Lecture 10 Introduction To Algebraic Graph Theory

  • June 2020
  • PDF

This document was uploaded by user and they confirmed that they have the permission to share it. If you are author or own the copyright of this book, please report to us by using this DMCA report form. Report DMCA


Overview

Download & View Lecture 10 Introduction To Algebraic Graph Theory as PDF for free.

More details

  • Words: 4,557
  • Pages: 11
CSE 713: Random Graphs and Applications SUNY at Buffalo, Fall 2003

Lecturer: Hung Q. Ngo Scribe: Hung Q. Ngo

Lecture 10: Introduction to Algebraic Graph Theory Standard texts on linear algebra and algebra are [2,14]. Two standard texts on algebraic graph theory are [3, 6]. The monograph by Fan Chung [5] and the book by Godsil [7] are also related references.

1

The characteristic polynomial and the spectrum

Let A(G) denote the adjacency matrix of the graph G. The polynomial pA(G) (x) is usually referred to as the characteristic polynomial of G. For convenience, we use p(G, x) to denote pA(G) (x). The spectrum of a graph G is the set of eigenvalues of A(G) together with their multiplicities. Since A (short for A(G)) is a real symmetric matrix, basic linear algebra tells us a few thing about A and its eigenvalues (the roots of p(G, x)). Firstly, A is diagonalizable and has real eigenvalues. Secondly, if λ is an eigenvalue of A, then the λ-eigenspace has dimension equal to the multiplicity of λ as a root of p(G, x). Thirdly, if n = |V (G)|, then Cn is the direct sum of all eigenspaces of A. Last but not least, rank(A) = n − m[0], where m[0] is the multiplicity of the 0-eigenvalue. Suppose A(G) has s distinct eigenvalues λ1 > · · · > λs , with multiplicities m[λ1 ], . . . , m[λs ] respectively, then we shall write   λ1 λ2 ... λs Spec(G) = m[λ1 ] m[λ2 ] . . . m[λs ] We also use λmax (G) and λmin (G) to denote λ1 and λs , respectively. Example 1.1 (The Spectrum of The Complete Graph). p(Kn , λ) = λI − J   λ −1 −1 ... −1   (λ+1)(λ−1) −(λ+1) −(λ+1) ... 0  λ λ λ   (λ+1)(λ−2) −(λ+1)   0 ... = det  0 λ (λ−1)  . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .   0 0 0 . . . (λ+1)(λ−(n−1)) (λ−(n−2)) = (λ + 1)n−1 (λ − n + 1) So,  Spec(Kn ) =

 n − 1 −1 1 n−1

Remark 1.2. Two graphs are co-spectral if they have the same spectrum. There are many examples of co-spectral graphs which are not isomorphic. There are also examples all the graphs with a particular spectral must be isomorphic. I don’t know of an intuitive example of co-spectral graphs (yet). Many examples can be found in the “bible” of graph spectra [15]. 1

A principal minor of a square matrix A is the determinant of a square submatrix of A obtained by taking a subset of rows and the same subset of columns. The principal minor is of order k if it has k rows and k columns. Proposition 1.3. Suppose p(G, x) = xn + c1 xn−1 + · · · + cn , then (i) c1 = 0. (ii) −c2 = |E(G)|. (iii) −c3 is twice the number of triangles in G. Proof. It is not difficult to see that (−1)i ci is the sum of the principal minors of A(G) of order i. Given this observation, we can see that (i) c1 = 0 since trA(G) = 0.   (ii) −c2 = |E(G)| since each non-zero principal minor of order 2 of A(G) corresponds to det 01 10 , and there is one such minor for each pair of adjacent vertices in G. (iii) Of all possible order-3 principal minors of A(G), the only non-zero minor is   0 1 1 det 1 0 1 = 2 1 1 0 which corresponds to a triangle in G.

