October 30, 2024
Exact solutions for excited states in non-integrable quantum Hamiltonians have revealed novel dynamical phenomena that can occur in quantum many-body systems. This work proposes a method to analytically construct a specific set of volume-law-entangled exact excited eigenstates in a large class of spin Hamiltonians. In particular, we show that all spin chains that satisfy a simple set of conditions host exact volume-law eigenstates in the middle of their spectra. Examples of physically relevant spin chains of this type include the transverse-field Ising model, PXP model, spin-\(S\) \(XY\) model, and spin-\(S\) Kitaev chain. Although these eigenstates are highly atypical in their structure, they are thermal with respect to local observables. Our framework also unifies many recent constructions of volume-law entangled eigenstates in the literature. Finally, we show that a similar construction also generalizes to spin models on graphs in arbitrary dimensions.
Introduction. The emergence of thermodynamics from unitary dynamics in isolated quantum many-body systems [1], [2] provides a fascinating puzzle in contemporary physics. At the heart of this enigma lies the Eigenstate Thermalization Hypothesis (ETH) [3]–[7], which posits that all the eigenstates of a quantum many-body system exhibit thermal properties for macroscopic observables. Thus, a comprehensive understanding of thermalization necessitates venturing beyond the traditional focus on the ground- and low-lying-excited- state properties of quantum many-body systems, and developing a thorough understanding of the structure of states in the bulk of their energy spectrum. Although the special class of integrable systems [8]–[11] offers a pathway to obtain solutions for excited states owing to the existence of extensive conserved quantities in them, they paradoxically serve as prime examples wherein the ETH is violated.
Obtaining analytical expressions for excited eigenstates remains a formidable challenge for generic non-integrable systems. This daunting task has spurred a plethora of research to identify tractable non-integrable models where a description beyond the low-energy states can be given [12]–[25]. This quest was further fueled by the fact that many of these exact eigenstates violate the ETH leading to anomalous non-thermal dynamics similar to those seen in experiments realizing the so-called PXP model [26], [27]. Hence, these eigenstates were dubbed as quantum many-body scars (QMBS) [28]–[32]. Examples of exact QMBS include towers of quasiparticle states in the Affleck-Kennedy-Lieb-Tasaki spin chain [33]–[36], \(\pi{-}\)bimagnon states in the spin-\(1\) \(XY\) model [20], [37], \(\eta{-}\)pairing states in the Hubbard model [38], [39], isolated area-law states in a variety of models [16] including the PXP [19] and others [40]–[50]. Multiple constructions that produce models hosting exact QMBS eigenstates have also been proposed [16], [38], [39], [51]–[57].
However, most of these QMBS states exhibit sub-volume entanglement entropy in contrast to the volume-law predicted by ETH. For some time, the main exception was a class of QMBS called rainbow scars [58], which do exhibit volume entanglement for some bipartitions, but the expectation value of appropriate local observables in it are still non-thermal. However, recent results have showcased exact volume-law entangled states hosted in various PXP-type models [59] and several spin-\(1/2\) chains with nearest-neighbor interactions [60], [61], which are “thermal" for local observables. These examples are few and far between and have been demonstrated in specific spin models, and the precise conditions for the existence of such solvable states are still unclear.
In this work, we provide a generic approach and explicit sufficient conditions for identifying a class of exact excited eigenstates in a large class of spin models. We show that any translation-invariant Hamiltonian whose every term formally anti-commutes with an on-site invertible operator times complex conjugation possesses these eigenstates. We start by constructing an exact zero energy eigenstate for spin-\(S\) chains with periodic boundary conditions (PBC) and then generalize it to higher dimensions. We further establish that these states show a volume-law scaling of entanglement entropy as predicted by ETH for all contiguous bipartitions. Moreover, these states exhibit thermal properties for any local observable, although they are atypical and athermal for non-local (but few-body) observables. Our construction recovers previously proposed volume-entangled states in various spin-\(1/2\) models [59]–[61]. Additionally, we construct analytical expressions of many new exact zero-energy states for a diverse set of spin-\(S\) Hamiltonians in arbitrary dimensions for both non-integrable and integrable models.
Figure 1: Schematic illustration of the exact zero-energy eigenstate \(|{\Lambda}{\rangle}_{\mathcal{C}}\) [defined in Eq. ?? ] for a spin-\(S\) system of \(2L\) sites: \((a)\) on a PBC chain, and \((b)\) in arbitrary dimension. The colored dotted lines indicate entangled “dimers" between spins at sites \(i\) and \(\overline{i}\). The entanglement in these dimers varies from \(0\) to \(\text{log}(2S{+}1)\), depending on the on-site invertible operator \(\mathcal{C}_{i}\). For example, in the case of spin-\(1/2\) and unitary \(\mathcal{C}_{i}\), the dotted lines represent pairs of spins in one of the Bell states {\(|\Psi^{\pm}\rangle\),\(|\Phi^{\pm}\rangle\)} [defined in Eq. 4 ]. Any arbitrary contiguous subsystem (shaded region) of size \(N{\leq}L\) cuts exactly \(N\) correlated bonds leading to an entanglement entropy \(\mathcal{S}_{N}{\propto}N\).
General construction in spin chains. We are interested in the class of spin-\(S\) chains of size \(L\) with a translation invariant interaction (with PBC) of the form \[H(L){=}\sum_{\alpha,i{=}1}^{L} \hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}{=}\sum_{\alpha,i{=}1}^{L}\prod_{k{=}1}^{n} (S_{i{+}j_{k}}^{\mu})^{q}, \label{eq:32generic95Hamiltonian}\tag{1}\] where \(S_{j}^{\mu}(\mu{=}x,y,z)\) is the spin-\(S\) operator at site \(j\) and \(q\) is a non-negative integer. The term \(\hat{H}^{\alpha}_{i+j_{1},{\cdots},i{+}j_{n}}\) represents interaction between any \(n{-}\)spins at sites \(\{i{+}j_{1},{\cdots},i{+}j_{n}\}\) with \(0{\leq}j_{k}{<}L\). The Hamiltonian is defined in the Hilbert space \(\mathcal{H}_{L}\) of dimension \(\mathcal{D}_{L}{=}(2S{+}1)^{L}\) on the global computational basis \(|\vec{S}\rangle{=}{\bigotimes}_{i{=}1}^{L}|s_{i}\rangle\) \((\mathcal{H}_{L}{\equiv}\text{span}\{|\vec{S}\rangle\})\), where \(|s_{i}\rangle\) denotes the \((2S{+}1)\) dimensional local Hilbert space on site \(i\). In this setting, we establish the following theorem.
Theorem 1. If each \(\hat{H}^{\alpha}_{i{+}j_{1},{\cdots}, i{+}j_{n}}\) in the Hamiltonian \(H(L)\) anti-commutes with an operator \(\mathcal{C}\mathcal{K}\), where \(\mathcal{C}{=}\prod_{i{=}1}^{L}\mathcal{C}_{i}\) with \(\mathcal{C}_{i}\) on-site invertible operators, and \(\mathcal{K}\) is the complex conjugation operator, i.e., \[\{\hat{H}^{\alpha}_{i+j_{1},\cdots,i+j_{n}},\mathcal{C}\mathcal{K}\}{=}0\text{for} i{\in}\{1,{\cdots}, L\}, \label{eq:32anti95commuation}\qquad{(1)}\] then the state [illustrated schematically in Fig. 1\((a)\)] \[\label{eq:32zero95energystate951d} \begin{align} |\Lambda\rangle_{\mathcal{C}}{=}{\sum\limits_{|\vec{S}\rangle\in\mathcal{H}_{L}}}\mathcal{C}|\vec{S}\rangle_{1,{\cdots},L} {\otimes} |\vec{S}\rangle_{\overline{1},{\cdots},\overline{L}} {=}{\bigotimes_{i{=}1}^{L}}{\sum_{s{=}1}^{2S{+}1}}\mathcal{C}_{i}|s_{i}{\rangle}|s_{\overline{i}}{\rangle}, \end{align}\qquad{(2)}\] where \(\overline{i}{\equiv}i{+}L\), is an exact zero-energy eigenstate of the Hamiltonian \(H(2L)\) defined in Eq. 1 on a system of size \(2L\) with PBC.
The anti-commutation relation in Eq. ?? implies \[\mathcal{C}(\hat{H}^{\alpha}_{i{+}j_{1}, \cdots, i{+}j_{n}})^{\ast}{=}-\hat{H}^{\alpha}_{i{+}j_{1}, \cdots, i{+}j_{n}}\mathcal{C}, \label{eq:32condition}\tag{2}\] where the complex conjugation is defined with respect to the computational basis \(|\vec{S}\rangle\). The proof of the theorem proceeds by using the mapping \(\overline{i}{\equiv}i{+}L\), and expressing the Hamiltonian for the \(2L\)-site system with PBC as \(H(2L){=}\sum_{\alpha,i{=}1}^{L} (\hat{H}^{\alpha}_{i{+}j_{1},{\cdots},i{+}j_{n}}{+}\hat{H}^{\alpha}_{\overline{i{+}j_{1}},{\cdots},\overline{i{+}j_{n}}})\), and then showing that for the state \(|\Lambda\rangle_{\mathcal{C}}\), the action of the term \(\hat{H}^{\alpha}_{i{+}j_{1}, {\cdots}, i{+}j_{n}}\) is canceled by that of \(\hat{H}^{\alpha}_{\overline{i{+}j_{1}},{\cdots},\overline{i{+}j_{n}}}\) for all \(i\) [see 1 of supplemental material (SM) [62] for details], i.e., \[\hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}|\Lambda\rangle_{\mathcal{C}}{=}{-}\hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}}|\Lambda\rangle_{\mathcal{C}}. \label{eq:32cancellation}\tag{3}\] By noting that Eq. 2 always holds for \(\mathcal{C} {=} \mathbb{I}\), we obtain the following corollary to Theorem 1.
