Non-abelian Quantum Statistics on Graphs
- 376 Downloads
Abstract
We show that non-abelian quantum statistics can be studied using certain topological invariants which are the homology groups of configuration spaces. In particular, we formulate a general framework for describing quantum statistics of particles constrained to move in a topological space X. The framework involves a study of isomorphism classes of flat complex vector bundles over the configuration space of X which can be achieved by determining its homology groups. We apply this methodology for configuration spaces of graphs. As a conclusion, we provide families of graphs which are good candidates for studying simple effective models of anyon dynamics as well as models of non-abelian anyons on networks that are used in quantum computing. These conclusions are based on our solution of the so-called universal presentation problem for homology groups of graph configuration spaces for certain families of graphs.
1 Introduction
The first part of this paper (Sects. 1–3) contains a meta analysis of literature concerning connections between topology of configuration spaces and the existence of different types of quantum statistics. Because the relevant literature is rather scarce, it was a nontrivial task to make such a meta analysis and we consider it an essential step in describing our results. This is because we see the need of introducing in a systematic and concise way the framework for studying quantum statistics which is designed specifically for graphs. The most challenging part in formulating such a framework is to avoid the language of differential geometry, as graph configuration spaces are not manifolds, whereas the great majority of results in the field concerns quantum statistics on manifolds. As a result, we obtain a universal framework whose many features can be utilised for a very wide class of topological spaces. The framework relies on the following mains steps: (i) defining flat bundles as quotients of the trivial bundle over the universal cover of the configuration space (Theorem 5), (ii) defining Chern characteristic classes solely by pullbacks of the universal bundle (Sect. 3.1), (iii) pointing out the role of the moduli space of flat U(n)-bundles as an algebraic variety in \(U(n)^{\times r}\), r being the rank of the fundamental group of the respective configuration space (Sect. 3.3).
- 1.
Topological classification of wave functions. Classify isomorphism classes of flat hermitian vector bundles of rank k over \(C_n(X)\). Here we also point out that in fact physically meaningful is the classification of vector bundles with respect to the so-called stable equivalence, as nonisomorphic but stable equivalent vector bundles have identical Chern numbers. An important role is played by the reduced K-theory and (co)homology groups of \(C_n(X)\). Calculation of those groups for various graph configuration spaces is the main problem we solve in Sect. 5.
- 2.
Classification of statistical properties. If X is a manifold, for each flat hermitian vector bundle, classify the flat connections. The parallel transport around loops in \(C_n(X)\) determines the statistical properties. For general paracompact X, this point can be phrased as classification of the U(k) - representations of the corresponding braid group, i.e. the fundamental group of \(C_n(X)\).
General methods that we describe in the first three sections of this paper, are applied to a special class of configuration spaces of particles on graphs (treated as 1-dimensional CW complexes). Graph configuration spaces serve as simple models for studying quantum statistical phenomena in the context of abelian anyons [7, 8] or multi-particle dynamics of fermions and bosons on networks [9, 10, 11]. Quantum graphs already proved to be useful in other branches of physics such as quantum chaos and scattering theory [12, 13, 14]. Of particular relevance to this paper are explicit physical models of non-abelian anyons on networks. One of the most notable directions of studies in this area is constructing models for Majorana fermions which can be braided thanks to coupling together a number of Kitaev chains [15, 16]. Such models lead to new robust proposals of architectures for topological quantum computers that are based on networks. Another general way of constructing models for anyons is via an effective Chern–Simons interaction [17, 18]. Such models can also be adapted to the setting of graphs as self-adjoint extensions of a certain Chern–Simons hamiltonian which is defined locally on cells of the graph configuration space [19]. All such physical models realise some unitary representations of a graph braid group.
a set of universal generators which generate all homology groups of graph configuration spaces
a set of universal relations which generate all relations between universal generators.
While solving the universal presentation problem we used not only the state-of-the-art methods that have been used previously in a different context by us and other authors, but also developed new computational tools. The already existing methods were in particular (i) discrete models of graph configuration spaces by Abrams and Świątkowski [20, 21], (ii) the product-cycle ansatz introduced in our previous paper concerning tree graphs [22], (iii) the vertex blow-up method introduced by Knudsen et. al. [23], (iv) discrete Morse theory for graph configuration spaces introduced by Farley and Sabalka [24]. However, these methods have not been used before to tackle the universal presentation problem. New computational tools we used mainly relied on (i) introducing explicit techniques for calculating homology groups appearing in the homological exact sequence stemming from the vertex blow-up, (ii) demonstrating a new strategy of decomposing a given graph by a sequence of vertex blow-ups and using inductive arguments to compute the homology groups, (iii) further formalising and developing the product-cycle ansatz so that it can be adapted for more general graphs than just tree graphs (iv) new ansatz for non-product universal generators which are homeomorphic to triple tori, (v) implementing discrete Morse theory for graph configuration spaces in a computer code. A non-trivial combination of the above methods that we have applied has proved to be very effective in tackling the universal presentation problem. Nevertheless, while formulating our general framework for studying quantum statistics we already arrive at a number of new very general corollaries. This in particular concerns the structure of abelian statistics on spaces with a finitely-generated fundamental group and pointing out the role of K-theory in studying non-abelian statistics of a high rank.
1.1 Quantum kinematics on smooth manifolds
Example 1
Exchange of two particles on the plane and the resulting loop in \(C_2({\mathbb {R}}^2)\)
Two types of loops in \({\mathrm{RP}}^2\) pictured as a half-sphere with the opposite points on the circumference of the base identified. Black loop and red loop are contractible, while blue loop is non-cntractible. Blue loop becomes homotopy equivalent to the red loop when crossed twice
Bundle \(E_0\) corresponds to the trivial representation of \(\pi _1\), while \(E'\) corresponds to the alternating representation that acts with multiplication by a phase factor \(e^{i\pi }\). Consequently, the holonomy group changes the sign of the wave function from \(E'\) when transported along a non-contractible loop, while the transport of a wave function from the trivial bundle results with the identity transformation. Therefore, bundle \(E_0\) is called bosonic bundle, whereas bundle \(E'\) is called the fermionic bundle.
As we have seen in the above examples, there is a fundamental difference between anyons in \({\mathbb {R}}^2\) and bosons and fermions in \({\mathbb {R}}^3\). Anyons emerge as different flat connections on the trivial line bundle over \(C_2({\mathbb {R}}^2)\), while fermions and bosons emerge as flat connections on non isomorphic line bundles over \(C_2({\mathbb {R}}^3)\). As we explain in section 3, these results generalise to arbitrary numbers of particles.
1.2 Quantum kinematics on graphs
Example 2
Configuration space of two particles on graphY. In \(Y\times Y\) there are 9 two-cells. Six of them are products of distinct (but not disjoint) edges of Y. Their intersect with \(\varDelta _2\) is a single point which we denote by (2, 2). The three remaining two-cells are of the form \(e\times e\). They have the form of squares which intersect \(\varDelta _2\) along the diagonal. Graph Y and space \(C_2(Y)\) are shown on Fig. 3.
