Also available at http://amc.imfm.si ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 6 (2013) 171–185 Sharp spectral inequalities for connected bipartite graphs with maximal Q-index Milica And̄elić ∗ Department of Mathematics, University of Aveiro, 3810-193 Aveiro, Portugal Faculty of Mathematics, University of Belgrade, 11 000 Belgrade, Serbia C. M. da Fonseca † Department of Mathematics, University of Coimbra, 3001-501 Coimbra, Portugal Tamara Koledin Faculty of Electrical Engineering, University of Belgrade, 11 000 Belgrade, Serbia Zoran Stanić ‡ Faculty of Mathematics, University of Belgrade, 11 000 Belgrade, Serbia Received 31 October 2011, accepted 26 December 2011, published online 11 July 2012 Abstract The Q-index of a simple graph is the largest eigenvalue of its signless Laplacian. As for the adjacency spectrum, we will show that in the set of connected bipartite graphs with fixed order and size, the bipartite graphs with maximal Q-index are the double nested graphs. We provide a sequence of (in)equalities regarding the principal eigenvector of the signless Laplacian of double nested graphs and apply these results to obtain some lower and upper bounds for their Q-index. In the end, we give some computational results in order to compare these bounds. Keywords: Double nested graph, signless Laplacian, largest eigenvalue, spectral inequalities. Math. Subj. Class.: 05C50 ∗Research supported by CIDMA - Center for Research and Development in Mathematics and Applications, FCT - Fundação para a Ciência e a Tecnologia, through European program COMPETE/FEDER and Serbian Ministry of Education and Science, Project 174033. †Research supported by CMUC - Centro de Matemática da Universidade de Coimbra and FCT - Fundação para a Ciência e a Tecnologia, through European program COMPETE/FEDER ‡Research supported by Serbian Ministry of Education and Science, Projects 174012 and 174033. E-mail addresses: milica.andelic@ua.pt (Milica And̄elić), cmf@mat.uc.pt (C. M. da Fonseca), tamara@etf.rs (Tamara Koledin), zstanic@math.rs (Zoran Stanić) Copyright c© 2013 DMFA Slovenije 172 Ars Math. Contemp. 6 (2013) 171–185 1 Introduction Let G = (V,E) be a (simple) graph, of order ν = |V | and size  = |E|. The signless Laplacian of G is defined to be the matrix Q = A+D, where A(= A(G)) is the adjacency matrix of G, while D(= D(G)) is the diagonal matrix of its vertex degrees. The largest eigenvalue (or spectral radius) of Q is usually called the Q-index of G, and is denoted by κ(= κ(G)). Much interest has been paid recently to this very important spectral invariant. Let us recall that ∆ + 1 6 κ 6 2(ν − 1) , (1.1) where ∆ denotes the maximal vertex degree of the graph, with equality for stars, on the lower bound, and complete graphs, on upper bound [7]. In 2007, Cvetković, Rowlinson, and Simić [6] conjectured that κ 6 ν − 1 + d̄ , (1.2) where d̄ is the average (vertex) degree of a graph. Later Feng and Yu [9] proved that (1.2) is true (cf. also [1]). Many other bounds on Q-index for arbitrary (connected) graphs can be found in [8]. We will now describe in brief the structure of a connected double nested graph (or DNG for short). It was first considered in [3, 4] and, independently, under the name of chain graph, in [5], in studying graphs whose least eigenvalue is minimal among the connected (bipartite) graphs of fixed