Corollary 1. The state \[|\Lambda\rangle_{\mathbb{I}}{=} \sum_{|\vec{S}\rangle} |\vec{S}\rangle {\otimes} |\vec{S}\rangle{=}\bigotimes_{i{=}1}^{L}\sum_{s_{i}{=}1}^{2S{+}1}|s_{i}{\rangle}|s_{\overline{i}}{\rangle} \label{eq:32zero95energystate95for95imaginary95Hamiltonian}\qquad{(3)}\] is always a zero-energy eigenstate of any Hamiltonian of the form Eq. 1 defined on an even number of lattice sites (\(2L\)) where each \(\hat{H}^{\alpha}_{i{+}j_{1}, {\cdots}, i{+}j_{n}}\) has a purely imaginary matrix representation in the computational basis.
Next, we show that the state \(|\Lambda\rangle_{\mathcal{C}}\) is volume-law entangled for an arbitrary contiguous bipartition. By definition, the state \(\ket{\Lambda}_{\mC}\) is a product of “dimers" between sites \(i\) and \(\overline{i}\). The entanglement between the spins on the sites \(i\) and \(\overline{i}\) is therefore strictly greater than \(0\). It can be as high as \(\text{log}(2S{+}1)\) depending upon the nature of \(\mathcal{C}_{i}\) (the reduced density matrix of each dimer is just \(\mC_i \mC_i^\dagger\), whose eigenvalues are just the singular values of \(\mC_i\)), and is maximally entangled if and only if \(\mathcal{C}_{i}\) is unitary.
An arbitrary contiguous subsystem of size \(N{\leq}L\) [shaded grey in Fig. 1\((a)\)] cuts exactly \(N\) bonds between the correlated spins leading to the entanglement entropy of the subsystem \(\mathcal{S}_{N}{\propto} N\) thereby obeying volume-law scaling. Moreover, when \(\mC_i\) is unitary, the reduced density matrix over any contiguous subsystem is just proportional to the identity matrix, which is just the infinite-temperature Gibbs density matrix, hence the state is thermal for strictly local observables [59], [61]. However, for tailored subsystems chosen to consist solely of sites \(\{i,\overline{i}\}\), the entanglement entropy is strictly zero, and non-local observables with support over these sub-systems exhibit highly athermal properties.
We now illustrate the power of our general construction by illustrating exact excited zero-energy states for several integrable and non-integrable spin chains. We begin by considering spin-\(1/2\) systems for which any Hamiltonian can be recast to the form given in Eq. 1 . The spin operators at site \(i\) can be represented by Pauli matrices \(\sigma_{i}^{\mu} ({\mu}{=}x,y,z)\) and the local Hilbert space is defined by the \(\sigma^{z}_{i}\)-eigenvectors \(|0\rangle_{i}\) and \(|1\rangle_{i}\). In what follows, the state \(|\Lambda\rangle_{\mathcal{C}}\) can be expressed in terms of Bell states between sites \(i,j\) defined as \[|\Psi^{\pm}\rangle_{i, j}{=}\frac{(|00\rangle {\pm} |11\rangle)_{i,j}}{\sqrt{2}};~ |\Phi^{\pm}\rangle_{i,j} {=} \frac{(|01\rangle {\pm} |10\rangle)_{i,j}}{\sqrt{2}}. \label{eq:32bell32states}\tag{4}\] For instance, consider the following non-integrable Hamiltonian discussed in Ref. [61] \[\begin{align} H_{1}(L){=}\sum_{i{=}1}^{L}\big(J_{1}\sigma_{i}^{x}\sigma_{i{+}1}^{y}+J_{2}\sigma_{i}^{y}\sigma_{i+1}^{z}\big), \end{align} \label{eq:32spin-147232example2}\tag{5}\] with arbitrary values of the coupling constants \(J_{1}\) and \(J_{2}\). Here \(n{=}2,j_{k}{=}k{-}1\) and each \(\hat{H}^{\alpha}_{i, i{+}1}\) (i.e., \(\sigma_{i}^{x}\sigma_{i{+}1}^{y}\) and \(\sigma_{i}^{y}\sigma_{i+1}^{z}\)) is purely imaginary in the computational basis. From Corollary 1, the state \(|\Lambda\rangle_{\mathbb{I}}\) given in Eq. ?? is a zero-energy eigenstate of \(H_{1}(2L)\) and takes a simple form \(|\Lambda(S{=}1/2)\rangle_{\mathbb{I}}{=}\bigotimes_{i{=}1}^{L}|\Psi^{+}\rangle_{i,\overline{i}}\). This state is depicted in Fig. 1\((a)\) where all the dotted lines represent the Bell state \(|\Psi^{+}\rangle\). Since the Hamiltonian \(H_{1}(2L)\) has global spectral-reflection symmetry generated by the operator \(\mathcal{M}_{1}{=}{\prod}_{i{=}1}^{2L}\sigma_{i}^{y}\), the zero-energy state mentioned above is a mid-spectrum excited state. Note that \(|\Lambda(S{=}1/2)\rangle_{\mathbb{I}}\) is a zero-energy eigenstate of any spin-\(1/2\) Hamiltonian with even number of sites of form Eq. 1 that has an odd number of \(\sigma_{y}\)’s in each \(\hat{H}^{\alpha}_{i{+}j_{1}, {\cdots}, i{+}j_{n}}\).
Furthermore, when the half-chain length \(L\) is a multiple of three, the operator \(\mathcal{C}_{1}{=}{\prod}_{p{=}1}^{L/3}\sigma_{3p{-}2}^{x} \sigma_{3p{-}1}^{y}\sigma_{3p}^{z}\) satisfies the conditions given in Eq. 2 . This allows us to identify another zero-energy eigenstate (and its translations) of \(H_{1}(2L)\) of the form \(|\Lambda\rangle_{\mathcal{C}_{1}}{=}\bigotimes_{i{=}1}^{L{-}2}|\Phi^{+}\rangle_{i,\overline{i}}|\Phi^{-}\rangle_{i{+}1,\overline{i{+}1}}|\Psi^{-}\rangle_{i{+}2,\overline{i{+}2}}\). The state \(|\Lambda\rangle_{\mathcal{C}_{1}}\) is precisely the exact zero-energy eigenstate dubbed as Entangled Antipodal Pair (EAP) state presented in Ref. [61]. While Ref. [61] derives the Hamiltonian \(H_{1}(2L)\) via an exhaustive search of all possible nearest-neighbor spin-\(1/2\) interactions for which the predefined EAP state has zero-energy, we show that \(H_1(2L)\) satisfies a very general sufficient condition for the existence of such an EAP state. Similarly, other zero energy eigenstates found in different spin-\(1/2\) Hamiltonians of Ref [60], [61] including the PXP chain [59] can be understood using our construction (see 3 of SM [62] for details with other examples).
Next, we investigate a spin-\(S\) \(XY\) chain, described by the Hamiltonian \[H_{XY}(L){=} \sum_{i=1}^{L} (J_{1}S_{i}^{x}S_{i+1}^{x}+J_{2}S_{i}^{y}S_{i+1}^{y}+hS_{i}^{z}), \label{eq:32spin-S32XY}\tag{6}\] where \(J_{1},J_{2}\) and \(h\) are arbitrary and \(S_{i}^{\mu}\)(\(\mu{=}x,y,z\)) are spin\({-}S\) operators at site \(i\). The spin components define site parity operators \(\mathcal{P}_{i}^{\mu}{=}e^{i\pi S_{i}^{\mu}}\). One can show that \(\mathcal{P}_{i}^{\mu}\) commutes with \(S_{i}^{\mu}\) but anti-commutes with \(S_{i}^{\mu'} (\mu{\neq}\mu')\). Hence when \(L\) is even, in the orthonormal basis \(|m\rangle_{i}\) that are eigenstates of \(S_{i}^{z}\) with eigenvalues \(m{=}{-}S, {-}S{+}1, {\cdots}, S\), the operator defined by \(\mathcal{C}_{XY}{=}\prod_{p{=}1}^{L/2}\mathcal{P}_{2p{-}1}^{x}\mathcal{P}_{2p}^{y}\), satisfies Eq. 2 . As a result of Theorem 1, the Hamiltonian \(H_{XY}(2L)\) hosts an exact zero-energy eigenstate \(|\Lambda\rangle_{\mathcal{C}_{XY}}\) (and its translations). Using the action of the on-site parity operators on the basis states \(\mathcal{P}_{i}^{x}|m\rangle_{i}{=}({-}1)^{S}|{-}m\rangle_{i}\) and \(\mathcal{P}_{i}^{y}|m\rangle_{i}{=}({-}1)^{S{+}m}|{-}m\rangle_{i}\), we can express the state as \[\begin{align} &|\Lambda\rangle_{\mathcal{C}_{XY}}=\bigotimes_{i{=}1}^{L{-}1}|\kappa_{S}^{-}\rangle_{i,\overline{i}}|\kappa_{S}^{+}\rangle_{i{+}1,\overline{i{+}1}},\text{where}\\ &|\kappa_{S}^{\pm}\rangle_{i,j}{=}\frac{1}{\sqrt{2S{+}1}}\sum_{p{=}0}^{2S}({\pm}1)^{p}(S^{-})^{p}|S\rangle_{i}{\otimes}(S^{+})^{p}|{-}S\rangle_{j}. \end{align} \label{eq:32Lambda95S95XY32state}\tag{7}\] For \(S{>}1/2\), \(H_{XY}\) is non-integrable [20] for generic values of the couplings, while for \(S{=}1/2\), \(H_{XY}\) is integrable (reduces to the transverse field Ising model for \(J_{2}{=}0\)) [63]. For \(S{=}1/2\), this state reads \(|\Lambda\rangle_{\mathcal{C}_{XY}}{=}\bigotimes_{i{=}1}^{L{-}1}|\Phi^{-}\rangle_{i,\overline{i}}|\Phi^{+}\rangle_{i{+}1,\overline{i{+}1}}\) and shows that volume-law eigenstates exist even in integrable models.
