Also available at http://amc.imfm.si ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 6 (2013) 261–278 Interlacing–extremal graphs∗ Irene Sciriha †, Mark Debono , Martha Borg Department of Mathematics, Faculty of Science, University of Malta, Msida, Malta Patrick W. Fowler , Barry T. Pickup Department of Chemistry, Faculty of Science, University of Sheffield, Sheffield S3 7HF, UK Received 18 November 2011, accepted 13 September 2012, published online 19 November 2012 Abstract A graph G is singular if the zero-one adjacency matrix has the eigenvalue zero. The multiplicity of the eigenvalue zero is called the nullity of G. For two vertices y and z of G, we call (G, y, z) a device with respect to y and z. The nullities of G, G − y, G − z and G−y− z classify devices into different kinds. We identify two particular classes of graphs that correspond to distinct kinds. In the first, the devices have the minimum allowed value for the nullity of G− y − z relative to that of G for all pairs of distinct vertices y and z of G. In the second, the nullity of G− y reaches the maximum possible for all vertices y in a graph G. We focus on the non–singular devices of the second kind. Keywords: Adjacency matrix, singular graphs, nut graphs, uniform–core graphs, nuciferous graphs, interlacing. Math. Subj. Class.: 05C50,05C35, 05C60, 05B20, 92E10, 74E40 1 Introduction A graphG = G(V, E) of order n has a labelled vertex set V = {1, 2, ..., n}. The set E ofm edges consists of unordered pairs of adjacent vertices. We write V(G) for a graph G when the graph G needs to be specified. A subset of V is said to be independent if no two of its vertices are adjacent, i.e., no two are connected by an edge. For a subset V1 of V , G − V1 denotes the subgraph of G induced by V\V1. The subgraph of G obtained by deleting a particular vertex y is denoted by G − y and that obtained by deleting two distinct vertices ∗This research was supported by the Research Project Funds MATRP01-02 Graph Spectra and Fullerene Molecular Structures of the University of Malta. †Corresponding Author, http://staff.um.edu.mt/isci1/ E-mail addresses: irene.sciriha-aquilina@um.edu.mt (Irene Sciriha), debonomark@gmail.com (Mark Debono), bambinazghira@gmail.com (Martha Borg), P.W.Fowler@sheffield.ac.uk (Patrick W. Fowler), b.t.pickup@sheffield.ac.uk (Barry T. Pickup) Copyright c© 2013 DMFA Slovenije 262 Ars Math. Contemp. 6 (2013) 261–278 y and z is denoted by G− y − z. A graph is said to be bipartite if its vertex set V may be partitioned into two independent subsets V1 and V2. The cycle and the complete graph on n vertices are denoted by Cn and Kn, respectively. The complete bipartite graph Kn1,n2 has a vertex partition into two subsets V1 and V2 of independent vertices of sizes n1 and n2, respectively, and has edges between each member of V1 and each member of V2. 1.1 The adjacency matrix The graphs we consider are simple, that is, without loops or multiple edges. We use A(G) (or just A when the context is clear) to denote the 0-1 adjacency matrix of a graph G, where the entry aik of the symmetric matrix A is 1 if {i, k} ∈ E and 0 otherwise. We note that the graph G is determined, up to isomorphism, by A. The adjacency matrix AC of the complement GC of G is J− I−A, where each entry of J is one and I is the identity matrix. The degree of a vertex i is the number of non–zero entries in the ith row of A. If the adjacency matrix A of a n-vertex graph G satisfies Ax = λx for some non-zero vector x then x is said to be an eigenvector belonging to the eigenvalue λ. There are n linearly independent eigenvectors. The eigenvalues of A are said to be the eigenvalues of G and to form the spectrum of G. They are obtained as the roots of the characteristic polynomial φ(G,λ) of the adjacency matrix of G, defined as the polynomial det(λI−A) in λ. Cauchy’s inequalities for a Hermitian matrix M (also collectively known as the In- terlacing Theorem) place restrictions on the multiplicity of the eigenvalues of principal submatrices relative to those of M (See [6] for instance). When they are applied to graphs we have: Theorem 1.1. Interlacing Theorem: LetG be an n-vertex graph and w ∈ V . If the eigen- values of G are λ1, λ2, . . . , λn and those of G − w are ξ1, ξ2, . . . , ξn−1, both in non- increasing order, then λ1 ≥ ξ1 ≥ λ2 ≥ ξ2 ≥ . . . ≥ ξn−1 ≥ λn. 1.2 Cores of singular graphs For the linear transformation A, the kernel, ker(A), of A is defined as the subspace of Rn mapped to zero by A. It is also referred to as the nullspace of A. A graph G is said to be singular of nullity ηG if the dimension of the nullspace ker(A) of A is ηG and ηG > 0. If there exists a non-zero vector x in the nullspace of the adjacency matrix A, then x is said to be a kernel eigenvector of the singular graph G and satisfies Ax = 0. It is therefore an eigenvector of A for the eigenvalue zero whose multiplicity ηG is also the number of roots of φ(G,λ) equal to zero. A vertex corresponding to a non-zero entry of x is said to be a core vertex CV of G. The core vertices corresponding to x induce a subgraph of G termed the core of G with respect to x. The core structure of a singular graph will be the basis of our classification of all graphs relative to ηG. A core graph is a singular graph in which every vertex is a core vertex. The empty graph (K4)C and the four cycle C4 are examples of 4–vertex core graphs of nullity four and two, respectively. A core graph of order at least three and nullity one is known as a nut graph. It is connected and non–bipartite [12]. For singular graphs, the vertices can be partitioned into core and core-forbidden ver- tices. The