Example 1.4. All principal minors of A(Km,n ) of order k 6= 2 are 0. Hence, p(Km,n , x) = xm+n + c2 xm+n−2 . By previous proposition, c2 = −mn. Thus, √ √  mn 0 − mn Spec(Km,n ) = 1 m+n−2 1 Notice that Spec(Km,n ) is symmetric above the eigenvalue 0. This beautiful property turns out to be true for all bipartite graphs, as the following lemma shows. Lemma 1.5 (The Spectrum of a Bipartite Graph). The following are equivalent statements about a graph G (a) G is bipartite. (b) The non-zero eigenvalues of G occurs in pairs λi , λj such that λi + λj = 0 (with the same multiplicity). (c) p(G, x) is a polynomial in x2 after factoring out the largest common power of x. Pn 2t+1 (d) = 0 for all t ∈ N. i=1 λi Proof. (a ⇒ b). First of all, we could assume that the bipartitions of G have the same size, otherwise adding more isolated vertices into one of the bipartitions only give  x us more 0 eigenvalues. We can 0 B permute the vertices of G so that A = A(G) = B T 0 . Let v = y be a λ-eigenvector. We have λv =      x −By x T 0 0 Av = B0T B0 xy = BBy T x . So, By = λx and B x = λy. Let v = −y then Av = B T x = −λ −y . Hence, v 0 is a (−λ)-eigenvector of A. The multiplicity of λ is the dimension of its eigenspace. The 2

mapping v → v 0 just described is clearly an invertible linear transformation, so the λ-eigenspace and the (−λ)-eigenspace have the same dimension. (b ⇒ c). Easy as (x − λi )(x + λi ) = x2 − λ2i . (c ⇒ d). When p(G, x) is a polynomial in x2 , its roots come in pairs λi + λj = 0, so that λ2t+1 + i 2t+1 λj = 0 for each pair. P (d ⇒ a). = ni=1 λ2t+1 = trA2t+1 by Proposition ??. Also, trA2t+1 is at least the total number of i closed walks of length 2t + 1 in G. So G does not have any cycle of odd length. It must be bipartite. Proposition 1.6 (A Reduction Formula for p(G, x)). Suppose vi is a vertex of degree 1 of G, and vj is v1 ’s neighbor. Let G1 = G − vi , and G2 = G − {vi , vj }, then p(G, x) = (xp(G1 , x) − p(G2 , x)) . Proof. Expanding the determinant of (xI − A) along row i and then column j yields the result. Example 1.7 (The Characteristic Polynomial of a Path). Let Pn be the path with n vertices {v1 , . . . , vn }, then p(Pn , x) = xp(Pn−1 , x) − p(Pn−2 , x), n ≥ 3; which is a straightforward application of the previous proposition. Note that this implies p(Pn , x) = Un (x/2) where Un is the Chebyshev polynomial of the second kind. For the sake of completeness, recall that the Chebyshev polynomial of the second kind has generating function ∞ X 1 u(t, x) = = Un (x)tn , 1 − 2xt + t2 n=0

for |x| < 1 and |t| < 1; which gives the three-term recurrence Un (x) = 2xUn−1 (x) − Un−2 (x).(why?) Proposition 1.8 (The Derivative of p(G, x)). For i = 1, . . . , n, let Gi be G − vi where V (G) = {v1 , . . . , vn }. Then, X p0 (G, x) = p(Gi , x). i

Proof. Write p0 (G, x) = (xn + c1 xn−1 + · · · + ci xn−i + · · · + cn )0 = nxn−1 +

n−j X

(n − j)cj xn−j−1 .

j=1

Now, nxn−1 distributes to n leading terms of p(Gi , x). We show that the terms (n − j)cj xn−j−1 also distribute to the corresponding terms of p(Gi , x). We know cj is (−1)j times the sum of all order-j principle minors of A. We want to show that (n − j)cj (−1)j is the sum of all order-j principle minors of all Ai = A(Gi ). An order-j principle minor of any Ai is an order-j principle minor of A. An order-j principle minor of A is an order-j principle minor of precisely (n − j) of the Ai . The j exceptions are the Ai obtained from A by removing one of the j rows (and columns) corresponding to the minor under consideration.

3

Example 1.9. Suppose A(G) has r identical columns indexed {i1 , . . . , ir }, i.e. those r vertices share the same set of neighbors. Let x be a vector all of whose components are 0 except at two components is and it where xis = −xit 6= 0. Then x is a 0-eigenvector of A. The vector space spanned by all these x has dimension r − 1 (why?), so the 0-eigenspace of A has dimension at least r − 1. This fact could be obtained by seeing that rank(A) ≤ n − r + 1 due to the r identical columns, then apply rank(A) = n − m[0]. P Example 1.10. It’s easy to see that theP number of closed of length k of G is trAk = λki . Hence, P walks if G has n vertices and m edges then λi = 0 and λ2i = 2m. (Here we let λ1 ≥ λ2 ≥ · · · ≥ λn be the eigenvalues of G.) It follows trivially that λ21 = (λ2 + · · · + λn )2 ≤ (n − 1)(2m − λ21 ). So, r 2m 2m(n − 1) ≤ λ1 ≤ , n n where the lower bound is shown in the next section.

