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 Z2 as well as Z3 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.

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.

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.

Consider a finite group G. Let X be a finite (left) G-set, i.e., there exist a map ⊳: G × XX representing the G-multiplication on X,
(1)
Given a G-set X, we define an algebra A over CG×X via the multiplication
(2)
with Ψ:G×G×XU(1)C. For this multiplication to define an algebra it has to be associative. This imposes a non-trivial condition on the phase Ψ,
(3)
This can be seen as a two-cocycle condition over U(1)X as a G-module with non-trivial action defined via the G-action on X. Moreover, we require that Ψx is normalized, i.e., Ψx(1G, h) = Ψx(h, 1G) = 1 ∀gG, xX. We can in fact redefine the basis states of A with ξ:GC× by (g, x) ↦ ξ(g)(g, x) to effectively map Ψx to a normalized two-cocycle22,23 so we do not loose generality with this assumption. We call such an algebra twisted group algebra with action. For more details on G-modules and group cohomology, see  Appendix B.

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.

First, we note that due to the delta in the multiplication in Eq. (2), A decomposes over transitive subsets of X, or equivalently, over G-orbits {XiX}, so we have an isomorphism
(4)
For any Xi there exists a subgroup KiG such that Xi is isomorphic to G/Ki, the set of left Ki-cosets. After this isomorphism, G acts via left translation, ghKi = (gh)Ki, g, hG, onto Xi. Since Xi is a transitive G-set, the stabilizer groups of any element xXi,
(5)
are isomorphic. In particular, for xX, gG,
(6)
Hence, we can define an abstract stabilizer group of X, StabG(X) ≃ StabG(x) for any xX. In fact, Ki is isomorphic to StabG(X). As a subgroup of G, Ki depends on the chosen identification XiG/Ki. We pick a representative x̂iXi and define
(7)
Note that we can obtain the stabilizer group of any other element in Xi from Ki with Eq. (6).
Next, we show how to further decompose each Ai into irreducible components. In fact, we find that they are in one-to-one correspondence with irreducible unitary Ψx̂i-projective representations (IPRs) of Ki. To see this, we explicitly construct the indecomposable central idempotents in Ai from IPRs of Ki. Let {ρx̂i} be the irreducible Ψx̂i-projective representations of Ki, defined via a representative x̂iXi. In particular, they fulfill
(8)
and are unitary. If not stated otherwise, any representation considered in this manuscript is assumed to be unitary. We can construct an equivalent IPR of any other isomorphic stabilizer group StabG(y) for yXi. Concretely, starting from an irrep ρx of StabG(x) and k′ ∈ StabG(x) we can define the irreducible representations on all of A via
(9)
which by construction fulfill
(10)
when the group label is restricted on the respective (isomorphic) stabilizer groups. Note that we act with the inverse group element on the left-hand-side of Eq. (9) because we pull of the G-action on X back onto ρ.

This allows us to uniquely construct the IPRs of all stabilizer groups StabG(x) for xXi from the IPRs of the stabilizer group of a single representative x̂iXi.

Given all the equivalent IPRs, we can construct the indecomposable central idempotents in Ai. They are labeled by IPRs {ρi} of Ki and given by
(11)
where χ̃ρixTr(ρix):GC is the irreducible Ψx-projective character derived from the IPR ρix. Using unitarity and irreducibility within Ai,
(12)
we can show straightforwardly that the ci,ρis are idempotent,
(13)
and central,
(14)
For now, we have found a set of invariants, {i, ρi} labeling independent central idempotents. To show that they are complete, i.e., correspond to irreducible invariant sub-spaces of A, we have to show that the ci,ρis are not only central and idempotent but also indecomposable. To see this, consider the following set of smaller idempotents
(15)
that are irreducible (by definition of χ̃) but not central. In fact, we can get to any value of xXi via left- or right-multiplication of an element in A which makes ci,ρi irreducible. Furthermore, the isomorphism discussed in  Appendix C provides a 1-1 mapping from irreducible representations of the twisted algebra with action in Eq. (2) and Ψx̂-twisted group algebras without action of (a collection of) subgroups. In  Appendix E we give an interpretation of this isomorphism in terms of an invertible domain wall mapping between two different kinds of state-sums with boundaries.

We summarize this section by giving a recipe on how to find the irreducible representations of an algebra of the form of Eq. (2).

  1. Decompose X into transitive G-orbits {Xi}.

  2. For each Xi:

    • Pick representative x̂iXi and calculate its stabilizer group Ki=StabG(x̂i).

    • Find irreducible Ψx̂i-projective representations of Ki, denote them with ρi.

    • Use Eq. (9) to derive irreducible representations and character functions, χ̃ρix(g)=Tr(ρix(g)) for all xXi.

  3. The indecomposable central idempotents in A are labeled by pairs (i, ρi) ∈ (G-orbits of X, Ψx-projective Irreps of Ki) and given by
    (16)
    Note 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.

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 C. 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 C-module category.

In the bulk, the microscopic model is defined by a spherical fusion category C. We denote the set of (finitely many) simple objects in C by Obj(C)={1C,i,j,,k} and fusion multiplicities NijkZ0. These define a fusion-operation
(17)
A fusion category is called Abelian if for every pair (i,j)Obj(C) there exists only one kObj(C) such that Nijk>0. In other words, the fusion of i and j is unique. Note that the term Abelian does not refer to the commutativity of the fusion operation. Moreover, the fusion above can be equipped with a non-trivial associator capturing the isomorphism between objects obtained from fusing in different orders. The associator is part of the input category C and will be described in terms of so-called F-symbols in microscopic models.

A microscopic (topological) model is defined on a framed trivalent graph (tessellating some two-dimensional manifold) with local Hilbert spaces Hi=span(Obj(C)) 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).

The total Hilbert space is the tensor product space of all the local spaces, Htot=iHi. Obj(C) defines a natural basis on Hi and with that on Htot. The fusion multiplicities define a local constraint at each vertex, defining the physical subspaceHphysHtot. Hphys is defined by the span of basis states for which the local labels (i, j, k) at every vertex fulfill Nijk0. For Nijk>1 the vertex itself carries in additional degree of freedom, span({0,1,,Nijk1}). For the rest of this work, we will work within Hphys and depict a vertex and its adjacent edge labels (i, j, k) in an allowed basis configuration with
(18)
Note that, when going around a vertex, one frame has to point in a different direction than the other two. We call this condition local acyclicity and is related to the fact that one has to distinguish the left and the right hand side of Eq. (17) when interpreting it as vertex. Given this prescription, one can tessellate (the bulk of) any two-dimensional manifold with these trivalent vertices, each of which enforce a local constraint on the edge labels. Every allowed configuration defines a basis vector in the state space assigned to the cellulation. To construct a topological fixed-point model, topological invariance is imposed exactly. In particular, one can relate different cellulations (with the same input category C) via Pachner moves, for example an F-move
(19)
The fusion category C defines linear transformations that relate vector spaces of different cellulations to each other. The matrix entries of this linear map in the basis of string labels are complex numbers {Fcdf,νρaabe,μη},
(20)
These F-symbols are part of the input category C. Any sequence of topological moves can be represented by a three-dimensional triangulation derived from the dual cellulation indicated in gray in Eqs. (19) and (20). In the multiplicity-free case, where μ, ν, ρ, η are fixed, the F-move above, for example, is represented by a tetrahedron,
(21)
In general, the faces would carry the additional multiplicity labels, but we will suppress those multiplicity labels for the rest of this paper. One can view the bottom (dotted) edge as the “initial,” vertical, edge and the top edge as the “final,” horizontal, edge in Eq. (20). The remaining four edges correspond to the outgoing/incoming edges in Eq. (20). Note that the boundary of the complex is the union of the initial and final cellulation. In a similar way, any sequence of Pachner moves can be represented by a three-dimensional cellulation whose boundary is the union of the initial and the final (two-dimensional) triangulation. The associated amplitude is evaluated by taking the product of numbers associated to the subsimplices. For details on the space-time picture for topological moves in our models, see  Appendix D.
The numbers associated to the labeled space-time simplices have to fulfill several consistency conditions to implement exact topological moves. For example, if one combines more than two vertices, there are different sequences of Pachner moves that have the same effect on the tesselation, see, for example, Fig. 1. This results in the non-trivial pentagon equation on the F-symbols,
(22)
for all a, b, c, d, e, f, g, h, i which can be depicted as
(23)
On the left hand side the diamond formed by the five vertices is tessellated with two tetrahedra (top and bottom) whereas on the right hand side the same diamond is tessellated with three tetrahedra around the middle axis. Both should give the rise to the same number when summed over labels in the interior.
FIG. 1.

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.

FIG. 1.

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.

Close modal

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 C26 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.

At the boundary, the bulk model is terminated along boundary edges which can host different degrees of freed. However, the there is some (possibly non-trivial) action of the bulk edges connected to the boundary on these new degrees of freedom. In order to have full topological invarince of the model, the action has to fulfill certain consistency conditions. As we will show in this section, a boundary to a bulk with input category C is given by a (left) C-module category MC.8,11,26 The module category MC is a (semi-simple) category equipped with a (left) C action. On the level of the simple objects of MC, the C action is defined via module fusion multiplicites Ma,αβZ0,
(24)
This action has to be compatible with the fusion operation in C, i.e.,
(25)
This gives consistency conditions on {Maαβ} in terms of {Na,bc}.
Graphically, the C-action can be represented by trivalent boundary vertices
(26)
where μ is non-trivial for Maαβ>1. We will later resort to the space-time picture where the dual edges of the above cellulation enter. As before, we depict them by dashed edges. Similar to the F-symbols in the bulk, there are linear maps corresponding to deformations of the boundary, whose entries we refer to as L-symbols,
(27)
In the space-time picture this move can be represented by a triangle with additional boundary labels at each of its vertices. Omitting the multiplicity labels it can be represented by
(28)
The L-symbols have to fulfill an associativity condition similar to the bulk pentagon equation. This can be expressed as a commutative diagram in Fig. 2. The resulting condition on the L- (in relation to the F-symbols) reads
(29)
The equation on the right shows the corresponding space-time recellulation. On the left, there are two boundary triangles, whereas on the right there is one bulk tetrahedron with two boundary triangles. Note that the edge labeled by e on the right side is not part of a boundary triangle but extends into the bulk. Again, Eq. (29) does not yield full topological invariance, but we have to add additional axioms as well as Hermiticity. For the models we are going to study, these additional axioms will be fulfilled automatically.
FIG. 2.

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.

FIG. 2.

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.

Close modal

Taken together, the mathematical structure describing the bulk-boundary microscopics is a C-module category. Given C 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

Given consistent F and L symbols – a fusion category C and a C-module category MC – we have defined a topological fixed-point model for a topological phase with boundary. In fact, this data can also describe interfaces between topological models each of which is described by fusion categories C and C. Such interfaces are often called domain walls and are defined by a C-C-bimodule category.8,11,26 A C-C-bimodule category is defined by a left C-, and a right C-action, each of which is equipped with L-, respectively R-symbols that are both compatible with fusion in C, respectively in C. Graphically, the associated moves can be depicted by
(30)
where C-labeled strings are depicted in green and C-labeled strings in blue. Moreover, there are moves including strings on both sides of the domain wall. The associated linear maps are given by so-called C-symbols and can be graphically depicted as
(31)
Again, the L, R and C symbols have to fulfill consistency conditions coming from different sequences of moves with the same overall effect. Another way of studying domain walls is by “folding” one side of the domain wall onto the other side. With this, the C-C domain wall becomes a boundary of a CC̄ model, where the bulk strings are labeled by tuples (a,a)Obj(C)×Obj(C) and the F-symbols are inherited from C and C but with the CF-symbols complex conjugated. Omitting the multiplicities in C and C individually, the folding trick can be depicted graphically as
(32)
Note that the boundary vertex can get a multiplicity even if the two input fusion categories are multiplicity-free. This is less of a feature of the folding trick than a consequence of combining two connected tri-valent vertices into a single four-valent one. We will see that for the model class we are interested in this paper, there is no additional multiplicity coming in. The L-symbols of the folded model (see Fig. 3) are straightforwardly obtained by a combination of the (L, R, C) symbols using the correspondence above to resolve the L move of the folded model to a sequence of L, R and C moves in the unfolded model,
(33)
FIG. 3.

Via the folding trick, a C-C domain wall, classified by a C-C bimodule, is equivalent to a CC 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.

FIG. 3.

Via the folding trick, a C-C domain wall, classified by a C-C bimodule, is equivalent to a CC 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.

Close modal
Let G be a finite group. For the rest of the section, we focus on C=Vecω(G), the category of G-graded vector spaces. In Vecω(G), the simple objects are group elements and the fusion multiplicities are given by the group multiplication,
(34)
With that, the vertices take the following form
(35)
Since there are just two free labels at any vertex, the F-moves are defined by a single function ω: G × G × GU(1),
(36)
In this case, the pentagon equation reduces to a three-cocycle condition on ω,
(37)
Moreover, we have a gauge freedom. Namely, we can multiply the F-symbols by product of local unitaries. In our model these unitaries are simple phases for every vertex, respectively face in the dual picture. In general this phase can depend on the local configuration around the vertex, given by a pair of group elements. Hence, any function η: G × GU(1) defines a gauge transformation. Applying the corresponding gauge in (36) maps the associated three-cocycle to
(38)
In fact, the product of ηs with which the three-cocycle gets multiplied is exactly the coboundary of the two-cochain η. This shows that topological lattice models based on Vecω(G) are classified by the third cohomology group H3(G, U(1)), see  Appendix B.