Graph Y and its two-particle configuration space. White dots and dashed lined denote the diagonal \(\varDelta _2\)
Theorem 1
Fix n—the number of particles. If \(\varGamma \) has the following properties: (i) each path between distinct vertices of degree not equal to 2 passes through at least \(n-1\) edges, (ii) each nontrivial loop passes through at least \(n+1\) edges, then \(C_n(\varGamma )\) deformation retracts to a CW-complex \(D_n(\varGamma )\) which is a subspace of \(C_n(\varGamma )\) and consists of the n-fold products of disjoint cells of \(\varGamma \).
More formally, we classify all homomorphisms \(\rho \in {{\mathrm{Hom}}}(\pi _1(C_n(\varGamma )),U(k))\) and consider the vector bundles that are induced by the action of \(\rho \) on the trivial principal U(k)-bundle over the universal cover of \(C_n(\varGamma )\). For more details, see Sect. 3.
Deformation of a loop from \(C_2(Y)\) to \(D_2(Y)\)
Example 3
Quantum kinematics of rank 1 of two particles on graphY. The two-particle discrete configuration space of graph Y consists of 6 edges that form a circle (Fig. 4). Therefore, any non-contractible loop in \(C_2(Y)\) is homotopic with \(D_2(Y)\).
2 Methodology
All topological spaces that are considered in this paper have the homotopy type of finite CW complexes. This is due to the following two theorems.
Theorem 2
[20, 21] The configuration space of any graph \(\varGamma \) can be deformation retracted to a finite CW complex which is a cube cumplex.
Theorem 3
[32, 33] The configuration space of n particles in \({\mathbb {R}}^k\) has the homotopy type of a finite CW-complex.
Using the structure of a CW-complex makes some computational problems more tractable. This is especially useful, while computing the homology groups of graph configuration spaces, because the corresponding CW-complexes have a simple, explicit form.
- 1.
The n-strand braid group of \({\mathbb {R}}^3\) is the permutation group, \(Br_n({\mathbb {R}}^3)=S_n\).
- 2.The n-strand braid group of \({\mathbb {R}}^2\) is often simply called braid group and denoted by \(Br_n\). It has \(n-1\) generators denoted by \(\sigma _1,\dots ,\sigma _{n-1}\). One can illustrate the generators by arranging particles on a line. In such a setting, \(\sigma _i\) corresponds to exchanging particles i and \(i+1\) in a clockwise manner. By composing such exchanges, one arrives at the following presentation of \(Br_n({\mathbb {R}}^2)\)$$\begin{aligned}&Br_n({\mathbb {R}}^2)=\langle \sigma _1,\dots ,\sigma _{n-1}:\ \sigma _i\sigma _{i+1}\sigma _i=\sigma _{i+1}\sigma _i\sigma _{i+1}\ {{\mathrm{for\ }}}i=1,\dots ,n-2,\\&\quad \sigma _i\sigma _j=\sigma _j\sigma _i{\mathrm{\ for\ }}|i-j|\ge 2\rangle . \end{aligned}$$
- 3.
The n-strand braid group of a sphere \(S^2\) has the same set of generators and relations as \(Br_n({\mathbb {R}}^2)\), but with one additional relation: \(\sigma _1\sigma _2\dots \sigma _{n-1}\sigma _{n-1}\dots \sigma _2\sigma _1=e\).
- 4.
The n-strand braid group of a torus \(T^2\). Group \(Br_n(T^2)\) is generated by (i) generators \(\sigma _1,\dots ,\sigma _{n-1}\) where the relations are the same as in the case of \({\mathbb {R}}^2\) and (ii) generators \(\tau _i,\ \rho _i\), \(i=1,\dots ,n\) that transport particle i around one of the two fundamental loops on \(T^2\) respectively. As the full set of relations defining \(Br_n(T^2)\) is quite long, we refer the reader to [35].
- 5.
Fundamental groups of n-particle configuration spaces of graphs, also called graph braid groups [24, 36]. The study of integral homology of graph braid groups is a central point of this paper.
Example 4
Group \(Br_2(\varGamma _\varTheta )\) is a free group with three generators: \(\alpha _U,\alpha _D,\gamma _L\)
A physical model for a U(2) representation of \(Br_2(\varGamma _\varTheta )\) can be constructed using general theory of exchanging Majorana fermions on networks of quantum wires presented in [16]. Here we only briefly sketch the main ideas of this construction. The role of particles is played by two Majorana fermions placed on the spots of black dots from Fig. 5. The two fermions are at the endpoints of the so-called topological region in a network of superconducting quantum wires. Majorana fermions are braided by adiabatically changing physical parameters of the quantum wire.
Theorem 4
If X has the homotopy type of a finite CW complex, then ranks of \(H^k(X,{\mathbb {Z}})\) and \(H_k(X,{\mathbb {Z}})\) are equal and the torsion of \(H^k(X,{\mathbb {Z}})\) is equal to the torsion of \(H_{k-1}(X,{\mathbb {Z}})\).
3 Vector Bundles and Their Classification
The main motivation for studying (co)homology groups of configuration spaces comes from the fact that they give information about the isomorphism classes of vector bundles over configuration spaces. In the following section, we review the main strategies of classifying vector bundles and make the role of homology groups more precise. Throughout, we do not assume that the configuration space is a differentiable manifold, as the configuration spaces of graphs are not differentiable manifolds. We only assume that \(C_n(X)\) has the homotopy type of a finite CW-complex. This means that \(C_n(X)\) can be deformation retracted to a finite CW-complex. As we explain in Sect. 4, configuration spaces of graphs are such spaces. The lack of differentiable structure means that the flat vector bundles have to be defined without referring the notion of a connection and all the methods that are used have to be purely algebraic. We provide such an algebraic definition of flat bundles in Sect. 3.3.
In this paper, we consider only complex vector bundles \(\pi :\ E\rightarrow B\), where E is a total space and B is the base. Two vector bundles are isomorphic iff there exists a homeomorphism between their total spaces which preserves the fibres. If two vector bundles belong to different isomorphism classes, there is no continuous function which transforms the total spaces to each other, while preserving the fibres. Hence, the wave functions stemming from sections of such bundles must describe particles with different topological properties. The classification of vector bundles is the task of classifying isomorphism classes of vector bundles. The set of isomorphism classes of vector bundles of rank k will be denoted by \(\mathcal {E}_k^{\mathbb {K}}(B)\).
Before we proceed to the specific methods of classification of vector bundles, we introduce an equivalent way of describing vector bundles which involves principal bundles (principal G-bundles). A principal G-bundle \(\xi :\ P\rightarrow B\) is a generalisation of the concept of vector bundle, where the total space is equipped with a free action of group G1 and the base space has the structure of the orbit space \(B\cong P/G\). Fibre \(\pi ^{-1}(p)\) is isomorphic to G is the sense that map \(\pi :\ P\rightarrow B\) is G-invariant, i.e. \(\pi (ge)=\pi (e)\). Moreover, all relevant morphisms are required to be G-equivariant. The set of isomorphism classes of principal G-bundles over base space B will be denoted by \(\mathcal {P}_G(B)\).
3.1 Universal bundles and Chern classes
3.2 Reduced K-theory
We start with recalling the definition of stable equivalence of vector bundles.
