On some matrices related to a tree with attached graphs

Similar documents
Linear estimation in models based on a graph

Determinant of the distance matrix of a tree with matrix weights

On minors of the compound matrix of a Laplacian

DETERMINANTAL DIVISOR RANK OF AN INTEGRAL MATRIX

Determinantal divisor rank of an integral matrix

PRODUCT DISTANCE MATRIX OF A GRAPH AND SQUARED DISTANCE MATRIX OF A TREE. R. B. Bapat and S. Sivasubramanian

On Euclidean distance matrices

~ g-inverses are indeed an integral part of linear algebra and should be treated as such even at an elementary level.

Identities for minors of the Laplacian, resistance and distance matrices

On the adjacency matrix of a block graph

Generalized Principal Pivot Transform

Resistance distance in wheels and fans

Elementary maths for GMT

RESISTANCE DISTANCE IN WHEELS AND FANS

CLASSIFICATION OF TREES EACH OF WHOSE ASSOCIATED ACYCLIC MATRICES WITH DISTINCT DIAGONAL ENTRIES HAS DISTINCT EIGENVALUES

For other titles published in this series, go to Universitext

Product distance matrix of a tree with matrix weights

The effect on the algebraic connectivity of a tree by grafting or collapsing of edges

GENERALIZED POWER SYMMETRIC STOCHASTIC MATRICES

ELA

MOORE-PENROSE INVERSE IN AN INDEFINITE INNER PRODUCT SPACE

MATH 2331 Linear Algebra. Section 2.1 Matrix Operations. Definition: A : m n, B : n p. Example: Compute AB, if possible.

Group inverse for the block matrix with two identical subblocks over skew fields

1 Counting spanning trees: A determinantal formula

The Distance Spectrum

Linear Algebra and its Applications

MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS. + + x 1 x 2. x n 8 (4) 3 4 2

The Singular Acyclic Matrices of Even Order with a P-Set of Maximum Size

n n matrices The system of m linear equations in n variables x 1, x 2,..., x n can be written as a matrix equation by Ax = b, or in full

Linear graph theory. Basic definitions of linear graphs

Week 4. (1) 0 f ij u ij.

Predrag S. Stanimirović and Dragan S. Djordjević

On Euclidean distance matrices

Review of Linear Algebra

The product distance matrix of a tree and a bivariate zeta function of a graphs

The Laplacian spectrum of a mixed graph

SPECIAL FORMS OF GENERALIZED INVERSES OF ROW BLOCK MATRICES YONGGE TIAN

Math Camp Lecture 4: Linear Algebra. Xiao Yu Wang. Aug 2010 MIT. Xiao Yu Wang (MIT) Math Camp /10 1 / 88

Laplacian Integral Graphs with Maximum Degree 3

Block distance matrices

Dragan S. Djordjević. 1. Introduction and preliminaries

Sang Gu Lee and Jeong Mo Yang

Matrix operations Linear Algebra with Computer Science Application

Laplacian eigenvalues and optimality: II. The Laplacian of a graph. R. A. Bailey and Peter Cameron

A lower bound for the Laplacian eigenvalues of a graph proof of a conjecture by Guo

SAMPLE OF THE STUDY MATERIAL PART OF CHAPTER 1 Introduction to Linear Algebra

On Hadamard Diagonalizable Graphs

Matrices Gaussian elimination Determinants. Graphics 2009/2010, period 1. Lecture 4: matrices

Matrix Algebra Determinant, Inverse matrix. Matrices. A. Fabretti. Mathematics 2 A.Y. 2015/2016. A. Fabretti Matrices

A revisit to a reverse-order law for generalized inverses of a matrix product and its variations

MATH 213 Linear Algebra and ODEs Spring 2015 Study Sheet for Midterm Exam. Topics

Combinatorial and physical content of Kirchhoff polynomials

Chapter 4 - MATRIX ALGEBRA. ... a 2j... a 2n. a i1 a i2... a ij... a in

x + 2y + 3z = 8 x + 3y = 7 x + 2z = 3

ECON 186 Class Notes: Linear Algebra

Generic degree structure of the minimal polynomial nullspace basis: a block Toeplitz matrix approach

SAMPLE OF THE STUDY MATERIAL PART OF CHAPTER 1 Introduction to Linear Algebra

FORBIDDEN MINORS FOR THE CLASS OF GRAPHS G WITH ξ(g) 2. July 25, 2006

GEOMETRY OF MATRICES x 1

On the construction of asymmetric orthogonal arrays

7 Matrix Operations. 7.0 Matrix Multiplication + 3 = 3 = 4

Dragan S. Djordjević. 1. Introduction

ORIE 6334 Spectral Graph Theory September 8, Lecture 6. In order to do the first proof, we need to use the following fact.