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 HG 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 Cψ[H] – a subgroup H and a two-cocycle ψ on it – in our calculations.

Given a subgroup H, the labels at the boundary are cosets in G/H = {aH|aG}. The bulk G-action is then defined by left action on the coset,
(39)
where aα is a representative of the coset α.
Since every vertex is multiplicity-free and only two of the labels of its incident edges are independent, the L-symbols only have three open indices. They are defined by ψ: G/H × G × GU(1) with
(40)
The phase function has to satisfy the mixed pentagon equation in Eq. (29). In this case, it simplifies to
(41)
where δ̃ is the twisted coboundary operator (for more details, see  Appendix B). This means ω has to be the twisted coboundary of ψα if we interpret ω as a twisted cochain which is constant in α. By setting α to the trivial coset H and restricting all arguments to the subgroup H, we see that ω has to be a (untwisted) coboundary when restricted to H. Given any solution ψ to the equation above and a twisted cocycle α (that is, δ̃α=0), ψ′ = ψ · α is also a solution. Furthermore, two different boundaries ψ can be considered equivalent if they differ by on-site unitary gauge on the boundary fusion vertices. In our case, those fusion spaces are one-dimensional (or zero-dimensional if the fusion rules are not obeyed), so such an isomorphism is defined by a phase ξ: G/H × GU(1) depending on the labels of the strings adjacent to a boundary fusion vertex. Since ψ contains three boundary fusion vertices (two at its input and one at its output), the isomorphism acts as
(42)
We see that the gauge corresponds to a multiplication of a coboundary with the same coboundary operator as in Eq. (41) but acting on the one-cochains {ξα}. In total, we see that although the different ψ are not twisted two-cocycles, their differences are, whereas ψs differing by 2-coboundaries are considered equivalent. Hence, the gauge equivalence classes of boundary models cannot be directly identified with the twisted cohomology group H2(G,(U(1)G/H)G), but are equipped with a regular action of the latter. Here, the subscript G indicates the non-trivial right action of G. That is, the set of equivalence classes forms a torsor over the second twisted cohomology group.
In  Appendixes C and  E, we show that this classification indeed coincides with known results, i.e.,
(43)
In particular, this isomorphism holds for every cohomology group Hn for n ≥ 1 and is induced by the map m(n): G/H × G×nH×n that – to the best of our knowledge – has first been mentioned by Lawson in Ref. 27.
A topologically ordered phase is called Abelian if the fusion outcome of any pair of topological point defects – anyons – is unique, i.e. they form an Abelian fusion category. Importantly, not every Abelian fusion category C gives rise to an Abelian phase D(C) (see Sec. IV D). However, as we will see later, the Abelianess of the anyons can be traced back to properties of the input category if it is of the form Vecω(G). In this case, the group has to be Abelian and the twisting three-cocycle of a certain form. Any finite Abelian group is isomorphic to a product of cyclic groups,
(44)
where pi is prime and mi positive integers for any i = 1, …, N. For the rest of this section it suffices to consider G having N independent cyclic factors. Inequivalent F-symbols are classified by three-cocycle classes on the group above. In fact, any three-cocycle class on such a product group can be represented as a product of so-called type-I, type-II and type-III cocycles (see  Appendix B). The anyons in our model are only Abelian for a cocycle that is cohomologous to a product of type-I and type-II cocycles only, see Sec. IV D.
From the above decomposition into cyclic factors any subgroup H and the associated cosets G/H can be easily determined. Moreover, its second cohomology group decomposes similarly, see  Appendix B. Given a subgroup H (with k factors), we can construct a non-trivial two-cocycle by taking products of two-cocycles on each pair Hij=Zhi×Zhj in the factor decomposition of H. In particular, there are gcd(hi, hj) inequivalent two-cocycle classes on Hij represented by normalized two-cocycles of the form
(45)
where ai(aj) is the component of a in Zhi(Zhj)Hij. Following the previous subsection, the associated L symbol is obtained by precomposition with m(2),
(46)

In this section, we give some examples of the input data for Vecω(G) models with boundary.

1. Vecω(ZN)

The cyclic group of order N, ZN={0,1,,N1} with group operation ap ba + b mod N has N three-cocycle classes each of which are generated from a single three-cocycle class. The normalized three-cocycle
(47)
is a canonical representative for the nth cocycle class of H3(ZN,U(1))=ZN.28 We say that a three-cocycle of that form, only supported on a single cyclic factor, is of type I. Using ZN and ωIn as an input gives an Abelian bulk theory.

Let us look at possible boundaries of such a bulk model. For simplicity, we take N = p prime. In this case, Zp only has two subgroups, H0 = {0} and H1=Zp. 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, ωIn becomes cohomologically trivial only on H0. Hence, these models only have one “standard” boundary.

2. Vecω(ZN×ZM)

An Abelian group with two cyclic factors ZN×ZM={(a,b)|aZp,bZq} has NM three-cocycle classes of type I. Their normalized representatives decompose into a product of cocycles as in Eq. (47), each supported on one factor only. Additionally, there are non-trivial type-II cocycle classes represented by28 
(48)
Three-cocycles of type II are supported on two cyclic factors and are gauge inequivalent to any type I cocycle. Together with the two subgroups generated by type I cocycles we have H3(ZN×ZM,U(1))=ZN×ZM×Zgcd(N,M). Using ZN×ZM and a cocycle of type I or II as in input gives an Abelian bulk model.
For simplicity, consider N = M = p prime. The possible boundary models are given by subgroups of Zp×Zp. They are H0={0},H1=(0,1)Zp,H2=(1,0)Zp,H3,l=(1,l)Zp and H4=Zp×Zp. In the untwisted case any of these subgroups defines a boundary model. Additionally, H4 has non-trivial two-cocycles of the form
(49)
Since H4 is the whole group, there is only one coset such that ψ only depends on group labels. Note that if H has two cyclic factors and is a proper subgroup of G the form of ψ will be different because the coset label α can be non-trivial, see Eq. (46).

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)

The smallest non-Abelian group is the permutation group of three elements S3=t,r|t2=r3=e,tr=r2tZ3Z2 (e being the identiy element). Its third cohomology group is given by28 
(50)
Interestingly, the third cohomology group is the product of the cohomology groups of the two non-trivial subgroups of S3. However, since S3 is a a semidirect product of the two, the three-cocycles are not simple products of three-cocycles of the subgroups. The six inequivalent classes can be represented by the normalized cocycles28 
(51)
where the elements in S3 are represented by tAra with AZ2, aZ3 and ⊕ denotes addition mod 3. When comparing the individual factors with a three-cocycle of a cyclic group [see Eq. (47)], we note that the second factor, (−1)ABC, corresponds to the non trivial three-cocycle of the (Abelian) subgroup Z2S3 and the first factor corresponds to the non-trivial three-cocycle of the subgroup Z3S3 precomposed with the automorphism ρA,B,CAut(Z3×3) defined by ρA,B,C(a, b, c) = ((−1)B+Ca, (−1)Cb, c) for any (A,B,C)Z2×3.

S3 has four subgroups up to conjugation, H0 = {e}, Ht=tZ2, Hr=rZ3 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.

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.

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.

In principle, one can choose any tesselation of the annulus for the upcoming analysis. It is beneficial to choose a simple representative. For the rest of this section, we define a basis element of the tube algebra T, labeled by (a,b,c,d)Obj(C)4 (again, omitting the multiplicity labels at the vertices), via the following cellulation:
(52)
We can define an (associative) multiplication on the vector space spanned by the diagrams of the above form by gluing two tessellated tubes together and using topological moves to map back to the initial cellulation, see Fig. 4. Evaluating the space-time complex corresponding to the recellulation, we obtain
(53)
FIG. 4.

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.

FIG. 4.

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.

Close modal

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.

Before deducing new quantities of the topological phase of the lattice model, let us take a step back and put the tube algebra into a larger context. The fact that the topological vector spaces associated to a tube, S1 × (0, 1), forms an algebra, comes from the fact that gluing a tube onto a tube again gives a tube. Hence, we can define an action of T onto itself, i.e., there exist a map
(54)
defining an (multiplicative) action of T onto itself. Similarly, when considering cellulations (modulo local moves) of other manifolds, one can define endomorphisms on the topological vector space from gluing operations that leave the manifold invariant. We will now turn our attention to how this allows us to calculate the fusion multiplicities in the UMTC formed by the anyons. Consider a fusion process in 2 + 1-dimensional space-time. It can represented as a vertex where three anyon world-lines meet. The boundary of the regular neighborhood of this vertex is a pair of pants – or three-punctured sphere. Just as before we can define a vector space (in terms of its basis) by tessellating this manifold with trivalent vertices from the input fusion category [see (18)]. Let’s call this space F. In fact, gluing a tube onto any of the three holes of the pair of pants does not change its topology. Hence, we can define a “tri-representation”
(55)
As a representation, F decomposes into a direct sum of triples of irreducible representations of T,
(56)
where a, b, c label the irreducible representations of T, i.e., the anyons of the bulk. The multiplicities in this decomposition Ñabc coincide with the fusion multiplicities of the bulk anyons. In order to calculate them, we consider the tri-representation in Eq. (55) for a triple of central idempotents (PaT,PbT,PcT)T×3 yielding a projector PF, and then take the trace of PF. We will see that we use a similar approach to get the fusion multiplicities into the boundary, see Sec. VI. In  Appendix D we give a formula for Ñabc for an input category of the form Vecω(G). We will use the same method to calculate the dimension of the “fusion” vector spaces for another type of space-time event, namely, the partial condensation of an anyon in the bulk to an anyon in the boundary, see Sec. VI.
To complete the UMTC, we have to define a braiding and an associator on the fusion spaces of the irreducible representations of T. One way to fully define a braiding is via so-called R-symbols, defining a half-braiding on anyon world-lines,
(57)
The associator on the other hand is defined by so-called F-symbols changing the order of fusion. Both R- and F-symbols have to obey consistency conditions known as pentagon and hexagon conditions.26 Given this set of data, one can derive the topological quantum numbers like self-exchange statistics (topological spin θa) and the mutual-exchange statistics (S matrix) for any anyon with a simple calculation. For more details, see Appendix E of Ref. 3.

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 (PaT,PbT,PcT)T×3. 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.

For gauge theory models the input fusion category is Vecω(G), each vertex is multiplicity-free. The basis state in Eq. (52) are parametrized by two group elements. We label them by (g,h)T=(hgh1,h,hg,g)T. Plugging in the defining data from Sec. III C, we obtain the tube algebra T for the group case as
(58)
defined over the vector space CG×G, where βg(h, h′) is the phase assigned to the sequence of moves from the left to the right hand side of Fig. 4. Evaluating the space-time complex representing the moves in Fig. 4, we get
(59)
which can be seen as a slant product igω, see  Appendix B. The tube algebra represents Dω(G), the twisted quantum double of G, introduced by Dijkgraaf et al. in Ref. 34 and plays an important role in the study of (finite) gauge theory models for topological phases and its applications to topological stabilizer codes.6,30,35,36
In fact, we find that Eq. (58) is of the form of Eq. (2) with X = G acting on itself via conjugation, i.e., hg = hgh−1. Indeed, βg plays the role of Ψx and fulfills
(60)
in analogy to Eq. (3). Choosing a normalized three-cocycle ω makes βg normalized and with that we can use the algorithm from Sec. II.
To construct the irreducible representations of T and the associated central idempotents we proceed as described in Sec. II. First, the transitive subsets {Xi} are the conjugacy classes {c}. For each conjugacy class c, we pick a representative ĉ. It is stabilized by its centralizer Z(ĉ)={gG|gc=cg}. Following the general considerations from Sec. II the irreducible sub-spaces of the tube algebra are additionally labeled by irreducible βĉ-projective representation of Z(ĉ). Combined, we identify the irreducible sub-spaces of T with a pair (c, ρc), a conjugacy class and an irreducible βg-projective representation of Z(ĉ). The central idempotents are given by
(61)
where χ̃ρcg(h)Tr(ρcg(h)) denotes the projective character of the IPR ρcg of Z(g). Note that the stabilizer group of any element in c is given by conjugating the centralizer of the chosen representative, Z(gĉg1)=gZ(ĉ)g1.
In practice, finding the irreducible projective character functions is not a straightforward task. In the special case, where βg is a coboundary, i.e.,
(62)
the IPRs of Z(g) are in bijection with the irreducible linear representations. In particular, one can “gauge” away the twist βg with the cochain ɛg and the projective characters are given by
(63)
where χρg(h) is the linear (not projective) character of the irrep ρ of Z(g). With that, the central idempotents simplify to
(64)
We will see in the following sections that the examples considered in this paper are all of the above form. A well-studied example where βg defines a non-trivial projective irrep is obtained from G=ZN×3 and ω being cohomologous to a type-III three-cocycle. In this case, there are higher-dimensional irreducible sub-spaces in the tube algebra even though the input group is Abelian, see, for example, Refs. 6 and 28.
In an Abelian anyon model the fusion of two anyons a and b yields a unique anyon c, i.e., Ñabc=1 for exactly one c and 0 otherwise. In other words, the fusion space of any pair of anyons is one-dimensional. If we want our microscopic model to yield an Abelian anyon theory, we thus require |c| = dim(ρc) = 1 for all conjugacy classes and IPR ρc. The fact that all conjugacy classes consist of a single element implies that G is Abelian. All linear irreducible representations of an Abelian group are one-dimensional and can be labeled by group elements. In order for the IPRs to be one-dimensional as well, βg has to be trivial in the sense of Eq. (62). In this case, the central idempotents simplify to
(65)
With these |G|2 independent central idempotents in the |G|2-dimensional algebra T we have found all irreducible sub-spaces. Each is labeled by a pair of group elements (g, k) and one-dimensional. The latter shows that the associated anyons are indeed Abelian and is directly related to the fact that βg is a coboundary. If it is a non-trivial two-cocycle, the resulting anyon theory is non-Abelian.6,28

