Topologically ordered phases in 2 + 1 dimensions are generally characterized by three mutually related features: fractionalized (anyonic) excitations, topological entanglement entropy, and robust ground state degeneracy that does not require symmetry protection or spontaneous symmetry breaking. Such a degeneracy is known as topological degeneracy and can be usually seen under the periodic boundary condition regardless of the choice of the system sizes L1 and L2 in each direction. In this work, we introduce a family of extensions of the Kitaev toric code to N level spins (N ≥ 2). The model realizes topologically ordered phases or symmetry-protected topological phases depending on the parameters in the model. The most remarkable feature of topologically ordered phases is that the ground state may be unique, depending on L1 and L2, despite that the translation symmetry of the model remains unbroken. Nonetheless, the topological entanglement entropy takes the nontrivial value. We argue that this behavior originates from the nontrivial action of translations permuting anyon species.

In the studies of many-body systems, one is often interested in the properties of ground states and low energy excitations. Ground state degeneracy that does not originate from spontaneous symmetry breaking or fine-tuning of parameters is called topological degeneracy.1–3 Such a degeneracy is robust against any local perturbations, including symmetry-breaking ones. In two dimensions, the order of topological degeneracy Ndeg depends on the genus g of the manifold on which the system is defined.

In topologically ordered phases with U(1) symmetry (e.g., fractional quantum Hall systems), Ndegqg when the filling is ν = 1/q. This degeneracy can be proven by a flux-threading type argument,4 assuming the appearance of fractional excitations with U(1) charge 1/q. More generally, there usually exist closed string operators that describe processes of creating a pair of anyonic excitations, dragging them apart, and pair-annihilating them again after forming a non-contractible loop. These loops commute with the Hamiltonian but not among them. Non-commutativity of loop operators implies the topological degeneracy. In particular, the topological degeneracy Ndeg on a torus (g = 1) is often equal to the number of distinct anyonic excitations. The topological degeneracy is also tied with the topological entanglement entropy,5,6 which is given by Stopo=logD, where D is the total quantum dimension and D2 is nothing but the number of distinct anyonic excitations for Abelian topological order. Therefore, it is often stated that the ground state degeneracy on torus, anyonic excitations, and topological entanglement entropy appear all at the same time.

In this work, we introduce a family of extensions of the Kitaev toric code7,8 to N-level spins (N = 2, 3, 4, …), which contains an integer parameter a (1 ≤ aN). The original model corresponds to the (N, a) = (2, 1) case. The model describes topologically ordered phases when a is not a multiple of rad(N) (the radical of N; see Sec. IV) and phases with no topological order when a is a multiple of rad(N). In particular, when N and a are coprime, these phases are characterized by the topological entanglement entropy Stopo = −log N, independent of system sizes L1, L2.

For a generic integer N, the case of a = 1 is the standard ZN toric code8 discussed widely, for example, in Refs. 912, which shows topological degeneracy Ndeg = N2 regardless of the choice of L1 and L2. The case of a = N − 1 (N ≥ 3) was discussed in Refs. 1316, although the ground state degeneracy on torus was not fully investigated. When N is odd and a = N − 1, we find that the topological degeneracy occurs only when both L1 and L2 are even,
(1)
The most striking situation of our model arises when N is a prime number and a is a primitive root modulo N (see Sec. III A). In this case, Ndeg is given by (see Sec. V A for the proof)
(2)
This means that the minimum system size to observe the degeneracy is L1 = L2 = N − 1, for which the Hilbert space dimension is N2(N1)2 (for example, 11200 for N = 11, for which a = 2, 6, 7, 8 are the primitive roots). It is thus nearly impossible to see the topological degeneracy for a large N in any numerical studies. Therefore, the uniqueness of the ground state for a sequence of Li cannot be used as a proof of the absence of topological order, although the converse might still be the case: topological degeneracy Ndeg > 1 in a sequence of Li implies a nontrivial topological order. Note that, if an open boundary condition is assumed instead of the periodic one, a unique ground state can be realized even in the original Z2 toric code due to the absence of any Wilson loops or constraints among stabilizers.

There is a more famous example, called Wen’s plaquette model,17 in which topological degeneracy depends on the system size. There are also more recent examples of this type behavior.18–22 However, in these examples, the ground state degeneracy on torus is at least two. Our example demonstrates that there are even cases where the ground state is unique and excitations are all gapped in a sequence of Li, despite their nontrivial topological order. It is interesting to contrast with a known theorem about topological quantum field theory (TQFT), according to which the phase is invertible (i.e., no topological order) if Ndeg = 1 on torus (and technically, on sphere as well).23 Our example shows that the relation between lattice models and corresponding effective field theories can be quite subtle. We will also show that the degeneracy can be understood in terms of the TQFT if the finite-size torus in the lattice system is viewed as a torus in continuum but with symmetry defect lines (or twisted boundary conditions) corresponding to the translation symmetry action in the low-energy theory.

The rest of this work is organized as follows. We summarize the definition and basic properties of our model in Sec. II. We review basic mathematical facts in number theory in Sec. III. Overall properties of our model for given integers N and a are summarized in Sec. IV. Then, the ground state degeneracy of the model in topologically ordered phases is studied in Sec. V. The relation of our model to the standard ZN toric code model is clarified in Sec. VI. Topological properties such as the topological entanglement entropy and anyon statistics in our model are discussed in Sec. VII. Finally, we study the cases with no topological order in Sec. VIII. We then conclude in Sec. IX.

In this section, we explain the definition and basic properties of the ZN toric code. Throughout this work, N is an integer greater than 1.

In our model, an N-level spin is placed on each link of square lattice. See Fig. 1 for the illustration. The action of operators X̂r and Ẑr on the N-level spin at r is represented by N-dimensional unitary matrices,
(3)
(4)
(5)
which generalize the Pauli matrices. All matrix elements left blank are zero. They satisfy
(6)
(7)
Operators for different spins commute so that
(8)
(9)
Note that, when N ≥ 3, matrices X and Z are not Hermitian and X̂rX̂r and ẐrẐr.
FIG. 1.

(a) The definition of a vertex operator Âv and a plaquette operator B̂p. (b) Pairs of magnetic excitations created by X̂ra. (c) Pairs of electric excitations created by Ẑra.

FIG. 1.

(a) The definition of a vertex operator Âv and a plaquette operator B̂p. (b) Pairs of magnetic excitations created by X̂ra. (c) Pairs of electric excitations created by Ẑra.