order and size. The vertex set of any such graph G consists of two colour classes (or co-cliques). To specify the nesting, both of them are partitioned into h non-empty cells h⋃ i=1 Ui and h⋃ i=1 Vi, respectively; all vertices in Us are joined (by cross edges) to all vertices in h+1−s⋃ k=1 Vk, for s = 1, 2, . . . , h. Denote by NG(w) the set of neighbors of a vertex w. Hence, if u′ ∈ Us+1, u′′ ∈ Us, v′ ∈ Vt+1, and v′′ ∈ Vt, then NG(u ′) ⊂ NG(u′′) and NG(v′) ⊂ NG(v′′). This claim makes precise the double nesting property. Observe that 1 6 s, t 6 h. If ms = |Us| and ns = |Vs|, with s = 1, 2, . . . , h, then G is denoted by DNG(m1,m2, . . . ,mh;n1, n2, . . . , nh) . Note that G is connected whenever m1, n1 > 0. Additionally, if some of the remain- ing parameters are equal to zero, we again get a DNG with a smaller value of h. Thus, throughout we assume that all these parameters are greater than zero. We now introduce some notation to be used later on. Let Ms = s∑ i=1 mi and Nt = t∑ j=1 nj , for 1 6 s, t 6 h. Thus G is of order ν = Mh+Nh and size  = ∑h s=1msNh+1−s. Observe that Nh+1−s is the degree of a vertex u ∈ Us; the degree of a vertex v ∈ Vt is equal to Mh+1−t. We also set Ms,t = Mt −Ms−1 and, additionally, M1,t = Mt. M. And̄elić et al.: Sharp spectral inequalities for connected bipartite graphs. . . 173 j z ppp ppp j z j z     Q Q Q Q Q L L L L L L L L L L L L Uh Vh U2 V2 U1 V1 Figure 1: The structure of a double nested graph. 2 Extremal bipartite graphs Let G be a bipartite graph with colour classes U and V . First, we state the main result of this section. Theorem 2.1. If G is a graph for which κ(G) is maximal among all connected bipartite graphs of order ν and size , thenG is a DNG with all pendant edges attached to a common vertex. Theorem 2.1 means that double nested graphs play the same role among bipartite graphs (with respect to the signless Laplacian index) as nested split graphs among non-bipartite graphs. The same classes of graphs appear as extremal with respect to the adjacency spectra as well, i.e., in the class of all connected (resp. all connected bipartite) graphs of fixed order and size, those with maximal radius with respect to the adjacency matrix are NSGs (resp. DNGs). The proof of Theorem 2.1 is based on the following lemmas, the first of which is taken from [6]. Recall that there exists a unique unit eigenvector corresponding to κ(G) having only positive entries; this eigenvector is called the principal eigenvector of G. Lemma 2.2. Let G′ be the graph obtained from a connected graph G by rotating the edge rs around r to the non-edge position rt. Let x = (x1, x2, . . . , xν)T be the principal eigenvector of G. If xt > xs then κ(G′) > κ(G). The next lemma will be very helpful when we find a bridge in a graph whose index is assumed to be maximal. Given two rooted graphs P (= Pu) and Q(= Qv) with u and v as roots, let G be the graph obtained from the disjoint union P ∪̇Q by adding the edge uv. Let G′ be the graph obtained from the coalescence of Pu and Qv by attaching a pendant edge at the vertex identified with u and v. Lemma 2.3. With the above notation, if P and Q are two non-trivial connected graphs then κ(G) < κ(G′). Proof. Let (x1, x2, . . . , xν)T be the principal eigenvector of G. Without loss of generality, we may suppose that xu 6 xv . Let NP (u) be the neigbourhood of u in P ; since P is 174 Ars Math. Contemp. 