Topological quantum error correction based on the manipulation of the anyonic defects constitutes one of the most promising frameworks towards realizing fault-tolerant quantum devices. Hence, it is crucial to understand how these defects interact with external defects such as boundaries or domain walls. Motivated by this line of thought, in this work, we study the fusion events between anyons in the bulk and at the boundary in fixed-point models of 2 + 1-dimensional non-chiral topological order defined by arbitrary fusion categories. Our construction uses generalized tube algebra techniques to construct a bi-representation of bulk and boundary defects. We explicitly derive a formula to calculate the fusion multiplicities of a bulk-to-boundary fusion event for twisted quantum double models and calculate some exemplary fusion events for Abelian models and the (twisted) quantum double model of S3, the simplest non-Abelian group-theoretical model. Moreover, we use the folding trick to study the anyonic behavior at non-trivial domain walls between twisted S3 and twisted as well as models. A recurring theme in our construction is an isomorphism relating twisted cohomology groups to untwisted ones. The results of this work can directly be applied to study logical operators in two-dimensional topological error correcting codes with boundaries described by a twisted gauge theory of a finite group.
I. INTRODUCTION
Topological phases of matter are intriguing zero-temperature quantum phases that are accompanied by robust ground state degeneracy and patterns of long-range quantum entanglement. They constitute cornerstones of modern condensed matter physics.1,2 At the same time, they play a central role in notions of topological quantum error correction (QEC) that is widely seen as one of the most promising paradigms for scalable quantum computing, a paradigm in which anyonic defects in the code are suitably manipulated to perform quantum information processing.3,4 In more abstract terms, a common high-level description of topologically ordered models is via the codimension-2 defects or “excitations.” The most famous example for this is the anyons in a 2 + 1-dimensional model, which are characterized by their fusion and braiding statistics.3 Such a description via higher-level invariant data can be extended to boundaries or domain walls, by also considering excitations within the boundary.
A more concrete lower-level description of topological order is via microscopic fixed-point models, which are exactly solvable due to a notion of discrete topological invariance.5 Those models allow for the computation of any higher-level invariants usind only finite-dimensional linear algebra. In 2 + 1 space-time dimensions any non-chiral topological order can be represented by such a microscopic fixed-point model, with finite-dimensional Hilbert spaces.
There are two major pictures in which those fixed-point models can be formulated, namely, the space (or string-net) picture, and the space-time picture. In the space picture, we start by assigning a Hilbert space to a cellulation of a two-dimensional manifold. The most well-known examples are string-net models defined on a trivalent cellulation2 or its dual formulation in terms of a triangulation of the same manifold.6 To achieve topological invariance, these models are also equipped with partial isometries mapping between the vector spaces of different cellulations. In fact, arbitrary changes of the cellulation are generated by local moves such as Pachner moves, so the model is defined by the associated local isometries. A local ground-state projector, or Hamiltonian, can be defined via a local sequence of moves/isometries which have no net effect on the cellulation.
In the space-time picture, a model is given by a discrete state-sum path integral in Euclidean space-time, the most famous one being the Turaev–Viro state-sum.7 Such a state-sum is defined on any three-dimensional cellulation of that space-time, and assigns a finite label set to any edge in that cellulation. The highest-dimensional cells carry weights depending on those labels. The state-sum is evaluated by taking the product of all weights and summing over all label configurations, and it must be invariant under changes in the cellulation. The cellulations in the space picture can be interpreted as codimension-1 sections of the cellulations in the space-time picture, giving rise to the equivalence of the two pictures. In this work, we study topological phases on manifolds with boundary using both pictures, since they each have their own up- and downsides. The space/string-net picture is more illustrative to most physicists since it is directly related to the usual quantum mechanical language of Hilbert spaces, states, and Hamiltonians. Hence, we mainly present our constructions in this space picture. However, the mapping between different cellulations in this picture can be quite tedious to work out, so we fall back on the space-time picture to evaluate complicated sequences of moves.
Topological fixed-point models on manifolds with boundaries have been studied in various places in the literature to understand properties of the defects on the boundary and how they interact with each other. References 8–10 focus on corners, defects between different types of boundaries or domain walls. In particular, they calculate vertical fusion events of codimension 2 defects along the same domain wall and horizontal fusion of defects on neighboring domain walls. Combining their methods allows them to also calculate the associator of these defects. Reference 11 gives an algebraic description of what happens to bulk defects when approaching the boundary in terms of a forgetful functor on tensor category describing the bulk anyons. However, they do not give a constructive framework to calculate these properties. Reference 12 gives a formula to calculate the set of condensable anyons in gauge theory models with trivial three-cocycle. In general, it is known that the set of condensable anyons have to form a Lagrangian algebra object in the modular tensor category that describes the bulk anyons.13–15
What we have found to be missing is a full description of the fusion of bulk anyons to boundary anyons. More explicitly, we are interested in the dimensions mij ≥ 0 of the fusion spaces between an (ingoing) bulk anyon i and an (outgoing) boundary anyon j. Importantly, we are interested in constructive formulas to calculate mi,j.
Apart from a mathematical interest, these fusion events find application in topological quantum error correction and computing with boundaries.16 For example, the logical operators of a topological code on a manifold with boundary are associated to ribbon operators17 of anyons which connect different boundary segments through the bulk. Moreover, any lattice-surgery-based computation scheme ultimately relies on deforming non-transparent domain walls between code patches into (partly-)transparent ones in a systematic way.18–20 Understanding how the anyon ribbon operators precisely behave close to these domain walls is therefore essential in the design of novel computational protocols in topological codes. In this sense, the work done here is also expected to provide guidance when devising novel schemes of topological quantum computing involving notions of lattice-surgery with codes beyond untwisted quantum doubles.
With the framework established in this work we aim to contribute to a further understanding of topological phases with boundaries by formulating a framework to describe bulk-to-boundary anyon fusion events in topological fixed-point models. We explicitly derive a closed formula for fusion multiplicities of fusion events between bulk and boundary anyons for 2 + 1-dimensional twisted gauge theory models, also known as Dijkgraaf–Witten state-sums.21 In this case, the anyons in the bulk as well as on the boundary are classified by irreducible sub-spaces of some special type of algebras, which we call twisted group algebras with action. We show how such an algebra is diagonalized and discover that there is an intimate connection to a group cohomological isomorphism that also appears in the classification of topological boundaries of gauge theories.
This manuscript is structured as follows. In Sec. II, we give a general recipe to find the irreducible sub-spaces of twisted group algebras with action, which characterize both the bulk and boundary anyons in gauge theory models of a finite group. In Sec. III, we give a self-contained introduction into string-net fixed-point models for topological phases with boundaries in two spatial dimensions and illustrate the equivalence to space-time state-sum models. This section is mainly addressed to readers not yet familiar with these models. Readers with a background in both formulations of fixed-point models might want to use that section to get familiar with our notation and conventions in the upcoming sections. In Secs. IV and V, we classify the anyons in the bulk and boundary. In particular, we focus on gauge theory models of a finite group. The main result of this paper is presented in Sec. VI, where we define a bimodule that allows to calculate the dimensions of bulk-to-boundary fusion events in any fixed-point model and explicitly derive a closed formula for the gauge theory case. Lastly, we give many examples, that partly already appeared in the literature, but combine them with new calculations to show the wide applicability of our formula to boundaries as well as domain walls. Finally, we conclude the results and give an outlook into possible continuations of this work in Sec. VII. For a reader interested in the technical details going into the derivation of the bimodule used in Sec. VI and tools to diagonalize the group algebras characterizing point defects, we refer to Appendixes A–E for further details.
II. DIAGONALIZING TWISTED GROUP ALGEBRAS WITH ACTION
Before we introduce topological fixed-point models, we want to highlight the technical tools used to classify bulk and boundary anyons of gauge theory models in Secs. IV and V. In both cases topological invariance of the anyonic subspaces naturally defines a finite-dimensional algebra. In most parts of this paper we focus on gauge theory models, derived from a finite group and a three-cocycle on it. For these models both bulk and boundary anyons are classified by a special type of algebra. In this section, we will present the technical tools used to find the irreducible sub-spaces of, i.e., block-diagonalize, these algebras.
The goal of this section is to find the irreducible sub-spaces of A. In particular, we find faithful invariants classifying the sub-spaces, determine their dimension and find the associated central idempotents whose representations project onto the respective sub-spaces. We give a comprehensive summary in terms of a recipe to construct the central idempotents at the end of this section.
This allows us to uniquely construct the IPRs of all stabilizer groups StabG(x) for x ∈ Xi from the IPRs of the stabilizer group of a single representative .
We summarize this section by giving a recipe on how to find the irreducible representations of an algebra of the form of Eq. (2).
Decompose X into transitive G-orbits {Xi}.
For each Xi:
Pick representative and calculate its stabilizer group .
Find irreducible -projective representations of Ki, denote them with ρi.
Use Eq. (9) to derive irreducible representations and character functions, for all x ∈ Xi.