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. Vecω(Zp)

Any three-cocycle class of ZN={0,1,,N1} is represented by a type-I three-cocycle as in Eq. (47). Note that this cocycle is symmetric in the latter two arguments, ωIn(a,b,c)=ωIn(a,c,b)a,b,c. Using this and the fact that Zp is an Abelian group, βa(b, c) reduces to ωIn(a,b,c) [see Eq. (59)] and the multiplication in T is given by
(66)
This tube algebra is Abelian in the sense described above since for any every aZp, we can define the one-cochain
(67)
such that βa(b,c)=ωIn(a,b,c)=(δεan)(b,c). Together with the group character function χa(b)=e2πipab, the central idempotents of the tube algebra – labeled by (a,k)Zp×2 – are given by
(68)

2. Vecω(Zp×Zp)

For a type-I cocycle on either of the two factors of Zp×Zp, the central idempotents will have the same form as the ones in Eq. (68). In this example, we consider a type-II cocycle defined in Eq. (48). The multiplication in T is given by
(69a)
(69b)
where we abbreaviated δa1,a1δa2,a2 with δa,a. Again, we can find a one-cochain
(70)
with the coboundary (δεan12)(b,c)=ωIIn12(a,b,c). Together with the character function χa(b)=e2πip(a1b1+a2b2) the central idempotents – labeled by (a,k)(Zp×Zp)×2 – are given by
(71)

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

S3 has two independent generators, r and t. They satisfy t2 = r3 = e, where e is the identity element, and tr = r2t = r−1t. The subgroup Z3 generated by r is normal. S3 has three conjugacy classes
(72)
with the respective centralizers
(73)
To obtain the central idempotents of the tube algebra, we need to find the irreducible representations of the centralizers. Let us first consider the trivial conjugacy class A. Its centralizer is all of S3 and has three irreducible representations, a trivial one Γ0, a one-dimensional one Γ1 and a two-dimensional one Γ2. The characters are given by
We obtain three anyons associated to the trivial conjugacy class. The corresponding idempotents read
(74a)
(74b)
(74c)
Note that (ē,Γ0) is the trivial anyon – or vacuum – subspace. The dimensions of the respective sub-spaces coincide with the dimensions of the irreducible representations Γ0, Γ1 and Γ2 respectively, i.e., d(ē,Γ0)=d(ē,Γ1)=1 and d(ē,Γ2)=dim(Γ2)=2.
Now consider the two non-trivial conjugacy classes. Their centralizers are cyclic groups, so their irreducible representations are one-dimensional and labeled by group elements. The characters read
Evaluating Eq. (64) for the two conjugacy classes and the character functions shown above gives the projectors
(75a)
(75b)
(75c)
(75d)
(75e)
The dimensions of the associated irreducible representations are
(76)
So in total we have found eight different anyons and the corresponding central idempotents of the tube algebra.
Next, let us consider a non-trivial three-cocycle from Eq. (51). This yields the phase [see Eq. (58)]
(77)
in the multiplication in T, which turns out to be a (twisted) coboundary28,
(78)
The cochain εtAra defines the projective characters in terms of the linear ones, see Eq. (63). With that, the central idempotents of the ωp-twisted S3 theory are given by Eq. (64). The number and dimensions of the irreducible sub-spaces are the same as in the untwisted case but the phases in the central idempotents are modified by εgp. This in turn changes their fusion and braiding (see Refs. 6, 28, and 38).

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 

Boundary anyons are determined by string-nets on a “semi-tube,” a tube that is cut in half by the physical boundary. One might also think of one half being replaced by vacuum. Note that for a domain wall between two phases, modeled by fusion categories C and C, the other half would not be vacuum but another bulk for the fusion category C. However, as illustrated in Sec. III B 1, these domain wall diagrams can be related to an equivalent boundary diagram of the form considered in this section. This semi-tube can be tessellated by a string diagram as follows,
(79)
which defines an orthonormal basis of the defect space S. In the following, we will omit the multiplicity labels at the vertices. Again, we can use gluing operations and topological moves to define a multiplication on S. In Eq. (5) we show the sequence of local moves that defines the multiplication in terms of its associated space-time complex. Evaluating the complex depicted in Fig. 5 gives rise to the multiplication
(80)
By analogy to the tube algebra, we call the resulting algebra semi-tube algebra. Similar structures have been already introduced in various places in the literature as “module tube algebra,” “module annular algebra” or “ladder category.”8,11,39,40
FIG. 5.

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.

FIG. 5.

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.

Close modal
In the case where the bulk is defined via Vecω(G), the boundary model is given by a subgroup HG [and a two-cocycle in Z2(H, U(1))], the boundary labels are cosets in G/H and every vertex is multiplicity-free, see Sec. III C. Hence, a semi-tube basis element is parameterized by (α,β,g)SG/H × G/H × G. The associated basis state from Eq. (79) are given by (α,β,g)S ≔ (α,β,gβ,gα,g)S. Plugging in the defining data form Sec. III C into Eq. (80), we get the multiplication
(81)
over the vector space CG/H×G/H×G.

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.

Again, S is an algebra of the form discussed in Sec. II where X = G/H × G/H with G acting via simultaneous left-translation: g ⊳ (aH, bH) = (gaH, gbH). Using the (twisted) two-cocycle condition of ψ we can easily see that
(82)
is a twisted two-cocycle fulfilling
(83)
in analogy to Eq. (3). Following Sec. II, we first have to find the G-orbits of G/H×2. Let α−1H\G be the left inverse of α, i.e., α−1α = H. Then we find that the double coset xα−1βH\G/H, is invariant under the G action. Moreover, all the subsets Sx ≔ {(α, β) ∈ G/H|α−1β = x} are transitive as we will show now.

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 h1a11b1h=a21b2. Hence, there exists a group element, namely, g=b2h1b11=a2h1a11, such that gα1 = α2 and gβ1 = β2. So any two coset pairs (α1, β1), (α2, β2) ∈ Sx are related via the action of some gG.

Next, we find the stabilizer subgroup for any element in Sx. In particular, we can pick a representative x̂=(α̂,β̂), explicitly calculate its stabilizer subgroup and then use Eq. (6) to derive the stabilizer groups of the other elements in Sx. Let (α̂,β̂)=(aH,bH)Sx. By construction, (H, a−1bH) will also be in Sx. The stabilizer of this element,
(84)
is particularly easy to compute and given by
(85)
This formula holds for any element in Sx, replacing a−1b with the respective double-coset representative of x. From these general considerations above, we can see that Kx simplifies in some special cases:
  • 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.

Combined, the irreducible sub-spaces of S are labeled by a doble-coset xH\G/H and an irreducible Ψα̂,β̂-projective representation of Kx. The associated central idempotents are given by
(86)
where χ̃κx(α,β)(g)Tr(κx(α,β)(g)) is the projective character of the IPR κx of StabG((α, β)).

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.

The fusion of the boundary anyons is described by how the irreps of the semi-tube algebra multiply, respectively fuse, to a direct sum of other irreps. If every fusion product only consist of a single irrep, we say the boundary anyons are Abelian. As discussed in Sec. IV C for the tube algebra there is a direct correspondence to the dimensions of every irrep being one-dimensional, meaning that the multiplcation in the tube algebra is commutative. In general, however, this correspondence does not hold. One can have a non-commutative algebra, captured by the existence of higher-dimensional irreps, that describe Abelian anyons. At the same time, every irrep being one-dimensional gives a sufficient condition on the associated anyons being Abelian. For models with Abelian bulk anyons (Abelian G and no type-III three-cocycles) every irrep is one-dimensional iff Ψα,αx is cohomologous to a trivial two-cocycle. In this case, we can express the irreducible projective characters in terms of linear ones,
(87)
Here, ηx(α):GU(1) is the cochain that trivializes the two-cocycle Ψα,αx. Linear characters χκ of (finite) Abelian groups are well known, see Sec. IV E. Combining the observations above, the central idempotents read
(88)
We believe that any boundary models to a bulk that leads to Abelian anyons only result in an Abelian fusion of the boundary anyons, also when the semi-tube algebra has higher-dimensional irreps.

Note the different notions of “Abelian.” In particular, the semi-tube algebra might not be Abelian (commutative) as an algebra but its irreducible representations still be described by an Abelian fusion category. For a recipe to calculate that fusion data explicitly, we refer to Refs. 8 and 10.

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

Before we consider specific groups (and their subgroups), let us take a closer look on how Eq. (86) simplifies in the case of the two trivial subgroups {1G} and G. In the former case, we denote the resulting semi-tube algebra S. The central idempotents are labeled by x ∈ {1G}\G/{1G} ≃ G alone. All the prefactors in Eq. (86) evaluate to 1 resulting in
(89)
In the latter case, where H = G, we denote the resulting semi-tube algebra by S. In this case the central idempotents are labeled by irreducible (linear) representations of G since there is only one double coset, namely, the trivial one. Moreover, since H = G only defines a boundary in the case of a trivial three-cocycle in the bulk, S is isomorphic to an untwisted group algebra. With that, only the G-character functions χκ enter into the prefactor in Eq. (86) and the central idempotents take the form
(90)
Note that with H = {1G}, the fusion category of boundary anyons is the same as the input fusion category Vec(G), whereas for H = G, they form the Morita equivalent category Rep(G).

2. Vecω(Zp)

Consider Zp={0,1,,p1} for p prime with the type-I three-cocycle ωIn in Eq. (47). Zp has only the trivial and full subgroup which were already discussed in Sec. V D 1. For the trivial subgroup H0 = {1G}, the p central idempotents are given as in Eq. (89) by
(91)
where ⊕p denotes addition modulo p. The full subgroup only defines a boundary for the trivial three-cocycle n = 0. The central idempotents are then given as in Eq. (90).

3. Vecω(Zp×Zp)

Consider the type-II cocycle ωIIn12 on Zp×Zp 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 Zp subgroups H1, H2, H3Zp from Sec. III E 3 define valid boundaries for any n12.

Zp only has trivial two-cocycles so it defines a unique boundary. Let us consider the particularly simple case of H2=(0,1)ZpZp×Zp, on which ωII directly evaluates to 1. In the microscopic model, the corresponding boundary labels are double cosets H2\Zp×Zp/H2, which can be identified with the integers {0, 1, …, p − 1}. With that, the p2 central idempotents are labeled by (x,κ)Zp×Zp and are given by
(92)

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).

In the case of a trivial bulk three-cocycle, any other subgroup defines a boundary as well. Consider the non-trivial normal subgroup Hr=rZ3. It defines a boundary of a bulk model twisted by ωp for p = 0, 3. Since Hr is normal in G the double cosets H2\G/H2 are in one-to-one correspondence with cosets G/H2Z2. The central idempotents are labeled by xG/HrZ2 and an irreducible representation of Hr, κZ3. Moreover, Hr only has trivial two-cocycles, so we use linear characters as in Eq. (86). Taken together, the central idempotents for an Hr-boundary read
(93)
for both p = 0 and 3. This shows that there are six inequivalent boundary anyons. In fact, when considering their fusion on the boundary, they form the fusion category Vec(S3) which is the same as the input category for the bulk model and with that – as expected – Morita equivalent to it, compare Ref. 29.

Let us now consider the more subtle case of the non-normal subgroup HtZ2. 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 xHt\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 Z2={0,1}. For the non-trivial double coset x = HtrHt we pick the representative rS3. 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.

Combined, the St boundary can host three different anyons, associated to the indecomposable central idempotents
(94a)
(94b)
In contrast to the Hr-boundary the Ht-boundary has only three anyons. Including fusion along the boundary, they form the fusion category Rep(S3) which is indeed Morita-equivalent to Vec(S3) (compare Ref. 29).

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.

As we have seen in Secs. IV and V, the (boundary) anyons can be computed by considering string-net diagrams on a (semi-)tube. Stacking a (semi-)tube to another one defines an action of the vector space of (semi-)tubes on that same vector space, which can be interpreted as an algebra, namely, the (semi-)tube algebra. Since a bulk-to-boundary fusion event maps a bulk to a boundary anyon, it should be related to a one-manifold which connects between tube segments and semi-tube segments. This can be achieved by an annulus where one boundary circle connects to tubes, and the other boundary circle half consists of the physical boundary and half connects to semi-tubes,
(95)
A tube can be attached to the small gray circle in the center, and a semi-tube to the large gray half circle at the bottom. As shown, we choose a minimal string diagram to cover the two-manifold. This two-manifold defines a bimodule C with an action of the tube algebra T and an action of the semi-tube algebra S, such that the representations commute with each other. The vector space of C is spanned by labeled string diagrams of the above form. The T-action is defined by gluing a tube into the inner hole and using F moves to rearrange the string diagrams to the above form. Similarly, the S-action is defined by gluing a semi-tube segment to the exterior gray half circle and using F and L moves. The general actions are given by
(96a)
(96b)
where we again omitted the multiplicity labels. In Fig. 6 we show the S-action as a space-time complex. The T-action is essentially given by the complex in Fig. 4 with two additional boundary labels at two of the vertices.
FIG. 6.

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.

FIG. 6.

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.