Close modal
The role of X̂r and Ẑr can be interchanged by the global unitary transformation ÛY, whose action on each spin is represented by
(10)
We have UYXUY=Z and UYZUY=X.
The positions of spins on the horizontal and vertical links are set to r=(m1+12,m2) and (m1,m2+12), respectively, where mi = 0, 1, …, Li − 1 (i = 1, 2) and Li is a positive integer,
(11)
We impose the periodic boundary condition and identify X̂r+(n1L1,n2L2) with X̂r and Ẑr+(n1L1,n2L2) with Ẑr for any n1,n2Z and r ∈ Λ. The sets of vertices V and plaquettes P are given by
(12)
(13)
The total number of spins in the system is 2L1L2, and the dimension of the Hilbert space is N2L1L2.
The Hamiltonian of the model reads as
(14)
which is invariant under translation T̂i (i = 1, 2), defined by
(15)
As illustrated in Fig. 1(a), a vertex operator Âv (vV) is defined by
(16)
and a plaquette operator B̂p (pP) is
(17)
Integers a1 and a2 (1 ≤ a1, a2N) are important parameters of this model. It is easy to verify that Âv’s (vV) and B̂p’s (pP) all commute with each other regardless of a1 and a2. For brevity, we set a1 = a2 = a in the following, but a1a2 cases can be treated in the same way.
The eigenstates of the Hamiltonian can be chosen as simultaneous eigenstates of all Âv’s (vV) and B̂p’s (pP). Since
(18)
eigenvalues of operators Âv and B̂p are N-fold, 1, ω, …, ωN−1.
A ground state of the Hamiltonian Ĥ can be constructed explicitly following the discussion for the original toric code, for example, in Ref. 24. Let |ϕ0⟩ be the “ferromagnetic” product state, satisfying
(19)
(20)
It has the eigenvalue +1 for all plaquette operators,
(21)
Now, we introduce a projection operator,
(22)
which satisfies
(23)
(24)
(25)
(26)
Then, the state
(27)
satisfies both
(28)
(29)
suggesting that |Φ0⟩ is a ground state with the energy eigenvalue EGS = −2L1L2. Here, NC > 0 is the normalization factor given by
(30)
It can be shown that NC counts the number of global constraints among the vertex operators of the form
(31)
and there is an equal number of constraints among the plaquette operators,
(32)
In our model, the total number of vertex operators and plaquette operators, 2L1L2, coincides with the total number of spins in the system. Hence, there would be no ground state degeneracy if all stabilizers were independent. Indeed, as we demonstrate in Sec. V, the number of constraints NC is related to the ground state degeneracy as Ndeg=NC2.
The state |Φ0⟩ is translation invariant,
(33)
because |ϕ0⟩ is translation invariant and P̂ commutes with T̂i. As we show below, the model has a nonzero excitation gap. Furthermore, all correlation functions of X̂r and Ẑr are short-ranged. For example,
(34)
These observations imply the absence of translation symmetry breaking.
We introduce (open) string operators X̂p,p(i) (i = 1, 2) by
(35)
(36)
and Ẑv,v(i) (i = 1, 2) by
(37)
(38)
In these expressions, we assumed 0m1m1L11 and 0m2m2L21.
The state
(39)
contains a pair of plaquettes operators with eigenvalues not equal to 1 [see Fig. 1(b) for the illustration], which are called magnetic excitations. In this state, the eigenvalues of B̂(m112,m2+12) and B̂(m1+12,m2+12) are ω and ωam1m1+1, respectively. The eigenvalues of other plaquette operators remain +1. In the derivation of these relations, we used the general property of exponents (zm)n=zmn for zC and m,nZ. Similarly, the state
(40)
has the eigenvalues ωam1m1+1 and ω−1 for the vertex operators Â(m1,m2) and Â(m1+1,m2), respectively. String operators along x2 direction also create pairs of electric or magnetic excitations at their ends [see Fig. 1(c)].
A single plaquette operator or a vertex operator with eigenvalue ωq (q = 1, 2, …, N − 1) costs an energy
(41)
(42)
The excitation energy of a pair Δpair can thus be bounded by
(43)

These electric and magnetic excitations can be further divided into equivalence classes up to local excitations (i.e., excitations that can be created locally), which are called anyon types. They will be discussed in Sec. VII C.

In this section, we review basic mathematical facts in number theory to set up notations for Sec. IV and below.

Given a positive integer n and a positive integer a coprime to n, the multiplicative order of a modulo n is defined as the smallest positive integer such that
(44)
which we denote by Mn(a) in this work. For example, Mn(a) = 1 if and only if a = 1 (mod n). In addition, for n ≥ 3, Mn(a) = 2 if a = −1 (mod n). Conversely, the relation in Eq. (44) implies that n and a are coprime. In the following applications, the integer n is chosen to be N itself or a divisor of N that is coprime to a.

The multiplicative order is related to Euler’s totient function φ(n), which is defined as the number of positive integers smaller than n that are relatively prime to n. By definition, 1 ≤ φ(n) ≤ n − 1. If and only if n is prime, φ(n) = n − 1.

Euler’s theorem,25  aφ(n) = 1 mod n, implies that Mn(a) is a divisor of φ(n). Thus,
(45)
Integers a that saturate the upper bound, i.e., Mn(a) = φ(n), are called the primitive roots modulo n. The primitive roots exist if and only if n is 2, 4, pk, or 2pk, where p is an odd prime number and k is a positive integer. It follows that, when n is a prime number, there exists an integer a such that
(46)
Finally, suppose that n′ is also a positive integer coprime to a. In this case, Mnn(a) is a multiple of both Mn(a) and Mn(a) because aMnn(a)=1 (mod nn′) also implies aMnn(a)=1 (mod n) and aMnn(a)=1 (mod n′). In particular,
(47)
when n″ is a multiple of n.

These mathematical facts underlie our results quoted in Eqs. (1) and (2).

Suppose that the integer N (N ≥ 2) can be prime factorized into
(48)
where pj’s (j = 1, 2, …, n) are prime numbers and rj’s are positive integers. The radical of N is defined as the product of all distinct prime factors of N,
(49)
We denote the set of all (positive) divisors of N by DN,
(50)
which includes 1, N, and rad(N), for example.
Without loss of generality, let us arrange prime factors pj’s of N in Eq. (48) in such a way that
(51)
Then, the largest divisor of N that is coprime to a is given by
(52)
By definition, we have
(53)
(54)
Here, gcd(p, q, r, …) for integers p, q, r, … represents their greatest common divisor. By definition, p/gcd(p, q) and q/gcd(p, q) are positive integers. Since Na is a multiple of any dDN that is coprime to a, MNa(a) is a multiple of Md(a).

Our model describes two distinct types of phases with or without topological order depending on whether a (1 ≤ aN) is a multiple of rad(N) or not. Here, we provide a brief summary of the main features of the two phases.

  1. Case 1: When a is not a multiple of rad(N), our model exhibits topological degeneracy for some sequences of L1 and L2. The ground state degeneracy on the torus is given by
    (55)
    For example, Ndeg=Na2 when both L1 and L2 are multiples of MNa(a) and Ndeg = 1 when L1 and L2 are not simultaneously multiples of Md(a) for any dDN coprime to a, except for d = 1. Correspondingly, there are Na2 species of anyons. This class thus falls into topologically ordered phases. It contains the important class of a being coprime to N. Examples include the cases of a = 1 and a = N − 1 previously discussed in the literature. The size dependence of Ndeg can be understood from the translation symmetry action on the anyon excitations, which will be discussed in Sec. VII D.
  2. Case 2: When a is a multiple of rad(N), the ground state is unique regardless of the choice of L1 and L2. The model thus realizes a trivial phase with regard to topological orders, but it still might be a nontrivial symmetry protected topological phase. As simplest examples, we discuss the cases of N = a and N = a2.

We study these two cases separately in Secs. V and VII and in Sec. VIII.

In this section, we show that, when a is not a multiple of rad(N), the order of ground state degeneracy Ndeg is greater than one for some sequences of L1 and L2.

We start with the simplest case where a is coprime to N.

1. When both L1 and L2 are multiples of MN(a)

Suppose that both L1 and L2 are multiples of MN(a) so that aL1=aL2=1 mod N. In this case, the ground state degeneracy and the low-energy excitations are basically the straightforward extension of the original toric code. For example, when a = 1, MN(a) = 1 and the assumption automatically holds for any L1 and L2. In contrast, when N ≥ 3 and a = N − 1, MN(a) = 2 and both L1 and L2 need to be even.

When both L1 and L2 are multiples of MN(a), there are two sets of global constraints among the stabilizers Âv’s and B̂p’s,
(56)
and
(57)
implying that NC in Eq. (30) is N. In the derivation, we used definitions in Eqs. (16) and (17) and the periodic boundary conditions, such as X̂(L112,m2)=X̂(12,m2) and X̂(m1,L212)=X̂(m1,12). These constraints imply that not all vertex operators and plaquettes operators are independent. For example, the eigenvalues of Âv0 [v0 := (0, 0)] and B̂p0 [p0:=(L112,L212)] are automatically fixed once the eigenvalues of other Âv’s and B̂p’s are chosen.
Correspondingly, there are four independent closed string operators, illustrated in Figs. 2(a) and 2(b), which commute with every term in the Hamiltonian,
(58)
(59)
and
(60)
(61)
where X̂p,p(i) and Ẑv,v(i) are defined in Eqs. (35)–(38). These operators satisfy [see Eq. (9)]
(62)
(63)
and
(64)
(65)
Hence, as the set of independent stabilizers commute with Ĥ, one can choose the following set of operators:
  • The vertex operators Âv (vV, vv0) and the plaquette operators B̂p (pP, pp0). There are in total N2(L1L21) different combinations of eigenvalues.

  • Closed string operators Ẑ(i) (i = 1, 2). There are N2 different combinations of eigenvalues.

FIG. 2.

Illustration of (a) closed string operators X̂(1) and X̂(2), (b) closed string operators Ẑ(1) and Ẑ(2), (c) open string operators that control the eigenvalues of plaquette operators B̂p (pp0), and (d) open string operators that control the eigenvalues of vertex operators Âv (vv0).

FIG. 2.

