We construct new families of spin chain Hamiltonians that are local, integrable, and translationally invariant. To do so, we make use of the Clifford group that arises in quantum information theory. We consider translation invariant Clifford group transformations that can be described by matrix product operators (MPOs). We classify translation invariant Clifford group transformations that consist of a shift operator and an MPO of bond dimension two—this includes transformations that preserve locality of all Hamiltonians and those that lead to non-local images of particular operators but, nevertheless, preserve locality of certain Hamiltonians. We characterize translation invariant Clifford group transformations that take single-site Pauli operators to local operators on at most five sites—examples of Quantum Cellular Automata—leading to a discrete family of Hamiltonians that are equivalent to the canonical XXZ model under such transformations. For spin chains solvable by the algebraic Bethe ansatz, we explain how conjugating by an MPO affects the underlying integrable structure. This allows us to relate our results to the usual classifications of integrable Hamiltonians. We also treat the case of spin chains solvable by free fermions.

Integrable models are fundamental tools in our understanding of quantum many-body systems and statistical mechanics.1–4 While a general definition of quantum integrability is open to debate,5 there are many examples of well-established integrable models. Two families of integrable models that are particularly well studied and that are of interest to us are Bethe ansatz solvable models,2 including the Heisenberg and XXZ spin chains,1–3,6,7 and free-fermion models, including the XY and Ising spin chains.8 An important topic is the classification of Bethe ansatz solvable models,9–14 and there are conjectured conditions implying integrability of a Hamiltonian.13,15 There has, moreover, been significant recent interest in finding new models and deriving conditions where a spin chain can be solved by free-fermion (and free-parafermion) methods.16–24 

Given known integrable Hamiltonians, we can consider transformations applied to them and between them. This can lead to finding new models that can be solved in the same way as known models. Transformations between exactly solvable models are, moreover, important in the field of symmetry-protected topological (SPT) phases; then, it is useful to identify transformations that take one from a simple, exactly solvable model in the trivial phase to an analogous exactly solvable model with non-trivial symmetry properties.25–27 In this paper, we will construct new families of spin chain Hamiltonians using Clifford group transformations, a family of transformations arising in quantum information theory that we describe below.

To set the scene, consider the following Hamiltonian for a spin-1/2 chain:

H0=nsitesZn3Yn2Xn1YnYn+1Xn+2Yn+3Zn+4+Zn3Yn2Yn1Yn+2Yn+3Zn+4+ΔZn1YnYn+1Zn+2
(1)

with −1 < Δ < 1, where X, Y, and Z are the standard Pauli operators. Perhaps surprisingly, this Hamiltonian is integrable. In particular, it can be analyzed using Bethe ansatz methods, and from that, one can derive highly non-trivial results. For example, letting angle brackets denote the ground state expectation value, one can rigorously prove the following large M asymptotics:

Z1X2Z3ZM+1XM+2ZM+3=18η2M2(1+o(1)),0<η<π/4,(1)MbMπ2η(1+o(1)),π/4<η<π/2,
(2)

where 2η = arccos(Δ) and b is independent of M. As we will explain, this integrability follows by giving a unitary transformation from H0 to the canonical XXZ spin chain,

HXXZ=nsitesXnXn+1+YnYn+1+ΔZnZn+1.
(3)

We can, then, use established results for the ⟨Z1ZM+1⟩ correlation function in the XXZ chain.28 Moreover, using conformal field theory (CFT) methods [H0 has low energy behavior described by a compactified free boson CFT for −1 < Δ < 1, with the SU(2)-point at Δ = −1], one can derive further ground state correlation functions in this model.2,28,29 We will show that, unlike HXXZ, H0 is at a transition between SPT phases26,30–32 (indeed, unitary transformations preserve the structure of the phase diagram, but not necessarily any symmetry properties).

Now, the fact that, say, the eigenstates of the Hamiltonians H and H′ = UHU, for some unitary U, are simply related is elementary. However, a typical unitary will not preserve the locality of the Hamiltonian, by which we mean that the Hamiltonian is a sum of terms, each of which acts non-trivially on a number of sites that is independent of the length of the chain. Moreover, a given local Hamiltonian is typically not integrable. The purpose of this paper is to construct new translationally invariant and local Hamiltonians that are integrable by virtue of being unitarily related to canonical integrable models, such as HXXZ. We do this through making connections to techniques from quantum information theory. In particular, we study interesting examples from the Clifford group of unitary transformations.

The Clifford group is a discrete subgroup of the full unitary group and consists of those transformations that map products (or strings) of Pauli operators to other products of Pauli operators under conjugation; for example,

X1X1,X2Z2,Z1Z1Z2,Z2X1X2
(4)

is a two-site Clifford group transformation. Since these transformations are unitary, they are constrained by the necessity of preserving the commutation relations between Pauli operators. One reason for considering the Clifford group is that many integrable spin-1/2 chains have a simple form written as a sum of Pauli strings and these transformations preserve that—for example, HXXZ has three types of interaction term, as does H0. Symmetries that are Pauli strings, such as the spin-flip ∏nZn, remain of this form after the transformation. This can help in the analysis of symmetries in the context of SPT phases,33,34 and so this can be used to find examples of integrable models (both gapped and gapless) with non-trivial SPT physics. For the case of H0, this analysis is given in  Appendix A.

Consider, then, a spin-1/2 chain with L sites. One approach to finding Clifford group transformations on such a chain is to start with a simple transformation for each site, such as the following that appears frequently in the literature on SPT phases:25–27 

XnZn1XnZn+1,ZnZn.
(5)

This transformation maps the trivial paramagnet with Hamiltonian −∑nXn to the SPT cluster model with Hamiltonian −∑nZn−1XnZn+1 (for this reason, it is sometimes referred to as an SPT entangler). One can check that (5) preserves the commutation relations, giving a Clifford group transformation.35 One nice feature of this transformation is that it preserves translation invariance—this is a physically interesting case and is a common, though inessential,3,28 feature of many integrable models. A key concern of this paper is to identify translation invariant transformations that, in turn, lead to translation invariant Hamiltonians.

Note that transformation (5), taking single-site operators to (at most) three-site operators for each site n, defines a Clifford group transformation on the whole chain. Let us consider, instead, the following partial definition of a transformation taking single-site operators to three-site operators:

X1X0X1Z2,Z1Z1.
(6)

There are many ways of extending this to a valid three-site Clifford transformation and, moreover, to a Clifford transformation on the whole chain. However, any way we do so will break translation invariance. In particular, there cannot be a valid Clifford transformation on the whole chain that acts as (6) on each site,

Xn?Xn1XnZn+1,Zn?Zn.
(7)

Indeed, such a transformation would not be unitary since, say, mapping X1X0X1Z2 and X3X2X3Z4 is inconsistent with the commutation relation [X1, X3] = 0. We, thus, see that it is a non-trivial problem to identify those local Clifford transformations that can apply in a translation invariant fashion on the whole chain. We present families of such transformations in Fig. 1.

FIG. 1.

Translation invariant Clifford group transformations: (a) Uchain = … U3,4U2,3U1,2 and (b) Uchain = … U3,4,5U2,3,4U1,2,3.

FIG. 1.

Translation invariant Clifford group transformations: (a) Uchain = … U3,4U2,3U1,2 and (b) Uchain = … U3,4,5U2,3,4U1,2,3.

Close modal

General Clifford transformations can be built from a discrete set of one- and two-site operations: CNOT, Hadamard, and Phase36,37 with matrix forms

1000010000010010,121111,and100i,
(8)

respectively. It is natural, then, to consider how to construct a translation invariant Clifford group transformation out of basic local Clifford transformations. One way to extend a single basic Clifford transformation U to a Clifford transformation Uchain acting on the whole chain is given in Fig. 1 for two- and three-site U, respectively. In fact, transformation (5) takes this form with a basic Clifford transformation that acts by conjugation as

XnXnZn+1,ZnZn,Xn+1ZnXn+1,Zn+1Zn+1.
(9)

While this construction allows us to build translation invariant Clifford transformations from simple basic Clifford transformations U, the transformations Uchain of Fig. 1 have a non-local causal structure and so, for a general choice of U, will not necessarily preserve the locality of the Hamiltonian that we transform. Indeed, consider applying (4) sequentially to pairs of sites, as in Fig. 1(a): this means that Uchain is a product of basic transformations that act by conjugation as

XnXn,Xn+1Zn+1,ZnZnZn+1,Zn+1XnXn+1.
(10)

We can see that on applying Uchain, the operator X2, for example, will transform into a non-local string: X2Z2Z2Z3Z2Z3Z4 → ⋯. One part of our analysis is identifying which cases maintain locality for all choices of Hamiltonian. We will call these transformations, which take all local operators to other local operators, locality-preserving transformations (where a local operator acts non-trivially on a number of sites independent of the length of the chain). The other Clifford transformations, which we will call locality-non-preserving, are, however, of real physical interest. In particular, while they take some local operators to infinite strings, they can transform certain local Hamiltonians to other local Hamiltonians (in the bulk). For example, the operator Uchain derived from (10) can be recognized as the Kramers–Wannier duality38 of the Ising model (again, for operators in the bulk; we will discuss this further below; see also Ref. 39).

On a spin-1/2 chain, transformations that map local strings of Pauli operators to local strings of Pauli operators are locality-preserving. In fact, the purely mathematical question that we arrive at by asking which translation invariant Clifford transformations maintain locality for all Pauli operators arises in the field of Clifford Quantum Cellular Automata (QCA). To be precise, a Clifford QCA is a translation invariant, reversible, and local transformation that takes Pauli operators to Pauli operators.40,41 The motivation in that setting is rather different: a QCA is a discrete time dynamical system, and the purpose of the locality condition is to ensure that the system has a finite propagation speed. There is an established theory of Clifford QCA,40–42 and certain claims made below are consequences of these more general results—we will point these out as appropriate, but aim to give a self-contained presentation and so provide elementary proofs as needed. Moreover, the condition that locality is preserved for every Hamiltonian is stronger than needed for our purposes. Allowing for locality-non-preserving transformations, which, nevertheless, preserve the locality of certain Hamiltonians, leads to a broader class of transformations than those that define a QCA.

Another important connection to make is to the theory of matrix product operators (MPOs).43,44 We explain in Sec. II B that the transformation Uchain in Fig. 1 can be understood as an MPO. This is particularly useful because both MPOs and the algebraic Bethe ansatz1,2 can be formulated in the intuitive graphical notation used in the theory of tensor networks.6 This allows us to begin to analyze, for models solved by the algebraic Bethe ansatz, how the integrable structure changes when we conjugate the Hamiltonian by the transformation Uchain.

This paper is organized as follows. In Sec. II, we motivate and analyze certain families of Clifford group transformations. In Secs. II A and II B, we define Clifford group transformations in general, and then the translation invariant Clifford group transformations of interest. We also explain the connection to MPOs. In Secs. III and IV, we, then, demonstrate how these results can be used to construct families of integrable models equivalent to HXXZ, including H0 given above. Models related by locality-preserving transformations built from a single two-site basic Clifford transformation are given in Sec. III A, while those related by locality-non-preserving transformations are given in Sec. III B. These are a consequence of a classification, given in Sec. III C, of all the possible transformations that arise from transformations Uchain constructed from a single two-site Clifford transformation. We see that, up to changes of on-site basis, only one new Hamiltonian appears in Sec. III A. In Sec. III A 1, we show that this Hamiltonian has some interesting connections to another integrable model, the folded XXZ chain,13,45–48 and to the recently introduced concept of a pivot Hamiltonian.49,50 In Sec. IV, we discuss translation invariant Clifford transformations directly rather than through the Uchain construction. We give, in Sec. IV A, all transformations that map single-site Pauli operators to local operators on up to five sites and show how these can be written using Uchain based on a single two- or three-site Clifford transformation. We give the corresponding families of Hamiltonians related to HXXZ in Sec. IV B. In Sec. V, we proceed to a more general discussion regarding families of integrable Hamiltonians. We focus our discussion on canonical models that can be solved by the quantum inverse scattering method or algebraic Bethe ansatz1,2,6 and explain how transformations of the MPO form can be applied to the integrable structures underlying these models. This includes, as a special case, the Clifford transformations that we have analyzed and, thus, gives our results a context by relating them to classifications of integrable models. We find interesting modifications to the underlying integrable structures, but there remain questions to resolve here. We also treat the case of spin chains solvable by transforming to free fermions, making connections to recent papers on families of exactly solvable spin-1/2 chains.17,19,20,24 Finally, we discuss further avenues of research.

We now introduce the Clifford group in the spirit of Ref. 36. Let P={I,X,Y,Z} denote the set of single-site Pauli operators. Consider the set of Pauli strings on N sites; this is given by

PN=P1P2PNPjP.
(11)

As above, we will omit the tensor product symbol and write elements of PN as j=1NPj. Furthermore, non-trivial Pauli strings on N sites are elements of

PN*=PN\{In}.
(12)

As mentioned, informally, the Clifford group consists of transformations that take Pauli strings to Pauli strings under conjugation. More carefully, the Clifford group on N sites is defined as the following subgroup of U(2N):

CN=UU(2N)P±PN*,UPU±PN*/U(1),
(13)

where ±PN* consists of ±P for PPN*.