set CV of core vertices consists of those vertices lying on some core of G. A core-forbidden vertex (CFV) corresponds to a zero entry in every kernel eigenvector. The set V\CV is the set of CFVs. It follows that, in a core graph, the set of CFVs is empty. I. Sciriha et al.: Interlacing–extremal graphs 263 Let y and z be two distinct vertices of a graph G. By interlacing, when a vertex y or z is deleted from G, the nullity ηG−y or ηG−z , that is the multiplicity of the eigenvalue zero of G − y or G − z, respectively, may take one of three values from ηG − 1 to ηG + 1. If the two distinct vertices y and z are deleted, then the nullity ηG−y−z of G − y − z may take values in the range from ηG − 2 to ηG +2. Let us call the graph having two particular distinct vertices y and z a device (G, y, z). The set of devices can be partitioned into three main varieties, namely variety 1 when both vertices are CVs, variety 2 when one vertex is a CFV and one a CV and variety 3 when both vertices are CFVs. A device (G, y, z) is said to be of kind (ηG, ηG−y, ηG−z, ηG−y−z). Since ηG−y and ηG−z can take three values each and ηG−y−z can take five values, there are potentially 45 kinds of graphs relative to ηG. Interlacing further restricts the values of ηG−y−z . Moreover, there are kinds of graphs that exclude certain combinatorial properties, such as that of being bipartite, as we shall see in Section 5. In Section 2, we express the characteristic polynomial of φ(G−y, λ) as the sum of two terms in ληG and ληG − 1 with coefficients fa(λ) and fb(λ), respectively, each of which is a polynomial expanded in terms of the entries of the eigenvectors of A forming an orthonormal basis for Rn. By comparing the diagonal entries of the adjugate of (λI −A) and of the spectral decomposition of (λI − A)−1 we obtain, in Section 3, an expression for φ(G− y − z) as the sum of three terms in ληG , ληG − 1 , ληG − 2, respectively, with polynomial coefficients. Moreover, the well known Jacobi’s identity (see, for instance, [4]), relating the entries of the adjugate of (λI−A) with the characteristic polynomials of a graph G and those of particular subgraphs of G, is used to determine which kinds are not realized by any graph G. In Section 4, the vertices of a graph are partitioned into three subsets of type lower, middle or upper, respectively, according to the vanishing or otherwise of fa(0) and fb(0). The Interlacing Theorem and Jacobi’s identity impose restrictions on the 45 kinds, so that not all are possible. In Sections 5 and 6, we show why there exist exactly twelve kinds of device (G, y, z) and how they are partitioned into the three main varieties. In Section 7, we identify two interesting classes of graphs that in a certain sense have extremal nullities. The first one has the minimum possible nullity ηG−y−z , that is ηG − 2, for all pairs of distinct vertices y and z in a graph G. A graph G in the second class has the maximum possible nullity ηG−y , that is ηG + 1, for all vertices y of G. We show that devices within the second class can reach the maximum allowed ηG + 2 for the nullity ηG−y−z for some but not for all pairs of distinct vertices y and z in a graph G. A characterization is given of the non–singular devices within the second class having the inverse A−1 of the adjacency matrix A with zero entries only on the diagonal. 2 Characteristic polynomials We first need to define some necessary notation. Associated with the n× n adjacency matrix A of a n–vertex graph of nullity ηG, there is an ordered orthonormal basis xr, 1 ≤ r ≤ n, for Rn, consisting of eigenvectors of A, with the ηG eigenvectors in the nullspace being labelled first. Let the n× 1 column vector 264 Ars Math. Contemp. 6 (2013) 261–278 xr be (xry), where for vertex y, 1 ≤ y ≤ n. If P =  x11 x 2 1 · · · xn1 x12 x 2 2 · · · xn2 ... ... ... ... x1n x 2 n · · · xnn  , where the ith column of P is the eigenvector xi belonging to the eigenvalue λi in the spectrum of A, diagonalization of A is given by P−1AP = D[λi], where D[λi] is the diagonal matrix having λi as the ith entry on the main diagonal. Expressing A in terms of D and P leads to the spectral decomposition theorem, which can also be applied to (λI−A)−1. This leads to an expression for the characteristic polynomial of the adjacency matrix φ(G − y, λ) of G − y which is given explicitly in terms of the eigenvector entries {xiy}. Together with Jacobi’s identity, it will serve as a basis for the characterization of graphs according to those kinds that can exist. Lemma 2.1. φ(G− y, λ) = n∑ i=1 (xiy) 2 (λ− λi) φ(G,λ). Proof. The characteristic polynomial of the adjacency matrix φ(G− y, λ) of G− y is the yth diagonal entry (adj(λI−A))yy of the adjugate of (λI−A). For arbitrary λ, the matrix (λI−A) is invertible and φ(G−y, λ) = ((λI−A)−1)yyφ(G,λ). Since P−1AP = D[λi], it follows that adj(λI−A) φ(G,λ) = (λI−A)−1 = PD[ 1 λ− λi ]P−1. Taking the yth diagonal entry, φ(G− y, λ) φ(G,λ) = (x1y x 2 y · · ·xny )D[ 1 λ− λi ]  x1y x2y ... xny  = n∑ i=1 (xiy) 2 (λ− λi) . (2.1) For a graph G with adjacency matrix A of nullity ηG, let s(λ) denote φ(G,λ). If the spectrum of A is λ1, λ2, · · · , λn, starting with the zero eigenvalues (if any), we write s(λ) = n∏ `=1 (λ− λ`) = ληGs0(λ) with s0(0) 6= 0. (2.2) Partitioning the range of summation in Equation (2.1), φ(G− y, λ) φ(G,λ) = ηG∑ i=1 (xiy) 2 λ + n∑ i=ηG+1 (xiy) 2 λ− λi I. Sciriha et al.: Interlacing–extremal graphs 265 Hence φ(G− y, λ) = ηG∑ k=1 (xky) 2s0(λ)λ ηG−1 + n∑ k=ηG+1 (xky) 2s0(λ)λ ηG λ− λk (2.3) which we shall express as φ(G− y, λ) = fbληG−1 + faληG . (2.4) 3 Jacobi’s Identity Relative to (G, y, z), let us denote by j(λ), or j, the