- The indecomposable central idempotents in A are labeled by pairs (i, ρi) ∈ (G-orbits of X, Ψx-projective Irreps of Ki) and given byNote that the above is not a complete algorithm for finding the algebra irreducible representations but merely reduces it to finding the irreducible representations of a much smaller algebra in step 2b. Those irreducible representations are equivalent to projective group representations which can be found in the literature in many cases.(16)
III. MODELS FOR TOPOLOGICAL PHASES WITH BOUNDARY
Topological phases which possess gapped boundaries can be studied using fixed-point models on a discretized space-time. Topological invariance highly restricts the microscopic constituents of these models. In the Turaev–Viro state-sum,7,24 or tensor-network path integrals5,25 for 2 + 1-dimensional topological order, the topological invariance corresponds to recellulations in a three-dimensional space-time, and the algebraic constraints correspond to the ones defining spherical fusion categories. A different but equivalent picture are Levin–Wen string-net models,2 where recellulations on a two-dimensional space triangulation are represented by linear operators acting on the local degrees of freedom. In this picture, the topological invariance in space-time takes the form of coherence axioms between different equivalent space recellulations. Since these models are equivalent, one can construct the linear operators implementing topological invariance in Levin–Wen models from a state-sum on particular space-time cellulations.
In this section, we introduce microscopic models for gapped topological phases on manifolds with boundaries. We mainly use the string-net picture, but occasionally refer to the space-time picture where we find it more instructive. An overview of the space-time picture can be found in Appendix D. First, we introduce the bulk degrees of freedom and how states hosting exact topological invariance are constructed based on a spherical fusion category . Secondly, we extend the models to boundaries and show how the model is constrained by the bulk data close to the boundary, leading to a description of the boundary in terms of a -module category.
A. Bulk
A microscopic (topological) model is defined on a framed trivalent graph (tessellating some two-dimensional manifold) with local Hilbert spaces on each edge. A framed graph has “flags” on each edge pointing perpendicular to it. The orientations have to be chosen such that the flags do not point in the same direction around any vertex. This induces a local ordering of the faces around any vertex by the number of flags pointing into the faces. Analogously, one can think of such a framing as a branching structure on the dual triangulation, see for example (19).
Different sequences of F-moves evaluating to the same transformation on of the graph have to compose to the same map on the associated vector spaces. In particular, the above diagram has to commute, i.e., the composite map corresponding to the top path has to evaluate to the same as composing the maps corresponding to the bottom path. The resulting condition on the F-symbols is called pentagon equation.
Different sequences of F-moves evaluating to the same transformation on of the graph have to compose to the same map on the associated vector spaces. In particular, the above diagram has to commute, i.e., the composite map corresponding to the top path has to evaluate to the same as composing the maps corresponding to the bottom path. The resulting condition on the F-symbols is called pentagon equation.
Strictly speaking, the pentagon equation in Eq. (22) is not the only constraint to the F-symbols. Firstly, for complete topological invariance, we would need to impose the move in Eq. (23) for all possible branching structure configurations. Equivalently, one can add simpler auxiliary axioms specifically targeted to change the branching structure.5 Secondly, for a physical model we have to impose Hermiticity/unitarity, which means that all triangulations carry an orientation, and orientation reversal equals complex conjugation.
A spherical fusion category 26 is the mathematical object giving all the data for a consistent definition of topological moves in the bulk. The topological (ground) space on a given manifold is modeled by the space of all labeled modulo topological moves. In practice, one chooses a minimal reference cellulation whose labelings form the basis of the associated vector space. One can use topological moves to relate any (basis) state on a given cellulation to an equivalent one on the reference cellulation.
B. Boundaries and domain walls
Different sequences of L- and F-moves evaluating to the same transformation on of the graph have to compose to the same map on the associated vector spaces. In particular, the above diagram has to commute, i.e., the composite map corresponding to the top path has to evaluate to the same as composing the maps corresponding to the bottom path. The resulting condition on the L- and F-symbols is called boundary pentagon equation.
Different sequences of L- and F-moves evaluating to the same transformation on of the graph have to compose to the same map on the associated vector spaces. In particular, the above diagram has to commute, i.e., the composite map corresponding to the top path has to evaluate to the same as composing the maps corresponding to the bottom path. The resulting condition on the L- and F-symbols is called boundary pentagon equation.
Taken together, the mathematical structure describing the bulk-boundary microscopics is a -module category. Given and module categories associated to (distinct) boundaries, we can now describe a topologically ordered ground state on manifolds with boundaries by the space of its cellulations modulo topological moves. The exact topological invariance allows us to work with a minimal reference cellulation and use moves to relate a state on a different cellulation to an equivalent one on the chosen reference cellulation.
1. Domain walls and the folding trick
Via the folding trick, a - domain wall, classified by a - bimodule, is equivalent to a module, defined by L symbols of the above form. In Eq. (33) we give the equation relating the data defining the bimodule to the boundary data after the fold.
Via the folding trick, a - domain wall, classified by a - bimodule, is equivalent to a module, defined by L symbols of the above form. In Eq. (33) we give the equation relating the data defining the bimodule to the boundary data after the fold.
C. Topological gauge theory models
Boundaries in topological gauge theory models correspond to module categories of Vecω(G) and have been studied in detail in the mathematical literature and classified by a (possibly twisted) group algebra of a subgroup H ⊂ G on which the input three-cocycle ω is cohomologically trivial.8,15 In this section, we see how topological boundaries are classified in our model and connect it to microscopically different, but equivalent, classifications in the literature. In particular, we explain how to model a boundary associated to a twisted group algebra – a subgroup H and a two-cocycle ψ on it – in our calculations.
D. Models for Abelian phases
E. Examples
In this section, we give some examples of the input data for Vecω(G) models with boundary.
1.
Let us look at possible boundaries of such a bulk model. For simplicity, we take N = p prime. In this case, only has two subgroups, H0 = {0} and . Both H1 and H2 only have trivial two-cocycles, i.e., only one potential boundary associated to either of them. In the untwisted case both give rise to a boundary and correspond to rough and smooth boundaries of Kitaev’s toric code on qupits.9,17 In the twisted case, for n ≠ 0, becomes cohomologically trivial only on H0. Hence, these models only have one “standard” boundary.
2.
The three-cocycle in Eq. (48) becomes cohomologically trivial on H0 and the isomorphic subgroups H1, H2 and H3,l ∀l. All of them have no non-trivial two-cocycle class which gives 3 + l inequivalent topological boundaries.
3. Vecω(S3)
S3 has four subgroups up to conjugation, H0 = {e}, , and HG = S3. All of them have a trivial second cohomology group.8,29 Hence, there is one boundary type associated to each of these three subgroups in the untwisted case. For a twisted bulk model, only the subgroups on which the three-cocycle in Eq. (51) is trivial, define a consistent boundary. Specifically, ωp is cohomologically trivial on Hr for p = 0, 3, and is trivial on Ht for p = 0, 2, 4.
IV. BULK ANYONS: TUBE ALGEBRA
In this section, we will revisit the question of how to add anyons to the fixed-point models from Sec. III A. Anyons are (irreducible) point-defects in the bulk of a topological phase. Their world-lines live in a three-dimensional space-time and are equipped with compatible fusion and braiding structure. In mathematical terms, anyons form the simple objects in a (unitary) modular tensor category (U)MTC. In fact, all their defining data can be calculated with fixed-point models. In this manuscript, we mainly focus on finding the anyons themselves, i.e., the set of irreducible sub-spaces of the point-defects and comment shortly on how to derive the fusion and braiding data with similar methods in Sec. IV B.
A. General fixed-point models
Anyons are point-like topological defects in the bulk, and are known to be characterized by string-nets on an annulus, respectively a “tube,” (0, 1) × S1. In fact, the associated vector space can be equipped with a multiplicative action on itself, rendering it an algebra. The irreducible representations of this tube algebra30–32 can be associated to the simple objects in the unitary modular tensor category (UMTC) describing the bulk anyons. In this section, we will illustrate this concept by first calculating the irreducible sub-spaces of the tube algebra in topological gauge theory models and sketch how to obtain a consistent fusion and braiding on them.
The multiplication of two tube algebra basis elements (a,b,c,d)T and (a′,b′,c′,d′)T is defined via gluing the two associated string diagrams together (left) and using F-moves to reduce it to the cellulation on the right. The phase acquired by the sequence of moves can be derived by evaluating the space-time complex that maps the two dual triangulations to each other, which is composed of three tetrahedra (middle). Note that the front- and the back-side edges of the space-time complex above are identified.
The multiplication of two tube algebra basis elements (a,b,c,d)T and (a′,b′,c′,d′)T is defined via gluing the two associated string diagrams together (left) and using F-moves to reduce it to the cellulation on the right. The phase acquired by the sequence of moves can be derived by evaluating the space-time complex that maps the two dual triangulations to each other, which is composed of three tetrahedra (middle). Note that the front- and the back-side edges of the space-time complex above are identified.
B. Fusion and braiding in the bulk
The anyons in topological fixed-point models form simple objects of a unitary modular tensor category (UMTC). This means they are equipped with additional data/topological quantum numbers related to their fusion and braiding. In fact, this data together is known as the Drinfeld center of the fusion category defining the lattice model. Although it is outside of the scope of this paper to compute the full center, i.e., the fusion and braiding data, we want to comment on how they can be calculated with similar techniques that we have already introduced. For related discussions we refer to Appendix D and Refs. 6 and 30.
