Homologies of digraphs and Künneth formulas

Similar documents
On a cohomology of digraphs and Hochschild cohomology

CELLULAR HOMOLOGY AND THE CELLULAR BOUNDARY FORMULA. Contents 1. Introduction 1

The Hurewicz Theorem

p,q H (X), H (Y ) ), where the index p has the same meaning as the

Math 6510 Homework 10

Math 530 Lecture Notes. Xi Chen

7.3 Singular Homology Groups

EILENBERG-ZILBER VIA ACYCLIC MODELS, AND PRODUCTS IN HOMOLOGY AND COHOMOLOGY

Partial cubes: structures, characterizations, and constructions

ALGEBRAIC PROPERTIES OF BIER SPHERES

DISCRETIZED CONFIGURATIONS AND PARTIAL PARTITIONS

Linear Algebra March 16, 2019

LECTURE 3: RELATIVE SINGULAR HOMOLOGY

Simplicial Homology. Simplicial Homology. Sara Kališnik. January 10, Sara Kališnik January 10, / 34

HOMOLOGY AND COHOMOLOGY. 1. Introduction

Manifolds and Poincaré duality

arxiv: v1 [math.co] 25 Jun 2014

58 CHAPTER 2. COMPUTATIONAL METHODS

Formal power series rings, inverse limits, and I-adic completions of rings

ALGEBRAIC GEOMETRY COURSE NOTES, LECTURE 2: HILBERT S NULLSTELLENSATZ.

Topological properties

and this makes M into an R-module by (1.2). 2

Course 311: Michaelmas Term 2005 Part III: Topics in Commutative Algebra

LECTURE IV: PERFECT PRISMS AND PERFECTOID RINGS

The Topology of Intersections of Coloring Complexes

Discrete Mathematics. Benny George K. September 22, 2011

2. Prime and Maximal Ideals

L E C T U R E N O T E S O N H O M O T O P Y T H E O R Y A N D A P P L I C AT I O N S

Chapter 1 Preliminaries

CW-complexes. Stephen A. Mitchell. November 1997

Unmixed Graphs that are Domains

Where is matrix multiplication locally open?

6 Axiomatic Homology Theory

Homotopy and homology groups of the n-dimensional Hawaiian earring

Eilenberg-Steenrod properties. (Hatcher, 2.1, 2.3, 3.1; Conlon, 2.6, 8.1, )

EXT, TOR AND THE UCT

V (v i + W i ) (v i + W i ) is path-connected and hence is connected.

QUATERNIONS AND ROTATIONS

Cographs; chordal graphs and tree decompositions

Chapter 1. Sets and Numbers

Topological Data Analysis - Spring 2018

An Algebraic View of the Relation between Largest Common Subtrees and Smallest Common Supertrees

MATH8808: ALGEBRAIC TOPOLOGY

De Rham Cohomology. Smooth singular cochains. (Hatcher, 2.1)

PERVERSE SHEAVES ON A TRIANGULATED SPACE

Elementary linear algebra

ALGEBRAICALLY TRIVIAL, BUT TOPOLOGICALLY NON-TRIVIAL MAP. Contents 1. Introduction 1

Groups of Prime Power Order with Derived Subgroup of Prime Order

121B: ALGEBRAIC TOPOLOGY. Contents. 6. Poincaré Duality

Linear Algebra Notes. Lecture Notes, University of Toronto, Fall 2016

Representations of quivers

We simply compute: for v = x i e i, bilinearity of B implies that Q B (v) = B(v, v) is given by xi x j B(e i, e j ) =

Bordism and the Pontryagin-Thom Theorem

Mathematics Course 111: Algebra I Part I: Algebraic Structures, Sets and Permutations

Let X be a topological space. We want it to look locally like C. So we make the following definition.

30 Surfaces and nondegenerate symmetric bilinear forms

Intersections of Leray Complexes and Regularity of Monomial Ideals

1 Differentiable manifolds and smooth maps

Generalized Pigeonhole Properties of Graphs and Oriented Graphs

Lax embeddings of the Hermitian Unital

Topology Hmwk 6 All problems are from Allen Hatcher Algebraic Topology (online) ch 2

Math 210B. Artin Rees and completions

Smith theory. Andrew Putman. Abstract

arxiv: v2 [math-ph] 24 Feb 2016

08a. Operators on Hilbert spaces. 1. Boundedness, continuity, operator norms

The Cubical Cohomology Ring: An Algorithmic Approach

4. Images of Varieties Given a morphism f : X Y of quasi-projective varieties, a basic question might be to ask what is the image of a closed subset

Math 341: Convex Geometry. Xi Chen

Chapter 3: Proving NP-completeness Results

Course MA2C02, Hilary Term 2010 Section 4: Vectors and Quaternions

Equivalence of the Combinatorial Definition (Lecture 11)

8. Prime Factorization and Primary Decompositions

Math Homotopy Theory Hurewicz theorem

ON THE NUMBER OF COMPONENTS OF A GRAPH

A duality on simplicial complexes

Part II. Algebraic Topology. Year

UNIVERSAL DERIVED EQUIVALENCES OF POSETS

Chapter 1. Measure Spaces. 1.1 Algebras and σ algebras of sets Notation and preliminaries

Topological automorphisms of modified Sierpiński gaskets realize arbitrary finite permutation groups

DERIVATIONS. Introduction to non-associative algebra. Playing havoc with the product rule? BERNARD RUSSO University of California, Irvine

Derived Algebraic Geometry I: Stable -Categories

Sets and Motivation for Boolean algebra

X-MA2C01-1: Partial Worked Solutions

Math 203, Solution Set 4.

Automata on linear orderings

Math 6802 Algebraic Topology II

Infinite-Dimensional Triangularization

Dimension of the mesh algebra of a finite Auslander-Reiten quiver. Ragnar-Olaf Buchweitz and Shiping Liu

TOWARDS A SPLITTER THEOREM FOR INTERNALLY 4-CONNECTED BINARY MATROIDS VII

12. Hilbert Polynomials and Bézout s Theorem

Topics in linear algebra

(iv) Whitney s condition B. Suppose S β S α. If two sequences (a k ) S α and (b k ) S β both converge to the same x S β then lim.

NOTES ON BASIC HOMOLOGICAL ALGEBRA 0 L M N 0

Math 121 Homework 5: Notes on Selected Problems

arxiv: v2 [math.ag] 24 Jun 2015

THE ENVELOPE OF LINES MEETING A FIXED LINE AND TANGENT TO TWO SPHERES

IDEAL CLASSES AND RELATIVE INTEGERS

SECTION 5: EILENBERG ZILBER EQUIVALENCES AND THE KÜNNETH THEOREMS

1. Simplify the following. Solution: = {0} Hint: glossary: there is for all : such that & and

Relations Graphical View

Transcription:

Homologies of digraphs and Künneth formulas Alexander Grigor yan Department of Mathematics University of Bielefeld 33501 Bielefeld, Germany Yuri Muranov Department of Mathematics University of Warmia and Mazury Olsztyn, Poland Shing-Tung Yau Department of Mathematics Harvard University Cambridge MA 02138, USA September 2015 Abstract We state and prove Künneth formulas for path homologies of Cartesian product and join of two digraphs. Contents 1 Introduction 2 2 Path homologies 3 2.1 Paths on finite sets................................... 3 2.2 Regular paths...................................... 5 2.3 Paths on digraphs................................... 6 2.4 Cyclic digraphs..................................... 8 3 Join of digraphs 10 3.1 Join of two digraphs.................................. 10 3.2 Some auxiliary results................................. 14 4 Cartesian product 14 4.1 Step-like paths..................................... 15 4.2 Cartesian product of digraphs............................. 19 4.3 Some auxiliary results................................. 23 5 -Invariant elements on abstract products 25 6 Proof of Theorems 3.3 and 4.7 32 Partially supported by SFB 701 of German Research Council and by Visiting Grants of Harvard University and MSC, Tsinghua University Partially supported by SFB 701 of German Research Council. Partially supported by the grant Geometry and Topology of Complex Networks, no. FA-9550-13-1-0097 1

