Average Mixing Matrix of Trees

Similar documents
Note on deleting a vertex and weak interlacing of the Laplacian spectrum

Refined Inertia of Matrix Patterns

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

The least eigenvalue of the signless Laplacian of non-bipartite unicyclic graphs with k pendant vertices

When can perfect state transfer occur?

A note on a conjecture for the distance Laplacian matrix

No Laplacian Perfect State Transfer in Trees

The Matrix-Tree Theorem

Bipartite graphs with at most six non-zero eigenvalues

On the adjacency matrix of a block graph

Laplacian Integral Graphs with Maximum Degree 3

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

On the maximum positive semi-definite nullity and the cycle matroid of graphs

Lower bounds for the Estrada index

State transfer in strongly regular graphs with an edge perturbation

Ma/CS 6b Class 23: Eigenvalues in Regular Graphs

Sparse spectrally arbitrary patterns

Bounds for the Zero Forcing Number of Graphs with Large Girth

A Simple Proof of Fiedler's Conjecture Concerning Orthogonal Matrices

Strongly Regular Decompositions of the Complete Graph

Triply Regular Graphs

New feasibility conditions for directed strongly regular graphs

Periodicity & State Transfer Some Results Some Questions. Periodic Graphs. Chris Godsil. St John s, June 7, Chris Godsil Periodic Graphs

Average mixing of continuous quantum walks

Linear algebra and applications to graphs Part 1

Chapter 2 Spectra of Finite Graphs

Eigenvectors and Reconstruction

A Questionable Distance-Regular Graph

The spectrum of the edge corona of two graphs

Which Quantum Walk on Graphs?

Topics in Graph Theory

On the distance signless Laplacian spectral radius of graphs and digraphs

Matching Polynomials of Graphs

Kernels of Directed Graph Laplacians. J. S. Caughman and J.J.P. Veerman

Ma/CS 6b Class 20: Spectral Graph Theory

Potentially nilpotent tridiagonal sign patterns of order 4

Graph coloring, perfect graphs

ELA A LOWER BOUND FOR THE SECOND LARGEST LAPLACIAN EIGENVALUE OF WEIGHTED GRAPHS

Intriguing sets of vertices of regular graphs

Bicyclic digraphs with extremal skew energy

Lecture 7: Positive Semidefinite Matrices

SIMPLICES AND SPECTRA OF GRAPHS

Smith Normal Form and Acyclic Matrices

Peter J. Dukes. 22 August, 2012

Quantum Walks on Strongly Regular Graphs

ALGEBRAIC MATCHING THEORY. C. D. Godsil 1

LINEAR ALGEBRA BOOT CAMP WEEK 1: THE BASICS

Graphs with Integral Spectrum

Locating-Total Dominating Sets in Twin-Free Graphs: a Conjecture

Normalized rational semiregular graphs

Lecture Summaries for Linear Algebra M51A

arxiv: v1 [math.co] 20 Sep 2012

Properties of θ-super positive graphs

Regular graphs with a small number of distinct eigenvalues

The third smallest eigenvalue of the Laplacian matrix

Nilpotent matrices and spectrally arbitrary sign patterns

A linear algebraic view of partition regular matrices

Ma/CS 6b Class 20: Spectral Graph Theory

PROBLEMS ON LINEAR ALGEBRA

1 Last time: least-squares problems

The Local Spectra of Regular Line Graphs

Minimum rank of a graph over an arbitrary field

An algebraic proof of the Erdős-Ko-Rado theorem for intersecting families of perfect matchings

Near-domination in graphs

Spectra of Semidirect Products of Cyclic Groups

Bulletin T.CXXII de l Académie Serbe des Sciences et des Arts Classe des Sciences mathématiques et naturelles Sciences mathématiques, No 26