6 (2013) 171–185 non-trivial, NP (u) 6= ∅. Now G′ is obtained from G by replacing the edges uw, with w ∈ NP (u) by the edges vw and, therefore, κ(G) < κ(G′), by Lemma 2.2, as required. In what follows we assume that G has maximal index among the connected bipartite graphs of fixed order and size. Lemma 2.4. LetG be a graph satisfying the above assumptions, and let x = (x1, . . . , xν)T be the principal eigenvector of G. If v and w are vertices in the same colour class such that xv > xw, then deg(v) > deg(w). Proof. Let U, V be the colour classes of G. Assuming that v and w are vertices in V such that xv > xw and deg(v) < deg(w), then deg(w) > 1 and there exists u ∈ U such that v 6∼ u ∼ w. By Lemma 2.3, we may rotate uw to uv to obtain a graph G′ such that κ(G′) > κ(G). If uw is a bridge, then deg(u) = 1 and, again by Lemma 2.3, G′ is necessarily connected; but now the maximality of κ(G) is contradicted and the proof is complete. From now on we take the colour classes to be U = {u1, u2, . . . , um} and V = {v1, v2, . . . , vn}, with xu1 > xu2 > · · · > xum and xv1 > xv2 > · · · > xvn . By Lemma 2.4, this ordering coincides with the ordering by degrees in each colour class. In the next lemma we note some consequences of those facts. Lemma 2.5. Let G be a graph satisfying the above assumptions including those on vertex ordering. Then (i) the vertices u1 and v1 are adjacent; (ii) u1 is adjacent to every vertex in V , and v1 is adjacent to every vertex in U ; (iii) if the vertex u is adjacent to vk, then u is adjacent to vj , for all j < k, and if the vertex v is adjacent to uk, then v is adjacent to uj , for all j < k. Proof. First we consider bridges in G. By Lemma 2.3, all bridges are pendant edges. By Lemma 2.2, all pendant edges are attached at the same vertex, and this vertex w is such that xw is maximal. Without loss of generality, xu1 > xv1 and w = u1. It follows that the result holds if G is a tree and, consequently, G is a star. Accordingly, we suppose that G is not a tree. To prove (i), suppose by way of contradiction that u1 6∼ v1. Then v1 is adjacent to some vertex u ∈ U , and uv1 is not a bridge. By Lemma 2.2, we may rotate v1u to v1u1 to obtain a connected bipartite graph G′ such that κ(G′) > κ(G), contradicting the maximality of κ(G). To prove (ii), suppose that u is a vertex of U not adjacent to v1. Then u 6= u1 by (i), uv is not a bridge, and u is adjacent to some vertex v in V other than v1. Now we can rotate uv to uv1 to obtain a contradiction as before. Secondly, suppose that v is a vertex of V not adjacent to u1. Then v 6= v1 by (i), again vu1 is not a bridge, and a rotation about v yields a contradiction. To prove (iii), suppose that u ∈ U , u ∼ vk and u 6∼ vj for some j < k. Now u 6= u1 by (ii), and so uvk is not a bridge. Then we can rotate uvk to uvj to obtain a contradiction. Finally, suppose that v ∈ V, v ∼ vk and v 6∼ uj for some j < k. In this case, vuk is not a bridge because k > 1, and the rotation of vuk to vuj yields a contradiction. The proof is now finished. M. And̄elić et al.: Sharp spectral inequalities for connected bipartite graphs. . . 175 Taking into account Lemma 2.5 and the definition of a DNG the first part of Theorem 2.1 follows. It remains only to prove that all cut-edges in the observed DNG are pendant edges attached to a common vertex. This easily comes from Lemma 2.3. 