Topological fixed-point models allow for a direct calculation of both R and F-symbols.33 In Appendix D, we give an explicit prescription of how to obtain the F-symbols and give a formula for the case of a Vecω(G) model. Let us here shortly lay out how to obtain the R-symbols with microscopic models. For this we again consider the vector space F defined via a tessellated pair of pants. This space represents the fusion space of three bulk anyons. The half-braiding acts on this vertex by interchanging two of its legs as in Eq. (57). This induces a non-trivial action on the cellulation. In the same spirit as before we can use local moves to map it back to the original cellulation and thereby define an action on F. To express this action as a tensor in the basis of anyons, i.e., irreducible representation of T, we have to project the R-action onto this associated sub-spaces by precomposing it with a triple of central projectors . Similarly, one can calculate the F-symbols from microscopic models by mapping between the two ways of decomposing a sphere with four holes into two pairs of pants.
C. Topological gauge theory models
D. Tube algebra for Abelian anyons
E. Examples
In this section, we will first give two examples for Abelian models that are representative for any Abelian lattice model. To illustrate the generic procedure of explicitly finding the central idempotents also for non-Abelian models we further discuss twisted versions of a S3 lattice model. For any Vecω(G) model, the tube algebra is diagonalized by the irreducible β-projective representations of the centralizers of all the conjugacy classes of G.
1.
2.
3. Vecω(S3)
In this section, we will first consider the tube algebra of an untwisted Vec(S3) model in detail and then sketch how the twisting by a three-cocycle affects the tube algebra. Note that the untwisted model has been studied in various contexts in the past.8,29,37
V. BOUNDARY ANYONS: SEMI-TUBE ALGEBRA
After having modeled point-defects (anyons) in the bulk, we continue with point-defects at the boundary. Even though these defects are not equipped with a braiding structure (they can only move along a boundary), we call them boundary anyons. However, they form a fusion category. A framework to calculate the fusion multiplicities thereof was introduced in Refs. 8 and 10 as “vertical fusion.” In fact, the boundary anyons have to form a Morita equivalent category to the one that defines the bulk strings.29
A. General fixed-point models
The multiplication of two basis elements in S is defined via gluing the two associated semi-tube string diagrams together (left) and using L (and F) moves to reduce it to the cellulation on the right. The associated phase can be derived by considering the space-time complex that maps from the initial to the final cellulation.
The multiplication of two basis elements in S is defined via gluing the two associated semi-tube string diagrams together (left) and using L (and F) moves to reduce it to the cellulation on the right. The associated phase can be derived by considering the space-time complex that maps from the initial to the final cellulation.
B. Topological gauge theory models
The anyons on the boundary correspond to the irreducible sub-spaces of S and we again want to find the associated indecomposable central idempotents in S.
Consider two pairs of cosets (α1, β1), (α2, β2) ∈ Sx represented by (a1, b1), (a2, b2) ∈ G. Since both pairs define the same double coset, there exist h, h′ ∈ H such that . Hence, there exists a group element, namely, , such that g ⊳ α1 = α2 and g ⊳ β1 = β2. So any two coset pairs (α1, β1), (α2, β2) ∈ Sx are related via the action of some g ∈ G.
For the trivial double coset x = H, Kx = H.
For H normal in G, Kx = H ∀x.
This result already appeared in Ref. 16 in the derivation of boundary defects in untwisted Quantum Double Models. In the twisted case, boundary anyons are labeled by irreducible Ψα,β-projective representations of Kx instead of linear ones. In the following, if not stated otherwise, we label these representations with κx.
C. Semi-tube algebra with Abelian anyons
In this section, we will consider models for which both the fusion of the bulk and the boundary anyons is Abelian. As we have seen, the bulk is then defined by a finite Abelian group G with N cyclic factors and a three-cocycle cohomologous to a product of type-I and type-II cocycles. The boundary is defined by a subgroup H on which the bulk three-cocycle is cohomologically trivial. Since G is an Abelian group any subgroup H is normal and there is a one-to-one correspondence of double cosets and cosets. In particular, α−1β = x implies that we can rewrite β = αx and the sum over the two coset labels in Eq. (86) reduces to one. Moreover, Kx = H for all x such that we can omit the x-label from the irrep label κx.
D. Examples
In this subsection we will sketch the construction for some exemplary cases. First, we note how the form reduces in the two extreme cases where the subgroup is one of the two trivial subgroups. Then, we will focus on small Abelian groups and as a simple example for an non-Abelian model we will give a full description of how to obtain the central idempotents in the case of G = S3. In particular, we will see how Eq. (85) helps significantly in constructing the stabilizer group whose irreducible representations labels parts of the boundary anyons.
1. Trivial subgroups
2.
3.
Consider the type-II cocycle on from Eq. (48). The cases of the boundary given by a trivial and full subgroups are discussed in Sec. V D 1. Note that the latter only defines a boundary for n12 = 0. The three types of subgroups H1, H2, from Sec. III E 3 define valid boundaries for any n12.
4. Vecω(S3)
As an examplary case for non-Abelian models, consider G = S3. It has four conjugacy classes of subgroups, see Sec. III E 3. The trivial subgroup H0 = {e} defines a boundary for any choice of bulk three-cocycle. In this case, the central idempotents of S the take the form of Eq. (89).
Let us now consider the more subtle case of the non-normal subgroup . It defines a boundary of a bulk model defined by the three-cocycle ωp for p = 0, 2, 4 [see Eq. (51)]. There are two double cosets x ∈ Ht\G/Ht = {{e, t} = Ht, {r, r2, tr, tr2} = HtrHt}. We obtain the stabilizer subgroup of both double cosets with Eq. (85). For the trivial double coset it is given by Ht itself whose irreducible representations are labeled by . For the non-trivial double coset x = HtrHt we pick the representative r ∈ S3. Noting that rHtr−1 = ⟨tr⟩ only shares the identity element with Ht, the stabilizer group for this double coset is the trivial subgroup {e}. Hence, we find only one boundary anyon for x = HtrHt.
VI. BULK-TO-BOUNDARY FUSION EVENTS
After defining our model in the bulk of a spacial two-manifold and on its boundary, we modeled point-like defects, namely, anyons, both in the bulk and on the boundary. In this section, we will look at fusion events between bulk and boundary anyons, which in space-time can be interpreted as point defects on the boundary where a boundary anyon world-line meets a (bulk) anyon world-line. In space such a fusion event can be interpreted as the process of moving a bulk anyon to the boundary and turning it into a boundary anyon.
A. Bimodule construction in general fixed-point models
The S-action on C is defined by gluing a semi-tube element onto a C element from the top and using local moves to reduce the cellulation back to the reference one (95). These moves can be represented by the space-time complex shown above. It decomposes into three bulk tetrahedra and two boundary triangles.
The S-action on C is defined by gluing a semi-tube element onto a C element from the top and using local moves to reduce the cellulation back to the reference one (95). These moves can be represented by the space-time complex shown above. It decomposes into three bulk tetrahedra and two boundary triangles.
B. Bulk-boundary fusion in gauge theory models
Equation (105) describes the fusion events at the boundary of any topological twisted gauge theory model. The condensation formula in Ref. 12 can be seen as a special case where the bulk three-cocycle is trivial and the boundary defect is set to the trivial one. Especially for non-Abelian models this formula gives insight into how non-Abelian bulk anyons split into boundary anyons when approaching the boundary.
1. Abelian models
C. Calculating Lagrangian algebras in gauge theory models
1.
As first example we consider the simplest class of Abelian models, when , the cyclic group of order N. Its character function reads . The bulk can be twisted with a type-I three-cocycle of the form of Eq. (47) giving rise to .
2.
Consider the group with the character function and a bulk twist of type-II ωq from Eq. (48), giving rise to .
3. Vecω(S3)
D. Condensation domain walls for Abelian phases
It is known that any Abelian twisted quantum double of a finite group can be obtained from boson condensation in an untwisted quantum double of a group of larger cardinality.36,42 The condensation process can be viewed as a non-invertible domain wall from the parent phase (an untwisted quantum double of an Abelian group) to the condensate (the twisted quantum double in question). In this section, we illustrate how such a domain wall is described on the microscopic level in two exemplary cases.
1. Condensation domain walls between type-I twisted and untwisted quantum doubles
The simplest examples of the above condensation is the double semion phase. On the one hand, it is realized by the twisted quantum double of . On the other hand it can be described as the phase obtained when condensing e2 m2 in the toric code, the (untwisted) quantum double of . More generally, a quantum double twisted by a type-I cocycle [see Eq. (47)] is equivalent to a condensate of a quantum double of (where eNmN is condensed). Such a condensation process can be viewed as a non-invertible domain wall between the parent phase and the condensate . In this section, we will see how these condensation domain walls can be obtained from our microscopic models. Interestingly, we see how the non-trivial type-I cocycle of becomes trivial when considered as a cocycle of a larger subgroup of .