entry of the adjugate adj(λI−A) in the yz position, obtained by taking the determinant of the submatrix of (λI−A) after deleting row y and column z and multiplying it by (−1)y+z . We use the convention that ηG−y ≥ ηG−z . Throughout the paper, where the context is clear, we may write s0 for s0(λ), j for j(λ), etc. Let s(λ), t(λ), u(λ), v(λ), often referred to simply as s, t, u and v respectively, be the characteristic polynomials φ(G,λ), φ((G−y), λ), φ((G− z), λ), φ((G−y− z), λ) of the graphs G, G− y, G− z and G− y − z, respectively, that is, the determinants s(λ) = det(λI−A(G)) t(λ) = det(λI−A(G− y)) u(λ) = det(λI−A(G− z)) v(λ) = det(λI−A(G− y − z)). (3.1) From Lemma 2.1, t(λ) = n∑ k=1 (xk y )2 ∏ ` 6=k (λ− λ`) (3.2) and u(λ) = n∑ k=1 (xk z )2 ∏ ` 6=k (λ− λ`). (3.3) We shall see that the characteristic polynomial v(λ) of G− y− z can also be expressed in terms of the eigenvector entries {xry} and {xrz} associated with distinct vertices y and z. Lemma 3.1. For y 6= z, Jacobi’s identity expresses the entry j of the adjugate of λI−A in the yz position, for a symmetric matrix A, in terms of the characteristic polynomials s, u, t and v: j2 = ut− sv Expressing Equations (3.2) and (3.3) as in (2.4), t(λ) = ηG∑ k=1 (xky) 2s0(λ)λ ηG−1 + n∑ k=ηG+1 (xky) 2s0(λ)λ ηG λ− λk = tbλ ηG−1 + taλ ηG , (3.4) and u(λ) = ηG∑ k=1 (xkz) 2s0(λ)λ ηG−1 + n∑ k=ηG+1 (xkz) 2s0(λ)λ ηG λ− λk (3.5) 266 Ars Math. Contemp. 6 (2013) 261–278 = ubλ ηG−1 + uaλ ηG Now we consider pairs of vertices of G. Since adj(λI −A) φ(G,λ) = (λI −A)−1 = PD[ 1 λ− λi ]P−1, j(λ) = n∑ k=1 (xk y xk z ) ∏ ` 6=k (λ− λ`.) (3.6) We can write j(λ) = ηG∑ k=1 xkyx k zs0(λ)λ ηG−1 + n∑ k=ηG+1 xkyx k zs0(λ)λ ηG λ− λk (3.7) = jbλ ηG−1 + jaλ ηG The characteristic polynomial v(λ) can be written as v(λ) = u(λ)t(λ)− j2(λ) s(λ) , that is v(λ) = vaληG + vbληG−1 + vcληG−2, where vc = 1 s0 (ubtb − j2b ) = 1 2 s0 ηG∑ i=1 ηG∑ `=1 (xizx ` y − x`zxiy)2 vb = 1 s0 (uatb + ubta − 2jajb) = s0 ηG∑ i=1 n∑ `=ηG+1 (xizx ` y − xiyx`z)2 λ− λ` va = 1 s0 (uata − j2a) = 1 2 s0 n∑ i=ηG+1 n∑ `=ηG+1 (xiyx ` z − x`yxiz)2 (λ− λi)(λ− λ`) (3.8) 4 Three types of vertex By interlacing, we can identify three types of vertex according to the effect on the nullity on deletion. We call a vertex y lower, middle or upper if the nullity of G− y is ηG − 1, ηG or ηG + 1, respectively. We shall distinguish among these three types of vertex according to the values of the functions fa and fb in Equation (2.4). In Table 1 we show the entries of the orthonormal eigenvectors {xr} in an ordered basis for Rn as presented in Section 2. We choose a vertex labelling such that the core vertices are labelled first. Note the zero submatrix corresponding to the CFVs. We consider φ(G− y, λ) s0ληG from Equation 2.3. It has poles at λ = µi, 1 ≤ i ≤ h, where, for 1 ≤ i ≤ h, the µi are the h distinct non-zero eigenvalues of G. Moreover, the gradient of n∑ k=ηG+1 (xky) 2 λ− λk is less than 0 for all λ 6= µi. It follows that φ(G− y, λ) s0ληG has at most (h − 1) roots strictly interlacing the h distinct eigenvalues of A. Note that ηG∑ k=1 (xky) 2 ≥ 0 with equality if and only if y is a CFV. Thus at λ = 0, fb is non–zero if y is a CV and zero I. Sciriha et al.: Interlacing–extremal graphs 267 hhhhhhhhhhhhhhhvertex-entries eigenvector x1 . . . xηG xηG+1 . . . xn x1 ∗ . . . ∗ ∗ . . . ∗ x2 ∗ . . . ∗ ∗ . . . ∗ ... ∗ . . . ∗ ∗ . . . ∗ x|CV | ∗ . . . ∗ ∗ . . . ∗ x|CV |+1 0 . . . 0 ∗ . . . ∗ ... ... ... ... ... ... ... xn 0 . . . 0 ∗ . . . ∗ Table 1: Ordered orthonormal basis of eigenvectors of A with * representing a possibly non–zero entry. if it is a CFV. For a CFV y, n∑ k=ηG+1 (xky) 2 λ− λk vanishes at λ = 0 when y is upper, and does not vanish when y is middle. Note that when ηG∑ k=1 (xky) 2 = 0, one of the (h− 1) interlacing roots may be zero. (†) Different cases occur depending on the vanishing or otherwise of the real constant ηG∑ k=1 (xky) 2 and n∑ k=ηG+1 (xky) 2 λ− λk at λ = 0. Equation (2.3) and the analysis in the previous paragraph (marked (†)) lead to the result that ηG − 1 ≤ ηG−y ≤ ηG + 1. This can be generalized for the multiplicity of any eigenvalue of G other than zero by replacing the cores and the nullspace of G by the µi–cores and µi–eigenspace of G (concepts introduced in [10]), thus giving another proof of the Interlacing Theorem. Proposition 4.1. The values of fb and fa of Expression (2.4) for φ(G − y, λ) at λ = 0 distinguish the three types of vertex as follows: Vertex y Status of y The values of fb and fa Lower CV fb(0) 6= 0 Middle CF fb(0) = 0 and fa(0) 6= 0 Upper CFV fb(0) = 0 and fa(0) = 0 Proof. Let y be a core vertex of a graph of nullity ηG > 0. There exists xky 6= 0 for some k, 1 ≤ k ≤ ηG. Then fb(0) 6= 0, which is a necessary and sufficient condition for the multiplicity of the eigenvalue zero to be ηG − 1 for G− y. It follows that a vertex is lower if and only if it is a CV. If y is a CFV, then fb(0) = 0. For G − y, the multiplicity of the eigenvalue zero is at least ηG. If one of the roots of n∑ k=ηG+1 (xky) 2 λ− λk is zero, then λ divides n∑ k=ηG+1 (xky) 2 λ− λk , the multiplicity of the eigenvalue zero is exactly ηG + 1 for G − y and the vertex y is upper. Otherwise the multiplicity of the eigenvalue zero remains ηG for G− y and the vertex y is middle. 268 Ars Math. Contemp. 