3 Q-eigenvectors of DNGs Here we consider the principal eigenvector of the signless Laplacian of DNGs. In this section (and in the next one, if not told otherwise) we will assume that x = (x1, . . . , xν) T is a Q-eigenvector of G with all positive entries, which is usually normalized, i.e., ν∑ i=1 xi = 1 . The entries of x are also called the weights of the corresponding vertices. We first observe that all vertices within the sets Us or Vt, for 1 6 s, t 6 h, have the same weights, since they belong to the same orbit of G. Let xu = as, if u ∈ Us, while xv = bt, if v ∈ Vt. From the eigenvalue equations for κ, applied to any vertex from Us or Vt, we get κas = Nh+1−sas + h+1−s∑ j=1 njbj , for s = 1, . . . , h, (3.1) and κbt = Mh+1−tbt + h+1−t∑ i=1 miai , for t = 1, . . . , h. (3.2) By normalization we have h∑ i=1 miai + h∑ j=1 njbj = 1 , (3.3) and, from (3.1), we easily get as = 1 κ−Nh+1−s h+1−s∑ j=1 njbj , for s = 1, . . . , h. (3.4) From (3.2) we have bt = 1 κ−Mh+1−t h+1−t∑ i=1 miai, for t = 1, . . . , h, (3.5) and therefore, using (3.3), we have as = 1 κ−Nh+1−s 1− h∑ i=1 miai − h∑ j=h+2−s njbj  , for s = 1, . . . , h, (3.6) 176 Ars Math. Contemp. 6 (2013) 171–185 or, using (3.2) for t = 1, as = 1 κ−Nh+1−s 1− (κ−Mh)b1 − h∑ j=h+2−s njbj  , for s = 1, . . . , h. Similarly, bt = 1 κ−Mh+1−t ( 1− (κ−Nh)a1 − h∑ i=h+2−t miai ) , for t = 1, . . . , h. Setting ah+1 = bh+1 = 0 and N0 = 0, from (3.4) and (3.6), together with (3.3), we get successively (κ−Nh−s)as+1 − (κ−Nh+1−s)as = −nh+1−sbh+1−s, for s = 1, . . . , h− 1, and (κ− n1)ah = n1b1, for s = h. Since all components of x are positive and κ > ∆ + 1 (1.1), it comes as+1 6 as, for s = 1, . . . , h− 1, (3.7) and bt+1 6 bt, for t = 1, . . . , h− 1. (3.8) Furthermore, by setting s = h in (3.2), we obtain (κ−m1)bh = m1a1. (3.9) Moreover, substituting s = 1 in (3.1) and t = 1 in (3.2) and applying in (3.3) we get (κ−Nh)a1 + (κ−Mh)b1 = 1, and finally as = 1 κ−Nh+1−s (κ−Nh)a1 − h∑ j=h+2−s njbj  . (3.10) Next we focus our attention on bounding ai’s and bj’s. Lemma 3.1. For any s = 1, . . . , h, we have Nh+1−sbh+1−s κ−Nh+1−s 6 as 6 Nh+1−sb1 κ−Nh+1−s . (3.11) Proof. From (3.4), we have as = 1 κ−Nh+1−s h+1−s∑ j=1 njbj . Therefore, (3.11) immediately follows since bj’s are strictly decreasing, from (3.8). M. And̄elić et al.: Sharp spectral inequalities for connected bipartite graphs. . . 177 Lemma 3.2. For any s = 1, . . . , h, as 6 a1 ( 1− Nh+2−s,h κ−Nh+1−s ( 1 + m1 κ−m1 )) , (3.12) Proof. The inequality (3.12) follows from (3.10), since bi’s are strictly decreasing, bearing in mind (3.9) as well. Lemma 3.3. For any s = 1, . . . , h, as > a1 κ−Nh+1−s ( 1− s−1∑ i=1 nh+1−iMi κ−Mi ) . Proof. By induction on s. For s = 1, the inequality holds trivially. Assume next that as > a1 κ−Nh+1−s ( 1− s−1∑ i=1 nh+1−iMi κ−Mi ) , for s > 1. Then as+1 = 1 κ−Nh−s h−s∑ j=1 njbj = 1 κ−Nh−s ((κ−Nh+1−s)as −Nh+1−sbh+1−s) > a1 κ−Nh−s ( 1− s−1∑ i=1 nh+1−iMi κ−Mi ) − Nh+1−sMsa1 (κ−Nh−s)(κ−Ms) = a1 κ−Nh−s ( 1− s∑ i=1 nh+1−iMi κ−Mi ) . This ends the proof. Lemma 3.4. For any s = 1, . . . , h, we have as 6 b1 κ−Nh+1−s ( Nh+1−s − κfh+1−s (κ− n1)(κ−Ms) ) , (3.13) where fh+1−s = h+1−s∑ j=1 njMh+2−j,h . Proof. From (3.4) and (3.12) applied to bj , we get as = 1 κ−Nh+1−s h+1−s∑ j=1 njbj 6 1 κ−Nh+1−s h+1−s∑ j=1 njb1 ( 1− Mh+2−j,h κ−Mh+1−j ( 1 + n1 κ− n1 )) 6 b1 κ−Nh+1−s ( Nh+1−s − κfh+1−s (κ− n1)(κ−Ms) ) . The proof is now complete. 