Via the folding trick (see Sec. III B 1), a domain wall between an untwisted and a quantum double twisted by a type-I cocycle [see Eq. (47)] corresponds to a boundary of a model, where is trivial on and of the form of Eq. (47) on . In the following we show that the subgroup defines a valid boundary and that it corresponds to the condensation domain wall discussed in the previous paragraph when unfolded.
Fusion multiplicities at the the domain wall implementing the condensation from a toric code to the double-semion phase. We see that e2 m2 is condensed since it fuses with the vacuum charge in the double-semion phase. Moreover, for a fixed double-semion anyon, the two toric-code anyons with which it has non-zero fusion multiplicity differ by e2 m2. An all-zero row shows that the associated toric-code anyon is confined to one side of the domain wall, i.e., there is no valid fusion event with any of the double-semion anyons.
. | 1 . | b . | . | s . |
---|---|---|---|---|
1 = (0, 0) | 1 | 0 | 0 | 0 |
e = (0, 1) | 0 | 0 | 0 | 0 |
e2 = (0, 2) | 0 | 1 | 0 | 0 |
e3 = (0, 3) | 0 | 0 | 0 | 0 |
m = (1, 0) | 0 | 0 | 0 | 0 |
me = (1, 1) | 0 | 0 | 0 | 1 |
me2 = (1, 2) | 0 | 0 | 0 | 0 |
me3 = (1, 3) | 0 | 0 | 1 | 0 |
m2 = (2, 0) | 0 | 1 | 0 | 0 |
m2e = (2, 1) | 0 | 0 | 0 | 0 |
m2e2 = (2, 2) | 1 | 0 | 0 | 0 |
m2e3 = (2, 3) | 0 | 0 | 0 | 0 |
m3 = (3, 0) | 0 | 0 | 0 | 0 |
m3e = (3, 1) | 0 | 0 | 1 | 0 |
m3e2 = (3, 2) | 0 | 0 | 0 | 0 |
m3e3 = (3, 3) | 0 | 0 | 0 | 1 |
. | 1 . | b . | . | s . |
---|---|---|---|---|
1 = (0, 0) | 1 | 0 | 0 | 0 |
e = (0, 1) | 0 | 0 | 0 | 0 |
e2 = (0, 2) | 0 | 1 | 0 | 0 |
e3 = (0, 3) | 0 | 0 | 0 | 0 |
m = (1, 0) | 0 | 0 | 0 | 0 |
me = (1, 1) | 0 | 0 | 0 | 1 |
me2 = (1, 2) | 0 | 0 | 0 | 0 |
me3 = (1, 3) | 0 | 0 | 1 | 0 |
m2 = (2, 0) | 0 | 1 | 0 | 0 |
m2e = (2, 1) | 0 | 0 | 0 | 0 |
m2e2 = (2, 2) | 1 | 0 | 0 | 0 |
m2e3 = (2, 3) | 0 | 0 | 0 | 0 |
m3 = (3, 0) | 0 | 0 | 0 | 0 |
m3e = (3, 1) | 0 | 0 | 1 | 0 |
m3e2 = (3, 2) | 0 | 0 | 0 | 0 |
m3e3 = (3, 3) | 0 | 0 | 0 | 1 |
2. Condensation domain walls between fully twisted and untwisted quantum doubles
As a second example we consider the “fully twisted” quantum double of , i.e., we consider a three-cocycle cohomologous to a product of type-I cocycles on both factors, see Eq. (47), and the type-II cocycle on the pair, see Eq. (48), with n1 = n2 = n12 =: n. For N = 2, it describes the six-semion phase where all elementary fluxes have semionic self-exchange statistics, see Ref. 36. This twisted quantum double can also be reppresented as a non-trivial condensate of the untwisted quantum double of .
Snippet of tunneling table from two Toric Codes to six-semion model for and trivial two-cocycle on it (top) and m = 1, 2, 3 cocycle on it (from top to bottom) and m = 3, see Eq. (49) with p = N2. m = 2 seems to have the right condensation pattern of and two inequivalent semions m2e1e2 and m1e1 get mapped to valid generators of the six-semion model.
. | 1 . | b1 . | b2 . | b3 . | . | s1 . | . | s2 . | s3 . | . | s3 . | s4 . | s5 . | . | . | s6 . |
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
1 = (0, 0, 0, 0) | 1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 |
m2e1e2 = (0, 1, 1, 1) | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 |
1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
0 | 0 | 1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
0 | 0 | 0 | 0 | 0 | 1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
m1e1 = (1, 0, 1, 0) | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 1 | 0 | 0 | 0 | 0 | 0 |
1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 1 | 0 | 0 | 0 | 0 | |
⋮ |
. | 1 . | b1 . | b2 . | b3 . | . | s1 . | . | s2 . | s3 . | . | s3 . | s4 . | s5 . | . | . | s6 . |
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
1 = (0, 0, 0, 0) | 1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 |
m2e1e2 = (0, 1, 1, 1) | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 |
1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
0 | 0 | 1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
0 | 0 | 0 | 0 | 0 | 1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
m1e1 = (1, 0, 1, 0) | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 1 | 0 | 0 | 0 | 0 | 0 |
1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 1 | 0 | 0 | 0 | 0 | |
⋮ |
E. Non-Abelian islands in Abelian phases
Topological stabilizer-based quantum error correction (QEC) is believed to be a promising framework to protect a logical qubit against (local) noise. Any Abelian phase can be used to construct a topological stabilizer code.35,36 In such codes, the logical Pauli operators can be associated with certain (non-trivial) loops of anyon ribbon operators. However, one cannot topologically protect a universal gate set with an Abelian phase alone.43,44 One approach to achieve a topologically protected universal gate set is to go beyond stabilizer-based topological QEC and use non-Abelian phases. However, even though first analysis show that – in principle – non-Abelian QEC is possible,45,46 it does not appear to be the best approach to simply store a qubit and perform simple operations, like Clifford gates. In Ref. 37, Laubscher et al. have introduced the concept of non-Abelian islands within an Abelian phase. The authors construct a protocol that allows to teleport a qubit encoded within punctures of an untwisted model into a puncture-encoded qubit within an untwisted S3 model (and vice versa). This allows to perform a non-Clifford gate within the S3 phase that can be teleported back into the phase. In this section, we show that the same method can be used to interface twisted models on a microscopic level by constructing the associated tunneling tables from Eq. (107).
In the following, consider puncture-encoded qubits within a topological phase, as in Ref. 37. To be able to teleport logical information from one phase to the other, one has to interface the two codes with a suitable domain wall. It has to tunnel exactly the right anyons whose ribbon operators span the logical Pauli group. To connect the macroscopic physics described by the tunneling of anyons with a microscopic model, Eq. (105) is of essence. Starting from a microscopic description in terms of a subgroup and a two-cocycle on it, it allows to check which domain walls can be used to teleport logical information from side of the domain wall to the other. In this section, we show how twisted non-Abelian models can be interfaced with twisted Abelian models to teleport a qubit, respectively qudit. In particular, we use Eq. (107) to construct tunneling tables that show which anyons can be transported though a given domain wall with local operations. We illustrate the concept via two examples, namely, how different twisted S3 models can be interfaced with a twisted (double-semion) model and a twisted model. After that we comment on how to find suitable microscopic models for more general phases that allow for teleportation of logical information from an Abelian to a non-Abelian phase. Moreover, we comment on the origin of universal quantum computation with non-Abelian islands in an Abelian codes.
Imagine we want to interface a non-Abelian model, for example Vecω(S3), with an Abelian one, for example . Any domain wall between such two models is equivalent to a boundary of a stacked model, for example , see Sec. III B 1. In the following, we will illustrate how this folding trick is used to construct the tunneling tables in two examples, where Vecω(S3) is interfaced with for N = 2, 3. In particular, the three-cocycles ω on S3 have to be in different cohomology classes to tunnel anyons non-trivially through the domain wall.
1.
Tunneling table for a domain wall defined by .
. | 1 . | b . | s . | . |
---|---|---|---|---|
1 | 0 | 0 | 0 | |
0 | 1 | 0 | 0 | |
1 | 1 | 0 | 0 | |
0 | 0 | 0 | 0 | |
0 | 0 | 0 | 0 | |
0 | 0 | 0 | 0 | |
0 | 0 | 1 | 0 | |
0 | 0 | 0 | 1 |
. | 1 . | b . | s . | . |
---|---|---|---|---|
1 | 0 | 0 | 0 | |
0 | 1 | 0 | 0 | |
1 | 1 | 0 | 0 | |
0 | 0 | 0 | 0 | |
0 | 0 | 0 | 0 | |
0 | 0 | 0 | 0 | |
0 | 0 | 1 | 0 | |
0 | 0 | 0 | 1 |
Note that the generator of H is supported on S3 as well as on which in turn makes the domain wall (partly) transparent.
2.
Tunneling table for a domain wall defined by .