Definition 1
3.3 Flat bundles and quantum statistics
An important conclusion regarding flat bundles on spaces that do not have a differential structure comes from the second part of correspondence (5). This is the reconstruction of a flat principal bundle from a given homomorphism \({{\mathrm{Hom}}}(\pi _1(B),G)\). It turns out that any flat bundle over B can be realised as a particular quotient bundle of the trivial bundle over the universal cover of B. In order to formulate the correspondence, we first introduce the notion of a covering space and a universal cover.2 The following theorem is also a definition of a flat principal bundle for spaces that are not differential manifolds.
Theorem 5
Fact 3.1
Two points in \(\mathcal {M}(B,G)\) that correspond to two non-isomorphic flat bundles, belong to different path-connected components of \(\mathcal {M}(B,G)\).
Equivalently, if two flat structures, i.e. points in \(\mathcal {M}(B,G)\), belong to the same path-connected component of \(\mathcal {M}(B,G)\), then the corresponding vector bundles are isomorphic. A path connecting the two points in \(\mathcal {M}(B,G)\) gives a homotopy between the corresponding flat structures.
Example 5
The moduli space of flat U(1) bundles a) for n particles on a plane, b) n particles in \({\mathbb {R}}^3\). Homomorphisms from \({\mathbb {Z}}\) to U(1) are parametrised by points from \(S^1\) via the map \(\phi \mapsto e^{\iota \phi }\). The corresponding homomorphism reads \(n\mapsto e^{\iota n\phi }\). There is only one path-connected component in \({{\mathrm{Hom}}}({\mathbb {Z}},U(1))\) which reflects the fact that there is only one flat U(1) bundle over \(C_n({\mathbb {R}}^2)\) (the trivial one) and points form the circle parametrise different flat connections. For particles in \({\mathbb {R}}^3\), there are two homomorphisms of \({\mathbb {Z}}_2=\{1,-1\}\)—the trivial one and \(1\mapsto e^{2\pi \iota }\), \(-1\mapsto e^{\iota \pi }\). They correspond to two isolated points on the torus \(T^2=U(1)\times U(1)\). The trivial homomorphism corresponds to the bosonic bundle, while the other homomorphism corresponds to the fermionic bundle. The fundamental difference between these two types of quantum statistics is that anyons arise as different flat connections on the trivial bundle, whereas bosons and fermions arise as canonical flat connections on two non-isomorphic flat bundles
Theorem 6
Remark 3.1
Theorem 6 in particular means that if B is a finite CW-complex, then by the universal coefficient theorem for cohomology (see e.g. [43]), the image of the characteristic map \(f_\xi ^*:\ H^*(BG,{\mathbb {Z}})\rightarrow H^*(B,{\mathbb {Z}})\) consists only of torsion elements of \(H^*(B,{\mathbb {Z}})\).
Specifying the above results for U(n)-bundles, we get that the lack of nontrivial torsion in \(H^{2i}(B,{\mathbb {Z}})\) has the following implications for the stable equivalence classes of flat vector bundles.
Proposition 7
Let B be a finite CW complex. If the integral homology groups of B are torsion-free, then every flat complex vector bundle over B is stably equivalent to a trivial bundle.
Proof
If the integral cohomology of B is torsion-free, then by the Chern character we get that the reduced Grothendieck group is isomorphic to the direct sum of even cohomology of B. Thus, if all Chern classes of a given bundle vanish, this means that this bundle represents the trivial element of the reduced Grothendieck group, i.e. is stably equivalent to a trivial bundle. \(\quad \square \)
- 1.
Configuration space of n particles on a plane. Space \(C_n({\mathbb {R}}^2)\) is aspherical, i.e. is an Eilenberg–Maclane space of type \(K(\pi _1,1)\), where the fundamental group is the braid group on n strands \(Br_n\). Cohomology ring \(H^*(C_n({\mathbb {R}}^2),{\mathbb {Z}})=H^*(Br_n,{\mathbb {Z}})\) is known [45, 46]. Its key properties are (i) finiteness—\(H^{i}(Br_n,{\mathbb {Z}})\) are cyclic groups, except \(H^{0}(Br_n,{\mathbb {Z}})=H^{1}(Br_n,{\mathbb {Z}})={\mathbb {Z}}\), (ii) repetition—\(H^{i}(Br_{2n+1},{\mathbb {Z}})=H^{i}(Br_{2n},{\mathbb {Z}})\), (iii) stability—\(H^{i}(Br_{n},{\mathbb {Z}})=H^{i}(Br_{2i-2})\) for \(n\ge 2i-2\). Description of nontrivial flat U(n) bundles over \(C_n({\mathbb {R}}^2)\) for \(n> 2\) is an open problem.
- 2.
Configuration space of n particles in \({\mathbb {R}}^3\). Much less is known about \(H^*(C_n({\mathbb {R}}^3))\). Some computational techniques are presented in [47, 48], but little explicit results are given. Ring \(H^*(C_3({\mathbb {R}}^3)\) is equal to \({\mathbb {Z}},0,{\mathbb {Z}}_2,0,{\mathbb {Z}}_3\) [49] and \(H^q(C_3({\mathbb {R}}^3))=0\) for \(q>4\). However, it has been shown that there are no nontrivial flat SU(n) bundles over \(C_3({\mathbb {R}}^3)\).
- 3.
Configuration space of n particles on a graph (a 1-dimensional CW-complex \(\varGamma \)). Spaces \(C_n(\varGamma )\) are Eilenberg–Maclane spaces of type \(K(\pi _1,1)\). The calculation of their homology groups is a subject of this paper. Group \(H_1(C_n(\varGamma ),{\mathbb {Z}})\) is known [8, 37] for an arbitrary graph. We review the structure of \(H_1(C_n(\varGamma ))\) in Sect. 4.1. By the universal coefficient theorem, the torsion of \(H^2(C_n(\varGamma ))\) is equal to the torsion of \(H_1(C_n(\varGamma ))\) which is known to be equal to a number of copies of \({\mathbb {Z}}_2\), depending on the structure of \(\varGamma \). We interpret this result as the existence of different bosonic or fermionic statistics in different parts of \(\varGamma \). The existence of torsion in higher (co)homology groups of \(C_n(\varGamma )\) which is different than \({\mathbb {Z}}_2\), is an open problem. In this paper, we compute homology groups for certain canonical families of graphs. However, the computed homology groups are either torsion-free, or have \({\mathbb {Z}}_2\)-torsion.
4 Configuration Spaces of Graphs
- 1.
Abram’s discrete configuration space [21]. The Abram’s deformation retract of \(C_n(\varGamma )\) is denoted by \(D_n(\varGamma )\). We use Abram’s discrete model mainly in the first part of this paper, where we apply discrete Morse theory to the computation of homology groups of some small canonical graphs (Sect. 5.2).
- 2.
The discrete model by Świątkowski [20] that we denote by \(S_n(\varGamma )\). We use this model in Sects. 5.3–5.6 to compute homology groups of configuration spaces of wheel graphs and some families of complete bipartite graphs.
each path between distinct vertices of degree not equal to 2 passes through at least \(n-1\) edges,
each nontrivial loop passes through at least \(n+1\) edges.