One can specify51 a Clifford group transformation by its action on each of {X1, …, XN, Z1, …, ZN}. These operators all mutually commute, apart from operators on the same site where we have the anticommutator {Xj, Zj} = 0. These commutation relations are preserved under conjugation by unitary transformations and, in particular, by Clifford group transformations. Respecting these commutation relations, one can define a Clifford group transformation sequentially, first specifying, say, X1P, where P±PN*; then defining Z1P′, where P±PN* and {P, P′} = 0; and so on. It can be shown that this procedure gives a unique element of the Clifford group,36 and so, in this way, one can derive the order of CN. For example, to define a transformation in C1, one can choose to take X to any of the six elements of ±P1* and Z to any of the four elements of ±P1* that anticommute with that choice, giving |C1|=24. The result |CN|=2N2+2Nj=1N(4j1) can be derived similarly, as shown in Ref. 36. For our spin chain of length L, we will consider many copies of, say, P2. These will correspond to each pair (or, more generally, in the case of PN>2, set) of neighboring sites. For example, we can view both X1X2 and X2X3 as elements of P2 for different subsets of sites on the chain.

We define a notation that corresponds to translating a Pauli operator by j sites: if P=P1(1)P2(2)Pn(n)PN, then let P(j)=Pj+1(1)Pj+2(2)Pj+N(n), with P(k)P fixed. Given a transformation in CN, we specify the sites it acts on and so we can, then, find the appropriate action on P(j). To illustrate this, let UC2 be defined by X1X1X2, X2Z1Z2, Z1Z1, and Z2X2. Then, for Q = X1X2, we have

U12QU12=Y1Y2,U23QU23=X1X2X3,Uj+1,j+2Q(j)Uj+1,j+2=Yj+1Yj+2.
(14)

Local Pauli operators are those that are elements of PN, where N is independent of L, the length of the chain. The translation invariant spin chain Hamiltonians that we consider take the form

H=nαJαPα(n),JαR,PαPN,
(15)

where α labels the different terms that appear and n indexes the sites, where each site has a translated Pauli operator. The couplings Jα are all real since H is Hermitian. For example, H0 defined in (1) would have α ∈ {1, 2, 3}, nZ, and P1 = Z1Y2X3Y4Y5X6Y7Z8.

While we can look at transformations in CL directly, we focus on those Clifford transformations that are of the form Uchain, illustrated in Fig. 1. This is given by

Uchain=nsitesUn,n+1,,n+k1,UCk,
(16)

where the product is ordered so that we act first with the operator acting on the left-most sites of the chain. Such transformations are illustrated for k = 2 and k = 3 in Fig. 1. For k = 1, this is an on-site change of basis. Note that throughout, we are thinking of a large system size L but will not concern ourselves with any subtleties arising in the thermodynamic limit L → ∞ (see, for example, the rigorous approach in Ref. 40). Now, as mentioned in the Introduction, this transformation has the form of an MPO, and this interpretation will be particularly helpful in Sec. V, where we discuss connections to Bethe ansatz methods. We use the standard tensor network graphical notation below52 although we will also write out some formulas explicitly. In Fig. 2, we graphically relate the translation invariant Clifford group transformations of Fig. 1 to MPOs. This particular argument appears, for example, in the theory of matrix product states (MPSs), showing equivalence between MPS and circuits of the form Uchain (not necessarily constructed from basic Clifford transformations).52–55 A translation invariant MPO on a chain n=1LHn is an operator M on the chain of the form

Mj1,,jLi1,,iL=tr(Vj1i1Vj2i2VjLiL)=α1,,αLVj1i1α2α1Vj2i2α3α2VjLiLα1αL.
(17)

For fixed ik, jk (physical indices corresponding to Hk), each of the Vjkik are χ × χ matrices acting on a “virtual space” as Vjkik:VV, where virtual indices are, then, summed over (i.e., we matrix multiply in the virtual space).43,44,52 The virtual space V has dimension χ, referred to as the bond dimension.56 Alternatively, one can think of V as a four index tensor, with two physical indices and two virtual indices. The trace over the virtual space, illustrated in Fig. 2(f), reflects periodic boundary conditions (i.e., H1 and HL are neighboring sites). For open boundary conditions, we do not take the trace and can simply alter the tensors at the boundary (for fixed physical indices, the boundary tensors are vectors rather than matrices)—this is illustrated in Fig. 2(e). To see that Uchain defined in (16) is an MPO, we show in Fig. 2(d) how to find the MPO tensor V from the basic Clifford transformation U. We see that for k = 2, we have57 a χ = 4 MPO, and the generalization for k = 3 gives a χ = 16 MPO.

FIG. 2.

Relating the translation invariant Clifford group transformations of Fig. 1 and MPOs. (a) Uchain = … U3,4U2,3U1,2 and (b) Uchain = … U3,4,5U2,3,4U1,2,3. (c) A periodic analog of (a) given by Uchain = … U3,4U2,3U1,2… . This is interpreted in the main text. In (a)–(c), all lines represent two-dimensional spaces. (d) An MPO tensor V in terms of U that appears in (a) and (c). (e) An MPO with open boundary; the virtual index α has dimension χ. (f) An MPO with periodic boundary; the virtual index α has dimension χ. This shows that (a) and (c) are of the MPO form with χ = 4.

FIG. 2.

Relating the translation invariant Clifford group transformations of Fig. 1 and MPOs. (a) Uchain = … U3,4U2,3U1,2 and (b) Uchain = … U3,4,5U2,3,4U1,2,3. (c) A periodic analog of (a) given by Uchain = … U3,4U2,3U1,2… . This is interpreted in the main text. In (a)–(c), all lines represent two-dimensional spaces. (d) An MPO tensor V in terms of U that appears in (a) and (c). (e) An MPO with open boundary; the virtual index α has dimension χ. (f) An MPO with periodic boundary; the virtual index α has dimension χ. This shows that (a) and (c) are of the MPO form with χ = 4.

Close modal

Now, if the operator Mj1,,jLi1,,iL defined in (17) is unitary, then it is called a matrix product unitary (MPU)—note that we take the trace, and so this operator is of the periodic boundary condition form. It has been shown that MPUs are equivalent to QCA58–60 and so will necessarily be locality-preserving.

Let us now relate this equivalence to the MPOs considered in this work. First, using the graphical argument of Fig. 2, our definition of Uchain in (16) is an MPO with open boundary (there is no trace corresponding to summing over a virtual index between site L and site 1). Since the basic Clifford transformation U is unitary, the resulting MPO is unitary and is, moreover, a Clifford transformation in CL. However, this does not define an MPU since the definition of an MPU is for the periodic boundary condition form of the MPO. Hence, there is no equivalence between this construction61 and QCA. Indeed, for Uchain derived from (10)—the Kramers–Wannier duality—we saw that this family of MPOs is not necessarily locality-preserving. In fact, this duality is discussed in detail in Ref. 39. In that paper, a periodic MPO form for the Kramers–Wannier duality is given. One can show that this MPO fits the construction we are considering and is built out of a basic Clifford transformation,62 

XnZn,Xn+1ZnXn+1,ZnXnZn+1,Zn+1Zn+1.
(18)

Using this basic Clifford to construct Uchain leads to the (bulk) transformation XnZn−1Zn and ZnZn+1Xn. The corresponding periodic MPO [as in Fig. 2(c)] is more subtle. The analysis in Ref. 39 shows that it is not invertible (and hence not unitary) by analyzing the fusion algebra with the spin-flip operator ∏nXn. More generally, if we take a basic Clifford transformation U and use this to construct a periodic boundary MPO [given in Fig. 2(c)], this is not necessarily unitary and so is not necessarily an MPU. However, if Uchain is locality preserving, then conjugation by Uchain acts as a Clifford QCA. This means by the general arguments in Refs. 58 and 59 that there exists a periodic MPU form. Our analysis below allows us, using elementary calculations, to give such a periodic form directly for transformations of interest (see Sec. IV C).

Our calculations focus on the open boundary case, which explicitly breaks translation invariance by fixing a left-most site. However, such transformations preserve bulk translation invariance in the following way. Let us index the sites here so that the first site has n ≃ −L/2 and the last site has nL/2. Then, given P, a local Pauli operator in the bulk, if Uchain transforms PQ, where Q is local, then we have that the translated operators transform as P(j) → Q(j) for j independent of L. Note that the causal structure of Uchain means that we can have a locality-non-preserving transformation such that, on conjugating local operators, the non-local images extend to the right. Instead of conjugating by Uchain, we could conjugate by Uchain, giving locality-non-preserving transformations where the non-local images extend to the left. We do not analyze this case, but analogous calculations would go through.

In this section, we characterize all local spin chains that are equivalent to HXXZ under conjugation by any Uchain that is built from a basic two-site Clifford transformation UC2. We first summarize the results for locality-preserving transformations in Sec. III A and then summarize the results for locality-non-preserving transformations in Sec. III B. We, then, prove our claims by classifying all Uchain with k = 2 in Sec. III C. Of course, this classification does not depend on the Hamiltonian, so these results could be easily applied to other canonical models. For example, one could straightforwardly generalize the analysis to the integrable XYZ chain3 simply by allowing independent couplings for each term in (3). Note also that HXXZ is integrable for all ΔR. For Δ > 1 (Δ < −1), the model is in a gapped (anti-)ferromagnetic phase, while for |Δ| < 1, we have a gapless paramagnetic phase.1 If we include an external field term, hnZn, the model remains integrable. The phase diagram in the Δ − h plane has three regions corresponding to the three phases mentioned above for h = 0; the value of Δ for the boundaries of these regions varies with h.

Consider all locality-preserving Uchain built from a two-site Clifford transformation, as defined in (16) for k = 2 [see also Fig. 1(a)]. In Sec. III C, we will show that acting with any such locality-preserving transformation has the effect on HXXZ of either (a) performing an on-site change of basis or (b) transforming HXXZHXXZ, where

HXXZ=nsitesJ1Zn1XnXn+1Zn+2+J2Zn1YnYn+1Zn+2+J3ZnZn+1,
(19)

up to an on-site change of basis. Here, two of the Ji are equal to 1 and the third is equal to Δ. Note that the transformation acts as (5) for this particular choice of the on-site basis and for J3 = Δ.

A sum over four-spin interaction terms of the form Zn−1XnXn+1Zn+2 gives the Hamiltonian of a generalized cluster model.26 Such terms appear in models dual to free-fermion chains63–68 through a standard Jordan–Wigner transformation (see the discussion in Sec. V B). However, applying this Jordan–Wigner transformation to HXXZ will lead to four-fermion interactions for Δ ≠ 0. We also note that a model with these interaction terms appears in the literature as a “twisted” XXZ chain.27,69,70 However, that model is not translation invariant and, moreover, contains additional interactions, meaning that it is not equivalent to HXXZ.

Note that HXXZ has a U(1) symmetry corresponding to rotations about the Z axis, and so this direction is distinguished. After the transformation, the distinguished direction corresponds to the term with Ji = Δ. In all the models above, we preserve integrability if we add an external field term of the form nUchainZnUchain. For example,

H̃XXZ=nsitesΔZn1XnXn+1Zn+2+Zn1YnYn+1Zn+2+ZnZn+1+hZn1YnZn+1
(20)

is unitarily equivalent to HXXZhnZn. To derive this, we conjugate the Hamiltonian by Uchain built from the basic Clifford transformation UC2 given by

X1X1Z2,X2Z2,Z1Z1,Z2Z1Y2.
(21)

1. The folded XXZ chain and pivot Hamiltonians

As an aside, we note an interesting connection between the model HXXZ, the folded XXZ chain,13,45–48 and the idea of pivot Hamiltonians.49,50 The folded XXZ chain is an integrable model with many interesting properties, including Hilbert space fragmentation,71,72 corresponding to exponentially large degeneracies in the spectrum. Pivot Hamiltonians generate SPT entanglers [an example of which is transformation (5)] and have been used to generate webs of dualities between models and to find interesting critical points with enhanced symmetry.49,50 Indeed, the critical points “half-way between” two models often have the pivot Hamiltonian as an additional U(1) symmetry.49 In this section, we show that using two different Ising Hamiltonians as a pivot, starting with HXXZ, we arrive at the model HXXZ. The folded XXZ chain is at the half-way point between these two models, and we see that the two Ising Hamiltonians give two additional U(1) symmetries of the folded XXZ chain. While these symmetries were already known, we hope that this connection will lead to new insights.

Let us define the following commuting charges:

Q1=12nsites(1Zn),Q2=12nsites(1ZnZn+1),Q4=14nsites(Xn+1Xn+2+Yn+1Yn+2)1+ZnZn+3.
(22)

The Hamiltonian, HF, of the folded XXZ chain is given by

HF=Q4+hQ1+ΔQ2.
(23)

The model HF also commutes with the operator Q2=n(1)nZnZn+1.46 Now, considering the XXZ chain (3) with an external field and the transformed chain (19), we have (up to unimportant constants)

4HF=nsitesXnXn+1+YnYn+1+ΔZnZn+1+hZnHXXZ(h)+U±HXXZ(h)U±HXXZ(h),
(24)

where U± is a Clifford transformation that can be written as a Uchain with k = 2 and acts by conjugation as

Xn±Zn1XnZn+1,ZnZn.
(25)

We see that HF is at the half-way point on a path of models interpolating between the usual XXZ model (with an external field) and the model HXXZ discussed above.

To make the connection to pivot Hamiltonians, we can write these choices of U± as U±(π), where

U±(θ)=eiθHpivot±forHpivot+=14nsites(1)nZnZn+1,Hpivot=14nsitesZnZn+1.
(26)

Then, define H±(θ)=U±(θ)H0U±(θ). These are particular examples of the pivot Hamiltonian construction given in Ref. 49. More generally, we define H(θ) in this way for some choice of Hpivot, where there is a further requirement that H(2π) = H(0) and that H(π) is an SPT phase for some symmetry group. An example would be the trivial paramagnet H0 = −∑nXn; then, H+(π) is the SPT cluster model. In our case, we will take H0 = HXXZ, which is a critical model; the analog of the SPT requirement would, then, be that H±(π) is a symmetry-enriched gapless model34 for some symmetry group. Analyzing the symmetry properties of these models73 is beyond our scope; we defer it to future work. Understanding this could lead to new perspectives on the folded XXZ model.