References 35 1 Introduction The purpose of this paper is to prove Künneth formulas for homology groups of digraphs (directed graphs). We use the notion of path homology of digraphs that was introduced in [9] and [10]. A digraph G is a pair (V, E) where V is a set of vertices and E the set of edges that is a subset of V V \ diag. An n-path on G is a formal linear combination of sequences v 0,..., v n of n + 1 vertices. If in any sequence, involved in the linear combination, all pairs (v i, v i+1 ) are edges, then the n-path is called allowed. There is a natural definition of the boundary of an n-path that is an (n 1)-path. However, the boundary of an allowed path does not have to be allowed. Those allowed paths whose boundaries are also allowed, are called -invariant. The linear space of -invariant n-paths is an element of the chain complex whose homologies are called the path homologies of the digraph. An undirected graph can also be considered as a digraph by replacing each undirected edge by two oppositely directed edges. In this way we obtain path homologies of undirected graphs. There has been a number of attempts to define the notions of homology and cohomology for graphs. At a trivial level, any graph can be regarded as an one-dimensional simplicial complex, so that its simplicial homologies are defined. However, all homology groups of order 2 and higher are trivial, which makes this approach uninteresting. Another way to make a graph into a simplicial complex is to consider all its cliques (complete subgraphs) as simplexes of the corresponding dimensions (cf. [5], [15]). Then higher dimensional homologies may be non-trivial, but in this approach the notion of graph looses its identity and becomes a particular case of the notion of a simplicial complex. Besides, some desirable functorial properties of homologies fail, for example, the Künneth formula is not true for Cartesian product of graphs. Yet another approach to homologies of digraphs can be realized via Hochschild homology. Indeed, allowed paths on a digraph have a natural operation of product, which allows to define the notion of a path algebra of a digraph. The Hochschild homology of the path algebra is a natural object to consider. However, it was shown in [14] that Hochschild homologies of order 2 are trivial, which makes this approach less attractive. More recently there have been a number of attempts to define singular homologies of graphs [2], [17]. In these theories one uses predefined small graphs as basic cell elements and defines singular chains using maps of the basic cells into the graph. However, simple examples show that the homology groups obtained in this way, depend essentially on the choice of the basic cells and, moreover, lack some the functorial properties of homologies. Not to say that computation of singular homologies for the graphs beyond the trivial ones is very hard. The path homologies of digraphs that are dealt with in this paper have many advantages in comparison with other notions of graph homologies. Firstly, the homologies of all dimensions could be non-trivial. Also, the chain complex may have a rich structure. It contains not only cliques but also binary hypercubes and many other interesting shapes. Besides, path homologies can be relatively easily computed, by definition for small graphs and by means of any conventional linear algebra computing package for larger graphs. Secondly, there is an independently developed notion of homotopy of digraphs [ 10] (similar to homotopy theory on graphs [1], [3]) that is compatible with path homology. For example, homotopy equivalent digraphs have the same homology groups, and the first homology group is abelization of the fundamental group. 2

Thirdly, there is a dual cohomology theory with the coboundary operator that arises independently and naturally as an exterior derivative on the algebra of functions on the vertex set of the digraph. This approach to the cohomology of digraphs, that is based on the classification of Bourbaki [4] of exterior derivations on algebras, was introduced by Dimakis and Müller-Hoissen in [6], [7] and further developed in [13]. Here we do not touch cohomologies of digraph and refer the reader to [9] for details. Finally, the notion of path homology has good functorial properties with respect to graphtheoretical operations. The main result of this paper goes exactly in this direction: we prove the Künneth formulas for the path homologies of the Cartesian product and of the join of two digraphs. We feel that the notion of path homology of digraphs has a rich mathematical content and hope that it will become a useful tool in various areas of pure and applied mathematics. For example, this notion was employed in [12] to give a new elementary proof of a theorem of Gerstenhaber and Schack [8] that identifies simplicial homology as a Hochschild homology. A link between path homologies of digraphs and cubical homologies was revealed in [ 11]. On the other hand, it is conceivable that the notion of path homology can be used in practical applications such as coverage verification in sensor networks (cf. [16]), and many others. Let us briefly describe the structure of the paper. In Section 2 we define the notion of the boundary operator, path homology and give some simple examples. In Section 3 we introduce the operation join of two digraphs and state a Künneth formula for it (Theorem 3.3). Particular cases of join are operations of building a cone and suspension over a digraph, which behave homologically in the same way as those in the classical algebraic topology. In Section 4 we introduce the notions of cross product of paths and Cartesian product of digraphs. We state a Künneth formula for Cartesian product (Theorem 4.7) and give some examples. In Sections 5, 6 we prove both Theorems 3.3 and 4.7 in a unified way. Note that in the both cases of join and Cartesian product we prove not only Künneth formulas for homologies but also similar formulas for chain complexes that have no analog in the classical algebraic topology. The main difficulty in the proof lies in distinction between the notions of allowed paths and -invariant paths. This difficulty does not occur in the classical algebraic topology and in order to overcome it we have developed a new tool of homological algebra that is stated in Theorem 5.1. 2 Path homologies 2.1 Paths on finite sets Let V be an arbitrary non-empty finite set whose elements will be called vertices. For any non-negative integer p, an elementary p-path on a set V is any sequence {i k } p k0 of p + 1 vertices of V (the vertices in the path do not have to be distinct). For p 1, an elementary p-path is an empty set. The p-path {i k } p k0 will also be denoted simply by i 0...i p, without delimiters between the vertices. Fix a field K and consider a K-linear space Λ p Λ p (V ) that consists of all formal linear combinations of all elementary p-paths with the coefficients from K. The elements of Λ p are called p-paths on V. An elementary p-path i 0...i p as an element of Λ p will be denoted by e i0...i p. The empty set as an element of Λ 1 will be denoted by e. By definition, the family { e i0...i p : i 0,..., i p V } is a basis in Λ p. Hence, each p-path v has a 3

unique representation in the form v v i 0...ip e i0...i p, (2.1) i 0,...,i p V where v i 0...ip K are the components of v. For example, Λ 0 consists of all linear combinations of elements e i that are the vertices of V, Λ 1 consists of all linear combinations of the elements e ij that are pairs of vertices, etc. Note that, Λ 1 consists of all multiples of e so that Λ 1 K. Definition 2.1 For any p 0, the boundary operator : Λ p Λ p 1 is a K-linear operator that acts on elementary paths by e i0...i p p ( 1) q e i0...î q...i p, (2.2) q0 where the hat î q means omission of the index i q. For example, we have It follows that, for any v Λ p, e i e, e ij e j e i, e ijk e jk e ik + e ij. (2.3) ( v) j 0...j p 1 k V p ( 1) q v j 0...j q 1 k j q...j p 1. (2.4) q0 Set also Λ 2 {0} and define : Λ 1 Λ 2 to be zero. Lemma 2.2 We have 2 v 0 for any v Λ p with p 0. Proof. For p 0 this is trivial. For p 1 we have by (2.2) 2 e i0...i p p ( 1) q e i0...î q...i p q0 p q 1 ( 1) q ( 1) r e i0...î r...î q...i p + q0 r0 ( 1) q+r e i0...î r...î q...i p p ( 1) r 1 e i0...î q...î r...i p rq+1 0 r<q p 0 q<r p ( 1) q+r e i0...î q...î r...i p. After switching q and r in the last sum we see that the two sums cancel out, whence 2 e i0...i p 0. This implies 2 v 0 for all v Λ p. Definition 2.3 For all p, q 1 and for any two paths u Λ p and v Λ q define their join uv Λ p+q+1 as follows: (uv) i 0...i pj 0...j q u i 0...i p v j 0...j q. (2.5) Clearly, join of paths is a bilinear operation that satisfies the associative law (but not commutative). It follows from (2.5) that e i0...i p e j0...j q e i0...i pj 0...j q. (2.6) Let us extend the definition of uv to the case when either p 2 and q 1 or q 2 and p 1, just by setting uv 0 Λ p+q+1. 4