Definition 2
Świątkowski complex of the Y-graph and of the lasso graph, where vertices of degree 1 have been reduced. a Świątkowski complex of \(C_2(Y)\). Only vertices of \(S_2(\varGamma )\) are captioned. The Y-cycle reads \(e_1(h_2-h_3)+e_2(h_3-h_1)+e_3(h_1-h_2)\). b Świątkowski complex of \(C_2(\varGamma )\) for the lasso graph. Vertices and some chosen edges of \(S_2(\varGamma )\) are captioned. The O-cycles are \(e_1(h_2-h_3)\) and \(e_2(h_2-h_3)\). The Y-cycle is their sum, hence can be written as \((e_1-e_2)(h_2-h_3)\)
Fact 4.1
Vertex blow up at vertex v in \(\varGamma \)
4.1 O-cycles and Y-cycles
There are some particular types of cycles that play an important role in this work. These are O-cycles and Y-cycles. We specify them for the Abram’s model. The construction for \(S_n(\varGamma )\) is fully analogous.
Definition 3
Definition 4
A Y-graph, its configuration space (b) and its discrete configuration space \(D_2(\varGamma )\) (a)
The fundamental relation between the two-particle cycle on a Y-graph and the AB-cycle and a two-particle cycle \(c_2\) in the lasso graph
Cycles \(c_{Y_1}\) and \(c_{Y_2}\) are homologically equivalent
5 Calculation of Homology Groups of Graph Configuration Spaces
This section contains the techniques that we use for computing homology groups of graph configuration spaces. We tackle this problem from the ‘numerical’ and the ‘analytical’ perspective. The numerical approach means using a computer code for creating the boundary matrices and then employing the standard numerical libraries for computing the kernel and the elementary divisors of given matrices. The procedures for calculating the boundary matrices of \(D_n(\varGamma )\), \(S_n(\varGamma )\) and the Morse complex (see Sect. 5.2) were written by the authors of this paper, based on papers [24, 37]. The analytical approach means computing the homology groups for certain families of graphs by suitably decomposing a given graph into simpler components and using various homological exact sequences. Recently in the mathematical community, there has been a growing interest in computing the homology groups of graph configuration spaces. A significant part of the recent work has been devoted to explaining certain regularity properties of the homology groups of \(C_n(\varGamma )\) [50, 51, 52, 53, 54, 55].
5.1 Product cycles
5.2 Discrete Morse theory for Abrams model
In this subsection, we apply a version of Forman’s discrete Morse theory [58] for Abram’s discrete model that was formulated in [24] (see also [59]). The results are listed in Tables 1 and 2.
Betti numbers for chosen graphs computed using the discrete Morse theory [24]
\(\varGamma \) | n | \(\beta _2(C_n(\varGamma ))\) | \(\beta _3(C_n(\varGamma ))\) | \(\beta _4(C_n(\varGamma ))\) |
---|---|---|---|---|
\(K_4\) | 3 | 3 | 0 | – |
4 | 9 | 0 | 0 | |
5 | 15 | 0 | 0 | |
6 | 21 | 4 | 0 | |
7 | 27 | 16 | 0 | |
8 | 33 | 40 | 1 | |
9 | 39 | 80 | 6 | |
\(K_{3,3}\) | 2 | 0 | – | – |
3 | 8 | 0 | – | |
4 | 19 | 1 | 0 | |
5 | 28 | 10 | 0 | |
6 | 37 | 39 | 0 | |
7 | 46 | 88 | 0 | |
8 | 55 | 157 | 15 | |
\(K_5\) | 2 | 0 | – | – |
3 | 30 | 0 | – | |
4 | 76 | 1 | 0 | |
5 | 116 | 77 | 0 | |
6 | 156 | 381 | 0 | |
7 | 196 | 961 | 0 |
The first regular homology groups of order 2 and 3 for the Petersen family
\(K_6\) | \(P_7\) | \(K_{3,3,1}\) | \(K_{4,4}\) | \(P_8\) | \(P_9\) | \(P_{10}\) | |
---|---|---|---|---|---|---|---|
\(\beta _2(C_4(\varGamma ))\) | 264 | 177 | 172 | 144 | 114 | 70 | 40 |
\(T_2(C_4(\varGamma ))\) | \({\mathbb {Z}}_2\) | \({\mathbb {Z}}_2\) | \({\mathbb {Z}}_2\) | \(\left( {\mathbb {Z}}_2\right) ^2\) | \({\mathbb {Z}}_2\) | \({\mathbb {Z}}_2\) | \({\mathbb {Z}}_2\) |
\(\beta _3(C_6(\varGamma ))\) | 4137 | 2058 | 1919 | 1460 | 986 | 452 | 191 |
\(T_3(C_6(\varGamma ))\) | 0 | 0 | 0 | \(\left( {\mathbb {Z}}_2\right) ^{73}\) | 0 | 0 | 0 |
Graphs that form the Petersen family
Table 2 presents the results for the second and third homology groups for graphs from the Petersen family (Fig. 12). These graphs serve as examples, where torsion in higher homology groups appears. Interestingly, the torsion subgroups are always equal to a number of copies of \({\mathbb {Z}}_2\). This phenomenon can be explained by embedding a nonplanar graph in \(\varGamma \) and considering suitable product cycles. The question about the existence of torsion different than \({\mathbb {Z}}_2\) in higher homologies remains open.
5.3 Wheel graphs
In this section, we deal with the class of wheel graphs. A wheel graph of order m is a simple graph that consists of a cycle on \(m-1\) vertices, whose every vertex is connected by an edge (called a spoke) to one central vertex (called the hub). We provide a complete description of the homology groups of configuration spaces for wheel graphs. In particular, we show that all homology groups are free. Therefore, in addition to tree graphs, wheel graphs provide another family of configuration spaces with a simplified structure of the set of flat complex vector bundles. The general methodology of computing homology groups for configuration spaces of wheel graphs is to consider only the product cycles and describe the relations between them. We justify this approach in Sect. 5.4.
The simplest example of a wheel graph is graph \(K_4\) which is the wheel graph of order 4. Let us next calculate all homology groups of graph \(K_4\) and then present the general method for any wheel graph.