6 (2013) 261–278 We consider three varieties of devices {(G, y, z)} with pairs (y, z) of vertices, namely variety 1 with both y and z being CVs, variety 2 with z being a CV and y a CFV and variety 3 with both y and z being CFVs. Since a CFV can be upper or middle, varieties 2 and 3 are subdivided further, as seen in Table 3. From Proposition 4.1, for variety 1, ub 6= 0; tb 6= 0; for variety 2, ub 6= 0; jb = tb = vc = 0; for variety 3: ub = jb = tb = vb = vc = 0. Some of these varieties can be further subdivided according to the values at λ = 0 of vc, vb and va or ja. From Proposition 4.1, tb(0) 6= 0 if and only if y is a core vertex. Similarly ub 6= 0 if and only if z is a core vertex. If at least one of z or y is core forbidden, then jb(0) = 0. However, there are ‘accidental’ cases where jb(0) vanishes when both z and y are CVs, for example in C4 andK2,3 if the vertices y and z are connected by an edge. Indeed this is true for all bipartite core graphs of nullity at least two, since each of u and t has zero as a root. It follows that E2η is a factor of j2 = ut − sv = (jbEη−1 + jaEη)2 and therefore jb(0) = 0. 5 Restrictions on the nullity of G − y − z It is our aim to classify all graphs according to their kind defined by the quadruple (ηG, ηG−y, ηG−z, ηG−y−z). Not all the 45 kinds mentioned in Section 1 exist, as we shall discover. The classification will be given in Table 3 on Page 272. It is best possible since each kind is realized by some graph. 5.1 Restrictions arising from interlacing In a device (G, y, z) of kind (ηG, ηG−y, ηG−z, ηG−y−z), interlacing restricts the values that ηG−y−z can take. The following result shows an instance when ηG−y−z is determined by interlacing alone. Lemma 5.1. For (ηG, ηG−y, ηG−z, ηG−y−z) = (ηG, ηG +1, ηG − 1, ηG−y−z), the nullity ηG−y−z of G− y − z is ηG. Hence, (ηG, ηG + 1, ηG − 1, ηG) is the only kind where the nullities ηG−y and ηG−z differ by two. We say that it belongs to variety 2a. In kinds where the nullities ηG−y and ηG−z differ by one, interlacing allows ηG−y−z to take either the value ηG−y or ηG−z . All three possible values of ηG−y−z are allowed by interlacing when ηG−y = ηG−z . The symmetry about zero of the spectrum of a bipartite graph G (See for instance [8]) requires that the number of zero eigenvalues is 2k, if G has an even number of vertices and 2k+1 if G has an odd number of vertices, for some k ≥ 0. This implies that on deleting a vertex from a bipartite graph, the nullity changes parity. Therefore if the nullity of a graph G and of its vertex–deleted subgraph G − y are the same, then G is not bipartite. Since on deleting a vertex a bipartite graph remains bipartite, it follows that a graph G of a kind where ηG = ηG−y or ηG−y = ηG−y−z cannot be bipartite. Lemma 5.2. If a vertex of a graph is middle, then the graph is not bipartite. Figure 1 shows a device (G, y, z) with a middle vertex z which becomes upper inG−y. I. Sciriha et al.: Interlacing–extremal graphs 269 Figure 1: A graph with two middle vertices y and z. 5.2 Restrictions arising from Jacobi’s Identity Lemma 3.1 requires that ut− sv which is j2 has 2k, k ≥ 0, zero roots. Let gf denote the number of zero roots of the real function f . Therefore, for kinds of graph that imply (i) gut = gsv − 1 and gu 6= gt or (ii) gut = gsv + 1 and gu = gt, there is a contradiction and those kinds of graphs do not exist. Lemma 5.3. The following kinds of graphs do not exist: (i) (ηG, ηG, ηG − 1, ηG); (ii) (ηG, ηG + 1, ηG + 1, ηG + 1); (iii) (ηG, ηG, ηG, ηG − 1). Furthermore, if gut = gsv and gut is odd, then a graph of that kind exists if ut − sv is zero at λ = 0, otherwise j2 would have an odd number of zeros. Therefore, if gut = gsv and gut is odd, jb = 0 at λ = 0. Lemma 5.4. Graphs with gut = gsv and gut odd exist provided jb = 0 at λ = 0. They are non–bipartite and of one of the following kinds: (i) (ηG, ηG, ηG − 1, ηG − 1); (ii) (ηG, ηG + 1, ηG, ηG + 1). We shall call kinds (i) and (ii), in Lemma 5.4 above, variety 2b and 3b(i), respectively (See Table 3). Lemma 5.5. If (G, y, z) is a singular graph with gut < gsv and gsv odd, then (G, y, z) is non–bipartite and of kind (ηG, ηG − 1, ηG − 1, ηG − 1). Proof. If y and z are CVs, gut < gsv , then (ηG, ηG−y, ηG−z, ηG−y−z) is (i) (ηG, ηG − 1, ηG − 1, ηG) or (ii) (ηG, ηG − 1, ηG − 1, ηG − 1). Now if furthermore, gsv is given to be odd, then ηG−y−z = ηG − 1. It follows that ηG−y = ηG−y−z Therefore, G is not bipartite. We shall call the graphs in Lemma 5.5 above, variety 1(iii) (See Table 3). 6 Kinds of graphs In this section we determine the properties of a kind (ηG, ηG−y, ηG−z, ηG−y−z) within each of the three varieties. 270 Ars Math. Contemp. 6 (2013) 261–278 6.1 Graphs of variety 1 Graphs of variety 1, are necessarily singular and therefore have at least one core. There are at least two vertices in a core. Lemma 6.1. For a device (G, y, z) of variety 1 and nullity one, jb(0) 6= 0 for core vertices y and z. Proof. For ηG = 1, a non-zero column of the adjugate adj(A) is a kernel eigenvector of G [9]. The non–zero entries occur only at core vertices. Therefore, jb(0) 6= 0. There are three types of pairs of vertices (CV,CV) for graphs of variety 1, depending on the nullity of G − y − z. Since ηG ≥ 1 and gu = gt = ηG − 1, then the nullity gv of G− y− z can take any of the three values ηG− 2, ηG and ηG− 1, corresponding to variety 1(i), 1(ii) and 1(iii), respectively. Theorem 6.2. For a device (G, y, z) of variety 1(iii), j(0) 6= 0 for core vertices y and z. Proof. For nullity one the result follows from Lemma 6.1. Now consider a graph with ηG > 1 of variety 1(iii), that is when gv = ηG − 1. The number of zeros gut of ut is 2ηG− 2 and less than that of sv which is odd. If j2, which is ut− sv, is not to have an odd number of zeros, it follows, from j = jbληG−1 + jaληG , that jb 6= 0 at λ = 0. For variety 1(i), the vertices y and z are CVs. Moreover, without loss of generality, the vertex z is a CV of the subgraph G− y. Only for variety 1(i) is vc 6= 0. Definition 6.3. The connected graphsG in the devices {(G, y, z)}with all pairs of vertices (y, z) ∈ V × V being of variety 1(i) are said to form the class of uniform–core graphs. Equivalently, ηG−y−z = ηG − 2, that is z is a CV of G − y for all vertex pairs (y, z). It is clear that all vertices of a uniform–core graph are CVs, and that they remain so even in a vertex–deleted subgraph G − y for any vertex y of G. Note that this is not the case in general; if y and z are two distinct core vertices of a graph G, then z need not remain a core vertex of G− y. We shall consider uniform–core graphs in more detail in Section 7. 