The sufficient conditions we derive allow us to construct a large family of Hamiltonians that will inherently host the specific volume-law entangled eigenstate of the particular form of Eq. ?? . For instance, the zero-energy state \(|\Lambda_{S}\rangle_{\mathcal{C}_{XY}}\) described in Eq. 7 is also a zero-energy eigenstate of the spin-\(S\) Kitaev chain [64]–[66] given by the Hamiltonian \(H_{K}(2L){=}\sum_{i{=}1}^{L}(J_{1}S_{2i{-}1}^xS_{2i}^x {+}J_{2} S_{2i}^y S_{2i{+}1}^y)\) for arbitrary values of \(J_{1}\) and \(J_{2}\). Furthermore, Hamiltonians involving longer-range (beyond nearest neighbors) interactions between spins, e.g. \(H_{\rm NNNN}(2L){=}\sum_{i{=}1}^{2L} (J_{1}S_{i}^{x}S_{i{+}3}^{x}{+}J_{2}S_{i}^{y}S_{i{+}3}^{y})\) or models with more than two-spin interactions, e.g. \(H_{3{-}4}(2L){=}\sum_{i{=}1}^{2L} (J_{1}S_{i}^{z}S_{i{+}1}^{z}S_{i{+}2}^{z}{+}J_{2}S_{i}^{x}S_{i{+}1}^{z}S_{i{+}2}^{y}S_{i{+}3}^{z})\) host the same zero-energy state \(|\Lambda_{S}\rangle_{\mathcal{C}_{XY}}\) in their spectrum as they satisfy Eq. 2 . Many such Hamiltonians exist physically or can be engineered experimentally.
Generalization to arbitrary dimensions. To investigate higher-dimensional spin systems, it is convenient to revisit the one-dimensional case and rearrange the \(2L\) sites labeled by indices \({1}, \overline{1}, {\cdots},{L}, \overline{L}\) into the configuration shown in Fig. 1\((b)\). Our goal is to determine local Hamiltonians that host the state \(|\Lambda\rangle_\mathcal{C}\) defined in Eq. ?? as a zero energy eigenstate provided the operator \(\mathcal{C}\) satisfies conditions analogous to those in Eq. 2 .
Similar to the one-dimensional case, for all the interaction terms \(\hat{H}_{j_{1},{\cdots}, j_{n}}^{\alpha}\) between any \(n{-}\)sites \(j_{1},{\cdots}, j_{n}\) (where \(j_k{\in}\{1,\overline{1},{\cdots}, L,\overline{L}\}\) and \(j_{k}{\neq} \overline{j}_{k'}\) for all \(k,k'\)), restricted to the form \(\hat{H}_{j_{1},{\cdots}, j_{n}}^{\alpha}{=}\prod_{k{=}1}^{n}(S_{j_{k}}^{\mu})^{q}\) and satisfying \(\{\hat{H}^{\alpha}_{j_{1},{\cdots},j_{n}},\mathcal{C}\mathcal{K}\}{=}0\), we find that \[(\hat{H}^{\alpha}_{j_{1},\cdots,j_{n}} + \hat{H}^{\alpha}_{\overline{j}_{1},\cdots,\overline{j}_{n}})|\Lambda\rangle_{\mathcal{C}}=0, \label{eq:32cancelation95in95arb95dimension}\tag{8}\] where we have considered the identification \(\overline{\overline{j}}_{k}{=}j_{k}\). Consequently, all Hamiltonians of the form \(H{=}\sum_{\alpha,\{j\}} J_{\alpha}(\hat{H}^{\alpha}_{j_{1},{\cdots},j_{n}} {+} \hat{H}^{\alpha}_{\overline{j}_{1},{\cdots},\overline{j}_{n}})\) with arbitrary coefficients \(J_{\alpha}\), host \(|\Lambda\rangle_{\mathcal{C}}\) as a zero energy eigenstate.
For instance, consider the Hamiltonian of the nearest-neighbor spin-\(S\) \(XY\) model with a transverse field defined on a \(2L{\times}2L\) square lattice with PBC in both directions, \[\begin{align} H_{XY}^{2D}{=}\sum_{\mu,i,j}J_{\mu}(S_{i,j}^{\mu}S_{i{+}1,j}^{\mu}{+}S_{i,j}^{\mu}S_{i,j{+}1}^{\mu}){+}h\sum_{i,j}S_{i,j}^{z}, \end{align} \label{eq:322D32XY32model32Hamiltonian}\tag{9}\] where \(S_{i,j}^{\mu}(\mu{=}x,y,z)\) are the spin operators at site \((i,j)\) with \(i,j{\in}\{1, {\cdots}, 2L\}\) with \(i\) and \(j\) labeling the horizontal and vertical directions, respectively, with \((1,1)\) at the bottom-left corner, and \(J_{\mu}\) and \(h\) are arbitrary couplings.
Figure 2: Schematic representation of the zero energy eigenstate \(|\Lambda\rangle_\mathcal{C}\) for spin-\(S\) \(XY\) model with transverse field in different geometries. The black dots denote the sites and the solid lines between them indicate the \(XY\) interaction. The spins at red (blue) sites \(r\) are entangled only with their partners \(\overline{r}\) in the state \(|\kappa_{S}^{-}\rangle\) (\(|\kappa_{S}^{+}\rangle\)) [defined in Eq. 7 ]. \((a),(b)\) Distinct zero energy states on a \(2L{\times}2L\) square lattice with PBC, with \(L{=}2\). \((c)\) Eigenstate on a \(2L{\times}M\) cylinder, with \((L,M){=}(3,2)\). \((d)\) Eigenstate on a \(2L{\times}2L{\times}M\) cuboid, with either PBC or open boundaries, with \((L,M){=}(2,2).\)
Now relabeling the \(4L^{2}\) sites with a single index \(r_{i,j}{\in}\{1, \overline{1}, {\cdots}, 2L^{2}, \overline{2L^{2}}\}\) as shown in Fig. 2\((a)\), where \[r_{i,j}{\equiv} \begin{cases} i{+}2L{\times}(j{-}1)&\text{for} j{\leq}L \\ \overline{1{+}[(i{-}1{+}L){\rm~mod~}2L]{+}2L{\times}(j{-}1{-}L)}&\text{for} j{>}L, \end{cases} \label{eq:322D32XY-relabelling}\tag{10}\] the Hamiltonian \(H_{XY}^{2D}\) reduces to the convenient form \[\begin{align} H&_{XY}^{2D}{=}\sum_{\mu,i,j}^{j{<}L}\Big[J_{\mu}(S_{r_{i,j}}^{\mu}S_{r_{i{+}1,j}}^{\mu}{+}S_{\overline{r_{i,j}}}^{\mu}S_{\overline{r_{i{+}1,j}}}^{\mu})\\ &{+}J_{\mu}(S_{r_{i,j}}^{\mu}S_{r_{i,j{+}1}}^{\mu}{+}S_{\overline{r_{i,j}}}^{\mu}S_{\overline{r_{i,j{+}1}}}^{\mu}){+}h(S_{r_{i,j}}^{z}{+}S_{\overline{r_{i,j}}}^{z})\Big]. \end{align} \label{eq:322D32XY32model32Hamiltonian95relabelled}\tag{11}\] In the computational basis \(|m\rangle_{r_{i,j}}\) that are eigenstates of \(S_{r_{i,j}}^{z}\), each \(\hat{H}_{\{r\}}^{\alpha}\) of the Hamiltonian, i.e., \(S_{r_{i,j}}^{x}S_{r_{i{+}1,j}}^{x},S_{r_{i,j}}^{y}S_{r_{i{+}1,j}}^{y}, S_{r_{i,j}}^{x}S_{r_{i,j{+}1}}^{x}\), \(S_{r_{i,j}}^{y}S_{r_{i,j{+}1}}^{y}\) and \(S_{r_{i,j}}^{z}\) are real and anticommute with the operator \(\mathcal{C}_{XY}^{2D}{=}\prod_{i,j}(\mathcal{P}_{r_{i,j}}^{x})^{\delta_{(i{+}j){\rm~mod~}2,0}}(\mathcal{P}_{r_{i,j}}^{y})^{\delta_{(i{+}j){\rm~mod~}2,1}}\) [thus satisfy Eq. 2 ]. Thus, using Theorem 1 and Eq. 8 , the Hamiltonian \(H_{XY}^{2D}\) hosts a zero energy eigenstate of the form \(|\Lambda\rangle_{C_{XY}}^{2D}\) which can be expressed as \[|\Lambda\rangle_{C_{XY}}^{2D}{=}\bigotimes_{i,j}^{j{<}L}(|\kappa_{S}^{-}\rangle_{r,\overline{r}})^{\delta_{(i{+}j){\rm~mod~}2,0}} (|\kappa_{S}^{+}\rangle_{r,\overline{r}})^{\delta_{(i{+}j){\rm~mod~}2,1}}, \label{eq:32zeroenergy95eigstate952DXY}\tag{12}\] where \(\ket{\kappa_S^\pm}\) are defined in Eq. 7 . The state \(|\Lambda\rangle_{C_{XY}}^{2D}\) is shown in Fig. 2\((a)\), where the spins at sites \(r_{i,j}\) and \(\overline{r_{i,j}}\) are in the state \(|\kappa_{S}^{-}\rangle\) when \(i{+}j\) is even (marked in red) and in the state \(|\kappa_{S}^{+}\rangle\) when \(i{+}j\) is odd (marked in blue). Similarly, with different choices of labeling of the sites [e.g., see Fig. 2\((b)\) and more in the 2 of SM [62]] we can find many more zero-energy eigenstates \(|\Lambda\rangle_{\mathcal{C}}\) for \(H_{XY}^{2D}\) with different pairing patterns. Using the same strategy, we can also find zero-energy eigenstates of spin-\(S\) \(XY\) model in different geometries as shown in Figs. 2\((c)\),\((d)\) (and more in the SM [62]). Many such zero-energy eigenstates can be found for other spin Hamiltonians in arbitrary dimensions using our construction. Unlike the one-dimensional case, in higher dimensions, there can be special contiguous bipartitions for which the entanglement is not a volume-law, although it is for most of them.