A lower bound for the spectral radius of graphs with fixed diameter

ZERO FORCING PARAMETERS AND MINIMUM RANK PROBLEMS. March 2, 2010

Spectral results on regular graphs with (k, τ)-regular sets

Central Groupoids, Central Digraphs, and Zero-One Matrices A Satisfying A 2 = J

Applicable Analysis and Discrete Mathematics available online at GRAPHS WITH TWO MAIN AND TWO PLAIN EIGENVALUES

arxiv:quant-ph/ v1 22 Aug 2005

Lecture 5 January 16, 2013

Extremal Graphs for Randić Energy

An Introduction to Algebraic Graph Theory

ALGEBRAIC COMBINATORICS. c1993 C. D. Godsil

Extremal graph on normalized Laplacian spectral radius and energy

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

The symmetric minimal rank solution of the matrix equation AX=B and the optimal approximation

(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 =

An Introduction to Spectral Graph Theory

Complex Hadamard matrices and 3-class association schemes

The Minimum Rank, Inverse Inertia, and Inverse Eigenvalue Problems for Graphs. Mark C. Kempton

An upper bound for the minimum rank of a graph

arxiv: v4 [math.co] 12 Feb 2017

The edge-density for K 2,t minors

Applied Mathematics Letters

The spectra of super line multigraphs

arxiv: v2 [math.co] 27 Jul 2013

Lecture 13 Spectral Graph Algorithms

Distance-regular graphs where the distance-d graph has fewer distinct eigenvalues

DISTINGUISHING PARTITIONS AND ASYMMETRIC UNIFORM HYPERGRAPHS

Iowa State University and American Institute of Mathematics

SVD, Power method, and Planted Graph problems (+ eigenvalues of random matrices)

Matrix functions that preserve the strong Perron- Frobenius property

Uniform Star-factors of Graphs with Girth Three

A New Spectral Technique Using Normalized Adjacency Matrices for Graph Matching 1

Mathematical Programs Linear Program (LP)

Hadamard matrices and strongly regular graphs with the 3-e.c. adjacency property

235 Final exam review questions

Transcription:

Electronic Journal of Linear Algebra Volume 34 Volume 34 08 Article 9 08 Average Mixing Matrix of Trees Chris Godsil University of Waterloo, cgodsil@uwaterloo.ca Krystal Guo Université libre de Bruxelles, kguo@uwaterloo.ca John Sinkovic University of Waterloo, john.sinkovic@uwaterloo.ca Follow this and additional works at: Part of the Discrete Mathematics and Combinatorics Commons Recommended Citation Godsil, Chris; Guo, Krystal; and Sinkovic, John. 08, "Average Mixing Matrix of Trees", Electronic Journal of Linear Algebra, Volume 34, pp. 69-8. DOI: https://doi.org/0.300/08-380, 537-958.3746 This Article is brought to you for free and open access by Wyoming Scholars Repository. It has been accepted for inclusion in Electronic Journal of Linear Algebra by an authorized editor of Wyoming Scholars Repository. For more information, please contact scholcom@uwyo.edu.

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. AVERAGE MIXING MATRIX OF TREES CHRIS GODSIL, KRYSTAL GUO, AND JOHN SINKOVIC Abstract. The rank of the average mixing matrix of trees with all eigenvalues distinct, is investigated. The rank of the average mixing matrix of a tree on n vertices with n distinct eigenvalues is bounded above by n/. Computations on trees up to 0 vertices suggest that the rank attains this upper bound most of the times. An infinite family of trees whose average mixing matrices have ranks which are bounded away from this upper bound, is given. A lower bound on the rank of the average mixing matrix of a tree, is also given. Key words. Algebraic graph theory, Eigenspaces, Quantum walks. AMS subject classifications. 8P68, 05C50, 8P6.. Introduction. We investigate the rank of the average mixing matrix of continuous time quantum walks on trees with all simple eigenvalues. The continuous-time quantum walk is a universal computational primitive [4]. Since their introduction in [7], many properties of quantum walks have been studied. Some topics include state transfer [5,, 4, 6] and uniform mixing [, 3, 3]. Let X be a graph and let A be the adjacency matrix of X. The transition matrix of a continuous-time quantum walk on X is a matrix-valued function in time, denoted Ut, and is given as follows: Ut = expita. Unlike a classical random walk on a connected graph, a continuous quantum walk does not reach a stationary distribution. For example, the transition of the complete graph K n is a periodic function of t. The average mixing matrix is, intuitively, a distribution that the quantum walk adheres to, on average, over time, and thus, may be thought of as a replacement for a stationary distribution. The average mixing matrix has been studied in [, 6, ]. We will now proceed with a few preliminary definitions. The mixing matrix Mt of the quantum walk, given by Mt = Ut Ut. where denotes the Schur also known as the Hadamard or element-wise product of two matrices. The average mixing matrix, denoted M, is defined as follows:. M = lim T T T 0 Mt dt. Received by the editors on February 7, 08. Accepted for publication on April 6, 08. Handling Editor: Bryan L. Shader. Corresponding Author: Krystal Guo. University of Waterloo, Waterloo, Canada cgodsil@uwaterloo.ca. C. Godsil gratefully acknowledges the support of the Natural Sciences and Engineering Council of Canada NSERC, grant no. RGPIN-9439. Université libre de Bruxelles, Brussels, Belgium guo.krystal@gmail.com. This work was done when K. Guo was a postdoctoral fellow at the University of Waterloo. University of Waterloo, Waterloo, Canada john.sinkovic@uwaterloo.ca.