Close modal
As a bimodule, C is a equivalent to a representation of TS, and thus decomposes into pairs (i, j) of irreducible representations i of T and j of S with multiplicities mij,
(97)
where Ti and Sj are the irreducible representations of the tube, respectively rectangle, algebra. Physically, mij the dimension of the vector space of the bulk-boundary fusion event “i fuses to j.” Acting with the associated central idempotents of these irreducible representations from the left and right constructs a projector onto the respective sub-spaces of dimension mij which we will denote with PijC. The multiplicities are obtained by taking the trace of this projector,
(98)
In gauge theory models the bulk strings are given by Vecω(G) and the boundary strings by a Vecω(G)-module category, classified by a subgroup H and a 2-cocyle class in H2(H, U(1)). The basis of C can by fully parameterized by two group and one coset element. We choose the following assignment:
(99)
with g, hG and αG/H. The T action reduces to
(100)
Similarly, the S-action is given by
(101)
Given these two actions we can use the central idempotents derived in Secs. IV C and V B respectively their regular representation to project onto the associated irreducible representations of T and R within C. In combination, Eq. (98) defines a projector P(c,ρc),(x,κx)CEnd(C), projecting onto the pair of Irreps (c, ρc) and (x, κx) in C. With the T and S action above, the consecutive action of P(c,ρc)T and P(x,κx)S onto a basis element (g,h,α)C is given by
(102a)
(102b)
where we have used Eq. (9) to relate the projective characters of different – but isomorphic—stabilizer groups. Taken together,
(103)
where we have used that h′ ∈ Z(g) in every term of the sum above. The multiplicities m(c,ρc),(x,κx) can now be obtained using Eq. (98). Using the dimensions of the irreps,
(104)
we obtain
(105)
where we have used the twisted two-cocycle condition for βg, Eq. (60), and Eq. (9) to simplify the trace to the expression in the last line.

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

For Abelian models, the above expression simplifies significantly. First note, that each group element is its own conjugacy class and the double coset x corresponds to a unique left coset. Moreover, the projective representations are in one-to-one correspondence to the linear representations (see Sec. IV D) which can be labeled by group elements. Similarly, the boundary anyon labels simplify (see Sec. V C). We get
(106)
where χkG denotes an irreducible G-character and χκH an irreducible H-character.
Macroscopically a (topological) boundary is defined via the set of anyons that can condense at the boundary, i.e., fuse to the trivial boundary charge. This set has to form consistency conditions involving their fusion and exchanges statistics which render the set of condensable anyons a Lagrangian algebra13–15 in the UMTC describing the anyon theory. Our framework connects the microscopic description of the boundary as a lattice model to the macroscopic picture of anyon condensation. In particular, Eq. (105) allows us to explicitly calculate the Lagrangian algebra object41 corresponding to a (H, ψ)-boundary by setting (x, κx) = (H, Γ0), the trivial boundary defect. We get
(107)
noting that α−1 = StabG(α) = {gG |gα = α}. Note that for a trivial three-cocycle, this formula can be derived from Ref. 12. In particular, we have shown that in twisted models the linear character gets replaced by a projective character.
In the Abelian case, it simplifies further to
(108)
where χkG denotes an irreducible G-character.

1. Vecω(ZN)

As first example we consider the simplest class of Abelian models, when G=ZN, the cyclic group of order N. Its character function reads χk(g)=e2πiNkg. The bulk can be twisted with a type-I three-cocycle of the form of Eq. (47) giving rise to εg(h)q=e2πiN2qgh,qZN.

For any q, the trivial subgroup H0 = {0} defines a boundary. Plugging the defining data into Eq. (108) yields the associated Lagrangian subgroup
(109)
i.e., the boundary condenses all the pure charges.
For an untwisted bulk, q = 0, the other trivial subgroup H1=ZN defines a boundary as well. Plugging in its defining data into Eq. (108) gives rise to the associated Lagrangian subgroup
(110)
consisting of all the pure fluxes.
If N is not prime ZN has more subgroups. Depending on q, these might also define a boundary. For example, for N = 4n for nN and q = N/2 = 2n, the subgroup H2=N/2Z2 also defines a boundary, since ωIN/2H21. Using Eq. (108) yields the associated Lagrangian subgroup
(111)

2. Vecω(Zp×Zp)

Consider the group G=Zp×2 with the character function χ(k1,k2)((g1,g2))=e2πip(k1g1+k2g2) and a bulk twist of type-II ωq from Eq. (48), giving rise to ε(g1,g2)q((h1,h2))=e2πip2g1h2.

A non-trivial subgroup defining a boundary is H3 = ⟨(0, 1)⟩ on which ωq evaluates to 1. It only has a trivial two-cocycle class, giving rise to the lagrangian subgroup
(112)
independent of q.
For q = 0, there are p boundaries associated to the subgroup H4=G=Zp×2 [see Eq. (49)], parametrized by mZp. Plugging a non-trivial two-cocycle Ωm into Eq. (108) yields the associated Lagrangian subgroups
(113)
We see that the non-trivial two-cocycle on H4 (m ≠ 0) “couples” the flux with the charge that is condensed at the associated boundary.

3. Vecω(S3)

For the untwisted Vec(S3) model, there are four inequivalent boundaries, one for each subgroup of S3: H0, H1, H2 and H3 (as defined in Sec. III E 3). Let us first consider the normal subgroup H1Z3. It defines a valid boundary for the trivial bulk cocycle as well as the cocycle ω3 [see Eq. (51)]. The associated Lagrangian algebras in these two cases are
(114a)
(114b)
The boundary anyons are labeled by a double coset xH1\S3/H1 = {H1, H1tH1} and irreducible representations of H1, κ ∈ {0, 1, 2} all of which are one-dimensional. The multiplicities for the bulk-boundary fusion events can be found in Table IX.
As a second example, we consider the other non-trivial subgroup H2. It defines a boundary in the untwisted case as well as for the non-trivial three-cocycles ω2 and ω4 [see Eq. (51)]. The boundary anyons are labeled by a double coset xH2\S3/H2 = {H2, H2rH2} and an irrep of H2, κ ∈ {0, 1}. Interestingly, the fusion multiplicities are the same for both twisted models and the untwisted model. The associated Lagrangian algebra is
(115)
The explicit fusion multiplicities, also for non-trivial boundary defects, can be found in Table X.

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 Z2. On the other hand it can be described as the phase obtained when condensing e2 m2 in the Z4 toric code, the (untwisted) quantum double of Z4. More generally, a ZN quantum double twisted by a type-I cocycle ωIn [see Eq. (47)] is equivalent to a condensate of a quantum double of ZN2 (where eNmN is condensed). Such a condensation process can be viewed as a non-invertible domain wall between the parent phase D(Vec(ZN2)) and the condensate D(VecωI(ZN)). 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 ZN becomes trivial when considered as a cocycle of a larger subgroup of ZN×ZN2.

Via the folding trick (see Sec. III B 1), a domain wall between an untwisted ZN2 and a ZN quantum double twisted by a type-I cocycle ωIn [see Eq. (47)] corresponds to a boundary of a VecωIn(ZN2×ZN) model, where ωIn is trivial on ZN2 and of the form of Eq. (47) on ZN. In the following we show that the subgroup HcondI=(1,1)ZN2ZN×ZN2 defines a valid boundary and that it corresponds to the condensation domain wall discussed in the previous paragraph when unfolded.

On HcondI, the three-cocycle reads
(116)
and A, B, C are a, b, c mod N. In fact, it is a coboundary of the cochain
(117)
which we can check explicitly
(118a)
(118b)
where we have used that addition of the integers in the exponent is modulo N2. This shows that ωIn is a coboundary on HcondI and with that, that this subgroup defines a valid boundary with L-symbols given by ψn [see Eq. (41)]. In Table I we list the fusion multiplicities for N = 2 and n = 1 which align with what one would expect from the anyon condensation picture, a transition from the Z4 toric code to the double-semion phase. In general, if H has non-trivial two-cocycles, we can multiply the trivializing cochain with such a two-cocycle to obtain a boundary of a different type.
TABLE I.

Fusion multiplicities at the the domain wall implementing the condensation from a Z4 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.

1bs̄s
1 = (0, 0) 
e = (0, 1) 
e2 = (0, 2) 
e3 = (0, 3) 
m = (1, 0) 
me = (1, 1) 
me2 = (1, 2) 
me3 = (1, 3) 
m2 = (2, 0) 
m2e = (2, 1) 
m2e2 = (2, 2) 
m2e3 = (2, 3) 
m3 = (3, 0) 
m3e = (3, 1) 
m3e2 = (3, 2) 
m3e3 = (3, 3) 
1bs̄s
1 = (0, 0) 
e = (0, 1) 
e2 = (0, 2) 
e3 = (0, 3) 
m = (1, 0) 
me = (1, 1) 
me2 = (1, 2) 
me3 = (1, 3) 
m2 = (2, 0) 
m2e = (2, 1) 
m2e2 = (2, 2) 
m2e3 = (2, 3) 
m3 = (3, 0) 
m3e = (3, 1) 
m3e2 = (3, 2) 
m3e3 = (3, 3) 

2. Condensation domain walls between fully twisted and untwisted quantum doubles

As a second example we consider the “fully twisted” quantum double of ZN×ZN, i.e., we consider a three-cocycle cohomologous to a product of type-I cocycles on both ZN 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 ZN2×ZN2.

Via the folding trick, the domain wall in question is equivalent to a boundary of a Vecω(ZN2×2×ZN×2) model with ω being a product of type-I cocycles [see Eq. (47)] and a type-II cocycle [see Eq. (48)] on both ZN factors,
(119)
where ai,bi,ciZN2 and Ai,Bi,CiZN.
The domain wall that models the condensation transition is associated to the subgroup HcondI,II=(1,0,1,0),(0,1,0,1)ZN2×2×ZN×2. We parametrize it by two ZN2 variables, HcondI,II={(a1,a2,a1modN,a2modN)|aiZN2}. Similar to the previous example we can find a coboundary βn((a1,a2),(b1,b2))=e2πiN2niAi(biBi)+A1(b2B2) whose coboundary is ωnHcondI,II,
(120a)
(120b)
where Ai, Bi, Ci stand for ai, bi, ci mod N. Hence, this subgroup indeed defines a valid boundary. In contrast to HcondI, HcondI,II has N2 − 1 non-trivial two-cocycle classes, represented by two-cocycles of the form of Eq. (49) with p = q = N2. We can obtain the two-cochains ψ defining all different domain walls by multiplying the cochain βn defined above with one of those two-cocycles. We find that a condensing domain wall from Vec(ZN2×ZN2) to Vecω(ZN×ZN) corresponds to non-trivial two-cocycle of order N. Plugging in the defining data into Eq. (108) and reverting the fold, we recover that the associated domain wall indeed implements the condensation from one phase to the other, compare Ref. 36. In Table II we printed a snippet of the condensation table for the domain wall implementing the transition from two copies of the Z4 toric code to the six-semion phase. In particular, we find that the correct toric code anyons either (1) condense, i.e., fuse to 1 in the double-semion phase, (2) confine, i.e., do not fuse to any anyon in the double-semion phase and (3) tunnel to the expected anyon in the double-semion phase.
TABLE II.

Snippet of tunneling table from two Z4 Toric Codes to six-semion model for HcondI,II 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 e22m22,m12e12e221 and two inequivalent semions m2e1e2 and m1e1 get mapped to valid generators of the six-semion model.

1b1b2b3s1̄s1s2̄s2s3s4̄s3s4s5s5̄s6̄s6
1 = (0, 0, 0, 0) 
m2e1e2 = (0, 1, 1, 1) 
m22e22=(0,2,0,2) 
e12=(0,0,2,0) 
m23e13e23=(0,3,3,3) 
m1e1 = (1, 0, 1, 0) 
m12e12e22=(2,0,2,2) 
m13e13=(3,0,3,0) 
⋮                 
1b1b2b3s1̄s1s2̄s2s3s4̄s3s4s5s5̄s6̄s6
1 = (0, 0, 0, 0) 
m2e1e2 = (0, 1, 1, 1) 
m22e22=(0,2,0,2) 
e12=(0,0,2,0) 
m23e13e23=(0,3,3,3) 
m1e1 = (1, 0, 1, 0) 
m12e12e22=(2,0,2,2) 
m13e13=(3,0,3,0) 
⋮                 

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 Z2 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 Z2 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 Z2 (double-semion) model and a twisted Z3 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 Vecω(ZN). Any domain wall between such two models is equivalent to a boundary of a stacked model, for example Vecωω̄(S3×ZN), 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 Vecω(ZN) 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. Vecω(S3)Vecω(Z2)

As a first example, we consider a domain wall between a ω3-twisted S3 model [see Eq. (51)] and the double-semion phase, the twisted Z2 model. After folding, the domain wall corresponds to a boundary of a S3×Z2 model twisted by the three-cocycle
(121)
The non-trivial subgroup H={(x,0,x)|xZ2}Z2 defines a boundary since ω evaluates to 1 on it. It has only trivial two-cocycles. Plugging in the defining data (G=S3×Z2, ω, H, ψ ≡ 1) into Eq. (107) and unfolding yields the tunneling Table III.
TABLE III.

Tunneling table for a Vecω3(S3)Vecω(Z2) domain wall defined by H=(1,0,1)Z2.

