Graphs with five eigenvalues

15
This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and education use, including for instruction at the authors institution and sharing with colleagues. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elsevier’s archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright

description

Two spectral characterizations of regular, bipartite graphs with five eigenvalues

Transcript of Graphs with five eigenvalues

Page 1: Graphs with five eigenvalues

This article appeared in a journal published by Elsevier. The attachedcopy is furnished to the author for internal non-commercial researchand education use, including for instruction at the authors institution

and sharing with colleagues.

Other uses, including reproduction and distribution, or selling orlicensing copies, or posting to personal, institutional or third party

websites are prohibited.

In most cases authors are permitted to post their version of thearticle (e.g. in Word or Tex form) to their personal website orinstitutional repository. Authors requiring further information

regarding Elsevier’s archiving and manuscript policies areencouraged to visit:

http://www.elsevier.com/copyright

Page 2: Graphs with five eigenvalues

Author's personal copy

Linear Algebra and its Applications 435 (2011) 2612–2625

Contents lists available at ScienceDirect

Linear Algebra and its Applications

journal homepage: www.elsevier .com/locate/ laa

Two spectral characterizations of regular, bipartite graphs

with five eigenvalues

Dragan Stevanovic∗,1FAMNIT, University of Primorska, Glagoljaška 8, 6000 Koper, Slovenia

Faculty of Sciences and Mathematics, University of Niš, Višegradska 33, 18000 Niš, Serbia

A R T I C L E I N F O A B S T R A C T

Article history:

Available online 14 May 2011

Submitted by E. van Dam

Dedicated to Dragoš Cvetkovic on the

occasion of his 70th birthday

AMS classification:

05C50

05C25

05C75

05D15

Keywords:

Regular graphs

Integral graphs

Mutually orthogonal Latin squares

Graphs with a few distinct eigenvalues usually possess an interest-

ing combinatorial structure. We show that regular, bipartite graphs

with atmost six distinct eigenvalues have theproperty that eachver-

tex belongs to the constant number of quadrangles. This enables to

determine, from the spectrum alone, the feasible families of num-

bers of common neighbors for each vertex with other vertices in

its part. For particular spectra, such as [6, 29, 06,−29,−6] (where

exponents denote eigenvalue multiplicities), there is a unique such

family, which makes it possible to characterize all graphs with this

spectrum.

Using this lemma we also to show that, for r � 2, a graph has

spectrum [r,√rr(r−1)

, 02(r−1),−√rr(r−1)

,−r] if and only if it is a

graph of a 1-resolvable transversal design TD(r, r), i.e., if it corre-sponds to the complete set of mutually orthogonal Latin squares of

size r in a well-defined manner.

© 2011 Elsevier Inc. All rights reserved.

1. Introduction

One of the basic and hardest problems of spectral graph theory is to reconstruct a graph from its

spectrum alone. This problem has been solved for a few well defined families of spectra (for example,

for the spectra of paths, cycles, complete and complete bipartite graphs [1], complements of paths [2]),

as well as for some additional spectra (for detailed surveys, see [3,4]). The hardness of the problem is

easily illustrated by the observation that the construction of strongly regular graphswith a given set of

∗ Address: Faculty of Sciences and Mathematics, University of Niš, Višegradska 33, 18000 Niš, Serbia.

E-mail address: [email protected] The author was supported by Grant 174033 of Serbian Ministry of Education and Science, Research Programme P1-0285 of

Slovenian Research Agency and Serbian-Slovak Bilateral Research Project SK-SRB-0005-09.

0024-3795/$ - see front matter © 2011 Elsevier Inc. All rights reserved.

doi:10.1016/j.laa.2011.04.032

Page 3: Graphs with five eigenvalues

Author's personal copy

D. Stevanovic / Linear Algebra and its Applications 435 (2011) 2612–2625 2613

parameters is equivalent to their reconstruction from the spectrum alone. Our goal is to add another

spectrum and another family of spectra to the set of spectra for which all corresponding graphs are

known.

Graphs with a few distinct eigenvalues usually possess an interesting combinatorial structure. The

strongly regular graphs are prime example of such graphs [5, Chapter VII.11], however, other graphs

with a few eigenvalues have been considered in the literature as well. van Dam [6] and van Dam

and Spence [7] studied properties and feasibility conditions for regular graphs with four eigenvalues,

giving several characterizations and constructing a number of examples, using both theoretical results

and computer search. They have also noted the correspondence of regular bipartite graphs with five

eigenvalues to the so-called partial geometric designs [8].

Another example of graphs with a few eigenvalues are integral graphs of small degree. Integral

graphs have been one of recurring research topics of Professor Cvetkovic, ever since Bussemaker and

Cvetkovic [9] determined all cubic integral graphs. Constructions of integral graphs have recently

attracted attention of several researchers, with over 50 papers having been published in last 10 years

[10–60].

We are interested here in regular, bipartite graphs with a few eigenvalues. Since a bipartite graph

with four eigenvalues must be the incidence graph of a symmetric design [1, p. 166], we will restrict

ourselves to regular, bipartite graphs with five eigenvalues. Our starting point is a simple lemma,

given in Section 2, which says that each vertex belongs to the constant number of quadrangles in a

regular, bipartite graph with at most six distinct eigenvalues. This makes it possible to determine,

from the spectrum alone, the feasible families of numbers of common neighbors for each vertex with

other vertices in its part. For particular spectra, such as [6, 29, 06, −29, −6] (where exponents denote

eigenvalue multiplicities), there is a unique such family, which further enables to characterize all

graphs with this spectrum in Section 3.

Wealsouse this lemmatoshowinSection4 that, for r � 2, agraphhasspectrum [r,√rr(r−1)

, 02(r−1),

−√rr(r−1)

, −r] if and only if it is a graph of a 1-resolvable transversal design TD(r, r), i.e., if it corre-sponds to the complete set of mutually orthogonal Latin squares of size r in a well-defined manner.

Some concluding remarks on constructing new examples of regular, bipartite graphs with five

distinct eigenvalues are given in Section 5.

2. Families of numbers of common neighbors

As usual, let J be the all-one matrix and I the unit matrix. We rely upon the following

Lemma 1 (Hoffman [61]). For a graph G with adjacency matrix A there exists a polynomial P(x), suchthat P(A) = J, if and only if G is regular and connected. In such case, we have

P(A) = n∏k

i=2(A − λiI)∏ki=2(r − λi)

, (1)

wheren is thenumberof vertices, r = λ1 is the largest eigenvalueandλ2, . . . , λk areall distinct eigenvalues

of G.