Discussion. In this work, we have introduced a general method for constructing a specific class of exact zero-energy excited eigenstates in spin-\(S\) models that satisfy certain conditions. This construction only uses the existence of an operator, which, along with complex conjugation, anti-commutes with each term of the Hamiltonian. The existence of such an operator is easy to guess and check, and this directly allows us to construct exact eigenstates without analyzing the complex microscopic details of the model. We demonstrated the broad applicability of our construction by finding new exact eigenstates for various natural examples of spin-\(S\) Hamiltonians. We also recovered some previously known results for spin-\(1/2\) models [59], [61] as special cases of our construction.
In one dimension, the states we identified exhibit volume-law scaling of entanglement entropy for general contiguous bipartitions. In cases when the operator \(\mC\) is unitary, they also exhibit infinite-temperature thermal properties for all strictly local observables. This phenomenology mostly also extends to higher dimensions, although a few contiguous bipartitions might not display a volume-law scaling of entanglement. However, in any dimension, these states are always athermal for non-local but few-body observables and hence can be considered examples of highly entangled QMBS. Indeed, by construction, they are simultaneous eigenstates of non-commuting local operators, consistent with the proposed definition of QMBS based on commutant algebras [56]. It would be interesting to explore whether other examples of QMBS that exhibit superficially similar structures, e.g., the exponentially many exact eigenstates in the 2D PXP model [22], fit into this construction.
Building on this foundation, it is natural to wonder if extensions of these ideas to models beyond spins such as fermionic and bosonic systems would work, or generalizations such as going beyond pairwise entanglement to entangled clusters of \(l\)-spins, with \(l{\geq}3\) can be engineered. Along with the theoretical interest in states close to completely satisfying ETH, the existence of such states can also aid in the experimental measurement of Out-of-Time-Ordered Correlators (OTOC), as discussed in [59].
Finally, on a different note, while we focused on constructing non-integrable models with these states, there are also integrable models, such as the one-dimensional transverse-field Ising model, which host such states. While these states are athermal and violate the ETH [3] in non-integrable models (at least with respect to some non-local but few-body observables), it is worth exploring the implications of their existence in integrable models to the generalized ETH [10], [11], [67], the natural analog of ETH for integrable models.
Acknowledgments. We thank Andrew Ivanov, Maurizio Fagotti, and Diptiman Sen for useful discussions. This work was undertaken on the Nandadevi supercomputer maintained and supported by the Institute of Mathematical Sciences’ High-Performance Computing Center, India. This research was supported by the Munich Center for Quantum Science and Technology (MCQST), and in part by the International Centre for Theoretical Sciences (ICTS) for participating in the program - Stability of Quantum Matter in and out of Equilibrium at Various Scales (code: ICTS/SQMVS2024/01). S. Moudgalya acknowledges the hospitality of the Institute of Mathematical Sciences (IMSc), Chennai, where this collaboration was initiated. S. Mohapatra is grateful to the School of Natural Sciences, Technical University of Munich (TUM) for its hospitality in the summer of 2024 where a part of this work was completed.
Supplemental Material for “Exact volume-law entangled eigenstates in a large class of spin models"
Sashikanta Mohapatra\(^{1,2}\), Sanjay Moudgalya\(^{3,4}\) and Ajit C. Balram\(^{1,2}\)
\(^{1}\)Institute of Mathematical Sciences, CIT Campus, Chennai 600113, India
\(^{2}\)Homi Bhabha National Institute, Training School Complex, Anushaktinagar, Mumbai 400094, India
\(^{3}\)School of Natural Sciences, Technische Universität München (TUM), James-Franck-Str. 1, 85748 Garching, Germany
\(^{4}\)Munich Center for Quantum Science and Technology (MCQST), Schellingstr. 4, 80799 München, Germany
This supplemental material contains some technical details and additional examples. In Sec. 1, we prove Theorem 1 of the main text. In Sec 2, we present additional examples of zero energy eigenstates of nearest neighbor spin-\(S\) \(XY\) model in arbitrary dimensions that we considered in the main text. Finally, in Sec 3, we reproduce some zero energy eigenstates of previously studied Hamiltonians [59]–[61] using our framework.
For convenience, let us restate the form of the Hamiltonian and Theorem 1 of the main text here. We are considering the class of spin-\(S\) chains of size \(L\) with a translation invariant interaction [with periodic boundary condition (PBC)] of the form \[H(L){=}\sum_{\alpha,i{=}1}^{L} \hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}{=}\sum_{\alpha,i{=}1}^{L}\prod_{k{=}1}^{n} (S_{i{+}j_{k}}^{\mu})^{q}, \label{suppeq:32generic95Hamiltonian}\tag{13}\] where \(S_{j}^{\mu}(\mu{=}x,y,z)\) is the spin-\(S\) operator at site \(j\) and \(q\) is a non-negative integer. The term \(\hat{H}^{\alpha}_{i+j_{1},\cdots,i+j_{n}}\) represents interaction between any \(n{-}\)spins at sites \(\{i{+}j_{1},{\cdots},i{+}j_{n}\}\) with \(0{\leq}j_{k}{<}L\). The Hamiltonian is defined in the Hilbert space \(\mathcal{H}_{L}\) of dimension \(\mathcal{D}_{L}{=}(2S{+}1)^{L}\) on the global computational basis \(|\vec{S}\rangle{=}{\bigotimes}_{i{=}1}^{L}|s_{i}\rangle\) \((\mathcal{H}_{L}{\equiv}\text{span}\{|\vec{S}\rangle\})\), where \(|s_{i}\rangle\) denotes the \((2S{+}1)\) dimensional local Hilbert space on site \(i\). For this setting, we stated the following theorem in the main text:
Theorem 1. If each \(\hat{H}^{\alpha}_{i{+}j_{1},{\cdots}, i{+}j_{n}}\) in the Hamiltonian \(H(L)\) anti-commutes with an operator \(\mathcal{C}\mathcal{K}\), where \(\mathcal{C}{=}\prod_{i{=}1}^{L}\mathcal{C}_{i}\) with \(\mathcal{C}_{i}\) on-site invertible operators, and \(\mathcal{K}\) is the complex conjugation operator, i.e., \[\{\hat{H}^{\alpha}_{i+j_{1},\cdots,i+j_{n}},\mathcal{C}\mathcal{K}\}{=}0\text{for} i{\in}\{1,{\cdots}, L\}, \label{suppeq:32anti95commuation}\qquad{(4)}\] then the state \[\label{suppeq:32zero95energystate951d} \begin{align} |\Lambda\rangle_{\mathcal{C}}{=}\sum_{|\vec{S}\rangle\in\mathcal{H}_{L}} \mathcal{C}|\vec{S}\rangle_{1,\cdots,L} \otimes |\vec{S}\rangle_{\overline{1},\cdots,\overline{L}} {=}\bigotimes_{i{=}1}^{L}\sum_{s{=}1}^{2S{+}1}\mathcal{C}_{i}|s_{i}{\rangle}|s_{\overline{i}}{\rangle}, \end{align}\qquad{(5)}\] where \(\overline{i}{\equiv} i{+}L\), is an exact zero energy eigenstate of the Hamiltonian \(H(2L)\) defined in Eq. 13 on a system of size \(2L\) with PBC.
Proof. Throughout this proof, we use the identification \(\overline{i}{\equiv}i{+}L\). We can then express the Hamiltonian for the \(2L\)-site system with PBC as \[H(2L){=}\sum_{\alpha,i{=}1}^{L} (\hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}{+}\hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}})\] and then the proof proceeds by showing that the action of the term \(\hat{H}^{\alpha}_{i{+}j_{1}, \cdots, i{+}j_{n}}\) on the state \(|\Lambda\rangle_{\mathcal{C}}\) is exactly canceled by that of \(\hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}}\) on the same state for all \(i\).
