ(s)(u+kη)n=02sκn[Q(u)P(kη)+Q(kη)P(u)]2sn[1Q2(u)Q2(kη)]n|sn.\displaystyle\psi^{(s)}(u+k\eta)\propto\sum_{n=0}^{2s}\kappa_{n}\left[Q(u)P(k\eta)+Q(k\eta)P(u)\right]^{2s-n}\left[1-Q^{2}(u)Q^{2}(k\eta)\right]^{n}\ket{s-n}. (E46)

From Eq. (E45), we see that P2(u)P^{2}(u) depends on Q(u)Q(u). As a consequence, we can expand the spin helix state with Qn(u)Q^{n}(u) and P(u)Qn(u)P(u)Q^{n}(u) as follows

|Ψ(s)(u,ϵ)\displaystyle\ket{\Psi^{(s)}(u,\bm{\epsilon})} jψj(s)(u+ηϵ𝒏j)\displaystyle\propto\bigotimes_{j}\psi^{(s)}_{j}\left(u+\eta\,\bm{\epsilon}\cdot\bm{n}_{j}\right)
=nQn(u)|Φ~n(s)(ϵ)+nP(u)Qn(u)|Φ¯n(s)(ϵ).\displaystyle=\sum_{n}Q^{n}(u)\ket{\tilde{\Phi}^{(s)}_{n}(\bm{\epsilon})}+\sum_{n}P(u)Q^{n}(u)\ket{\bar{\Phi}^{(s)}_{n}(\bm{\epsilon})}. (E47)

Since uu is an arbitrary parameter, it can be straightforwardly proven that these uu-independent states |Φ~n(s)(ϵ)\ket{\tilde{\Phi}^{(s)}_{n}(\bm{\epsilon})} and |Φ¯n(s)(ϵ)\ket{\bar{\Phi}^{(s)}_{n}(\bm{\epsilon})} are indeed eigenstates of the Hamiltonian. However, an elegant universal expression for |Φ~n(s)(ϵ)\ket{\tilde{\Phi}^{(s)}_{n}(\bm{\epsilon})} and |Φ¯n(s)(ϵ)\ket{\bar{\Phi}^{(s)}_{n}(\bm{\epsilon})} is lacking currently.

Let us consider the simplest case: the one-dimensional spin-12\frac{1}{2} model. The SHS is

n=1Nψn(12)(u+nη).\displaystyle\bigotimes_{n=1}^{N}\psi^{(\frac{1}{2})}_{n}\left(u+n\eta\right). (E48)

After some calculations, we can derive the expression of |Φ~0,1(s)(+1),|Φ¯0,1(s)(+1)\ket{\tilde{\Phi}^{(s)}_{0,1}(+1)},\ket{\bar{\Phi}^{(s)}_{0,1}(+1)}, specifically as follows

|Φ~0(12)(+1)=\displaystyle\ket{\tilde{\Phi}^{(\frac{1}{2})}_{0}(+1)}= evenmn1<n2<nmj=1mQ(njη)σn1+σn2+σnm+|ξ¯,\displaystyle\sum_{{\rm even}\,m}\,\sum_{n_{1}<n_{2}\cdots<n_{m}}\,\prod_{j=1}^{m}Q(n_{j}\eta)\sigma_{n_{1}}^{+}\sigma_{n_{2}}^{+}\cdots\sigma_{n_{m}}^{+}\ket{\bar{\xi}}, (E49)
|Φ~1(12)(+1)=\displaystyle\ket{\tilde{\Phi}^{(\frac{1}{2})}_{1}(+1)}= evenmn1<n2<nmk{n1,,nm}P(nη)j=1mQ(njη)σk+σn1+σn2+σnm+|ξ¯,\displaystyle\sum_{{\rm even}\,m}\,\sum_{n_{1}<n_{2}\cdots<n_{m}}\,\sum_{k\notin\{n_{1},\ldots,n_{m}\}}P(n\eta)\prod_{j=1}^{m}Q(n_{j}\eta)\sigma_{k}^{+}\sigma_{n_{1}}^{+}\sigma_{n_{2}}^{+}\cdots\sigma_{n_{m}}^{+}\ket{\bar{\xi}}, (E50)
|Φ¯0(12)(+1)=\displaystyle\ket{\bar{\Phi}^{(\frac{1}{2})}_{0}(+1)}= oddmn1<n2<nmj=1mQ(njη)σn1+σn2+σnm+|ξ¯,\displaystyle\sum_{{\rm odd}\,m}\,\sum_{n_{1}<n_{2}\cdots<n_{m}}\,\prod_{j=1}^{m}Q(n_{j}\eta)\sigma_{n_{1}}^{+}\sigma_{n_{2}}^{+}\cdots\sigma_{n_{m}}^{+}\ket{\bar{\xi}}, (E51)
|Φ¯1(12)(+1)=\displaystyle\ket{\bar{\Phi}^{(\frac{1}{2})}_{1}(+1)}= oddmn1<n2<nmk{n1,,nm}P(kη)j=1mQ(njη)σk+σn1+σn2+σnm+|ξ¯,\displaystyle\sum_{{\rm odd}\,m}\,\sum_{n_{1}<n_{2}\cdots<n_{m}}\,\sum_{k\notin\{n_{1},\ldots,n_{m}\}}P(k\eta)\prod_{j=1}^{m}Q(n_{j}\eta)\sigma_{k}^{+}\sigma_{n_{1}}^{+}\sigma_{n_{2}}^{+}\cdots\sigma_{n_{m}}^{+}\ket{\bar{\xi}}, (E52)

where |ξ¯=|121|122|12N\ket{\bar{\xi}}=\ket{-\tfrac{1}{2}}_{1}\otimes\ket{-\tfrac{1}{2}}_{2}\cdots\otimes\ket{-\tfrac{1}{2}}_{N}.