1bss̄
(ē,Γ0) 
(ē,Γ1) 
(ē,Γ2) 
(r̄,0) 
(r̄,1) 
(r̄,2) 
(t̄,0) 
(t̄,1) 
1bss̄
(ē,Γ0) 
(ē,Γ1) 
(ē,Γ2) 
(r̄,0) 
(r̄,1) 
(r̄,2) 
(t̄,0) 
(t̄,1) 

Note that the generator of H is supported on S3 as well as on Z2 which in turn makes the domain wall (partly) transparent.

2. Vecω(S3)Vecω(Z3)

Similarly, one can interface an S3 and a Z3 model non-trivially. In this example, we consider a ω4-twisted S3 model [see Eq. (51)] and a type-I twisted Z3 model. After folding, the boundary in question is a boundary of a S3×Z3 model twisted by a three-cocycle of the form
(122)
It evaluates to 1 on the subgroup H=(0,1,1)Z3. There are only trivial two-cocycles on H so it defines a unique boundary. Plugging in the defining data (G=S3×Z2,ω,H,Ψ1) into Eq. (107) and reinterpreting the results via unfolding, we obtain the tunneling table in Table IV. Comparing the two tunneling Tables III and IV, we see that – in a way – r̄ and t̄ have changed roles.
TABLE IV.

Tunneling table for a Vecω4(S3)Vecω(Z3) domain wall defined by H=(0,1,1)Z3.

(0,0)(0,1)(0,2)(1,0)(1,1)(1,2)(2,0)(2,1)(2,2)
(ē,Γ0) 
(ē,Γ1) 
(ē,Γ2) 
(r̄,0) 
(r̄,1) 
(r̄,2) 
(t̄,0) 
(t̄,1) 
(0,0)(0,1)(0,2)(1,0)(1,1)(1,2)(2,0)(2,1)(2,2)
(ē,Γ0) 
(ē,Γ1) 
(ē,Γ2) 
(r̄,0) 
(r̄,1) 
(r̄,2) 
(t̄,0) 
(t̄,1) 

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:

  1. Coupling via the subgroup: The subgroup H cannot be generated a set of generators {(gi,gi)} 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.

  2. Coupling via a two-cocycle: Let H have two subfactors H1G ×{1G} and H2 ⊆{1GG′. 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 Z2 and a Z3 subgroup which allow it to be interfaced with Z2 and Z3 models. The two-cocycles fulfilling condition 2 are in most cases 2-coycles on an (Abelian) subgroup isomorphic to ZN×ZM 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 em duality domain wall in a Vec(Z2) model. Microscopically, it is defined by subgroup of the folded model HZ2×Z2 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 Vec(Z2) 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 

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 Vec(Z2) 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.

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).

The authors have no conflicts to disclose.

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

In this appendix, we summarize the fusion multiplicities at boundaries of an untwisted Vec(Z2×Z2) and Vecω(S3) models with different three-cocycles.

1. Vec(Z2×Z2)

For the Z2×Z2 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 H=G=Z2×Z2. 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 Z2×Z2 model (which is in the phase as the topological color code53) or—equivalently—as domain wall tunneling events between two Vec(Z2) models (toric code phase).

TABLE V.

The fusion multiplicities for all pairs of bulk and boundary anyons for G=Z2×Z2 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/HG, represented by G elements. The fact that they are actually coset labels is marked with an overline.

(0,1)̄(1,0)̄(0,0)̄(1,1)̄
[(0, 1), (0, 1)] 
[(0, 1), (1, 0)] 
[(0, 1), (1, 1)] 
[(0, 1), (0, 0)] 
[(1, 0), (0, 1)] 
[(1, 0), (1, 0)] 
[(1, 0), (1, 1)] 
[(1, 0), (0, 0)] 
[(1, 1), (0, 1)] 
[(1, 1), (1, 0)] 
[(1, 1), (1, 1)] 
[(1, 1), (0, 0)] 
[(0, 0), (0, 1)] 
[(0, 0), (1, 0)] 
[(0, 0), (1, 1)] 
[(0, 0), (0, 0)] 
(0,1)̄(1,0)̄(0,0)̄(1,1)̄
[(0, 1), (0, 1)] 
[(0, 1), (1, 0)] 
[(0, 1), (1, 1)] 
[(0, 1), (0, 0)] 
[(1, 0), (0, 1)] 
[(1, 0), (1, 0)] 
[(1, 0), (1, 1)] 
[(1, 0), (0, 0)] 
[(1, 1), (0, 1)] 
[(1, 1), (1, 0)] 
[(1, 1), (1, 1)] 
[(1, 1), (0, 0)] 
[(0, 0), (0, 1)] 
[(0, 0), (1, 0)] 
[(0, 0), (1, 1)] 
[(0, 0), (0, 0)] 
TABLE VI.

Fusion multiplicities at the boundary of a model where the bulk is defined by G=Z2×Z2 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.

[(1,0)̄, (0, 1)][(1,0)̄, (0, 0)][(0,0)̄, (0, 1)][(0,0)̄, (0, 0)]
[(0, 1), (0, 1)] 
[(0, 1), (1, 0)] 
[(0, 1), (1, 1)] 
[(0, 1), (0, 0)] 
[(1, 0), (0, 1)] 
[(1, 0), (1, 0)] 
[(1, 0), (1, 1)] 
[(1, 0), (0, 0)] 
[(1, 1), (0, 1)] 
[(1, 1), (1, 0)] 
[(1, 1), (1, 1)] 
[(1, 1), (0, 0)] 
[(0, 0), (0, 1)] 
[(0, 0), (1, 0)] 
[(0, 0), (1, 1)] 
[(0, 0), (0, 0)] 
[(1,0)̄, (0, 1)][(1,0)̄, (0, 0)][(0,0)̄, (0, 1)][(0,0)̄, (0, 0)]
[(0, 1), (0, 1)] 
[(0, 1), (1, 0)] 
[(0, 1), (1, 1)] 
[(0, 1), (0, 0)] 
[(1, 0), (0, 1)] 
[(1, 0), (1, 0)] 
[(1, 0), (1, 1)] 
[(1, 0), (0, 0)] 
[(1, 1), (0, 1)] 
[(1, 1), (1, 0)] 
[(1, 1), (1, 1)] 
[(1, 1), (0, 0)] 
[(0, 0), (0, 1)] 
[(0, 0), (1, 0)] 
[(0, 0), (1, 1)] 
[(0, 0), (0, 0)] 
TABLE VII.

The bulk-to-boundary fusion multiplicities in a model where the bulk is defined by G=Z2×Z2 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.

[(1,0)̄, (1, 1)][(1,0)̄, (0, 0)][(0,0)̄, (1, 1)][(0,0)̄, (0, 0)]
[(0, 1), (0, 1)] 
[(0, 1), (1, 0)] 
[(0, 1), (1, 1)] 
[(0, 1), (0, 0)] 
[(1, 0), (0, 1)] 
[(1, 0), (1, 0)] 
[(1, 0), (1, 1)] 
[(1, 0), (0, 0)] 
[(1, 1), (0, 1)] 
[(1, 1), (1, 0)] 
[(1, 1), (1, 1)] 
[(1, 1), (0, 0)] 
[(0, 0), (0, 1)] 
[(0, 0), (1, 0)] 
[(0, 0), (1, 1)] 
[(0, 0), (0, 0)] 
[(1,0)̄, (1, 1)][(1,0)̄, (0, 0)][(0,0)̄, (1, 1)][(0,0)̄, (0, 0)]
[(0, 1), (0, 1)] 
[(0, 1), (1, 0)] 
[(0, 1), (1, 1)] 
[(0, 1), (0, 0)] 
[(1, 0), (0, 1)] 
[(1, 0), (1, 0)] 
[(1, 0), (1, 1)] 
[(1, 0), (0, 0)] 
[(1, 1), (0, 1)] 
[(1, 1), (1, 0)] 
[(1, 1), (1, 1)] 
[(1, 1), (0, 0)] 
[(0, 0), (0, 1)] 
[(0, 0), (1, 0)] 
[(0, 0), (1, 1)] 
[(0, 0), (0, 0)] 
TABLE VIII.

G=Z2×Z2H=G 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, 1)] 
[(0, 1), (1, 0)] [(0, 1), (1, 0)] 
[(0, 1), (1, 1)] [(0, 1), (1, 1)] 
[(0, 1), (0, 0)] [(0, 1), (0, 0)] 
[(1, 0), (0, 1)] [(1, 0), (0, 1)] 
[(1, 0), (1, 0)] [(1, 0), (1, 0)] 
[(1, 0), (1, 1)] [(1, 0), (1, 1)] 
[(1, 0), (0, 0)] [(1, 0), (0, 0)] 
[(1, 1), (0, 1)] [(1, 1), (0, 1)] 
[(1, 1), (1, 0)] [(1, 1), (1, 0)] 
[(1, 1), (1, 1)] [(1, 1), (1, 1)] 
[(1, 1), (0, 0)] [(1, 1), (0, 0)] 
[(0, 0), (0, 1)] [(0, 0), (0, 1)] 
[(0, 0), (1, 0)] [(0, 0), (1, 0)] 
[(0, 0), (1, 1)] [(0, 0), (1, 1)] 
[(0, 0), (0, 0)] [(0, 0), (0, 0)] 
(0, 1)(1, 0)(1, 1)(0, 0)(0, 1)(1, 0)(1, 1)(0, 0)
[(0, 1), (0, 1)] [(0, 1), (0, 1)] 
[(0, 1), (1, 0)] [(0, 1), (1, 0)] 
[(0, 1), (1, 1)] [(0, 1), (1, 1)] 
[(0, 1), (0, 0)] [(0, 1), (0, 0)] 
[(1, 0), (0, 1)] [(1, 0), (0, 1)] 
[(1, 0), (1, 0)] [(1, 0), (1, 0)] 
[(1, 0), (1, 1)] [(1, 0), (1, 1)] 
[(1, 0), (0, 0)] [(1, 0), (0, 0)] 
[(1, 1), (0, 1)] [(1, 1), (0, 1)] 
[(1, 1), (1, 0)] [(1, 1), (1, 0)] 
[(1, 1), (1, 1)] [(1, 1), (1, 1)] 
[(1, 1), (0, 0)] [(1, 1), (0, 0)] 
[(0, 0), (0, 1)] [(0, 0), (0, 1)] 
[(0, 0), (1, 0)] [(0, 0), (1, 0)] 
[(0, 0), (1, 1)] [(0, 0), (1, 1)] 
[(0, 0), (0, 0)] [(0, 0), (0, 0)] 

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.

TABLE IX.

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 Hr\G/Hr×Z3. 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)
(ē,Γ0) 
(ē,Γ1) 
(ē,Γ2) 
(r̄,0) 2(0) 0(1) 0(1) 
(r̄,1) 0(1) 2(0) 0(1) 
(r̄,2) 0(1) 0(1) 2(0) 
(t̄,0) 
(t̄,1) 
(Hr, 0)(Hr, 1)(Hr, 2)(HrtHr, 0)(HrtHr, 1)(HrtHr, 2)
(ē,Γ0) 
(ē,Γ1) 
(ē,Γ2) 
(r̄,0) 2(0) 0(1) 0(1) 
(r̄,1) 0(1) 2(0) 0(1) 
(r̄,2) 0(1) 0(1) 2(0) 
(t̄,0) 
(t̄,1) 
TABLE X.

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 Z2, for HtrHt, the only non-trivial double coset, the stabilizer group is trivial, i.e., Z1. 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)
(ē,Γ0) 
(ē,Γ1) 
(ē,Γ2) 
(r̄,0) 
(r̄,1) 
(r̄,2) 
(t̄,0) 
(t̄,1) 
(Ht, 0)(Ht, 1)(HtrHt, 0)
(ē,Γ0) 
(ē,Γ1) 
(ē,Γ2) 
(r̄,0) 
(r̄,1) 
(r̄,2) 
(t̄,0) 
(t̄,1) 

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.

Definition 1

[(left) G-module]. Consider a finite group (G, ·). An Abelian group (M, *) together with a (left) group action ⊳: G × MM, (g, a) ↦ ga that ∀ a, bM, g, hG fulfills

  • g ⊳ (a*b) = (ga)*(gb) and

  • (g · h) ⊳ a = g ⊳ (ha),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.

Remark.

Any Abelian group can be made into a G-module with a trivial group action by defining ga = a ∀aM, gG.

Definition 2

(n-cochain). Let (M, *) be a G-module. A map ηn: G×nM is called n-cochain (of G over M). We denote the space of all such n-cochains by Cn(G,M). In fact, Cn(G,M) is a group with the group multiplication inherited from M.

Definition 3
[(twisted) n-coboundary]. Let (M, *) be a G-bimodule, potentially with trivial (left/right) G-action. The map δn:Cn(G,M)Cn+1(G,M) defined by
(B1)
is called n-coboundary operator.

Most importantly, for any n and ηCn(G,M), (δn+1δn)(η)=1Cn(G,M), where 1Cn(G,M) 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).

Remark.

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 δ̃.

Definition 4
[(twisted) cohomology groups]. Let (M, *) be a G-module. The cochains in the kernel of δn,
(B2)
are called n-cocycles. They form a subgroup of the group of cochains Cn(G,U(1)). The n + 1 cochains in the image of δn,
(B3)
are called n-coboundaries.
The quotient group
(B4)
is called nth cohomology group of G over M.

Remark.

If two n-cocycles are in the same equivalence class in Hn(G, M) we call them cohomologous.

In any n-cocycle class we can find a representative ωZn(G,M) that fulfills
(B5)
We call such a cocycle normalized.

Definition 5
(slant product). Let (M, *) be a G-module. For any xG and integer n > 1, the map ixn:Cn(G,M)Cn1(G,M),
(B6)
is called slant product. Again, we write ix=nixn to abbreviate the collection of all slant products.