2

Eigenvalues and some basic parameters of a graph

The eigenvalues of a graph gives pretty good bounds on certain parameters of a graph. I include here several representative results. More relationships of this kind shall be presented later (e.g. the chromatic number in section 5). Lemma 2.1. If G0 is an induced subgraph of G, then λmin (G) ≤ λmin (G0 ) ≤ λmax (G0 ) ≤ λmax (G) Proof. Follows directly from the theorem about interlacing of eigenvalues Lemma 2.2. For every graph G, δ(G) ≤ λmax (G) ≤ ∆(G). Proof. Let x be a λ-eigenvector for some eigenvalue λ of G. Let |xj | = maxi |xi | be the largest absolute coordinate value in x, then X |λ||xj | = |(Ax)j | = |xi | ≤ deg(j)|xj | ≤ ∆(G)|xj | i | ij∈E(G)

For the lower bound, let 1 be the all-1 vector. Applying Rayleigh’s principle yields λmax ≥

1T A1 1X 2|E(G)| = aij = T 1 1 n n i,j

Thus, actually λmax is at least the average degree. Proposition 2.3 (Largest eigenvalue of regular graphs). If G is a k-regular graph, then (i) k is an eigenvalue of G. (ii) if G is connected, then m[k] = 1. (iii) for any other eigenvalue λ of G, λ ≤ k. 4

Proof. Let ~1 denote the all 1 vector, then A~1 = k~1, showing (i). Now, let x = [x1 , . . . , xn ]t be any k-eigenvector of G, then (Ax)i is the sum of k of the xj for which j is a neighbor of i. Moreover, (kx)i is kxi . If xi was the largest among all components of x, then it follows that all k neighboring xj must have the same value as xi . Tracing this neighboring relation we conclude that all of x’s components are the same. In fact, if G is a union of m k-regular graphs, then the multiplicity of the eigenvalue k of G is m. The fact that λ ≤ k can be shown by a similar argument, we just have to pick a component with largest absolute value. Theorem 2.4 (Alon, Milman (1985, [1])). Suppose G is a k-regular connected graph with diameter d, then &r ' 2k d≤2 log2 n . k − λ2 Proof. An improvement was given by Mohar: Theorem 2.5 (Mohar (1991, [11])). Suppose G is a k-regular connected graph with diameter d, then   2k − λ2 ln(n − 1) . d≤2 4(k − λ2 ) Proof.

3

The Coefficients of the Characteristic Polynomial

Theorem 3.1 (Harary, 1962 [8]). Let H be the collection of spanning subgraphs of a simple graph G such that for all H ∈ H, every component of H is either an edge or a cycle. Let c(H) and y(H) be the number Pof components and the number of components that are cycles of H, respectively. Then, det A(G) = H∈H (−1)n−c(H) 2y(H) , where n = |V (G)|. Q P Proof. We use det A = π∈Sn sgn(π) ni=1 aiπ(i) . A term corresponding to π of this product is not zero iff aiπ(i) = 1 for all i, namely π is a permutation such that (i, π(i)) ∈ E(G). In other words, if H(π) is the functional digraph of π with edges undirected, then H(π) ∈ H. Hence, there is a one-to-many mapping between H and the set of π which contribute 1 to det A. We can group the indices of the sum according to H instead, and count how many π with H(π) = H. Given H ∈ H, each cycle of length ≥ 3 has 2 choices of direction to construct the corresponding π, this gives the factor 2y(H) . The sign is readily verified. As we have noticed in the proof of the Matrix Tree theorem, sgn(π) = (−1)n−c(π) where c(π) is the number of cycles of π, which is the number of components of its functional digraph. Corollary 3.2 (Sachs, 1967 [13]). Let Hi denotes collection of i-vertex subgraphs of G whose P then−i components are edges or cycles. If p(G, λ) = c λ is the characteristic polynomial of G, then i i P c(H) y(H) ci = H∈Hi (−1) 2 . Proof. We already noticed that (−1)i ci is the sum of all order i principal minors of A(G). Each principal minor correspond uniquely to an induced subgraph of G on some i vertices. Applying Harary’s theorem completes our proof.