Illustration of (a) closed string operators X̂(1) and X̂(2), (b) closed string operators Ẑ(1) and Ẑ(2), (c) open string operators that control the eigenvalues of plaquette operators B̂p (pp0), and (d) open string operators that control the eigenvalues of vertex operators Âv (vv0).

Close modal
Starting from the ground state |Φ0⟩ in Eq. (27), which has the eigenvalue +1 for all of these 2L1L2 operators, one can generate all N2L1L2 states in the Hilbert space by using the open string operators illustrated in Figs. 2(c) and 2(d) and the closed loop operators X̂(i) (i = 1, 2). They can be distinguished by N2(L1L21)×N2=N2L1L2 distinct combinations of eigenvalues of these stabilizers. In particular, all degenerate ground states can be written as [X̂(1)]j1[X̂(2)]j2|Φ0 (j1, j2 = 0, 1, …, N − 1), which has the eigenvalue ωji of Ẑ(i). Hence, the order of topological degeneracy is
(66)

The closed loop operators in Eqs. (58) and (59) create a pair of magnetic excitations at xi=±12, dragging the one at xi=12 all the way to xi=Li12=12 and annihilating them in pair. The pair annihilation requires that magnetic excitations with eigenvalues ω and ω−1 meet. This is possible only when Li is a multiple of MN(a).

2. When L1 or L2 is not a multiple of MN(a)

Next, we consider the case where L1 or L2 is not a multiple of MN(a). Without loss of generality, we assume that L1 is not a multiple of MN(a).

Let us introduce a product of string operators associated with the plaquette p=(m1+12,m2+12)P,
(67)
The first factor creates magnetic excitations with eigenvalues ωaL11m1 and ωaL1 at the plaquettes (12,m2+12) and p, respectively. The second factor creates magnetic excitations with eigenvalues ω and ωaL11m1 at the plaquettes p and (L112,m2+12), respectively. Combining these two effects, the operator X̂p(1) create a single magnetic excitation with eigenvalue ω1aL1 at the plaquette p. In fact, X̂p(1) satisfies
(68)
(69)
for any vV and pP. Since L1 is not a multiple of MN(a), ω1aL11.
Similarly, the following operator can be introduced for each vertex v=(m1,m2)V:
(70)
which satisfies
(71)
(72)
for any vV and pP. Hence, Ẑv(1) creates a single electric excitation with eigenvalue ωaL111 at the vertex v.
To proceed, let us assume further that aL11 is coprime to N. In this case,
(73)
are all different and are not equal to 0 mod N. Therefore, the eigenvalue of the plaquette operator B̂p (the vertex operator Âv) can be freely controlled by [X̂p(1)] ([Ẑv(1)]) without affecting others, implying the absence of global constraints involving B̂p or Âv, such as the ones of the form in Eq. (31) (i.e., NC = 1).
Moreover, operators X̂v(1) (vV) and Ẑp(1) (pP) all commute with each other. Hence, starting from the ground state |Φ0⟩ satisfying Eqs. (28) and (29), one can generate all N2L1L2 states in the Hilbert space by successively applying Ẑv(1)’s and X̂p(1)’s. In particular, there is no state other than |Φ0⟩ that has eigenvalue +1 for all vertex operators and plaquette operators. This proves the uniqueness of the ground state
(74)
given that aL11 is coprime to N. This condition is satisfied, for example, (i) when N is prime and L1 is not a multiple of MN(a) (in this case, aL110 mod N) and (ii) when N is odd, a = N − 1, and L1 is not a multiple of MN(a) = 2 (in this case, aL11=2 mod N). This completes the proof of Eqs. (1) and (2).

The gap to the first excited states is given by Δq in Eq. (41), although these states are created by nonlocal operators X̂p(1) and Ẑv(1). Local excitations are still given by pairs of magnetic excitations and electric excitations, for which the exaction gap is bounded by Eq. (43).

Next, we discuss the most general case where a/rad(N)Z but a is not necessarily coprime to N. In this case, we will see that Na in Eq. (52) plays the role of N in the above discussion.

Let us list up all constraints among Âv’s and B̂p’s of the form of Eqs. (31) and (32). We have
(75)
and
(76)
Here, nDN is a parameter specified shortly. In order to set the products in Eqs. (75) and (76) to be 1, we need
(77)
To solve this equation, let us define di,aDN (i = 1, 2) by
(78)
This is the largest divisor d of N such that (i) d is coprime to a and (ii) Li is a multiple of Md(a). For example, di,a = N when a = 1 and di,a = 1 when aLi1 is coprime to N. The smallest positive integer nDN satisfying Eq. (77) is given by
(79)
where
(80)
After all, we find the following set of global constraints:
(81)
(82)
where n′ = 0, 1, 2, …, da − 1, suggesting that NC in Eq. (30) is given by da. These constraints imply that not all vertex operators and plaquettes operators are independent. For example, the eigenvalues of Âv0na and B̂p0na can be automatically fixed once the eigenvalues of other Âv’s and B̂p’s are chosen. Then, as the set of independent stabilizers commuting with Ĥ, one can choose the following set of operators:
  • The vertex operators Âv (vV, vv0) and the plaquette operators B̂p (pP, pp0). There are in total N2(L1L21) different combinations of eigenvalues.

  • The residual free parts of Âv0 and B̂p0. The eigenvalues of these operators can be written as ωx+da ( = 0, 1, …, na − 1), where the value of x (x = 0, 1, …, da) is automatically determined by the constraints in Eqs. (81) and (82). Hence, there are effectively na2 different combinations of eigenvalues.

  • Loop operators [X̂(1)]n1,a and [Ẑ(1)]n1,a (or [X̂(2)]n2,a and [Ẑ(2)]n2,a), where ni,a := N/di,a (i = 1, 2). Their eigenvalues are di,a-fold: ωni,aj (j = 0, 1, …, di,a − 1), which include ωnaj(j=0,1,,da1) as a subset. As detailed below, only da2 different eigenvalues of these operators can be manipulated without affecting the eigenvalues of other stabilizers.

Hence, starting from the ground state |Φ0⟩ in Eq. (27), one can generate all N2L1L2 states in the Hilbert space, which can be distinguished by N2(L1L21)×na2×da2=N2L1L2 distinct combinations of eigenvalues of these stabilizers. This implies that the order of the ground state degeneracy is given by Eq. (55).

It remains to show that the eigenvalues of stabilizers can be manipulated as stated above. Clearly, open string operators illustrated in Figs. 2(c) and 2(d) can be used to control the eigenvalues of Âv (vV, vv0) and B̂p (pP, pp0). The remaining operators satisfy the following algebra:
(83)
(84)
(85)
(86)
where αi (0 ≤ αiN − 1) is defined by
(87)
which is coprime to a and a multiple of di,a = gcd(αi, Na). All these operators commute with Âv (vV, vv0) and B̂p (pP, pp0) and thus do not change their eigenvalues.

1. Case 1: α1 = α2 = 0

When α1 = α2 = 0, N is coprime to a and both L1 and L2 are multiples of MN(a). This case was covered in Sec. V A 1.

2. Case 2: Either α1 = 0 or α2 = 0

Next, we discuss the case when either α1 = 0 or α2 = 0. Without loss of the generality, here, we assume α1 ≠ 0 and α2 = 0. In this case, N is again coprime to a, and we have d1,a = da = gcd(α1, N) and d2,a = N.

Since α1/d1,a is coprime to n1,a = N/d1,a, there exists an integer 1 (1 ≤ 1n1,a − 1) such that
(88)
Then, we can control the eigenvalues of Âv0 and B̂p0 by [X̂(1)]1 and [Ẑ(1)]1,
(89)
(90)
(91)
(92)
without affecting the eigenvalues of [X̂(1)]n1,a and [Ẑ(1)]n1,a. We can also control the eigenvalues of [X̂(1)]n1,a and [Ẑ(1)]n1,a by X̂(2) and Ẑ(2),
(93)
(94)
(95)
(96)
without affecting the eigenvalues of Âv0 and B̂p0. Since n1,a = na = N/da, this is what we needed.

3. Case 3: α10 and α20

Finally, we discuss the case when α1 ≠ 0 and α2 ≠ 0. We define operators X̂(1,2):=[X̂(1)]1[X̂(2)]2 and Ẑ(1,2):=[Ẑ(1)]1[Ẑ(2)]2.