Lemma 2.4 (Product rule for join) For all p, q 1 and u Λ p, v Λ q we have (uv) ( u)v + ( 1) p+1 u v. (2.7) Proof. Assume first that p, q 0. It suffices to prove (2.7) for u e i0...i p and v e j0...j q. We have (uv) e i0...i pj 0...j q e i1...i pj 0...j q e i0 i 2...i pj 0...j q +... + ( 1) p+1 ( e i0...i p j 1...j q e i0...i p j 0 j 2...j q +... ) ( e i0...i p ) ej0...j q + ( 1) p+1 e i0...i p e j0...j q, whence (2.7) follows. If p 1 then it suffices to prove (2.7) for u e. In this case uv v, ( u) v 0, and u v v so that the both sides of (2.7) are equal to v. The case q 1 is similar. 2.2 Regular paths Definition 2.5 An elementary p-path e i0...i p k 0,..., p 1, and non-regular otherwise. on a set V is called regular if i k i k+1 for all For any p 1, consider the following subspace of Λ p R p R p (V ) : span { e i0...i p : i 0...i p is regular }, whose elements are called regular p-paths. We would like to consider the operator on the spaces R p. However, in the present form is not invariant on the spaces R p. For example, e iji R 2 for i j while e iji e ji e ii + e ij contains a non-regular component e ii. The same applies to the notion of join of paths: the join of two regular path does not have to be regular, for example, e ij e ji e ijji. To overcome this difficulty, consider the complementary subspace I p I p (V ) : span { e i0...i p : i 0...i p is non-regular } and observe the following property of I p. Lemma 2.6 Let u I p. Then u I p 1 and uv I p+q+1 for any v Λ q. Proof. It suffices to prove both claims for u e i0...i p. Since this path is non-regular, there exists an index k such that i k i k+1. Then we have e i0...i p e i1...i p e i0 i 2...i p +... + ( 1) k e i0...i k 1 i k+1 i k+2...i p + ( 1) k+1 e i0...i k 1 i k i k+2...i p (2.8) +... + ( 1) p e i0...i p 1. By i k i k+1 the two terms in the middle line of (2.8) cancel out, whereas all other terms are non-regular, whence e i0...i p I p 1. If v e j0...j q then uv e i0...i pj 0...j q is obviously non-regular. It follows from Lemma 2.6 that the both operations and join are well defined on the quotients Λ p /I p. On the other hand, it is clear that Λ p R p I p and, hence, R p Λp /I p, where each element of Λ p /I p has a unique representative in R p. 5

Definition 2.7 Define the regular operations and join on the spaces R p as pullbacks of those from Λ p /I p using the natural linear isomorphism R p Λp /I p. The previously defined operations and join on Λ p will be then referred to as non-regular. Of course, Lemmas 2.2 and 2.4 remain true for the regular operations. When applying the formulas (2.2), (2.4) for the regular boundary operator and (2.5), (2.6) for regular join, one should make the following adjustments: (I) all the components v i 0...i p of v R p for non-regular paths i 0...i p are equal to 0 by definition; (II) all non-regular elementary paths e i0...i p, should they arise as a result of an operation, are treated as zeros. For example, for non-regular operator : Λ 2 Λ 1 we have e iji e ji e ii + e ij whereas for the regular operator : R 2 R 1 we have e iji e ji + e ij since e ii is set to be zero. Similarly, for non-regular join we have e ij e ji e ijji whereas for the regular join e ij e ji 0. Hence, we obtain the regular chain complex of the set V : 0 K R 0 R 1... R p 1 R p... (2.9) where all the arrows are given by regular operator. We will need also the truncated regular chain complex 0 R 0 R 1... R p 1 R p... (2.10) where we follow a different convention for R 1, by setting R 1 0. In this case v for v R 0 is redefined by v 0. Note that for the latter definition the product rule (2.7) for join breaks down if p 0 or q 0. However, (2.10) will be useful when dealing with cross product of paths in Section 4. In the rest of this paper means always the regular boundary operator acting on R p, although in two modifications: (2.9) and (2.10), depending on the context. More precisely, in Section 3 we use (2.9) whereas in Section 4 the version (2.10). 2.3 Paths on digraphs A digraph is a pair G (V, E) where V is an arbitrary set, called the set of vertices, and E V V \ {diag} is the set of directed edges. For vertices a, b V the fact that (a, b) E will be denoted by a b. Note that a b excludes a b. The set V will always be assume finite. Let n 1 be an integer. Definition 2.8 An elementary n-path i 0...i n Λ n (V ) is called allowed on (V, E) if i k 1 i k for all k 1,..., n. Denote by E n the set of all allowed n-paths on (V, E). We have E 1 {e}, E 0 V and E 1 E. For any integer n 1 denote by A n the K-linear space that is spanned by all the paths from E n, that is A n A n (G) span {e i0...i n : i 0...i n is allowed}. Set also A 2 {0}. Definition 2.9 The elements of A n (G) are called allowed n-paths. 6

By construction, A n is a subspace of R n. For example, A p Λ p R p for p 0, while A 1 is spanned by all the edges from E and can be smaller than R 1. Sometimes we will need also the space N n N n (G) span {e i0...i n : i 0...i n is regular and non-allowed}. (2.11) Clearly, we have R n A n N n. We would like to restrict the regular boundary operator to the spaces A n. For some digraphs it can happen that A n A n 1, so that the restriction of to A n is straightforward. However, in general A n does not have to be a subspace of A n 1. For example, this is the case for a digraph 1 0 2 where the 2-path e 012 is allowed, while e 012 e 12 e 02 + e 01 is non-allowed because e 02 is non-allowed. For a general digraph G (V, E) and for any n 1, consider the following subspaces of A n : Ω n Ω n (G) : {v A n : v A n 1 }. (2.12) Set also Ω 2 {0}. Note that Ω n A n for n 1; in particular Ω 0 consists of all K-linear combinations of the vertices and Ω 1 consists of all R-linear combination of the edges, so that dim Ω 0 V and dim Ω 1 E. For n 2 the space Ω n can be actually smaller that A n. Claim. We have Ω n Ω n 1 for all n 1. Proof. Indeed, if v Ω n then v A n 1 and ( v) 2 v 0 A n 2 whence it follows that v Ω n 1, which was to be proved. Definition 2.10 The elements of Ω n (G) are called -invariant n-paths. Thus, we obtain the chain complex of -invariant paths: 0 K Ω 0 Ω 1... Ω n 1 Ω n Ω n+1... (2.13) where all arrows are given by, which is a subcomplex of (2.9). Consider also its truncated version 0 Ω 0 Ω 1... Ω n 1 Ω n Ω n+1... (2.14) that is a subcomplex of (2.10). In the case (2.13) we have Ω 1 A 1 K, whereas in the case (2.14) we redefine Ω 1 {0}. Definition 2.11 The homology groups of (2.14) are referred to as the path homology groups of the digraph G and are denoted by H n (G, K), n 0, that is, H n (G, K) ker Ωn / Im Ωn+1. The homology groups of (2.13) are called the reduced path homology groups of G and are denoted by H n (G, K), n 1. 7