178 Ars Math. Contemp. 6 (2013) 171–185 4 Some bounds on the Q-index of a DNG In this section we obtain some upper and lower bounds on the Q-index of DNGs using the eigenvalue and the matrix technique. We also emphasize that our main goal is to consider the estimation of theQ-index of large graphs. Before we proceed, we provide the following observations. First, if h = 1 we get a complete bipartite graph Km1,n1 , whose Q-index is equal to m1 + n1 = ν [2]. Furthermore, since the Q-index of an arbitrary graph increases by inserting edges (cf. [6]), we have κ 6 ν, (4.1) for any (not necessarily connected) DNG. Otherwise, if h > 1 is fixed but the graph size is not, using the same previous ar- guments, the maximal Q-index would appear in DNG(m1, 1, . . . , 1;n1, 1, . . . , 1). The computational results suggest this will happen when |m1−n1| 6 1. So, these cases are not interesting for our research and, therefore, we will assume that h > 1 and the size is fixed. 4.1 Eigenvalue technique Now we establish some bounds for the Q-index of DNGs using the eigenvalue technique. We start with lower bounds. Proposition 4.1. If G is a connected DNG, then κ > max 16k6h {Mh+1−k +Nk}. Proof. On the one hand, from (3.2), we get bk = 1 κ−Mh+1−k h+1−k∑ i=1 miai > Mh+1−kah+1−k κ−Mh+1−k , since, from (3.7), ai’s are decreasing. On the other hand, from (3.1), we get ah+1−k = 1 κ−Nk k∑ j=1 njbj > Nkbk κ−Nk , since bj’s are decreasing, from (3.8). From the last two inequalities we get κ(κ− (Mh+1−k +Nk)) > 0, which is equivalent to κ >Mh+1−k +Nk. In particular, for k = h and k = 1, we obtain the following corollary. Corollary 4.2. If G is a connected DNG, then κ > m1 +Nh and κ > n1 +Mh . M. And̄elić et al.: Sharp spectral inequalities for connected bipartite graphs. . . 179 Proposition 4.3. If G is a connected DNG, then κ > 1 2 t+  Nh + √( t−  Nh )2 + 4ê∗h  , where t = ∑h i=1miN 3 h+1−i∑h i=1miN 2 h+1−i and ê∗h = h∑ i=1 mi N2h+1−i Nh . Proof. Let y = (y1, . . . , yν)T be a vector whose components are indexed by the vertices of G, and let yu = Nh+1−i if u ∈ Ui, for some i ∈ {1, . . . , h}, or, otherwise, yv = q = κ− t, for some t, if v ∈ Vj for some j ∈ {1, . . . , h}. Substituting y into the Rayleigh quotient (see, e.g., [8, p. 49]) we obtain κ > 2 ∑h i=1miN 2 h−1+iq + ∑h i=1miN 3 h+1−i + ∑h i=1 niMh−1+iq 2∑h i=1miN 2 h+1−i +Nhq 2 due to Rayleigh’s principle which reads yTQy yT y 6 κ. Since q = κ− t, we get Nhq 3 + (Nht− )q2 − h∑ i=1 miN 2 h+1−iq > h∑ i=1 miN 3 h+1−i − t h∑ i=1 miN 2 h+1−i . Choosing t = ∑h i=1miN 3 h+1−i∑h i=1miN 2 h+1−i , and having in mind that N1 6 t 6 Nh, we immediately get a quadratic inequality in q and the proof is concluded. Proposition 4.4. If G is a connected DNG, then κ 6 1 2 ( ν + √ ν2 − 4(MhNh − ) ) . (4.2) Proof. From (3.1), with s = h, and from (3.3), recalling (3.11), we get (κ−Mh)b1 = h∑ i=1 miai 6 h∑ i=1 mi Nh+1−i κ−Nh+1−s b1 . Then, we obtain (κ−Mh)(κ−Nh) 6 , and, therefore, from the quadratic inequality κ2 − (Mh +Nh)κ+MhNh −  6 0, we obtain κ1 6 κ 6 κ2 where κ1 and κ2 are the solutions of the associated quadratic equality, and this completes the proof. 180 Ars Math. Contemp. 