Since a and αi are coprime, da in Eq. (80) can also be written as gcd(α1, α2, N). It follows that gcd(α1, α2)/da is coprime to na = N/da. Thus, there exists an integer b0 such that
(97)
Furthermore, Bézout’s lemma tells us the existence of integers b1 and b2 such that
(98)
Therefore, we have
(99)
with i = b0bi mod na (0 ≤ ina − 1). The eigenvalues of Âv0 and B̂p0 can be controlled by X̂(1,2) and Ẑ(1,2),
(100)
(101)
(102)
(103)
This process might affect the eigenvalues of the closed loop operators [X̂(i)]ni,a and [Ẑ(i)]ni,a.
Next, suppose that
(104)
In this case, X̂(1,2) and Ẑ(1,2) commute with Âv0 and B̂p0. For example, one can set 1=(α2+b2N)/gcd(α1+b1N,α2+b2N) and 2=(α1+b1N)/gcd(α1+b1N,α2+b2N), with b1,b2Z being free parameters. Choosing 1 and 2 properly, we can realize
(105)
See  Appendix A for the proof. Assuming this and using the relations
(106)
(107)
(108)
(109)
we can control the eigenvalues of [X̂(1)]n1,a and [Ẑ(1)]n1,a by a multiple of ωna without affecting the eigenvalues of Âv0 and B̂p0. This completes the proof of Eq. (55).

In this section, we clarify the relation of our model in Eq. (14) to the a=1ZN toric code with a twisted boundary condition. This connection for a prime N is implied by the result in Ref. 26, but our discussion goes more generally whenever N and a are coprime.

Let us consider a modified Hamiltonian
(110)
We still assume the periodic boundary condition. This model is equivalent to Ĥ in Eq. (14) in the sense that it is written as the sum of the same set of stabilizers Âv (vV) and B̂p (pP) in Eqs. (16) and (17) with a1 = a2 = a. The ground states are still given by those who have eigenvalue +1 for all Âv (vV) and B̂p (pP), and the ground state degeneracy remains unchanged.
We introduce a local unitary operator Ûr (r ∈ Λ), whose action on the local spin is given by a unitary matrix Ui,j := δj,1+mod[(i−1)a,N]. This operator satisfies
(111)
(112)
(113)
Here and hereafter, X̂r±a [ = 1, 2, …, MN(a)] should be understood as X̂r±aMN(a) (recall that aMN(a)=1 mod N). The global operator rΛÛr is a symmetry of Ĥ as it commutes with Ĥ. When gcd(N, a) ≠ 1, such a unitary operator does not exist.
Now, let us define a twist operator
(114)
The twist operator converts the stabilizers Âv and B̂p away from the boundary (i.e., 1 ≤ m1L1 − 2 and 1 ≤ m2L2 − 2) to those for a = 1,
(115)
(116)
Here, Âv(1) and B̂p(1) represent Âv and B̂p in Eqs. (16) and (17) for a1 = a2 = 1, respectively. Therefore, except for boundary terms,
(117)
is equivalent to the standard ZN toric code (a = 1). Boundary terms are given by
(118)
(119)
(120)
and
(121)
(122)
(123)
See Fig. 3 for the illustration. These boundary terms can be understood as a result of twisted boundary condition,
(124)
(125)
(126)
(127)
This boundary condition modifies the translation symmetries to T̂1:=m2=0L21[Û(12,m2)Û(0,m2+12)]L1T̂1 and T̂2:=m1=0L11[Û(m1,12)Û(m1+12,0)]L2T̂2, and the original translation symmetries in Eq. (15) are broken.
FIG. 3.

The standard ZN toric code with a twisted boundary condition.

FIG. 3.

The standard ZN toric code with a twisted boundary condition.

Close modal

When N and a are not coprime, our model in Eq. (14) cannot be mapped to the standard ZN toric code in this way. In fact, as we shall see in Sec. VII, they have different topological orders and cannot be mapped to each other by local unitary transformations.

In Sec. V, we showed that the order of ground state degeneracy under the periodic boundary condition can be 1 depending on the system size. Then, one might suspect that the system is not in a topologically ordered phase. In this section, we show this is not the case by demonstrating nontrivial topological entanglement entropy and anyonic excitations in the system. In addition, we discuss how the size dependence of the ground state degeneracy can be understood by viewing the lattice system as a continuum torus but with lattice translation symmetry defects.

Here, we compute the topological entanglement entropy Stopo of the ground state of our model. We use the Kitaev–Preskill prescription,5 
(128)
where
(129)
is the von Neumann entropy of the subregion R of the system and ρ̂R:=trR̄|Φ0Φ0| (trR̄ represents the partial trace over the complement of the region R) is the reduced density matrix of the ground state |Φ0⟩. The von Neumann entropy shows the area law behavior SR = α∂R + Stopo (R is the length of the boundary of the region R). The formula in Eq. (128) is designed in such a way that contributions from the area law term cancel.
The von Neumann entropy SR for a stabilizer Hamiltonian can be computed easily.10,27 Let G be the multiplicative group generated by all Âv’s (vV), B̂p’s (pP) and possible closed string operators for which |Φ0⟩ has the eigenvalue +1. Suppose that |Φ0⟩ is the unique state that has the eigenvalue +1 for all operators in G. Then, the projector onto |Φ0⟩ can be written as
(130)
We have ĝP̂G=P̂Gĝ=P̂G for any ĝG due to the rearrangement theorem. To see Eq. (130), it is enough to check that P̂G|Φ0=|Φ0 and P̂G|Ψ=0 if there exists ĝ*G such that ĝ*|Ψ=λ*|Ψ with λ* ≠ 1. The former is simply the definition of |Φ0⟩. The latter follows by applying P̂G=P̂Gĝ* to the state |Ψ⟩,
(131)
As tr[ĝ] is nonzero only when ĝ is identity, the order of the group G is given by |G|=N2L1L2. Similarly, trR̄[ĝ] can be nonzero only when ĝ is identity over R̄. Thus,
(132)
where nR is the number of N-level spins in R and GR is the subgroup of G supported in R. In the last step, we introduced the projector
(133)
Therefore, ρ̂R has only one nonzero eigenvalue λ=|GR|/NnR, whose order of degeneracy is nλ=NnR/|GR|=1/λ. Therefore,10,27
(134)
Up to this point, no assumption has been made on a.
When a is coprime to N, |GR| is given by NmR, where mR is the number of generators of G supported in R.10 Therefore, the formula in Eq. (134) reduces to
(135)
Using this formula, we find that the topological entanglement entropy of our model is
(136)
regardless of L1 and L2, as far as a is coprime to N. For example, for the subregions A, B, and C illustrated in Fig. 4(a), we have
(137)
We confirm this result by the exact diagonalization up to L1 = L2 = 3 and N = 3.
FIG. 4.

(a) Subregions A, B, and C used in the computation of Stopo. (b)–(g) Generators of GC. (b)–(d) Âv and B̂p themselves, and (e)–(g) ÂvNgcd(N,a) and B̂pNgcd(N,a). When N = N1N2 and a = N1a′ (N1, N2, and a′ are mutually coprime), then Ngcd(N,a)=N2.

FIG. 4.

(a) Subregions A, B, and C used in the computation of Stopo. (b)–(g) Generators of GC. (b)–(d) Âv and B̂p themselves, and (e)–(g) ÂvNgcd(N,a) and B̂pNgcd(N,a). When N = N1N2 and a = N1a′ (N1, N2, and a′ are mutually coprime), then Ngcd(N,a)=N2.

Close modal
When N and a have a common divisor, one needs to directly use the formula in Eq. (134). For example, let us take positive, mutually coprime integers N1, N2, a′ and set N = N1N2 and a = N1a′. For the subregions A, B, and C illustrated in Fig. 4(a), we find
(138)
Generators of GR used in the calculation are shown in Figs. 4(b)4(g) using region C as an example. This result is what one would expect from the ZNa topological order. However, more generally, we have
(139)
By definition [see Eqs. (48) and (52)], N/[Nagcd(N, a)] is a positive integer. When it is larger than one, Stopo is shifted from the expected value −log Na. We examine this additional contribution to Stopo in detail below.

It is known that the topological entanglement entropy may suffer from spurious contributions and may become nonzero even when the ground state does not have a topological order.10,28,29 Thus, we need to verify that the nonzero topological entanglement entropy found in Sec. VII A is the legitimate one.