6.2 Graphs of variety 2 In a device (G, y, z) of variety 2, (y, z) is a mixed vertex pair, that is exactly one vertex z of the pair (y, z) is a CV. From Lemmas 5.1 and 5.3, the following result follows immediately. Proposition 6.4. In a device (G, y, z) of variety 2, (i) there is only one kind when y is upper, namely kind (ηG, ηG + 1, ηG − 1, ηG) in variety 2a and (ii) only one kind when y is middle, namely kind (ηG, ηG, ηG − 1, ηG − 1) in variety 2b. From Lemma 5.2, the graphs of variety 2b are non-bipartite. Theorem 6.5. In a device (G, y, z) of variety 2b, the term in λ2ηG−1 of j2 is identically equal to zero. I. Sciriha et al.: Interlacing–extremal graphs 271 Proof. In variety 2b, a graph is of kind (ηG, ηG, ηG−1, ηG−1). The parameter vc vanishes and vb(λ) = ubta s0 6= 0. The number of zeros of ut is the same as that of sv. Therefore, j2 = ut− sv has at least 2ηG − 1 zeros. In variety 2b, the term in λ2ηG−1 in its expansion is ubta − s0vb. Also vc vanishes and vb(λ) = ubta s0 6= 0. Hence, s0vb = ubta and the term in λ2ηG−1 in the expansion of j2 is identically equal to zero, as expected from the fact that j2 is a perfect square. The parameter vb distinguishes between a graph in variety 2a and one in variety 2b. Theorem 6.6. For a graph in variety 2a, vb vanishes at λ = 0. For a graph in variety 2b, vb 6= 0 at λ = 0. Proof. For both kinds in variety 2, ub 6= 0. For an upper vertex, ta = 0 at λ = 0 and for a middle vertex ta 6= 0 at λ = 0. Since s0 6= 0, it follows that for a graph in variety 2a vb = 0 at λ = 0 and, for a graph in variety 2b, vb 6= 0 at λ = 0. 6.3 Graphs of variety 3 We now consider variety 3 for (CFV,CFV) pairs, when tb, ub, jb, vb and vc all vanish. Interlacing provides three types of vertex pairs depending on whether a CFV in the pair (y, z) is upper or middle. When both vertices are upper (variety 3a), by Lemma 5.3 only variety 3a(i) for gv = ηG and variety 3a(ii), when gv = ηG + 2 are allowed. The values at λ = 0 of va or ja suffice to distinguish between graphs of variety 3(i) and 3(ii). Theorem 6.7. For variety 3a(i), both va and ja are non–zero at λ = 0. For variety 3a(ii), both va and ja vanish at λ = 0. Proof. For variety 3, vb = 0. Variety 3a(i) is (ηG, ηG + 1, ηG + 1, ηG). Since v = vaληG and ηG−y−z = ηG, va 6= 0 at λ = 0. Also gj2 = 2ηG so that ja 6= 0 at λ = 0. Variety 3a(ii) is (ηG, ηG+, ηG − 1, ηG − 1). Since gv = ηG + 2, λ2 divides va and λ divides all of the functions ta, ua and ja. For variety 3b, one vertex is upper and one is middle. Interlacing allows only gv = ηG + 1 and ηG, corresponding to variety 3b(i) and variety 3b(ii), respectively. Both vb and jb vanish at λ = 0. The value of ja at λ = 0 distinguishes between variety 3b(i) and variety 3b(ii). Theorem 6.8. For variety 3b(i), ja vanishes at λ = 0. For variety 3a(ii), ja is non–zero at λ = 0. Proof. For variety 3b(i), λ divides ja, as otherwise ut − sv is not the perfect square j2. variety 3b(ii) gv = ηG requires ja 6= 0 at λ = 0. For variety 3c, both vertices are middle. The values at λ = 0 of ta and ua are non– zero. By Lemma 5.3, gv = ηG + 1 or ηG, corresponding to variety 3c(i) and variety 3c(ii), respectively. For variety 3c(ii), when gv = ηG, va is non–zero at λ = 0. Two cases may occur. Either ja 6= 0 at λ = 0 or the number of zeros of ja is at least one. The former case is 272 Ars Math. Contemp. 6 (2013) 261–278 Vertex y Vertex z variety 1 7 variety 1(i) 1 4 variety 1(ii) 1 2 variety 1(iii) 1 15 variety 2a 1 5 variety 2b 17 18 variety 3a(i) 15 17 variety 3a(ii) 5 15 variety 3b(i) 15 16 variety 3b(ii) 11 16 variety 3c(i) 5 6 variety 3c(iiA) 5 17 variety 3c(iiB) Table 2: All varieties and kinds for the same graph G illustrated in Figure 2. denoted by variety 3c(iiA). The latter case is variety 3c(iiB) for which the terms in λ2ηG−2 and in λ2ηG−1 of j2 vanish. The remaining case is for variety 3c(i) when gv = ηG + 1 and λ divides va. Figure 2: A device (G, y, z) of all possible kinds for various (y, z). The graph in Figure 2 exhibits a device (G, y, z) of all varieties and kinds for different choices of (y, z). The classification of devices into kinds and varieties has an application in chemistry in the identification of molecules (with carbon atoms in particular) that conduct or else bar conduction at the Fermi level. In the chemistry paper [3], conductors and insulators are classified into eleven cases that are essentially the twelve kinds of Table 3, with case 7 in [3] corresponding to the kinds (ηG, ηG, ηG, ηG) in variety 3c(iiA) and (ηG, ηG, ηG, ηG) in variety 3c(iiB). The latter two varieties are distinguishable by the non–vanishing or other- wise of ja(0). 