One interesting consequence of making this connection is as follows. We can consider a family of models Hα±=αH±(0)+(1α)H±(π). At the half-way point H1/2±=12H±(0)+H±(π), there is an enhanced symmetry. Indeed, we clearly have [H1/2±,U±(π)]=0 [in our case of interest U±(π) = U±, we have U(2π) = 1, and this is a Z2 symmetry]. If H0 is a trivial paramagnet, in many cases, this is enhanced to a U(1) symmetry generated by Hpivot±, i.e., [H1/2±,Hpivot±]=0. This follows from a general result given in Ref. 49 based on a well-chosen Clifford transformation (in this context also called a generalized Kramers–Wannier duality or the gauging map in quantum information theory) that reveals a Zn symmetry and shows that neutrality of the Hamiltonian under this Zn implies neutrality under the full U(1).

Now, taking H0 = HXXZ, we have that H1/2±=HF. The result given above for trivial H0 no longer applies, but the property that [H1/2±,Hpivot±]=0 in this case corresponds to the commutativity of Q1,Q2,Q2, and Q4. This can be checked straightforwardly in this model and naturally raises the question of how to extend the argument given in Ref. 49 to further families of Hamiltonian H0, including gapless models.

We now consider locality-non-preserving transformations Uchain, again built from a basic Clifford transformation UC2. By considering all locality-non-preserving Uchain with k = 2, we show in Sec. III C that there are three types of non-local Pauli strings that can appear. Two of these types of string will not preserve locality of HXXZ, while the third does—this includes the Kramers–Wannier duality given in (10). Using this transformation, we have that the following Hamiltonian is unitarily equivalent74 to HXXZhnZn:

HXXZ=nsitesZn+ΔXnXn+2Xn1ZnXn+1+hXnXn+1.
(27)

We conclude that the Kramers–Wannier duality is the only locality-non-preserving transformation (in the class under consideration and up to on-site basis changes) that takes HXXZhnZn to a local Hamiltonian. Note that HXXZ is related to the anisotropic next-nearest-neighbor Ising (ANNNI) model.75 Using the Kramers–Wannier duality to map the XXZ chain to this “U(1)-enhanced” ANNNI model is discussed in Ref. 76. The Hamiltonian (27) has also appeared recently in Ref. 13 as an example of a medium-range integrable model (the model that appears is, in fact, the bond-site transformed XYZ chain, rather than XXZ chain; the bond-site transformation corresponds to the Kramers–Wannier transformation used here). There, it was derived as part of a classification of “interaction-round-a-face” models with three-site interaction terms.

We give two other families of locality-non-preserving transformations below. In the Kramers–Wannier case, we have strings of the form XnZnZn+1Zn+2⋯ and, hence, XnXn+1Zn. As mentioned, the other locality-non-preserving transformations have different strings, and these do not cancel when we transform operators such as XnXn+1 that appear in HXXZ. We can, however, use these locality-non-preserving transformations to map between local integrable Hamiltonians. For example, consider HXXZ with h = 0. Then, the transformation given in (39) transforms this model to

HXXZ=nsitesZnXnXn+2+ΔXn1ZnXn+1.
(28)

To prove these results, we now analyze the effect of conjugating once by Uchain on single-site Pauli operators. Using the formula given in Sec. II A, we have that |C2|=11520 and, hence, can write down that many different Uchain. We will show that these fall into seven classes: three (L1, L2, and L3) that are locality-preserving and four (NL1, NL2, NL3, and NL4) that are locality-non-preserving.

Consider any two anti-commuting single-site Pauli operators P,QP, we prove in Sec. III D that the locality-preserving transformations are of the following form:

  • (L1)

    On-site change of basis: Xn → ±Pn, Zn → ±Qn.

  • (L2)

    On-site change of basis and shift by one site to the left:77Xn → ±Pn−1, Zn → ±Qn−1.

  • (L3)

    Decorating transformations25 

Xn±Sn1PnSn+1,Zn±Sn1QnSn+1,whereSniPnQn.
(29)

Transformation (L3) could, of course, be preceded by an on-site change of basis—this will correspond to Uchain given by a different basic Clifford transformation in C2. Then, for example, we can also have

Xn±Sn1PnSn+1,Yn±Sn1QnSn+1,whereSniPnQn.
(30)

Below, we will take as given that one can always do this initial on-site change of basis.

We now consider locality-non-preserving transformations. These all map at least one Pauli operator to a non-local string of Pauli operators. For a finite system, this string will end at the right-most site (this site has index L). As we are interested in transforming local Hamiltonians to local Hamiltonians, the precise details of the boundary terms are not so important: they must drop out in local expressions—we use S′ and T to represent the Pauli operators that are included in this boundary term. We prove in  Appendix B that the locality-non-preserving transformations are of forms (NL1)–(NL4). (Note that there exist transformations Uchain for all choices of anti-commuting P,QP; the operator appearing in the string, S, is then fixed.)

  • (NL1)
    Transformations leading to a string ∏jSn+j,
    Xn(1)an+bSnSn+1Sn+2Sn+3SL1TL,Zn(1)an+cPn1QnSn+1Sn+2Sn+3SL1TL,
    (31)
    where Sn ≔ ±iPnQn and a, b, c ∈ {0, 1}.
  • (NL2)
    Transformations leading to a string ∏jSn+2j,
    Xn(1)x(n)Pn1SnSn+2Sn+4Sn+6SL1TL,Zn(1)z(n)Qn1SnSn+2Sn+4Sn+6SL1TL,
    (32)
    where Sn ≔ ±iPnQn and the oscillatory terms depend on x(n) and z(n) that are discussed in  Appendix B.
  • (NL3)
    Transformations leading to a string given by ∏jSn+3jSn+3j+1,
    Xn(1)x(n)Pn1SnSn+1Sn+3Sn+4Sn+6Sn+7SL1TL,Zn(1)z(n)Qn1SnSn+1Sn+3Sn+4Sn+6Sn+7SL1TL,
    (33)
    where Sn ≔ ±iPnQn and the oscillatory terms depend on x(n) and z(n) that are discussed in  Appendix B.
  • (NL4)
    On-site change of basis and shift by one site to the left, with boundary term and oscillatory factor,
    Xn(1)an+bPn1TL,Zn(1)an+cQn1TL,
    (34)
    where a, b, c ∈ {0, 1}.

As mentioned, S′ and T are the boundary terms—these are unimportant for transforming local Hamiltonians to local Hamiltonians. The operator S′ is either equal to S or to I depending on whether the site L − 1 would appear in the string;78T is not fixed [and, in general, depends on S′, L, and details of the particular Clifford transformation—for example, we show in  Appendix B that one can write down a Clifford transformation, leading to (31) with any choice of TP]. Note also that these are bulk transformations, so the transformation will, in general, act differently on sites at the boundary.

To prove the claim about locality-preserving transformations, let us define79 a two-site Clifford transformation Un,n+1 acting non-trivially on sites n and n + 1 by

XnsXLn(X)Rn+1(X),ZnsZLn(Z)Rn+1(Z)YnsXsZiLn(X)Ln(Z)Rn+1(X)Rn+1(Z).
(35)

L and R are single-site Pauli operators on the left and right sites80 and sX, sZ ∈ {±1}. It is enough to consider how the transformation Uchain acts left to right since that is the direction where Uchain has a non-local causal structure. There are three cases to consider:

  1. First, if R(X)=R(Z)=I, then this is a single-site Clifford transformation and we simply have an on-site change of basis: the transformations of the form (L1).

  2. The second case is where R(X)R(Z) and both are non-trivial, implying R(Z)R(X)R(Z)R(X). Hence, all three single-site Pauli operators appear in the R-terms on site n + 1 in (35). This means that Xn, Yn, and Zn will keep growing into non-local strings of operators as we continue to apply the transformations making up Uchain. The only exception to this is when, in addition, L(X)=L(Z)=I—this does keep bulk operators local, corresponding to an on-site change of basis and a shift by one site to the left.81 Hence, the locality-preserving transformations in this case give the transformations of the form (L2).

  3. The final case is where one of R(X), R(Z), and R(X)R(Z) is equal to identity (and then, the other two must be equal). Let us consider a non-trivial R(X) = R(Z) = R. If R = Y, then we have a locality-preserving Uchain. This follows since R(Y)=I, and so

    Un+1,n+2Un,n+1XnUn,n+1Un+1,n+2=sXUn+1,n+2Ln(X)Yn+1Un+1,n+2=sZLn(X)iLn+1(X)Ln+1(Z).
    (36)
    Furthermore, two-site Clifford transformations making up Uchain act on sites to the right of n + 1, so we see that it transforms Xn to a local operator. An analogous calculation holds for transforming Zn, and so for R = Y, we have a locality-preserving transformation. If RY, then we will have non-local strings by iterating the transformation. To see this explicitly,
    Un+1,n+2Un,n+1XnUn,n+1Un+1,n+2=sXUn+1,n+2Ln(X)Rn+1Un+1,n+2=sXsRLn(X)Ln+1(R)Rn+2,
    (37)
    where further two-site Clifford transformations making up Uchain will continue to act non-trivially. We can make an analogous argument for the case that R(X)=I and also for the case that R(Z)=I. For example, if R(Y) = R(Z) = X, we have locality-preserving transformations, while if R(Y) = R(Z) = Y, we will not.
    In (36), we saw that if R = Y, then the transformation is locality-preserving. We now consider how Uchain acts in this case. First, R = Y means that in (35), we must have L(X)L(Z). We can, then, consider how Un,n+1 in (35) acts on Xn+1 and Zn+1. One can show that,82 up to signs, the images of Xn+1 and Zn+1 are any two distinct elements of {Yn+1,iLn(X)Ln(Z)Xn+1,iLn(X)Ln(Z)Zn+1}. Note that the choice of which operator goes to which operator is simply the choice of the initial basis for our transformation Uchain. Let us define the product S ≔ iL(X)L(Z). We see that, without loss of generality, conjugating by Uchain acts as
    Xn±Sn1Ln(X)Sn+1,Zn±Sn1Ln(Z)Sn+1.
    (38)
    In this way, we reach the locality-preserving transformations (L3).

To prove the claim about locality-non-preserving transformations, we simply have to consider all the cases we excluded in the above discussion. In particular, we need to consider the cases where R(X) = R(Z)Y and where R(X)R(Z) and both are non-trivial. These different cases are analyzed in  Appendix B. Here, we give examples of a basic Clifford transformation that generates each type.

  • (NL1)

    Up to an on-site basis change, this is the Kramers–Wannier duality with basic Clifford transformation (10).

  • (NL2)
    Putting P = X, Q = Y, and S = Z, we have the basic Clifford transformation,
    XnZnXn+1,Xn+1XnZn+1,ZnZnYn+1,Zn+1YnZn+1.
    (39)
    This example has x(n) = z(n) = 0, so the string is not oscillatory.
  • (NL3)
    Putting P = Z, Q = Y, and S = X, we have the basic Clifford transformation,
    XnXnZn+1,Xn+1ZnXn+1,ZnXnYn+1,Zn+1YnXn+1.
    (40)
    This example has x(n) = z(n) = 0, so the string is not oscillatory.
  • (NL4)
    Putting P = X and Q = Z, we have
    XnYnXn+1,Xn+1XnYn+1,ZnYnZn+1,Zn+1ZnYn+1.
    (41)
    Since Yn → −Yn+1, we have an oscillatory term ±(−1)n, and moreover, T = Y.

We note that the results for locality-preserving transformations given above are consistent with the theory of Clifford QCA,40,41 as they should be. However, to prove our results using that theory, one would first need to show that the operators under consideration have images on up to three neighboring sites.

In this section, we give a classification of all translation invariant Clifford transformations on the spin chain that map Xn and Zn to Pauli strings on at most five neighboring sites (i.e., each image is of the form Pm2(1)Pm1(2)Pm(3)Pm+1(4)Pm+2(5)). This is the same mathematical problem as classifying Clifford QCA where the images of Xn and Zn are non-trivial on at most five neighboring sites.83 We, furthermore, show that all these transformations are of the form Uchain for some basic Clifford UC2 or UC3. We, then, use these results to write down a form for all Hamiltonians that are equivalent to HXXZ under translation invariant Clifford group transformations that take Xn and Zn to local operators on at most five sites.

In Sec. III, we started by considering basic Clifford transformations and showed that certain families of Uchain were locality-preserving. An alternative approach is to consider locality-preserving transformations directly. As discussed in Sec. II A, unitary transformations preserve commutation relations. If our unitary is a translation invariant Clifford group transformation, this gives a map XnP and ZnQ, with P and Q being some Pauli strings. Commutators such as [Xn, Xn+m] = 0 ⇔ [P, P(m)] = 0 give us non-trivial constraints on P and Q [recall that P(m) is defined in Sec. II A as a translation of P by m sites]. In  Appendix C, we analyze these constraints in the case that P and Q are Pauli strings on up to five sites. The outcome is that if we map any single-site Pauli operator to a two- or four-site operator, then it must map the other on-site Pauli operators to non-local strings (this follows from the theory of Clifford QCA—see Ref. 83). We also prove the results for locality-preserving transformations given in Subsection IV A.

If we map to a three- or five-site operator, we show below that we have the transformations that we already found using basic Clifford transformations in C2, along with the following additional transformations:

  • (L4)
    Decorating transformation,
    Xn±An2An1CnAn+1An+2,Zn±An2An1CnAn+1An+2,
    (42)
    with A,C,CP\{I}, CC′, and A = iCC′.
  • (L5)
    Decorating transformation,
    Xn±An2CnAn+2,Zn±An2CnAn+2,
    (43)
    with A,C,CP\{I}, CC′, and A = iCC′.
  • (L6)
    Decorating transformation,
    Xn±An2Bn1AnBn+1An+2,Zn±An2Cn1CnCn+1An+2,
    (44)
    with A,B,CP\{I}, AB, and C = iAB.