We will also use short notations H n (G) and H n (G) since the field K is usually fixed. Clearly, H n (G) H n (G) for n 1, H 0 (G) H 0 (G) /K and H 1 (G) {0}. Although we have assumed from the very beginning that K is a field, the notion of path homology can be defined in the same way when K is a commutative unital ring. However, the main theorems of the present paper are proved under the assumption that K is a field. The notion of path homologies of digraphs is the main object of this paper. It is easy to prove that dim H 0 (G) is equal to the number of (undirected) connected components of the G. In particular, for connected digraphs dim H 0 (G) 1 and, hence, H0 (G) {0}. Example 2.12 Consider a digraph on Fig. 1. A direct computation shows that H 1 (G) {0} and H 2 (G) K; moreover, H 2 (G) is generated by e 124 + e 234 + e 314 (e 125 + e 235 + e 315 ), which will be proved in Example 3.8. It is easy to see that G is a planar digraph but nevertheless its second homology group is non-trivial. 4 3 1 5 2 Figure 1: An example of a planar digraph G with non-trivial H 2 2.4 Cyclic digraphs To give an example of computation of Ω p (G) and H p (G), we consider here a class of cyclic digraphs. We say that a digraph G (V, E) is a cycle if it is connected (as an undirected graph), every vertex had the degree 2, and there are no double edges. We refer to G as an n-cycle if the number V of its vertices is n. For an n-cycle we have dim Ω 0 (G) dim Ω 1 (G) n and dim H 0 (G) 1. Consider two specific examples of cycles. Let us call by a triangle the following digraph 2 0 1 (2.15) (2.16) Note that the triangle contains a 2-path e 012 Ω 2 as e 012 A 2 and e 012 e 12 e 02 + e 01 A 1. Let us called by a square the following digraph 2 3 0 1 8

The square contains a 2-path v : e 013 e 023 Ω 2 because v A 2 and v (e 13 e 03 + e 01 ) (e 23 e 03 + e 02 ) e 13 + e 01 e 23 e 02 A 1, where the non-allowed path e 03 cancels out. Proposition 2.13 Let G be a cycle. Then Ω p (G) {0} for all p 3 and H p (G) {0} for all p 2. If G is a triangle or a square then dim Ω 2 (G) 1, dim H 1 (G) 0 whereas otherwise dim Ω 2 (G) 0, dim H 1 (G) 1. Proof. Let G be a triangle (2.15). For p 3 there are no allowed p-paths, in particular, Ω p {0} and H p {0}. Obviously, A 2 is spanned by a single 2-path e 012, and this path is also -invariant, so that Ω 2 span {e 012 }. Since e 012 0, we see that ker Ω2 0 and, hence, H 2 {0}. Clearly, Ω 1 span {e 01, e 02, e 12 } and it is easy to see that ker Ω1 is spanned by e 12 e 02 + e 01, which coincides with Im Ω2 ; hence, H 1 {0}. Let G be a square (2.16). As above, we obtain for p 3 that Ω p {0} and H p {0}. The space A 2 is now 2-dimensional: A 2 span {e 013, e 023 }. A 2-path v αe 013 + βe 023 A 2 has the boundary v α (e 13 e 03 + e 01 ) + β (e 23 e 03 + e 02 ) that is allowed if and only if the terms e 03 cancel out, that is, when α + β 0. Hence, Ω 2 is one-dimensional and Ω 2 span {e 013 e 023 }. As in the case of a triangle, we obtain ker Ω2 0 and H 2 {0}. Also, ker Ω1 is spanned by e 13 + e 01 e 23 e 02, which coincides with Im Ω2 ; hence, H 1 {0}. Assume first that G is neither triangle nor square. Then G contains neither triangle nor square as subgraph. Let us show that Ω p {0} for any p 2. Indeed, let v be a -invariant p-path. Consider one of the elementary paths e i0...i p that enters v with non-zero coefficients. Then we have i 0 i 1 i 2... but i 0 i 2 because otherwise i 0 i 1 i 2 would be a triangle. Note that e i0...i p contains the term e i0 î 1 i 2...i p that is not allowed. Hence, the latter term should cancel out with a similar term that comes from another elementary path e i0 i 1 i 2...i n path being also a part of v. But then we have i 0 i 1 i 2 so that the vertices i 0, i 1, i 1, i 2 form a square, which is impossible. Consequently, we have H p {0} for p 2. Finally, let us compute H 1 ker Ω1. Set n V. By the definition of a cycle, the set of vertices of G can be identified with Z n so that there is an edge between i and i + 1 for any i Z n. Denote this edge by v i, that is, v i is either e i(i+1) or e (i+1)i. Then we have v i σ i (e i+1 e i ), where σ i { 1, if i (i + 1) 1, if (i + 1) i, 9 (2.17)

For any allowed ( -invariant) 1-path v i α iv i we have v i α i σ i (e i+1 e i ) i (α i 1 σ i 1 α i σ i ) e i, which vanishes if and only if for all i α i 1 σ i 1 α i σ i. (2.18) Hence, H 1 ker Ω1 is one-dimensional. The condition (2.18) is, in particular, satisfied if α i σ i for all i, which implies that H 1 is spanned by v i σ iv i. An example of such a path is shown on Fig. 2. 0 +1 5-1 4 +1-1 +1 3 1-1 2 Figure 2: The 1-path v e 01 e 12 + e 23 + e 34 e 45 + e 50 spans H 1. 3 Join of digraphs To simplify notation, we denote the set of vertices of a digraph by the same letter as the digraph itself. In this section we always use the version (2.13) of the chain complex {Ω p } p 1 and, hence, the reduced homologies { H p } p 1. 3.1 Join of two digraphs Definition 3.1 Let X, Y be two digraphs whose sets of vertices are disjoint. Consider the digraph Z with the set of vertices X Y and with the set of edges, that consists of all the edges of X and Y, as well as of all the edges of the form x y for all x X and y Y. The digraph Z is called the join of X and Y and is denoted by X Y. An example of a join of two digraphs is shown on Fig. 3(right). The operation on the digraphs is obviously non-commutative but associative. Since X and Y are subgraphs of Z X Y, every (regular) path on X or Y can be considered as a (regular) path on Z. In particular, we can consider the operation join uv of regular paths u on X and v on Y, so that the result uv is a regular path on Z (see Fig. 3(left)). Clearly, if u R p (X) and v R q (Y ) then uv R p+q+1 (Z). It is clear from construction that an elementary path e z on Z is allowed if and only if it has the form e x e y where e x is an elementary allowed path on X and e y is that on Y. Moreover, x and y in the representation e z e x e y are uniquely defined. Proposition 3.2 Let p, q 1 and r p + q + 1. (a) If u A p (X) and v A q (Y ) then uv A r (Z). If u A p (X) and v N q (Y ) then uv N r (Z), and the same is true for u N p (X) and v A p (Y ) (where N is defined in (2.11)). 10

X u 2 1 0 X Y v 6 5 4 3 Y Figure 3: Join of two paths (left) and join of two digraphs (right) (b) If u Ω p (X) and v Ω q (Y ) then uv Ω r (Z). Moreover, the operation u, v uv extends to that for the homology classes u H p (X) and v H q (Y ) so that uv H r (Z). Proof. (a) For u e x, v e y the both claims are obvious, for general u, v they follow by linearity. (b) If u Ω p (X) and v Ω q (Y ) then by (a) the path uv is allowed. Since u and v are allowed, by (a) also ( u) v and u ( v) are allowed, whence (uv) is allowed by the product rule (2.7). It follows that uv is -invariant. If u, v are representatives of homology classes, that is, closed paths, then by (2.7) the join uv is also closed, so that uv represents a homology class of Z. We are left to verify that the class of uv depends only on the classes of u and v. For that it suffices to prove that if either u or v is exact then so is uv. Indeed, if u w then (wv) ( w) v + ( 1) p w ( v) uv so that uv is exact. One of our main results is the following theorem. Here we denote by Ω the full chain complex (2.13) and deal with its homologies H. Set also, for any p 0, Ω p : Ω p 1 so that Ω is the same chain complex as Ω but with the shifted index p. Theorem 3.3 Let X, Y be two finite digraphs and Z X Y. Then we have the following isomorphism of the chain complexes: Ω (Z) Ω (X) Ω (Y ), (3.1) which is given by the map u v uv with u Ω (X) and v Ω (Y ). In particular, for any r 1, Ω r (Z) (Ω p (X) Ω q (Y )). (3.2) Consequently, we have, for any r 0, (a Künneth formula for join). H r (Z) {p,q 1:p+qr 1} {p,q 0:p+qr 1} ( Hp (X) H q (Y )) (3.3) 11