Now, let G be an r-regular bipartite graph (U, V, E) and let nu,v denote the number of common

neighbors of vertices u, v ∈ U. The number of walks of length two starting at u is equal to r2. On the

other hand, it is also equal to∑

v∈U nu,v. Since nu,u = r, we get∑v∈U\{u}

nu,v = r(r − 1). (2)

If G has at most six distinct eigenvalues (k � 6), Lemma 1 yields that the polynomial P(A) is of degreeat most five. Considering (u, u)-entries on both sides of the identity P(A) = J, and taking into account

that for bipartite graphs holds Au,u = (A3)u,u = (A5)u,u = 0, we see that for some a0, a1, a2, which

Page 4: Graphs with five eigenvalues

Author's personal copy

2614 D. Stevanovic / Linear Algebra and its Applications 435 (2011) 2612–2625

depend only on r = λ1, λ2, …, λk , it holds

a0(A4)u,u + a1(A

2)u,u + a2 =k∏

i=2

(r − λi).

Since (A2)u,u = r, we get, in particular, that (A4)u,u does not depend on u. Now, (A4)u,u denotes the

number of closed walks of length four starting at u and it is straightforward to see that

(A4)u,u = r2 + r(r − 1) + 2qu,

where qu denotes the number of quadrangles to which u belongs. An important consequence is that qudoes not depend on u, and therefore qu = 4q/n, where q is the total number of quadrangles in G. On

the other hand, if we partition the closed walks of length four starting at u according to their middle

vertex, we get

(A4)u,u = ∑v∈U

n2u,v.

From nu,u = r it follows that∑v∈U\{u}

n2u,v = r(r − 1) + 2qu. (3)

Thus, we have proved.

Lemma 2. Let G = (U, V, E) be an r-regular, bipartite graph having at most six distinct eigenvalues. If

n and q are the numbers of vertices and quadrangles of G, respectively, then each vertex u ∈ U belongs to

qu = 4q/n quadrangles. Moreover, the family of numbers of common neighbors {nuv : v ∈ U\{u}} (the

nuv family) satisfies Eqs. (2) and (3).

Suppose now that G is a connected, r-regular, bipartite integral graph with adjacency matrix A

and 2p vertices (thus, p vertices in each part). Suppose that A has five distinct eigenvalues, and let its

spectrum be [r, xy, 02z, −xy, −r] (where exponents denote the eigenvalue multiplicities). Further, let

q and h denote the number of quadrangles and hexagons in G, respectively. Recall that the number of

closed walks of length k in G is equal to the trace of Ak , i.e., to the sum of kth powers of eigenvalues

of A. By counting the numbers of closed walks of lengths 0, 2, 4 and 6 in G according to the starting

vertex and the type of the closed walk, we can get the following Diophantine equations (where each

side represents one half of the number of closed walks of the corresponding length):

1 + y + z = p, (4)

r2 + yx2 = pr, (5)

r4 + yx4 = pr(2r − 1) + 4q, (6)

r6 + yx6 = pr(5r2 − 6r + 2) + 24(r − 1)q + 6h. (7)

Eq. (7) follows from the enumeration of closed walks of length six (note that in (7) we take only a half

of the total number):

Lemma 3. The number of closed walks of length six in a regular, bipartite graph G with n vertices, q

quadrangles and h hexagons is equal to nr(5r2 − 6r + 2) + 48(r − 1)q + 12h.

Proof. Each closed walkW : v0v1 . . . v5v0 in G has one of the types illustrated in Fig. 1. The numbers

of closed walks of each of these types are as follows:

Page 5: Graphs with five eigenvalues

Author's personal copy

D. Stevanovic / Linear Algebra and its Applications 435 (2011) 2612–2625 2615

Fig. 1. Types of closed walks of length 6 in a bipartite graph.

(i) There are n possible starting vertices, and r possible neighbors for A, B, C, so there are nr3 walks

of this type.

(ii) There are n possible starting vertices, r possible neighbors for A, r − 1 possible neighbors for B

(as B has to be different from the starting vertex), and r possible neighbors for C, so there are

nr2(r − 1) walks of this type.

(iii) There are n possible starting vertices, r possible neighbors for A and r − 1 possible neighbors

for each of B and C, so there are nr(r − 1)2 walks of this type.

(iv) There are n possible starting vertices, r possible neighbors for each of A and B and r − 1 possible

neighbors for C, so there are nr2(r − 1) walks of this type.

(v) There are n possible starting vertices, r possible neighbors for A and r − 1 possible neighbors

for each of B and C, so there are nr(r − 1)2 walks of this type.

(vi) For each of q quadrangles, there are four possible starting vertices, two possible directions and

six choices for quadrangle edge to be repeated (if quadrangle vertices are visited as ABCD, then

these six choices are ADA at the beginning of the walk, then ABA, BCB, CDC, DAD and ABA at the

end of the walk), so there are 48q walks of this type.

(vii) For eachofqquadrangles, there are fourpossible startingvertices at thequadrangle, twopossible

directions and 5(r − 2) choices for an extra edge (r − 2 edges incident with A at the beginning

of the walk, r − 2 edges incident with B, C and D, and r − 2 edges incident with A at the end of

the walk), so there are 40q(r − 2) walks of this type.

(viii) For each of q quadrangles, there are two possible directions and r − 2 neighbors of each of four

quadrangle vertices as possible starting vertices, so there are 8q(r − 2) walks of this type.

(ix) For each of h hexagons there are six possible starting vertices and two possible directions, so

there are 12h walks of this type.

In total, there are nr(5r2 − 6r + 2) + 48(r − 1)q + 12h closed walks of length six, as claimed.

Eqs. (4)–(7) may be used to determine the feasible spectra of regular, bipartite, integral graphs

with five distinct eigenvalues. The same approach has been applied in [9,44] to determine the feasible

spectra of the cubic and 4-regular integral graphs, respectively.

Page 6: Graphs with five eigenvalues

Author's personal copy

2616 D. Stevanovic / Linear Algebra and its Applications 435 (2011) 2612–2625

Table 1

Examples of feasible spectra determining unique nuv families.