As above, we can always precede these transformations by an on-site change of basis, equivalent to taking a different Uchain. Furthermore, one can show that (L4)–(L6) can all be derived using Uchain built from UC3. This is demonstrated84 in Table I. Finally, note that by considering only the constraints from translation invariance and the commutation relations, we can include an arbitrary translation (or shift operator). We have fixed this translation so that the images are symmetric about site n—transformations Uchain do not include such an arbitrary translation.

TABLE I.

Transformations Uchain of the type shown in Fig. 1 that take HXXZ to Hi by Hi=ŨiHXXZŨi.

Basic Clifford(U12 or U123)Clifford group transformation (Uchain)
Ũ1 X1X1Z2 X2Z1Y2  XnZn−1YnZn+1 
Z1Z1 Z2Z2  ZnZn 
Ũ2 X1Y1Z2Z3 X2Z1Y2 X3Z1Y3 Xn → −Zn−2Zn−1YnZn+1Zn+2 
Z1Z1 Z2Z2 Z3Z3 ZnZn 
Ũ3 X1X1Z2Z3 X2Z1X2Z3 X3Z1Z2X3 XnZn−2XnZn+2 
Z1Z1 Z2Z2 Z3Z3 ZnZn 
Ũ4 X1X1X2X3 X2Z1Y2X3 X3Z1X2Y3 Xn → −Zn−2Yn−1YnYn+1Zn+2 
Z1Z1 Z2X2 Z3X3 Zn → −Zn−1XnZn+1 
Basic Clifford(U12 or U123)Clifford group transformation (Uchain)
Ũ1 X1X1Z2 X2Z1Y2  XnZn−1YnZn+1 
Z1Z1 Z2Z2  ZnZn 
Ũ2 X1Y1Z2Z3 X2Z1Y2 X3Z1Y3 Xn → −Zn−2Zn−1YnZn+1Zn+2 
Z1Z1 Z2Z2 Z3Z3 ZnZn 
Ũ3 X1X1Z2Z3 X2Z1X2Z3 X3Z1Z2X3 XnZn−2XnZn+2 
Z1Z1 Z2Z2 Z3Z3 ZnZn 
Ũ4 X1X1X2X3 X2Z1Y2X3 X3Z1X2Y3 Xn → −Zn−2Yn−1YnYn+1Zn+2 
Z1Z1 Z2X2 Z3X3 Zn → −Zn−1XnZn+1 

We can summarize some of the above results as follows: We have that the spin chains,

H1=nsitesJ1Zn1XnXn+1Zn+2+J2Zn1YnYn+1Zn+2+J3ZnZn+1,
(45)
H2=nsitesJ1Zn2XnXn+1Zn+3+J2Zn2YnYn+1Zn+3+J3ZnZn+1,
(46)
H3=nsitesJ1Zn2Zn1XnXn+1Zn+2Zn+3+J2Zn2Zn1YnYn+1Zn+2Zn+3+J3ZnZn+1,
(47)
H4=nsitesJ1Zn2Xn1Xn+2Zn+3+J2Zn2Yn1YnYn+1Yn+2Zn+3+J3Zn1YnYn+1Zn+2,
(48)

where two of the Ji are equal to 1 and the third is equal to Δ, are equivalent to the XXZ chain under Clifford group transformations. Furthermore, these are all unitarily related to HXXZ by a transformation of the form given in Fig. 1. Moreover, up to an on-site change of basis, this is a complete list of Hamiltonians that are equivalent to HXXZ under translation invariant Clifford group transformations that take Xn and Zn to local operators on at most five sites—in Table I, we give the relevant transformations for J3 = Δ. For the cases with J1 or J2 = Δ, we do an on-site change of basis to HXXZ and then use a transformation from Table I.

It is simple to generate further models by applying a sequence of such transformations (including single-site unitaries). This is illustrated in Fig. 3(a) for the case of two transformations UchainUchain. The transformations Ũi defined in Table I are such that Hi=ŨiHXXZŨi. Then, H0=Ũ4Ũ2HXXZŨ2Ũ4 is the Hamiltonian given in (1). In  Appendix A, we use this unitary transformation to understand the nearby phases to H0 and to show that it is at a phase transition between SPT phases.

FIG. 3.

(a) Applying Uchain followed by Uchain generates further transformations. (b) A depth-four local unitary quantum circuit; Fig. 1(a) is of this form for U given in (49). (c) An MPU tensor corresponding to (b) (note that we block neighboring spins, so the physical dimension is four). (d) Reordering Uchain built from U that commute on different pairs of sites. (e) A depth-eight circuit that has the same effect as transforming with Ũ4.

FIG. 3.

(a) Applying Uchain followed by Uchain generates further transformations. (b) A depth-four local unitary quantum circuit; Fig. 1(a) is of this form for U given in (49). (c) An MPU tensor corresponding to (b) (note that we block neighboring spins, so the physical dimension is four). (d) Reordering Uchain built from U that commute on different pairs of sites. (e) A depth-eight circuit that has the same effect as transforming with Ũ4.

Close modal

Note that the two-site Clifford U12,

X1X1Y2,X2±Y1X2,Z1Z1Y2,Z2±Y1Z2,
(49)

satisfies [Un,n+1,Un+1,n+2]=0 and the corresponding transformation Uchain takes Xn → ±Yn−1XnYn+1 and Zn → ±Yn−1ZnYn+1. Reordering the U as in Fig. 3(d), we, then, have a depth-two local unitary quantum circuit for Uchain—illustrated in Fig. 3(b) (on setting U=U=I). By using on-site basis transformations beforehand and afterward, we can, then, do any of the transformations (L3). This means that for any decorating transformation of the form (L3), we have a depth-four local unitary quantum circuit, as shown in Fig. 3(b). (Note that translation invariance is retained due to the commutativity of the Un,n+1.) This tells us, in particular, that the transformation Ũ1 appearing in Table I can be written in this way.

We can similarly find local unitary circuits of depth at most eight for the remaining transformations of Table I. One can check that the basic Clifford transformations that appear in Ũ2 and Ũ3 commute when applied to different sites. This means that, by analogy with Fig. 3(d), Ũ2 and Ũ3 can be written as depth-three local unitary quantum circuits. The basic Clifford transformation appearing in Ũ4 does not commute when applied to different sites; however, note that applying two consecutive two-site decorating transformations gives the same effect. In particular, let V1 take XnZn−1YnZn+1, YnZn−1XnZn+1 and V2 take XnZn−1XnZn+1, ZnZn−1YnZn+1. Then, conjugation by Ũ4 is the same as conjugation by V2V1. This means that Ũ4 can be written as a local unitary circuit of depth at most eight—this is illustrated in Fig. 3(e), where U′, U″ and V′, V″ are the appropriate basis changes for the decorating transformations V1 and V2, respectively.

This leads to the following conclusion: any Uchain that takes each Xn and each Zn to a local operator on at most five sites can be written as a local unitary quantum circuit of depth at most eight. (We include on-site changes of basis before and after applying transformations in Table I. For Ũ4, we can change U′ and V″ to give the correct basis.) Moreover, in this form, one can see how to explicitly make this transformation periodic. Then, as in going from Fig. 3(b) to Fig. 3(c), one can find an MPU tensor. Having such an MPU tensor is helpful in understanding how the integrable structure of the XXZ model transforms, as we will see in Sec. V.

Thus far, we have considered translation invariant Clifford transformations and used them to construct families of Hamiltonians that are unitarily equivalent to HXXZ. The XXZ model, defined in (3), is an integrable model, and this integrability is related to some underlying structure. It is intuitive that this integrable structure will be roughly unchanged under Clifford transformations or, indeed, other unitary transformations. It is the purpose of this section to begin to address in what sense the transformed XXZ chains constructed above share this structure and to see how this work fits more generally into the classification of integrable models. We will not comment on the precise meaning of the term quantum integrable model; rather, we refer the interested reader to the discussion in Ref. 5. Pragmatically, if one knows something useful about the states of some Hamiltonian, the ideas introduced above can be used to understand some other Hamiltonian. Two classes about which we can say a lot are free models and models solvable by the Bethe ansatz, such as the XXZ model. Free models are those where there is no quasiparticle scattering, while Bethe ansatz solvable models have the property that many-quasiparticle scattering can be factorized into two-body scattering.

We will first analyze how conjugation by a general MPU affects the integrable structure of a Bethe ansatz solvable model. Then, we will explain what this means for the Clifford transformations that we consider. Finally, we will discuss free models.

1. Generalities

The algebraic Bethe ansatz is a well-developed method that allows us, by solving a system of “Bethe equations,” to find the eigenstates and eigenvalues of an integrable Hamiltonian. It is a further non-trivial step to derive correlation functions in these eigenstates; however, there are many known results.1–3,6,28 We point out that certain results apply at finite system sizes, while others apply only on taking the thermodynamic limit. The central object in this method is the R-matrix. This is an operator that solves the Yang–Baxter equation, depicted graphically85 in Fig. 4(a). Using such a solution to the Yang–Baxter equation, one can generate a set of local conserved charges, and one of these is the Hamiltonian of the corresponding fundamental integrable model. There is no prescription to go from an integrable Hamiltonian (which may or may not be fundamental) to the R-matrix1 although there are interesting tests for integrability of spin chain Hamiltonians.13,15 The XXZ spin chain is a Hamiltonian that is the fundamental model derived from the XXZ R-matrix and can be analyzed using the algebraic Bethe ansatz.

FIG. 4.

(a) Yang–Baxter equation for R-matrices. (b) Yang–Baxter algebra for L-operators.

FIG. 4.

(a) Yang–Baxter equation for R-matrices. (b) Yang–Baxter algebra for L-operators.

Close modal

2. The fixed-point of an MPU

Before proceeding, we briefly review some further technical results about MPUs that are needed in the following discussion. These results and corresponding proofs are found in Refs. 58 and 59. All of the results are given diagrammatically in Fig. 5.

FIG. 5.

(a) Blocking an MPO; the two shaded tensors V are combined (by contraction of a virtual index) to give a single tensor W. (b) and (c) Definition of the vectors ⟨l| and |r⟩ as left- and right-eigenvectors of a matrix formed from the tensors V and V̄; d is the dimension of the physical index. (d) The first MPU fixed-point equation. (e) The second MPU fixed-point equation. (f) and (g) The “pulling through conditions,” a consequence of the fixed-point equations.

FIG. 5.

(a) Blocking an MPO; the two shaded tensors V are combined (by contraction of a virtual index) to give a single tensor W. (b) and (c) Definition of the vectors ⟨l| and |r⟩ as left- and right-eigenvectors of a matrix formed from the tensors V and V̄; d is the dimension of the physical index. (d) The first MPU fixed-point equation. (e) The second MPU fixed-point equation. (f) and (g) The “pulling through conditions,” a consequence of the fixed-point equations.

Close modal

We can combine tensors by blocking physical indices; this simply involves viewing multiple physical sites as one physical site with a larger on-site dimension. This is illustrated in Fig. 5(a); if i1 and i2 correspond to a two-dimensional spin-1/2 site, then ĩ1 will be a four-dimensional index corresponding to the tensor product. An important result is that if we have an MPU with bond dimension (dimension of the virtual index) equal to χ, then by blocking at most χ4 sites, we have a fixed-point (or “simple”) form for the MPU.58,59

Now, let us suppose that we have blocked sites such that the MPU is in the fixed-point form. From the corresponding blocked tensors V and V̄, we can form a “transfer matrix.” This transfer matrix has just one non-zero eigenvalue, and this is equal to one.58 Then, we have two vectors ⟨l| and |r⟩, the unique left- and right-eigenvectors86 with eigenvalue one—see Figs. 5(c) and 5(d). The blocked tensor V satisfies two fixed-point equations, given in Figs. 5(d) and 5(e). These equations imply two further equations called pulling through conditions, given in Figs. 5(f) and 5(g).

3. Integrable models transformed by an MPU

Recall that in Sec. IV C we showed that for a given locality-preserving Uchain given in Table I, we can find an MPU tensor V such that conjugating by the MPU Mj1,,jLi1,,iL=tr(Vj1i1VjLiL) gives the same transformation as conjugating by Uchain. In the discussion below, we show that, after blocking sites, if we conjugate our integrable spin chain by any MPU, then given the R-matrix for the original integrable model, we can find another R-matrix depending on the MPU that satisfies the Yang–Baxter equation. The new R-matrix contains the original R-matrix acting on a subspace of an enlarged virtual Hilbert space—see Fig. 6. On the other part of the enlarged space, there is a projection I|rl|I constructed from the eigenvectors of the MPU transfer matrix described in Sec. II. This projection means that this new R-matrix does not satisfy a technical condition called regularity and is, moreover, not invertible. However, we will show below that one can still deduce certain features of the model using this modified R-matrix,87 including the conserved charges. Thus, we see that the underlying integrability is modified in a relatively straightforward way depending on the original R-matrix and the MPU. The Hamiltonian itself is, in fact, more closely related to another object called the L-operator, introduced below. We will see that the L-operator is acted on by the operators V that form the MPU, illustrated in Fig. 7.

FIG. 6.

(a) The R-matrix in the enlarged virtual space (the virtual space of the original R-matrix and an additional virtual space). (b) The Yang–Baxter equation in the additional virtual space. This identity holds for any vectors |r⟩ and ⟨l| and implies that R(λ,μ) satisfies the Yang–Baxter equation if R(λ,μ) does. Note that this is not a regular solution of the Yang–Baxter equation.

FIG. 6.