5.3.1 Graph \(K_4\)
Graph \(K_4\) and the relevant Y-subgraphs and cycles. We omit the subdivision of edges in the picture
Graph \(K_4\) subdivided for \(n=4\). Differences \(c_{AB}^u-c_{AB}^v\) and \(c_{AB}^u-c_{AB}^w\) are homologically equivalent to combinations of \(Y\times Y\)-cycles. \(c_{Y_1}\otimes c_{Y_h}-c_{Y_1}\otimes c_{Y_2}\) and \(c_{Y_1}\otimes c_{Y_h}-c_{Y_1}\otimes c_{Y_3}\) respectively
Graph \(K_4\) after removing two Y-subgraphs
5.3.2 General wheel graphs
Betti numbers of configuration spaces for chosen wheel graphs computed using the discrete Morse theory
\(\varGamma \) | n | \(\beta _2(D_n(\varGamma ))\) | \(\beta _3(D_n(\varGamma ))\) | \(\beta _4(D_n(\varGamma ))\) |
---|---|---|---|---|
\(W_5\) | 3 | 8 | 0 | – |
4 | 22 | 0 | 0 | |
5 | 34 | 4 | 0 | |
6 | 46 | 30 | 0 | |
7 | 58 | 90 | 0 | |
8 | 70 | 196 | 13 | |
\(W_6\) | 3 | 15 | 0 | – |
4 | 40 | 0 | 0 | |
5 | 60 | 15 | 0 | |
6 | 80 | 90 | 0 | |
7 | 100 | 250 | 5 | |
\(W_7\) | 3 | 24 | 0 | – |
4 | 63 | 0 | 0 | |
5 | 93 | 36 | 0 | |
6 | 123 | 197 | 0 | |
7 | 153 | 527 | 24 |
Relations between different pairs of \(Y_h\times Y\)-cycles in a wheel graph. a Cycles, where \(Y_h\) and Y share an edge of the graph are independent. b Cycle, where \(Y_h\) and Y do not share any edges is in the same homology class as cycle \(Y'\times Y\)
The fan graph that is created after removing two neighbouring Y-subgraphs from the perimeter of \(W_5\). It has \(\mu =3\) leaves. There are two types of Y-cycles at the hub: a cycles, where the Y-graph is spanned in three different leaves—the number of such cycles for \(k+2\) particles is \(\beta _1^{(k+2)}(S_3)\), b cycles, where the Y-graph is spanned in two different leaves—the number of such cycles for \(k+2\) particles is \({{k+3}\atopwithdelims (){k+1}}-1\), see [8]
The possibilities of choosing a number of Y-subgraphs from the perimeter of a wheel graph
\(\varGamma \) | Groups of Y-subgraphs—\(\mathbf{n}\) | Number of possible choices—\(N_\mathbf{n}\) | Number of leaves—\(\mu _\mathbf{n}\) |
---|---|---|---|
\(W_5\) | (1) | 4 | 1 |
(1,1) | 2 | 4 | |
(2) | 4 | 3 | |
(3) | 4 | 4 | |
(4) | 1 | 4 | |
\(W_6\) | (1) | 5 | 2 |
(1,1) | 5 | 4 | |
(2) | 5 | 3 | |
(2,1) | 5 | 5 | |
(3) | 5 | 4 | |
(4) | 5 | 5 | |
(5) | 1 | 5 | |
\(W_7\) | (1) | 6 | 2 |
(1,1) | 9 | 4 | |
(2) | 6 | 3 | |
(1,1,1) | 2 | 6 | |
(2,1) | 12 | 5 | |
(3) | 6 | 4 | |
(2,2) | 3 | 6 | |
(3,1) | 6 | 6 | |
(4) | 6 | 5 | |
(5) | 6 | 6 | |
(6) | 1 | 6 |
5.4 Wheel graphs via Świątkowski discrete model
In this section we show that the homology of configuration spaces of wheel graphs is generated by product cycles. The strategy is to consider two consecutive vertex cuts that bring any wheel graph to the form of a linear tree.
Lemma 1
Proof
Vertex blowup at the hub of wheel \(W_{m+1}\) resulting with net graph \(N_{m}\)
Blowup of a vertex in net graph \(N_{m}\) resulting with linear tree graph \(T_{m}\)
5.5 Graph \(K_{3,3}\)
Graph \(K_{3,3}\)
Graph \(K_{3,3}\) has the property that all its vertices are of degree three. High homology groups of graphs with such a property have been studied in [23]. In particular, we have the following result.
Theorem 8
As we show in Sect. 5.7, the second homology group of configuration spaces of such graphs is also generated by product cycles. Later in this section, by comparing the ranks of homology groups computed via the discrete Morse theory, we argue that \(H_4(C_n(K_{3,3}))\) is also generated by product cycles. Interestingly, in \(H_3(C_n(K_{3,3}))\) there is a new non-product generator. Using this knowledge, we explain the relations between the product and non-product cycles that give the correct rank of \(H_3(C_n(K_{3,3}))\).
Graph \(K_{3,3}\) after removing two Y-subgraphs
Graph \(K_{3,3}\) sufficiently subdivided for \(n=4\). The deleted edges are marked with dashed lines
Graph \(K_{3,3}\) after removing three Y-subgraphs
5.6 Triple tori in \(C_n(K_{2,p})\)
a Graph \(K_{2,p}\). b Graph \(\varTheta _p\)
Lemma 2
The first homology group of \(C_n(K_{2,p})\) is equal to \({\mathbb {Z}}^{p(p-1)}\) for \(n\ge 2\) and \(p-1\) for \(n=1\).
By counting the number of 0-, 1- and 2-cells in \(S_n(K_{2,p})\), we compute the Euler characteristic (see also [61]).
Lemma 3
Example 6
Example 7
A triple torus
Proposition 9
Proposition 10
Hence, all \(\varTheta \)-cycles generate a subgroup of \(H_2(S_n(\varTheta _p))\) which is isomorphic to \({\mathbb {Z}}^{{p-1}\atopwithdelims (){3}}\). The last type of relations we have to account for5 are the new relations between products of Y-cycles.
Proposition 11
5.7 When is \(H_2(C_n(\varGamma ))\) generated only by product cycles?
In this section we prove the following theorem.
Theorem 12
Let \(\varGamma \) be a simple graph, for which \(|\{v\in V(\varGamma ):\ d(v)>3\}|=1\). Then group \(H_2(C_n(\varGamma ))\) is generated by product cycles.
In the proof we use the Świątkowski discrete model. The strategy of the proof is to first consider the blowup of the vertex of degree greater than 3 and prove theorem 12 for graphs, whose all vertices have degree at most 3. For such a graph, we choose a spanning tree \(T\subset \varGamma \). Next, we subdivide once each edge from \(E(\varGamma )-E(T)\). We prove the theorem inductively by showing in lemma 4 that the blowup at an extra vertex of degree 2 does not create any non-product generators. The base case of induction is obtained by doing the blowup at every vertex of degree 2 in \(\varGamma -T\). This way, we obtain graph which is isomorphic to tree T and we use the fact that for tree graphs the homology groups of \(S_n(T)\) are generated by products of Y-cycles.
Lemma 4
Let \(\varGamma \) be a simple graph, whose all vertices have degree at most 3. Let T be a spanning tree of \(\varGamma \). Let \(v\in V(\varGamma )\) be a vertex of degree 2 and \(\varGamma _v\) the graph obtained from \(\varGamma \) by the vertex blowup at v. If \(H_2(S_n(\varGamma _v))\) is generated by product cycles, then \(H_2(S_n(\varGamma ))\) is also generated by product cycles.