7 Graphs with analogous vertex pairs In general, vertex pairs in a graph may be of different varieties and kinds. We shall explore two interesting classes of graphs with the same extremal nullity (allowed by interlacing) for all vertex–deleted subgraphs. These emerge in the classification of devices {(G, y, z)} according to their kind (ηG, ηG−y, ηG−z, ηG−y−z). A pair of vertices y and z for which ηG−y = ηG−z is said to be an analogous vertex pair. I. Sciriha et al.: Interlacing–extremal graphs 273 Kind Characterization Variety G bipartite Two CVs 1 (gs, gt, gu) = (ηG, ηG − 1, ηG − 1) gv = ηG − 2 vc 6= 0 & tb 6= 0 & ub 6= 0 1(i) Allowed & ηG ≥ 2 gv = ηG vc = 0 & tb 6= 0 & ub 6= 0 & vb(0) = 0 1(ii) Allowed & ηG ≥ 1 gv = ηG − 1 vc = 0 & tb 6= 0 & ub 6= 0 & vb(0) 6= 0 1(iii) Forbidden & ηG ≥ 1 CV and CFV 2 (gs, gt, gu) = (ηG, ηG + 1, ηG − 1) vc = 0 & tb = 0 & ub 6= 0 & vb(0) = 0 2a Allowed gv = ηG & ηG ≥ 1 (gs, gt, gu, gv) = (ηG, ηG, ηG − 1) vc = 0 & tb = 0 & ub 6= 0 & vb(0) 6= 0 2b Forbidden gv = ηG − 1 & ηG ≥ 1 Two CFVs 3 (gs, gt, gu) = (ηG, ηG + 1, ηG + 1) 3a gv = ηG vc = 0 & tb = 0 & ub = 0 & vb(0) = 0 3a(i) Allowed & ta(0) = 0 & ua(0) = 0 & va(0) 6= 0 gv = ηG + 2 vc = 0 & tb = 0 & ub = 0 & vb(0) = 0 3a(ii) Allowed & ta(0) = 0 & ua(0) = 0 & va(0) = 0 (gs, gt, gu) = (ηG, ηG + 1, ηG) 3b gv = ηG + 1 vc = 0 & tb = 0 & ub = 0 & vb(0) = 0 3b(i) Forbidden & ta(0) = 0 & ua(0) 6= 0 & va(0) = 0 gv = ηG vc = 0 & tb = 0 & ub = 0 & vb(0) = 0 3b(ii) Forbidden & ta(0) = 0 & ua(0) 6= 0 & va(0) 6= 0 (gs, gt, gu) = (ηG, ηG, ηG) 3c gv = ηG + 1 vc = 0 & tb = 0 & ub = 0 & vb(0) = 0 3c(i) Forbidden & ta(0) 6= 0 & ua(0) 6= 0 & va(0) = 0 gv = ηG vc = 0 & tb = 0 & ub = 0 & vb(0) = 0 3c(ii) Forbidden & ta(0) 6= 0 & ua(0) 6= 0 & va(0) 6= 0 gv = ηG &ja(0) 6= 0 vc = 0 & tb = 0 & ub = 0 & vb(0) = 0 3c(iiA) Forbidden & ta(0) 6= 0 & ua(0) 6= 0 & va(0) 6= 0 & ja(0) 6= 0 gv = ηG& ja(0) = 0 vc = 0 & tb = 0 & ub = 0 & vb(0) = 0 3c(iiB) Forbidden & ta(0) 6= 0 & ua(0) 6= 0 & va(0) 6= 0 & ja(0) = 0 Table 3: A characterization of all devices (G, y, z) according to their variety and kind. 274 Ars Math. Contemp. 6 (2013) 261–278 The first of these two classes consists of graphs G with the minimum possible nullity ηG−y−z for all pairs of distinct vertices y and z, (i.e., ηG − 2) and therefore also the minimum possible nullities ηG−y and ηG−z (i.e., ηG − 1). By Definition 6.3, these graphs form precisely the class of uniform–core graphs. On the other hand, the second of the two classes consists of graphs with the maximum possible nullity ηG−y−z , that is ηG + 2, for some pair of distinct vertices y and z, and therefore also the maximum possible nullities ηG−y and ηG−z (i.e., ηG + 1). 7.1 Uniform–core graphs By Definition 6.3, each vertex pair in a uniform–core graph corresponds to a graph of variety (1i). Since the nullity of a graph is non–negative, and ηG−y−z = ηG − 2 for all vertex pairs y, z of a uniform–core graph G, then the nullity of G is at least two. To understand better the core–structure of uniform–core graphs and be able to characterize them as a subclass of singular graphs, it is necessary to use their core structure with respect to a basis for their nullspace. Let B be a basis for the η–dimensional nullspace of A of a singular graph G (with no isolated vertices) of nullity η ≥ 1. As seen in [11], Hall’s Marriage problem for sets, or the Rado–Hall Theorem for matroids, guarantees a vertex–subset S of distinct vertex representatives [1, 11], to represent a system SCores of cores corresponding to the vectors of B. This implies that deleting a vertex v representing a core F eliminates the core F from G− v, which will now have a new system of η − 1 cores. Also any k ≥ 1 cores in a system SCores of ηG cores cover at least k + 1 vertices. Theorem 7.1. A device (G, y, z) is of variety 1(i) if and only if the two vertices y and z do not lie in one core only, i.e. at least two cores are needed to cover the vertices y and z. Proof. Consider a basis B for the nullspace of A. The vertices y and z lie on at least one core of G. There are two possibilities. Firstly, B has exactly one vector with non–zero entries at positions associated with y and z. In this case ηG−y−z = ηG−y = ηG− 1, which does not correspond to variety 1(i). Secondly, B has at least two vectors with non–zero entries at positions associated with y or z, when ηG−y−z = ηG−y − 1 = ηG − 2, which corresponds to variety 1(i). The two core vertices must represent two distinct cores in a system SCores of ηG cores corresponding to a basis B for the nullspace [11]. A subclass U of uniform–core graphs can be constructed from nut graphs. A graph G ∈ U is obtained from a nut graph H on n vertices and m edges by duplicating each of the n vertices of H . Then G has 2n vertices and 4m edges. Figure 3 shows the uniform– core graph G ∈ U obtained from the smallest nut graph H . The nullity of G is |V(G)| 2 +1. Deletion of any |V(G)| 2 + 1 vertices reduces the graph to a non–singular graph. Let the vertices of G be labelled 1, 2, ..., n, 1′, 2′, ...n′ where {1, 2, ...} are the vertices of the nut graph H and {1′, 2′, ...} are the duplicate vertices of {1, 2, ...} in that order in G. Note that a vertex labelled r for 1 ≤ r ≤ n is adjacent to the original neighbours in H and also to precisely those primed vertices with the same numeric label. For instance, vertex 1 is adjacent to 2 and 7 in H and to 2, 2’, 7 and 7’ in G. The following result, expressing the adjacency matrix of G ∈ U in terms of the adjacency matrix of H , is immediate. I. Sciriha et al.: Interlacing–extremal graphs 275 Figure 3: The smallest nut graph H and the uniform–core graph G derived from H . Theorem 7.2. If H is the adjacency matrix of the nut graph H , then the adjacency matrix of the uniform–core graph G ∈ U is ( H H H H ) . The spectrum of G consists of n eigen- values equal in value to double the eigenvalues of H and an additional n zero eigenvalues corresponding to the n duplicate vertex pairs. If (x1, x2, · · · , xn)t is an eigenvector of H for an eigenvalue µ, then (x1, x2, · · · , xn, x1, x2, · · · , xn)t