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. Chris Godsil, Krystal Guo, and John Sinkovic 70 A tree is a connected graph with no cycles. The class of trees is a natural class of graphs to study, since it is a smaller class of graphs with well-understood combinatorial properties and eigenvalues. Trees have the desirable property of being sparse and initial data for the rank of the average mixing matrix prompted further study of the class. The rank of the average mixing matrix of a tree on n vertices with n distinct eigenvalues is at most n /, see [6]. We computed the rank of the average mixing matrix of all trees up to 0 vertices; the computational results suggest that the rank of M of a tree with all simple eigenvalues attains the upper bound most of the time. We give an infinite family of trees whose average mixing matrices have ranks which are bounded away from this upper bound. Theorem.. For every positive real number c, there exists a tree T on n vertices with all simple eigenvalues such that n rk MT > c. We then give the following lower bound on the rank of the average mixing matrix of a tree: Theorem.. If T is a tree with all simple eigenvalues on at least four vertices and is not isomorphic to P 4, then rk MT 3. We give some preliminary results about the average mixing matrix in Section. We also find the ranks and traces of the average mixing matrices of star graphs, as an example. In Section 3, we define the rooted product of graphs and we find the average mixing matrix of the rooted product of X and Y = {K,..., K } in terms of the eigenprojectors of X. We use this result in Section 4 to give a family of trees whose average mixing matrices have ranks which are upper-bounded away from the maximum. In Section 5, we prove the lower bound on the ranks of the average mixing matrices of trees which are not isomorphic to the path on four vertices.. Preliminaries and examples. Let X be a graph on n vertices and let A be the adjacency matrix of X. Let θ,..., θ d be the distinct eigenvalues of A and, for r =,..., d, let E r be the idempotent projection onto the θ r eigenspace of A; the spectral decomposition of A is as follows: A = d θ r E r. The following is an important theorem as it allows us to understand the average mixing matrix. r= Theorem.. [] Let X be a graph and let A be the adjacency matrix of X. Let A = d r= θ re r be the spectral decomposition of A. The average mixing matrix of X is M = d E r E r. r= As an example, we find the average mixing matrix of the star and deduce that it always has full rank. Lemma.. The average mixing matrix of the star graph K,n has full rank and has tr M = n 3n + 3. n

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. 7 Average Mixing Matrix of Trees Proof. We will proceed by finding an explicit expression for MK,n. Observe that the eigenvalues of K,n are ± n and 0 with multiplicity n. We give the spectral idempotent matrices as follows: E n = n T n n J, E n = n T n n J and E 0 = 0 0 0 I n J, where I is the n n identity matrix and J is the n n all ones matrix. Note that I nj always has rank n. Thus we have that MK,n = E n + E = = n T n n 3 n + E 0 n J + n T n J + n I 0 0 0 n I + n J. We can now observe that this matrix has full rank; we can easily see that the lower n n matrix does not have 0 as an eigenvalue and adding the first row and column does not introduce any linear dependencies. rk M = 3 P 6 E 6 ˆD5 rk M = 5 rk M = 6 D 6 T K,5 Figure. All trees on 6 vertices, organized by the rank of their average mixing matrices. The path graphs P and P 3 are the only trees on two and three vertices, respectively. There are six trees on six vertices; three of them have all simple eigenvalues and rank 3, two have rank 4 and one graph the star graph of full rank. They are shown in Figure.