This is straightforward to show when each \(\hat{H}^{\alpha}_{i{+}j_{1}, {\cdots}, i{+}j_{n}}\) acts entirely on either the first \(L\) sites (\(1 {\leq} i{+}j_{1}{\leq}L\) and \(1 {\leq} i{+}j_{n}{\leq}L\)), or the last \(L\) sites (\(\overline{1}{\leq}i{+}j_{1}{\leq}\overline{L}\) and \(\overline{1}{\leq}i{+}j_{n}{\leq}\overline{L}\)). By the anti-commutation relation in Eq. ?? , we simply mean \[\mathcal{C}(\hat{H}^{\alpha}_{i{+}j_{1}, \cdots, i{+}j_{n}})^{\ast}{=}-\hat{H}^{\alpha}_{i{+}j_{1}, \cdots, i{+}j_{n}}\mathcal{C}, \label{suppeq:32condition}\tag{14}\] where the complex conjugation is defined with respect to the computational basis \(|\vec{S}\rangle\). Hence, we obtain \[\begin{align} &\hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}|\Lambda\rangle_{\mathcal{C}}{=}{-}\sum_{|\vec{S}\rangle} \mathcal{C}(\hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}})^{\ast} |\vec{S}\rangle{\otimes} |\vec{S}\rangle \tag{15}\\ &\hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}}|\Lambda\rangle_{\mathcal{C}}{=}{\sum_{|\vec{S}\rangle}}\mathcal{C}|\vec{S}\rangle{\otimes} \hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}}|\vec{S}\rangle. \tag{16} \end{align}\] Next, in Eqs. 15 and 16 we insert the resolution of the identity operator \(\mathbb{I}_{1,{\cdots}, L}{=}(\sum_{\vec{S'}}|\vec{S'}\rangle\langle \vec{S'}|)_{1,\cdots, L}\) and \(\mathbb{I}_{\overline{1},{\cdots}, \overline{L}}{=}(\sum_{\vec{S'}}|\vec{S'}\rangle\langle \vec{S'}|)_{\overline{1},{\cdots}, \overline{L}}\) respectively to get
\[\begin{align} &\hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}|\Lambda\rangle_{\mathcal{C}}{=}{-}\sum_{|\vec{S}\rangle,|\vec{S'}\rangle} f^{\ast}(S',S)\; \mathcal{C}|\vec{S'}\rangle\otimes|\vec{S}\rangle, \tag{17}\\ &\hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}}|\Lambda\rangle_{\mathcal{C}}{=}{\sum_{|\vec{S}\rangle,|\vec{S'}\rangle}}f(S',S)\; \mathcal{C}|\vec{S}\rangle\otimes |\vec{S'}\rangle, \tag{18} \end{align}\] where the function \(f(S',S){=}\langle \vec{S'}| (\hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}) |\vec{S}\rangle\) with \(f^{\ast}(S',S){=}\langle \vec{S'}| (\hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}})^{\ast} |\vec{S}\rangle\). Using the hermiticity of \(\hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}\) in this basis, i.e., \(f^{\ast}(S',S){=}f(S,S')\), we interchange \(|\vec{S}\rangle\) and \(|\vec{S'}\rangle\) in Eq. 17 to get \[\hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}|\Lambda\rangle_{\mathcal{C}}{=}{-}\hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}}|\Lambda\rangle_{\mathcal{C}}\text{for} i{+}j_{n}{\leq}L.\] For terms that straddle the first \(L\) sites and last \(L\) sites, i.e., when \(i{+}j_{1}{\leq}L\) and \(i{+}j_{n}{>}L\) (or \(i{+}j_{1}{\leq}\overline{L}\) and \(i{+}j_{n}{<}L\)), given the form of the Hamiltonian in Eq. 13 , we can decompose the terms \(\hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}\) and \(\hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}}\) as \[\hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}{=}\mathcal{A}^{\alpha}_{i{+}j_{1}, \cdots, L}{\otimes}\mathcal{B}^{\alpha}_{\overline{1}, \cdots, \overline{i{+}j_{n}{-}L}}~~\text{and}~~\hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}}{=}\mathcal{B}^{\alpha}_{1, \cdots, i{+}j_{n}{-}L}{\otimes}\mathcal{A}^{\alpha}_{\overline{i{+}j_{1}}, \cdots, \overline{L}}, \label{suppeq:32partition32of32H32into32A32and32B}\tag{19}\] where \(\mathcal{A}^{\alpha}_{i{+}j_{1}, \cdots, L}{=}\prod_{k{=}1}^{i{+}j_{k}{=}L} (S_{i{+}j_{k}}^{\mu})^{q}\) and \(\mathcal{B}^{\alpha}_{1, \cdots, i{+}j_{n}{-}L}{=}\prod_{i{+}j_{k}{=}1}^{i{+}j_{n}{-}L} (S_{i{+}j_{k}}^{\mu})^{q}\) acts on the first \(L\) sites while \(\mathcal{A}^{\alpha}_{\overline{i{+}j_{1}}, \cdots, \overline{L}}{=}\) \(\prod_{k{=}1}^{i{+}j_{k}{=}L} (S_{\overline{i{+}j_{k}}}^{\mu})^{q}\) and \(\mathcal{B}^{\alpha}_{\overline{1}, \cdots, \overline{i{+}j_{n}{-}L}}{=}\prod_{i{+}j_{k}{=}1}^{i{+}j_{n}{-}L} (S_{\overline{i{+}j_{k}}}^{\mu})^{q}\) acts on the last \(L\) sites. Since both \(\hat{H}^{\alpha}_{i{+}j_{1},\cdots, i{+}j_{n}}\) and the operator \(\mC\) are just products of on-site operators, the condition stated in Eq. 14 implies that for each of the spins we should have \[\mC_{i{+}j_{k}}[(S_{i{+}j_{k}}^{\mu})^{q}]^{\ast}=K_{i{+}j_{k}}(S_{i{+}j_{k}}^{\mu})^{q}\mC_{i{+}j_{k}}~~\text{with}~~\prod_{k{=}1}^{n}K_{i{+}j_{k}}{=}{-}1, \label{suppeq:32on95site95condition}\tag{20}\] where \(K_{i{+}j_{k}}\) is a non-zero complex number. Thus for terms of the form given in Eq. 19 , we should have \[\mathcal{C}(\mathcal{A}^{\alpha}_{i{+}j_{1}, \cdots, L})^{\ast}{=}K_{\rm{left}}\mathcal{A}^{\alpha}_{i{+}j_{1}, {\cdots}, L}\mathcal{C}~~\text{and}~~ \mathcal{C}(\mathcal{B}^{\alpha}_{\overline{1}, \cdots, \overline{i{+}j_{n}{-}L}})^{\ast}{=}K_{\rm{right}}\mathcal{B}^{\alpha}_{\overline{1}, \cdots, \overline{i{+}j_{n}{-}L}}\mathcal{C}~~\text{with}~~K_{\rm{left}}{\times}K_{\rm{right}}{=}{-}1, \label{suppeq:32splitting95condition}\tag{21}\] where \(K_{\rm{left}}{=}\prod_{l{=}i{+}j_{1}}^{L}K_{l}\) and \(K_{\rm{right}}{=}\prod_{l{=}L{+}1}^{i{+}j_{n}}K_{l}\). As we show in Lemma 1 below, the only non-zero value of \(K_{l}\) can be \({\pm}1\). From Eq. 21 we then get either \(K_{\rm{left}}{=}1\) and \(K_{\rm{right}}{=}{-}1\) or \(K_{\rm{left}}{=}{-}1\) and \(K_{\rm{right}}{=}1\), i.e., \[\mathcal{C}(\mathcal{A}^{\alpha}_{i{+}j_{1}, \cdots, L})^{\ast}{=}{\pm}\mathcal{A}^{\alpha}_{i{+}j_{1}, {\cdots}, L}\mathcal{C},\;\; \mathcal{C}(\mathcal{B}^{\alpha}_{\overline{1}, \cdots, \overline{i{+}j_{n}{-}L}})^{\ast}{=}{\mp}\mathcal{B}^{\alpha}_{\overline{1}, \cdots, \overline{i{+}j_{n}{-}L}}\mathcal{C}. \label{suppeq:32spilting95condition9564}\tag{22}\] Using the decompositions shown in Eq. (21 ), and then using Eq. (22 ), we obtain \[\begin{align} &\hat{H}^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}|\Lambda\rangle_{\mathcal{C}}{=}\sum_{|\vec{S}\rangle}\mathcal{A}^{\alpha}_{i{+}j_{1}, \cdots, L}\mathcal{C}|\vec{S}\rangle {\otimes} \mathcal{B}^{\alpha}_{\overline{1}, \cdots, \overline{i{+}j_{n}{-}L}}|\vec{S}\rangle{=}{\pm}\sum_{|\vec{S}\rangle}\mathcal{C}(\mathcal{A}^{\alpha}_{i{+}j_{1}, \cdots, L})^{\ast}|\vec{S}\rangle {\otimes} \mathcal{B}^{\alpha}_{\overline{1}, \cdots, \overline{i{+}j_{n}{-}L}}|\vec{S}\rangle \tag{23}\\ &\hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}}|\Lambda\rangle_{\mathcal{C}}{=}\sum_{|\vec{S}\rangle}\mathcal{B}^{\alpha}_{1, \cdots, i{+}j_{n}{-}L}\mathcal{C}|\vec{S}\rangle{\otimes} \mathcal{A}^{\alpha}_{\overline{i{+}j_{1}}, \cdots, \overline{L}} |\vec{S}\rangle{=}{\mp}\sum_{|\vec{S}\rangle}\mathcal{C}(\mathcal{B}^{\alpha}_{1, \cdots, i{+}j_{n}{-}L})^{\ast}|\vec{S}\rangle{\otimes} \mathcal{A}^{\alpha}_{\overline{i{+}j_{1}}, \cdots, \overline{L}} |\vec{S}\rangle. \tag{24} \end{align}\] Inserting in Eqs. 23 and 24 , the resolution of identity \(\mathbb{I}_{1,{\cdots}, L}{\otimes}\mathbb{I}_{\overline{1},{\cdots}, \overline{L}}{\equiv}\sum_{|\vec{F}\rangle,|\vec{G}\rangle \in\mathcal{H}_{L}}(|\vec{F}\rangle\langle \vec{F}|)_{1,{\cdots}, L}{\otimes} (|\vec{G}\rangle\langle \vec{G}|)_{\overline{1},{\cdots}, \overline{L}}\), we get \[\begin{align} \Big( H^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}{+}\hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}}\Big) |\Lambda\rangle_\mathcal{C}= {\pm}& \sum_{|\vec{S}\rangle,|\vec{F}\rangle,|\vec{G}\rangle} \Bigg[ \langle (\vec{F}|(\mathcal{A}^{\alpha}_{i{+}j_{1}, \cdots, L})^{\ast}|\vec{S}\rangle)_{1,{\cdots}, L} (\langle \vec{G}|\mathcal{B}^{\alpha}_{\overline{1}, \cdots, \overline{i{+}j_{n}{-}L}}|\vec{S}\rangle)_{\overline{1},{\cdots}, \overline{L}} -\\ &(\langle \vec{F}|(\mathcal{B}^{\alpha}_{1, \cdots, i{+}j_{n}{-}L})^{\ast}|\vec{S}\rangle)_{1,{\cdots}, L} (\langle \vec{G}|\mathcal{A}^{\alpha}_{\overline{i{+}j_{1}}, \cdots, \overline{L}}|\vec{S}\rangle)_{\overline{1},{\cdots}, \overline{L}}\Bigg]\mathcal{C}|\vec{F}\rangle\otimes|\vec{G}\rangle. \end{align}\] Using \((\langle \vec{G}|\mathcal{A}^{\alpha}_{\overline{i{+}j_{1}}, \cdots, \overline{L}}|\vec{S}\rangle)_{\overline{1},\cdots,\overline{L}}{=}(\langle \vec{G}|\mathcal{A}^{\alpha}_{{i{+}j_{1}}, \cdots, {L}}|\vec{S}\rangle)_{1,\cdots,L}\) (similarly for \(\mathcal{B}^{\alpha})\) we obtain \[\begin{align} \Big( H^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}{+}\hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}}\Big) |\Lambda\rangle_\mathcal{C}= {\pm} \sum_{|\vec{S}\rangle,|\vec{F}\rangle,|\vec{G}\rangle} \Bigg[ &\langle \vec{F}|(\mathcal{A}^{\alpha}_{i{+}j_{1}, \cdots, L})^{\ast}|\vec{S}\rangle\langle \vec{G}|\mathcal{B}^{\alpha}_{{1}, \cdots, {i{+}j_{n}{-}L}}|\vec{S}\rangle -\\ &\langle \vec{F}|(\mathcal{B}^{\alpha}_{1, \cdots, i{+}j_{n}{-}L})^{\ast}|\vec{S}\rangle \langle \vec{G}|\mathcal{A}^{\alpha}_{{i{+}j_{1}}, \cdots, {L}}|\vec{S}\rangle\Bigg]\mathcal{C}|\vec{F}\rangle\otimes|\vec{G}\rangle. \end{align} \label{suppeq:32Proof195part1}\tag{25}\] Now the hermiticity of \(\mathcal{A}^{\alpha}\) and \(\mathcal{B}^{\alpha}\) implies \(\langle \vec{F}|(\mathcal{A}^{\alpha}_{i{+}j_{1}, \cdots, L})^{\ast}|\vec{S}\rangle{=}\langle \vec{S}|\mathcal{A}^{\alpha}_{i{+}j_{1}, \cdots, L}|\vec{F}\rangle\) and \(\langle \vec{F}|(\mathcal{B}^{\alpha}_{1, \cdots, i{+}j_{n}{-}L})^{\ast}|\vec{S}\rangle{=}\) \(\langle \vec{S}|\mathcal{B}^{\alpha}_{1, \cdots, i{+}j_{n}{-}L}|\vec{F}\rangle\) respectively, leading us to \[\begin{align} \Big(H^{\alpha}_{i{+}j_{1},\cdots,i{+}j_{n}}{+}\hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}}\Big) |\Lambda\rangle_{\mC} = {\pm} \sum_{|\vec{F}\rangle,|\vec{S}\rangle,|\vec{G}\rangle} \Bigg[& \langle \vec{G}|\mathcal{B}^{\alpha}_{{1}, \cdots, {i{+}j_{n}{-}L}}|\vec{S}\rangle \langle \vec{S}|\mathcal{A}^{\alpha}_{i{+}j_{1}, \cdots, L}|\vec{F}\rangle-\\ & \langle \vec{G}|\mathcal{A}^{\alpha}_{{i{+}j_{1}}, \cdots, {L}}|\vec{S}\rangle\langle \vec{S}|\mathcal{B}^{\alpha}_{1, \cdots, i{+}j_{n}{-}L}|\vec{F}\rangle\Bigg]\mathcal{C}|\vec{F}\rangle\otimes|\vec{G}\rangle. \end{align} \label{suppeq:32Proof195part2}\tag{26}\] Since \(\sum_{|\vec{S}\rangle \in \mathcal{H}_{L}}|\vec{S}\rangle\langle \vec{S}|{=}\mathbb{I}\), Eq. 26 implies \(\Big(\hat{H}^{\alpha}_{i+j_{1},\cdots,i+j_{n}}{+}\hat{H}^{\alpha}_{\overline{i{+}j_{1}},\cdots,\overline{i{+}j_{n}}}\Big) |\Lambda\rangle_{\mathcal{C}}{=}0\), completing the proof. ◻
Lemma 1. For a spin-\(S\) system, if \(\mC_{l}\) is an invertible operator at site \(l\) and \(K_{l}\) is a non-zero complex number satisfying the condition \[\mC_{l}[(S_{l}^{\mu})^{q}]^{\ast}=K_{l}(S_{l}^{\mu})^{q}\mC_{l}, \label{suppeq:32condition95K}\qquad{(6)}\] then \(K_{l}\) can take only values \(\pm 1\).
Proof. To simplify the analysis, we express each operator in Eq. ?? as \((2S{+}1){\times}(2S{+}1)\) matrix in the basis where \(S_{i}^{\mu}\) is diagonal. In this diagonal basis of \(S_{i}^{\mu}\), comparing row-\(m\) and column-\(n\) element of both sides of Eq. ?? gives the equation \[(\mathfrak{s}_{n})^{q}(\mathcal{C}_{l})_{mn}{=}K_{l}(\mathfrak{s}_{m})^{q}(\mathcal{C}_{l})_{mn} \text{for all m,n\in\{0,\cdots,2S\} }, \label{suppeq:32comparing95each95entry}\tag{27}\] where \(\mathfrak{s}_{m}{=}{-}S{+}m\) is the eigenvalue of \(S_{i}^{\mu}\). Now to determine the possible values of \(K_{l}\), we begin by dividing the potential values of \(K_{l}\) (which are all non-zero complex numbers) into two distinct sets and analyzing the solutions separately:
The set containing all the non-zero finite \(q\)th power of the ratio of eigenvalues of \(S_{i}^{\mu}\) i.e. \[K_{l}=\Big(\frac{\mathfrak{s}_{a}}{\mathfrak{s}_{b}}\Big)^q\text{where}a,b\in\{0,\cdots,S{-}1,S{+}1,\cdots,2S\}. \label{suppeq:32all95Kvalues95without95C}\tag{28}\] (Note that neither \(a\) nor \(b\) can equal \(S\) since we are only considering non-zero finite \(K_{l}\) in this set.)
The set of all other non-zero complex numbers.
If \(K_{l}\) is not of the form given in Eq. 28 , for Eq. 27 to hold for all values \(m\) and \(n\), we must have \((\mC_{l})_{mn}{\times}(\mathfrak{s}_{n})^{q}{=}0\) [implies only non-zero element in \(\mathcal{C}_{l}\) can be \((\mathcal{C}_{l})_{m,S}\) since \(\mathfrak{s}_{S}{=}0\)] and \((\mC_{l})_{mn}{\times}(\mathfrak{s}_{m})^{q}{=}0\) [implies only non-zero element in \(\mathcal{C}_{l}\) can be \((\mathcal{C}_{l})_{S,n}\) since \(\mathfrak{s}_{S}{=}0\)] for all \(m,n\). This implies that the matrix elements of \(\mC_{l}\) are of the form \((\mC_{l})_{mn}{=}C_{S}\delta_{m, S}\delta_{n, S}\) where \(C_{S}\) is an arbitrary complex number [i.e., the only non-zero element in \(\mathcal{C}_{l}\) can be (\(\mathcal{C}_{l})_{S, S}\)], and hence \(\det[\mC_{l}]{=}0\) for any value of \(C_{S}\). This means that \(\mC_l\) must be non-invertible, violating the assumptions of the Lemma 1. Thus the values of \(K_{l}\) in set (II) are not valid solutions.
