Graph homology

Last updated

In algebraic topology and graph theory, graph homology describes the homology groups of a graph, where the graph is considered as a topological space. It formalizes the idea of the number of "holes" in the graph. It is a special case of a simplicial homology, as a graph is a special case of a simplicial complex. Since a finite graph is a 1-complex (i.e., its 'faces' are the vertices - which are 0-dimensional, and the edges - which are 1-dimensional), the only non-trivial homology groups are the 0-th group and the 1-th group. [1]

Contents

The 1st homology group

The general formula for the 1st homology group of a topological space X is:

The example below explains these symbols and concepts in full detail on a graph.

Example

Let X be a directed graph with 3 vertices {x,y,z} and 4 edges {a: x→y, b: y→z, c: z→x, d: z→x}. It has several cycles:

If we cut the plane along the loop a+b+d, and then cut at c and "glue" at d, we get a cut along the loop a+b+c. This can be represented by the following relation: (a+b+d) + (c-d) = (a+b+c). To formally define this relation, we define the following commutative groups: [2] :6:00

Most elements of C1 are not cycles, for example a+b, 2a+5b-c, etc. are not cycles. To formally define a cycle, we first define boundaries. The boundary of an edge is denoted by the operator and defined as its target minus its source, so So is a mapping from the group C1 to the group C0. Since a,b,c,d are the generators of C1, this naturally extends to a group homomorphism from C1 to C0. In this homomorphism, . Similarly, maps any cycle in C1 to the zero element of C0. In other words, the set of cycles in C1 generates the null space (the kernel) of . In this case, the kernel of has two generators: one corresponds to a+b+c and the other to a+b+d (the third cycle, c-d, is a linear combination of the first two). So ker is isomorphic to Z2.

In a general topological space, we would define higher-dimensional chains. In particular, C2 would be the free abelian group on the set of 2-dimensional objects. However, in a graph there are no such objects, so C2 is a trivial group. Therefore, the image of the second boundary operator, , is trivial too. Therefore:

This corresponds to the intuitive fact that the graph has two "holes". The exponent is the number of holes.

General case

The above example can be generalized to an arbitrary connected graph G = (V, E). Let T be a spanning tree of G. Every edge in E \ T corresponds to a cycle; these are exactly the linearly independent cycles. Therefore, the first homology group H1 of a graph is the free abelian group with |E \ T| generators. This number equals |E|-|V|+1; so: [1]

In a disconnected graph, when C is the set of connected components, a similar computation shows:

In particular, the first group is trivial if and only if X is a forest.

The 0-th homology group

The general formula for the 0-th homology group of a topological space X is:

Example

We return to the graph with 3 vertices {x,y,z} and 4 edges {a: x→y, b: y→z, c: z→x, d: z→x}. Recall that the group C0 is generated by the set of vertices. Since there are no (−1)-dimensional elements, the group C−1 is trivial, and so the entire group C0 is a kernel of the corresponding boundary operator: = the free abelian group generated by {x,y,z}. [3]

The image of contains an element for each pair of vertices that are boundaries of an edge, i.e., it is generated by the differences {y−x, z−y, x−z}. To calculate the quotient group, it is convenient to think of all the elements of as "equivalent to zero". This means that x, y and z are equivalent - they are in the same equivalence class of the quotient. In other words, is generated by a single element (any vertex can generate it). So it is isomorphic to Z.

General case

The above example can be generalized to any connected graph. Starting from any vertex, it is possible to get to any other vertex by adding to it one or more expressions corresponding to edges (e.g. starting from x, one can get to z by adding y-x and z-y). Since the elements of are all equivalent to zero, it means that all vertices of the graph are in a single equivalence class, and therefore is isomorphic to Z.

In general, the graph can have several connected components. Let C be the set of components. Then, every connected component is an equivalence class in the quotient group. Therefore:

It can be generated by any |C|-tuple of vertices, one from each component.

Reduced homology

Often, it is convenient to assume that the 0-th homology of a connected graph is trivial (so that, if the graph contains a single point, then all its homologies are trivial). This leads to the definition of the reduced homology. For a graph, the reduced 0-th homology is:

This "reduction" affects only the 0-th homology; the reduced homologies of higher dimensions are equal to the standard homologies.

Higher dimensional homologies

A graph has only vertices (0-dimensional elements) and edges (1-dimensional elements). We can generalize the graph to an abstract simplicial complex by adding elements of a higher dimension. Then, the concept of graph homology is generalized by the concept of simplicial homology .