Remark.

ign 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 (δ̃n1in)=(in+1δn). 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.

Example B.1
[ZN]. The cyclic group ZN is one of the few finite groups where one can simply derive all its cohomology groups. Following Ref. 23 we get
(B7)

Example B.2
[ZN×ZM]. Using the Künneth formula for group cohomology,23 we can relate the cohomology groups of ZN×ZM to the ones of ZN and ZM. For n = 0, 1, 2, 3, we obtain
(B8a)
(B8b)
(B8c)
(B8d)

Example B.3
[Zn1×Zn2××Znk]. To calculate the cohomology groups of product groups, like Zn1×Zn2××Znk, we can use the Künneth formula.23 For n = 0, 1, 2, 3, we obtain
(B9a)
(B9b)
(B9c)
(B9d)

Remark.

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.

Example B.4
[S3]. Following Refs. 8 and 28 the first three cohomology groups of S3 are given by
(B10a)
(B10b)
(B10c)
(B10d)

Let (G, ·) be a finite group and X a left G-set. In particular, there exist a map ⊳: G × XX such that
(C1)
Given such a G-action we consider the G-module U(1)X with the non-trivial (right) action defined by
(C2)
An n-cochain for this G-module can be represented by a map ψ: G×n × G/HU(1). In this representation, the twisted coboundary operator acts as
(C3)
Every G-set X is isomorphic to a disjoint union of cosets
(C4)
for a collection of subgroups {HiG}. G acts transitively on each component by left-translation g ⊳ (xHi) = (g · x)Hi. This makes U(1)X a Cartesian product of decoupled modules, so the total cohomology group is isomorphic to a product
(C5)
The goal of this appendix is to show the following isomorphism that holds for each component:
(C6)
This isomorphism is induced by the following two maps
(C7)
The easy direction of the isomorphism is given by , the embedding of H×n into G×n × G/H, i.e.,
(C8)
The map in the converse direction is obtained from first choosing a representative for each of the cosets. Let (gn, …, g1, α) ∈ G×n × G/H and consider the cosets α, g1α, …, (gng1) ⊳ α. We denote their respective representatives by r0, r1, …, rn. With that, m(n) is defined by
(C9)
The image is guaranteed to lie in H×n because, by construction, gk(rk−1H) = rkH ∀k and hence ri1giri1H,i.

Remark.

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

We can apply m(n), respectively , on cochains ψCn(G,U(1)G/H) and ρCn(H,U(1)), via precomposition
(C10)
In fact, both are chain maps, i.e. mapping cocycles onto cocycles. This can be seen with a straightforward calculation
(C11a)
(C11b)
(C11c)
where ri is the chosen representative for coset (gigi−1g0) ⊳ α and r the representative of α. The converse, δρ̂=δ̃ρ̂ follows simply from the fact that the G-action on U(1)G/H becomes trivial when every argument of the cochain is restricted to H.
To complete the proof, we have to show that ̃ and ̂ induced by m(n) and are inverses of each other up to coboundaries and thereby preserve the cohomology class. One direction is easy to see from
(C12)
To show the converse, we explicitly construct a n − 1-cochain Ω: Gn−1 × G/HU(1) such that
(C13)
namely,
(C14)
with r0 being the representative for α and ri for (gn−1g1) ⊳ α. This can be seen explicitly by rewriting ψα(gn, …, g1) with (δ̃ψ)α(gn,gn1,,g1,r0)=1 in terms of ψH.

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.

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

The bulk of the twisted quantum double model is a state-sum path integral on three-dimensional space-time triangulations with a branching structure consisting of tetrahedra,
(D1)
At every edge of the triangulation there is a state-sum variable (such as a, b, and c above) taking values in the group G. On every triangle,
(D2)
the group labels a, b, and c have to satisfy the constraint
(D3)
At every tetrahedron with labels as in Eq. (D1), we have a weight
(D4)
Note that the three labels a, b, and c determine all other labels of the tetrahedron via the triangle constraint. If the tetrahedron has the opposite orientation from Eq. (D1), we instead associate the complex conjugate ω̄. This is how unitarity/Hermiticity is implemented in the state-sum language. Retriangulation invariance is then imposed via Pachner moves as shown in Eq. (23) in the main text. This makes ω a group three-cocycle representing an element of H3(G, U(1)).

2. Boundary

Next we include boundaries. The normal space to a point in the boundary is a half-line, and an ɛ-sphere restricted to that half line is a single point. So the link of the boundary is a point, and an extended cellulation is essentially just a three-dimensional triangulation with boundary. However, it better to “thicken” the boundary triangles into three-cells which look like triangle prisms,
(D5)
The bottom triangle formed by the three edges labeled a, b, c is attached to a bulk tetrahedron, the three rectangles on the sides are attached to other boundary three-cells, and the top triangle formed by the three dotted edges corresponds to the actual boundary. By replacing boundary triangles with boundary three-cells, we can also make sense of two boundary triangles separated by an “infinitely thin” bulk as for example later in Eq. (D27). At every boundary vertex [or more precisely, every of the short thick blue lines in Eq. (D5)], there is a state-sum variable taking values in some finite set A, which is equipped with a right action of G,
(D6)
At each “thickened” boundary edge,
(D7)
such as at the three sides of the boundary triangle in Eq. (D5), the labels fulfill the constraint
(D8)
To each triangle as in Eq. (D5), we associate a weight
(D9)
noting that the labels c, β, and γ are determined from α, a, and b via the constraints. If the triangle has opposite orientation from Eq. (D5), we associate the complex conjugate instead, just as for the bulk tetrahedra. Note that this Hermiticity condition holds generally for all weights/cells in a state-sum, and we will in the following always assume it without explicitly saying so.
Topological invariance can be imposed by a move attaching/removing a tetrahedron to the boundary,
(D10)
The same move was depicted slightly differently in Eq. (29) in the main text. On the left side, there are two boundary triangles only, drawn with the bulk at the back such that the dashed blue lines are in front of the black lines with arrows. On the right hand side, we have two boundary triangles, and a bulk tetrahedron attached to both at the back. The boundary triangles on both sides alone are related by a 2-2 Pachner move. So ψ has to fulfill
(D11)
which is the same (apart from slightly different conventions) as in Eq. (41) in the main text.

3. Anyons

Anyons in the state-sum can be represented by considering triangulations of three-manifolds with an embedded one-manifold representing the anyon world-line. The link of an embedded one-manifold is a circle, which can be triangulated with a single “looping” edge whose end vertices coincide,
(D12)
Since the one-manifold itself is triangulated with edges, we need to trianglate an edge times the looping edge. We pick a triangulation of this “tube segment” with two triangles,
(D13)
Even though such a tube segment is formally something two-dimensional, it is good to think of it as a cylinder-like three-cell, with a red shaded 1-gon at the top and bottom. An extended cellulation is a three-manifold triangulation into which we embed sequences of tube segments as above. The two red shaded 1-gons attach to other tube segments, whereas the two triangles wrapping round the side attach to bulk tetrahedra. Those triangles may also be directly attached to the triangles of other tube segments with an infinitely thin bulk in between.
The state-sum associates one additional variable to each red shaded 1-gon where two tube segments meet,
(D14)
The dimension of this variable μ may to depend on the group label a at the corresponding loop edge. To each tube segment with labels as in Eq. (D13), we associate a weight
(D15)
The topological invariance is again imposed by combinatorial moves. This time, we have two tube segments stacked on one side, and one tube segment on the other side. The one tube segment needs to be surrounded by bulk tetrahedra such that the two sides become compatible.
In order to concisely state the axioms giving rise to topological invariance, and to classify the different anyons, it is useful to apply a dimensional reduction, more specifically a compactification. To this end, consider a mapping from two-dimensional triangulations to three-dimensional ones by taking the Cartesian product with the circle. Pulling back the three-dimensional state-sum along this mapping we obtain a two-dimensional state-sum. On a combinatorial level, we take the Cartesian product of a two-dimensional triangulation with the triangulation of the circle link, i.e., the looping edge in Eq. (D12). Every triangle then becomes a “triangle prism” with top and bottom identified,
(D16)
We need to choose a simple three-dimensional triangulation for this triangle prism which can be done with three tetrahedra as shown. We then redistribute the (independent) three-dimensional state-sum variables onto the two-dimensional triangulation as shown. This way, we get a two-dimensional state-sum with one G-variable at each edge and one at each vertex. At each edge,
(D17)
we have a constraint y = c−1xc. The state-sum weight at each triangle is
(D18)
One of the tetrahedra above has a different orientation as in Eq. (D1), thus the according weight ω is complex conjugated. Up to differences in conventions, this is Eq. (58) from the main text. By construction, the topological invariance of the two-dimensional state-sum follows from the topological invariance of the three-dimensional state-sum. The resulting state-sum is a thick state-sum as discussed in  Appendix E, whose vertex label set is G, with G-action
(D19)
Accordingly, β is a G two-cocycle with module U(1)G, such that gG acts on ϕU(1)G by gϕ(x) = ϕ(xg−1).
Next, we extend the above mapping to a mapping from two-manifolds with boundary to three-manifolds with anyon world-lines. We do this by mapping the one-dimensional boundary to a one-dimensional anyon world-line. Combinatorially, we map a “thickened” boundary edge to an anyon tube segment. Pulling back the anyon state-sum along this mapping yields a two-dimensional boundary state-sum,
(D20)
Now, the boundary retriangulation invariance of the compactified boundary,
(D21)
is equivalent to the retriangulation invariance of the anyon world-lines. Thus, we see that line-like defects are in one-to-one correspondence with the boundaries of the compactified two-dimensional state-sum. So the weights (ρa(c))μν are determined by the equation
(D22)
Note that there is a finite number of irreducibleρ, which are what is commonly understood by anyons. To see this we notice that a two-dimensional state-sum with variables only on the edges is an associative (in fact, †-Frobenius-) algebra, and is a direct sum of full matrix algebras. We can copy the vertex variables of our state-sum onto each of the adjacent edges, such that at every edge we have a triple (x, a, y). The triples have to satisfy y = xa and are thus fully specified by (x, a). The resulting algebra equals the tube algebra discussed in the literature and in the main text in Sec. IV. Its irreducible matrix blocks, or the according irreducible representations, are the anyons.

4. Boundary anyons

To describe anyons within the boundary we need to define the state-sum on triangulations of manifolds with a one-manifold embedded into the boundary. The normal space to the one-manifold is a half-plane, and an ɛ-sphere within this half-plane is an interval. Thus the link is an interval, and it can be triangulated with only two (thickened) boundary vertices separated by an infinitely thin bulk,
(D23)
The tick on the right short blue line is in order to remove the symmetries of the link. Next we take the product of that link with a single edge,
(D24)
It is again helpful to think of this as a three-cell which looks like a little triangle wedge, by imagining another rectangle face between the two dotted lines at the back, such that there is a red shaded triangle on the bottom and top. Then an extended cellulation contains sequences of such “semi-tube segments.” The red shaded triangles are attached to other semi-tube segments, whereas the two rectangular faces on the left and right are attached to the sides of nearby boundary triangles. At every red shaded triangle where two semi-tube segments meet,
(D25)
there is one additional variable μ. Those variables take values in some finite set X which may depend on the value of the adjacent boundary vertex labels α and α′. To each semi-tube segment with labels as in Eq. (D24), we associate a weight
(D26)
noting that the values of β and β′ are determined by the other labels.
Again, for stating the topological invariance of the boundary anyons and classifying them, we apply a compactification by pulling back the state-sum via the Cartesian product with the interval link in Eq. (D23). Each triangle is mapped to a sandwich of two boundary triangles,
(D27)
We obtain a two-dimensional state-sum with one G-label at each edge and two A-labels at each vertex. The weight associated to each triangle is
(D28)
Note that one of the boundary triangle cells is reflected compared to Eq. (D5) and thus complex conjugated. The resulting two-dimensional state-sum is a thick state-sum with vertex set A × A and right action given by
(D29)
Accordingly Ψ is a G two-cocycle with module U(1)A×A and left action determined by the above.
As for bulk anyons, we extend the above to a mapping from two-manifolds with boundary to three-manifolds with boundary and boundary anyons, where the one-dimensional boundary is mapped to a one-dimensional boundary anyon world-line. Combinatorially, we map a boundary edge of a two-dimensional triangulation to a semi-tube segment,
(D30)
Again, the topological invariance of boundary-anyon world-lines is equivalent to the topological invariance of the boundary above as in Eq. (D21). So boundary anyons are determined by the equation
(D31)
Again, what is commonly meant by boundary anyons are the irreducible κ. When we copy vertex labels onto edges in our two-dimensional state-sum we obtain the semi-tube algebra discussed in the main text. The boundary anyons are the irreducible representations of this semi-tube algebra.