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. Chris Godsil, Krystal Guo, and John Sinkovic 7 3. Rooted products. Let X be a graph with vertices {v,..., v n } and let Y be a disjoint union of rooted graphs Y,..., Y n, rooted at y,..., y n, respectively. The rooted product of X and Y, denoted XY, is the graph obtain by identifying v i with the root vertex of Y i. The rooted product was first introduced by Godsil and McKay in [8]. Let φy i, t denote the characteristic polynomial of Y i and let BX, Y be the matrix given as follows: φy i, t, if i = j; BX, Y i,j = φy i \ y i, t, if i j; 0, otherwise. Theorem 3.. [8] φxy, t = detb. In particular, we will consider the special case where Y is a sequence of n copies of K. In this case, we will write XK to denote the rooted product and the following is found in [8] as a consequence of the above theorem 3. φxk, t = t n φ H, t. t The following lemma follows directly from 3.. Lemma 3.. Let X be a graph with all simple eigenvalues. The rooted product XK also has all simple eigenvalues which are roots of for λ an eigenvalue of X. t λt = 0, Lemma 3.3. Let X be a graph with all simple eigenvalues and let {v,..., v n } be an orthonormal eigenbasis of AX with eigenvalues λ,..., λ n, respectively. For i =,..., n, let µ i and ν i be the two roots of t λ i t = 0. Then { µi v i µ i + v i is an orthonormal eigenbasis for XK., ν i + Proof. We may write the adjacency matrix of XK as follows: AXK = } νi v i i =,..., n v i AX I. I 0 For i [n] and µ {µ i, ν i }, we see that µ 0 and µ = λµ +. We obtain that and the lemma follows. AX I µvi λi µv = i + v i µ v = i µvi = µ, I 0 v i µv i µv i v i

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. 73 Average Mixing Matrix of Trees Theorem 3.4. Let X be a graph with all simple eigenvalues and let F,..., F n the orthogonal projections onto the eigenspaces of AX with corresponding eigenvalues λ,..., λ n. Then MX N N MXK =, N MX N where N = n i= λ i + 4 F i F i. Proof. Following the notation of the previous lemma, we see that F i = v i v T i. For i [n], let µ {µ i, ν i }. The orthogonal projection E µ onto the µ eigenspace of AXK is given by µ + µvi v i µv T i v T i µ v = i vi T µv i vi T µ + µv i vi T v i vi T = µ + µ F i µf i µf i F i. Thus, E µi E µi + E νi E νi = µ 4 i + ν 4 i µ i + ν i + µ i + ν i µ i + νi + F i F i F i F i µ i + ν i µ i + νi + + µ i + νi + F i F i. F i F i Observe that µ i + ν i = λ i and µ i ν i =, and so µ i + ν i = µ i + ν i µ i ν i = λ i + and µ 4 i + ν 4 i = λ 4 i + 4µ i + 4ν i 6 = λ 4 i + 4λ i +. We obtain that µ 4 i µ + νi 4 i + ν = µ4 i ν i + + ν 4 i µ i + = λ4 i + 6µ i + 6ν i 4 i + µ i + νi + λ i + = λ i + 4 λ i + 4, and µ i µ + νi i + ν = µ i ν i + + ν i µ i + = λ i + 8 i + µ i + νi + λ i + = 4 λ i + 4 µ i + + ν i + = ν i + + µ i + µ i + ν i + = λ i + λ i + 4.