Now from the set of values of \(K_{l}\) given in Eq. 28 , we want to find the values of \(K_{l}\) consistent with \(\mC_l\) being invertible. The eigenvalues \(\mathfrak{s}_{a}\) and \(\mathfrak{s}_{b}\) satisfy one of the following relations
\(\mathfrak{s}_{a}{=}\mathfrak{s}_{b}\) (when \(a{=}b\)). This would mean that \(K_l{=}{+}1\).
\(\mathfrak{s}_{a}{=}{-}\mathfrak{s}_{b}\) (when \(a{=}2S{-}b\)). This would mean that \(K_l{=}{-}1\) when \(q\) is odd and \(K_l = +1\) when \(q\) is even.
\(|\mathfrak{s}_{a}|{\neq}|\mathfrak{s}_{b}|\) (otherwise). This would mean that \(K_l{\neq}{\pm}1\).
Therefore, we can divide the range of values of \(K_{l}\) in Eq. 28 into three categories. We then solve for \(\mC_{l}\) using Eq. 27 for each \(K_{l}\) in these categories and check if the invertibility condition is met. The detailed procedure is outlined below.
Category (i): Using \(K_{l}{=}1\) in Eq. 27 we get, \[(\mathfrak{s}_{n})^{q}(\mathcal{C}_{l})_{mn}{=}(\mathfrak{s}_{m})^{q}(\mathcal{C}_{l})_{mn} \text{for all m,n\in\{0,\cdots,2S\} }. \label{suppeq:32comparing95each95entry95for95K611}\tag{29}\] Now solving Eq. 29 for all \(m,n\), we find the matrix elements of \(\mC_{l}\) can be given by \((\mC_{l})_{mn}{=}C_{m}\delta_{mn}\), where \(\{C_{m}\}\) are arbitrary complex numbers. For any non-zero set of \(\{C_{m}\}\), the invertibility condition \(\text{det}[\mC_{l}]{\neq}0\) can be satisfied, hence \(K_{l}{=}1\) is a potentially valid solution for \(K_{l}\).
Category (ii): Substituting \(K_{l}{=}{\pm}1\) into Eq. 27 for all values of \(m,n\), we find the matrix elements \((\mC_{l})_{mn}{=}C_{n}\delta_{2S{-}m,n}\), where \(\{C_{n}\}\) are arbitrary complex numbers. Similarly, any non-zero set of \(\{C_{n}\}\) also satisfies \(\text{det}[\mC_{l}]{\neq}0\) consistent with \(\mathcal{C}\) being invertible, so \(K_{l}{=}{\pm}1\) is also a potentially valid solution for \(K_{l}\).
Category (iii): From Eq. 28 let us consider an arbitrary value of \(K_{l}\) given by \(K_{l}{=} (\mathfrak{s}_{a_{1}})^{q}/(\mathfrak{s}_{b_{1}})^{q}\) in this category (i.e., with \(a_{1}{\neq}b_{1}\) and \(a_{1}{\neq}2S{-}b_{1}\)). Substituting this value of \(K_{l}\) back into Eq. 27 , we get the matrix element of \(\mC_{l}\): \[(\mC_{l})_{mn}{=}C_{mn}(\delta_{m,a_{1}}\delta_{n,b_{1}}{+}\delta_{m,2S{-}a_{1}}\delta_{n,2S{-}b_{1}}), \label{suppeq:32non-inverible32C}\tag{30}\] where \(\{C_{mn}\}\) are arbitrary complex numbers. Any set of values of \(\{C_{mn}\}\) leads to \(\text{det}[\mC_{l}]{=}0\) making it an invalid solution for \(K_{l}\) under the assumptions of the Lemma 1.
Thus, with the invertibility constraint on \(\mC_{l}\), the only valid non-zero values for \(K_{l}\) are \({\pm}1\), completing the proof of the Lemma 1. ◻
Figure 3: Schematic representation of some more \(|\Lambda\rangle_\mathcal{C}\) type zero energy eigenstates for the spin-\(S\) \(XY\) model with transverse field in different geometries. The black dots denote the sites and the solid lines between them indicate the \(XY\) interaction. The spins at the red (blue) sites \(r\) are correlated with their partners \(\overline{r}\) in the state \(|\kappa_{S}^{-}\rangle\) (\(|\kappa_{S}^{+}\rangle\)) [defined in Eq. 31 ]. \((a)\), \((b)\), and \((c)\) illustrate distinct zero energy states on a \(2L{\times}2L\) square lattice with PBC, with \(L{=}2\). \((d)\), \((e)\), and \((f)\) show the eigenstate on a \(2L{\times}2L{\times}M\) cuboid with PBC, with \((L, M){=}(2, 2)\).
In this section, we provide additional examples of exact zero energy eigenstates of the spin-\(S\) \(XY\) model with a transverse field in geometries different from those we considered in the main text. The states in Fig. 3\((a{-}c)\) are different zero energy states on a \(2L{\times}2L\) square lattice with PBC, while states in Fig. 3\((d{-}f)\) depict distinct zero energy states on a \(2L{\times}2L{\times}M\) cuboid with PBC. The spins at sites \(r\) and \(\overline{r}\) share the state \(|\kappa_{S}^{-}\rangle\) (when marked in red) and the state \(|\kappa_{S}^{+}\rangle\) (when marked in blue), where \[|\kappa_{S}^{\pm}\rangle_{i,j}{=}\frac{1}{\sqrt{2S{+}1}}\sum_{p{=}0}^{2S}({\pm}1)^{p}(S^{-})^{p}|S\rangle_{i}{\otimes}(S^{+})^{p}|{-}S\rangle_{j}. \label{suppeq:32kappa}\tag{31}\] Note that by interchanging the dimers (i.e., changing the red color to blue color and vice-versa), we get a different zero energy state (which can be verified to be orthogonal to the previous ones in this case) in each case. The examples presented in Fig. 3 serve as representatives of classes of zero energy eigenstates in the spin-\(S\) \(XY\) model in higher dimensions. A systematic investigation of the various classes, in particular, enumerating the number of distinct classes and the corresponding number of states within each class, is an interesting direction worth exploring in the future.
In this section, we apply our formalism to previously studied spin-\(1/2\) Hamiltonians in the literature to demonstrate that it can reproduce the same zero energy states as reported in those works. Note that for all of the following examples, we consider the Hilbert space defined by computational basis \(|\vec{b}\rangle{=}{\bigotimes}_{i{=}1}^{L}|b_{i}\rangle\) \((\mathcal{H}_{L}{\equiv}\text{span}\{|\vec{b}\rangle\})\), where \(\vec{b}{=}(b_{1}, b_{2}, {\cdots}, b_{L})\) is either a string of spins, i.e., \(b_j \in \{\uparrow, \downarrow\}\), or a “bit-string" of length \(L\), i.e., \(b_j \in \{0, 1\}\).
First, we consider the following Hamiltonian with a three-spin interaction and a transverse magnetic field for a system of size \(L\) (with PBC) that was studied in Ref. [60] \[\begin{align} H_{1}(L){=}\sum_{i{=}1}^{L}\big(h\sigma_{i}^{z}+J\sigma_{i}^{x}\sigma_{i+1}^{x}\sigma_{i+2}^{x}\big), \end{align} \label{suppeq:32three95spin95ising}\tag{32}\] where \(h\) and \(J\) are arbitrary constants. Here each \(\hat{H}^{\alpha}_{i+j_{1},\cdots,i+j_{n}}\), i.e., \(\sigma_{i}^{x}\sigma_{i+1}^{x}\sigma_{i+2}^{x}\) and \(\sigma_{i}^{z}\) are real in the chosen basis and anti-commute with the operator \(\mathcal{C}_{1}{=}\prod_{i}^{L} \sigma_{i}^{y}\) [hence they satisfy Eq. 14 ]. Thus using Theorem 1, a zero energy eigenstate of Hamiltonian \(H_{1}(2L)\) is given by \[|\Lambda\rangle_{\mathcal{C}_{1}}{=}\bigotimes_{i{=}1}^{L}|\Phi^{-}\rangle_{i,\overline{i}}, \label{suppeq:32Lambda95state1}\tag{33}\] where \(|\Phi^{\pm}\rangle_{i,j}{=}(|01\rangle {\pm} |10\rangle)_{i,j}/\sqrt{2}\). Similarly, when the half-chain length \(L\) is a multiple of three, the operator \(\mathcal{C}_{1}'{=}\prod_{p{=}1}^{\frac{L}{3}} \sigma_{3p{-}2}^{x}\sigma_{3p{-}1}^{y}\sigma_{3p}^{x}\) anti-commutes with each \(\hat{H}_{i{+}j_{1}, {\cdots}, i{+}j_{n}}^{\alpha}\) [thus satisfying Eq. 14 ] resulting in another zero energy eigenstate of the form \[|\Lambda\rangle_{\mC_{1}'}{=}\bigotimes_{i{=}1}^{L{-}2}|\Phi^{+}\rangle_{i,\overline{i}}|\Phi^{-}\rangle_{i{+}1,\overline{i{+}1}}|\Phi^{+}\rangle_{i{+}2,\overline{i{+}2}}. \label{suppeq:32Lambda95state1p}\tag{34}\] Note that due to PBC the translations of the state \(|\Lambda\rangle_{\mC_{1}'}\) by one site, i.e., \(|\Lambda^{\mathcal{T}_1}\rangle_{\mC_{1}'}{=}\bigotimes_{i{=}1}^{L{-}2}|\Phi^{-}\rangle_{i,\overline{i}}|\Phi^{+}\rangle_{i{+}1,\overline{i{+}1}}|\Phi^{+}\rangle_{i{+}2,\overline{i{+}2}}\) and two sites i.e. \(|\Lambda^{\mathcal{T}_2}\rangle_{\mC_{1}'}{=}\bigotimes_{i{=}1}^{L{-}2}|\Phi^{+}\rangle_{i,\overline{i}}|\Phi^{+}\rangle_{i{+}1,\overline{i{+}1}}|\Phi^{-}\rangle_{i{+}2,\overline{i{+}2}}\) are also distinct zero energy eigenstates of \(H_{1}(2L)\). The states of Eqs. 33 and 34 are exactly the volume-law entangled states reported in Ref. [60].