Feasible spectrum Unique nuv family Comment

[3, 29, 010, −29, −3] {02, 16} Constructed in [9]

[4, 28, 06, −28, −4] {01, 18, 22} Constructed in [62]

[4, 212, 06, −212, −4] {03, 112} Constructed in [62]

[6, 29, 04, −29, −6] {23, 38} Constructed in Section 3

[12, 320, 012, −320, −12] {2, 520, 65} Does not exist (see text)

Table 2

Examples of feasible spectra with q = 0.

Feasible spectrum Feasible spectrum Feasible spectrum

[3, 29, 010, −29, −3] [15, 5441, 0616, −5441, −15] [25, 5600, 048, −5600, −25][4, 212, 06, −212, −4] [16, 4240, 030, −4240, −16] [28, 71944, 02970, −71944, −28][6, 350, 060, −350, −6] [16, 5384, 0462, −5384, −16] [29, 71682, 02376, −71682, −29][7, 349, 040, −349, −7] [19, 63249, 05850, −63249, −19] [31, 6961, 0370, −6961, −31][9, 372, 016, −372, −9] [21, 5441, 0208, −5441, −21] [33, 817424, 032800, −817424, −33][9, 4324, 0520, −4324, −9] [21, 6980, 01440, −6980, −21] [35, 71445, 01224, −71445, −35][13, 4169, 0102, −4169, −13] [22, 6847, 01120, −6847, −22] [36, 61260, 070, −61260, −36]

There are quite a few feasible spectra satisfying the aboveDiophantine equations. Themost promis-

ing among them are those that determine a small number, preferably only one, feasible nuv family for

each vertex. Small such spectra and corresponding nuv families (with exponents denoting multiplici-

ties) are given in Table 1.

The nonexistence of the graphwith spectrum [12, 320, 012, −320, −12] is easy to show: its unique

nuv-family {2, 520, 65} clearly shows that each vertex u in such hypothetical graph has a unique

counterpart vertex v in its part with nuv = 2. Thus, each part has to be divisible in pairs (as the

counterpart of v is u again), but this is impossible, as each part has 27 vertices.

The construction of all graphs with spectrum [6, 29, 04, −29, −6] is given in Section 3.

A separate set of feasible spectra are those with q = 0, which automatically determine unique nuv

family. Namely, for such spectra

∑v∈U\{u}

nuv = ∑v∈U\{u}

n2uv = r(r − 1),

and so each vertex has the same nuv family {1r(r−1), 0p−1−r(r−1)}. Such feasible spectra with r � 36

are shown in Table 2.

Among these spectra, particularly interesting are those of form

[s2, ss2(s2−1), 02(s2−1), −ss

2(s2−1), −s2], for some natural s � 2.

They are the subject of Section 4.

3. Graph with spectrum [6, 29, 04,−29,−6]Theorem 4. There exists exactly three nonisomorphic graphs with spectrum [6, 29, 04, −29, −6].Proof. Let G = (U, V, E) be a regular, bipartite graph with the above spectrum, and let A be its

adjacency matrix. For this particular spectrum, we have p = 12 and q = 162, and consequently

Page 7: Graphs with five eigenvalues

Author's personal copy

D. Stevanovic / Linear Algebra and its Applications 435 (2011) 2612–2625 2617

qu = 2q/p = 27, so that for each u ∈ U its 11-element nuv family satisfies

∑v∈U\{u}

nuv = 30,

∑v∈U\{u}

n2uv = 84.

This system of Diophantine equations has a unique solution in nonnegative integers:

{nuv|v ∈ U\{u}} = {38, 23}.Let a ∈ U be an arbitrary vertex ofG. Let c1, . . . , c8 ∈ U be vertices having three commonneighbors

with a each, and let b1, b2, b3 ∈ U be vertices having two common neighbors with a each. Hoffman’s

equation for G reads

A4 + 6A3 − 4A2 − 24A = 96J.

Select b1 arbitrarily from {b1, b2, b3}. Then(A4)a,b1 − 4(A2)a,b1 = 96.

From na,b1 = (A2)a,b1 = 2, we have (A4)a,b1 = 104. Among these 104 walks of length four starting

in a and ending in b1, there are 12na,b1 = 24 walks which have either a or b1 as their middle vertex.

Remaining 80 walks have as their middle vertex one of the vertices c1, . . . , c8 or b2, b3. If d is the

middle vertex of a walk of length four between a and b1, then the number of such walks is na,dnd,b1 .

Therefore,

80 =8∑

i=1

na,ci nci,b1 + na,b2nb2,b1 + na,b3nb3,b1

= 3

8∑i=1

nci,b1 + 2nb2,b1 + 2nb3,b1

� 3 · 8 · 3 + 2 · 2 + 2 · 2 = 80.

The equality has to hold in the above inequality, which means that

nb1,ci = 3 for each 1 � i � 8 and nb1,b2 = nb1,b3 = 2.

Since b1 was chosen arbitrarily from {b1, b2, b3}, the analogous conclusion follows for b2 and b3:

nb2,ci = nb3,ci = 3 for each 1 � i � 8 and nb2,b3 = 2.

Thus, 12 vertices of U may be divided in three fourtuples, such that the number of common neigh-

bors is equal to two for vertices fromthe same fourtuple and three for vertices fromdifferent fourtuples.

One of these fourtuples is (a, b1, b2, b3), andwithout loss of generality, wemay suppose that the other

two fourtuples are (c1, c2, c3, c4) and (c5, c6, c7, c8).Next, let V = {d1, . . . , d6, e1, . . . , e6}, where d1, . . . , d6 are adjacent to a, and e1, . . . , e6 are

not. Since na,bi = 2 for i = 1, 2, 3, it follows that the bipartite subgraph induced by b1, b2, b3 and

e1, . . . , e6 is (4,2)-semiregular. As only one (4,2)-semiregular graph on 3 + 6 vertices exists, we may,

without loss of generality, suppose that:

• b1 is adjacent to e1, e2, e3, e4,

Page 8: Graphs with five eigenvalues

Author's personal copy

2618 D. Stevanovic / Linear Algebra and its Applications 435 (2011) 2612–2625

Fig. 2. The subgraph induced by the vertices in V and a fourtuple in U, each pair of whose vertices has two common neighbors in V .

• b2 is adjacent to e1, e2, e5, e6, and• b3 is adjacent to e3, e4, e5, e6.