is an eigenvector of G for an eigenvalue 2µ. We shall now characterize uniform–core graphs by requiring that a set of vertex repre- sentatives of a system SCores of cores be an arbitrary subset of the vertices for all systems of cores. Theorem 7.3. A graph of nullity ηG is a uniform–core graph if and only if it is a singular graph such that the deletion of any subset of ηG vertices produces a non–singular graph. Proof. Let us relate the nullspace of A to the vertices of a uniform–core graph G of nullity ηG. Let S be any subset of ηG vertices of G labelled {1, 2, · · · , ηG} and let B be an ordered basis for the nullspace of A. If all pairs of vertices give a graph of variety 1(i), then no two vertices lie in only one core of SCores. Therefore, it is possible to obtain a new ordered basis B′ for the nullspace of A, by linear combination of the vectors in B, such that, for 1 ≤ i ≤ ηG, only the vector i of B′ has a non–zero entry at position i [11]. Removal of any vertex in S destroys precisely one eigenvector of B′ reducing the nullity by one. Deletion of all the vertices in S destroys all the kernel eigenvectors and leaves a non–singular graph. A characterization of the subclass G ∈ U of uniform–core graphs uses the operation NEPS (non–complete extended p-sum) of a nut graph and K2. The graph product NEPS is described for instance in [2]. Definition 7.4. Given a collection {G1, G2, · · · , Gk, · · · , Gn} of graphs and a correspond- ing set B ⊆ {0, 1}n\{(0, 0, ..., 0)}, called the basis, of non–zero binary n-tuples, the NEPS of G1, G2, ..., Gn is the graph with vertex set V(G1) × V(G2) × · · · × V(Gn) in which two vertices {w1, w2, · · · , wn} and {y1, y2, · · · , yn} are adjacent if and only if there ex- ists (β1, β2, · · · , βn) ∈ B such that wi = yi whenever βi = 0 and wi is adjacent to yi whenever βi = 1. 276 Ars Math. Contemp. 6 (2013) 261–278 Lemma 7.5. [2] If for 1 ≤ i ≤ n, λi1, λi2, · · · , λini is the spectrum of Gi, of order ni for 1 ≤ i ≤ n, then the spectrum of the NEPS of G1, G2, · · · , Gn with respect to basis B is { ∑ β∈B λβ11i1 , λ β2 2i2 , · · · , λβnnin : ik = 1, 2, ..., nk & k = 1, 2, ..., n}. The following result follows from the construction of a uniform–core graph G ∈ U . Theorem 7.6. A uniform–core graph G ∈ U is the NEPS of a nut graph G1 and K2 with basis {(1, 0), (1, 1)}. From Lemma 7.5 and Theorem 7.6, the spectrum of the uniform–core graph G ∈ U is λi + λiλj where {λi} is the spectrum of the nut graph H and {λj} = {1,−1} is the spectrum of K2. This agrees with the result in Theorem 7.2. 7.2 Non-singular graphs with a complete weighted inverse We shall now look into the second class of devices. Such a graph G is a device (G, y, z), of variety 3a(ii), for some pair of distinct vertices y and z. Graphs which are devices (G, y, z), of variety 3a(ii), for a particular pair of vertices y and z exist, as shown in the example of Figure 2 for vertex connections 15 and 17. Can a graphG be a device (G, y, z), of variety 3a(ii), for all vertex pairs {y, z}? The question amounts to determining whether it is possible to have η(G− y − z) equal to the maximum allowed nullity relative to η(G), that is η(G) + 2, for all vertex pairs {y, z}. The answer is in the negative. Lemma 7.7. It is impossible that a graph G is a device (G, y, z) of variety 3a(ii) for all pairs of distinct vertices y and z. Proof. Suppose G is a graph which is a device (G, y, z) of variety 3a(ii) for all pairs of distinct vertices y and z. This requires that each of the graphs G− y and G− z is singular and therefore has CVs. Deletion of a CV from G − y, restores the nullity back to η(G). Hence it is impossible to achieve η(G− y − z) = η(G) + 2, for all vertex pairs {y, z}. By Lemma 5.3(ii), the kind (ηG, ηG−y, ηG−z, ηG−y−z)=(ηG, ηG+1, ηG, ηG+1) is im- possible. Hence the only devices (G, y, z) within the second class that have the maximum value of η(G − y) relative to ηG, for all vertices y, are of kind (ηG, ηG + 1, ηG + 1, ηG). Our focus is on the non–singular graphs of this kind having the inverse A−1 equal to the adjacency matrix of the complete graph with real non–zero weighted edges and no loops. The smallest candidate is K2. Indeed A(K2) = A(K2))−1 = ( 0 1 1 0 ) . Definition 7.8. Let G be a non–singular graph G with the off–diagonal entries of the in- verse A−1 of its adjacency matrix A being non–zero and real, and all the diagonal entries of A−1 being zero. Then G is said to be a nuciferous graph. The motivation for the name nuciferous graph (meaning nut–producing graph) will become clear from Theorem 7.9. To characterize this class of graphs, let us consider the deck {G−v : v ∈ V} of subgraphs, as in the investigation of the polynomial reconstruction problem [10]. Theorem 7.9. Let G be a nuciferous graph. Then G is either K2 or each vertex–deleted subgraph G− v is a nut graph. I. Sciriha et al.: Interlacing–extremal graphs 277 Proof. Let Q be the n − 1 × n matrix obtained from A−1 by suppressing the diagonal entry from each column. Therefore each entry of Q is non–zero. Let the ith column of Q be qi := (q(1)i, q(2)i, ..., q(i−1)i, q(i+1)i, q(i+2)i, ..., q(n)i)t for 2 ≤ i ≤ n − 1. The first and last columns are q1 := (q(2)1, q(3)1, ..., q(n)1)t and qn := (q(1)n, q(2)n, ..., q(n−1)n) t, respectively. Since AA−1 is the identity matrix I, then A(G−i)qi = 0 for all 1 ≤ i ≤ n. Therefore qi is a kernel eigenvector (with non–zero entries) of G − i for all the vertices i. Hence G− i is a core graph. By interlacing, it has nullity one. It follows that each vertex–deleted subgraph is a nut graph. From Lemma 7.7, nuciferous devices (G, y, z) are not of type of variety 3a(ii) for