Proof
This way, we obtained that \(H_2\left( {\tilde{S}}^v_n(\varGamma )\right) \cong \ker \left( \delta _{n,1}\right) \oplus \mathrm{coker}\left( \delta _{n,2}\right) \) and that elements of \(\ker \left( \delta _{n,1}\right) \) are represented by product \(c_O\otimes c_Y\) cycles. By the inductive hypopaper, elements of \(\mathrm{coker}\left( \delta _{n,2}\right) \) are the product cycles that generate \(H_2\left( S_{n-1}(\varGamma _v)\right) \) subject to relations \(ce\sim ce'\). \(\square \)
The last step needed for the proof of theorem 12 is showing that the blowup of \(\varGamma \) at the unique vertex of degree greater than 3 does not create any non-product cycles. Here we only sketch the proof of this fact which is analogous to the proof of lemma 4. Namely, using the knowledge of relations between the generators of \(H_{1}\left( S_{n-1}(\varGamma _v)\right) \), one can show that the elements of \(\ker \left( \delta _{n,1}\right) \) are of two types: i) the ones that are of the form \(\partial (c\otimes b_{p(c)})\), where \([c]\in H_{1}\left( S_{n-1}(\varGamma _v)\right) \) and \(b_{p(c)}\) is the 1-cycle corresponding to path \(p(c)\subset \varGamma _v\) which is disjoint with \({\mathrm{Supp}}(c)\) and whose boundary are edges incident to v, ii) pairs of cycles of the form \((c(e_j-e_0),c(e_0-e_j))\), where \(e_0, e_i, e_j\) are edges incident to v and \([c]\in H_{1}\left( S_{n-2}(\varGamma _v)\right) \). Such pairs are mapped by \(\delta _{n,1}\) to \(c\otimes \left( (e_j-e_0)(e_0-e_i)+(e_0-e_i)(e_0-e_j)\right) \) which is equal to \(\partial (c\otimes c_{0ij})\), where \(c_{0ij}\) is the Y-cycle corresponding to the Y-graph in \(\varGamma \) centred at v and spanned by edges \(e_0,e_i,e_j\). Next, in order to show splitting of the homological short exact sequences, we consider a homomorphism \(f:\ \ker \left( \delta _{n,1}\right) \rightarrow H_2\left( {\tilde{S}}^v_n(\varGamma )\right) \), for which \(\varPsi _{n,2}\circ f=id_{\ker (\delta _{n,1})}\). Such a homomorphism maps [c] to \([c\otimes c_{O_{p(c)}}]\), where \(O_{p(c)}\) is the cycle which contains path p(c) and vertex v. Pairs \(([c(e_j-e_0)],[c(e_0-e_j)])\) are mapped by f to cycles \(c\otimes c_{0ij}\). We obtain that \(H_2\left( {{\tilde{S}}}^v_n(\varGamma )\right) \cong \ker \left( \delta _{n,1}\right) \oplus \mathrm{coker}\left( \delta _{n,2}\right) \), where the generators of \(\ker \left( \delta _{n,1}\right) \) are in a one-to-one correspondence with the product homology classes of \(H_2\left( {{\tilde{S}}}^v_n(\varGamma )\right) \) described above. Elements of \(\mathrm{coker}\left( \delta _{n,2}\right) \) are also represented by product cycles. These cycles are the generators of \(H_2\left( S_{n}(\varGamma _v)\right) \) subject to relations \(ce_0\sim ce_i\), \(i=1,\dots ,d(v)\), where \(e_0,e_1,\dots ,e_{d(v)}\) are edges incident to v.
The task of characterising all graphs, for which \(H_2(S_n(\varGamma ))\) is generated by product cycles requires taking into account the existence of non-product generators from Sect. 5.6. As we show in Sect. 5.6 the existence of pairs of vertices of degree greater than 3 in the graph implies that there may appear some multiple tori in the generating set of \(H_2(C_n(\varGamma ))\) stemming from subgraphs isomorphic to graph \(K_{2,4}\). Furthermore, the class of graphs, for which higher homologies of \(C_n(\varGamma )\) are generated by product cycles is even smaller. Recall graph \(K_{3,3}\) whose all vertices have degree 3, but \(H_3(C_n(K_{3,3}))\) has one generator which is not a product of 1-cycles (see Sect. 5.5).
6 Summary
Configuration spaces of tree graphs, wheel graphs and complete bipartite graphs \(K_{2,p}\) have no torsion in their homology. This means that the set of flat bundles over configuration spaces of such graphs has a simplified structure, namely every flat vector bundle is stably equivalent to a trivial vector bundle. Hence, these families of graphs are good first candidates for a class of simplified models for studying the properties of non-abelian statistics.
Computation of the homology groups of configuration spaces of some small canonical graphs via the discrete Morse theory shows that in some cases there is a \({\mathbb {Z}}_2\)-torsion in the homology. However, we were not able to provide an example of a graph which has a torsion component different than \({\mathbb {Z}}_2\) in the homology of its configuration space.
It is a difficult task to accomplish a full description of the homology groups of graph configuration spaces using methods presented in this work. One fundamental obstacle is that such a task requires the knowledge of possible embeddings of d-dimensional surfaces in \(C_n(\varGamma )\) which generate the homology. However, cycles generating the homology in dimension 2 of graph configuration spaces have the homotopy type of tori or multiple tori. This fact allowed us to find all generators of the second homology group of configuration spaces of a large family of graphs in Sect. 5.7.
Trivial bundles. They are relevant in the context of quantum computing where one is interested mainly in universality of unitary representations of braid groups and the dimension of the representations grow exponentially with the number of particles. In that context, the fact that one can have different isomorphism classes of vector bundles does not seem to play a significant role. In fact, it is even better not to have many isomorphism classes. If we know that there is just one isomorphism class (all bundles are isomorphic to the trivial bundle) then all representations of the braid group are related to each other via the isomorphism of the corresponding bundles and the problem of classifying them should become more tractable. As we show in this paper, this happens when one considers high-dimensional representations (stable range) of graph braid groups where there is no torsion in the homology of \(C_n(\varGamma )\).
Non-trivial bundles. They become relevant in situations where the rank of the bundle is not too high, i.e. if the considered bundles are not in the stable range. The corresponding representations of braid groups appear in the effective models of non-abelian Chern-Simons particles which are point-like sources mutually interacting via a topological non-Abelian Aharonov–Bohm effect. A model of such particles constrained to move on graphs would be constructed by defining a separate Chern–Simons hamiltonian for each cell of the closure of \(C_n(\varGamma )\) viewed as a subset of \(\varGamma ^{\times n}\). The non-abelian braiding would show up as proper gluing conditions for the wave-functions on the boundaries of cells from \(C_n(\varGamma )\) while studying self-adjoint extensions of such a hamiltonian. The moduli space of flat U(n) bundles over \(C_n(\varGamma )\) is the space of possible parameters that appear as the gluing conditions (see e.g. [19]). This area is still quite unexplored and some more progress has to be made to see how this theory works explicitly for concrete graphs.
A graph that consists of three 3-connected components (depicted as boxes). The schematically pictured abelian representation of \(Br_2(\varGamma )\) is such that exchange of particles in each of the components results with a fermionic (F) or bosonic (B) phase factor. The situation is more complicated for non-abelian representations, but the general characteristic survives—one can choose different R-matrices for each of the components
One expects that the fusion rules will nevertheless significantly restrict the number of admissible representations of graph braid groups.
Footnotes
- 1.