5. Ground states

We can also consider zero-dimensional defects in space-time, i.e., embedded points. If the link of such a defect is not a two-sphere, then the embedded points correspond to some kind of singularity. Such defects are in one-to-one correspondence with the ground states of the model on the two-manifold which is the link. To represent such a defect combinatorially in the most concise way, we pick a minimal triangulation of the two-manifold link. For a torus, for example, we can choose a triangulation with two triangles,
(D32)
with left/right and bottom/top identified. It might be intuitive to think of this triangulation as a solid-torus-like three-cell, but we should keep in mind formally there is only the two-dimensional triangulation, and a more canonical filling would be given by a cone with a singularity in the center. A combinatorial extended manifold is thus a three-manifold with some torus three-cells. At every torus three-cell there is a weight
(D33)
We want the zero-dimensional defects defined by the torus three-cell to be topological. Roughly, this means that a layer of bulk padding the torus three-cell can be absorbed into it. To concisely impose this topological invariance, we again use compactification. We pull back the state-sum along a mapping from one-manifolds to three-manifolds. Combinatorially, an edge is mapped to an edge times the torus link,
(D34)
where front/back and left/right are identified. We thus get a one-dimensional state-sum with one G-variable at every edge and two G-variables at every vertex, subject to the constraints d = x−1bx and c = x−1ax at every edge with labels as above. The weight associated to every edge is
(D35)
which can be obtain from a triangulation of Eq. (D34) with six tetrahedra which is not shown. The topological invariance of this state-sum follows by construction from the topological invariance of the original state-sum. This is a thick one-dimensional state-sum with vertex labels in G × G and right action
(D36)
Accordingly, P is a G one-cocycle with module U(1)G×G, and left action determined by the above. We extend the mapping to a mapping from one-manifolds with boundary to three-manifolds with torus-link point defects, by mapping a boundary point to a torus three-cell,
(D37)
such that topological invariance
(D38)
of the one-dimensional state-sum boundary is equivalent to that of the torus-link point defects. Spelled out in letters, both are given by
(D39)
which can be rephrased as dS = P, if we interpret S as a twisted zero-cochain. This equation is linear in S, thus the ground states form a vector space, which is a feature of all zero-dimensional defects. one-dimensional state-sums with variables only on the vertices are given by projectors, and our state-sum can be brought into such a form by simply summing over x in Pa,b(x). Ground states Sa,b are then vectors in the support of this projector.

6. Bulk fusion events

The link of a zero-dimensional defect might itself have defects. For example, take as link a two-sphere with three embedded points,
(D40)
Such defects correspond to points in space-time where three anyon world-lines meet. The points in the link are just the intersections of a sphere around the zero-dimensional defect with those three anyon world-lines. At the same time, those point-like defects are ground states on a sphere with three anyons.55 Here we will look at the fusion vertices between three different anyon world-lines which weights ρ0, ρ1 and ρ2, in which case it would be natural to color the points differently.
This sphere link can be triangulated by one triangle (formed on the “outside” enclosed by all three edges in the following picture) with three red shaded 1-gons,
(D41)
Those three 1-gons are attached to three tube segments corresponding to the anyon world-lines, and the outside triangle is attached to a bulk tetrahedron. As usual we may think of this as a “junction” three-cell. So extended cellulations are cellulations including sequences of tube segments (of three different colors) and junction three-cells where three tube segments meet. To each junction three-cell, we associate a weight
(D42)
As usual, we consider the one-dimensional state-sum arising from compactification with the link above. Combinatorially, we take the triangulation in Eq. (D41) times an edge, and choose a three-dimensional triangulation for the resulting volume,
(D43)
On the right-hand side, the three edges labeled x are identified, such that the edges a, b, c, d, e, and f form loops, which are the start and end loops of three tube segments as in Eq. (D13). The labels α, β, γ, δ, ɛ, and ϕ are the labels at the bottom and top of the three tube segments, whose red 1-gons have not been drawn.. The resulting one-dimensional state sum has one G-label at every edge, and two G-labels and three X-labels at every vertex, subject to the constraints d = x−1ax, e = x−1bx at every edge. The weight associated to each edge is
(D44)
coming from the cellulation using three tetrahedra and three tube segments. The resulting one-dimensional state-sum has vertex labels in G × G with action as in Eq. (D36), as well as three free vertex labels. The fusion events are in one-to-one correspondence with boundaries of this one-dimensional state-sum. The topological invariance is thus determined by
(D45)
As in the previous section, summing over x yields a projector, and S are vectors in the support of this projector. The dimension of this support vector space is the ground space dimension on the sphere with three anyons, or equivalently the fusion multiplicity Nρ2ρ0ρ1. It can be calculated by taking the trace of the projector, or another compactification to a zero-dimensional state-sum via the Cartesian product with the circle in Eq. (D12). Plugging Eq. (D43) into this compactification means summing and identifying a and d, b and e, c and f, α and δ, β and ɛ, γ and ϕ, yielding
(D46)

7. Bulk-boundary fusion events

Next, let us consider fusion events at the boundary, between a bulk anyon and a boundary anyon. Those are zero-dimensional defects with link
(D47)
A simple triangulation of this link consists of only one (thickened) boundary edge, an anyon 1-gon and a boundary-anyon triangle,
(D48)
Combinatorial extended manifolds are three-manifolds with bulk tube segments and boundary wedge segments, meeting at “junction” three-cells like above. At every such junction, we have a weight
(D49)
noting that α′ is determined by α′ = αa.
As usual, we pull back the three-dimensional state-sum along the Cartesian product with the link in Eq. (D47). Concretely, an edge of a one-dimensional triangulation is mapped to its Cartesian product with Eq. (D48),
(D50)
The right hand side consists of one tube segment, one semi-tube segment, as well as one thickened boundary 4-gon, which can be triangulated with two boundary triangles. The resulting one-dimensional state-sum has one G-label at each edge, and one G-label, one A-label, one anyon-space label, and one boundary-anyon-space label at each vertex. Those labels are subject to the constraints β = αc and b = c−1ac. For an anyon ρ and boundary anyon κ, the weight associated to an edge is
(D51)
So bulk-to-boundary anyon fusion events are determined by
(D52)
If we want to know the fusion multiplicity for a bulk anyon ρ and boundary anyon κ, we need to evaluate the one-dimensional model on a circle. Equivalently, we identify the top and bottom in Eq. (D50), yielding
(D53)
where S(α) is the stabilizer of α, i.e., the set of G-elements c such that αc = α. When both the boundary as well as ρ and κ are irreducible, we have A = A\G, ρ corresponding to a conjugacy class C and κ corresponding to a double-coset x, we get
(D54)
Note that x labels subsets of labels (α, α′) ∈ H\G × H\G with transitive action via x = αα−1. The condition aα−1 arises from this together with α′ = αa. Moreover, cZ(a) arises from b = c−1ac = a, cα−1α from β = αc = α, and ca−1α−1αa from β′ = αc = α′. Eq. (D54) is equal to Eq. (105) from the main text, apart from some choices of conventions.

8. Bulk anyon F-symbol

The F-symbol of the resulting anyon theory does not correspond to a defect like in all the previous sections, but simply to the evaluation of the state-sum on a specific extended cellulation. The corresponding extended manifold is a three-sphere with six anyon world-lines meeting at 4 three-valent fusion point defects, forming the one-skeleton of an embedded tetrahedron. Now, the F-symbol is not a single state-sum evaluation, but the collection of such evaluations for all different anyons and fusion point defects. For the anyons, we choose one irreducible representation of the tube algebra from each such isomorphism class. For the fusion point defects, we use an isomorphism from some abstract fusion vector space into the vector space of all point defects. I.e., instead of a single Sαβγa,b as in Eq. (D42) fulfilling Eq. (D45), we consider
(D55)
fulfilling
(D56)
More precisely, we can choose S[ρa, ρb, ρc] to be an isometry (and we should do so in practice), such that the dimension of χ is the rank of P, namely, Nρcρa,ρb.
Now we choose an explicit extended cellulation. The smallest one consists of only four fusion point defects and one bulk tetrahedron,
(D57)
Above we only drew the bulk tetrahedron. All four vertices are identified such that all six edges form loops. Each of the triangles is the triangle in one of the four junction three-cells shown in Eq. (D41). With this, the evaluation is given by
(D58)

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

Recall that a two-dimensional group-cohomology state-sum is defined on two-dimensional branching structure triangulations with one group label at every edge, and one constraint ab = c and one weight ω(a, b) at every triangle as in Eq. (D2). The thin-boundary state-sum is defined for a subgroup HG and a H one-cochain ψ such that = ω|H, or more explicitly
(E1)
for all a, bH. At every boundary edge,
(E2)
the variables only take values in H and not G, and we have an associated weight ψ(h). The simplicity of the boundary comes at the expense of having a slightly more complicated topological invariance. A move with only one bulk triangle on one side as for the thick boundary in Eq. (D21) does not hold. Instead, the moves
(E3)
and
(E4)
do follow from Eq. (E1).

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 = αg 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.

An irreducible thick boundary ψ̂ (i.e., one with transitive action) can be obtained from a thin boundary ψ with the same subgroup H as follows. We start by padding each thin boundary edge with little bulk rectangle, and flatten those rectangles to get a generalized thick boundary,
(E5)
which has labels in G instead of H\G at the boundary vertices. It is easy to see that the moves for the thick boundary such as Eq. (D21) follow from the thin boundary moves such as in Eqs. (E3) and (E4), after we plug in Eq. (E22). We say “generalized” because the label y is not determined by x and a via some action, but only constrained through h := xay−1H.
In order to obtain a proper thick boundary ψ̂, we realize that the state-sum on the right of Eq. (E5) has a gauge symmetry acting on the edges adjacent to a boundary vertex,
(E6)
for all γH, where the equality is for the weights at the surrounding triangles and boundary edges. On the left of Eq. (E5), the same gauge symmetry is only acting on a single vertex label,
(E7)
Thus, we remove the gauge freedom by replacing the vertex labels x, yG by left cosets in α, βH\G. The constraint β = αa of the thick boundary then directly follows from xa = hy and hH on the right. The weight ψ̂ after gauge fixing can be obtained from the weight before gauge fixing by choosing a standard representative R(α) ∈ G for each coset α, and using
(E8)
The weight can then be read off from the right-hand side of Eq. (E5),
(E9)
using the shortcut in Eq. (E8).
Vice versa, a thin boundary ψ̃ can be obtained from an irreducible thick boundary ψ by realizing that the thick boundary has the following gauge symmetry around a boundary vertex: Change every ingoing adjacent edge by a, every outgoing adjacent edge by aγ−1a, and the coset label at the vertex itself by ααγ, for any γG. Since every such gauge symmetry only affects a single boundary vertex, we can use it to fix all the coset labels to the trivial coset H. So ψ̃ is obtained from ψ by simply setting the coset label to H, which automatically restricts the group label to the subgroup H,
(E10)
for aH.
It is easy to see that if we first transform a thin boundary into a thick one, and go back to a thin one, we end up with the same thin boundary again,
(E11)
However, going from a thick to a thin and then back to a thick boundary does not yield the same boundary, but one that is in the same phase/cohomology class. In order to see this, we consider the following domain wall η between the thick boundaries ψ and ,
(E12)
using x = R(α) and y = R(β) = R(αa) as usual. On the right-hand side, we read off
(E13)
in accordance with Eqs. (E10) and (E9). The weight
(E14)
associated to the vertical thick boundary edge on the right-hand side defines the weight of the domain wall on the left-hand side. With this domain wall, ψ and are related by
(E15)
That is, ψ and are twisted one-cochains which differ by a twisted one-coboundary .

2. 2 + 1 dimensions

Let us now look at the thin boundary in 2 + 1 dimensions. Again, it is defined for HG and a H two-cochain ψ with = ω|H, that is,
(E16)
Again, we restrict the group labels at the boundary edges to H, and associate to every boundary triangle,
(E17)
the constraint ab = c and the weight
(E18)
Again, the topological moves that hold are slightly more complicated. In fact, the move in Eq. (D10) does still hold for the thin boundary,
(E19)
This is because all involved edges labels are constrained to H and so this is just Eq. (E16). However, the thick boundary is invariant under additional moves, such as
(E20)
This move consists of one boundary triangle and one bulk tetrahedron on the left, and three boundary triangles only on the right. This move does not hold for the thin boundary, since the three edges adjacent to the central vertex are constrained to H on the right, but can take values in all of G on the left. This can be fixed by padding the right-hand side with three bulk tetrahedra.

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.

Let us discuss the relation between the thin and thick boundary. We can construct a transitive thick boundary ψ̂ from a thin boundary ψ as follows. Analogous to the 1 + 1-dimensional case, we start with a generalized thick boundary obtained by padding each thin boundary triangle with a triangle prism,
(E21)
It is easy to see that the moves for the thick boundary in Eqs. (D10) and (E20) follows form the thin boundary moves such as in Eq. (E19), after we plug in Eq. (E21).
The H boundary group labels on the right are determined by the other labels α, a, bG so they do not appear on the left. Just as in the 1 + 1-dimensional case, we have a gauge symmetry on the right involving all edges adjacent to a fixed boundary vertex, which becomes a gauge symmetry acting on a single boundary vertex label on the left. Again, this allows us to replace the G-elements x, y, z on the left by left cosets α, β, γ. Then using a triangulation of the above prism similar to Eq. (D16), we find
(E22)
using
(E23)
where R denotes a choice of standard representative of every coset as in the previous paragraph.
Vice versa, we can construct a thin boundary ψ̃ from a thick boundary ψ by realizing that the thick boundary has a gauge symmetry acting on a single boundary vertex label. This gauge freedom can be fixed by setting all the cosets to the trivial coset H, yielding
(E24)
for a, bH. As in the 1 + 1-dimensional case, we have
(E25)
but not vice versa. Similar to the 1 + 1-dimensional case, and ψ are separated by a one-dimensional domain wall which associates a weight η to the according edges. This weight comes from flattening the vertical side of a “step” representing the domain wall. While in Eq. (E12) the side step is an edge flattened to a point, here we have a rectangle flattened to an edge,
(E26)
So the weight is given by
(E27)
using Eq. (E23). The relation between ψ and is now
(E28)
The generalization of thick vs thin boundaries to higher dimensions is straightforward. To map a thin to a thick boundary of an n-dimensional space-time bulk analogous to Eq. (E5), pad the n − 1-simplex with the n − 1-simplex times an edge. The resulting n-cell can be triangulated using n n-simplices. The edge labels of the edges perpendicular to the boundary are set to standard representatives of the cosets of the thick boundary. Evaluation of the space-time volume yields an expression with n bulk weigths ω and one thin-boundary weight ψ. Moreover, the domain wall η between ψ and analogous to Eq. (E27) is obtained from triangulations of a boundary n − 2-simplex times an edge, yielding a formula with n − 1 times ψ.