all pairs of distinct vertices y and z. Moreover, from Theorem 7.9, for G 6= K2, each vertex– deleted subgraph is a nut graph and therefore has nullity one. On deleting a vertex from a nut graph, the nullity becomes zero. Hence a candidate graph G cannot be of variety 3a(ii) for any pair of vertices y and z. Theorem 7.10. Let G be a nuciferous graph G. If G is not K2, then (i) it has order at least eight; (ii) the device (G, y, z) is of variety 3a(i) for all pairs of distinct vertices y and z; (iii) the graph G is not bipartite. Proof. (i) Since nut graphs exist for order at least seven [12], it follows, from Theorem 7.9, that a nuciferous graph G, of order at least three, has at least eight vertices. (ii) From the proof of Lemma 7.7, a nuciferous graph G is of kind (ηG, ηG + 1, ηG + 1, ηG). Thus G is a device (G, y, z) of variety 3a(i) for all pairs of distinct vertices y and z. (iii) From Theorem 7.9,G−y andG−z are nut graphs and therefore cannot be bipartite [12]. Hence G has odd cycles and cannot be bipartite. To date, no graph (except K2) has been found to satisfy the condition of Theorem 7.9. An exhaustive search on all graphs on up to 10 vertices and all chemical graphs on up to 16 vertices reveals no counter example. We conjecture the following result. Conjecture 7.11. There are no graphs for which every vertex–deleted subgraph is a nut graph. 8 Chemical implications Graph theory has strong connections with the study of physical and chemical properties of all-carbon frameworks such as those in benzenoids, fullerenes and carbon nanotubes. The eigenvalues and eigenvectors of the adjacency matrix of the molecular graph (the graph of the carbon skeleton) are used in qualitative models of the energies and spatial distributions of the mobile π electrons of such systems. Specifically, graphs and their nullities figure in simple theories of ballistic conduction of electrons by conjugated systems. In the sim- plest formulation [3] of the SSP (Source and Sink Potential ) [5] approach to molecular conduction, the variation of electron transmission with energy is qualitatively modelled in terms of the characteristic polynomials of G, G − y, G − z, G − y − z, where G is the molecular graph and vertices y and z are in contact with wires. (This is the motivation for the definition of a device in the present paper.) As a consequence, the transmission at the Fermi level (corresponding here to λ = 0) obeys selection rules couched in terms of 278 Ars Math. Contemp. 6 (2013) 261–278 the nullities ηG, ηG−y , ηG−z , and ηG−y−z [7], motivating the definition of kinds here. In terms of the varieties defined here, the SSP theory predicts conduction at the Fermi level for connection across the vertex pair (y, z) for 1(ii), 1(iii), 3a(i), 3b(ii), 3c(i) and 3c(iiA), and, conversely, insulation at the Fermi level for 1(i), 2a, 2b, 3a(ii), 3b(i) and 3c(iiB). The two classes of graphs with analogous vertex pairs and certain extremal conditions on the nullity of their vertex-deleted subgraphs, explored in Section 7 are envisaged to have interesting developments in spectral graph theory. Moreover, the classification of graphs into varieties and kinds has an application in chemistry in the identification of molecules (with carbon atoms in particular) that conduct or else bar conduction at the Fermi level that has already been investigated in [3]. According to the SSP theory, the first class, the uniform–core graphs, corresponds to insulation at the Fermi–level for all two vertex con- nections and the second class, the nuciferous graphs, to Fermi–level conducting devices (G, y, z) for all pairs of distinct vertices y and z. The latter class has the additional prop- erties that it consists of devices corresponding to non-singular graphs that are Fermi–level insulators when y = z. Therefore nuciferous graphs have no non–bonding orbital and are conductors for all distinct vertex connection pairs and insulators for all one vertex connec- tions. We conjecture that the only nuciferous graph is K2. References [1] R. A. Brualdi, The mutually beneficial relationship of graphs and matrices, volume 115 of CBMS Regional Conference Series in Mathematics. American Mathematical Society, Provi- dence, RI, 2011. [2] D. M. Cvetković, P. Rowlinson and S. Simić, Eigenspaces of Graphs, CUP, Cambridge, 1997. [3] P. W. Fowler, B. T. Pickup, T. Z. Todorova and W. Myrvold, A selection rule for molecular conduction, J. Chem. Phys. 131 (2009), 044104 (7 pages). [4] F. R. Gantmacher, The Theory of Matrices, volume 1, AMS Chelsea, 2010. [5] F. Goyer, M. Ernzerhof and M. Zhuang, Source and sink potentials for the description of open systems with a stationary current passing through. J. Chem. Phys., 126 (2007), 144104. [6] M. Marcus and H. Minc, A Survey of Matrix Theory and Matrix Inequalities, volume 115 of CBMS Regional Conference Series in Mathematics, Allyn and Bacon, Inc., Boston, Mass., 1964. [7] B. T. Pickup and P. W. Fowler, An analytical model for steady-state currents in conjugated systems, Chem. Phys. Lett. 459 (2008), 198–202. [8] A. J. Schwenk and R. J. Wilson, On the eigenvalues of a graph, in: L. W. Beineke, R. J. Wilson (eds), Selected Topics in Graph Theory I, Academic Press, London, 1978. [9] I. Sciriha, On the coefficient of λ in the characteristic polynomial of singular graphs, Util. Math., 52 (1997), 97–111. [10] I. Sciriha, Graphs with a common eigenvalue deck, Linear Algebra Appl. 430 (2009), 78–85. [11] I. Sciriha, Maximal core size in singular graphs, Ars Math. Contemp. 2 (2009), 217–229. [12] I. Sciriha and I. Gutman, Nut graphs – maximally extending cores, Util. Math. 54 (1998), 257–272.