The action of G on P can be left or right. In this work we pick up the convention of right action. This means that \(g(h(p))=(gh)(p)\) for \(g,h\in G\), \(p\in P\). Group action is free iff for all \(g\in G\) and \(p\in P\), \(gp\ne p\).
- 2.
Universal covers of graph configuration spaces have a particularly nice structure, as they have the homotopy type of a CAT(0) cube complex [21] which is contractible.
- 3.
The \(\varTheta \) graph consists of two vertices which are connected by three edges. It can be also viewed as complete bipartite graph \(K_{2,3}\).
- 4.
\(W_{m+1}-(v_h(Y_1)\cup \dots \cup v_h(Y_d))\) is a disconnected topological space. We give this space the structure of a graph by adding a vertex to the open end of each open edge.
- 5.
We do not mention here the typical relations between different Y-cycles on Y-subgraphs of the \(S_p\) graphs which are met while computing the first homology group of the configuration spaces of star graphs (see [8]). Such relations are also inherited by the products of Y-cycles.
Notes
Acknowledgements
TM acknowledges the financial support of the National Science Centre of Poland – grants Etiuda no. 2017 / 24 / T / ST1 / 00489 and Preludium no. 2016 / 23 / N / ST1 / 03209. AS was supported by the National Science Centre of Poland grant Sonata Bis no. 2015 / 18 / E / ST1 / 00200.
References
- 1.Leinaas, J.M., Myrheim, J.: On the theory of identical particles. Nuovo Cim. 37B, 1–23 (1977)ADSGoogle Scholar
- 2.Souriau, J.M.: Structure des systmes dynamiques. Dunod, Paris (1970)zbMATHGoogle Scholar
- 3.Wilczek, F.: Fractional Statistics and Anyon Superconductivity. World Scientific, Singapore (1990)zbMATHGoogle Scholar
- 4.Berry, M.V., Robbins, J.M.: Indistinguishability for quantum particles: spin, statistics and the geometric phase. Proc. R. Soc. Lond. A 453, 1771 (1997)ADSMathSciNetzbMATHGoogle Scholar
- 5.Chruściński, D., Jamiołkowski, A.: Geometric Phases in Classical and Quantum Mechanics. Progress in Mathematical Physics. Birkhäuser, Boston (2004)zbMATHGoogle Scholar
- 6.Aharonov, Y., Bohm, D.: Significance of electromagnetic potentials in quantum theory. Phys. Rev. 115, 48549 (1959)MathSciNetzbMATHGoogle Scholar
- 7.Harrison, J.M., Keating, J.P., Robbins, J.M.: Quantum statistics on graphs. Proc. R. Soc. A 467(2125), 212-23 (2011)ADSMathSciNetzbMATHGoogle Scholar
- 8.Harrison, J.M., Keating, J.P., Robbins, J.M., Sawicki, A.: n-Particle quantum statistics on graphs. Commun. Math. Phys. 330(3), 1293–1326 (2014)ADSMathSciNetzbMATHGoogle Scholar
- 9.Bolte, J., Garforth, G.: Exactly solvable interacting two-particle quantum graphs. J. Phys. A 50(10), 105101, 27 (2017)MathSciNetzbMATHGoogle Scholar
- 10.Bolte, J., Kerner, J.: Quantum graphs with singular two-particle interactions. J. Phys. A: Math. Theor. 46, 045206 (2013)ADSMathSciNetzbMATHGoogle Scholar
- 11.Bolte, J., Kerner, J.: Quantum graphs with two-particle contact interactions. J. Phys. A: Math. Theor. 46(4), 045207 (2013)ADSMathSciNetzbMATHGoogle Scholar
- 12.Gnutzmann, S., Smilansky, U.: Quantum graphs: applications to quantum chaos and universal spectral statistics. Adv. Phys. 55, 527–625 (2006)ADSGoogle Scholar
- 13.Band, R., Sawicki, A., Smilansky, U.: Scattering from isospectral quantum graphs. J. Phys. A-Math. Theor. 43, 41 (2010)MathSciNetzbMATHGoogle Scholar
- 14.Hul, O., Ławniczak, M., Bauch, S., Sawicki, A., Kuś, M., Sirko, L.: Are scattering properties of graphs uniquely connected to their shapes? Phys. Rev. Lett. 109, 040402 (2012)ADSGoogle Scholar
- 15.Kitaev, A.: Unpaired Majorana fermions in quantum wires. Phys.-Usp. 44, 131 (2001)ADSGoogle Scholar
- 16.Alicea, J., Oreg, Y., Refael, G., von Oppen, F., Fisher, M.P.A.: Non-abelian statistics and topological quantum information processing in 1D wire networks. Nat. Phys. 7, 412–417 (2011)Google Scholar
- 17.Lee, T., Phillial, OhP: Non-abelian Chern–Simons quantum mechanics. Phys. Rev. Lett. 72, 1141 (1994)ADSMathSciNetzbMATHGoogle Scholar
- 18.Wilczek, F.: Fractional Statistics and Anyon Superconductivity. World Scientific, Singapore (1990)zbMATHGoogle Scholar
- 19.Balachandran, A.P., Ercolessi, E.: Statistics on networks. Int. J. Mod. Phys. A 7, 4633–4654 (1992)ADSMathSciNetGoogle Scholar
- 20.Świątkowski, J.: Estimates for homological dimension of configuration spaces of graphs. Colloq. Math. 89(1), 69–79 (2001). https://doi.org/10.4064/cm89-1-5 MathSciNetCrossRefzbMATHGoogle Scholar
- 21.Abrams, A.: Configuration spaces and braid groups of graphs, Ph.D. paper, UC Berkley (2000)Google Scholar
- 22.Maciazek, T., Sawicki, A.: Homology groups for particles on one-connected graphs. J. Math. Phys., 58(6) (2017)ADSMathSciNetzbMATHGoogle Scholar
- 23.Byung Hee An, Drummond-Cole, G.C., Knudsen, B.: Subdivisional spaces and graph braid groups, preprint (2017). arXiv:1708.02351
- 24.Farley, D., Sabalka, L.: Discrete Morse theory and graph braid groups. Algebr. Geom. Topol. 5, 1075–1109 (2005)MathSciNetzbMATHGoogle Scholar
- 25.Doebner, H.-D., Mann, H.-J.: Vector bundles over configuration spaces of nonidentical particles: topological potentials and internal degrees of freedom. J. Math. Phys 38, 3943 (1997)ADSMathSciNetzbMATHGoogle Scholar
- 26.Doebner, H.-D., Groth, W., Hennig, J.D.: On quantum mechanics of n-particle systems on 2-manifolds - a case study in topology. J. Geom. Phys. 31, 35-50 (1999)ADSMathSciNetzbMATHGoogle Scholar
- 27.Tolar, J.: Borel quantization and the origin of topological effects in quantum mechanics. In: Hennig, J.D., Lücke, W., Tolar, J. (eds.) Differential Geometry, Group Representations, and Quantization. Lecture Notes in Physics, vol 379. Springer, Berlin (1991)Google Scholar