. | (0,0) . | (0,1) . | (0,2) . | (1,0) . | (1,1) . | (1,2) . | (2,0) . | (2,1) . | (2,2) . |
---|---|---|---|---|---|---|---|---|---|
1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
0 | 1 | 1 | 0 | 0 | 0 | 0 | 0 | 0 | |
0 | 0 | 0 | 0 | 1 | 0 | 0 | 0 | 1 | |
0 | 0 | 0 | 0 | 0 | 1 | 1 | 0 | 0 | |
0 | 0 | 0 | 1 | 0 | 0 | 0 | 1 | 0 | |
0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 |
. | (0,0) . | (0,1) . | (0,2) . | (1,0) . | (1,1) . | (1,2) . | (2,0) . | (2,1) . | (2,2) . |
---|---|---|---|---|---|---|---|---|---|
1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
1 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
0 | 1 | 1 | 0 | 0 | 0 | 0 | 0 | 0 | |
0 | 0 | 0 | 0 | 1 | 0 | 0 | 0 | 1 | |
0 | 0 | 0 | 0 | 0 | 1 | 1 | 0 | 0 | |
0 | 0 | 0 | 1 | 0 | 0 | 0 | 1 | 0 | |
0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | |
0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 |
3. Remarks on non-Abelian islands for quantum computation
In order to use islands of Non-Abelian codes within an Abelian code, one needs a way to transfer logical information from one code to the other fault-tolerantly. As layed out in Ref. 37 this can be achieved by a certain code-deformation protocol that effectively moves holes (that encode some logical information) through an interfacing region from the Non-Abelian code to the Abelian code. In fact, this interface can be understood as a domain wall between the two codes and as such, it is described by tunneling tables as the ones calculated above. To be able to transfer logical information, the domain wall has to be transparent for a subset of the anyons of both models. Consider interfacing two group theoretical models, each defined by a finite group G(G′) and a three-cocycle. In this case, the interface is defined by a subgroup H of G × G′ (on which the three-cocycle is trivial) and a two-cocycle ψ on H. From Eq. (107) we can derive conditions on (H, ψ) for the associated domain wall to be (semi-)transparent. We observe that there are two ways in which one can “couple” the two models non-trivially:
Coupling via the subgroup: The subgroup H cannot be generated a set of generators each of which is of the form (g, 1G′) or (1G, g′). We say, H does not factorize over the two sides of the domain wall.
Coupling via a two-cocycle: Let H have two subfactors H1 ⊆ G ×{1G′} and H2 ⊆{1G}× G′. A two-cocycle that is non-trivial on H1 × H2 but not cohologous to a product of two-cocycles on H1 and H2 individually.
Note that a (semi-)transparent domain wall can also fulfill both of the above conditions. The examples considered above are all transparent due to condition 1. For example, S3 has both a and a subgroup which allow it to be interfaced with and models. The two-cocycles fulfilling condition 2 are in most cases 2-coycles on an (Abelian) subgroup isomorphic to for some pair of integers N, M that are not coprime.47 The simplest example of a domain wall that is transparent for that reason is the e − m duality domain wall in a model. Microscopically, it is defined by subgroup of the folded model and the non-trivial two-cocycle on it.
Reference 37 nicely describes a computational scheme based on Non-Abelian island for the simplest case of Vec(S3) islands in a model and can be straightforwardly generalized. It relies on preparing an auxiliary non-stabilizer state within the Non-Abelian patches, encoded into certain punctures. Since the preparation involves topological charge measurements, universality can be achieved even in group-theoretical MTCs that wouldn’t be universal by braiding alone.48
VII. CONCLUSION AND OUTLOOK
In this manuscript, we have explored bulk-to-boundary anyon fusion events in non-chiral topologically ordered quantum models, aimed at understanding how anyonic defects interact with external defects such as boundaries or domain walls. It has been motivated by considerations both in the study of quantum phases of matter and quantum information theory. Specifically, we have calculated bulk-to-boundary fusion multiplicities in topological fixed-point models in 2 + 1 space-time dimensions. Our framework allows for a step-by-step calculation of the fusion multiplicities in all such models. Apart from the calculation of projective irreducible representations of the G-subgroups which define the action in the tube and semi-tube algebras, the calculation only involves the evaluation of linear expressions. The fusion multiplicities allow to characterize the behavior of anyonic bulk excitations when approaching a boundary and – via the folding trick – to calculate the effect an anyon has on a domain wall when moved through it.
At the core of our construction lies a bimodule that is a representation of the tube algebra defining the bulk anyons as well as the semi-tube algebra defining the boundary anyons. We have defined this bimodule for any fixed-point model and explicitly derived a closed formula for the fusion multiplicities in the subclass of topological lattice gauge theories, where the fusion category defining the bulk is given by Vecω(G). We have used this formula to calculate Lagrangian algebras in various gauge theory models without the need to explicitly solve the consistency conditions defining a Lagrangian algebra. This is particularly useful for more involved non-Abelian topological phases where finding Lagrangian algebras is more intricate. We showcase this in the derivation of the Lagrangian algebras for Vecω(S3) models with non-trivial three-cocycles. Moreover, using the folding trick we can use bulk-to-boundary fusion events to study the tunneling of anyons through domain walls. In particular, our formula allows to keep track of which anyon is left behind at a domain wall when moving a bulk anyon from one side to the other. As a proof of principle, we have calculated the fusion multiplicities of a special class of non-invertible domain walls which implement anyon condensation from Abelian untwisted to twisted quantum doubles. This is a well-known transition on the level of the anyon models36,42 but as far as we know we are the first to give a microscopic description of the corresponding domain walls in space-time. Moreover, we have shown how to interface Abelian twisted quantum double phases with non-Abelian ones on a microscopic level.
In our construction, we have observed that both bulk and boundary anyons are characterized by a special type of algebra which we call twisted group algebra with action. We give a step-by-step recipe to derive the irreducible representations of such algebras from simpler projective representations of the subgroups stabilizing the action. In fact, any sort of line-like defects in space-time in a Vecω(G) state-sum model will be classified by irreducible representations of a certain group algebra with action. Hence, our techniques can be used to study these defects and their interaction with membrane-like defects in 3 + 1-dimensional models. Generalizing this recipe to groupoid-like algebras (see Ref. 31) and thereby extending the methodology beyond twisted quantum double models might be an interesting avenue for further research.
Together with previous work8–11,49 the fusion multiplicities calculated in this paper contributes to the algebraic description of the anyons in topological models with boundaries and domain walls. An understanding of similar depth of defects in higher-dimensional models is lacking. Further research directions can include the application of the techniques used in this paper to line-defects in 3 + 1-dimensional models and extending the techniques to further understand the interaction of defects of different (co-)dimensions in higher-dimensional models.
Again, our work is not only interesting from the perspective of the mathematically minded study of topological phases of matter but it has also important applications in different areas of physics. Above all, any practical topological quantum error correction (QEC) scheme involves boundaries or domain walls in one way of the other. On the one hand, stabilizer-based topological QEC can be understood as storing a qudit in the ground space of an Abelian topological phase modeled by a Vecω(G) fixed-point model.35,36
On the other hand, going beyond stabilizer-based approaches allows to natively perform universal topological quantum computation.45 Given the overhead in resources of protocols that uplift non-universal stabilizer-based approaches of quantum computing to universal ones by means of magic-state distillation,4,50 such an avenue may well have its benefits. In both cases a thorough understanding of the interaction of anyons with other types of defects is important. For example, finding the logical operators in a given planar code including boundaries and domain walls reduces to characterizing which anyons can condense at which boundary. Our work shows how to calculate these quantities in the most general case, in particular we extend the results of Ref. 12 to twisted quantum doubles. The bulk-to-boundary fusion multiplicities can also be used to study computational protocols including boundaries and domain walls.18,37,51 This includes lattice surgery20,52 schemes in Abelian topological codes and allows for systematic study of the computational possibilities of a given code via anyon condensation.18 Moreover, domain walls between non-Abelian phases und Abelian ones can be used to generalize the scheme presented in Ref. 37 where a partly-transparent domain wall between a Vec(S3) and a phase is used to teleport a topologically encoded qubit from one model to the other. Starting there, it will be interesting to investigate universal computing schemes based on twisted quantum doubles, particularly when combined with a Pauli-based description of the Abelian phase from Ref. 36.
Lastly, to fully understand the computational capabilities of non-Abelian quantum error correction45,46 we want to investigate domain walls between non-Abelian phases to see how external defects can extend non-Abelian codes. This is again partially motivated by the quest to find schemes for quantum computing without the need for magic state distillation. These few examples should illustrate the wide applicability of bulk-to-boundary fusion events, especially in topological QEC. We hope our work sparks inspiration to develop new QEC and computing protocols based on more exotic topological phases. On a higher level, this work is aimed at contributing to building new interfaces between quantum information theory and mathematical condensed matter physics which seems a mutually inspiring intersection.
ACKNOWLEDGMENTS
The authors would like to thank Markus Kesselring for fruitful discussions on the application of our work to topological QEC schemes. J.C.M.d.l.F. also wants to thank Tyler Ellison for discussions on condensation domain walls. This work is supported by the DFG (Grant No. CRC 183) and the BMBF (RealistiQ, QSolid), and the BMWK (PlanQK).
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Julio C. Magdalena de la Fuente: Conceptualization (equal); Formal analysis (equal); Methodology (equal); Writing – original draft (lead); Writing – review & editing (lead). Jens Eisert: Conceptualization (supporting); Funding acquisition (lead); Project administration (equal); Writing – review & editing (supporting). Andreas Bauer: Conceptualization (equal); Methodology (equal); Supervision (supporting); Writing – original draft (equal); Writing – review & editing (equal).