5

4

The Adjacency Algebra

Recall that an algebra is a vector space with an associative multiplication of vectors (thus also imposing a ring structure on the space). The adjacency algebra A(G) of G is the algebra of all polynomials in A(G). In other words, A(G) is the set of all linear combination of powers of A. A(G) is the basic tool to study a class of graphs called distance-regular graphs (see, e.g. [4] for a comprehensive treatment). The theory of distance-regular graphs, in turn, has deep relations to Coding Theory (see [10], [?]) and Design Theory (see [?]). We found yet another great reason to study algebraic graph theory. Obviously, it makes sense to first study powers of A. Proposition 4.1. The number of walks of length l in G, from vi to vj , is the (i, j) entry of A(G)l . Proof. Easy to see by inspection or by induction Lemma 4.2. If G is a connected graph with diameter d, then deg(m(A)) = dim(A(G)) ≥ d + 1. Proof. Let x, y ∈ V (G) with distance d apart. Suppose x = v0 , v1 , . . . , vd = y is a path of length d joining x and y. Then, for all i ∈ [d] the distance from x to vi is i. Consequently, (Ai )x,vi > 0 but (Aj )x,vj = 0, ∀j < i. This implies that for all i ∈ [d] Ai is independent from {I, A, . . . , Ai−1 }, or {I, A, . . . , Ad } is a set of independent members of A(G). Corollary 4.3. A graph with diameter d has at least d + 1 distinct eigenvalues. In other words, the diameter of a graph is strictly less than the number of its distinct eigenvalues. Proof. If A(G) has s distinct eigenvalues, then by Lemma ??, the minimum polynomial of A(G) has degree s, making dim(A(G)) = s. So, s ≥ d + 1 by the previous lemma.

5

The Chromatic Number

The following theorem improves the greedy bound χ(G) ≤ 1 + ∆(G). Theorem 5.1 (Wilf, 1967 [16]). For every graph G, χ(G) ≤ 1 + λmax (G), where χ(G) is the chromatic number of G. Proof. If χ(G) = k, successively delete vertices of G until we obtain a k-critical subgraph H of G, i.e. χ(H − v) = k − 1, ∀v ∈ V (H). We claim δ(H) ≥ k − 1. Suppose δ(H) ≤ k − 2, let v be the vertex in H with deg(v) ≤ k − 2. H − v is (k − 1)-colorable, so H is also k − 1 colorable since adding back v wouldn’t require a new color. Consequently, k ≤ 1 + δ(H) ≤ 1 + λmax (H) ≤ 1 + λmax (G)

It must be noted that this bound is still a poor estimate for the chromatic number. A parallel result concerning the lower bound is as follows. Theorem 5.2 (Hoffman, 1970 [9]). For any graph G with non-empty edge set χ(G) ≥ 1 +

We first need two auxiliary results. 6

λmax (G) −λmin (G)

Lemma 5.3. Let X be a real symmetric matrix, partitioned in the form   P Q X= QT R where P and R are square symmetric matrices, then λmax (X) + λmin (X) ≤ λmax (P ) + λmax (R) Proof. Let λ = λmin (X). Let X 0 = X − λI, P 0 = P − λI and R0 = R − λI, then clearly  0  P Q X0 = QT R0 Let A and B be defined as follows  P0 0 A = QT 0   0 Q B = 0 R0 

then, every eigenvalue of A (B) is an eigenvalue of P 0 (R0 ) since Ax = µx ⇒ P 0 y = µy where x = [y z]T with y being the part corresponding to P 0 . Consequently, the eigenvalues of A and B are all real. Theorem ?? implies λmax (X) − λ = λmax (X 0 ) = λ1 (A + B) ≤ λ1 (A) + λ1 (B) ≤ λ1 (P 0 ) + λ1 (R0 ) = λ1 (P ) − λ + λ1 (R) − λ

Corollary 5.4. Let A be a real symmetric matrix, partitioned into t2 submatrices Aij in such a way that the rows and columns are partitioned in the same way, i.e. the diagonal submatrices Aii are all square matrices. Then t X λmax (A) + (t − 1)λmin (A) ≤ λmax (Aii ) i=1

Proof. Induction and apply previous lemma Proof of Theorem 5.2. Let c = χ(G) and partition V (G) into c color classes, inducing a partition of A(G) into c2 submatrices where all diagonal submatrices Aii consist entirely of 0’s. Thus, λmax (A) + (c − 1)λmin (A) ≤

c X

λmax (Aii ) = 0

i=1

But if G has at least one edge, pA (λ) = λn + c1 λn−1 + · · · + cn 6= λn , because c2 = −|E(G)|. Hence, λmin (A) < 0. This completes the proof.