Since each pair of vertices from {b1, b2, b3} already has two common neighbors among {e1, . . . , e6},then no such pair has common neighbors among {d1, . . . , d6}. In other words, the bipartite subgraph

induced by b1, b2, b3 and d1, . . . , d6 is (2,1)-semiregular, and we may further suppose that:

• b1 is adjacent to d1, d2,• b2 is adjacent to d3, d4, and• b3 is adjacent to d5, d6.

The subgraph of G induced by the vertices {a, b1, b2, b3, d1, . . . , d6, e1, . . . , e6} is shown in Fig. 2.

To construct the rest of this graph,we have to determine the neighbors of vertices c1, . . . , c8 among

{d1, . . . , d6, e1, . . . , e6}. For this task, we use the fact that for each i = 1, . . . , 8

nci,a = nci,b1 = nci,b2 = nci,b3 = 3. (8)

Let xj , j = 1, . . . , 6 and yj , j = 1, . . . , 6 be the variables indicating whether the vertex ci is adjacent

to dj (in which case xj = 1) or not (xj = 0), respectively, whether it is adjacent to ej (yj = 1) or not

(yj = 0). Thus, since ci has degree six, we have

(x1 + x2) + (x3 + x4) + (x5 + x6) + (y1 + y2) + (y3 + y4) + (y5 + y6) = 6. (9)

The following table shows which of these indicator variables have to be added together to determine

the number of common neighbors of ci and each of a, b1, b2, b3:

d1 d2 d3 d4 d5 d6 e1 e2 e3 e4 e5 e6

a x1 x2 x3 x4 x5 x6

b1 x1 x2 y1 y2 y3 y4

b2 x3 x4 y1 y2 y5 y6

b3 x5 x6 y3 y4 y5 y6

The conditions (8) are translated into the following system of equations:

(x1 + x2) + (x3 + x4) + (x5 + x6) = 3,

(x1 + x2) + (y1 + y2) + (y3 + y4) = 3,

(x3 + x4) + (y1 + y2) + (y5 + y6) = 3,

(x5 + x6) + (y3 + y4) + (y5 + y6) = 3,

Page 9: Graphs with five eigenvalues

Author's personal copy

D. Stevanovic / Linear Algebra and its Applications 435 (2011) 2612–2625 2619

which is equivalent to

(x1 + x2) + (x3 + x4) + (x5 + x6) = 3, (10)

x1 + x2 = y5 + y6, (11)

x3 + x4 = y3 + y4, (12)

x5 + x6 = y1 + y2. (13)

There are twopossible solution forms for theabove system:either the sums in (11)–(13) all havevalue1,

or they have values 2, 1 and 0 in some order. Since all variables x1, . . . , x6, y1, . . . , y6 are either 0 or 1,

there are 26 different solutions in the former case, and 3! · 22 different solutions in the latter case. In

total, there are 88 solutions which describe potential neighborhoods of vertices c1, . . . , c8.Beforewe proceed further, let us to introducemore compact notation for the solutions of (10)–(13):

first, we put the sums in the following order

x1 + x2 x3 + x4 x5 + x6 y1 + y2 y3 + y4 y5 + y6,

and then use the symbols 0, l, r, 2 to denote feasible options for each sum:

0 : both summands are equal 0,

l : left summand is 1, right summand is 0,

r : left summand is 0, right summand is 1,

2 : both summands are equal 1.

For example, one of the solutions

x1 = x2 = x3 = 1, x4 = x5 = x6 = 0, y1 = y2 = y3 = 0, y4 = y5 = y6 = 1,

may be shortly denoted as 2l00r2.

We can now create an auxiliary, bicolored compatibility graph with 88 solutions of (10)–(13) as its

vertices: two solutions are joined by a red edge in the compatibility graph if the corresponding neigh-

borhoods have twoneighbors in common, and by a blue edge if the corresponding neighborhoods have

three neighbors in common. The compatibility graph contains no loops, as each solution represents a

neighborhood with six elements. The neighborhoods of vertices {c1, . . . , c8} thus form an 8-clique in

the compatibility graph, consisting of two disjoint red 4-cliques {c1, . . . , c4} and {c5, . . . , c8}, mutu-

ally connected by blue edges. Our next task is to describe such cliques.

First, let us consider cliques in the part of the compatibility graph induced by solutions with the

sums (11)–(13) equal to 2, 1 and 0 in some order. If two distinct solutions have the same sum among

(11)–(13) equal to two, then they have at least four common neighbors, meaning that they are not

adjacent in the compatibility graph. Thus, this part of the compatibility graph has cliques of size at

most three, with each vertex in such clique having different sum among (11)–(13) equal to 2.

Therefore, an 8-clique in the compatibility graph has to contain at least five solutions inwhich each

of the sums (11)–(13) is equal to 1. Moreover, each of the red 4-cliques has to contain at least one such

solution. Suppose that the neighborhoods of c1 and c5 are described by such solutions. We can choose

two solutions to describe the neighborhoods of c1 and c5, respectively, without loss of generality:

c1 : llllll, c5 : lllrrr.Possible neighborhoods for c2, c3, c4 are now found among those solutions to (10)–(13)which have

two common neighbors with c1 and three common neighbors with c5. It is straightforward to see that

there are 15 such solutions, and the red colored subgraph of the compatibility graph induced by these

solutions is shown in Fig. 3. Vertices c2, c3 and c4 form one of 11 triangles in this subgraph.

Page 10: Graphs with five eigenvalues

Author's personal copy

2620 D. Stevanovic / Linear Algebra and its Applications 435 (2011) 2612–2625

Fig. 3. Vertices from the compatibility graph having two common neighbors with c1 and three common neighbors with c5.

Fig. 4. Vertices from the compatibility graph having three common neighbors with c1 and two common neighbors with c5.

Similarly, possible neighborhoods for c6, c7, c8 are found among those solutions to (10)–(13) which

have three common neighbors with c1 and two common neighbors with c5. Again, there are 15 such

solutions, and the red colored subgraph of the compatibility graph induced by these solutions is shown

in Fig. 4. Vertices c6, c7 and c8 form one of 11 triangles in this subgraph.

To fully reconstruct graph G, it remains to find pairs of the triangles from these two induced sub-

graphs which induce 6-clique in the compatibility graph. There are 22 such pairs, indicated in Table 3.