Linear algebra and applications to graphs Part 1

Topics in Graph Theory

The eigenvalues of an n x n complex matrix A are the roots of the n th degree polynomial det(a-λi) = 0. Further the positive square roots of the

Energy, Laplacian Energy and Zagreb Index of Line Graph, Middle Graph and Total Graph

Chapter 2 Notes, Linear Algebra 5e Lay

STRAND J: TRANSFORMATIONS, VECTORS and MATRICES

Root systems and optimal block designs

Combinatorial Batch Codes and Transversal Matroids

Linear Algebra and its Applications

Inverse Perron values and connectivity of a uniform hypergraph

Equality: Two matrices A and B are equal, i.e., A = B if A and B have the same order and the entries of A and B are the same.

(1) for all (2) for all and all

Linear Equations in Linear Algebra

Chapter 7. Linear Algebra: Matrices, Vectors,

The Matrix-Tree Theorem

Graphs and Their Applications (7)

Things we can already do with matrices. Unit II - Matrix arithmetic. Defining the matrix product. Things that fail in matrix arithmetic

Math Camp Notes: Linear Algebra I

Operators with Compatible Ranges

QUIVERS AND LATTICES.

APPROXIMATION OF MOORE-PENROSE INVERSE OF A CLOSED OPERATOR BY A SEQUENCE OF FINITE RANK OUTER INVERSES

Lectures on Linear Algebra for IT

ELEMENTARY LINEAR ALGEBRA

Relation of Pure Minimum Cost Flow Model to Linear Programming

a Λ q 1. Introduction

Algorithm to Compute Minimal Nullspace Basis of a Polynomial Matrix

Product of Range Symmetric Block Matrices in Minkowski Space

Mathematical Programs Linear Program (LP)

Lecture 5 January 16, 2013

Numerical Analysis Lecture Notes

Row Space, Column Space, and Nullspace

(a) If A is a 3 by 4 matrix, what does this tell us about its nullspace? Solution: dim N(A) 1, since rank(a) 3. Ax =

THE SPECTRUM OF THE LAPLACIAN MATRIX OF A BALANCED 2 p -ARY TREE

Discrete Optimization 23

Digital Workbook for GRA 6035 Mathematics

Review of Vectors and Matrices

Acyclic Digraphs arising from Complete Intersections

Transcription:

On some matrices related to a tree with attached graphs R. B. Bapat Indian Statistical Institute New Delhi, 110016, India fax: 91-11-26856779, e-mail: rbb@isid.ac.in Abstract A tree with attached graphs is a tree, together with graphs defined on its partite sets. In an earlier paper, we introduced the notion of the incidence matrix Q and the Laplacian L = QQ for a tree with attached graphs. Here we consider the case when the attached graphs are trees. We obtain formulas for the Moore-Penrose inverse of Q and for the inverse of K = Q Q, thereby extending some work in the literature, including a formula by Merris proved in the case of a tree. The case when the attached graphs are paths is related to the basic feasible solutions in a transportation problem. Key words and phrases: tree, incidence matrix, Moore-Penrose inverse, transportation problem, edge Laplacian matrix AMS Classification Number: Primary: 05C50; Secondary: 15A09, 15A15 1