6 (2013) 171–185 The next two bounds improve the bound (4.2). We recall that fh+1−i is defined in Lemma 3.4. Proposition 4.5. If G is a connected DNG, then κ 6 1 2 ( ν + √ ν2 − 4(MhNh − ′) ) , where ′ = − ν(ν −Nh) (ν − n1)2(ν −m1) h∑ i=1 mifh+1−i . Proof. As in the proof of Proposition 4.4, we have (κ−Mh)b1 = h∑ i=1 miai . Using (3.12), we get κ−Mh 6 h∑ i=1 mi κ−Nh+1−i ( Nh+1−i − κfh+1−i (κ− n1)(κ−Mi) ) , and therefore (κ−Mh)(κ−Nh) 6 − κ(κ−Nh) (κ− n1)2(κ−m1) h∑ i=1 mifh+1−i . Taking into account that κ 6 ν, from Proposition 4.4 it follows (κ−Mh)(κ−Nh) 6 ′, and the proof ends. The next result may be proved in a similar way. Proposition 4.6. If G is a connected DNG, then κ 6 1 2 ( ν + √ ν2 − 4(MhNh − ′′) ) , where ′′ = − κ ′(κ′ −Nh) (κ′ − n1)2(κ′ −m1) h∑ i=1 mifh+1−i, for κ′ = 1 2 ( ν + √ ν2 − 4(MhNh − ′) ) . M. And̄elić et al.: Sharp spectral inequalities for connected bipartite graphs. . . 181 4.2 Matrix technique The partition V = h⋃ k=1 Uk ∪ h⋃ k=1 Vk (4.3) is equitable since every vertex in Ui and every vertex in Vi have the same number of neigh- bors in Uj and Vj , for all i, j ∈ {1, 2, . . . , h}. Let AD be the signless Laplacian divisor matrix of a DNG(m1, . . . ,mh;n1, . . . , nh) with respect to the equitable partition (4.3). The matrix AD has the following form: AD =  Nh n1 n2 · · · nh−1 nh Nh−1 n1 n2 · · · nh−1 . . . ... ... ... N2 n1 n2 N1 n1 m1 m2 · · · mh−1 mh Mh m1 m2 · · · mh−1 Mh−1 ... ... ... . . . m1 m2 M2 m1 M1  , where the non-mentioned entries are to be read as zero. Setting N =  n1 n2 · · · nh−1 nh n1 n2 · · · nh−1 ... ... ... n1 n2 n1  , M =  m1 m2 · · · mh−1 mh m1 m2 · · · mh−1 ... ... ... m1 m2 m1  , D1 = diag(Nh, . . . , N1), and D2 = diag(Mh, . . . ,M1), AD can be rewritten in the com- pact block form AD = ( D1 N M D2 ) . In order to obtain more bounds we set P = ( 0 xI I 0 ) , for some x 6= 0. Since the matrices PADP −1 = ( D2 xM x−1N D1 ) and AD are similar, they have the same index. We choose x such that the sum in the first row and the (h + 1)-th row are equal. It leads to Mhx2 − (Nh −Mh)x − Nh = 0, i.e., x = NhMh . By Frobenius Theorem [11, Theorem 3.1.1], we have min 16i6n Ri 6 κ 6 max 16i6n Ri, 182 Ars Math. Contemp. 6 (2013) 171–185 where Ri stands for the sum of elements in the i-th row of PADP−1. Using this, we get the following result. Proposition 4.7. If G is a connected DNG, then min { n1 ( Nh Mh + 1 ) , n1 ( Mh Nh + 1 )} 6 κ 6 Nh +Mh = ν. (4.4) Clearly, the upper bound does not provide a decisive progress in our quest (recall (4.1)). We will establish some more interesting improvements next. Let Ri be the sum of the entries in row i of the matrix AD. It is easy to confirm that Ri = 2Nh−i+1, for i ∈ {1, . . . , h} = 2Mh−i+1, for i ∈ {h+ 1, . . . , 2h}, and, therefore, maxRi = max{2Nh, 2Mh}. By Frobenius Theorem min{2n1, 2m1} 6 κ 6 max{2Nh, 2Mh}. (4.5) Here the upper bound does not make any (general) improvement since max{2Nh, 2Mh} > ν (compare (4.1)), so next we use the result of Minc (see [10]), which for the matrix AD reads: min i ∑h j=1(AD)ijRj Ri 6 κ 6 max i ∑h j=1(AD)ijRj Ri . Proposition 4.8. If G is a connected DNG, then min{n1 +Mh,m1 +Nh} 6 κ 6 max {  Nh +Nh,  Mh +Mh } . (4.6) The bounds (4.6) obviously improve both (4.4) and (4.5), but the lower bound is still rough comparing with Corollary 4.2. 