These pairs yield exactly three nonisomorphic graphs with spectrum [6, 29, 04, −29, −6], whose

codes in Graph6 format [63] are given by:

W???????????{orKF‘_Fwii@UgBJGCio@YS?eX?BQo?Sq_?W???????????{orKF‘_Fwii@Ug?}_E_w@dc?eX?DIo?LQ_?W???????????{orKF‘_Fwii@Ug?}_E_w@dc?eX?BQo?Sq_?

4. Graphs with spectrum [r,√rr(r−1)

, 02(r−1),−√rr(r−1)

,−r]In this section, we characterize graphs with the above spectrum and show their relationship with

the complete sets of mutually orthogonal Latin squares of size r. Let us first recall the definition of a

transversal design from [5, Chapter III.3.2].

Definition 1. A transversal design of groupsize n, blocksize k, and index λ, denoted TDλ(k, n), is a

triple (U, G, B), where

(1) U is a set of kn elements.

(2) G is a partition of U into k classes (the groups), each of size n.

(3) B is a collection of k-subsets of U (the blocks).

Page 11: Graphs with five eigenvalues

Author's personal copy

D. Stevanovic / Linear Algebra and its Applications 435 (2011) 2612–2625 2621

Table 3

Encodings of theneighborhoods of vertices c2, c3, c4 and c6, c7, c8. Codes in the same cell produce isomorphic graphs.

c2 c3 c4 c6 c7 c8

rlrrlr rrllrr lrrrrl rlrlrl rrlrll lrrllr

rrlrrl lrrrlr rlrlrr rrlllr lrrlrl rlrrll

lrrlrr rlrrrl rrlrlr lrrrll rlrllr rrllrl

2r00r2 0r22r0 rlrrlr rlrlrl rrlllr lrrrll

02rr20 20rr02 rrllrr rrlrll lrrlrl rlrllr

r0220r r2002r lrrrrl lrrllr rlrrll rrllrl

2r00l2 0r22l0 rlrlrl rlrrlr rrlrrl lrrlrr

02rl20 20rl02 rrlrll rrllrr lrrrlr rlrrrl

r0220l r2002l lrrllr lrrrrl rlrlrr rrlrlr

2r00r2 02rr20 r0220r rlrlrl rrlrll lrrllr

0r22r0 20rr02 r2002r rlrlrl rrlrll lrrllr

2r00l2 02rl20 r0220l rlrrlr rrllrr lrrrrl

0r22l0 20rl02 r2002l rlrrlr rrllrr lrrrrl

rlrrlr rrlrrl lrrlrr rlrlrl rrlllr lrrrll

rrllrr lrrrlr rlrrrl rrlrll lrrlrl rlrllr

lrrrrl rlrlrr rrlrlr lrrllr rlrrll rrllrl

2r00r2 0r22r0 rlrrlr rlrlrl rrlrll lrrllr

02rr20 20rr02 rrllrr rlrlrl rrlrll lrrllr

r0220r r2002r lrrrrl rlrlrl rrlrll lrrllr

2r00l2 0r22l0 rlrlrl rlrrlr rrllrr lrrrrl

02rl20 20rl02 rrlrll rlrrlr rrllrr lrrrrl

r0220l r2002l lrrllr rlrrlr rrllrr lrrrrl

(4) every unordered pair of elements from U is contained either in exactly one group or in exactly

λ blocks, but not both.

A TDλ(k, n) (V, G, B) is α-resolvable if its blocks can be partitioned into sets B1, . . . , Bs, where every

element of V occurs exactly α times in each Bi.

When λ = 1, one writes simply TD(k, n).

Further, recall that the graph of the transversal design (U, G, B) is the bipartite graph on |U| + |B|vertices corresponding to the elements and the blocks of the design, with two vertices adjacent if and

only if one corresponds to a block and the other corresponds to an element contained in that block.

Thus, the edges of this graph represent incidences between the elements and the blocks of the design.

Theorem 5. A graph has spectrum [r,√rr(r−1)

, 02(r−1), −√rr(r−1)

, −r] if and only if it is a graph of a

1-resolvable TD(r, r).

Proof. Since the regularity and bipartiteness are determined from the spectrum [1, p. 87, 94], let us

first suppose that G = (U, V, E) is a regular, bipartite graph with the spectrum [r,√rr(r−1)

, 02(r−1),

−√rr(r−1)

, −r], and let A be its adjacency matrix. We have p = r2 and q = 0, and consequently

qu = 0, so that for each u ∈ U its nuv family satisfies∑v∈U\{u}

nuv = r(r − 1) = ∑v∈U\{u}

n2uv.

This system of Diophantine equations has a unique solution in nonnegative integers:

{nuv|v ∈ V\{u}} = {1r(r−1), 0r−1}.Hoffman’s equation for G reads

A4 + rA3 − rA2 − r2A = r(r − 1)J.

Page 12: Graphs with five eigenvalues

Author's personal copy

2622 D. Stevanovic / Linear Algebra and its Applications 435 (2011) 2612–2625

Let a1 ∈ U be an arbitrary vertex of U, and let a2, . . . , ar be the vertices having no common neighbors

with a1. Suppose thatnak,al = 1 for some k �= l, and let b a commonneighbor of ak and al . Then a1 is not

adjacent to b (since na1,ak = na1,al = 0), so that Aa1,b = (A2)a1,b = (A4)a1,b = 0 (last two equalities

hold since a1 and b are in different parts of G). Thus, from Hoffman’s equation A3a1,b

= r − 1. On the

other hand, we have that A3a1,b

� r − 2, as b is adjacent to at most r − 2 vertices from V\{a1, . . . , ar}and only one walk of length three from b to a1 can go over any such neighbor of b. This contradiction

shows that nak,al = 0 holds for any two k �= l.

Therefore, we can partition the set U into subsets U1, . . . ,Ur , having r vertices each, such that

any two vertices from the same subset Ui have no common neighbors, while any two vertices from

different subsets Ui and Uj have exactly one common neighbor. Similarly, the set V can be partitioned

into subsets V1, . . . , Vr with analogous properties.

Next, let Nv ⊂ U be the set of neighbors of each v ∈ V and let B be the collection of neighborhoods

by {Nv : v ∈ V}. The triple (U, {U1, . . . ,Ur}, B) is a transversal design TD(r, r), as every unordered

pair {s, t} of vertices inU either has no commonneighbor and is thus contained in exactly one groupUi,

or it has one common neighbor b and is thus contained in exactly one block Nb.