1 Introduction Minors of matrices associated with a graph have been an area of considerable interest, starting with the celebrated Matrix Tree Theorem of Kirchhoff, which asserts that any cofactor of the Laplacian matrix equals the number of spanning trees in the graph. Several papers have been devoted to the theme of Matrix Tree type theorems; see [3, 4, 8, 12] for more information and further references. Combinatorial interpretation of minors is closely related to interpreting the inverse, or a generalized inverse, of the matrix. As an example, the Moore- Penrose inverse of the Laplacian matrix has been investigated in [1, 10]. In a recent paper [2] we introduced the concept of a tree with attached graphs, which is simply a tree with graphs defined on its partite sets. The notions of incidence matrix, Laplacian and distance matrix were defined for a tree with attached graphs. In the present paper we consider the case when the attached graphs are trees. Formulas are obtained for the Moore-Penrose inverse of the incidence matrix Q and for the inverse of K = Q Q. When one of the partite sets has only one vertex, these results reduce to known formulae [1, 11] for the case of a tree. Another special case of a tree with attached graphs that is of interest is when the attached graphs are paths. It is related to basic feasible solutions in a transportation problem. This is illustrated in Section 5. 2 Preliminaries We consider simple graphs, that is, graphs that have no loops or parallel edges. The vertex set and the edge set of the graph G are denoted by V (G) and E(G) respectively. The transpose of a matrix A is denoted A. If G is a directed graph with n vertices and m edges, then its incidence matrix Q is the n m matrix defined as follows. The rows and the columns of Q are indexed by V (G) and E(G) respectively. If i V (G) and j E(G), then the (i, j)-entry of Q is 0 if vertex i and edge j are not incident and otherwise it is 1 or 1 according 2

as j originates or terminates at i respectively. The Laplacian matrix L of G is defined as L = QQ. The Laplacian does not depend on the orientation and thus is defined for an undirected graph. We assume familiarity with basic properties of the incidence matrix and the Laplacian; see, for example, [5, 12, 15]. The vector of appropriate size with each entry equal to one will be denoted by 1. We now introduce some more definitions. If A is an n m matrix, then an m n matrix H is called a generalized inverse (or a g-inverse) of A if AHA = A. The Moore-Penrose inverse of A is an m n matrix H satisfying the equations AHA = A, HAH = H, (AH) = AH and (HA) = HA. It is well-known that the Moore-Penrose inverse exists and is unique. We denote the Moore-Penrose inverse of A by A +. For background material on generalized inverses, see [6, 7]. The following notation will be used in the rest of the paper. Let T be a tree with V (T ) = n + 1. We assume n 2. Let X 1 and X 2 be partite sets of T with X 1 = p 1, X 2 = p 2, p 1 + p 2 = n + 1. Let G i be a connected, directed graph with V (G i ) = X i and E(G i ) = m i, i = 1, 2. We assume that the edges of T are directed as well and, for convenience, we take them to be oriented from X 1 to X 2. We think of the graph T G 1 G 2 as a tree (T ) with attached graphs (G 1, G 2 ). Let A be the (n + 1) n incidence matrix of T and let B i be the p i m i incidence matrix of G i, i = 1, 2. We set B = B 1 0 0 B 2. (1) We remark that if p 2 = 1, then B 2 is vacuous and the last row of B consists of all zeros. The columns of A are linearly independent. Thus there is a unique n (m 1 + m 2 ) matrix Q satisfying AQ = B. We call Q, the incidence matrix, and L = QQ, the Laplacian, of T G 1 G 2. The rows and the columns of Q are indexed by E(T ) and E(G 1 ) E(G 2 ) respectively. The incidence vector of an edge of G 1 G 2 is a unique linear combination, with coefficients 0, ±1, of the 3

columns of A, and the corresponding column of Q merely lists the coefficients in such a linear combination. In particular, since the edges of T are all oriented from X 1 to X 2, the sum of the entries in every column of Q is zero. We also note some elementary properties of L. Clearly L is n n, symmetric, positive semidefinite, and has rank n 1. The entries of L are integers, but in contrast to the Laplacian of a graph, it does not necessarily have nonpositive off-diagonal entries. In this paper we consider the case when G 1 and G 2 are trees. Thus let T 1 and T 2 be directed trees on X 1 and X 2 respectively, and consider the tree with attached graphs, T T 1 T 2. The orders of the matrices A, B and Q are, respectively, (n+1) n, (n+1) (n 1) and n (n 1). The rows and columns of Q are indexed by E(T ) and E(T 1 T 2 ) respectively. As observed in [2], Q is totally unimodular, and in particular, all its entries are 0 or ±1. 3 Moore-Penrose inverse of Q Let T be a directed tree and let e E( T ). Clearly, T \ e is a forest with two components. The component of T \e that contains the head (respectively, tail) of e will be called the head-tree (respectively, tail-tree) of T \ e. We denote the head-tree of T \ e by H( T \ e) and the tail-tree of T \ e by T ( T \ e). We continue to work with the notation introduced in the previous section. Lemma 1 Define the (n 1) n matrix G as follows. The rows and the columns of G are indexed by E(T 1 T 2 ) and E(T ) respectively. If e E(T 1 ) and f E(T ), then the (e, f)-element of G is 1 if f is incident to T (T 1 \ e) and 0 otherwise. If e E(T 2 ) and f E(T ), then the (e, f)-element of G is 1 if f is incident to H(T 2 \ e) and 0 otherwise. Then G is a left-inverse of Q. Proof: Let e E(T 1 T 2 ) and consider the (e, e)-element of GQ. The e-th column of Q is the incidence vector of the directed path from the tail of e to the head of e. Thus the inner product of the e-th column of Q and the e-th 4