DATA AVAILABILITY
Data sharing is not applicable to this article as no new data were created or analyzed in this study.
APPENDIX A: EXAMPLES OF FUSION MULTIPLICITIES
In this appendix, we summarize the fusion multiplicities at boundaries of an untwisted and Vecω(S3) models with different three-cocycles.
1.
For the model, we first show the table of fusion multiplicities at the trivial boundary defined by the order 1 subgroup H = {(0, 0)} (Table V). In Tables VI and VII we give the fusion multiplicities at the boundaries defined by the subgroups ⟨(0, 1)⟩ and ⟨(1, 1)⟩. The only subgroup with non-trivial two-cocycles is . It has two two-cocycle classes and hence two boundaries associated to it. The corresponding fusion multiplicities can be found in Table VIII. All these fusion events can either be seen as fusions into the boundary of a model (which is in the phase as the topological color code53) or—equivalently—as domain wall tunneling events between two models (toric code phase).
The fusion multiplicities for all pairs of bulk and boundary anyons for and the standard boundary modeled by the subgroup H = {(0, 0)}. The bulk anyons (rows) are labeled by G×2, the boundary anyons by cosets G/H ≃ G, represented by G elements. The fact that they are actually coset labels is marked with an overline.
. | . | . | . | . |
---|---|---|---|---|
[(0, 1), (0, 1)] | 1 | 0 | 0 | 0 |
[(0, 1), (1, 0)] | 1 | 0 | 0 | 0 |
[(0, 1), (1, 1)] | 1 | 0 | 0 | 0 |
[(0, 1), (0, 0)] | 1 | 0 | 0 | 0 |
[(1, 0), (0, 1)] | 0 | 1 | 0 | 0 |
[(1, 0), (1, 0)] | 0 | 1 | 0 | 0 |
[(1, 0), (1, 1)] | 0 | 1 | 0 | 0 |
[(1, 0), (0, 0)] | 0 | 1 | 0 | 0 |
[(1, 1), (0, 1)] | 0 | 0 | 0 | 1 |
[(1, 1), (1, 0)] | 0 | 0 | 0 | 1 |
[(1, 1), (1, 1)] | 0 | 0 | 0 | 1 |
[(1, 1), (0, 0)] | 0 | 0 | 0 | 1 |
[(0, 0), (0, 1)] | 0 | 0 | 1 | 0 |
[(0, 0), (1, 0)] | 0 | 0 | 1 | 0 |
[(0, 0), (1, 1)] | 0 | 0 | 1 | 0 |
[(0, 0), (0, 0)] | 0 | 0 | 1 | 0 |
. | . | . | . | . |
---|---|---|---|---|
[(0, 1), (0, 1)] | 1 | 0 | 0 | 0 |
[(0, 1), (1, 0)] | 1 | 0 | 0 | 0 |
[(0, 1), (1, 1)] | 1 | 0 | 0 | 0 |
[(0, 1), (0, 0)] | 1 | 0 | 0 | 0 |
[(1, 0), (0, 1)] | 0 | 1 | 0 | 0 |
[(1, 0), (1, 0)] | 0 | 1 | 0 | 0 |
[(1, 0), (1, 1)] | 0 | 1 | 0 | 0 |
[(1, 0), (0, 0)] | 0 | 1 | 0 | 0 |
[(1, 1), (0, 1)] | 0 | 0 | 0 | 1 |
[(1, 1), (1, 0)] | 0 | 0 | 0 | 1 |
[(1, 1), (1, 1)] | 0 | 0 | 0 | 1 |
[(1, 1), (0, 0)] | 0 | 0 | 0 | 1 |
[(0, 0), (0, 1)] | 0 | 0 | 1 | 0 |
[(0, 0), (1, 0)] | 0 | 0 | 1 | 0 |
[(0, 0), (1, 1)] | 0 | 0 | 1 | 0 |
[(0, 0), (0, 0)] | 0 | 0 | 1 | 0 |
Fusion multiplicities at the boundary of a model where the bulk is defined by and a trivial three-cocycle and the boundary by the non-trivial subgroup H = ⟨(0, 1)⟩. The bulk anyons (rows) are labeled by G×2 and the boundary anyons (columns) by G/H × G. The coset label marked with an overline.
. | [, (0, 1)] . | [, (0, 0)] . | [, (0, 1)] . | [, (0, 0)] . |
---|---|---|---|---|
[(0, 1), (0, 1)] | 0 | 0 | 1 | 0 |
[(0, 1), (1, 0)] | 0 | 0 | 0 | 1 |
[(0, 1), (1, 1)] | 0 | 0 | 1 | 0 |
[(0, 1), (0, 0)] | 0 | 0 | 0 | 1 |
[(1, 0), (0, 1)] | 1 | 0 | 0 | 0 |
[(1, 0), (1, 0)] | 0 | 1 | 0 | 0 |
[(1, 0), (1, 1)] | 1 | 0 | 0 | 0 |
[(1, 0), (0, 0)] | 0 | 1 | 0 | 0 |
[(1, 1), (0, 1)] | 1 | 0 | 0 | 0 |
[(1, 1), (1, 0)] | 0 | 1 | 0 | 0 |
[(1, 1), (1, 1)] | 1 | 0 | 0 | 0 |
[(1, 1), (0, 0)] | 0 | 1 | 0 | 0 |
[(0, 0), (0, 1)] | 0 | 0 | 1 | 0 |
[(0, 0), (1, 0)] | 0 | 0 | 0 | 1 |
[(0, 0), (1, 1)] | 0 | 0 | 1 | 0 |
[(0, 0), (0, 0)] | 0 | 0 | 0 | 1 |
. | [, (0, 1)] . | [, (0, 0)] . | [, (0, 1)] . | [, (0, 0)] . |
---|---|---|---|---|
[(0, 1), (0, 1)] | 0 | 0 | 1 | 0 |
[(0, 1), (1, 0)] | 0 | 0 | 0 | 1 |
[(0, 1), (1, 1)] | 0 | 0 | 1 | 0 |
[(0, 1), (0, 0)] | 0 | 0 | 0 | 1 |
[(1, 0), (0, 1)] | 1 | 0 | 0 | 0 |
[(1, 0), (1, 0)] | 0 | 1 | 0 | 0 |
[(1, 0), (1, 1)] | 1 | 0 | 0 | 0 |
[(1, 0), (0, 0)] | 0 | 1 | 0 | 0 |
[(1, 1), (0, 1)] | 1 | 0 | 0 | 0 |
[(1, 1), (1, 0)] | 0 | 1 | 0 | 0 |
[(1, 1), (1, 1)] | 1 | 0 | 0 | 0 |
[(1, 1), (0, 0)] | 0 | 1 | 0 | 0 |
[(0, 0), (0, 1)] | 0 | 0 | 1 | 0 |
[(0, 0), (1, 0)] | 0 | 0 | 0 | 1 |
[(0, 0), (1, 1)] | 0 | 0 | 1 | 0 |
[(0, 0), (0, 0)] | 0 | 0 | 0 | 1 |
The bulk-to-boundary fusion multiplicities in a model where the bulk is defined by and a trivial three-cocycle and the boundary by the non-trivial subgroup H = ⟨(1, 1)⟩. The bulk anyons (rows) are labeled by G×2 and the boundary anyons (columns) by G/H × G. The coset label marked with an overline.
. | [, (1, 1)] . | [, (0, 0)] . | [, (1, 1)] . | [, (0, 0)] . |
---|---|---|---|---|
[(0, 1), (0, 1)] | 1 | 0 | 0 | 0 |
[(0, 1), (1, 0)] | 1 | 0 | 0 | 0 |
[(0, 1), (1, 1)] | 0 | 1 | 0 | 0 |
[(0, 1), (0, 0)] | 0 | 1 | 0 | 0 |
[(1, 0), (0, 1)] | 1 | 0 | 0 | 0 |
[(1, 0), (1, 0)] | 1 | 0 | 0 | 0 |
[(1, 0), (1, 1)] | 0 | 1 | 0 | 0 |
[(1, 0), (0, 0)] | 0 | 1 | 0 | 0 |
[(1, 1), (0, 1)] | 0 | 0 | 1 | 0 |
[(1, 1), (1, 0)] | 0 | 0 | 1 | 0 |
[(1, 1), (1, 1)] | 0 | 0 | 0 | 1 |
[(1, 1), (0, 0)] | 0 | 0 | 0 | 1 |
[(0, 0), (0, 1)] | 0 | 0 | 1 | 0 |
[(0, 0), (1, 0)] | 0 | 0 | 1 | 0 |
[(0, 0), (1, 1)] | 0 | 0 | 0 | 1 |
[(0, 0), (0, 0)] | 0 | 0 | 0 | 1 |
. | [, (1, 1)] . | [, (0, 0)] . | [, (1, 1)] . | [, (0, 0)] . |
---|---|---|---|---|
[(0, 1), (0, 1)] | 1 | 0 | 0 | 0 |
[(0, 1), (1, 0)] | 1 | 0 | 0 | 0 |
[(0, 1), (1, 1)] | 0 | 1 | 0 | 0 |
[(0, 1), (0, 0)] | 0 | 1 | 0 | 0 |
[(1, 0), (0, 1)] | 1 | 0 | 0 | 0 |
[(1, 0), (1, 0)] | 1 | 0 | 0 | 0 |
[(1, 0), (1, 1)] | 0 | 1 | 0 | 0 |
[(1, 0), (0, 0)] | 0 | 1 | 0 | 0 |
[(1, 1), (0, 1)] | 0 | 0 | 1 | 0 |
[(1, 1), (1, 0)] | 0 | 0 | 1 | 0 |
[(1, 1), (1, 1)] | 0 | 0 | 0 | 1 |
[(1, 1), (0, 0)] | 0 | 0 | 0 | 1 |
[(0, 0), (0, 1)] | 0 | 0 | 1 | 0 |
[(0, 0), (1, 0)] | 0 | 0 | 1 | 0 |
[(0, 0), (1, 1)] | 0 | 0 | 0 | 1 |
[(0, 0), (0, 0)] | 0 | 0 | 0 | 1 |
with trivial (left) and non-trivial two-cocycle (right).