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.

The mapping in Eq. (E22) for a one-dimensional boundary of a two-dimensional bulk can also be used to map a standalone one-dimensional thin state-sum to a thick state-sum (with the same H),
(E29)
using the short-hand notation x: = R(α) and y: = R(αa) as in Eq. (E8). Up to some conventions, this is just precomposition with the cohomological isomorphism m(1) discussed in  Appendix C. Vice versa, Eq. (E10) can also be used to map a standalone thick state-sum ψ to a thin state-sum ψ̃. Furthermore, the relation in Eq. (E15) with Eq. (E14) still can be used to relate ψ and .
Also the mapping in Eq. (E21) for two-dimensional boundaries of three-dimensional bulks can be used for two-dimensional standalone bulks,
(E30)
with z := R(αab) as in Eq. (E23). This is m(2) up to conventions. Vice versa, the Eq. (E24) also maps from a standalone thick state-sum ψ to a thin state-sum ψ̃, and Eq. (E28) with Eq. (E27) relate ψ and . The generalization to higher dimensions is straightforward.

4. Boundaries of thick bulk from boundaries of thin bulk

Let us now consider the case where the bulk itself is a transitive thick state-sum in 1 + 1 dimensions given by ωα(a, b), which we map to a thin state-sum given by
(E31)
for a, b, cH. Now, consider a generalized thick boundary of the thin state-sum. Here, generalized means that the set of labels at the vertices is not equipped with G-action. At a boundary edge,
(E32)
the two vertex labels μ, ν are independent and do not have to satisfy any constraints. The label set of μ and ν (which can also be understood as a vector space since there is a unitary gauge symmetry acting on μ and ν) is allowed to depend on the value of the group label aG. The according weight is
(E33)
On the other hand, a generalized thick boundary of a transitive thick state-sum has edges,
(E34)
with aG, two bulk coset labels α and β, and two free labels μ, ν as before. The according weight is
(E35)
Now we can obtain a thick boundary of the thick state-sum from the thick boundary of the thin state-sum by padding it with a layer of thick bulk. When doing so, we restrict the bulk coset labels at the boundary to the trivial coset H,
(E36)
As in the previous sections, h on the right can be inferred from the other labels. Since Hx = α, x is a representative of α, and the same holds for y and β. Evaluating the right-hand side above yields
(E37)
where x and y are chosen representatives of α and β as in Eq. (E8).

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 ψ(h)μν 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.

1.
X.-G.
Wen
, “
Topological orders in rigid states
,”
Int. J. Mod. Phys. B
04
,
239
(
1990
).
2.
M. A.
Levin
and
X.-G.
Wen
, “
String-net condensation: A physical mechanism for topological phases
,”
Phys. Rev. B
71
,
045110
(
2005
).
3.
A.
Kitaev
, “
Anyons in an exactly solved model and beyond
,”
Ann. Phys.
321
,
2
111
(
2006
).
4.
B. M.
Terhal
, “
Quantum error correction for quantum memories
,”
Rev. Mod. Phys.
87
,
307
346
(
2015
).
5.
A.
Bauer
,
J.
Eisert
, and
C.
Wille
, “
A unified diagrammatic approach to topological fixed point models
,”
SciPost Phys. Core
5
,
038
(
2022
).
6.
Y.
Hu
,
Y.
Wan
, and
Y.-S.
Wu
, “
Twisted quantum double model of topological phases in two dimensions
,”
Phys. Rev. B
87
,
125114
(
2013
).
7.
V. G.
Turaev
and
O. Y.
Viro
, “
State sum invariants of 3-manifolds and quantum 6j-symbols
,”
Topology
31
,
865
902
(
1992
).
8.
J. C.
Bridgeman
and
D.
Barter
, “
Computing data for Levin-Wen with defects
,”
Quantum
4
,
277
(
2020
).
9.
D.
Barter
,
J. C.
Bridgeman
, and
C.
Jones
, “
Domain walls in topological phases and the Brauer–Picard ring for Vec(Z/pZ)
,”
Commun. Math. Phys.
369
,
1167
1185
(
2019
).
10.
J. C.
Bridgeman
,
D.
Barter
, and
C.
Jones
, “
Fusing binary interface defects in topological phases: The Vec(Z/pZ) case
,”
J. Math. Phys.
60
,
121701
(
2019
).
11.
A.
Kitaev
and
L.
Kong
, “
Models for gapped boundaries and domain walls
,”
Commun. Math. Phys.
313
,
351
373
(
2012
).
12.
S.
Beigi
,
P. W.
Shor
, and
D.
Whalen
, “
The quantum double model with boundary: Condensations and symmetries
,”
Commun. Math. Phys.
306
,
663
694
(
2011
).
13.
A.
Kapustin
and
N.
Saulina
, “
Topological boundary conditions in Abelian Chern–Simons theory
,”
Nucl. Phys. B
845
,
393
435
(
2011
).
14.
M.
Levin
, “
Protected edge modes without symmetry
,”
Phys. Rev. X
3
,
021009
(
2013
).
15.
A.
Davydov
and
D.
Simmons
, “
On Lagrangian algebras in group-theoretical braided fusion categories
,”
J. Algebra
471
,
149
175
(
2017
).
16.
I.
Cong
,
M.
Cheng
, and
Z.
Wang
, “
Topological quantum computation with gapped boundaries
,” arXiv:1609.02037 [quant-ph] (
2016
).
17.
A.
Kitaev
, “
Fault-tolerant quantum computation by anyons
,”
Ann. Phys.
303
,
2
30
(
2003
).
18.
M. S.
Kesselring
,
J. C.
Madgdalena de la Fuente
,
F.
Thomsen
,
J.
Eisert
,
S. D.
Bartlett
, and
B. J.
Brown
, “
Anyon condensation and the color code
,” arXiv:2212.00042 (
2022
).
19.
F.
Thomsen
,
M. S.
Kesselring
,
S. D.
Bartlett
, and
B. J.
Brown
, “
Low-overhead quantum computing with the color code
,” arXiv:2201.07806 (
2022
).
20.
D.
Litinski
, “
A game of surface codes: Large-scale quantum computing with lattice surgery
,”
Quantum
3
,
128
(
2019
).
21.
R.
Dijkgraaf
and
E.
Witten
, “
Topological gauge theories and group cohomology
,”
Commun. Math. Phys.
129
,
393
429
(
1990
).
22.
K. S.
Brown
,
Cohomology of Groups
(
Springer Science & Business Media
,
2012
), Vol.
87
.
23.
X.
Chen
,
Z.-C.
Gu
,
Z.-X.
Liu
, and
X.-G.
Wen
, “
Symmetry protected topological orders and the group cohomology of their symmetry group
,”
Phys. Rev. B
87
,
155114
(
2013
).
24.
J. W.
Barrett
and
B. W.
Westbury
, “
Invariants of piecewise-linear 3-manifolds
,”
Trans. Am. Math. Soc.
348
,
3997
4022
(
1996
).
25.
A.
Bauer
,
J.
Eisert
, and
C.
Wille
, “
Towards topological fixed-point models beyond gappable boundaries
,”
Phys. Rev. B
106
,
125143
(
2022
).
26.
P.
Etingof
,
S.
Gelaki
,
D.
Nikshych
, and
V.
Ostrik
,
Tensor Categories
(
American Mathematical Society
,
2016
), Vol.
205
.
27.
T.
Lawson
, “
Computing an explicit homotopy inverse for b(*, h, *) → b(*, g, g/h)
,” https://mathoverflow.net/q/288304 (
2017
); accessed 17 June 2022.
28.
M. d. W.
Propitius
, “
Topological interactions in broken gauge theories
,” Ph.D. thesis,
Instituut voor Theoretische Fysica
,
Amsterdam
,
1995
.
29.
I.
Cong
,
M.
Cheng
, and
Z.
Wang
, “
Defects between gapped boundaries in two-dimensional topological phases of matter
,”
Phys. Rev. B
96
,
195129
(
2017
).
30.
A.
Bullivant
and
C.
Delcamp
, “
Tube algebras, excitations statistics and compactification in gauge models of topological phases
,”
J. High Energy Phys.
2019
,
216
.
31.
A. L.
Bullivant
, “
Exactly solvable models for topological phases of matter and emergent excitations
,” Ph.D. thesis,
University of Leeds
,
2018
.
32.
D. E.
Evans
and
Y.
Kawahigashi
, “
On Ocneanu’s theory of asymptotic inclusions for subfactors, topological quantum field theories and quantum doubles
,”
Int. J. Math.
06
,
205
228
(
1995
).
33.

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.

34.
R.
Dijkgraaf
,
V.
Pasquier
, and
P.
Roche
, “
Quasi hope algebras, group cohomology and orbifold models
,”
Nucl. Phys. B, Proc. Suppl.
18
,
60
72
(
1991
).
35.
J. C.
Magdalena de la Fuente
,
N.
Tarantino
, and
J.
Eisert
, “
Non-Pauli topological stabilizer codes from twisted quantum doubles
,”
Quantum
5
,
398
(
2021
).
36.
T. D.
Ellison
,
Y.-A.
Chen
,
A.
Dua
,
W.
Shirley
,
N.
Tantivasadakarn
, and
D. J.
Williamson
, “
Pauli stabilizer models of twisted quantum doubles
,”
PRX Quantum
3
,
010353
(
2022
).
37.
K.
Laubscher
,
D.
Loss
, and
J. R.
Wootton
, “
Universal quantum computation in the surface code using non-Abelian islands
,”
Phys. Rev. A
100
,
012338
(
2019
).
38.
A.
Coste
,
T.
Gannon
, and
P.
Ruelle
, “
Finite group modular data
,”
Nucl. Phys. B
581
,
679
717
(
2000
).
39.
J. C.
Bridgeman
,
L.
Lootens
, and
F.
Verstraete
, “
Invertible bimodule categories and generalized schur orthogonality
,”
Commun. Math. Phys.
402
,
2691
(
2022
).
40.
D.
Barter
,
J. C.
Bridgeman
, and
R.
Wolf
, “
Computing associators of endomorphism fusion categories
,”
SciPost Phys.
13
,
029
(
2022
); arXiv:2110.03644.
41.

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.

42.
K.
Duivenvoorden
,
M.
Iqbal
,
J.
Haegeman
,
F.
Verstraete
, and
N.
Schuch
, “
Entanglement phases as holographic duals of anyon condensates
,”
Phys. Rev. B
95
,
235119
(
2017
).
43.
M. E.
Beverland
,
O.
Buerschaper
,
R.
Koenig
,
F.
Pastawski
,
J.
Preskill
, and
S.
Sijher
, “
Protected gates for topological quantum field theories
,”
J. Math. Phys.
57
,
022201
(
2016
).
44.
P.
Webster
,
M.
Vasmer
,
T. R.
Scruby
, and
S. D.
Bartlett
, “
Universal fault-tolerant quantum computing with stabilizer codes
,”
Phys. Rev. Res.
4
,
013092
(
2022
).
45.
W.
Feng
, “
Non-Abelian quantum error correction
,” Ph.D. thesis,
The Florida State University
,
2015
.
46.
A.
Schotte
,
G.
Zhu
,
L.
Burgelman
, and
F.
Verstraete
, “
Quantum error correction thresholds for the universal Fibonacci Turaev-Viro code
,”
Phys. Rev. X
12
,
021012
(
2022
).
47.

N, M have to share a divisor in order for there to exist a non-trivial 2-coycle class, see  Appendix B.

48.
D.
Naidu
and
E. C.
Rowell
, “
A finiteness property for braided fusion categories
,”
Algebras Representation Theory
14
,
837
855
(
2011
).
49.
J. C.
Bridgeman
and
D.
Barter
, “
Computing defects associated to bounded domain wall structures: The Z/pZ case
,”
J. Phys. A: Math. Theor.
53
,
235206
(
2020
).
50.
D.
Litinski
, “
Magic state distillation: Not as costly as you think
,”
Quantum
3
,
205
(
2019
).
51.
M. S.
Kesselring
,
F.
Pastawski
,
J.
Eisert
, and
B. J.
Brown
, “
The boundaries and twist defects of the color code and their applications to topological quantum computation
,”
Quantum
2
,
101
(
2018
).
52.
D.
Horsman
,
A. G.
Fowler
,
S.
Devitt
, and
R. V.
Meter
, “
Surface code quantum computing by lattice surgery
,”
New J. Phys.
14
,
123011
(
2012
).
53.
A.
Kubica
,
B.
Yoshida
, and
F.
Pastawski
, “
Unfolding the color code
,”
New J. Phys.
17
,
083026
(
2015
).
54.
A.
Bauer
, “
Disentangling modular Walker-Wang models via fermionic invertible boundaries
,”
Phys. Rev. B
107
,
085134
(
2023
); arXiv:2208.03397.
55.

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.

Published open access through an agreement with FU Berlin Fachbereich Physik