In Ref. 28, it was shown that such spurious contributions can be captured by another combination of entropies computed for a dumbbell-shaped configuration,
(140)
Regions A, B, and C must be chosen carefully,28 and here, we assume those illustrated in Fig. 5(a).
FIG. 5.

(a) Subregions A, B, and C used in the computation of Sdumb. (b)–(g) Generators of GABC. Those simply given by Âv and B̂p are omitted. (b) and (c) Subsystem symmetry operators. (d)–(f) B̂pa.

FIG. 5.

(a) Subregions A, B, and C used in the computation of Sdumb. (b)–(g) Generators of GABC. Those simply given by Âv and B̂p are omitted. (b) and (c) Subsystem symmetry operators. (d)–(f) B̂pa.

Close modal
When a is coprime to N, we find
(141)
implying that Stopo in Eq. (136) is physical. This remains true more generally when N/[Nagcd(N, a)] = 1.
This is no longer the case when N/[Nagcd(N, a)] > 1. For example, when N = a2 (a > 1), there are no anyons (Na = 1) and the phase must be topologically trivial as we will discuss in Sec. VIII. However, in this case, N/[Nagcd(N, a)] = a > 1 and Stopo in Eq. (139) becomes
(142)
This nonzero value comes from the spurious contribution originating from subsystem symmetries. Subsystem symmetries are rigid string operators that cannot be deformed freely, unlike the Wilson loop operators, but commute with the Hamiltonian. In our model, subsystem symmetries exist when NaN. They have nontrivial contribution to Stopo and Sdumb when their ends have a shape illustrated by dashed lines in Figs. 4(b) and 4(c), which occurs when N/[Nagcd(N, a)] > 1. Indeed, when N = a2 (a > 1), we find
(143)
We illustrate generators of GABC used in the calculation in Figs. 5(b)5(f). These behaviors imply that N = a2 cases realize subsystem symmetry-protected topological (SSPT) phases, and we will come back to this point in Sec. VIII C.

When a is coprime to N, all magnetic and electric excitations can be understood as anyons with nontrivial mutual braiding statistics. They are created in pairs by open string operators, as in Sec. II D, or by extended string operators in Eqs. (67) and (70) without forming a pair. The appearance of anyonic excitations is another hallmark of topologically ordered phases.

When a is not coprime to N, some of magnetic and electric excitations are trivial in the sense they can be created locally without forming a pair. To see this, let us focus on divisors of N given by
(144)
If k < k′, dk/dk is a positive integer because
(145)
In particular, dk = Na for every kmax{rj}j=m+1n, where Na is defined in Eq. (52) and rj’s are powers appearing in the prime factorization in Eq. (48). Therefore, all dk’s are multiples of Na.
The string operator
(146)
creates a single magnetic excitation with the eigenvalue ωdk of B̂v. The eigenvalue of B̂v+(k,0) remains ωdkak=1. We can do the same for electric excitations. Hence, a magnetic or electric excitation with the eigenvalue ωNa (Z) can be created locally by [X̂v,Na(1)] without forming a pair. Conversely, if q is not a multiple of Na, excitations with eigenvalue ωq need to be created in pairs. Therefore, only excitations with the charge q = 1, 2, …, Na − 1 are nontrivial.
Generally, we label the anyonic excitations by their electric and magnetic charges qe and qm, where qe, qm ∈ {0, 1, …, Na − 1}. The topological order of this model is thus identical to that of the standard ZNa toric code model, i.e., the same anyon types, fusion rules, and braiding statistics. In particular, they satisfy the following fusion rule:
(147)
Here, [x]Na means x mod Na. Thus, we may view the anyons as an Abelian group A=ZNa×ZNa, with the multiplication given by fusion.
However, if we take into account lattice translation symmetry, the system can have distinct translation symmetry-enriched topological phases30 as the standard ZNa toric code. More specifically, under a unit translation in x1 or x2, an anyon (qe, qm) becomes
(148)
This action is well-defined since for every q = 1, 2, …, Na − 1, there exists (1 ≤ Na − 1) such that q = a mod Na. Then, aq := a+1 and a−1q := a−1 mod Na. When q = 0, aq = a−1q = 0.

We should mention that to completely describe the symmetry-enriched topological order, there are further information beyond the permutation action.30 However, they are not relevant for our purpose, so we will not consider them in more details.

When a ≠ 1, the Ti action generally changes anyon types. We can also see that TiMNa(a) keeps all anyon types invariant, so effectively Ti generates a ZMNa(a) symmetry group of the low-energy topological theory. In this section, we will use ρak to denote the permutation
(149)

Before we continue, it will be very useful to understand the properties of (point-like) symmetry defects, i.e., dislocations in this case.30–36 Generally, each symmetry defect is uniquely associated with a group element, which determines the symmetry action that takes place when moving around the defect. We denote the set of all defects associated with symmetry group element g by Cg. Note that for g = 1, trivial defects are nothing but the anyons. Symmetry defects are always at the end points of defect lines, which can be intuitively thought of as branch cuts where the symmetry action takes place. Just like anyons, defects can fuse with each other to new defects, and the fusion rules must respect the group multiplication structure. Defects can also fuse with anyons, which do not change the associated group element. See Ref. 30 for a more systematic discussion of defect fusion rules.

Let us consider the ρak defects. We pick one of them as a reference and denote it by σak,0. The other defects can be obtained by fusing σak,0 with anyons. Naively, one might think that the number of different defect types is the same as the number of anyon types. However, due to the permutation action, we also have the following fusion rule:
(150)
for any qe, qm. To see this, one can locally create a pair of anyons (qe, qm) and (−qe, −qm) near the defect, move (qe, qm) around the defect so it becomes (akqe, akqm), and then fuse it again with (−qe, −qm) to give [(ak − 1)qe, (ak − 1)qm]. In other words, σak,0 and σak,0×((ak1)qe,(ak1)qm) are related by a local operation, so must be the same type of defect.
Therefore, the defect types should be identified with a quotient of the group of anyons A by the subgroup generated by (ak − 1, 0) and (0, ak − 1).30,35 We will denote by [qe, qm] the equivalence classes of anyons under this quotient. Define tak=gcd(ak1,Na)=gcd(ak1,Na) [the second equality follows from gcd(ak, Na) = 1]; then, we can label the defects by σak,[qe,qm] with qe,m=0,1,,tak as representatives of the equivalence classes,
(151)
These different types of defects can be uniquely labeled by the braiding phases of ρak-invariant anyons around the defect. We can now define σak,0 as the defect where all such braiding phases are 1.

As an example, if Na is a prime and a ≠ 1 mod Na, then the subgroup generated by ak − 1 for 0 < k < Na − 1 is basically the entire group ZNa. Hence, the quotient group has a single element, and there is only a unique type of defect.

We also need to know how the ρak defects transform under the ρak action. It is clear that σak,0 is invariant under ρak. Hence, the action on σak,[qe,qm] is given by
(152)

Let us now consider the ground state degeneracy on a torus, with a ρak1 defect line in one direction and a ρak2 defect line in the other direction. According to the general theory in Ref. 30, the ground state degeneracy is equal to the number of ρak defect types invariant under ρak action given in Eq. (152).

We now show that the number of such ρak1 defects is
(153)
To see why, first, note that the invariance of σak,[qe,qm] under ρak means that qe, qm satisfy
(154)
Without any loss of generality, we can restrict qe,qm{0,1,,tak}. Clearly, we can treat the electric and magnetic sector separately, so we will focus on the electric sector and suppress the subscript e. To shorten notations, define b1 = ak − 1, b2 = ak′ − 1, and ti = gcd(bi, Na). In the electric sector, Eq. (154) means that there exists an integer r such that
(155)
Given q, this is possible if and only if t1 = gcd(b1, Na) divides b2q. In other words, there exists an integer r′ such that
(156)
The smallest positive integer q that makes it solvable is t1gcd(t1,b2). Note that gcd(t1, b2) = gcd(gcd(b1, Na), b2) = gcd(b1, b2, Na). Therefore, the number of solutions is precisely gcd(b1, b2, Na). The same argument works for the magnetic sector, so together, we find that the total number of solutions to Eq. (154) is given by gcd(b1,b2,Na)2=gcd(ak1,ak1,Na)2.