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. Chris Godsil, Krystal Guo, and John Sinkovic 74 Thus, λ i + E µi E µi + E νi E νi = λ i +4 F i F i λ i +4 F i F i λ i +4 F i F i λ i +. λ i +4 F i F i Let We can see that N = n i= λ i + 4 F i F i. MX N N MXK =, N MX N and the lemma follows. We observe that N and M have the same kernel. 4. Trees with all simple eigenvalues with rk M bounded away from n /. The following lemma is found in [6]. Lemma 4.. [6] If X is a bipartite graph on n vertices with n distinct eigenvalues, then rk MX n /. In this section, we will give a construction for trees on n vertices with all simple eigenvalues where rk M is bounded away from n /. A computer search finds that for n =,..., 7, 9, 0, every tree on n vertices with all simple eigenvalues has rk M = n /. Up to isomorphism, there is one tree T on 8 vertices such that rk MT = 8, which is given in Figure. We will retain the notation of T to denote this for the rest of this section. 4 3 6 5 8 4 7 3 0 5 9 6 0 7 Figure. The tree T on 8 vertices such that MT = 8.

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. 75 Average Mixing Matrix of Trees We will need the following technical lemma. Lemma 4.. Let X be a graph with adjacency matrix A. Suppose the spectral decomposition AX = d r=0 θ re r. Let M be the average mixing matrix of X. We have that d ker M = kere r E r. Proof. Let v ker M. We see that d 0 = v T Mv = v T E r E r v. Since E r E r is a principal submatrix of E r E r and is hence positive semi-definite, we see that r=0 r=0 v T E r E r v 0 for all r. Since the sum of non-negative real number is equal to zero, every term must be zero, and thus, d ker M kere r E r. Containment in the other direction is clear. r=0 We will use Lemma 4. to bound the rank of iterated rooted products of T with K. We note that Lemma 4. implies that vertices u and v are strongly cospectral in X if and only if Me u = Me v, which is also proven in []. The characteristic polynomial of T is which has all simple roots. φt, x =x x + x x x + x ] x 3 x x + Let X 0 = T and X i+ = X i K for i > 0. x 3 + x x x 6 8x 4 + x, Lemma 4.3. For i 0, the graph X i has all simple eigenvalues and rk MX i i+3. Proof. We proceed by induction on i. We computed that rk MT = 8 and T has all simple eigenvalues. Suppose the statement is true for i >. By Lemma 3., the graph X i+ has all simple eigenvalues. For simplicity, let n = i 8, the number of vertices of X i. Let F,..., F n be the orthogonal projections onto the eigenspaces of AX i with corresponding eigenvalues λ,..., λ n. Theorem 3.4 gives that MXi N N MX i+ =, N MXi N

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. Chris Godsil, Krystal Guo, and John Sinkovic 76 where N = n λ j + 4 F j F j. j= Let v ker MX i. By Lemma 4., we see that v kerf j F j for all j, and thus, Nv = 0. Let 0 m denote the m dimensional all zero vector. We have v MX i+ = 0 n MXi N N N v = 0 n MXi N 0 n and 0n MXi N MX i+ = v N 0n N = 0 n. MXi N v Thus, if {v k } k is an orthogonal basis for ker MX i, then { vk is a set of dim ker MX i vectors in ker MX i+, and the statement follows. 0 n, Theorem 4.4. For every positive real number c, there exists a tree T with all simple eigenvalues such that V T 0n v k } k rk MT > c. Proof. Note that X i as constructed above is a tree on i 8 vertices such that V Xi rk MX i i 8 i 8 = i. For any c > 0, we pick i such that i > c, and the result follows. 5. Lower bound on rank. We denote by X \u the graph obtained from X by deleting vertex u. The coefficient matrix of a graph X is the matrix C with rows indexed by the vertices of X and columns indexed by integers [n] such that the u, r entry of C is the coefficient of t r in the characteristic polynomial φx\u, t. We will see that for graphs with all simple eigenvalues, the coefficient matrix is useful in studying the average mixing matrix. The following lemma appears in the proof of [, Theorem 3.]. Lemma 5.. [] Let X be a graph on n vertices with n distinct eigenvalues θ,..., θ n. Let be the n n diagonal matrix whose r-th diagonal entry is φ X, θ r. Let V be the n n Vandermonde matrix with ij-entry equal to θ i j. Then MX = CV V T C T. From this, we derive the following corollary as a direct consequence.