Example

In the above example graph, we can add a two-dimensional "cell" enclosed between the edges c and d; let's call it A and assume that it is oriented clockwise. Define C2 as the free abelian group generated by the set of two-dimensional cells, which in this case is a singleton {A}. Each element of C2 is called a 2-dimensional chain.

Just like the boundary operator from C1 to C0, which we denote by , there is a boundary operator from C2 to C1, which we denote by . In particular, the boundary of the 2-dimensional cell A are the 1-dimensional edges c and d, where c is in the "correct" orientation and d is in a "reverse" orientation; therefore: . The sequence of chains and boundary operators can be presented as follows: [4]

The addition of the 2-dimensional cell A implies that its boundary, c-d, no longer represents a hole (it is homotopic to a single point). Therefore, the group of "holes" now has a single generator, namely a+b+c (it is homotopic to a+b+d). The first homology group is now defined as the quotient group:

Here, is the group of 1-dimensional cycles, which is isomorphic to Z2, and is the group of 1-dimensional cycles that are boundaries of 2-dimensional cells, which is isomorphic to Z. Hence, their quotient H1 is isomorphic to Z. This corresponds to the fact that X now has a single hole. Previously. the image of was the trivial group, so the quotient was equal to . Suppose now that we add another oriented 2-dimensional cell B between the edges c and d, such that . Now C2 is the free abelian group generated by {A,B}. This does not change H1 - it is still isomorphic to Z (X still has a single 1-dimensional hole). But now C2 contains the two-dimensional cycle A-B, so has a non-trivial kernel. This cycle generates the second homology group, corresponding to the fact that there is a single two-dimensional hole:

We can proceed and add a 3-cell - a solid 3-dimensional object (called C) bounded by A and B. Define C3 as the free abelian group generated by {C}, and the boundary operator . We can orient C such that ; note that the boundary of C is a cycle in C2. Now the second homology group is:

corresponding to the fact that there are no two-dimensional holes (C "fills the hole" between A and B).

General case

In general, one can define chains of any dimension. If the maximum dimension of a chain is k, then we get the following sequence of groups:

It can be proved that any boundary of a (k+1)-dimensional cell is a k-dimensional cycle. In other words, for any k, (the group of boundaries of k+1 elements) is contained in (the group of k-dimensional cycles). Therefore, the quotient is well-defined, and it is defined as the k-th homology group:

Related Research Articles

<span class="mw-page-title-main">Exact sequence</span> Sequence of homomorphisms such that each kernel equals the preceding image

An exact sequence is a sequence of morphisms between objects such that the image of one morphism equals the kernel of the next.

In mathematics, a chain complex is an algebraic structure that consists of a sequence of abelian groups and a sequence of homomorphisms between consecutive groups such that the image of each homomorphism is included in the kernel of the next. Associated to a chain complex is its homology, which describes how the images are included in the kernels.

<span class="mw-page-title-main">Homological algebra</span> Branch of mathematics

Homological algebra is the branch of mathematics that studies homology in a general algebraic setting. It is a relatively young discipline, whose origins can be traced to investigations in combinatorial topology and abstract algebra at the end of the 19th century, chiefly by Henri Poincaré and David Hilbert.

In mathematics, homology is a general way of associating a sequence of algebraic objects, such as abelian groups or modules, with other mathematical objects such as topological spaces. Homology groups were originally defined in algebraic topology. Similar constructions are available in a wide variety of other contexts, such as abstract algebra, groups, Lie algebras, Galois theory, and algebraic geometry.

In mathematics, a self-adjoint operator on an infinite-dimensional complex vector space V with inner product is a linear map A that is its own adjoint. If V is finite-dimensional with a given orthonormal basis, this is equivalent to the condition that the matrix of A is a Hermitian matrix, i.e., equal to its conjugate transpose A. By the finite-dimensional spectral theorem, V has an orthonormal basis such that the matrix of A relative to this basis is a diagonal matrix with entries in the real numbers. This article deals with applying generalizations of this concept to operators on Hilbert spaces of arbitrary dimension.

In mathematics, specifically in homology theory and algebraic topology, cohomology is a general term for a sequence of abelian groups, usually one associated with a topological space, often defined from a cochain complex. Cohomology can be viewed as a method of assigning richer algebraic invariants to a space than homology. Some versions of cohomology arise by dualizing the construction of homology. In other words, cochains are functions on the group of chains in homology theory.