(a) The R-matrix in the enlarged virtual space (the virtual space of the original R-matrix and an additional virtual space). (b) The Yang–Baxter equation in the additional virtual space. This identity holds for any vectors |r⟩ and ⟨l| and implies that R(λ,μ) satisfies the Yang–Baxter equation if R(λ,μ) does. Note that this is not a regular solution of the Yang–Baxter equation.

Close modal
FIG. 7.

The monodromy matrix (a) before and (b) after conjugation by an MPU M constructed from tensors (Vji)βα. In (b), the shaded tensor on the left is T(λ), while the shaded tensor on the right is L(λ).

FIG. 7.

The monodromy matrix (a) before and (b) after conjugation by an MPU M constructed from tensors (Vji)βα. In (b), the shaded tensor on the left is T(λ), while the shaded tensor on the right is L(λ).

Close modal

4. Classification of integrable models

In order to explore these ideas in greater depth, let us review some further details about the algebraic Bethe ansatz. As mentioned, along with the R-matrix, we have the L-operator. The L-operator satisfies the Yang–Baxter algebra depicted in Fig. 4(b). Given an L-operator that satisfies this algebra, one can generate further L-operators satisfying the algebra. In this way, one generates the Hamiltonian and other conserved charges.6 

Now, the fundamental solution to the Yang–Baxter algebra is given by a “rotated” R-matrix. In particular, if we identify the horizontal (virtual) Hilbert space and the vertical (physical) Hilbert space, then identifying the upper left and lower right indices of the R-matrix as physical indices (rotating the picture by π/4) gives an L-operator that solves Fig. 4(b). This is, however, not a unique solution to the Yang–Baxter algebra.

In general, the L-operator is a tensor acting on a single site, with two physical indices and two uncontracted virtual indices. From L-operators, L(λ), one can form the monodromy matrix,88 

T(λ)βα=α2,,αLL1(λ)α2αL2(λ)α3α2LL(λ)βαL.
(50)

This is an MPO on L sites—depicted in Fig. 7(a)—and is a tensor formed by matrix multiplication of the virtual indices of the L operators. T(λ) acts as a matrix on physical indices and has two uncontracted virtual indices: by taking the trace, we reach the transfer matrix T(λ), where T(λ) has physical indices only. Transfer matrices at different values of the spectral parameter commute, [T(λ), T(μ)] = 0, and this means that the transfer matrix is the generating function of the conserved charges. In particular, the canonical Hamiltonian of the integrable model is the logarithmic derivative of the transfer matrix at λ = 0.1 

Now, let us address the question of classifying integrable models. One approach is to classify solutions to the Yang–Baxter equation, or R-matrices, of some type.9 For example, the case corresponding to nearest-neighbor Hamiltonians was studied recently in Refs. 11 and 12 for on-site Hilbert space C2 and C4, respectively (although it remains to apply Bethe ansatz techniques to find the spectrum and eigenstates of some of these examples). By the discussion above, this would classify fundamental integrable models. In those works, there is discussion of the fact that there are whole classes of equivalent spin chains that can be found by applying local basis transformations to the R-matrix—this is equivalent to applying single-site basis transformations to the fundamental spin chain Hamiltonian. In Ref. 10, the following analogy with representation theory is drawn: fixing a particular R-matrix is similar to fixing a particular group, while enumerating all monodromy matrices for a given R-matrix is analogous to constructing representations of the group. Representations are reducible, in general, so one can try to find the analog of irreducible representations—these are the L-operators studied in Ref. 10.

In Figs. 6 and 7(b), we demonstrate the action of conjugation by an MPU on the various operators. First, in our conjugated model, we can define a new L-operator, L, that includes the V and V̄ tensors that constitute the MPU [see the shaded tensor in Fig. 7(b)]. The dimension of the physical indices of L (corresponding to the dimension of the site of the spin chain) is unchanged, while the bond dimension of L is the bond dimension of L multiplied by χ2, where χ is the bond dimension of the tensor V. We also define a blocked operator L̃, where we block k physical sites so that the MPU reaches its fixed point. In Fig. 8, we show that L̃ satisfies a modified Yang–Baxter algebra, with a modified R-matrix: R; this is itself defined in Fig. 6. In Fig. 9, we show that despite this modification of the algebra, we can still use it to show the commutativity of the corresponding transfer matrices.

FIG. 8.

A variant of the Yang–Baxter algebra for the R matrix. Physical sites are blocked (k times) so that the MPU satisfies the fixed-point equations. We can interpret the left-hand side equaling the right-hand side as the usual Yang–Baxter algebra holding whenever we contract virtual indices with additional (sufficiently blocked) MPU tensors such that we can use the pulling through condition. Indeed, the first equality is an application of the pulling through condition for the MPU, the second equality follows from the usual Yang–Baxter algebra for the original R-matrix, and then the third equality uses the pulling through condition again to give the analog of the Yang–Baxter algebra.

FIG. 8.

A variant of the Yang–Baxter algebra for the R matrix. Physical sites are blocked (k times) so that the MPU satisfies the fixed-point equations. We can interpret the left-hand side equaling the right-hand side as the usual Yang–Baxter algebra holding whenever we contract virtual indices with additional (sufficiently blocked) MPU tensors such that we can use the pulling through condition. Indeed, the first equality is an application of the pulling through condition for the MPU, the second equality follows from the usual Yang–Baxter algebra for the original R-matrix, and then the third equality uses the pulling through condition again to give the analog of the Yang–Baxter algebra.

Close modal
FIG. 9.

A graphical proof that the transfer matrices commute for different values of the spectral parameter. The dashed lines indicate tracing over the virtual indices. In writing the first equality, we assume that the (original) R-matrix is invertible, and then, we use the fixed-point MPU equations to insert the projector |r⟩⟨l|. The modified Yang–Baxter algebra and the fixed-point MPU equation give the second equality.

FIG. 9.

A graphical proof that the transfer matrices commute for different values of the spectral parameter. The dashed lines indicate tracing over the virtual indices. In writing the first equality, we assume that the (original) R-matrix is invertible, and then, we use the fixed-point MPU equations to insert the projector |r⟩⟨l|. The modified Yang–Baxter algebra and the fixed-point MPU equation give the second equality.

Close modal

To summarize, we have shown that by applying MPU transformations to integrable models, we find a modified form of the underlying integrability. We have derived an underlying R-matrix, where the dependence on the spectral parameter is through the R-matrix of the original model only. This R-matrix of the transformed model is not regular and can be written as a tensor product of the original (assumed regular) R-matrix and a trivial (but not regular) solution of the Yang–Baxter equation that is itself derived from the MPU. Moreover, we find a transformed L-operator that satisfies a modified Yang–Baxter algebra. These transformed models fall outside the usual equivalence under basis transformations to the indices of the R-matrix; indeed, this must be the case since the Hamiltonian is no longer nearest neighbor, in general. We, hence, conclude that the essential features of the underlying integrability are unchanged, but our transformed models are not equivalent to the original model in the usual classification scheme. It would be most interesting to take this analysis further, for example, by showing the existence of an R-matrix and L-operator that give the transformed Hamilton and satisfy the unmodified Yang–Baxter algebra. One interesting point is whether such an L-operator would have a blocked physical index, and if so, how this blocking relates to the range of the transformed Hamiltonian.

5. Relation to translation invariant Clifford transformations

We argued above that the Clifford transformations Uchain in Table I be written in an MPU form. This means that we can directly use the above analysis to see that the integrable spin chains in Sec. IV B correspond to Bethe ansatz solvable models. These models have an R-matrix that is simply related to the XXZ R-matrix, and the L-operator is not fundamental, but is related to the fundamental L-operator of the XXZ chain by conjugating with the appropriate MPU tensor. These operators do not satisfy the usual Yang–Baxter algebra, but a variant of this algebra that, nevertheless, allows us to prove commutativity of transfer matrices in the usual way.

The locality-non-preserving transformations do not fit directly into this analysis, as we do not have an MPU form. However, we can consider the system on open boundaries. Then, the formalism of the algebraic Bethe ansatz is more involved, but can still be described in terms of tensor networks.6 Moreover, the locality-non-preserving transformations that we consider can be written as unitary MPOs. It would be interesting to generalize the above analysis to the open boundary condition case to better understand the locality-non-preserving transformations discussed in Sec. III B. We also note that the L-operator and (implicitly) the R-matrix for the bond-site transformed XYZ model [see also model (27)] are given in Ref. 13. It would be interesting to connect their results with the methods of this paper.

We now turn to free models. In this section, we introduce a standard spin chain that can be analyzed by transforming to a free-fermion model. Then, we point out a construction, using Clifford group transformations, of further spin chain Hamiltonians that can be solved by free fermions.

Now, focusing on translation invariant spin-1/2 chains, if we can find a transformation into fermionic modes such that the resulting Hamiltonian is quadratic, then this is a free-fermion chain.8,89 An example is the Ising model,

HIsing=12nsiteshZnXnXn+1,
(51)

and the Jordan–Wigner transformation takes us to a quadratic fermion model. In particular, define

γn=m=1n1ZmXn,γ̃n=m=1n1ZmYn,
(52)

where cn=(γn+iγ̃n)/2 is a canonical spinless fermion at site n. Then, HIsing=i2nγ̃nγn+1+hγ̃nγn. This is a nearest neighbor model. An interesting class of longer-range spin chains are the generalized cluster models,26,63,64

Hfree=12nsitest0Znt1XnXn+1t2XnZn+1Xn+2t3XnZn+1Zn+2Xn+3t1YnYn+1t2YnZn+1Yn+2t3YnZn+1Zn+2Yn+3,
(53)

where we can extend to longer range terms by including longer strings of Z operators. Using the same Jordan–Wigner transformation (52), this can be mapped into a quadratic fermion Hamiltonian with terms coupling sites beyond nearest neighbor: Hfermion=i/2n,αtαγ̃nγn+α. This can be solved similarly to the Ising model, giving the spectrum and eigenstates. Moreover, one can derive many results for the phase diagram, correlations, and topological edge modes.66,68,90,91 Note that these models have fermionic couplings only of the form iγ̃nγm; more generally, one could also include couplings iγ̃nγ̃m.

Now, one can apply Clifford group transformations Uchain to these spin chains and the transformed model will remain solvable by transforming to free fermions, and moreover, we retain all the mentioned results (appropriately transformed). The transformation Ũ4 of Table I, for example, leads to the following transformations of terms from (53) up to four-sites (one- and two-site terms appear already in Sec. IV B):

XnZn+1Zn+2Xn+3Zn2Yn1XnZn+1Zn+2Xn+3Yn+4Zn+5,XnZn+1Xn+2Zn2Yn1YnXn+1Yn+2Yn+3Zn+4,YnZn+1Yn+2Zn2Xn1ZnXn+1Zn+2Xn+3Zn+4,YnZn+1Zn+2Yn+3Zn2Xn1Xn+4Zn+5.
(54)

One way to approach these free-fermion models is to transform the definitions of the Majorana fermions (52),

γn=m=1n1(UZmU)(UXnU),γ̃n=m=1n1(UZmU)(UYnU),
(55)

and then insert these definitions into the Hamiltonian Hfermion. A related approach is explored in Ref. 17—there, the starting point is any two Pauli operators that obey the commutation algebra of the Ising model terms Zn and XnXn+1 and then building a Jordan–Wigner transformation from them. By applying Uchain to the operators Zn and XnXn+1, we preserve this algebra, so our results give a different perspective on that work. We also mention that the authors of Refs. 19, 20, and 24 discussed the question of whether a particular spin chain Hamiltonian can be transformed into a free-fermion form. In these papers, the commutation relations between the terms in the Hamiltonian define a graph, and then, graph-theoretic conditions for the existence of a free-fermion solution are given (see also earlier work on bond algebras92). Note that this graph depends on the basis of Pauli operators with which we expand the Hamiltonian. As pointed out in Ref. 19, this graph is invariant under Clifford transformations, and so given a Hamiltonian of the form (15), the transformations discussed in this paper map between families of spin chains with the same underlying graph.

In this paper, we have explained how to generate new families of integrable Hamiltonians using translation invariant Clifford transformations. As an example, we classified certain families of transformed XXZ models. These families were derived from a classification of translation invariant Clifford transformations that are built from UC2 and a classification of translation invariant Clifford transformations that take single-site Pauli operators to local operators on at most five sites. Finally, we put these results in context by discussing how they relate to usual classifications of integrable Hamiltonians.

This is, however, far from a complete picture. We do not know how to determine whether or not two given Hamiltonians are related by a Clifford transformation. Understanding how to do this is an avenue of future research. As mentioned, there do exist interesting (conjectural) tests for integrability of spin chains.13,15 These conjectures are based on the existence of a conserved charge. The models we have considered that are related to the XXZ chain will naturally have a conserved charge corresponding to a certain operator in the XXZ chain. However, the second conjecture of Ref. 15 relies on deriving this charge from a “boost operator,” and since the models in Sec. IV B are not nearest neighbor, the transformed boost operator will not be of the standard form given there. A generalization to the case of medium-range models was given recently in Ref. 13, and it would be interesting to understand whether our models satisfy the conjectures given in that paper. Furthermore, there are unresolved questions regarding the underlying integrable structures for the transformed models considered above. Finally, as discussed above, recent papers19,20,24 analyze the question of when a spin chain can be solved by transforming to a free-fermion model. It would certainly be of interest to find related results for integrable models beyond free fermions, such as those we have considered in this paper.