- 28.Doebner, H.-D., Nattermann, P.: Borel quantization: kinematics and dynamics. Acta Phys. Polon. B 27, 2327–2339 (1996)MathSciNetzbMATHGoogle Scholar
- 29.Doebner, H.-D., Šťovíček, P., Tolar, J.: Quantization of kinematics on configuration manifolds. Rev. Math. Phys. 13, 799 (2001)MathSciNetzbMATHGoogle Scholar
- 30.Qiang, Z.: Quantum kinematics and geometric quantization. J. Geom. Phys. 21(1), 34–42 (1996)ADSMathSciNetzbMATHGoogle Scholar
- 31.Kobayashi, S.: Differential Geometry of Complex Vector Bundles. Princeton University Press, Princeton (1987)zbMATHGoogle Scholar
- 32.Roth, F.: On the category of Euclidean configuration spaces and associated fibrations. Geom. Topol. Monogr. 13, 447–461 (2008)MathSciNetzbMATHGoogle Scholar
- 33.Fadell, E.R., Husseini, S.Y.: Geometry and Topology of Configuration Spaces. Springer Monographs in Mathematics. Springer, Berlin (2001)zbMATHGoogle Scholar
- 34.De La Harpe, P.: Topics in Geometric Group Theory, Chicago Lectures in Mathematics (2000)Google Scholar
- 35.Einarsson, T.: Fractional statistics on a torus. Phys. Rev. Lett. 64(17), 1995 (1990)ADSMathSciNetzbMATHGoogle Scholar
- 36.Farley, D., Sabalka, L.: Presentations of graph braid groups. Forum Mathematicum 24(4), 827–859 (2012)MathSciNetzbMATHGoogle Scholar
- 37.Ko, K.H., ParkH, W.: Characteristics of graph braid groups. Discrete Comput. Geom. 48(4), 915–963 (2012)MathSciNetzbMATHGoogle Scholar
- 38.Hatcher, A.: Algebraic Topology. Cambridge University Press, Cambridge (2001)zbMATHGoogle Scholar
- 39.Milnor, J.W., Stasheff, J.D.: Characteristic classes. Ann. Math. Stud. 76, (1974)Google Scholar
- 40.Husemoller, D.H.: Fibre Bundles. McGraw-Hill, New York (1966)zbMATHGoogle Scholar
- 41.Atiyah, M.-F., Hirzebruch, F.: Vector bundles and homogeneous spaces. Proc. Symp. Pure Maths Amer. Math. Soc. 3, 7–38 (1961)MathSciNetzbMATHGoogle Scholar
- 42.Kobayashi, S., Nomizu, K.: Foundations of Differential Geometry. Wiley Classics Library, vol. 1. Wiley, New York (1963)zbMATHGoogle Scholar
- 43.Spanier, E.H.: Algebraic Topology. ISBN 978-0-387-94426-5. Springer, New York (1966)zbMATHGoogle Scholar
- 44.Kamber, F., Tondeur, P.: The characteristic homomorphism of flat bundles. Topology 6, 153–159 (1967)MathSciNetzbMATHGoogle Scholar
- 45.Arnold, V.I.: Collected works, Volume II: On some topological invariants of algebraic functions (1970), Alexander B. Givental, Boris A. Khesin, Alexander N. Varchenko, Victor A. Vassiliev, Oleg Ya. Viro (Eds.), ISBN 978-3-642-31030-0, Springer-Verlag Berlin Heidelberg, (2014)Google Scholar
- 46.Vainshtein, F.V.: Cohomologies of braid groups. Funct. Anal. Appl. 12, 135 (1978). https://doi.org/10.1007/BF01076259. 1978CrossRefzbMATHGoogle Scholar
- 47.Bloore, F.J.: Configuration spaces of identical particles. In: García, P.L., Pérez-Rendón, A., Souriau, J.M. (eds.) Differential Geometrical Methods in Mathematical Physics, Lecture Notes in Mathematics, vol. 836. Springer, Berlin (1980)Google Scholar
- 48.Cohen, F.R., Lada, T.J., May, J.P.: The Homology of Iterated Loop Spaces. Lecture Notes in Math., vol. 533. Springer, Berlin (1976)zbMATHGoogle Scholar
- 49.Bloore, F.J., Bratley, I., Selig, J.M.: \(SU(n)\) bundles over the configuration space of three identical particles moving on \({{\mathbb{R}}}^3\). J. Phys. A: Math. Gen. 16, 729–736 (1983)ADSzbMATHGoogle Scholar
- 50.Ramos, E.: Stability phenomena in the homology of tree braid groups. Algebr. Geom. Topol. 18, 2305–2337 (2018)MathSciNetzbMATHGoogle Scholar
- 51.Ramos, E., White, G.: Families of nested graphs with compatible symmetric-group actions (2017). arXiv:1711.07456
- 52.Ramos, E.: An application of the theory of FI-algebras to graph configuration spaces (2018). arXiv:1805.05316
- 53.Chettih, S., Lütgehetmann, D.: The homology of configuration spaces of trees with loops. Algebr. Geom. Topol. 18, 2443–2469S (2018)MathSciNetzbMATHGoogle Scholar
- 54.Lütgehetmann, D.: Representation Stability for Configuration Spaces of Graphs (2017). arXiv:1701.03490
- 55.Byung Hee An, Drummond-Cole, G.C., Knudsen, B.: Edge stabilization in the homology of graph braid groups, preprint (2018). arXiv:1806.05585
- 56.Barnett, K., Farber, M.: Topology of configuration space of two particles on a graph. Algebr. Geom. Topol. 9, 593–624 (2009)MathSciNetzbMATHGoogle Scholar
- 57.Farley, D.: Homology of tree braid groups in Topological and Asymptotic Aspects of Group Theory. In: Rostislav Grigorchuk, Michael Mihalik, Mark Sapir, Zoran Sunik (Eds.) Contemporary Mathematics, vol. 394 (2006). https://doi.org/10.1090/conm/394 Google Scholar
- 58.Forman, R.: Morse theory for cell complexes. Adv. Math. 134, 90145 (1998)MathSciNetzbMATHGoogle Scholar
- 59.Sawicki, A.: Discrete Morse functions for graph configuration spaces. J. Phys. A: Math. Theor. 45, 505202 (2012)MathSciNetzbMATHGoogle Scholar
- 60.Maciążek, T.: An implementation of discrete Morse theory for graph configuration spaces (2019). www.github.com/tmaciazek/graph-morse
- 61.Gal, Ś.: Euler characteristic of the configuration space of a complex. Colloq. Math. 89(1), 61–67 (2001)MathSciNetzbMATHGoogle Scholar
- 62.Gallier, J., Xu, D.: A Guide to the Classification Theorem for Compact Surfaces. Springer, Berlin (2013)zbMATHGoogle Scholar
- 63.Kitaev, A.: Anyons in an exactly solved model and beyond. Ann. Phys. 321, 2–111 (2006)ADSMathSciNetzbMATHGoogle Scholar
- 64.Etingof, P., Nikshych, D., Ostrik, V.: On fusion categories. Ann. Math. 162, 581–642 (2005)MathSciNetzbMATHGoogle Scholar
Copyright information
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.