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. 77 Average Mixing Matrix of Trees Corollary 5.. If X is a graph with all simple eigenvalues, then rk MX is equal to the rank of the coefficient matrix. We will also use this standard fact, which can be found in [0], for example. Lemma 5.3. If T is a tree, then the characteristic polynomial of T is equal to the matching polynomial of T. Lemma 5.4 is Corollary 8.9. in [9]. Lemma 5.4. [9] If A is a symmetric matrix of rank r, then it has a principal submatrix of full rank. Corollary 5.5 follows from Lemma 5.4 and the fact that a tree has a perfect matching if and only if its adjacency matrix has full rank. Corollary 5.5. If T is a tree with all simple eigenvalues, then either T has a perfect matching or there exists a vertex v such that T \ v has a perfect matching. Proof. If T is a tree with all simple eigenvalues, then zero can have multiplicity at most one. If the multiplicity of zero is zero, then T has a perfect matching. If the multiplicity of zero is one, then the rank of AT is T. Thus, by Lemma 5.4, AT has T T principal submatrix of full rank. Thus, there exists a vertex v such that AT \ v has full rank. It follows that T \ v has a perfect matching. Lemma 5.6. If T is a tree with all simple eigenvalues, and T 3, then T has a leaf adjacent to a vertex of degree two. Proof. The path on three vertices satisfies the conclusion of the lemma. Let T be a tree with all simple eigenvalues and T 4. Consider a diametrical path P = {v 0,..., v d } of T. Since the star on n 4 vertices does not have all simple eigenvalues, d 3. Since P is a diametrical path of length at least three, v and v d are distinct and adjacent to exactly one non-leaf vertex of T. If either v or v d has degree two, then we are done. Suppose by way of contradiction, that both vertices have degree greater than two. Deleting v and v d from T yields a graph with at least four isolated vertices. As each isolated vertex contributes a zero eigenvalue, the multiplicity of zero is at least four. Since the eigenvalues of induced subgraphs interlace, the multiplicity of the zero eigenvalue of T is at least. This contradicts that T has all simple eigenvalues. Theorem 5.7. If T is a tree with all simple eigenvalues on at least four vertices and is not isomorphic to P 4, then rk MT 3. Proof. In light of Corollary 5., it is sufficient to show that the rank of the coefficient matrix of a tree is at least three. By Lemma 5.6, T has a leaf u adjacent to a vertex v of degree two. Let w be the non-leaf neighbor of v with degw = l. Let m α T denote the number of matchings of size α in T. By Lemma 5.3, the ith row of the coefficient matrix of T can be expressed in terms of m α T \i for the appropriate value of α. Case : T has a perfect matching. As T has a perfect matching, n = T is even. The only tree on four vertices with a perfect matching is P 4. It has two pairs of cospectral vertices, and thus, rk MP 4 =. For the remainder of this case, assume that n 6. Consider the 3 3 submatrix C of the coefficient matrix of T with rows corresponding to u, v, w

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. Chris Godsil, Krystal Guo, and John Sinkovic 78 and columns corresponding to t, t n 3, and t n. For simplicity in the notation let k = n /. Then k m k T \ u m T \ u m 0 T \ u C = m k T \ v m T \ v m 0 T \ v. k m k T \ w m T \ w m 0 T \ w As m 0 T \i is defined to be and m T \i is the number of edges in T \i, it remains to determine the values of the first column. In T \ u, the set of matchings of size k can be partitioned into two sets: the k matchings which use edge vw and those that do not. Thus m k T \ u = m k T \ {u, v, w} + m k T \ {u, v}. Since T has a perfect matching and every perfect matching of T contains the edge uv, we have that T \ {u, v} also contains a perfect matching. Thus m k T \ {u, v} =. Similarly, m k T \ v =. Since T \ w has uv as a component and any matching of size k uses uv, we see that m k T \ w = m k T \ {u, v, w}. Letting q = m k T \ {u, v, w}, we obtain k q + n C = k n 3 k q n l and det C = k + q l. Note that det C = 0 if and only if l = + /q. As l N, this occurs if and only if q = and l =. We claim that if l =, then q. Let y be the other neighbor of w. Since T has a perfect matching, so does T \ {u, v, w, y}. Thus m k T \ {u, v, w, y} = and T \ {u, v, w} has a k matching which uses no edge incident to y. Since n 6, degy. Thus T \ {u, v, w} has at least one k matching which uses an edge incident to y. Therefore q = m k T \ {u, v, w}. Further det C 0 and the coefficient matrix of T has rank at least 3. Case : T does not have a perfect matching. There are two possibilities to consider; either T \ u has a perfect matching or it does not. In the case that T \u has a perfect matching, n = T is odd and so T 5. Consider the 3 3 submatrix C of the coefficient matrix of T with rows corresponding to u, v, w and columns corresponding to t 0, t n 3, and t n. For simplicity in the notation let n / = j. Then