The proof will be given in Section 6. It follows from (3.2) that Ω r (Z) has a basis {u (p) i v (q) j } i,j, (3.4) where {u (p) i {p,q 1:p+qr 1} } is a basis in Ω p (X) and {v (q) j dim Ω r (Z) } is a basis in Ω q (Y ); in particular, dim Ω p (X) dim Ω q (Y ). {p,q 1:p+qr 1} In the same way one expresses a basis in H r (Z) via the basis in H p (X) and H q (Y ). Example 3.4 Consider the digraph Z X Y as on Fig. 3(right). In this case we have by Proposition 2.13 that all the homology groups H p (X) and H q (Y ) are trivial except for H 1 (X) span {e 01 + e 12 + e 20 }, H 1 (Y ) span {e 35 e 65 + e 64 e 34 }. It follows from (3.3) that all Hr (Z) are trivial except for H 3 (Z), and the latter is generated by a single element e 0135 e 0165 + e 0164 e 0134 + e 1235 e 1265 + e 1264 e 1234 + e 2035 e 2065 + e 2064 e 2034 that is the join of the generators of H 1 (X) and H 1 (Y ). Given a digraph X, the digraph Cone X is obtained from X by adding one more vertex a and all the edges of the form x a for all x X. The vertex a is called the cone vertex. Clearly, we have Cone X X Y where the digraph Y consists of a single vertex a. Observe that Ω 1 (Y ) K, Ω 0 (Y ) span {e a }, Ω q (Y ) {0} for q 1 and H q (Y ) {0} for all q 1. Hence, applying Theorem 3.3 with Y {a}, we obtain the following statement. Proposition 3.5 For any digraph X and for any r 0, we have Ω r (Cone X) Ω r (X) Ω r 1 (X), (3.5) where the isomorphism is given by the map u, v u + ve a, where u Ω r (X), v Ω r 1 (X) and a is the cone vertex. Furthermore, all the reduced homologies H r (Cone X) are trivial. Example 3.6 Let us define for any n 0 a simplex-digraph Sm n as follows: its set of vertices is {0, 1,..., n} and the edges are i j for all i < j. For example, we have 2 Sm 1 0 1, Sm 2 0 1, and Sm 3 is shown on Fig. 4. Clearly, we have Sm n Cone Sm n 1. Since Ω 0 (Sm 0 ) span {e 0 } and Ω n (Sm n 1 ) {0}, we obtain by induction from (3.5) that Ω n (Sm n ) span {e 01...n }. Of course, all the reduced homologies of Sm n are trivial. A suspension over a digraph X is a digraph Sus X that is obtained from X by adding two vertices a, b and all the edges x a and x b for all x X. The vertices a, b are called the suspension vertices. Clearly, we have Sus X X Y where Y is a digraph that consists of two vertices a, b and no edges. Observe that Ω 1 (Y ) K, Ω 0 (Y ) span {e a, e b }, Ω q (Y ) {0} for q 1, H0 (Y ) span (e b e a ) and H q (Y ) {0} for q 1. Applying Theorem 3.3 with Y {a, b}, we obtain the following statement. 12

2 3 0 1 Figure 4: A simplex-digraph Sm 3 Proposition 3.7 For any digraph X and for any r 0, we have Ω r (Sus X) Ω r (X) Ω r 1 (X) Ω r 1 (X), (3.6) and the isomorphism is given by the map u, v, w u + ve a + we b, where u Ω r (X), v, w Ω r 1 (X) and a, b are the suspension vertices. Furthermore, we have H r (Sus X) H r 1 (X), (3.7) and the isomorphism is given by the map u u (e a e b ), where u H r 1 (X). Example 3.8 Let S be any cycle that is neither triangle nor square, so that by Proposition 2.13 dim H 1 (S) 1. We regards S as an analog of a circle. Define S n inductively by S 1 S and S n+1 Sus S n, so that S n can be regarded as an analog of the n-dimensional sphere. Proposition 3.7 implies by induction that dim H n (S n ) dim H 1 (S) 1, which gives an example of a nontrivial H n with an arbitrary n. In the same way one shows that H p (S n ) {0} for all p 1, p n. Let v be an 1-path on S that spans H 1 (S) (see the proof of Proposition 2.13). Denoting by a n, b n the suspension vertices of S n+1 Sus S n, we obtain by induction that H n (S n ) is generated by the path v (e a1 e b1 ) (e a2 e b2 )... ( e an 1 e bn 1 ). For example, the digraph G on Fig. 1 is an S 2 based on 3-cycle S with the vertices 1, 2, 3. Since by Proposition 2.13 v e 12 + e 23 + e 31, we obtain that H 2 (G) is generated by v (e 4 e 5 ) (e 124 + e 234 + e 314 ) (e 125 + e 235 + e 315 ). Another example of an S 2 is shown on Fig. 5 in two ways. 0 4 3 1 5 1 2 3 0 4 2 5 Figure 5: An octahedron digraph Indeed, denoting by S the 4-cycle with vertices {0, 1, 2, 3}, we see that the digraph G on Fig. 5 is Sus S S 2. Hence, we obtain that H 2 (G) is generated by e 024 e 025 e 034 + e 035 e 124 + e 125 + e 134 e 135. 13

3.2 Some auxiliary results Given a digraph G, set E (G) p 1 E p (G), that is, E (G) is the set of all allowed elementary p-paths on G with p 1. Denote by A (G) the union of all A p (G) with p 1 and define in A (G) the K-scalar product as follows: for all u, v A (G) [u, v] : x E (X) u x v x. In particular, if u A p and v A p with p p then clearly [u, v] 0. As before, let Z X Y for two digraphs X, Y. Let us prove some simple properties of join of paths that we will need later for the proof of Theorem 3.3. Lemma 3.9 Any path w A (Z) admits a representation where c xy K. w x E (X),y E (Y ) c xy e x e y (3.8) Proof. By construction of join, any allowed elementary path on Z has the form e x e y, where x E (X) and y E (Y ). Hence, any path w A (Z) has the form (3.8). Lemma 3.10 If u A p (X), ϕ A p (X) and v A q (Y ), ψ A q (Y ) with p, q, p, q 1 then [uv, ϕψ] [u, ϕ] [v, ψ]. (3.9) Proof. Set r p + q + 1. If p + q + 1 r then then both sides of (4.22) vanish. Assuming p + q + 1 r we obtain [uv, ϕψ] (uv) z (ϕψ) z z E r(z) x E p(x),y Eq(Y ) u x v y ϕ x ψ y. If p p then ϕ x 0 and then both sides of (4.22) vanish. If p p and, hence, q q, then we obtain [uv, ϕψ] u x ϕ x v y ψ y [u, ϕ] [v, ψ], which finishes the proof. 4 Cartesian product x E p(x) y E q(y ) In this section we use the truncated chain complexes {R p } p 0 and {Ω p } p 0 (cf. (2.14)) and the homologies {H p } p 0 of the latter. (2.10) and 14