In terms of applications, we have already pointed out the appearance of Clifford transformations in the literature on SPT phases and have given a discussion of H0 as a transition between SPT phases. As another example, one can take the critical Ising model,

HIsing=nsitesZnZn+1+Xn,
(56)

and using transformation (5), one can map it to

HCluster-Ising=nsitesZnZn+1+Zn1XnZn+1.
(57)

One notable feature of this model is that its low energy physics is described by a symmetry-enriched Ising CFT.34 The symmetry in this case is Z2×Z2T, generated by the spin-flip ∏nXn and an antiunitary symmetry T that acts as complex-conjugation (in particular, Xn and Zn are real, while Yn is imaginary). In general, Clifford transformations will not preserve the symmetry of the Hamiltonian. A key point is that for a Clifford transformation on the whole chain that takes us from one symmetric Hamiltonian to another, the basic Clifford transformation that we use to construct it need not be symmetric. This property allows us to map between distinct symmetric phases as in the example above. Our methods give a general family of transformations that we could use to explore mapping standard integrable models to symmetry-enriched examples in a similar way. Conversely, given a Hamiltonian and a certain choice of symmetry group, one could classify the Clifford transformations considered above according to how they affect these symmetry properties. In Sec. I, we made connections to a recent study49 on SPT entanglers and pivot Hamiltonians. It would be interesting to understand more deeply these connections. For example, understanding the integrability properties and symmetry properties of pivoted integrable models would be most interesting. As mentioned, this could give new perspectives on the folded XXZ model.

In this work, we focus on integrable spin chains, including spin chains that can be solved by mapping into free-fermion chains. One may be directly interested in fermionic integrable models, for example, the Hubbard chain.2,93 In that case, it would be natural to consider generalizing our discussion above to include fermionic transformations. This would include Majorana translations γnγn+1 that do not have an analog in the spin chain.94 Moreover, in the fermionic case, the equivalence between MPUs and QCA that we discuss above no longer holds.95 

Finally, while we have focused our discussion on integrable models, the Clifford transformations Uchain could, of course, be applied to any Hamiltonian. A motivation for restricting to integrable models was that we could use the many results about integrable models to analyze the resulting transformed model. For general models, it may still be of interest to understand which families of models are equivalent under the transformations that we have analyzed here.

We thank R. Verresen for a number of stimulating conversations and for insightful comments on the manuscript. We are also grateful to P. Fendley, S. Piddock, L. Piroli, B. Pozsgay, D. Shepherd, and R. Thorngren for helpful discussions and correspondence. We thank Y. Miao for valuable discussions and for identifying an issue with Sec. V A of the first version of this manuscript. N.L. acknowledges support from the UK Engineering and Physical Sciences Research Council through Grant Nos. EP/R043957/1, EP/S005021/1, and EP/T001062/1.

The authors have no conflicts to disclose.

Nick G. Jones: Conceptualization (equal); Formal analysis (equal); Writing – original draft (lead); Writing – review & editing (equal). Noah Linden: Conceptualization (equal); Formal analysis (equal); Writing – original draft (supporting); Writing – review & editing (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

In this appendix, we will discuss nearby phases of the integrable model given in (1), which, for convenience, we repeat here, allowing for arbitrary (real) coupling constants,

H0=nsitesJ1Zn3Yn2Xn1YnYn+1Xn+2Yn+3Zn+4+J2Zn3Yn2Yn1Yn+2Yn+3Zn+4+J3Zn1YnYn+1Zn+2.
(A1)

Recall also that H0=U0HXYZU0, where U0=Ũ4Ũ2 are defined in Table I and HXYZ is given by

HXYZ=nsitesJ1YnYn+1+J2XnXn+1+J3ZnZn+1.
(A2)

For J1 = J2 = 1, J3 = Δ, this is the XXZ model.

Let us first discuss HXYZ setting J3 = 0 for convenience: then, we have the XY model.90 This model has a Z2×Z2 symmetry generated by PX = ∏nXn and PY = ∏nYn (we also have that PZ = ∏nZn = PXPY is a symmetry, where we assume that the chain has length a multiple of four, for convenience) and an anti-unitary Z2T time-reversal symmetry, T, that corresponds to complex conjugation. Now, for J1 > J2, the ground state spontaneously breaks PX and T with order parameter ⟨Yn⟩. Similarly, for J2 > J1, the ground state spontaneously breaks PY with order parameter ⟨Xn⟩. More generally, for |J3| < J2 < J1, the nearby phases are the same.96 Hence, the XXZ model is a transition between an X-ordered ferromagnetic phase and a Y-ordered ferromagnetic phase.

We can make the same analysis for H0. This model has the same symmetries as HXYZ. First, for J1 > J2 > |J3|, by applying the unitary U0, we have symmetry-breaking with order parameter ⟨Zn−3Xn−2YnXn+2Zn+3⟩. Hence, the ground state spontaneously breaks PX and T but preserves PY and PZT. In the fixed-point limit J1 → ∞, we have the following:

|Zn3Xn2YnXn+2Zn+3|=1|Z1Y2Y3X4n=5MYnXM+1YM+2YM+3ZM+4|=1.
(A3)

Hence, this correlator is a string-order parameter97 for the symmetry string PY, and the end-point operator (XM+1YM+2YM+3ZM+4) has a non-trivial charge under the unbroken symmetry PZT. This means that it is a non-trivial SPT phase, and this is a stable property for all J1 > J2 > |J3|.

Second, for J2 > J1 > |J3|, we have symmetry-breaking with order parameter

Zn3Xn2Zn1ZnZn+1Xn+2Zn+3.
(A4)

The ground state spontaneously breaks PX and PY but preserves PZ and T. Again, considering the fixed-point limit, we have that

|Zn3Xn2Zn1ZnZn+1Xn+2Zn+3|=1|Z1Y2X3Y4X5n=6MZnXM+1YM+2XM+3YM+4ZM+5|=1,
(A5)

giving a string order parameter for the phase. The end-point operator (XM+1YM+2XM+3YM+4ZM+5) has trivial charge under the unbroken symmetry PZT. Hence, we see that H0 (J1 = J2 > |J3|) is at a transition between two different SPT phases.

Note that for the XXZ chain, the analogous string orders are simply PX and PY, which have trivial end-points, and hence, no non-trivial charges appear.

Here, we prove the claims about the different locality-non-preserving transformations arising from two-site Clifford transformations, given in Sec. III C. Recall that we use the following notation for images of Xn and Zn when conjugating by Un,n+1:

XnsXLn(X)Rn+1(X),ZnsZLn(Z)Rn+1(Z)YnsXsZiLn(X)Ln(Z)Rn+1(X)Rn+1(Z),
(B1)

where sX, sZ ∈ {±1}. Above, we considered the case R(X) = R(Z) = R, which gave rise to local transformations for R = Y. If, instead, R = X or R = Z, then we reach a transformation of type (NL1) given in (31). More precisely, suppose R = X. Then, we must have (up to signs)

XnSnXn+1,P̃n+1PnYn+1,ZnQnXn+1,Q̃n+1PnZn+1,
(B2)

where P and Q are anti-commuting single-site Pauli operators and S ≔ iPQ. P̃ and Q̃ are also anti-commuting single-site Pauli operators. By applying Uchain, we can fix them to be P̃=Y and Q̃=Z by an initial choice of on-site basis. The string comes from repeated transformations of X. We see that under conjugation by Uchain, we have (31) with boundary term T = X. Now, suppose that in (B2) we take Xn → −SnXn−1. Then, we will have an oscillatory sign, depending on site-index: (−1)n. Other signs in (B2) will change overall signs, but not give this oscillatory behavior, which comes from repeated transformations of X. Note that if we take R = Z, we would have T = Z, while, for completeness, the two-site Clifford,

YnSnYn+1,Yn+1PnXn+1,ZnQnYn+1,Zn+1PnZn+1,
(B3)

will give the same transformation (31), but with T = Y.

We, then, need to analyze the cases where R(X)R(Z) and both are non-trivial, and so neither are equal to R(Y) = iR(X)R(Z). Now, if L(X)=L(Z)=I, then Uchain is a swap transformation (up to an on-site change of basis) that is local in the bulk. Without loss of generality, we, then, have L(X)=L(Z)I. Let us first consider L(X) = L(Z) = R(Y). This implies that the basic Clifford acts as

XnRn(Y)Rn+1(X),P̃n+1Rn(X)Rn+1(Y),ZnRn(Y)Rn+1(Z),Q̃n+1Rn(Z)Rn+1(Y),
(B4)

where P̃Q̃P\{I}. We, then, have three different cases that can be understood by98 

(NL2)XnsXZnXn+1,ZnsZZnYn+1,
(B5)
(NL3)XnsXXnZn+1,ZnsZXnYn+1,
(B6)
(NL4)XnsXYnXn+1,ZnsZYnZn+1.
(B7)

The choice of signs sX, sZ gives different oscillatory terms (see Table II). The three different cases correspond to the remaining transformations given in Sec. III C. In particular, (B5) leads to (32), (B6) leads to (33), and (B7) leads to (34). The oscillatory terms are determined by explicit calculation.

TABLE II.

Oscillatory terms that appear in non-local transformations depending on the signs (sX, sZ). For the image of Xn (Zn), the oscillation will be (−1)x(n) [(−1)z(n)], where x(n) [z(n)] is the appropriate entry. We have b ∈ {0, 1}, where b may be different for x(n) and z(n) and is fixed by the other signs in the basic Clifford transformation.

(sX, sZ(1, 1) (1, −1) (−1, 1) (−1, −1) 
Transformation (B5n/2⌋ + b n/2⌋ + n + b b n + b 
Transformation (B6b n/3⌋ + n + b n/3⌋ + ⌊2n/3⌋ + n + b ⌊2n/3⌋ + b 
Transformation (B7b n + b n + b b 
(sX, sZ(1, 1) (1, −1) (−1, 1) (−1, −1) 
Transformation (B5n/2⌋ + b n/2⌋ + n + b b n + b 
Transformation (B6b n/3⌋ + n + b n/3⌋ + ⌊2n/3⌋ + n + b ⌊2n/3⌋ + b 
Transformation (B7b n + b n + b b 

Finally, we consider the other possibility, which is without loss of generality, that L(X) = L(Z) = R(X). Then,

XnRn(X)Rn+1(X),P̃n+1Rn(Y)Rn+1(Y),ZnRn(X)Rn+1(Z),Q̃n+1Rn(Z)Rn+1(Y).
(B8)

This is similar to the previous case. The string depends on the choices of R(X) and R(Z), and we are again led to the three cases (NL2)–(NL4). One can take, for example,

(NL2)XnsXXnXn+1,ZnsZXnYn+1,
(B9)
(NL3)XnsXZnZn+1,ZnsZZnYn+1,
(B10)
(NL4)XnsXXnXn+1,ZnsZXnZn+1.
(B11)

The oscillatory factors for these transformations are also given in Table II—the relevant entry depends on the corresponding transformation [(B5)–(B7)] according to the type of the non-local string [(NL2)–(NL4)].

Note that this discussion shows that for all choices of basic Clifford transformation that is not one of the locality-preserving cases, we have a transformation of form (NL1)–(NL4). We also have all transformations of this form (ignoring boundary terms) since we can apply an on-site basis change after the transformation—this is equivalent to taking a different basic Clifford transformation.

In this appendix, we prove certain results about translation invariant Clifford transformations. As explained above, the results for two- and four-site images follow directly from results in the literature on Clifford QCA; we find it helpful to have a self-contained presentation here. In particular, we see explicitly the general rule that the images of Xn and Zn must be symmetric under reflection about a site.40,41

1. Two-site

Suppose that we have a translation invariant Clifford group transformation that maps Xn to a two-site Pauli operator, without loss of generality, say, sites n − 1 and n. Then, we must have that Xn → ±An−1An and that Zn±jQj(j) where this string is non-local.

To show this, let Xn → ±An−1Bn. Then, since [Xn, Xn+1] = 0, we see that [An, Bn] = 0. Since we have a genuine two-site Pauli operator, we must have that An = Bn, i.e., Xi → ±An−1An. Now, consider ZnjQj(j); this must anticommute with An−1An; hence, exactly one of Q(n−1) and Q(n) must be a non-trivial Pauli, different from A. Let us say that Q(n) anticommutes with A. Now,

[Zn1,Xn]=0[Qn1(n)Qn(n+1),An1An]=0,
(C1)

and so we must have that Q(n+1) anticommutes with A. By repeating this reasoning, we end up with a non-local string of non-trivial Pauli operators to the right. If we, instead, have that Q(n−1) anticommutes with A, we would get a non-local string to the left.

2. Three-site

Suppose that we have a translation invariant Clifford group transformation that takes

XnAn1BnCn+1,ZnAm1BmCm+1.
(C2)

Since [Xn, Xn+2] = 0, we must have [C, A] = 0. Hence, either XnAn−1BnAn+1 or XnBn. Similarly, [Zn, Zn+2] = 0 means ZnAm1BmAm+1 or ZnBm.

  • If XnBn and ZnBm, then clearly m = n and BB′.

  • If XnAn−1BnAn+1 and ZnBm, then m = n. (Since the image of Zn anticommutes with the image of Xn, we must have n − 1 ≤ mn + 1. If, say, m = n + 1, then since the image of Zn anticommutes with the image of Xn, it would also anticommute with the image of Xn+2, which is a contradiction.) Then, we have B′ ≠ B so that XnAn−1BnAn+1, ZnBn, and Yn±An1BnAn+1. Since [Yn+1, Xn] = 0, we must have that AB and AB′.

  • If XnAn−1BnAn+1 and ZnAm1BmAm+2, then again m = n. Any other option would mean that the image of Zn and the image of Zn both anticommute with Xn, where n′ ∈ {n ± 2, n ± 3}. Then, [Zn+2, Xn] = 0 implies that A′ = A. This is now the same as the previous case with Z and Y swapped.

Additional signs on the right-hand side of (C2) are analyzed in the same way. Hence, we can conclude that the most general translation invariant Clifford group transformation that takes us to strings of length at most three is of the form

Pn±Sn1PnSn+1,Qn±Sn1QnSn+1,Sn=iPnQn.
(C3)

3. Four-site

Suppose that we have a translation invariant Clifford group transformation that takes

XnAn1BnCn+1Dn+2.
(C4)

Then, we must have that

XnAn1BnBn+1An+2,
(C5)

and the image of Zn is infinite.