7

6

The Laplacian

This section is built upon the first chapter’s outline of Fan Chung’s book [5]. See has an entirely different system of notations and definitions (she normalized everything and defined the eigenvalues of a graph to be the eigenvalues of the Laplacian). So, I’ll try my best of map them back to our, I believe, more standard notations. However, the mapping isn’t so simple. It will take me some time to link the two definitions. Thus, courtesy Bill Gate : “the best is yet to come.”

6.1

The Laplacian and eigenvalues

Definition 6.1. Let G be a simple graph, D the diagonal matrix with (D)ii = deg(i), and A the adjacency matrix of G. Then, the matrix L := D − A is called the Laplacian matrix of G. We shall often use µ1 ≥ µ2 ≥ · · · ≥ µn to denote the eigenvalues of L. Definition 6.2. Let N be the incident matrix of any orientation H of G(V, E). Let L2 (V ) (L2 (E)) be the space p of real valued functions on V (E), with the usual inner product hf, gi and the usual norm kf k = hf, f i. Note that L2 (V ) is isomorphic to Rn and thus we can define the Rayleigh quotient for f similarly: i RA (f ) = hLf,f . Also note that kf k2 hLf, f i = hN T N f, f i = hN f, N f i X = (f (u) − f (v))2 (u,v)∈E(H)

=

X

(f (u) − f (v))2

u∼v

So, L is non-negative definite, which implies L has non-negative eigenvalues. We’ve just proved the first statement of the following proposition. Proposition 6.3. We have µ1 ≥ · · · ≥ µn−1 ≥ µn = 0, ∀i. Moreover, µn−1 = 0 iff G is not connected; and, when G is regular, m[0] is the number of connected components of G. Proof. Firstly, µn = 0 because L1 = 0, i.e. 1 is a 0-eigenvalue of L. Secondly, notice that any function y which is non-zero and constant on the connected components of G would make Ly = 0, and thus y is a 0-eigenvector of G. Hence, the multiplicity of 0, being the dimension of the 0-eigenspace, is ≥ 2 when G is disconnected. For the converse, we assume µn−1 = 0 so that the 0-eigenspace has dimension ≥ 2. Let f be any µn−1 -eigenvector orthogonal to 1 then X (f (u) − f (v))2 µn−1 =

u∼v

X

f 2 (v)

v

This means that f has to be constant on all connected components of G. If G has only 1 connected component, f has to be identically 0 contradicting the fact that it is an eigenvector. Lastly, also note that if each connected component of G is regular, then the multiplicity of 0 is equal to the number of connected components.

8

Theorem 6.4. Let f ∈ L2 (V ) such that then µn−1

P

f (v) = 0. Let µn−1 be the second smallest eigenvalue of L X (f (u) − f (v))2 hLf, f i ≤ = u∼v X kf k f 2 (v) v

v

In fact, a stronger statement holds µn−1 = min f 6=0

hLf, f i kf k

P with the min runs over all f satisfying v f (v) = 0. X (f (u) − f (v))2 is sometime called the Dirichlet sum of G. Note. u∼v

√ Proof. Let un = 1/ n be the unit µn -eigenvector, then by Theorem ?? we have µn−1 = min n RL (f ) = min n 06=f ∈C f ⊥un

The condition f ⊥ 1 is the same as

P

u f (u)

06=f ∈C f ⊥1

hLf, f i kf k

= 0.