4.1 Step-like paths Given two finite sets X, Y, consider their Cartesian product Z X Y. Fix r 0 and let z z 0 z 1...z r be a regular elementary r-path on Z, where z k (x k, y k ) with x k X and y k Y. We say that the path z is step-like if, for any k 1,..., r, either x k 1 x k or y k 1 y k. In fact, exactly one of these conditions holds as z is regular. Any step-like path z on Z determines regular elementary paths x on X and y on Y by projection. More precisely, x is obtained from z by taking the sequence of all X-components of the vertices of z and then by collapsing in it any subsequence of repeated vertices to one vertex. The same rule applies to y. By construction, the projections x and y are regular elementary paths on X and Y, respectively. If the projections of z are x x 0...x p and y y 0...y q then p + q r (cf. Fig. 6(left)). y q z r (x p,y q ) (0,q) S(z) (p,q) y j z k(x i,y j) (i,j) L(z) z 0 (x 0,y 0 ) x i x p (0,0) (p,0) Figure 6: Left: a step-like path z and its projections x and y. Right: a staircase S (z) and its elevation L (z) (here L (z) 30). Every vertex (x i, y j ) of a step-like path z can be represented as a point (i, j) of Z 2 so that the whole path z is represented by a staircase S (z) in Z 2 connecting the points (0, 0) and (p, q). Definition 4.1 Define the elevation L (z) of the path z as the number of the cells in Z 2 + below the staircase S (z) (cf. the shaded area on Fig. 6(right)). By definition, any p-path u on X is given by u x u x e x where the summation is taken over all elementary p-paths x on X and u x K are the components of u. It will be convenient to extend the summation here to all elementary paths x with arbitrary length, by setting u x 0 if the length of x is not equal to p. Definition 4.2 For any paths u R p (X) and v R q (Y ) with p, q 0 define their cross product u v as a path on Z by the following rule: for any step-like elementary path z on Z, the component (u v) z is defined by (u v) z ( 1) L(z) u x v y, (4.1) where x and y are the projections of z onto X and Y, while for the rest paths z set (u v) z 0. Hence, we have u v R p+q (Z). 15

In the context of this section we use the truncated regular chain complex (2.10), that is, the convention that R 1 {0}. Let us extend the definition of the cross product u v to the case when either p 1 and q 0 or p 0 and q 1 simply by setting u v 0 R p+q. For any regular elementary p-path x on X and q-path y on Y with p, q 0 denote by Π x,y the set of all step-like paths z on Z whose projections on X and Y are x and y respectively. Clearly, we have Π x,y ( ) p+q p. It follows from (4.1) that, for all regular elementary paths x, y, e x e y z (e x e y ) z e z z ( 1) L(z) (e x ) x (e y ) y e z, where x and y are projections of z, whence e x e y ( 1) L(z) e z. (4.2) z Π x,y It is not difficult to see that cross product is associative. Example 4.3 Let us denote the vertices of X by letters a, b, c etc. and the vertices of Y by integers 0, 1, 2, etc. Then the vertices of Z X Y will be denoted as chessboard fields, for example, a0, b1 etc. Here are some examples of cross products: 1. e a e 01 e a0a1, e ab e 0 e a0b0 2. e ab e 01 e a0b0b1 e a0a1b1 3. e abc e 01 e a0b0c0c1 e a0b0b1c1 + e a0a1b1c1 4. e abc e 012 e a0b0c0c1c2 e a0b0b1c1c2 + e a0b0b1b2c2 + e a0a1b1c1c2 e a0a1b1b2c2 + e a0a1a2b2c2 (cf. Fig. 7). a2 b2 c2 a1 b1 c1 a0 b0 c0 Figure 7: The staircase a0b0b1c1c2 has elevation 1. Hence, e a0b0b1c1c2 enters the product e abc e 012 with the negative sign. Proposition 4.4 (Product rule for cross product) If u R p (X) and v R q (Y ) where p, q 0, then (u v) ( u) v + ( 1) p u ( v). (4.3) Proof. It suffices to prove (4.3) for the case u e x and v e y where x x 0...x p and y y 0...y q are regular elementary p-path on X and q-path on Y, respectively. Set r p + q so that e x e y R r (Z). If p q 0 then all the terms in (4.3) vanish. Assume p 0 and q 1. Then Π x,y contains the only element z z 0...z q where z i (x 0, y i ). Since L (z) 0, we obtain by (4.2) that e x e y e z0...z q 16

By (2.2) obtain (e x e y ) e z0...z q e x e y0...y q, which is equivalent to (4.3), because u 0. In the same way (4.3) is proved if q 0 and p 1. Consider now the main case p, q 1. We have by (4.2) and (2.2) (e x e y ) where we use a shortcut z Π x,y ( 1) L(z) e z z Π x,y k0 z (k) z 0...ẑ k...z r z 0...z k 1 z k+1...z r. Switching the order of the sums, rewrite (4.4) in the form (e x e y ) r k0 r ( 1) L(z)+k e z(k), (4.4) z Π x,y ( 1) L(z)+k e z(k). (4.5) Given an index k 0,..., r and a path z Π x,y, consider the following four logically possible cases how horizontal and vertical couples combine around z k : (H) : z k 1 z k z k+1 (V ) : z k+1 z k z k 1 (R) : z k+1 (L) : z k 1 z k z k z k+1 Here (H) stands for a horizontal position, (V ) for vertical, (R) for right and (L) for left. If k 0 or k r then z k 1 or z k+1 should be ignored, so that one has only two distinct positions (H) and (V ). If z Π x,y and z k stands in (R) or (L) then consider a path z Π x,y such that z i z i for all i k, whereas z k stands in the opposite position (L) or (R), respectively, as on the diagrams: z k 1 z k z k+1 z k 1 z k z k z k+1 z k 1 z k Clearly, we have L (z ) L (z) ± 1 which implies that the terms e z(k) and e z in (4.5) cancel (k) out. Denote by Π k x,y the set of paths z Π x.y such that z k stands in position (V ) and by Πx,y k the set of paths z Π x,y such that z k stands in position (H). By the above observation, we can restrict the summation in (4.5) to those pairs k, z where z k is either in vertical or horizontal position, that is, r (e x e y ) ( 1) L(z)+k e z(k). (4.6) k0 z Π k x,y Π x,y k 17

Let us now compute the first term in the right hand side of (4.3): ( e x ) e y p ( 1) l e x e y l0 p ( 1) L(w)+l e w. (4.7) l0 w Π x(l),y Fix some l 0,..., p and w Π x(l),y. Since the projection of w on X is x (l) x 0...x l 1 x l+1...x p, there exists a unique index k such that w k 1 projects onto x l 1 and w k projects onto x l+1. Then w k 1 and w k have a common projection onto Y, say y m (cf. Fig. 8). y q y m z k (x l,y m ) w k-1 (x l-1,y m ) w k (x l+1,y m ) y 0 x 0 x l-1 x l x l+1 x p Figure 8: Step-like paths w and z. The shaded area represents the difference L (z) L (w). Define a path z Π k x,y by setting w i for i k 1, z i (x l, y m ) for i k, w i 1 for i k + 1. (4.8) By construction we have z (k) w. It also follows from the construction that Since k l + m, we obtain that L (z) L (w) + m. L (z) + k L (w) + l + 2m. We see that each pair l, w where l 0,..., p and w Π x(l),y gives rise to a pair k, z where k 0,..., r, z Π k x,y, and ( 1) L(z)+k e z(k) ( 1) L(w)+l e w. By reversing this argument, we obtain that each such pair k, z gives back l, w so that this correspondence between k, z and l, w is bijective. Hence, we conclude that ( e x ) e y p r ( 1) L(w)+l e w ( 1) L(z)+k e z(k). (4.9) l0 w Π x(l),y k0 z Πx,y k 18