5 Computational results In this final section, we provide several examples which can help to gain a better insight into the quality of the bounds obtained in the previous section. We compute the lower bounds of Propositions 4.1 and 4.3, and the upper bounds from Propositions 4.4, 4.5, 4.6, and 4.8. One observes that the bound from Proposition 4.1 is always integral. The number of vertices in the corresponding DNG is also given in every example since it makes another upper bound (cf. (4.1)). It can be easy checked that the lower bounds from Corollary 4.2 and Propositions 4.7 and 4.8, all of them having simple expressions, are rough in some cases, and therefore they are not considered in our examples. We also compute the relative errors in each case. Example 5.1. First we consider a randomly chosen DNG with small number of vertices and some larger DNGs derived from the previous one: G1 = DNG(2, 2, 5, 3; 2, 3, 1, 1); G2 = DNG(10, 10, 25, 15; 10, 15, 5, 5); G3 = DNG(200, 200, 500, 300; 200, 300, 100, 100); M. And̄elić et al.: Sharp spectral inequalities for connected bipartite graphs. . . 183 Prop. 4.3 Prop. 4.1 κ Prop. 4.8 Prop 4.6 Prop 4.5 Prop 4.4 ν G1 13.6785 14 15.6451 16.7500 17.0210 17.0550 17.4530 19 -12.57% -10.52% 7.06% 8.79% 9.01% 11.56% 21.44% G2 68.3923 70 78.2257 83.7500 85.1052 85.2749 87.2649 95 -12.57% -10.52% 7.06% 8.79% 9.01% 11.56% 21.44% G3 1367.8452 1400 1564.5133 1675.0000 1702.1030 1705.4985 1745.2987 1900 -12.57% -10.52% 7.06% 8.79% 9.01% 11.56% 21.44% Notice that κ(G2) (resp. κ(G3)) is very close to 5κ(G1) (resp. 100κ(G3)); we get 5κ(G1) − κ(G2) ≈ 10−7. Since the similar fact holds for all bounds obtained (compare the corresponding propositions), we get the same results for the relative errors. Example 5.2. Here we consider the DNGs obtained from G1 by multiplying some of its parameters: H1 = DNG(2000, 2, 5, 3; 2, 3, 1, 1000); H2 = DNG(2000, 2, 5, 3; 2, 3, 1000, 1); H3 = DNG(2000, 2, 5, 3; 2, 3000, 1, 1); H4 = DNG(2000, 2, 5, 3; 2000, 3, 1, 1); Prop. 4.3 Prop. 4.1 κ Prop. 4.8 Prop 4.6 Prop 4.5 Prop 4.4 ν H1 3006.0284 3006 3006.0287 3011.0164 3008.2682 3008.2960 3012.6750 3016 -8 · 10−6% -1 · 10−3% 0.17% 0.07% 0.08% 0.22% 0.33% H2 3008.0175 3007 3008.0177 3012.0104 3009.8145 3009.8323 3013.3388 3016 -5 · 10−6% -0.03% 0.13% 0.06% 0.06% 0.18% 0.27% H3 5010.9908 5009 5010.9909 5010.9980 5011.7199 5011.7199 5012.2008 5014 -5 · 10−7% -0.04% 1 · 10−4% 0.15% 0.15% 0.02% 0.06% H4 4014.9731 4010 4014.9732 4014.9866 4014.9800 4014.9800 4014.9933 4015 -3 · 10−6% -0.12% 3 · 10−4% 2 · 10−4% 2 · 10−4% 5 · 10−4% 7 · 10−4% In this example all bounds are (more or less) close to the exact value of Q-index. We already pointed that the bounds obtained in Propositions 4.5 and 4.6 are the improvements of the one obtained in Proposition 4.4. This example shows that the bound from Proposition 4.8 is incomparable to them. In opposition to the previous example, here Proposition 4.1 gives a better estimation than Proposition 4.3. Example 5.3. The parameters of the following DNGs are obtained by multiplying the parameters of G1 by 1, 10, 100 or 1000 ad hoc. I1 = DNG(2, 2, 5, 3; 2000, 300, 10, 1); I2 = DNG(2, 2, 5, 3; 2, 30, 100, 1000); I3 = DNG(2000, 200, 50, 30; 2000, 300, 10, 1); Prop. 4.3 Prop. 4.1 κ Prop. 4.8 Prop 4.6 Prop 4.5 Prop 4.4 ν I1 2255.0867 2314 2316.3632 2322.5716 2322.5716 2322.5733 2322.5737 2323 -2.65% -0.10% 0.26% 0.27% 0.27% 0.27% 0.29% I2 1118.5026 1134 1134.0007 1134.3799 1134.4002 1134.4000 1134.4002 1144 -1.37% -6 · 10−5% 0.03% 0.04% 0.04% 0.04% 0.88% I3 4562.6064 4550 4562.6584 4563.2717 4563.1367 4563.1369 4563.6312 4591 -0.28% -1 · 10−3% 0.01% 0.01% 0.01% 0.02% 0.62% 184 Ars Math. Contemp. 