Next, we reproduce the volume-law state constructed in the PXP model in Ref. [59] using our framework. The PXP model for a spin chain of size \(L\) with PBC is described by the Hamiltonian [29] \[H_{\rm PXP}(L)=\sum_{i{=}1}^{L} \hat{P}_{i{-}1}\sigma_{i}^{x}\hat{P}_{i{+}1}=\sum_{i{=}1}^{L} \Big[\sigma_{i}^{x}{+}\sigma_{i}^{x}\sigma_{i{+}1}^{z}{+}\sigma_{i{-}1}^{z}\sigma_{i}^{x}{+}\sigma_{i{-}1}^{z}\sigma_{i}^{x}\sigma_{i{+}1}^{z}\Big]. \label{suppeq:32PXP95Hamiltonian}\tag{35}\] Here, we used \(\hat{P}_{i}{=}(\mathbb{I}{+}\sigma_{i}^{z})/2\) to write the Hamiltonian in the form given in Eq. 13 . Now in our chosen computational basis the global particle-hole symmetry operator \(\mathcal{C}_{\rm PXP}{=}\prod_{i{=}1}^{L}\sigma_{i}^{z}\) satisfies the conditions stated in Eq. 14 . Thus, using Theorem 1, a zero energy eigenstate of \(H_{\rm PXP}(2L)\) is \[|\Lambda\rangle_{\rm PXP}{=}\bigotimes_{i{=}1}^{L}|\Psi^{-}\rangle_{i,\overline{i}}{=} \sum_{|\vec{b}\rangle\in\mathcal{H}_{L}} ({-}1)^{|b|}|\vec{b}\rangle_{1,\cdots,L} \otimes |\vec{b}\rangle_{\overline{1},\cdots,\overline{L}}, \label{suppeq:32Lambda95PXP}\tag{36}\] where we have used \(\mC_{\rm PXP}|\vec{b}\rangle{=}({-}1)^{|b|}|\vec{b}\rangle\) with \(|b|\) denoting the parity of the bit string \(\vec{b}\) and \(|\Psi^{-}\rangle_{i,j}{=}(|00\rangle {-} |11\rangle)_{i,j}/\sqrt{2}\)
While \(\ket{\Lambda}_{\rm PXP}\) is an eigenstate of the PXP model in the full \(2^L\)-dimensional Hilbert space, the Hilbert space of the PXP model is usually truncated to the so-called nearest-neighbor Rydberg blockade Hilbert space [29]. This truncated blockade subspace \(\mathcal{F}_{L}\) for chains of \(L\) spins with PBC is spanned by basis vectors \(|\vec{f}\rangle\) whose bit-strings \(\vec{f}\) do not contain the sub-bit-string “11". The projector on this subspace defined by \(\mathcal{P}_{\text{Ryd}}{=}\prod_{i}(\mathbb{I}{-}\hat{Q}_{i}\hat{Q}_{i{+}1})\) (i.e., \(\mathcal{P}_{\text{Ryd}}\mH_{L}{=}\mathcal{F}_{L})\), where \(\hat{Q}_{i}{=}\mathbb{I}{-}\hat{P}_{i}\), commutes with the Hamiltonian \(H_{\rm PXP}\), which enables the truncation of this Hilbert space for the PXP Hamiltonian. Thus, the projection of the state \(|\Lambda\rangle_{\rm PXP}\) into the Rydberg blockade sector should also be an eigenstate of the PXP model truncated into this Hilbert space, and this state reads \[|\Lambda\rangle_{\rm Ryd}{=}\mathcal{P}_{\text{Ryd}}|\Lambda\rangle_{\rm PXP}{=}\sum_{|\vec{f}\rangle\in \mathcal{F}_{L}}({-}1)^{|f|}|\vec{f}\rangle_{1,\cdots, L}\otimes |\vec{f}\rangle_{L{+}1,\cdots, 2L}. \label{suppeq:32PXP95eigenstate95Rydberg95subspace}\tag{37}\] This is precisely the volume-law eigenstate reported in Ref. [59] for the PXP chain. We note that this construction also works for the PXP model in arbitrary dimensions and in a wide variety of geometries, which then reproduce the results of Ref. [59].
Finally, we apply our formalism to a next-nearest-neighbor interacting Hamiltonian for a system of size \(L\) with PBC considered in Ref. [61] given by \[\begin{align} H_{\rm NNN}(L){=}\sum_{i{=}1}^{L} \Big( J^{xzx}\sigma_{i}^{x}\sigma_{i{+}1}^{z}\sigma_{i{+}2}^{x}{+}J^{yzy}\sigma_{i}^{y}\sigma_{i{+}1}^{z}\sigma_{i{+}2}^{y}{+}J^{xzy}\sigma_{i}^{x}\sigma_{i{+}1}^{z}\sigma_{i{+}2}^{y}{+}J^{yzx}\sigma_{i}^{y}\sigma_{i{+}1}^{z}\sigma_{i{+}2}^{x}{+}\\ J^{zzz}\sigma_{i}^{z}\sigma_{i{+}1}^{z}\sigma_{i{+}2}^{z}{+} J^{xx}\sigma_{i}^{x}\sigma_{i{+}1}^{x}{+} J^{yy}\sigma_{i}^{y}\sigma_{i{+}1}^{y}{+} J^{xy}\sigma_{i}^{x}\sigma_{i{+}1}^{y}{+} J^{yx}\sigma_{i}^{y}\sigma_{i{+}1}^{x}{+} J^{z}\sigma_{i}^{z}\Big), \end{align}\] where all \(J\)’s are arbitrary. In our chosen computational basis, the Hamiltonian consists of terms \(\hat{H}^{\alpha}_{i+j_{1},\cdots,i+j_{n}}\) that are either purely real (i.e., \(\sigma_{i}^{x}\sigma_{i{+}1}^{z}\sigma_{i{+}2}^{x}\), \(\sigma_{i}^{y}\sigma_{i{+}1}^{z}\sigma_{i{+}2}^{y}\), \(\sigma_{i}^{z}\sigma_{i{+}1}^{z}\sigma_{i{+}2}^{z}\), \(\sigma_{i}^{x}\sigma_{i{+}1}^{x}\), \(\sigma_{i}^{y}\sigma_{i{+}1}^{y}\) and \(\sigma_{i}^{z}\)) or purely imaginary (i.e., \(\sigma_{i}^{x}\sigma_{i{+}1}^{z}\sigma_{i{+}2}^{y}\), \(\sigma_{i}^{y}\sigma_{i{+}1}^{z}\sigma_{i{+}2}^{x}\), \(\sigma_{i}^{x}\sigma_{i{+}1}^{y}\) and \(\sigma_{i}^{y}\sigma_{i{+}1}^{x}\)). When \(L\) is even, all real \(\hat{H}^{\alpha}_{i+j_{1},\cdots,i+j_{n}}\) anticommute with an operator \(\mathcal{C}_{\rm NNN}{=}\prod_{p{=}1}^{\frac{L}{2}}\sigma_{2p{-}1}^{x}\sigma_{2p}^{y}\) while all purely imaginary \(\hat{H}^{\alpha}_{i+j_{1},\cdots,i+j_{n}}\) commute with \(\mathcal{C}_{\rm NNN}\) satisfying the condition given in Eq. 14 . Thus, as a result of Theorem 1, the Hamiltonian \(H_{\rm NNN}(2L)\) hosts a zero energy eigenstate given by \[|\Lambda\rangle_{\rm NNN}{=}\bigotimes_{i{=}1}^{L{-}1}|\Phi^{-}\rangle_{i,\overline{i}}|\Phi^{+}\rangle_{i{+}1,\overline{i{+}1}}, \label{suppeq:32EAP95eigenstate}\tag{38}\] where \(|\Phi^{\pm}\rangle_{i,j}{=}(|01\rangle {\pm} |10\rangle)_{i,j}/\sqrt{2}\). Note that due to PBC the translation of the state by one site given by \[|\Lambda^{\mathcal{T}_1}\rangle_{\rm NNN}{=}\bigotimes_{i{=}1}^{L{-}1}|\Phi^{+}\rangle_{i,\overline{i}}|\Phi^{-}\rangle_{i{+}1,\overline{i{+}1}}\] produces another distinct zero energy eigenstate (orthogonal to \(|\Lambda\rangle_{\rm NNN}\)). These are precisely the entangled antipodal paired (EAP) states [named so since there are Bell-pairs between antipodal points when the lattice sites are arranged uniformly on a circle] identified in Ref. [61].