To derive the possible images of Xn, note that [Xn, Xn+3] = 0 means that A = D. Then, [Xn, Xn+1] = [Xn, Xn+2] = 0 gives us that B = C. To, then, show that the image of Zn is infinite, there are two cases to consider, [B, A] = 0 and {B, A} = 0. Let us consider the first case. Then, either B = A or B=I. If XnAn−1An+2, then we can proceed as in the two-site case. The image of Zn must contain a Pauli operator either on site n − 1 or on site n + 2 that anticommutes with A. Then, since the image of Zn commutes with Am−1Am+2 for all mn, the image of Zn must be an infinite string to either the left or the right. If XnAn−1AnAn+1An+2, letting Zn=Pn1(n1)Pn(n)Pn+1(n+1)Pn+2(n+2), we have that either one or three of {P(n−1), …, P(n+2)} anticommute with A. In the case that only P(n−1) (P(n+2)) anticommutes with A, then the image of Zn is infinite to the left (right); otherwise, it is infinite in both directions. This follows easily from the commutativity with images of Xm for mn.

For the case {B, A} = 0, we use a different approach. In that case,

XnAn1BnBn+1An+2.
(C6)

Let us call our four-site Clifford transformation U4; we know from Appendix C 2 that we can find a translation invariant Clifford group transformation U3 such that AnAn and BnAn−1BnAn+1. Now, conjugating by U3U4 will take XnBnBn+1 and will take ZnU3U4ZnU4U3. Since U3 maps single-site operators to at most three-site operators, if U4ZnU4 is a finite string, then U3U4 contradicts the result of Appendix C 1. Hence, the image of Zn must be an infinite string.

4. Five-site

Suppose that we have a translation invariant Clifford group transformation that takes

XnAn2Bn1CnDn+1En+2,ZnAm2Bm1CmDm+1Em+2.
(C7)

Let us first consider the case where none of these operators are the identity. Then, [Xn, Xn+4] = [Zn, Zn+4] = 0 gives A = E and A′ = E′. That is,

XnAn2Bn1CnDn+1An+2,ZnAm2Bm1CmDm+1Am+2.
(C8)

Now, [Xn, Xn+3] = 0 tell us that either B = D = A or that both B and D are distinct from A. In the second case, [Xn, Xn+2] = 0 gives us that B = D. The same considerations apply to the image of Z. Since both images are symmetric about the middle site, if mn, we have an inconsistency: if m = n + k, then the image of Zn−2k will also anticommute with the image of Xn. Hence, we have that the image of Xn is either An−2Bn−1CnBn+1An+2 or An−2An−1CnAn+1An+2, while the image of Zn is either An2Bn1CnBn+1An+2 or An2An1CnAn+1An+2. In all cases, [Xn, Zn+4] = 0 implies A′ = A. Considering also [Xn, Zn+3] = 0, [Xn, Zn+2] = 0, and {Xn, Zn} = 0 gives us two options.

  • The first case is
    XnAn2An1CnAn+1An+2,ZnAn2An1CnAn+1An+2,
    (C9)
    where CnCnandAn=iCnCn.
  • The second case is
    XnAn2Bn1CnBn+1An+2,ZnAn2Bn1CnBn+1An+2,
    (C10)
    with A,B,C,CP such that BA, B′ ≠ A, CC′. Now, if B′ = B, then not only CA and C′ ≠ A but also CB and C′ ≠ B, which is inconsistent. Hence, B′ ≠ B, and either C = A or C′ = A. If C = A, then C′ = B′. Otherwise, if C′ = A, then C = B.

An example of this case is

XnXn2Zn1ZnZn+1Xn+2,ZnXn2Yn1XnYn+1Xn+2,YnXn1YnXn2.
(C11)

Now, let us consider the five-site case where we allow some operators to be the identity. Due to Appendixes C 1 and C 3, we can take that Xn maps to a five-site, three-site, or one-site operator. First, let

XnAn2Bn1CnDn+1An+2,

where AI. We can immediately exclude cases where each of B,C,D{I,A} since, similarly to Appendix C 3, one could, then, argue that the image of Zn is infinite. Since the images of different Xn commute, one can show that there are the following possibilities that include at least one identity and one operator not equal to A:

XnAn2CnAn+2,CA,
(C12)
XnAn2Bn1Bn+1An+2,BA.
(C13)

Let U5 be any transformation that acts as (C13), and using U3 of Appendix C 3, we have that XiBn−1Bn+1 under conjugation by U3U5. Clearly, any image of Zn that anticommutes with that image and commutes with the images of all other Xm will be infinite, and so the image of Zn under U5 must be infinite. We can, hence, ignore transformations of the form (C13).

For the other possibility, (C12), we can find images of Zn over five sites. First, the image of Zn must be either one-site, three-site, or five-site according to Appendixes C 1 and C 3. If the image of Zn is one site, the only consistent possibility is ZnAn. If the image of Zn is three-site, then we must have ZnAm1BmAm+1. Then, the form of (C12) means that the only consistent choice is A=I. Finally, if the image of Zn is five site, we have ZnAm2Bm1CmBm+1Am+2. Since only one image of Zn can anticommute with (C12), we must have A′ = A and that B{A,I}. Then, by symmetry, we must have m = n with {C′, C} = 0 and C′ ≠ A. By considering [Xn±1, Zn] = 0, we have that B=I. Note that in this final case, Yn → ±An, so all five-site possibilities involving identity reduce to the same case.

For completeness, we also consider the cases where the image of Xn is a three-site or a one-site operator. Excluding cases treated in earlier appendices, we may assume that the image of Zn is a five-site operator. We now show that all such cases reduce to previously considered five-site transformations up to some relabeling on the left-hand side. If XnPn, a one-site operator, then it is easy to see from the commutation relations that ZnAn−2Bn−1CnBn+1An+2, where CA and where [B, A] = 0. These possibilities have already appeared above. If the image of Xn is a three-site operator, then we have either XnAn−1An+1 or XnAn−1BnAn+1. The first option leads to an infinite image of Zn. The second option must have BA. Otherwise, again, we have an infinite image of Zn. In particular, we have that AB, A,BI, and ZnAn2Bn1CnBn+1An+2. We have already seen that B=IB=I. If BI, then in all consistent cases, we reach an example of the form (C10).

In conclusion, we have proved that all locality-preserving translation-invariant Clifford transformations that take single-site Pauli operators to at most five-site Pauli operators must be of the form (L1)–(L6) given in Secs. III and IV.

1.
F.
Franchini
,
An Introduction to Integrable Techniques for One-Dimensional Quantum Systems
(
Springer
,
2017
), Vol. 940.
2.
V. E.
Korepin
,
N. M.
Bogoliubov
, and
A. G.
Izergin
,
Quantum Inverse Scattering Method and Correlation Functions
(
Cambridge University Press
,
1997
).
3.
R. J.
Baxter
,
Exactly Solved Models in Statistical Mechanics
(
Elsevier Science
,
2016
).
4.
X.-W.
Guan
,
M. T.
Batchelor
, and
C.
Lee
, “
Fermi gases in one dimension: From Bethe ansatz to experiments
,”
Rev. Mod. Phys.
85
,
1633
1691
(
2013
).
5.
J.-S.
Caux
and
J.
Mossel
, “
Remarks on the notion of quantum integrability
,”
J. Stat. Mech.: Theory Exp.
2011
,
P02023
.
6.
V.
Murg
,
V. E.
Korepin
, and
F.
Verstraete
, “
Algebraic Bethe ansatz and tensor networks
,”
Phys. Rev. B
86
,
045125
(
2012
).
7.
S.
Sachdev
,
Quantum Phase Transitions
, 2nd ed. (
Cambridge University Press
,
2011
).
8.
E.
Lieb
,
T.
Schultz
, and
D.
Mattis
, “
Two soluble models of an antiferromagnetic chain
,”
Ann. Phys.
16
,
407
466
(
1961
).
9.
P. P.
Kulish
and
E. K.
Sklyanin
, “
Solutions of the Yang-Baxter equation
,”
J. Sov. Math.
19
,
1596
1620
(
1982
).
10.
A. G.
Izergin
and
V. E.
Korepin
, “
The most general L operator for the R-matrix of the XXX model
,”
Lett. Math. Phys.
8
,
259
265
(
1984
).
11.
M.
de Leeuw
,
A.
Pribytok
, and
P.
Ryan
, “
Classifying integrable spin-1/2 chains with nearest neighbour interactions
,”
J. Phys. A: Math. Theor.
52
,
505201
(
2019
).
12.
M.
de Leeuw
,
A.
Pribytok
,
A. L.
Retore
, and
P.
Ryan
, “
New integrable 1D models of superconductivity
,”
J. Phys. A: Math. Theor.
53
,
385201
(
2020
).
13.
T.
Gombor
and
B.
Pozsgay
, “
Integrable spin chains and cellular automata with medium-range interaction
,”
Phys. Rev. E
104
,
054123
(
2021
).
14.
P.
Ryan
, “
Integrable systems, separation of variables and the Yang-Baxter equation
,” arXiv:2201.12057 [math-ph] (
2022
).
15.
M. P.
Grabowski
and
P.
Mathieu
, “
Integrability test for spin chains
,”
J. Phys. A: Math. Theor.
28
,
4777
4798
(
1995
).
16.
K.
Minami
, “
Solvable Hamiltonians and fermionization transformations obtained from operators satisfying specific commutation relations
,”
J. Phys. Soc. Jpn.
85
,
024003
(
2016
).
17.
K.
Minami
, “
Infinite number of solvable generalizations of XY-chain, with cluster state, and with central charge c = m/2
,”
Nucl. Phys. B
925
,
144
160
(
2017
).
18.
P.
Fendley
, “
Free fermions in disguise
,”
J. Phys. A: Math. Theor.
52
,
335002
(
2019
).
19.
A.
Chapman
and
S. T.
Flammia
, “
Characterization of solvable spin models via graph invariants
,”
Quantum
4
,
278
(
2020
).
20.
M.
Ogura
,
Y.
Imamura
,
N.
Kameyama
,
K.
Minami
, and
M.
Sato
, “
Geometric criterion for solvability of lattice spin systems
,”
Phys. Rev. B
102
,
245118
(
2020
).
21.
Y.
Yanagihara
and
K.
Minami
, “
Exact solution of a cluster model with next-nearest-neighbor interaction
,”
Prog. Theor. Exp. Phys.
2020
,
113A01
.
22.
F. C.
Alcaraz
and
R. A.
Pimenta
, “
Free fermionic and parafermionic quantum spin chains with multispin interactions
,”
Phys. Rev. B
102
,
121101(R)
(
2020
).
23.
F. C.
Alcaraz
and
R. A.
Pimenta
, “
Integrable quantum spin chains with free fermionic and parafermionic spectrum
,”
Phys. Rev. B
102
,
235170
(
2020
).
24.
S. J.
Elman
,
A.
Chapman
, and
S. T.
Flammia
, “
Free fermions behind the disguise
,”
Commun. Math. Phys.
388
,
969
1003
(
2021
).
25.
X.
Chen
,
Y. M.
Lu
, and
A.
Vishwanath
, “
Symmetry-protected topological phases from decorated domain walls
,”
Nat. Commun.
5
,
3507
(
2014
).
26.
R.
Verresen
,
R.
Moessner
, and
F.
Pollmann
, “
One-dimensional symmetry protected topological phases and their transitions
,”
Phys. Rev. B
96
,
165124
(
2017
).
27.
T.
Scaffidi
,
D. E.
Parker
, and
R.
Vasseur
, “
Gapless symmetry-protected topological order
,”
Phys. Rev. X
7
,
041048
(
2017
).
28.
N. M.
Bogoliubov
,
A. G.
Izergin
, and
V. E.
Korepin
, “
Critical exponents for integrable models
,”
Nucl. Phys. B
275
,
687
705
(
1986
).
29.
P.
di Francesco
,
P.
Mathieu
, and
D.
Sénéchal
,
Conformal Field Theory
, Graduate Texts in Contemporary Physics (
Springer
,
New York
,
1999
).
30.
F.
Pollmann
,
A. M.
Turner
,
E.
Berg
, and
M.
Oshikawa
, “
Entanglement spectrum of a topological phase in one dimension
,”
Phys. Rev. B
81
,
064439
(
2010
).
31.
X.
Chen
,
Z.-C.
Gu
, and
X.-G.
Wen
, “
Classification of gapped symmetric phases in one-dimensional spin systems
,”
Phys. Rev. B
83
,
035107
(
2011
).
32.
N.
Schuch
,
D.
Pérez-García
, and
I.
Cirac
, “
Classifying quantum phases using matrix product states and projected entangled pair states
,”
Phys. Rev. B
84
,
165139
(
2011
).
33.
F.
Pollmann
, “
Symmetry-protected topological phases in one-dimensional systems
,” in
Topological Aspects of Condensed Matter Physics: Lecture Notes of the Les Houches Summer School: Volume 103, August 2014
, edited by
C.
Chamon
,
M. O.
Goerbig
,
R.
Moessner
, and
L. F.
Cugliandolo
(
Oxford University Press
,
2017
).
34.
R.
Verresen
,
R.
Thorngren
,
N. G.
Jones
, and
F.
Pollmann
, “
Gapless topological phases and symmetry-enriched quantum criticality
,”
Phys. Rev. X
11
,
041059
(
2021
).
35.