row of G is 1. Similarly, if e, f E(T 1 T 2 ), e f, then it can be seen that the (e, f)-element of GQ is zero. Thus GQ = I and the proof is complete. In the next result we present a formula for Q +. Theorem 2 The rows and the columns of Q + are indexed by E(T 1 T 2 ) and E(T ) respectively. Let e E(T 1 T 2 ) and f E(T ). Case (i): e E(T 1 ). If f is incident to a vertex in T (T 1 \ e), then the (e, f)-entry of Q + equals 1 n times the number edges of T incident to H(T 1 \ e). If f is incident to a vertex in H(T 1 \ e), then the (e, f)-entry of Q + equals 1 n times the number of edges of T incident to T (T 1 \ e). Case (ii): e E(T 2 ). If f is incident to a vertex in H(T 2 \e), then the (e, f)- entry of Q + equals 1 n times the number of edges of T incident to T (T 2 \ e). If f is incident to a vertex in T (T 2 \ e), then the (e, f)-entry of Q + equals 1 n times the number of edges of T incident to H(T 2 \ e). form Proof: The set of g-inverses of Q is given by (see [14], p.25) matrices of the G + X GQXGQ (2) where G is the g-inverse given in Lemma 1 and X is arbitrary. As noted in Lemma 1, GQ = I. Furthermore, I QG is the projection matrix on the nullspace of Q. Since the null-space of Q is one-dimensional and is spanned by the vector of all ones, I QG equals a scalar multiple of the matrix of all ones. Thus, in view of (2), the class of g-inverses of Q is seen to be matrices of the form G + where α 1,..., α n 1 are arbitrary. α 1. α n 1 [ 1 1 ], 5

Suppose Q + = G + β 1. [ 1 1 ]. (3) β n 1 Since Q + 1 = 0, then from the preceding equation, β 1. = 1 G1. (4) n β n 1 The result follows from Lemma 1 and the formulae (3), (4). We remark that when T 2 has only one vertex, Theorem 2 specializes to the formula for the Moore-Penrose inverse of the incidence matrix of a tree given in [1]. This will be elaborated in Section 5. 4 Inverse of K = Q Q. When Q is the incidence matrix of a tree, the matrix K = Q Q has been termed the edge Laplacian by Merris [11], who obtains a formula for the inverse of K. Later, Moon [13] presented another proof of the same formula. In this section we generalize the formula to the present setup. We continue to follow the notation introduced in Sections 2,3. Let K = Q Q. Clearly, K is of order (n 1) (n 1) with rank K = rank Q = n 1. Since the minors of Q of order (n 1) (n 1) are all equal to ±1, a simple application of the Cauchy-Binet formula shows that det K = n. We now present a formula for K 1. Theorem 3 The rows and the columns of K 1 are indexed by E(T 1 T 2 ). Let e, f E(T 1 T 2 ). Case (i): e, f E(T 1 ) and they are similarly oriented, i.e., the unique path in T 1 from e to f goes from the head of e to the tail of f, or vice versa. Let a 6