We now show that knowing the permutation action of Ti on anyons is enough to derive the topological degeneracy. Here, the key is to think of a L1 × L2 torus as a torus in continuum, but with a T1L1 defect line along x2 and a T2L2 defect line along x1. Intuitively, this is because traveling across the torus in the xi direction is the same as translating by Li. With this picture, the ground state degeneracy is obtained by substituting k = L1 and k′ = L2, which reproduces the result in Eq. (55). In our model, as shown in Sec. VI, when a and N are coprime, we can indeed explicitly map the Hamiltonian on a torus to the standard toric code (where translation symmetry acts trivially) with a twisted boundary condition or equivalently with symmetry defect lines wrapping around the two non-contractible cycles, in full agreement with the argument in this section. The standard toric code has a smooth continuum limit; thus, the finite-lattice effect is completely captured by the defect lines, establishing the continuum picture at the microscopic level.

In this section, we consider the case when a is a multiple of rad(N).

Let us demonstrate the uniqueness of the ground state regardless of the choice of the system size L1 and L2 although it is already implied by our general formula in Eq. (55) with Na = 1.

Let 0 be the smallest positive integer such that
(157)
To see that 0 indeed exists, let us write rM:=max{ri}i=1n, where ri’s are powers appearing in the prime factorization of N in Eq. (48). Because rad(N)rM=j=1npjrM is a multiple of N and also because arM is a multiple of rad(N)rM, we have
(158)
Therefore, 0 is in the range 1 ≤ 0rM.
Then, our discussion in Sec. II C implies that the state
(159)
contains a magnetic excitation with eigenvalue ω at the plaquette p=(m1+12,m2+12). The eigenvalue of the plaquette operator B̂(m1+0+12,m2+12) remains 1. Most importantly, the string operator in Eq. (159) is local in the sense that its length 0 does not depend on the system size. Hence, a single elementally magnetic excitation can be created locally. Similarly, the state
(160)
contains an electric excitation with the eigenvalues ω−1 at the vertex v = (m1, m2). The rest of the discussion proceeds exactly the same as in Sec. V A 2. Therefore,
(161)
for any L1 and L2.
As an example, let us discuss the case of N = a. In this case, the Hamiltonian is completely decoupled,
(162)
(163)
The ground state of ĥr is unique and has the energy gap Δ1. We denote the ground state by |ϕ0r. Then, the unique ground state of Ĥ is given by the product state rΛ|ϕ0r. Therefore, this phase is completely trivial. Indeed, the topological entablement entropy in Eq. (128) vanishes
(164)
for the subregions in Fig. 4(a).

Next, we discuss the case of N = a2. We argue that this case realizes a SSPT phase.37,38

To this end, we study the property of the model obtained by rotating the one introduced in Sec. II by 45° (see Fig. 6). Spins are now defined on square lattice sites r=(m̄1,m̄2) with m̄1,m̄2Z. Vertices and plaquettes can be associated with odd (even) sites,
(165)
(166)
The Hamiltonian is given by
(167)
where
(168)
(169)
FIG. 6.

The model rotated by 45°, corresponding to the L̄1=L̄2=6 case. Dashed lines represent the original square lattice before the rotation. (a) Plaquette operators and vertex operators. (b) Subsystem symmetries. The red jagged line represents the symmetry flux across the link between (L̄11,2j2+1) and (0, 2j2 + 1). (c) Edge zero modes under an open boundary condition protected by subsystem symmetries.

FIG. 6.

The model rotated by 45°, corresponding to the L̄1=L̄2=6 case. Dashed lines represent the original square lattice before the rotation. (a) Plaquette operators and vertex operators. (b) Subsystem symmetries. The red jagged line represents the symmetry flux across the link between (L̄11,2j2+1) and (0, 2j2 + 1). (c) Edge zero modes under an open boundary condition protected by subsystem symmetries.

Close modal

1. Charge pumping

Let us work with the periodic boundary condition first. We identify r+(n1L̄1,n2L̄2) with r for n1,n2Z. Both L̄1 and L̄2 are assumed to be even. Unit translation symmetries of the model shift r by either (1, 1) or (1, −1).

The model has subsystem symmetries,
(170)
(171)
for each m̄2 separately, as illustrated in Fig. 6(b), which act only on a single row. When N = a2, these operators can be rewritten in terms of stabilizers as
(172)
(173)
when m̄2 is even and
(174)
(175)
when m̄2 is odd. Hence, in the ground state |Φ0⟩ where all vertex operators and plaquette operators take the value +1, we have
(176)
for all m̄2.
Now, we insert a symmetry flux associated with the subsystem symmetry Ẑm̄2=2j2+1 at the link between m̄1=L̄11 and m̄1=0. This operation multiplies a factor ωa to the vertex term Â(L̄11,2j2) [the red shaded vertex in Fig. 6(b)],
(177)
In the ground state |Φ0 of Ĥ, the eigenvalue of Â(L̄11,2j2) is thus modified to ωa. Therefore, using Eqs. (172) and (174), we find
(178)
Namely, the charge ωa is pumped for the subsystem symmetry X̂2j2 upon inserting the symmetry flux associated with the subsystem symmetry Ẑ2j2+1. This pumped charge is a topological invariant that distinguishes this phase from product states.

2. Zero energy edge states

Next, let us consider the open boundary condition. We impose the subsystem symmetries X̂m̄2 and Ẑm̄2 in Eqs. (170) and (171) for every m̄2, including the edges.

We introduce two sets of generalized Pauli matrices,
(179)
(180)
and
(181)
(182)
which commute with all stabilizers in the bulk Hamiltonian. They satisfy
(183)
for s, s′ = L, R and j,j=1,2,,L̄221. A pair of σ̂2j,sx and σ̂2j,sz generates a ZN×ZN symmetry, implying N-fold degeneracy, and there are L̄22 such pairs. This NL̄22-fold degeneracy cannot be lifted by perturbations on the edges, as long as the subsystem symmetries are maintained. In contrast, the two edges at m̄2=0 and m̄2=L̄21 can be gapped by edge perturbations.

As a concluding remark, let us discuss implications of our example on the Lieb–Schultz–Mattis (LSM)-type theorems,39–53 which formulate necessary conditions for the unique ground state with a nonzero excitation gap under the periodic boundary condition. When one of these conditions are not satisfied, the appearance of ground state degeneracy or gapless excitations is guaranteed. The ground state degeneracy originates from either spontaneous symmetry breaking or topological degeneracy. Hence, violation of LSM-type conditions in symmetric and gapped phases can be used as a sufficient condition for a nontrivial topological order.45,47

There are a variety of such theorems applicable to quantum many-body systems in different settings. For example, in one dimension, an early version of LSM theorems for quantum spin chains with spin-rotation symmetry states that S needs to be an integer in the presence of the time-reversal symmetry.39,40 More generally, Sm (m is the magnetization per unit cell) must be an integer to realize a unique gapped ground state.41 Similarly, in fermionic systems with U(1) symmetry, the filling ν (the average number of fermions per unit cell) must be an integer.42 These results apply to any sequence of L1. One can even start with the infinite system from the beginning.52,53

In contrast, there is usually a restriction on the choice of the sequence of Li’s in higher-dimensional extensions of these theorems. In the formulation, one usually starts with a finite size system with the length Li in xi direction (i = 1, …, d) and considers the limit L1, …, Ld → +∞. For example, for spin systems, the arguments in Refs. 44 and 46 are effective only when L2, …, Ld are all odd. For particle systems, the discussions in Refs. 43 and 51 assume that L2, …, Ld are coprime to q when ν = p/q. There is a way to remove such a restriction by modifying the boundary condition to a tilted one,54 but this argument is not about the original periodic boundary condition. Namely, changing the boundary condition from the periodic one to the tilted one might affect the degeneracy or excitation gap.

As we demonstrated through an example, a topologically ordered phase may not show topological degeneracy on torus depending on the sequence of system size. Hence, even when all the LSM-type conditions are fulfilled and the ground state is indeed unique in some sequences of the system size, it still might be the case that the ground state is actually topologically ordered.