In algebraic topology, the Betti numbers are used to distinguish topological spaces based on the connectivity of n-dimensional simplicial complexes. For the most reasonable finite-dimensional spaces, the sequence of Betti numbers is 0 from some point onward, and they are all finite.

A CW complex is a kind of a topological space that is particularly important in algebraic topology. It was introduced by J. H. C. Whitehead to meet the needs of homotopy theory. This class of spaces is broader and has some better categorical properties than simplicial complexes, but still retains a combinatorial nature that allows for computation. The C stands for "closure-finite", and the W for "weak" topology.

In mathematics, particularly algebraic topology and homology theory, the Mayer–Vietoris sequence is an algebraic tool to help compute algebraic invariants of topological spaces, known as their homology and cohomology groups. The result is due to two Austrian mathematicians, Walther Mayer and Leopold Vietoris. The method consists of splitting a space into subspaces, for which the homology or cohomology groups may be easier to compute. The sequence relates the (co)homology groups of the space to the (co)homology groups of the subspaces. It is a natural long exact sequence, whose entries are the (co)homology groups of the whole space, the direct sum of the (co)homology groups of the subspaces, and the (co)homology groups of the intersection of the subspaces.

In algebraic topology, singular homology refers to the study of a certain set of algebraic invariants of a topological space X, the so-called homology groups Intuitively, singular homology counts, for each dimension n, the n-dimensional holes of a space. Singular homology is a particular example of a homology theory, which has now grown to be a rather broad collection of theories. Of the various theories, it is perhaps one of the simpler ones to understand, being built on fairly concrete constructions.

In functional analysis, a branch of mathematics, a compact operator is a linear operator , where are normed vector spaces, with the property that maps bounded subsets of to relatively compact subsets of . Such an operator is necessarily a bounded operator, and so continuous. Some authors require that are Banach, but the definition can be extended to more general spaces.

<span class="mw-page-title-main">Barycentric subdivision</span>

In mathematics, the barycentric subdivision is a standard way to subdivide a given simplex into smaller ones. Its extension on simplicial complexes is a canonical method to refine them. Therefore, the barycentric subdivision is an important tool in algebraic topology.

In mathematics, particularly in combinatorics, given a family of sets, here called a collection C, a transversal (also called a cross-section) is a set containing exactly one element from each member of the collection. When the sets of the collection are mutually disjoint, each element of the transversal corresponds to exactly one member of C (the set it is a member of). If the original sets are not disjoint, there are two possibilities for the definition of a transversal:

In algebraic topology, universal coefficient theorems establish relationships between homology groups (or cohomology groups) with different coefficients. For instance, for every topological space X, its integral homology groups:

In algebraic topology, simplicial homology is the sequence of homology groups of a simplicial complex. It formalizes the idea of the number of holes of a given dimension in the complex. This generalizes the number of connected components.

In algebraic topology, a branch of mathematics, the (singular) homology of a topological space relative to a subspace is a construction in singular homology, for pairs of spaces. The relative homology is useful and important in several ways. Intuitively, it helps determine what part of an absolute homology group comes from which subspace.

In complex geometry, a polar homology is a group which captures holomorphic invariants of a complex manifold in a similar way to usual homology of a manifold in differential topology. Polar homology was defined by B. Khesin and A. Rosly in 1999.

In mathematics, and more precisely in topology, the mapping class group of a surface, sometimes called the modular group or Teichmüller modular group, is the group of homeomorphisms of the surface viewed up to continuous deformation. It is of fundamental importance for the study of 3-manifolds via their embedded surfaces and is also studied in algebraic geometry in relation to moduli problems for curves.

In mathematics, the Bloch group is a cohomology group of the Bloch–Suslin complex, named after Spencer Bloch and Andrei Suslin. It is closely related to polylogarithm, hyperbolic geometry and algebraic K-theory.

In mathematics, a derivation of a commutative ring is called a locally nilpotent derivation (LND) if every element of is annihilated by some power of .

References

  1. 1 2 Sunada, Toshikazu (2013), Sunada, Toshikazu (ed.), "Homology Groups of Graphs", Topological Crystallography: With a View Towards Discrete Geometric Analysis, Surveys and Tutorials in the Applied Mathematical Sciences, Tokyo: Springer Japan, pp. 37–51, doi:10.1007/978-4-431-54177-6_4, ISBN   978-4-431-54177-6
  2. Wildberger, Norman J. (2012). "An introduction to homology". YouTube .
  3. Wildberger, Norman J. (2012). "Computing homology groups". YouTube .
  4. Wildberger, Norman J. (2012). "An introduction to homology (cont.)". YouTube .