Moreover, the blocks ofB can be partitioned into setsB1, . . . , Br , withBj = {Nv : v ∈ Vj}. Since notwo vertices in Vj have a common neighbor, and

∑v∈Vj

|Nv| = r · r = |U|, it follows that∪v∈VjNv = U.

Thus, every element ofU occurs exactly once in eachBj , so that the triple (U, {U1, . . . ,Ur}, B) is indeeda 1-resolvable transversal design TD(r, r).

On the other hand, let (U, G, B) be a 1-resolvable transversal design TD(r, r) and let G = (U, B, E)be its graph having the adjacency matrix A. Let G be a partition of U into groups U1, . . . ,Ur , each of

size r. Since each of(r2

2

)− r

(r

2

)unordered pairs of elements from different groups is contained in

exactly one block, and each block contains(r

2

)unordered pairs, the number of blocks is equal to r2.

Thus, the number of edges in E is r3.

Further, let u be an arbitrary element of U. Suppose that u ∈ Ui and consider r unordered pairs

{{u, u′} : u′ ∈ Uj} for some fixed j �= i. Each of these pairs is contained in exactly one block.Moreover,

these pairs have to belong to different blocks: otherwise, if {u, u′} and {u, u′′} belong to the same

block B, then the pair {u′, u′′} belongs both to the group Uj and to the block B, a contradiction. Thus,

u belongs to at least r blocks, or in other words, the degree of u in G is at least r. However, it cannot

be larger than r, otherwise the total number of edges in E would be larger than r3. Therefore, G is a

bipartite and r-regular graph with r2 vertices in each part.

Since (U, G, B) is 1-resolvable, let B be partitioned into sets B1, . . . , Br , consisting of r blocks each,

such that every element of U occurs exactly once in each Bi. Thus, any two blocks belonging to the

same set Bi have to be disjoint, or in other words, they have no common neighbors in G. On the other

hand, B′ ∈ Bi and B′′ ∈ Bj for i �= j cannot have two common neighbors u′ and u′′ in G, as then the

pair {u′, u′′} would belong to two different blocks, a contradiction. Thus, B′ and B′′ have at most one

common neighbor. Further, from (2)∑B′′∈B\Bi

nB′,B′′ = r(r − 1),

and since nB′,B′′ � 1 and |B\Bi| = r(r − 1), it follows that nB′,B′′ = 1 for each B′′ ∈ B\Bi.

As nu,v is actually equal to (A2)u,v, from the above we see that

A2 = rI2r2 +⎛⎝ (Jr − Ir) ⊗ Jr Or2

Or2 (Jr − Ir) ⊗ Jr

⎞⎠ (14)

holds for suitable ordering of vertices ofG, where⊗ is the Kronecker product. Specifically, the ordering

of vertices of G is given by U1, . . . ,Ur, B1, . . . , Br , while the vertices within any of these sets are or-

dered arbitrarily. Since thematrix on the right-hand side of (14) has spectrum [(r2)2, r2r(r−1), 02(r−1)]

Page 13: Graphs with five eigenvalues

Author's personal copy

D. Stevanovic / Linear Algebra and its Applications 435 (2011) 2612–2625 2623

(where (r2)2 clumsily denotes the eigenvalue r2 of multiplicity two), it follows that G as a bipartite

graph with spectrum symmetric with respect to 0 [1, p. 87], has the spectrum [r2, √rr(r−1)

, 02(r−1),

−√rr(r−1)

, −r2].The previous theorem does not exclude the possibility that there exists nonisomorphic cospectral

graphswith this spectrum. Itmerely says that each suchgraphbelongs to some1-resolvable transversal

design TD(r, r). Further, the existence of a 1-resolvable transversal design TD(r, r) is equivalent to the

existence of the complete set of mutually orthogonal Latin squares of size r [64, Lemma 4.15], [5,

Theorem III.3.18]. Accordingly, given the complete set M1, . . . ,Mr−1 of mutually orthogonal Latin

squares of size r, this equivalence leads to the following construction of a graph with the spectrum

[r2, √rr(r−1)

, 02(r−1), −√rr(r−1)

, −r2]:LetU1, . . . ,Ur and V1, . . . , Vr be the sets of r vertices each and letU = ∪iUi, V = ∪jVj . The bipartite

graph is constructed with U and V as its color classes, and with the following edges:

• for each j, k, vertex j of U1 is adjacent to vertex j of Vk;• for each i � 2 and each j, k, vertex j of Ui is adjacent to vertexMij,k of Vk .

5. Some constructions of new bipartite, regular graphs with five eigenvalues

In this section, we briefly mention a few ways of constructing bipartite, regular graphs with five

eigenvalues.

First, recall that the direct product G × H of graphs G and H has as its eigenvalues all prod-

ucts of the form λμ, where λ is an eigenvalues of G and μ is an eigenvalue of H [1]. New bipar-

tite, regular graphs with five eigenvalues can be obtained as K2 × H if H is either a regular graph

with four eigenvalues whose second largest and the smallest eigenvalues have equal absolute val-

ues, or the complement of aKm for some natural a,m. In latter case, K2 × aKm has the spectrum

[m(a − 1),ma−1, 02(m−1)a, −ma−1, −m(a − 1)].Next, if G is a graph with adjacency matrix A, then G⊗ Jm denotes the graph with adjacency matrix

A⊗ Jm [6]. If G is connected, bipartite and r-regular, then so is G⊗ Jm. If G has n vertices and spectrum

[r, xy, 02z, −xy, −r], where z is possibly zero, then G ⊗ Jm has nm vertices and spectrum

[rm, (xm)y, 02z+(m−1)n, (−xm)y, −rm].So, if we have a connected, bipartite, regular graph with four eigenvalues (which are necessarily the

incidence graphs of symmetric 2-(v, k, λ) designs [1, p. 166]), or a connected, bipartite, regular graph

with five eigenvalues, then this construction produces a bigger graph with five eigenvalues.

The last construction we mention is taking the bipartite complement. The bipartite complement

Gbof a graph G = (U, V, E) is the graph with the same vertex set such that u ∈ U and v ∈ V are

adjacent in Gbif and only if they are not adjacent in G. If G has adjacency matrix⎛⎝ 0 B

BT 0

⎞⎠ ,

then its bipartite complement Gbhas adjacency matrix⎛⎝ 0 J − B

J − BT 0

⎞⎠ .