H.W. thanks Hiroki Hamaguchi for informing us of the proof of Eq. (A4) in  Appendix A. The work of H.W. was supported by JSPS KAKENHI under Grant Nos. JP20H01825 and JP21H01789. M.C. acknowledges support from NSF under Award No. DMR-1846109. The work of Y.F. was supported by JSPS KAKENHI under Grant No. JP20K14402 and JST CREST under Grant No. JPMJCR19T2. H.W. acknowledges the hospitality and fruitful discussions at the Institute of Basic Science, Daejeon, Korea, during the week of Conference on Advances in The Physics of Topological and Correlated Matter.

The authors have no conflicts to disclose.

Haruki Watanabe: Formal analysis (lead); Investigation (lead); Writing – original draft (lead); Writing – review & editing (lead). Meng Cheng: Formal analysis (equal); Investigation (equal); Supervision (equal); Validation (lead); Writing – original draft (equal); Writing – review & editing (equal). Yohei Fuji: Formal analysis (equal); Investigation (equal); Supervision (lead); Validation (equal); Writing – original draft (equal); Writing – review & editing (equal).

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

Here, we demonstrate the validity of Eq. (105). As stated in the main text, we set
(A1)
Using the properties of the greatest common divisor, we find
(A2)
and
(A3)
In the last line, we used the fact that N/da is coprime to gcd(α1 + b1 N, α2 + b2 N)/da. Hence, Eq. (105) can be rewritten as
(A4)

Below, we prove the following statement: for any integer N ≥ 2 and integers α1 and α2 in the range 1 ≤ α1, α2N − 1, there always exist integers b1 and b2 such that Eq. (A4) holds.55 In particular, b2 can be set 0. Since this is trivially the case when da = 1, in the following, we assume da ≠ 1. We introduce shorthands α1:=α1/da, α2:=α2/da, and N′ := N/da.