be the number edges of T incident to both T (T 1 \ e) and T (T 1 \ f), and let c be the number of edges of T incident to both H(T 1 \ e) and H(T 1 \ f). Then the (e, f)-element of K 1 is given by ac n. Case (ii): e, f E(T 1 ) and they are oppositely oriented, i.e., the unique path in T 1 from e to f goes from the head of e to the head of f, or from the tail of e to the tail of f. Let a, c be as defined in Case (i). Then the (e, f)-element of K 1 is given by ac n. If e, f E(T 2 ), then two cases arise, depending on whether e and f are similarly or oppositely oriented. These are similar to Cases (i) and (ii). Case (iii): e E(T 1 ), f E(T 2 ). Let a be the number of edges of T incident to both T (T 1 \ e) and H(T 2 \ f), let b be the number of edges of T incident to both T (T 1 \ e) and T (T 2 \ f), let c be the number of edges of T incident to both H(T 1 \ e) and H(T 2 \ f), and let d be the number of edges of T incident to both H(T 1 \ e) and T (T 2 \ f). Then the (e, f)-element of K 1 is given by ad bc n. Proof: Since K 1 = Q + (Q + ), the (e, f)-element of K 1 is given by j E(T ) q + ejq + fj. (5) First consider Case (i). Let a and c be as defined in Case (i) and let b be the number edges of T incident to both H(T 1 \ e) and T (T 1 \ f). If j E(T ) is incident to both T (T 1 \e) and T (T 1 \f), then, using Theorem 2, q ejq + fj + = b+c c. If j E(T ) is incident to both H(T n n 1 \ e) and T (T 1 \ f), then q ejq + fj + = a c. Finally, if j E(T ) is incident to both H(T n n 1 \ e) and H(T 1 \ f), then q ejq + fj + = a a+b. Substituting in (5) we see that n n j E(T ) q ejq + fj + = 1 {(b + c)ca acb + a(a + b)c} n2 = 1 n 2 {abc + ac2 abc + a 2 c + abc} = 1 ac(a + b + c) n2 = ac n 7

and the proof is complete in this case. The proof is similar for Case (ii). Now suppose Case (iii) holds and let a, b, c and d be as defined in Case (iii). If j E(T ) is incident to both T (T 1 \ e) and H(T 2 \ f), then, using Theorem 2, n 2 q + ejq + fj incident to both T (T 1 \ e) and T (T 2 \ f), then, n 2 q + ejq + fj = a(c + d)(b + d). If j E(T ) is = b(c + d)(a + c). If j E(T ) is incident to both H(T 1 \ e) and H(T 2 \ f), then, n 2 q + ejq + fj = c(a + b)(b + d). If j E(T ) is incident to both H(T 1 \ e) and T (T 2 \ f), then, n 2 q ejq + fj + = d(a + b)(a + c). Substituting in (5) we see that j E(T ) q ejq + fj + = 1 {(c + d)(ab + ad ab bc) + (a + b)(ad + cd bc cd)} n2 and the proof is complete. = 1 {(c + d)(ad bc) + (a + b)(ad bc)} n2 = 1 (ad bc)(a + b + c + d) n2 ad bc = n 5 Special cases Two special cases of the general setup considered in Sections 2 4 are of interest. The first is the case when T is a star. Thus suppose X 2 = 1. As remarked earlier, the matrix B 2 in (1) is vacuous in this case. However, B continues to have the last row of all zeros. The n n submatrix of A formed by the first n rows is the identity matrix of order n. A left-inverse of A is given by the identity matrix of order n augmented by a column of zeros. Thus Q is seen to equal B 1. There is a one-to-one correspondence between the edges of T and X 1 = V (G 1 ). Thus the results obtained in this paper specialize to results about the incidence matrix of G 1. (Of course, in the present paper G 1 is a tree.) The second special case of interest arises when G 1 and G 2 are paths. We 8

1 2 5 6 3 4 Table 1: A transportation tableau first introduce some notation. Consider a transportation problem with a set of sources S, with S = p, and a set of destinations, D, with D = q. To any feasible solution of the problem we may associate a bipartite graph. The partite sets of the graph are S and D. We assume that the elements of S and D are numbered 1,..., p and 1,..., q respectively. If i S and j D, then there is an edge from i to j if and only if a positive quantity is shipped from source i to destination j. It is well-known (see, for example, [9]) that such a bipartite graph is a tree if and only if the corresponding feasible solution is a basic feasible solution. From now onwards we assume that T is a tree corresponding to a basic feasible solution with S and D as its partite sets. If P 1 and P 2 denote paths with vertex sets S and D respectively then T P 1 P 2 is a tree with attached graphs. Let Q be the incidence matrix of this tree with attached graphs, as defined in Section 2. We may write formulas for the Moore-Penrose inverse of the incidence matrix Q and the inverse of K = Q Q using Theorem 2 and Theorem 3 respectively. We conclude with an example. Consider a transportation problem with 3 sources and 4 destinations. A tableau corresponding to a basic feasible solution and the associated tree with attached graphs (which are paths) are shown in Table 1 and Figure 1 respectively. 9