. | (0, 1) . | (1, 0) . | (1, 1) . | (0, 0) . | . | (0, 1) . | (1, 0) . | (1, 1) . | (0, 0) . |
---|---|---|---|---|---|---|---|---|---|
[(0, 1), (0, 1)] | 0 | 1 | 0 | 0 | [(0, 1), (0, 1)] | 0 | 0 | 1 | 0 |
[(0, 1), (1, 0)] | 1 | 0 | 0 | 0 | [(0, 1), (1, 0)] | 0 | 0 | 0 | 1 |
[(0, 1), (1, 1)] | 0 | 0 | 1 | 0 | [(0, 1), (1, 1)] | 1 | 0 | 0 | 0 |
[(0, 1), (0, 0)] | 0 | 0 | 0 | 1 | [(0, 1), (0, 0)] | 0 | 1 | 0 | 0 |
[(1, 0), (0, 1)] | 0 | 1 | 0 | 0 | [(1, 0), (0, 1)] | 0 | 0 | 0 | 1 |
[(1, 0), (1, 0)] | 1 | 0 | 0 | 0 | [(1, 0), (1, 0)] | 0 | 0 | 1 | 0 |
[(1, 0), (1, 1)] | 0 | 0 | 1 | 0 | [(1, 0), (1, 1)] | 0 | 1 | 0 | 0 |
[(1, 0), (0, 0)] | 0 | 0 | 0 | 1 | [(1, 0), (0, 0)] | 1 | 0 | 0 | 0 |
[(1, 1), (0, 1)] | 0 | 1 | 0 | 0 | [(1, 1), (0, 1)] | 0 | 1 | 0 | 0 |
[(1, 1), (1, 0)] | 1 | 0 | 0 | 0 | [(1, 1), (1, 0)] | 1 | 0 | 0 | 0 |
[(1, 1), (1, 1)] | 0 | 0 | 1 | 0 | [(1, 1), (1, 1)] | 0 | 0 | 0 | 1 |
[(1, 1), (0, 0)] | 0 | 0 | 0 | 1 | [(1, 1), (0, 0)] | 0 | 0 | 1 | 0 |
[(0, 0), (0, 1)] | 0 | 1 | 0 | 0 | [(0, 0), (0, 1)] | 1 | 0 | 0 | 0 |
[(0, 0), (1, 0)] | 1 | 0 | 0 | 0 | [(0, 0), (1, 0)] | 0 | 1 | 0 | 0 |
[(0, 0), (1, 1)] | 0 | 0 | 1 | 0 | [(0, 0), (1, 1)] | 0 | 0 | 1 | 0 |
[(0, 0), (0, 0)] | 0 | 0 | 0 | 1 | [(0, 0), (0, 0)] | 0 | 0 | 0 | 1 |
. | (0, 1) . | (1, 0) . | (1, 1) . | (0, 0) . | . | (0, 1) . | (1, 0) . | (1, 1) . | (0, 0) . |
---|---|---|---|---|---|---|---|---|---|
[(0, 1), (0, 1)] | 0 | 1 | 0 | 0 | [(0, 1), (0, 1)] | 0 | 0 | 1 | 0 |
[(0, 1), (1, 0)] | 1 | 0 | 0 | 0 | [(0, 1), (1, 0)] | 0 | 0 | 0 | 1 |
[(0, 1), (1, 1)] | 0 | 0 | 1 | 0 | [(0, 1), (1, 1)] | 1 | 0 | 0 | 0 |
[(0, 1), (0, 0)] | 0 | 0 | 0 | 1 | [(0, 1), (0, 0)] | 0 | 1 | 0 | 0 |
[(1, 0), (0, 1)] | 0 | 1 | 0 | 0 | [(1, 0), (0, 1)] | 0 | 0 | 0 | 1 |
[(1, 0), (1, 0)] | 1 | 0 | 0 | 0 | [(1, 0), (1, 0)] | 0 | 0 | 1 | 0 |
[(1, 0), (1, 1)] | 0 | 0 | 1 | 0 | [(1, 0), (1, 1)] | 0 | 1 | 0 | 0 |
[(1, 0), (0, 0)] | 0 | 0 | 0 | 1 | [(1, 0), (0, 0)] | 1 | 0 | 0 | 0 |
[(1, 1), (0, 1)] | 0 | 1 | 0 | 0 | [(1, 1), (0, 1)] | 0 | 1 | 0 | 0 |
[(1, 1), (1, 0)] | 1 | 0 | 0 | 0 | [(1, 1), (1, 0)] | 1 | 0 | 0 | 0 |
[(1, 1), (1, 1)] | 0 | 0 | 1 | 0 | [(1, 1), (1, 1)] | 0 | 0 | 0 | 1 |
[(1, 1), (0, 0)] | 0 | 0 | 0 | 1 | [(1, 1), (0, 0)] | 0 | 0 | 1 | 0 |
[(0, 0), (0, 1)] | 0 | 1 | 0 | 0 | [(0, 0), (0, 1)] | 1 | 0 | 0 | 0 |
[(0, 0), (1, 0)] | 1 | 0 | 0 | 0 | [(0, 0), (1, 0)] | 0 | 1 | 0 | 0 |
[(0, 0), (1, 1)] | 0 | 0 | 1 | 0 | [(0, 0), (1, 1)] | 0 | 0 | 1 | 0 |
[(0, 0), (0, 0)] | 0 | 0 | 0 | 1 | [(0, 0), (0, 0)] | 0 | 0 | 0 | 1 |
2. Vecω(S3)
Depending on which three-cocycles we use in Vecω(S3) different subgroups define topological boundaries. In Table IX we give the fusion multiplicities for the cases when Hr defines a boundary. In Table X we contrast it with the fusion multiplicities for the cases where Ht defines a boundary.
All the fusion multiplicities for a model where the bulk is defined by G = S3 and the boundary by the non-trivial subgroup Hr = ⟨r⟩ (see Sec. III E 3). The bulk anyons (rows) are labeled as in Sec. IV E 3 and the boundary anyons by . We show the fusion multiplicites for the three-cocycles ωp [see Eq. (51)] with p = 0, 3. Where they differ we give the value for p = 3 in brackets. These are all models where Hr defines a valid boundary.
. | (Hr, 0) . | (Hr, 1) . | (Hr, 2) . | (HrtHr, 0) . | (HrtHr, 1) . | (HrtHr, 2) . |
---|---|---|---|---|---|---|
1 | 0 | 0 | 0 | 0 | 0 | |
1 | 0 | 0 | 0 | 0 | 0 | |
0 | 1 | 1 | 0 | 0 | 0 | |
2(0) | 0(1) | 0(1) | 0 | 0 | 0 | |
0(1) | 2(0) | 0(1) | 0 | 0 | 0 | |
0(1) | 0(1) | 2(0) | 0 | 0 | 0 | |
0 | 0 | 0 | 1 | 1 | 1 | |
0 | 0 | 0 | 1 | 1 | 1 |
. | (Hr, 0) . | (Hr, 1) . | (Hr, 2) . | (HrtHr, 0) . | (HrtHr, 1) . | (HrtHr, 2) . |
---|---|---|---|---|---|---|
1 | 0 | 0 | 0 | 0 | 0 | |
1 | 0 | 0 | 0 | 0 | 0 | |
0 | 1 | 1 | 0 | 0 | 0 | |
2(0) | 0(1) | 0(1) | 0 | 0 | 0 | |
0(1) | 2(0) | 0(1) | 0 | 0 | 0 | |
0(1) | 0(1) | 2(0) | 0 | 0 | 0 | |
0 | 0 | 0 | 1 | 1 | 1 | |
0 | 0 | 0 | 1 | 1 | 1 |
All the fusion multiplicities for a model where the bulk is defined by G = S3 and the boundary by the non-trivial subgroup Ht = ⟨t⟩ (see Sec. III E 3). The bulk anyons (rows) are labeled as in Sec. IV E 3 and the boundary anyons by Ht\G/Ht and the irreducible representations of the associated stabilizer group. For Ht as the trivial double coset, this is , for HtrHt, the only non-trivial double coset, the stabilizer group is trivial, i.e., . We show the fusion multiplicites for the three-cocycles ωp [see Eq. (51)] with p = 0, 2, 4. Interestingly, they coincide for all of these models. In fact, for other values of p Ht does not define a valid boundary.