For an integer m and a prime p, let us denote by νp(m) the largest non-negative integer ν such that pν divides m. Suppose that ej:=νpj(da)1 for j = 1, 2, …, J. In other words, da can be prime-factorized as da=j=1Jpjej. Then, Eq. (A4) holds if and only if
(A5)
for all j = 1, 2, …, J. In addition, by definition,
(A6)
Therefore, if
(A7)
simultaneously for all j = 1, 2, …, J, Eq. (A5) is fulfilled. In the following, we write nj:=νpj(N) and mj:=νpj(α1).
Let us derive the condition for Eq. (A7). When njmj, we need
(A8)
with α1/pjmj0 mod pj. In this case, we can set b1 = 0 mod pj. On the other hand, when mj > nj, we need
(A9)
with α1/pjnj=0 mod pj and N/pjnj0 mod pj. In this case, we can set b1 = 1 mod pj. After all, we found a condition of the form b1 = xj mod pj for each j = 1, 2, …, J. The Chinese remainder theorem guarantees the existence b1 in the range 0 to 1+j=1Jpj such that these conditions are simultaneously satisfied.
When N1 and N2 are coprime, the N = N1N2-level spin can be decomposed into the tensor product of N1- and N2-level spins. To see this, let us write the matrices in Eqs. (3) and (4) as X(N) and Z(N), respectively. We have
(B1)
(B2)
where [V]i1i2,i:=δi1,mod(N2(i11)+N1(i21),N) and [Z(N1)Z(N2)]i1i2,j1j2:=[Z(N1)]i1,j1[Z(N2)]i2,j2 for i1 = 1, …, N1, i2 = 1, …, N2, and i = 1, …, N. These reduction formulas can be readily shown by using the representations in Eqs. (3) and (4).
Let us discuss the implication of these relations. Suppose that N = N1N2 and N1 and N2 are coprime. We introduce another modified Hamiltonian,
(B3)
where Âv and B̂p are vertex and plaquette operators in Eqs. (16) and (17) with a1 = a2 = a. The eigenstates of this Hamiltonian are also identical to those for Ĥ in Eq. (14), and the ground state degeneracy remains unchanged.
Let V̂ be the global unitary operator whose action on each N level spin is given by the unitary matrix V above. Using the reduction formulas, we find
(B4)
where Âv(Ni) and B̂p(Ni)(i=1,2) are vertex and plaquette operators for Ni-level spins. Ground states have the eigenvalue +1 for all Âv(Ni)’s and B̂p(Ni)’s. This result indicates that, if we denote the ground state degeneracy of the our model for N-level spin by Ndeg(N),
(B5)
Indeed, this is consistent with our result in Eq. (55) because
(B6)
when N1 and N2 are coprime. More generally, for the form of N in Eq. (48), we have
(B7)
However, this decomposition alone is not sufficient to derive Eq. (55). One still has to compute Ndeg(pjrj), and this requires an investigation, which is almost as hard as what we did in this work.
1.
X.-G.
Wen
,
Quantum Field Theory of Many-Body Systems: From the Origin of Sound to an Origin of Light and Electrons
(
Oxford University Press
,
New York
,
2004
).
2.
X.-G.
Wen
, “
Colloquium: Zoo of quantum-topological phases of matter
,”
Rev. Mod. Phys.
89
,
041004
(
2017
).
3.
B.
Zeng
,
X.
Chen
,
D.-L.
Zhou
, and
X.-G.
Wen
,
Quantum Information Meets Quantum Matter: From Quantum Entanglement to Topological Phases of Many-Body Systems
(
Springer
,
2019
).
4.
M.
Oshikawa
and
T.
Senthil
, “
Fractionalization, topological order, and quasiparticle statistics
,”
Phys. Rev. Lett.
96
,
060601
(
2006
).
5.
A.
Kitaev
and
J.
Preskill
, “
Topological entanglement entropy
,”
Phys. Rev. Lett.
96
,
110404
(
2006
).
6.
M.
Levin
and
X.-G.
Wen
, “
Detecting topological order in a ground state wave function
,”
Phys. Rev. Lett.
96
,
110405
(
2006
).
7.
E.
Dennis
,
A.
Kitaev
,
A.
Landahl
, and
J.
Preskill
, “
Topological quantum memory
,”
J. Math. Phys.
43
,
4452
4505
(
2002
).
8.
A. Y.
Kitaev
, “
Fault-tolerant quantum computation by anyons
,”
Ann. Phys.
303
,
2
30
(
2003
).
9.
S. S.
Bullock
and
G. K.
Brennen
, “
Qudit surface codes and gauge theory with finite cyclic groups
,”
J. Phys. A: Math. Theor.
40
,
3481
3505
(
2007
).
10.
L.
Zou
and
J.
Haah
, “
Spurious long-range entanglement and replica correlation length
,”
Phys. Rev. B
94
,
075151
(
2016
).
11.
K.
Slagle
and
Y. B.
Kim
, “
Quantum field theory of X-cube fracton topological order and robust degeneracy from geometry
,”
Phys. Rev. B
96
,
195139
(
2017
).
12.
S.
Vijay
, “
Isotropic layer construction and phase diagram for fracton topological phases
,” arXiv:1701.00762 (
2017
).
13.
M. D.
Schulz
,
S.
Dusuel
,
R.
Orús
,
J.
Vidal
, and
K. P.
Schmidt
, “
Breakdown of a perturbed ZN topological phase
,”
New J. Phys.
14
,
025005
(
2012
).
14.
M.
Barkeshli
,
P.
Bonderson
,
M.
Cheng
,
C.-M.
Jian
, and
K.
Walker
, “
Reflection and time reversal symmetry enriched topological phases of matter: Path integrals, non-orientable manifolds, and anomalies
,”
Commun. Math. Phys.
374
,
1021
1124
(
2020
).
15.
J. C.
Bridgeman
,
S. D.
Bartlett
, and
A. C.
Doherty
, “
Tensor networks with a twist: Anyon-permuting domain walls and defects in projected entangled pair states
,”
Phys. Rev. B
96
,
245122
(
2017
).
16.
Y.
Fuji
, “
Anisotropic layer construction of anisotropic fracton models
,”
Phys. Rev. B
100
,
235115
(
2019
).
17.
X. G.
Wen
, “
Quantum orders in an exact soluble model
,”
Phys. Rev. Lett.
90
,
016803
(
2003
).
18.
D. J.
Williamson
,
Z.
Bi
, and
M.
Cheng
, “
Fractonic matter in symmetry-enriched u(1) gauge theory
,”
Phys. Rev. B
100
,
125150
(
2019
).
19.
Y.-T.
Oh
,
J.
Kim
,
E.-G.
Moon
, and
J. H.
Han
, “
Rank-2 toric code in two dimensions
,”
Phys. Rev. B
105
,
045128
(
2022
).
20.
S. D.
Pace
and
X.-G.
Wen
, “
Position-dependent excitations and UV/IR mixing in the ZN rank-2 toric code and its low-energy effective field theory
,”
Phys. Rev. B
106
,
045145
(
2022
).
21.
G.
Delfino
,
W. B.
Fontana
,
P. R. S.
Gomes
, and
C.
Chamon
, “
Effective fractonic behavior in a two-dimensional exactly solvable spin liquid
,”
SciPost Phys.
14
,
002
(
2023
).
22.
Y.-T.
Oh
,
J.
Kim
, and
J. H.
Han
, “
Effective field theory of dipolar braiding statistics in two dimensions
,”
Phys. Rev. B
106
,
155150
(
2022
).
23.
C.
Schommer-Pries
, “
Tori detect invertibility of topological field theories
,”
Geom. Topol.
22
,
2713
2756
(
2018
).
24.
H.
Tasaki
,
Physics and Mathematics of Quantum Many-Body Systems
(
Springer
,
2020
).
25.
W.
Stein
,
Elementary Number Theory: Primes, Congruences, and Secrets: A Computational Approach
(
Springer
,
2008
).
26.
J.
Haah
, “
Classification of translation invariant topological Pauli stabilizer codes for prime dimensional qudits on two-dimensional lattices
,”
J. Math. Phys.
62
,
012201
(
2021
).
27.
N.
Linden
,
F.
Matus
,
M. B.
Ruskai
, and
A.
Winter
, “
The quantum entropy cone of stabiliser states
,” in
8th Conference on the Theory of Quantum Computation, Communication and Cryptography (TQC 2013)
,
Leibniz International Proceedings in Informatics (LIPIcs) Vol. 22
, edited by
S.
Severini
and
F.
Brandao
(
Schloss Dagstuhl – Leibniz-Zentrum für Informatik
,
Dagstuhl, Germany
,
2013
), pp.
270
284
.
28.
D. J.
Williamson
,
A.
Dua
, and
M.
Cheng
, “
Spurious topological entanglement entropy from subsystem symmetries
,”
Phys. Rev. Lett.
122
,
140506
(
2019
).
29.
D. T.
Stephen
,
H.
Dreyer
,
M.
Iqbal
, and
N.
Schuch
, “
Detecting subsystem symmetry protected topological order via entanglement entropy
,”
Phys. Rev. B
100
,
115112
(
2019
).
30.
M.
Barkeshli
,
P.
Bonderson
,
M.
Cheng
, and
Z.
Wang
, “
Symmetry fractionalization, defects, and gauging of topological phases
,”
Phys. Rev. B
100
,
115147
(
2019
); arXiv:1410.4540.
31.
H.
Bombin
, “
Topological order with a twist: Ising anyons from an Abelian model
,”
Phys. Rev. Lett.
105
,
030403
(
2010
); arXiv:1004.1838 [cond-mat.str-el].
32.
Y.-Z.
You
and
X.-G.
Wen
, “
Projective non-Abelian statistics of dislocation defects in a ZN rotor model
,”
Phys. Rev. B
86
,
161107
(
2012
); arXiv:1204.0113 [cond-mat.str-el].
33.
M.
Barkeshli
,
C.-M.
Jian
, and
X.-L.
Qi
, “
Theory of defects in Abelian topological states
,”
Phys. Rev. B
88
,
235103
(
2013
); arXiv:1305.7203 [cond-mat.str-el].
34.
J. C. Y.
Teo
,
A.
Roy
, and
X.
Chen
, “
Unconventional fusion and braiding of topological defects in a lattice model
,”
Phys. Rev. B
90
,
115118
(
2014
).
35.
J. C. Y.
Teo
,
T. L.
Hughes
, and
E.
Fradkin
, “
Theory of twist liquids: Gauging an anyonic symmetry
,”
Ann. Phys.
360
,
349
445
(
2015
).
36.
N.
Tarantino
,
N. H.
Lindner
, and
L.
Fidkowski
, “
Symmetry fractionalization and twist defects
,”
New J. Phys.
18
,
035006
(
2016
).
37.
Y.
You
,
T.
Devakul
,
F. J.
Burnell
, and
S. L.
Sondhi
, “
Subsystem symmetry protected topological order
,”
Phys. Rev. B
98
,
035112
(
2018
).
38.
T.
Devakul
,
D. J.
Williamson
, and
Y.
You
, “
Classification of subsystem symmetry-protected topological phases
,”
Phys. Rev. B
98
,
235121
(
2018
).
39.
E.
Lieb
,
T.
Schultz
, and
D.
Mattis
, “
Two soluble models of an antiferromagnetic chain
,”
Ann. Phys.
16
,
407
466
(
1961
).
40.
I.
Affleck
and
E. H.
Lieb
, “
A proof of part of Haldane’s conjecture on spin chains
,”
Lett. Math. Phys.
12
,
57
69
(
1986
).
41.
M.
Oshikawa
,
M.
Yamanaka
, and
I.
Affleck
, “
Magnetization plateaus in spin chains: ‘Haldane gap’ for half-integer spins
,”
Phys. Rev. Lett.
78
,
1984
1987
(
1997
).
42.
M.
Yamanaka
,
M.
Oshikawa
, and
I.
Affleck
, “
Nonperturbative approach to Luttinger’s theorem in one dimension
,”
Phys. Rev. Lett.
79
,
1110
1113
(
1997
).
43.
M.
Oshikawa
, “
Commensurability, excitation gap, and topology in quantum many-particle systems on a periodic lattice
,”
Phys. Rev. Lett.
84
,
1535
1538
(
2000
).
44.
M. B.
Hastings
, “
Lieb-Schultz-Mattis in higher dimensions
,”
Phys. Rev. B
69
,
104431
(
2004
).
45.
M. B.
Hastings
, “
Sufficient conditions for topological order in insulators
,”
Europhys. Lett.
70
,
824
830
(
2005
).
46.
B.
Nachtergaele
and
R.
Sims
, “
A multi-dimensional Lieb-Schultz-Mattis theorem
,”
Commun. Math. Phys.
276
,
437
472
(
2007
).
47.
H.
Watanabe
,
H. C.
Po
,
A.
Vishwanath
, and
M.
Zaletel
, “
Filling constraints for spin-orbit coupled insulators in symmorphic and nonsymmorphic crystals
,”
Proc. Natl. Acad. Sci. U. S. A.
112
,
14551
14556
(
2015
).
48.
M.
Cheng
,
M.
Zaletel
,
M.
Barkeshli
,
A.
Vishwanath
, and
P.
Bonderson
, “
Translational symmetry and microscopic constraints on symmetry-enriched topological phases: A view from the surface
,”
Phys. Rev. X
6
,
041068
(
2016
).
49.
H. C.
Po
,
H.
Watanabe
,
C.-M.
Jian
, and
M. P.
Zaletel
, “
Lattice homotopy constraints on phases of quantum magnets
,”
Phys. Rev. Lett.
119
,
127202
(
2017
).
50.
D. V.
Else
and
R.
Thorngren
, “
Topological theory of Lieb-Schultz-Mattis theorems in quantum spin systems
,”
Phys. Rev. B
101
,
224437
(
2020
).
51.
S.
Bachmann
,
A.
Bols
,
W.
De Roeck
, and
M.
Fraas
, “
A many-body index for quantum charge transport
,”
Commun. Math. Phys.
375
,
1249
1272
(
2020
).
52.
Y.
Ogata
,
Y.
Tachikawa
, and
H.
Tasaki
, “
General Lieb–Schultz–Mattis type theorems for quantum spin chains
,”
Commun. Math. Phys.
385
,
79
99
(
2021
).
53.
H.
Tasaki
, “
The Lieb-Schultz-Mattis theorem: A topological point of view
,” in The Physics and Mathematics of Elliott Lieb, edited by R. L. Frank, A. Laptev, M. Lewin, and R. Seiringer (EMS Press, 2022), Vol. 2, pp. 405-446.
54.
Y.
Yao
and
M.
Oshikawa
, “
Generalized boundary condition applied to Lieb-Schultz-Mattis-type ingappabilities and many-body Chern numbers
,”
Phys. Rev. X
10
,
031008
(
2020
).
55.

The proof is due to Hiroki Hamaguchi.