The second term in the right hand side of (4.3) is computed similarly: ( 1) p e x e y q ( 1) m+p e x e y(m) m0 q m0 w Π x,y(m) ( 1) L(w)+m+p e w. Each pair m, w here gives rise to a pair k, z where k 0,..., r and z Π k x,y in the following way: choose k such that w k 1 projects onto y m 1 and w k projects onto y m+1. Then w k 1 and w k have a common projection onto X, say x l. y q y m+1 w k (x l,y m+1 ) y m y m-1 z k (x l,y m ) w k-1 (x l,y m-1 ) y 0 x 0 x l x p Figure 9: Paths w and z. The shaded area represents L (z) L (w). Define the path z Π k x,y as in (4.8) (cf. Fig. 9). Then we have w z (k) and Since k l + m, we obtain and ( 1) p e x e y q m0 L (z) L (w) + p l. L (z) + k L (w) + p + m w Π x,y(m) ( 1) L(w)+m+p e w Combining this with (4.6) and (4.9), we obtain (4.3). 4.2 Cartesian product of digraphs r k0 z Π k x,y ( 1) L(z)+k e z(k). Definition 4.5 Given two digraphs X and Y, consider the digraph Z with the set of vertices X Y and with the set of edges defined by the following rule: for x, x X and y, y Y, (x, y) ( x, y ) if either x x and y y or x x and y y. The digraph Z is called the Cartesian product of X and Y and is denoted by X Y. 19

A fragment of the graph Z is shown here: y (x,y... ) (x,y )... (x,y) y... (x,y)... Y X... x x... It is not difficult to see that the Cartesian product of digraphs is associative. Clearly, any regular elementary path on Z X Y is allowed if and only if it is step-like and its projections onto X and Y are allowed. Proposition 4.6 Let p, q 0 and r p + q. (a) If u A p (X) and v A q (Y ) then u v A r (Z). If u N p (X) and v A q (Y ) then u v N r (Z), and the same is true if u A p (X) and v N q (Y ). (b) If u Ω p (X) and v Ω q (Y ) then u v Ω r (Z). Moreover, the operation u, v u v extends to that for the homology classes u H p (X) and v H q (Y ) so that u v H r (Z). Proof. (a) It suffices to prove the both claims for u e x and v e y. By (4.2) e x e y is a linear combination of e z with z Π x,y. If x and y are allowed then any z Π x,y is allowed, which implies that e x e y A r (Z). Let x be non-allowed. Since the projection of z Π x,y onto X is x, it follows that z is nonallowed. Hence, e x e y is a linear combination of non-allowed paths, that is e x e y N r (Z). (b) The proof is based on the product rule (4.3) and goes the same way as the proof of Proposition 3.2(b). The next theorem gives a complete description of -invariant paths on Z. We denote by Ω {Ω p } p 0 the truncated chain complex (2.14) and by H {H p } p 0 its homologies. Theorem 4.7 Let X, Y be two finite digraphs and Z X Y. Then we have the following isomorphism of the chain complexes: Ω (Z) Ω (X) Ω (Y ), (4.10) which is given by the map u v u v with u Ω (X) and v Ω (Y ). In particular, for any r 0 Ω r (Z) (Ω p (X) Ω q (Y )). (4.11) Consequently, we have that is, for any r 0, H r (Z) (a Künneth formula for product). {p,q 0:p+qr} H (Z) H (X) H (Y ), (4.12) {p,q 0:p+qr} (H p (X) H q (Y )) (4.13) Example 4.8 Consider the digraph Z X Y (shown on Fig. 10), where X b a c and Y 20 2 3. 0 1

b2 b3 c2 c3 a2 a3 b0 b1 c0 c1 a0 a1 Figure 10: Cartesian product of a triangle and a square For r 4 we obtain from (4.11) that Ω 4 (Z) Ω 2 (X) Ω 2 (Y ) because on both digraphs X, Y we have Ω p {0} for p 3. By the proof of Proposition 2.13, Ω 2 (X) span (e abc ) and Ω 2 (Y ) span (e 013 e 023 ), whence it follows that Ω 4 (Z) is spanned by a single 4-path e abc (e 013 e 023 ) e a0b0c0c1c3 e a0b0b1c1c3 + e a0b0b1b3c3 +e a0a1b1c1c3 e a0a1b1b3c3 + e a0a1a3b3c3 e a0b0c0c2c3 + e a0b0b2c2c3 e a0b0b2b3c3 e a0a2b2c2c3 + e a0a2b2b3c3 e a0a2a3b3c3. Similarly one can compute Ω r (Z) for other values of r. For example, Ω 3 (Z) Ω 1 (X) Ω 2 (Y ) Ω 2 (X) Ω 1 (Y ), which implies dim Ω 3 (Z) 3 1 + 1 4 7 and the generators of Ω 3 (Z) are e ab (e 013 e 023 ), e ac (e 013 e 023 ), e bc (e 013 e 023 ) e abc e 01, e abc e 13, e abc e 02, e abc e 23 Since all the homology groups of X, Y are trivial except for H 0, we obtain that the same is true for homologies of Z. Example 4.9 Consider Z X Y where X, Y are cyclic digraphs: X b a c, Y 1 2. 0 3 By Proposition 2.13 all the homologies H p (X) and H q (Y ) are trivial for p, q 2 whereas It follows from (4.12) that H 2 (Z) H 1 (X) span (e ab + e bc + e ca ) H 1 (Y ) span (e 01 + e 12 + e 23 + e 30 ). {p,q 0:p+q2} (H p (X) H q (Y )) H 1 (X) H 1 (Y ), 21

in particular, dim H 2 (Z) 1. The generating element of H 2 (Z) is (e ab + e bc + e ca ) (e 01 + e 12 + e 23 + e 30 ). For any digraph X, define the cylinder over X by Cyl X : X Y with Y (0 1). Assuming that the vertices of X are enumerated by 0, 1,..., n 1, let us enumerate the vertices of Cyl X by 0, 1,..., 2n 1 using the following rule: (x, 0) is assigned the number x, while (x, 1) is assigned x + n. Define the operation of lifting paths from X to Cyl X as follows: for any regular path v on X, the lifted path v is defined by v v e 01. For example, if v e i0...i p then v e i0...i p e 01 ( 1) p Since e 01 Ω 1 (Y ), we see that if v Ω p (X) then v Ω p+1 (Cyl X). p ( 1) k e i0...i k (i k +n)...(i p+n). (4.14) Example 4.10 Let us define n-cube Cube n inductively as follows: Cube 0 {0} and k0 Cube n Cyl Cube n 1. For example, Cube 1 is Cube 2 is a square 0 1 2 3 0 1 and Cube 3 is shown on Fig. 11. 6 7 4 5 2 3 0 1 Figure 11: A 3-cube Since Cube n Cube n 1 Y, where Ω q (Y ) is non-trivial only for q 0, 1, and Ω n (Cube n 1 ) {0}, we obtain from (4.11) Ω n (Cube n ) Ω n 1 (Cube n 1 ) Ω 1 (Y ). Since Ω 1 (Y ) is generated by a single element v 1 e 01, we obtain by induction that dim Ω n (Cube n ) 1. A generating element v n of Ω n (Cube n ) can be computed inductively by v n v n 1 e 01 v n 1. 22