We ignore any boundary conditions here. In fact, accounting for the boundary in this transformation is one way to understand the edge mode characteristic of SPT phases in the cluster model.26 

36.
M.
Ozols
, “
Clifford group
” (unpublished) (
2008
); accessed 3 April 2020.
37.
D.
Gottesman
, “
Stabilizer codes and quantum error correction
,” Ph.D. thesis,
California Institute of Technology
,
1997
.
38.
H. A.
Kramers
and
G. H.
Wannier
, “
Statistics of the two-dimensional ferromagnet. Part I
,”
Phys. Rev.
60
,
252
262
(
1941
).
39.
D.
Aasen
,
R. S. K.
Mong
, and
P.
Fendley
, “
Topological defects on the lattice: I. The Ising model
,”
J. Phys. A: Math. Theor.
49
,
354001
(
2016
).
40.
B.
Schumacher
and
R. F.
Werner
, “
Reversible quantum cellular automata
,” arXiv:quant-ph/0405174 (
2004
).
41.
D.-M.
Schlingemann
,
H.
Vogts
, and
R. F.
Werner
, “
On the structure of Clifford quantum cellular automata
,”
J. Math. Phys.
49
,
112104
(
2008
).
42.
T.
Farrelly
, “
A review of quantum cellular automata
,”
Quantum
4
,
368
(
2020
).
43.
F.
Verstraete
,
J. J.
García-Ripoll
, and
J. I.
Cirac
, “
Matrix product density operators: Simulation of finite-temperature and dissipative systems
,”
Phys. Rev. Lett.
93
,
207204
(
2004
).
44.
B.
Zeng
,
X.
Chen
,
D.-L.
Zhou
, and
X.-G.
Wen
,
Quantum Information Meets Quantum Matter
(
Springer
,
2019
).
45.
Z.-C.
Yang
,
F.
Liu
,
A. V.
Gorshkov
, and
T.
Iadecola
, “
Hilbert-space fragmentation from strict confinement
,”
Phys. Rev. Lett.
124
,
207602
(
2020
).
46.
L.
Zadnik
and
M.
Fagotti
, “
The folded spin-1/2 XXZ model: I. Diagonalisation, jamming, and ground state properties
,”
SciPost Phys. Core
4
,
10
(
2021
).
47.
L.
Zadnik
,
K.
Bidzhiev
, and
M.
Fagotti
, “
The folded spin-1/2 XXZ model: II. Thermodynamics and hydrodynamics with a minimal set of charges
,”
SciPost Phys.
10
,
99
(
2021
).
48.
B.
Pozsgay
,
T.
Gombor
,
A.
Hutsalyuk
,
Y.
Jiang
,
L.
Pristyák
, and
E.
Vernier
, “
Integrable spin chain with Hilbert space fragmentation and solvable real-time dynamics
,”
Phys. Rev. E
104
,
044106
(
2021
).
49.
N.
Tantivasadakarn
,
R.
Thorngren
,
A.
Vishwanath
, and
R.
Verresen
, “
Pivot Hamiltonians as generators of symmetry and entanglement
,” arXiv:2110.07599 [cond-mat.str-el] (
2021
).
50.
N.
Tantivasadakarn
,
R.
Thorngren
,
A.
Vishwanath
, and
R.
Verresen
, “
Building models of topological quantum criticality from pivot Hamiltonians
,” arXiv:2110.09512 [cond-mat.str-el] (
2021
).
51.

If we know how each of these operators transform, we can use Uj=1NPjU=j=1N(UPjU), linearity, and that ZjXj = iYj to find the action on any element of ±PN*.

52.
J.
Ignacio Cirac
,
D.
Pérez-García
,
N.
Schuch
, and
F.
Verstraete
, “
Matrix product states and projected entangled pair states: Concepts, symmetries, theorems
,”
Rev. Mod. Phys.
93
,
045003
(
2021
).
53.
C.
Schön
,
E.
Solano
,
F.
Verstraete
,
J. I.
Cirac
, and
M. M.
Wolf
, “
Sequential generation of entangled multiqubit states
,”
Phys. Rev. Lett.
95
,
110503
(
2005
).
54.
A.
Smith
,
B.
Jobst
,
A. G.
Green
, and
F.
Pollmann
, “
Crossing a topological phase transition with a quantum computer
,”
Phys. Rev. Res.
4
,
L022020
(
2022
).
55.
F.
Barratt
,
J.
Dborin
,
M.
Bal
,
V.
Stojevic
,
F.
Pollmann
, and
A. G.
Green
, “
Parallel quantum simulation of large systems on small NISQ computers
,”
npj Quantum Inf.
7
,
79
(
2021
).
56.

Note that this need not coincide with the dimension of the physical sites.

57.

In fact, Uchain with k = 2 is the product of two χ = 2 MPOs. The left-moving shift operator, an MPO Mshift with χ = 2 that takes site n + 1 to site n,52 and a χ = 2 MPO MU constructed from U combine to give Uchain = MUMshift.

58.
J.
Ignacio Cirac
,
D.
Perez-Garcia
,
N.
Schuch
, and
F.
Verstraete
, “
Matrix product unitaries: Structure, symmetries, and topological invariants
,”
J. Stat. Mech.: Theory Exp.
2017
,
083105
.
59.
M.
Burak Şahinoğlu
,
S. K.
Shukla
,
F.
Bi
, and
X.
Chen
, “
Matrix product representation of locality preserving unitaries
,”
Phys. Rev. B
98
,
245122
(
2018
).
60.
L.
Piroli
and
J. I.
Cirac
, “
Quantum cellular automata, tensor networks, and area laws
,”
Phys. Rev. Lett.
125
,
190402
(
2020
).
61.

For a detailed discussion of the relationship between tensor networks and QCA, see Ref. 60.

62.

This basic Clifford transformation is the same as (10) followed by an on-site change of basis on site n: XnZn, ZnXn. In Uchain, this means we do this basis change on every site apart from the right-most site.

63.
M.
Suzuki
, “
Relationship among exactly soluble models of critical phenomena. I*): 2D Ising model, dimer problem and the generalized XY-model
,”
Prog. Theor. Phys.
46
,
1337
(
1971
).
64.
J. P.
Keating
and
F.
Mezzadri
, “
Random matrix theory and entanglement in quantum spin chains
,”
Commun. Math. Phys.
252
,
543
579
(
2004
).
65.
P.
Smacchia
,
L.
Amico
,
P.
Facchi
,
R.
Fazio
,
G.
Florio
,
S.
Pascazio
, and
V.
Vedral
, “
Statistical mechanics of the cluster Ising model
,”
Phys. Rev. A
84
,
022304
(
2011
).
66.
W.
DeGottardi
,
M.
Thakurathi
,
S.
Vishveshwara
, and
D.
Sen
, “
Majorana fermions in superconducting wires: Effects of long-range hopping, broken time-reversal symmetry, and potential landscapes
,”
Phys. Rev. B
88
,
165111
(
2013
).
67.
T.
Ohta
,
S.
Tanaka
,
I.
Danshita
, and
K.
Totsuka
, “
Topological and dynamical properties of a generalized cluster model in one dimension
,”
Phys. Rev. B
93
,
165423
(
2016
).
68.
R.
Verresen
,
N. G.
Jones
, and
F.
Pollmann
, “
Topology and edge modes in quantum critical chains
,”
Phys. Rev. Lett.
120
,
057001
(
2018
).
69.
D. E.
Parker
,
T.
Scaffidi
, and
R.
Vasseur
, “
Topological Luttinger liquids from decorated domain walls
,”
Phys. Rev. B
97
,
165114
(
2018
).
70.
C. M.
Duque
,
H.-Y.
Hu
,
Y.-Z.
You
,
V.
Khemani
,
R.
Verresen
, and
R.
Vasseur
, “
Topological and symmetry-enriched random quantum critical points
,”
Phys. Rev. B
103
,
L100207
(
2021
).
71.
P.
Sala
,
T.
Rakovszky
,
R.
Verresen
,
M.
Knap
, and
F.
Pollmann
, “
Ergodicity breaking arising from Hilbert space fragmentation in dipole-conserving Hamiltonians
,”
Phys. Rev. X
10
,
011047
(
2020
).
72.
V.
Khemani
,
M.
Hermele
, and
R.
Nandkishore
, “
Localization from Hilbert space shattering: From theory to physical realizations
,”
Phys. Rev. B
101
,
174204
(
2020
).
73.

For external field h = 0, there is a Z2×Z2 symmetry generated by nXn and nYn. There is also a Z2T antiunitary time-reversal symmetry.

74.

This is the bulk form of the Hamiltonian; we are not including any boundary terms.

75.
I.
Peschel
and
V. J.
Emery
, “
Calculation of spin correlations in two-dimensional Ising systems from one-dimensional kinetic models
,”
Z. Phys. B: Condens. Matter
43
,
241
249
(
1981
).
76.
R.
Verresen
,
A.
Vishwanath
, and
F.
Pollmann
, “
Stable Luttinger liquids and emergent U(1) symmetry in constrained quantum chains
,” arXiv:1903.09179 [cond-mat.str-el] (
2019
).
77.

Note that there is a corresponding transformation that shifts to the right if we, instead, conjugate by Uchain.

78.

For example, for the string of the form jSn+2j, then S′ = S if (L − 1 − n) is even and S=I if (L − 1 − n) is odd. One can see that S′ must fit the string pattern on site L − 1 since if we increase the length of the chain, then UchainUL,L+1Uchain and UL,L+1 does not act on this site.

79.

This is a partial definition: to fully define the Clifford transformation, we also need the image of Xn+1 and Zn+1.

80.

For example, for the transformation in (9), we have L(X) = X, R(X) = Z, L(Z) = Z, and R(Z)=I.

81.

If we have a periodic system, then this does shift every site one site to the left since this will take P1PL. The corresponding MPU is an on-site change of basis followed by the left-moving shift operator.52 

82.

To see this, consider Xn+1PnQn+1. If P ∈ {L(X), L(Z)}, we are led to an inconsistency since conjugation by Un,n+1 preserves the commutators [Xn, Xn+1] = [Zn, Xn+1] = 0.

83.

Note that this implies that the images of Xn and Zn are restricted to a single set of five neighbouring sites. This follows from the general result for Clifford QCA that the images of Xn and Zn must be symmetric under reflection about a site and that this is the same site for both images41 (we give a derivation for the case under consideration in  Appendix C). This means that the image of Yn is also symmetric about the same site, with support restricted to the support of the images of Xn and Zn.

84.

Note that a change of basis after the transformation is needed to reach the general transformations given above. To find the corresponding basic Clifford transformation, take the transformation from Table I and then transform the Pauli operators on the first site on the right-hand side of the transformation.

85.

The R-matrix acts on virtual Hilbert spaces ViVjViVj and depends on two “spectral parameters” or “rapidities” λ and μ: i.e., it is a tensor of the form R(λ,μ)αi,βjγi,δj, where each index ranges from 1 to χ (the bond dimension). These indices are always contracted in calculating physical quantities. We prefer the graphical notation and refer the reader to Ref. 6 for a detailed exposition of this.

86.

We have l|=n=1χn,n| if we gauge-fix the tensor V such that it is in a canonical form,58 but we will not need the explicit form of these vectors for our purposes.

87.

Note that we do not exclude the possibility that there is some other natural modification of the R-matrix.

88.

Typically, one includes on-site rapidities ξj, i.e., T(λ;ξ1,,ξL)βα=L1(λξ1)α2αL2(λξ2)α3α2LL(λξL)βαL. These are helpful in certain intermediate calculations. Then, one can restore translation invariance by putting ξj = ξ0 at the end.1 For the purposes of our discussion, we suppress these variables. In any case, our intention is to use the transformation to carry over known results rather than to rederive them from scratch.

89.

There are examples of chains that are free-fermion but cannot be solved by a Jordan–Wigner transformation—for discussion, see Refs. 18 and 24.

90.
E.
Barouch
and
B. M.
McCoy
, “
Statistical mechanics of the XY model. II. Spin-correlation functions
,”
Phys. Rev. A
3
,
786
804
(
1971
).
91.
N. G.
Jones
and
R.
Verresen
, “
Asymptotic correlations in gapped and critical topological phases of 1D quantum systems
,”
J. Stat. Phys.
175
,
1164
1213
(
2019
).
92.
Z.
Nussinov
and
G.
Ortiz
, “
Bond algebras and exact solvability of Hamiltonians: Spin S=12 multilayer systems
,”
Phys. Rev. B
79
,
214440
(
2009
).
93.
E. H.
Lieb
and
F. Y.
Wu
, “
Absence of Mott transition in an exact solution of the short-range, one-band model in one dimension
,”
Phys. Rev. Lett.
20
,
1445
1448
(
1968
).
94.
L.
Fidkowski
,
H. C.
Po
,
A. C.
Potter
, and
A.
Vishwanath
, “
Interacting invariants for Floquet phases of fermions in two dimensions
,”
Phys. Rev. B
99
,
085115
(
2019
).
95.
L.
Piroli
,
A.
Turzillo
,
S. K.
Shukla
, and
J. I.
Cirac
, “
Fermionic quantum cellular automata and generalized matrix-product unitaries
,”
J. Stat. Mech.: Theory Exp.
2021
,
013107
.
96.
R. J.
Baxter
, “
One-dimensional anisotropic Heisenberg chain
,”
Ann. Phys.
70
,
323
337
(
1972
).
97.
F.
Pollmann
and
A. M.
Turner
, “
Detection of symmetry-protected topological phases in one dimension
,”
Phys. Rev. B
86
,
125441
(
2012
).
98.

Other transformations fall into these three cases depending on whether the Pauli operator on site n + 1 in the image is the same as the Pauli operator before the transformation. For example, XnZnYn+1 and ZnZnXn+1 is of type (NL3) since the Pauli operator on site n + 1 in each image is different from the initial Pauli operator on site n. XnXnYn+1 and ZnXnZn+1 are of type (NL2) since exactly one Pauli operator (Z) has the same Pauli operator on site n + 1 in its image.