Theorem 6.4 gives us a very useful upper bound for µn−1 . However, sometime we need also a lower bound. The following Proposition fills our gap. Proposition 6.5. Let G be a connected graph, µ = µn−1 (G) and f ∈ L2 (V ) be any µ-eigenvalue. Let V + := {v ∈ V | f (v) > 0 and V − := V − V + , then define g ∈ L2 (V ) as follows. ( f (v) if v ∈ V + g(v) = 0 otherwise. Then, we have X µ≥

(g(u) − g(v))2

u∼v

X

g 2 (v)

v

Proof. Note that since G is connected, µ 6= 0, making f 6= 0. Hence, V + 6= ∅. By definition, we have (Lf )(v) = µf (v), ∀v ∈ V . Thus, X (Lf )(v)f (v) µ=

v∈V +

X

f 2 (v)

v∈V +

But, X

f 2 (v) =

v∈V +

X v∈V

9

g 2 (v)

and, 

 X

(Lf )(v)f (v) =

v∈V +

X

d(v)f 2 (v) −

v∈V +

X

=

X u∈Γ(v)

(f (u) − f (v))2 +

X

X

f (u)(f (u) − f (v))

uv∈E(V + ,V − )

uv∈E(V + )



f (v)f (u)

(g(u) − g(v))2

u∼v

completes our proof.

6.2

The Laplacian spectrum

6.3

Eigenvalues of weighted graphs

6.4

Eigenvalues and random walks

7

Cycles and cuts

8

More on spanning trees

9

Spectral decomposition and the walk generating function

10

Graph colorings

11

Eigenvalues and combinatorial optimization

This section shall be based on an article with the same title by Bojan Mohar and Svatopluk Poljak [12].

References [1] N. A LON AND V. D. M ILMAN, λ1 , isoperimetric inequalities for graphs, and superconcentrators, J. Combin. Theory Ser. B, 38 (1985), pp. 73–88. [2] M. A RTIN, Algebra, Prentice-Hall Inc., Englewood Cliffs, NJ, 1991. [3] N. B IGGS, Algebraic graph theory, Cambridge University Press, Cambridge, second ed., 1993. [4] A. E. B ROUWER , A. M. C OHEN , AND A. N EUMAIER, Distance-regular graphs, Springer-Verlag, Berlin, 1989. [5] F. R. K. C HUNG, Spectral graph theory, Published for the Conference Board of the Mathematical Sciences, Washington, DC, 1997. [6] D. M. C VETKOVI C´ , M. D OOB , AND H. S ACHS, Spectra of graphs, Johann Ambrosius Barth, Heidelberg, third ed., 1995. Theory and applications. [7] C. D. G ODSIL, Algebraic combinatorics, Chapman & Hall, New York, 1993. [8] F. H ARARY, The determinant of the adjacency matrix of a graph, SIAM Rev., 4 (1962), pp. 202–210. [9] A. J. H OFFMAN, On eigenvalues and colorings of graphs, in Graph Theory and its Applications (Proc. Advanced Sem., Math. Research Center, Univ. of Wisconsin, Madison, Wis., 1969), Academic Press, New York, 1970, pp. 79–91. [10] F. J. M AC W ILLIAMS AND N. J. A. S LOANE, The theory of error-correcting codes. II, North-Holland Publishing Co., Amsterdam, 1977. North-Holland Mathematical Library, Vol. 16.

10

[11] B. M OHAR, Eigenvalues, diameter, and mean distance in graphs, Graphs Combin., 7 (1991), pp. 53–64. [12] B. M OHAR AND S. P OLJAK, Eigenvalues in combinatorial optimization, in Combinatorial and graph-theoretical problems in linear algebra (Minneapolis, MN, 1991), Springer, New York, 1993, pp. 107–151. ¨ [13] H. S ACHS, Uber Teiler, Faktoren und charakteristischePolynome vonGraphen. II, Wiss. Z. Techn. Hochsch. Ilmenau, 13 (1967), pp. 405–412. [14] G. S TRANG, Linear algebra and its applications, Academic Press [Harcourt Brace Jovanovich Publishers], New York, second ed., 1980. [15] D. T SVETKOVICH , M. D UB , AND K. Z AKHS, Spektry grafov, “Naukova Dumka”, Kiev, 1984. Teoriya i primenenie. [Theory and application], Translated from the English by V. V. Strok, Translation edited by V. S. Korolyuk, With a preface by Strok and Korolyuk. [16] H. S. W ILF, The eigenvalues of a graph and its chromatic number, J. London Math. Soc., 42 (1967), pp. 330–332.

11

Related Documents