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. 79 Average Mixing Matrix of Trees j m j T \ u m T \ u m 0 T \ u C = m j T \ v m T \ v m 0 T \ v. j m j T \ w m T \ w m 0 T \ w As stated in Case, m 0 T \ i = and m T \ i is the number of edges in T \ i. Matchings of size j = n / are perfect matchings in a graph on n vertices. Further since T \u has a perfect matching so does T \ w. Lastly, T \ v does not have a perfect matching as u is isolated by deleting v. Thus, j n C = 0 n 3. j n l Note that det C = j l. Since degw = l, C has full rank and so the coefficient matrix of T has rank at least three. We now consider the other possibility mentioned at the start of this case. Assume that T \ u does not have a perfect matching. By Corollary 5.5 there exists a vertex z of degree f such that T z has a perfect matching. Thus, n = T is odd and n 5. Consider the 3 3 submatrix C 3 of the coefficient matrix of T with rows corresponding to u, v, z and columns corresponding to t 0, t n 3, and t n. For simplicity in the notation let n / = j. Then j m j T \ u m T \ u m 0 T \ u C 3 = m j T \ v m T \ v m 0 T \ v. j m j T \ z m T \ z m 0 T \ z As in the other cases, the entries of the last two columns are easily filled. The entries in the first column are determined by whether T \ i has a perfect matching. Thus, 0 n C 3 = 0 n 3. j n f Note that det C 3 = j+ 0. Thus, C 3 has full rank and so the coefficient matrix of T has rank at least three. 6. Computational data. Computations of average mixing matrices of trees on up to 0 were carried out, using Sage Mathematics Software [5]. The results are recorded in Tables and. For each order n and rank r, we record the number of trees on n vertices whose average mixing matrix has rank r and, amongst those, the number with n distinct eigenvalues.

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. Chris Godsil, Krystal Guo, and John Sinkovic 80 n rank # trees # simple eigenvalues 3 4 4 0 5 3 5 0 6 3 3 5 0 6 0 7 4 5 5 5 0 6 4 0 7 0 8 4 5 4 5 4 0 6 8 0 7 4 0 8 0 9 5 9 8 6 3 0 7 5 0 8 7 0 9 3 0 0 4 0 5 4 6 9 0 7 30 0 8 0 9 6 0 0 5 0 n rank # trees # simple eigenvalues 5 0 6 64 6 7 8 0 8 79 0 9 40 0 0 6 0 7 0 5 0 6 44 37 7 06 0 8 9 0 9 9 0 0 93 0 48 0 0 3 6 0 7 64 50 8 07 0 9 4 0 0 3 0 86 0 87 0 3 0 4 6 4 0 7 46 6 8 55 0 9 59 0 0 694 0 6 0 34 0 3 7 0 4 37 0 Table Ranks of average mixing matrices of trees on to 4 vertices. 7. Open problems. In Section 5, we give a constant lower on the rank of the average mixing matrix of trees with all simple eigenvalues on at least 4 vertices. Table 3 shows the minimum ranks among the average mixing matrices of trees of n vertices.