By (4.14) we obtain successively v 2 v 1 e 013 e 023, v 3 v 2 e 0457 e 0157 + e 0137 e 0467 + e 0267 e 0237, etc. In general, v n is an alternating sum of n! elementary paths that correspond to partitioning of a solid n-cube into n! simplexes. By (4.13) all homology groups of Cube n are trivial except for H 0. 4.3 Some auxiliary results For any digraph G denote by E (G) the union of all E p (G) with p 0, where E p (G) is the set of all allowed elementary p-paths. Note that the notation E here is different from that in Section 3.2 where p 1 was also allowed. Denote by A (G) the union of all A p (G) with p 0 and by Ω (G) the similar union of all Ω p (G) with p 0. We will use the K-scalar product in A (G) defined by for all u, v A (G) by [u, v] : u x v x. x E (X) As in Section 4.2, we work with two digraphs X, Y and their Cartesian product Z X Y. Here we prove some auxiliary results. Lemma 4.11 The family of paths {e x e y } where x E (X) and y E (Y ) is linearly independent. Proof. Set for some c xy K w x E (X),y E (Y ) c xy e x e y and prove that w 0 implies that c xy 0 for all couple x, y as in the summation. Fix such a couple x, y and choose one z Π x,y. By (4.1) we have { (e x e y ) z ( 1) L(z), if x x and y y, 0, otherwise, which implies that w z ( 1) L(z) c x y. Hence, w 0 c x y 0, which was to be proved. Lemma 4.12 Any path w Ω (Z) admits a representation w x E (X), y E (Y ) c xy (e x e y ) (4.15) with some coefficients c xy K (note that c xy are uniquely defined by Lemma 4.11). Proof. Fix w Ω r (Z) for some r 0. For any x E (X) and y E (Y ) choose some z Π x,y and set c xy ( 1) L(z) w z. (4.16) Let us first show that the value of c xy in (4.16) is independent of the choice of z Π x,y. Set z i 0...i r. Let k be an index such that one of the couples i k 1 i k, i k i k+1 is vertical and the other 23

is horizontal. If i k 1 (a, b) and i k+1 (a, b ) where a, a X and b, b Y, then i k is either (a, b) or (a, b ). Denote the other of these two vertices by i k, as, for example, on the diagram:. b i k i k+1... b... i k 1 i k. y. x... a a.... Replacing in the path z i 0...i r the vertex i k by i k, we obtain the path z i 0...i k 1 i k i k+1...i r that clearly belongs to Π x,y and, hence, is allowed. Since the (r 1)-path i 0...i k 1 i k+1...i r is regular but non-allowed (as it is not step-like), while w is allowed, we have On the other hand, we have by (2.4) ( w) i 0...i k 1 i k+1...i r ( w) i 0...i k 1 i k+1...i r 0. (4.17) j Z ( k 1 m0 ( 1) m w i 0...i m 1 ji m...i k 1 i k+1...i r (4.18) + ( 1) k w i 0...i k 1 ji k+1...i r (4.19) ) r+1 + ( 1) m 1 w i 0...i k 1 i k+1...i m 1 ji m...i r. (4.20) mk+2 All the components of w in the sums (4.18) and (4.20) vanish since they correspond to nonallowed paths, while w is allowed. The path i 0...i k 1 ji k+1...i r in the term (4.19) is also nonallowed unless j i k or j i k (note that i k and i k are uniquely determined by i k 1 and i k+1 ). Hence, the only non-zero terms in (4.18)-(4.20) are w i 0...i k 1 i k i k+1...i r w z and w i 0...i k 1 i k i k+1...i r w z. Combining (4.17) and (4.18)-(4.20), we obtain Since L (z ) L (z) ± 1, it follows that 0 w z + w z. ( 1) L(z ) w z ( 1) L(z) w z. (4.21) The transformation z z described above, allows us to obtain from a given z Π x,y in a finite number of steps any other path in Π x,y. Since the quantity ( 1) L(z) w z does not change under this transformation, it follows that it does not depend on a particular choice of z Π x,y, which was claimed. Hence, the coefficients c xy are well-defined by (4.16). Finally, let us show that the identity (4.15) holds with the coefficients c xy from (4.16). By (4.2) we have e x e y ( 1) L(z) e z. z Π x,y 24

Using (4.16) we obtain x E (X), y E (Y ) c xy (e x e y ) x E (X), y E (Y ) x E (X), y E (Y ) z E (Z) w z e z w, c xy ( 1) L(z) e z z Π x,y w z e z z Π x,y which finishes the proof. Lemma 4.13 If u A p (X), ϕ A p (X) and v A q (Y ), ψ A q (Y ), where p, q 0, then [u v, ϕ ψ] ( ) p+q p [u, ϕ] [v, ψ]. (4.22) Proof. Set r p + q. If p + q r then the both sides of (4.22) vanish. Assume further that p + q r so that w : ϕ ψ A r (Z). We have then [u v, w] (u v) z w z z E r(z) z E r(z) ( 1) L(z) u x v y w z (x, y are projections of z) x E p(x) y E q(y ) By definition of cross product, we have z Π x,y ( 1) L(z) u x v y w z. (4.23) w z ( 1) L(z) ϕ x ψ y, where x and y are the projections of z. In (4.23) we have x E p (X) and y E q (Y ). Therefore, if p p then ϕ x 0, w z 0 and the sum in (4.23) vanishes. In this case, the both sides of (4.22) are equal to zero again. In the main case p p and, hence, q q, we obtain [u v, ϕ ψ] x E (X) y E (Y ) x E (X) y E (Y ) ( p+q p where we have used Π x,y ( ) p+q p. ) [u, ϕ] [v, ψ] z Π x,y ( 1) L(z) u x v y ( 1) L(z) ϕ x ψ y Π x,y u x ϕ x v y ψ y 5 -Invariant elements on abstract products The main part of the proof of Theorems 3.3 and 4.7 is contained in Theorem 5.1 below. This theorem will be stated in terms of an abstract chain complex. Fix some integer f and let {R p } p f be a sequence of finite dimensional K-linear spaces, that together with a boundary operator : R p R p 1 forms a chain complex 0 R f R f+1... R p 1 R p... 25

Fix a basis {e x } in R p where the index x varies in a finite set R p. For any u R p we have then a unique expansion u x R p u x e x where u x K. We extend the notation u x to any x R p f R p by setting u x 0 if x R q with q p. Define in R p f R p a K-scalar product by [u, v] x R u x v x. Fix a subset E p of R p and denote by A p the subspace of R p spanned by {e x } with x E p. The elements of A p f A p are called allowed. Define also N p span {e x : x R p \ E p } so that R p A p N p. The elements of N are called non-allowed. Define for any p f a subspace Ω p of A p by Ω p {u A p : u A p 1 }. Obviously, we have Ω p Ω p 1 so that {Ω p } p f is a chain complex. The elements of Ω p are called -invariant. Of course, the above definitions represent abstract version of the construction of -invariant paths. If we start with a digraph G then R p is the set of regular elementary p-paths on G and E p the set of allowed elementary p-paths. The value of f is a flag that distinguishes the full chain complex (2.13) from the truncated one (2.14). Therefore, f 1 in Theorem 3.3 and f 0 in Theorem 4.7. Assume now that we have three sets of the structure {R, A, Ω }, which we will distinguish by adding to these notations (X), (Y ), (Z). Of course, in application X, Y, Z will be digraphs where Z X Y or Z X Y, but abstractly X, Y, Z are nothing else but indices to distinguish the three structures. This convention makes the notation in the abstract setting identical to those for the digraph setting. Therefore, the reader who is interested only in digraphs may safely assume that X, Y, Z are digraphs. Assume that there exists a K-bilinear operation u R (X), v R (Y ) u v R (Z) with the following properties (everywhere p, q f). 1. If u R p (X) and v R q (Y ) then u v R r (Z) with r p + q f. 2. If u A (X) and v A (Y ) then u v A (Z). 3. If u A (X) and v N (Y ) then u v N (Z); the same is true if u N (X) and v A (Y ). 4. The sequence {e x e y } x E (X),y E (Y ) is linearly independent in A (Z). 5. If u R p (X) and v R q (Y ) then (u v) α u v + βu v, (5.1) where α and β are non-zero scalars depending on p, q. In the case p f we understand u v as zero, and the same convention applies to u v in the case q f. 26