. | (Ht, 0) . | (Ht, 1) . | (HtrHt, 0) . |
---|---|---|---|
1 | 0 | 0 | |
0 | 1 | 0 | |
1 | 1 | 0 | |
0 | 0 | 1 | |
0 | 0 | 1 | |
0 | 0 | 1 | |
1 | 0 | 1 | |
0 | 1 | 1 |
. | (Ht, 0) . | (Ht, 1) . | (HtrHt, 0) . |
---|---|---|---|
1 | 0 | 0 | |
0 | 1 | 0 | |
1 | 1 | 0 | |
0 | 0 | 1 | |
0 | 0 | 1 | |
0 | 0 | 1 | |
1 | 0 | 1 | |
0 | 1 | 1 |
APPENDIX B: COHOMOLOGY OF FINITE GROUPS
In this section, we provide an algebraic definition of cohomology theory for finite groups. For a more detailed background, see, for example, Refs. 22 and 23.
[(left) G-module]. Consider a finite group (G, ·). An Abelian group (M, *) together with a (left) group action ⊳: G × M → M, (g, a) ↦ g ⊳ a that ∀ a, b ∈ M, g, h ∈ G fulfills
g ⊳ (a*b) = (g ⊳ a)*(g ⊳ b) and
(g · h) ⊳ a = g ⊳ (h ⊳ a),is called a left G-module. Analogously, a right G-module is defined with a G-action from the right which we denote by ⊲. An Abelian group M equippped with both a G-action from the left and a G′-action from the right is called G-G′-bimodule.
Any Abelian group can be made into a G-module with a trivial group action by defining g ⊳ a = a ∀a ∈ M, g ∈ G.
(n-cochain). Let (M, *) be a G-module. A map ηn: G×n → M is called n-cochain (of G over M). We denote the space of all such n-cochains by . In fact, is a group with the group multiplication inherited from M.
Most importantly, for any n and , , where denotes the trivial (constant) map from G×n to M. We say that the coboundary of a coboundary is trivial. Often, we will shortly write δ = ⊕nδn which can act on any (combination of) cochain(s).
In this work we encounter two sorts of G-modules. First, U(1) with trivial action, and second, U(1)A for a finite set A whose action is determined by a permutation action of G on A. We will distinguish the latter twisted coboundary operator by explicitly writing .
If two n-cocycles are in the same equivalence class in Hn(G, M) we call them cohomologous.
induces a homomorphism on the cohomology groups Hn(G, M) → Hn−1(G, MG) where MG is a G-module whose action is a combination of the action on M plus conjugation of the G-label. Indeed, one can check that . This explains why βg in Sec. IV is a twisted two-cocycle.
In the following, we will give some examples for cohomology groups of Abelian and non-Abelian groups over U(1) as a G-module with trivial action.
The three-cocycles classes in Eq. (B9d) decompose into products of three-cocycle classes of three types: The ones only depending on the single tensor factor (type-I), a pair of factors (type-II) and depending on a triple of factors (type-III). Remarkably, the slant product maps three-cocycles of type I and II to trivial two-cocycles and only three-cocycles of type III to non-trivial ones.
APPENDIX C: THE ISOMORPHISM m(n)

For g1, …, gn ∈ H, m(n)(gn, …, g1, H) = (gn, …, g1). In the case of an Abelian group, any gi ∈ H itself is left invariant by m(n), independent of the remaining arguments.


This isomorphism is of particular importance for our work. It not only defines the L-symbols of boundaries in Sec. III C but also enters in the algebra diagonalization in Sec. II. Moreover, Eq. (C13) shows that the twisted two-cocycle defining the semi-tube algebra Ψα,αx is cohomologous for all α when restricted to H. This is implicitly used in the central idempotents since we sum over α. In Appendix E, we give a geometric interpretation of this isomoprhism as an invertible domain wall on a boundary state-sum.
APPENDIX D: STATE-SUM PICTURE
In this appendix we give a concise and systematic rederivation of what is discussed in the main text, using the language of state-sum models in space-time. We will make use of the notions of extended manifolds/cellulations as defined in Appendix B of Ref. 54. Roughly, an extended manifold is a composite of manifolds-with-boundary called regions of different dimension attached to each other in different ways. The link of a region is the intersection of a small-enough ɛ-sphere around a point within the space normal to that region, and has to be the same for all points of that region. For example, if we have a one-dimensional region embedded into a three-dimensional one, the normal space at a point is a plane, so the link of the one-dimensional region is a circle. To get an extended cellulation, we triangulate the Cartesian product of each region with its link. This triangulation is identified with the boundary triangulation of higher-dimensional regions.
We now associate state-sum variables and weights to different cells of the different regions, and demand their invariance under topology-preserving moves such as Pachner moves. The highest-dimensional region then defines the bulk of a state-sum model, and the lower-dimensional regions define topological boundaries, anyon world-lines, domain walls, and other sorts of defects. In the following, we will present the twisted quantum double model and its boundaries, anyons, etc. in this language.
1. Bulk
2. Boundary
3. Anyons
4. Boundary anyons
5. Ground states
6. Bulk fusion events
7. Bulk-boundary fusion events
8. Bulk anyon F-symbol
APPENDIX E: THIN vs THICK BOUNDARIES
In this appendix, we describe the space-time picture for a different microscopic way of defining a boundary. We will call this the thin boundary state-sum as opposed to the thick boundary described in the main text and in Appendix D. We will first discuss the thin boundary in 1 + 1 dimensions and then in 2 + 1 dimensions, such that the generalization to arbitrary space-time dimensions will be straightforward.
1. 1 + 1 dimensions
Let us now discuss the relation between the thick and thin boundary. Note that in Appendix D, a thick boundary was defined with values in an arbitrary right G-set A at the vertices. In 1 + 1 dimensions, there is a constraint β = α ⊲ g and a weight ψα(a) at every thickened edge as in Eq. (D7). Every set with right G action is isomorphic to a disjoint union of left coset sets H\G for different H. Each H is determined up to conjugation, and the right action on H\G is transitive and given by multiplication with the G-element from the right. Note that physically, boundaries with a transitive action are those which are irreducible, that is, robust to perturbations.



2. 2 + 1 dimensions
Another drawback of the thin boundary is that it is not directly compatible with the more general way of defining boundaries in terms of F and L symbols. That is, there is no analogue of a thin boundary for non-group-cocycle F-symbols or L-symbols.



3. Thin vs thick bulk
When the bulk ω is trivial, then the thick boundary ψ gives rise to a state-sum on its own, which we will refer to as a thick state-sum. Such a thick state-sum is a two-dimensional state-sum with vertex labels equipped with a right G-action, and G-elements on the edges. Examples for such state-sums-with-action arose in the compactification discussed around Eq. (D16). As discussed in the previous paragraph, the set of vertex labels is isomorphic to a direct sum left coset sets H\G on which G acts transitively, for different subgroups H determined up to conjugation. On the other hand, we will refer to the conventional state-sum with only H-elements on the edges as thin state-sum.


4. Boundaries of thick bulk from boundaries of thin bulk
Note that the two-dimensional state-sums arising from compactifications in Appendix D are thick state-sums, as they have labels on the vertices that are acted on by G. The (boundary) anyons are in one-to-one correspondence with the (irreducible) generalized thick boundaries of those thick compactified state-sums. The formula above provides a way to obtain such boundaries from boundaries of a thin state-sum. The computation of the latter is simpler in practice as it takes place on smaller vector spaces. We follow the following steps, which are also discussed in a more algebraic way in Sec. II in the main text.
Decompose the thick compactified state-sum into transitive ones, with vertex label set H\G for different subgroups H.
For each transitive part, calculate the corresponding thin state-sum, that is, the corresponding H two-cocycle ω(a, b).
For each transitive part, find the irreducible generalized thick boundaries of this thin state-sum. This corresponds to finding the irreducible representations of the ω-twisted group algebra, or in other words, the projective irreducible representations of H with two-cocycle ω.
Use Eq. (E37) to obtain the irreducible generalized thick boundaries of the thick state-sum.
REFERENCES
In fact, one can directly calculate the modular data of the anyons without deriving the R tensor first. For this, one considers the vector space defined by a cellulation of a torus and analyzes the endomorphism induced by the mapping class group of the torus, generated by S and T matrices. For a detailed derivation, see Ref. 6.
Note that the full Lagrangian algebra is not only described by an object in the UMTC, i.e. the set of condensable anyons, but also by an algebra morphism. This morphism can also be computed explicitly combining structures and techniques described in this manuscript. We plan to address this in future work.
N, M have to share a divisor in order for there to exist a non-trivial 2-coycle class, see Appendix B.
Note that here we think of anyons as defects, so “ground state with anyons” means ground states of a Hamiltonian which is altered at some points to enforce the existance of anyons.