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. 8 Average Mixing Matrix of Trees n rank # trees # simple eigenvalues 5 7 4 0 8 7 04 9 663 0 0 73 0 365 0 38 0 3 79 0 4 309 0 5 63 0 6 7 7 0 8 543 465 9 96 0 0 834 0 465 0 388 0 3 650 0 4 494 0 5 600 0 6 0 0 7 8 0 9 4889 445 0 435 0 653 0 8340 0 3 9347 0 4 574 0 5 300 0 6 46 0 7 9 0 8 7 0 8 5 9 08 77 0 5306 0 n rank # trees # simple eigenvalues 8 489 0 6545 0 3 494 0 4 949 0 5 980 0 6 609 0 7 4 0 8 368 0 9 8 0 9 5 0 0 59 9884 604 0 6470 0 3 5349 0 4 630 0 5 4383 0 6 7389 0 7 603 0 8 459 0 9 78 0 0 8 5 0 9 43 0 0 864 7055 8498 0 79080 0 3 6508 0 4 5309 0 5 39556 0 6 08 0 7 583 0 8 6098 0 9 8405 0 0 343 0 Table Ranks of average mixing matrices of trees on 5 to 0 vertices. n 3 4 5 6 7 8 9 0 3 4 5 6 7 8 9 0 min rank 3 3 4 4 5 4 5 5 6 6 7 7 8 7 8 8 Table 3 The minimum ranks of the average mixing matrices of trees on n vertices, for n =,..., 0.

Electronic Journal of Linear Algebra, ISSN 08-380 Volume 34, pp. 69-8, May 08. Chris Godsil, Krystal Guo, and John Sinkovic 8 Finally, we are motivated by the computational data to ask the following question: Conjecture 7.. For any k, the number of trees T such that the rank of the average mixing matrix of T is at most k is finite. REFERENCES [] W. Adamczak, K. Andrew, L. Bergen, D. Ethier, P. Hernberg, J. Lin, and C. Tamon. Non-uniform mixing of quantum walk on cycles. International Journal of Quantum Information, 5:78 793, 007. [] W. Adamczak, K. Andrew, P. Hernberg, and C. Tamon. A note on graphs resistant to quantum uniform mixing. Preprint, arxiv:quant-ph/0308073, 003. [3] A. Chan. Complex Hadamard matrices, instantaneous uniform mixing and cubes. Preprint, arxiv:305.58, 03. [4] A.M. Childs. Universal computation by quantum walk. Physical Review Letters, 0:8050, 009. [5] G. Coutinho, C. Godsil, K. Guo, and F. Vanhove. Perfect state transfer on distance-regular graphs and association schemes. Linear Algebra and its Applications, 478:08 30, 05. [6] G. Coutinho, C. Godsil, K. Guo, and H. Zhan. A new perspective on the average mixing matrix. Preprint, arxiv:709.0359, 07. [7] E. Farhi and S. Gutmann. Quantum computation and decision trees. Physical Review A, 58:95 98, 998. [8] C. Godsil and B. McKay. A new graph product and its spectrum. Bulletin of the Australian Mathematical Society, 8:-8, 978. [9] C. Godsil and G. Royle. Algebraic Graph Theory. Graduate Texts in Mathematics, Vol. 07, Springer-Verlag, New York, 00. [0] C.D. Godsil. Algebraic Combinatorics. Chapman & Hall, New York, 993. [] C.D. Godsil. State transfer on graphs. Discrete Mathematics, 3:9 47, 0. [] C.D. Godsil. Average mixing of continuous quantum walks. Journal of Combinatorial Theory, Series A, 0:649 66, 03. [3] C.D. Godsil, N. Mullin, and A. Roy. Uniform mixing and association schemes. Preprint, arxiv:30.5889, 03. [4] A. Kay. Perfect, efficient, state transfer and its application as a constructive tool. International Journal of Quantum Information, 8:64 676, 00. [5] W. Stein, et al. Sage Mathematics Software Version 6... The Sage Development Team, http://www.sagemath.org, 04. [6] L. Vinet and A. Zhedanov. How to construct spin chains with perfect state transfer. Physical Review A, 85:033, 0.