Figure 1: T P 1 P 2 The matrices A, B, Q, K for T P 1 P 2, as introduced in Sections 2 and 4, are given by A = 1 1 0 0 0 0 0 0 1 1 0 0 0 0 0 0 1 1 1 0 0 0 0 0 0 0 0 0 1 0 0 1 1 0 0 1 0 0 0 1 0 0, B = 1 0 0 0 0 1 1 0 0 0 0 1 0 0 0 0 0 1 0 0 0 0 1 1 0 0 0 0 1 1 0 0 0 0 1 Q = 0 0 1 0 0 1 0 1 0 0 1 1 0 0 1 0 0 0 0 1 0 0 1 1 0 0 1 1 1 0, K = Q Q = 2 1 1 0 1 1 2 1 1 1 1 1 4 2 0 0 1 2 2 0 1 1 0 0 2 The Moore-Penrose inverse of Q and the inverse of K are given by 10

Q + = 1 6 4 4 2 2 2 2 2 2 2 2 4 4 5 1 1 1 1 1 4 2 2 2 4 2 1 1 1 5 1 1, K 1 = 1 6 8 4 4 2 2 4 8 2 2 2 4 2 5 4 1 2 2 4 8 2 2 2 1 2 5 Note that the number of edges incident to the head-tree and the tail-tree, appearing in Theorems 2 and 3, are very easily seen from the transportation tableau. We leave it to the reader to verify that the expressions for Q + and K 1 given above confirm those given by Theorems 2 and 3. 11

REFERENCES 1. R. B. Bapat, Moore-Penrose inverse of the incidence matrix of a tree, Linear and Multilinear Algebra, 42 (1997), 159 167. 2. R. B. Bapat, Distance matrix and Laplacian of a tree with attached graphs, Linear Algebra and Its Applications, to appear. 3. R.B. Bapat, J.W. Grossman and D. M. Kulkarni, Generalized matrix tree theorem for mixed graphs, Linear and Multilinear Algebra, 46 (1999), 299 312. 4. R.B. Bapat and D. M. Kulkarni, Minors of some matrices associated with a tree, Algebra and Its Applications, D. V. Huynh, S. K. Jain, S. R. López-Permouth, Ed., Contemporary Mathematics, American Math. Society, 259 (2000), 45 66. 5. Béla Bollobás, Modern Graph Theory, Springer-Verlag, 1998. 6. A. Ben-Israel and T. N. E. Greville, Generalized Inverses: Theory and Applications, Wiley-Interscience, 1974. 7. S. L. Campbell and C. D. Meyer, Jr., Generalized Inverses of Linear Transformations, Pitman, 1979. 8. S. Chaiken, A combinatorial proof of the all minors matrix tree theorem, SIAM J. Algebraic Discrete Methods, 3 (1982), 319 329. 9. E. L. Kaplan, Mathematical Programming and Games, John Wiley, 1982. 10. S. J. Kirkland, M. Neumann and B. L. Shader, Distances in weighted trees and group inverse of Laplacian matrices, SIAM J. Matrix Anal. Appl. 18(4) (1997), 827 841. 11. R. Merris, An edge version of the matrix-tree theorem and the Wiener index, Linear and Multilinear Algebra, 25 (1989), 291 296. 12

12. R. Merris, Laplacian matrices of graphs: a survey, Linear Algebra Appl. 197,198 (1994), 143 176. 13. J. W. Moon, On the adjoint of a matrix associated with trees, Linear and Multilinear Algebra, 39 (1995), 191 194. 14. C. R. Rao, Linear Statistical Inference and Its Applications, John Wiley, 1973 15. D. B. West, Introduction to Graph Theory, 2nd Ed., Prentice-Hall Inc., 2001. 13