6 (2013) 171–185 Taking into account the lower bound of Proposition 4.3, one can conclude that its devia- tion from the exact value is expected for I1 (and other similar graphs). Note that Proposition 4.6 will often give better bound than Proposition 4.5, but not always – see graphs I2 or J2 in the next example. Example 5.4. Finally, we consider the extensions of the original graphs: J1 = DNG(2, 2, 5, 3, 2, 3, 1, 1; 2, 3, 1, 1, 2, 2, 5, 3); J2 = DNG(20000, 2, 5, 3, 10, 10, 10, 10; 2, 3, 1, 10000, 10, 10, 10, 10); J3 = DNG(2, 2, 5, 3, 1, 1, 1, 1; 2000, 300, 10, 1, 1, 1, 1, 1); Prop. 4.3 Prop. 4.1 κ Prop. 4.8 Prop 4.6 Prop 4.5 Prop 4.4 ν J1 23.1888 23 27.4601 29.0526 32.3032 32.3022 32.8203 38 -15.56% -16.24% 5.80% 17.64% 17.64% 19.52% 38.38% J2 30065.9176 30046 30065.9178 30080.9446 30072.6747 30072.7003 30085.9668 30096 -6 · 10−7% -0.07% 0.05% 0.02% 0.02% 0.07% 0.10% J3 2313.0140 2324 2327.5409 2330.8445 2330.8452 2330.8452 2330.8456 2331 -0.62% -0.15% 0.14% 0.14% 0.14% 0.14% 0.15% The bounds obtained will be used in a forthcoming research regarding the graphs with maximal Q-index and fixed (but high) orders and also a fixed particular size. We also remark that the results could be also compared to the corresponding bounds obtained for the adjacency spectra. Acknowledgment The authors thank two anonymous referees for providing constructive comments in im- proving the contents of this paper. References [1] M. And̄elić, C. M. da Fonseca, S. K. Simić and D. V. Tošić, Connected graphs of fixed order and size with maximal Q-index: Some spectral bounds, Discrete Appl. Math. 160 (2012), 448–459. [2] M. And̄elić, T. Koledin and Z. Stanić, Nested graphs with bounded second largest (signless Laplacian) eigenvalue, Electron. J. Linear Algebra 24 (2012), to appear. [3] F. K. Bell, D. Cvetković, P. Rowlinson and S. K. Simić, Graphs for which the least eigenvalue is minimal, I, Linear Algebra Appl. 429 (2008), 234–241. [4] F. K. Bell, D. Cvetković, P. Rowlinson and S. K. Simić, Graphs for which the least eigenvalue is minimal, II, Linear Algebra Appl. 429 (2008), 2168–2179. [5] A. Bhattacharya, S. Friedland and U. N. Peled, On the first eigenvalue of bipartite graphs, Electron. J. Combin. 15 (2008), R#144. [6] D. Cvetković, P. Rowlinson and S. K. Simić, Eigenvalue bounds for the signless Laplacian, Publ. Inst. Math. (Beograd), 81 (95) (2007), 11–27. [7] D. Cvetković, P. Rowlinson and S. K. Simić, Signless Laplacians of finite graphs, Linear Alge- bra Appl. 423 (2007), 155–171. [8] D. Cvetković, P. Rowlinson and S. K. Simić, An Introduction to the Theory of Graph Spectra, Cambridge University Press, 2009. M. And̄elić et al.: Sharp spectral inequalities for connected bipartite graphs. . . 185 [9] L.-H. Feng, G.-H. Yu, On three conjectures involving the signless Laplacian spectral radius of graphs, Publ. Inst. Math. (Beograd) 85 (99) (2009), 35–38. [10] S.-L. Liu, Bounds for the greatest characteristic root of a nonnegative matrix, Linear Algebra Appl. 239 (1996), 151–160. [11] M. Marcus and H. Minc, A Survey of Matrix Theory and Matrix Inequalities, Dover Publica- tions Inc., New York, 1992; unabridged reprint of the corrected 1969 printing, Prindle, Weber, & Schmidt, Boston.