Let G = (U, V, E) be a bipartite, r-regular graph with 2p vertices, and let its eigenvalues be r =λ1 � λ2 � · · · � λ2p = −r. (G is not necessarily connected, so r need not be a simple eigenvalue.)

Page 14: Graphs with five eigenvalues

Author's personal copy

2624 D. Stevanovic / Linear Algebra and its Applications 435 (2011) 2612–2625

The eigenvector of G corresponding to r is

⎛⎝ j

j

⎞⎠, and the eigenvector corresponding to −r is

⎛⎝ j

−j

⎞⎠.

Any other eigenvector

⎛⎝ x

y

⎞⎠, corresponding to an eigenvalue λi, 2 � i � n− 1, is orthogonal to these

two vectors and thus, satisfies jT x = jT y = 0.

The bipartite complement Gbis (p − r)-regular, and has

⎛⎝ j

j

⎞⎠ and

⎛⎝ j

−j

⎞⎠ as eigenvectors of p − r

and −(p − r), respectively. For any other eigenvector

⎛⎝ x

y

⎞⎠, corresponding to an eigenvalue λi of G,

2 � i � n − 1, we have⎛⎝ 0 J − B

J − BT 0

⎞⎠ ⎛⎝ x

y

⎞⎠ =⎛⎝ Jy − By

Jx − BTx

⎞⎠ =⎛⎝ −By

−BTx

⎞⎠ = −λi

⎛⎝ x

y

⎞⎠ .

Thus, Gbhas

⎛⎝ x

y

⎞⎠ as an eigenvector with eigenvalue −λi. Since the spectrum of G is symmetric

around 0, we conclude that Gbhas the same spectrum as G, except that the pair of eigenvalues r, −r

of G is replaced with eigenvalues p − r, −(p − r).

If G is connected, then r, −r are simple eigenvalues of G, and Gbis another connected, bipartite,

regular graph with five eigenvalues, unless G ∼= aKm,mbfor some natural a,m. If G is not connected,

then r, −r appear together with p − r, −(p − r) in the spectrum of Gb. For p �= 2r, this implies that

Gbhas five eigenvalues if and only if G ∼= aKr,r for some natural a, in which case G

bhas the spectrum

[r(2a − 1), ra−1, 02(r−1)a, −ra−1, −r(2a − 1)].

References

[1] D. Cvetkovic, M. Doob, H. Sachs, Spectra of Graphs—Theory and Applications, Academic Press, New York, 1980.[2] M. Doob, W.H. Haemers, The complement of the path is determined by its spectrum, Linear Algebra Appl. 356 (2002) 57–65.

[3] E.R. van Dam, W.H. Haemers, Which graphs are determined by their spectrum?, Linear Algebra Appl. 373 (2003) 241–272.[4] E.R. van Dam, W.H. Haemers, Developments on spectral characterizations of graphs, Discrete Math. 309 (2009) 576–586.

[5] C.J. Colbourn, J.H. Dinitz, Handbook of Combinatorial Designs, second ed., Chapmann & Hall/CRC, Boca Raton, 2007.

[6] E.R. van Dam, Regular graphs with four eigenvalues, Linear Algebra Appl. 226–228 (1995) 139–162.[7] E.R. van Dam, E. Spence, Small regular graphs with four eigenvalues, Discrete Math. 189 (1998) 233–257.

[8] E.R. van Dam, E. Spence, Combinatorial designs with two singular values. II. Partial geometric designs, Linear Algebra Appl.396 (2005) 303–316.

[9] F.C. Bussemaker, D.M. Cvetkovic, There are exactly 13 connected, cubic, integral graphs, Publ. Fac. Electrotech. Univ. BelgradeSer. Math. Phys. 544–576 (1976) 43–48.

[10] Y. Xu, J. Meng, Gaussian integral circulant graphs, Discrete Math. 311 (2011) 45–50.

[11] M. Petkovic, M. Bašic, Further results on the perfect state transfer in integral circulant graphs, Comput. Math. Appl., 61 (2011)300–312.

[12] Z. Tang, Y. Hou, The integral graphs with index 3 and exactly two main eigenvalues, Linear Algebra Appl. 433 (2010) 984–993.[13] A. Ilic, Distance spectra and distance energy of integral circulant graphs, Linear Algebra Appl. 433 (2010) 1005–1014.

[14] W. Klotz, T. Sander, Integral Cayley graphs over abelian groups, Electron. J. Combin. 17 (2010) 1–13.[15] M. Bašic, M. Petkovic, Perfect state transfer in integral circulant graphs of non-square-free order, Linear Algebra Appl. 433

(2010) 149–163.[16] A. Ilic, M. Bašic, On the chromatic number of integral circulant graphs, Comput. Math. Appl. 60 (2010) 144–150.

[17] P. Carvalho, P. Rama, Integral graphs and (κ , τ )-regular sets, Linear Algebra Appl. 432 (2010) 2409–2417.

[18] M.A.A. de Freitas, N.M.M. de Abreu, R.R. Del-Vecchio, S. Jurkiewicz, Infinite families of Q-integral graphs, Linear Algebra Appl.432 (2010) 2352–2360.

[19] S.K. Simic, Z. Stanic, On Q-integral (3, s)-semiregular bipartite graphs, Appl. Anal. Discrete Math. 4 (2010) 167–174.[20] G.R. Omidi, On integral graphs with few cycles, Graphs Combin. 25 (2010) 841–849.

Page 15: Graphs with five eigenvalues

Author's personal copy

D. Stevanovic / Linear Algebra and its Applications 435 (2011) 2612–2625 2625

[21] R.J. Angeles-Canul, R.M. Norton,M.C. Opperman, C.C. Paribello,M.C. Russell, C. Tamon, Perfect state transfer, integral circulants,

and join of graphs, Quantum Inf. Comput. 10 (2010) 0325–0342.[22] L. Wang, Q. Wang, Integral complete multipartite graphs Ka1 ·p1,a2 ·p2,...,as ·ps with s = 5, 6, Discrete Math. 310 (2010) 812–818.

[23] S. Kirkland, M.A.A. de Freitas, R.R. Del Vecchio, N.M.M. de Abreu, Split non-threshold Laplacian integral graphs, Linear andMultilinear Algebra 58 (2010) 221–233.

[24] K.C. Das, S.G. Lee, G.S. Cheon, On the conjecture for certain Laplacian integral spectrum of graphs, J. Graph Theory 63 (2010)

106–113.[25] M. Bašic, M. Petkovic, Some classes of integral circulant graphs either allowing or not allowing perfect state transfer, Appl.

Math. Lett. 22 (2009) 1609–1615.[26] A. Abdollahi, E. Vatandoost, Which Cayley graphs are integral? Electron. J. Combin. 16 (2009) R122.

[27] M. Bašic, A. Ilic, On the clique number of integral circulant graphs, Appl. Math. Lett. 22 (2009) 1406–1411.[28] T. Sander, Sudoku graphs are integral, Electron. J. Combin. 16 (2009) N25.

[29] M. Bašic, M. Petkovic, D. Stevanovic, Perfect state transfer in integral circulant graphs, Appl. Math. Lett. 22 (2009) 1117–1121.[30] O. Ahmadi, N. Alon, I.F. Blake, I.E. Shparlinski, Graphs with integral spectrum, Linear Algebra Appl. 430 (2009) 547–552.

[31] Z. Stanic, Some results on Q-integral graphs, Ars Combin. 90 (2009) 321–335.

[32] L. Wang, H. Broersma, C. Hoede, X. Li, G. Still, Some families of integral graphs, Discrete Math. 308 (2008) 6383–6391.[33] N.M.M. de Abreu, C.T.M. Vinagre, A.S. Bonifacio, I. Gutman, The Laplacian energy of some Laplacian integral graphs, MATCH 60

(2008) 447–460.[34] S.K. Simic, Z. Stanic, Q-integral graphs with edge-degrees at most five, Discrete Math. 308 (2008) 4625–4634.

[35] S. Kirkland, Laplacian integral graphs with maximum degree 3, Electron. J. Combin. 15 (2008) 1–42.[36] L. Wang, X. Liu, Integral complete multipartite graphs, Discrete Math. 308 (2008) 3860–3870.

[37] P. Hic, M. Pokorny, Integral complete 4-partite graphs, Discrete Math. 308 (2008) 3704–3705.

[38] R. Grone, R. Merris, Indecomposable Laplacian integral graphs, Linear Algebra Appl. 428 (2008) 1565–1570.[39] L. Wang, C. Hoede, Constructing fifteen infinite classes of nonregular bipartite integral graphs, Electron. J. Combin. 15 (2008)

R8.[40] L. Wang, H. Sun, Three classes of bipartite integral graphs, Lect. Notes Comput. Sci. 4381 (2007) 206–215.

[41] N. Saxena, S. Severini, I. Shparlinski, Parameters of integral circulant graphs and periodic quantum dynamics, Int. J. QuantumInf. 5 (2007) 417–430.

[42] P.E. Gunnells, R.A. Scott, B.L.Walden, On certain integral Schreier graphs of the symmetric group, Electron. J. Combin. 14 (2007)

1–26.[43] S. Kirkland, Constructably Laplacian integral graphs, Linear Algebra Appl. 423 (2007) 3–21.

[44] D. Stevanovic, N.M.M. de Abreu, M.A.A. de Freitas, R. Del-Vecchio, Walks and regular integral graphs, Linear Algebra Appl. 423(2007) 119–135.

[45] L.S. de Lima, N.M.M. de Abreu, C.S. Oliveira, M.A.A. de Freitas, Laplacian integral graphs in S(a, b), Linear Algebra Appl. 423(2007) 136–145.

[46] M. Petrovic, B. Borovicanin, Z. Radosavljevic, The integral 3-harmonic graphs, Linear Algebra Appl. 416 (2006) 298–312.

[47] W. So, Integral circulant graphs, Discrete Math. 306 (2006) 153–158.[48] M. Lepovic, On integral graphs which belong to the class αKa,a ∪ βKb,b , J. Appl. Math. Comput. 20 (2006) 61–74.

[49] L. Wang, X. Li, C. Hoede, Two classes of integral regular graphs, Ars Combin. 76 (2005) 303–319.[50] S. Kirkland, Completion of Laplacian integral graphs via edge addition, Discrete Math. 295 (2005) 75–90.

[51] S. Barik, S. Pati, On algebraic connectivity and spectral integral variations of graphs, Linear Algebra Appl. 397 (2005) 209–222.[52] K.T. Balinska, S.K. Simic, K.T. Zwierzynski, Which non-regular bipartite integral graphs withmaximum degree four do not have

±1 as eigenvalues? Discrete Math. 286 (2004) 15—24.

[53] M. Lepovic, On integral graphs which belong to the class αKa ∪ βKb,b , Discrete Math. 285 (2004) 183–190.[54] L. Wang, X. Li, C. Hoede, Integral complete r-partite graphs, Discrete Math. 283 (2004) 231–241.

[55] M. Lepovic, On integral graphs which belong to the class αKa ∪ βKb , J. Appl. Math. Comput. 14 (2004) 39–49.[56] M. Lepovic, On integral graphs which belong to the class αKa,b , Graphs Combin. 19 (2003) 527–532.

[57] Y.Z. Fan, On spectral integral variations of mixed graphs, Linear Algebra Appl. 374 (2003) 307–316.[58] F. Yizheng, Spectral integral variations of degree maximal graphs, Linear and Multilinear Algebra 51 (2003) 147–154.

[59] F. Yizheng, On spectral integral variations of graphs, Linear and Multilinear Algebra 50 (2002) 133–142.

[60] K.T. Balinska, S.K. Simic, The nonregular, bipartite, integral graphs with maximum degree 4. Part I: basic properties, DiscreteMath. 236 (2001) 13–24.

[61] A.J. Hoffman, On the polynomial of a graph, Amer. Math. Monthly 70 (1963) 30–36.[62] D. Stevanovic, Some compositions of graphs and integral graphs, Ph.D. Thesis (supervised by Dragoš Cvetkovic), University of

Niš, Serbia, 2000 (in Serbian).[63] B. McKay, graph6 and sparse6 graph formats. Available from: <http://cs.anu.edu.au/ bdm/data/formats.html>.

[64] C.J. Colbourn, J.H. Dinitz,Mutually orthogonal latin squares: a brief survey of constructions, J. Statist. Plann. Inference 95 (2001)9–48.