Weakly Distance-Regular Digraphs

F. Comellas1, M. A. Fiol, J. Gimbert2, and M. Mitjana3

September 9, 2002









Running Head:
weak distance-regularity

Mailing Address:
M.A. Fiol
Departament de Matemàtica Aplicada IV
Universitat Politècnica de Catalunya
Jordi Girona 1-3
Mòdul C3, Campus Nord
08034 Barcelona, SPAIN.
Email:
fiol@mat.upc.es

Abstract

We introduce the concept of weakly distance-regular digraph and study some of its basic properties. In particular, the (standard) distance-regular digraphs, introduced by Damerell, turn out to be those weakly distance-regular digraphs which have a normal adjacency matrix. As happens in the case of distance-regular graphs, the study is greatly facilitated by a family of orthogonal polynomials called the distance polynomials. For instance, these polynomials are used to derive the spectrum of a weakly distance-regular digraph. Some examples of these digraphs, such as the butterfly and the cycle prefix digraph which are interesting for their applications, are analyzed in the light of the developed theory. Also, some new constructions involving the line digraph and other techniques are presented.

1  Introduction

Let us first introduce some basic notation (see [4,17]) and discuss the relevant background. Let G =(V,E) be a (strongly) connected digraph with order N:=|V| and diameter D. We will denote by dist (u,v) the distance from vertex u to vertex v. Notice that dist (u,v) and dist (v,u) may not be equal since we are considering oriented graphs. For any fixed integer 0 £ k £ D, we will denote by G+k(v) (respectively, G-k(v)) the set of vertices at distance k from v (respectively, the set of vertices from which v is at distance k). Sometimes it is written, for short, G+(v) or G-(v) instead of G1+(v) or G1-(v), respectively. Thus, the out-degree and in-degree of v are d+(v):=|G+(v)| and d-(v):=|G-(v)|; and the digraph G is (D-)regular if d+(v)=d-(v)=D for every v Î V.

Distance-transitivity and distance-regularity

The concept of a distance-regular digraph was introduced by Damerell [9] in the late seventies. He defined distance-regular digraphs by using a regularity type condition concerning the cardinality of some vertex subsets. More precisely, a digraph G with diameter D is distance-regular if, for any pair of vertices u,v Î V such that dist (u,v)=k, 0 £ k £ D, the numbers
si1k(u,v):=|Gi+ (u)ÇG1+ (v)|,
(1)
for each i such that 0 £ i £ k+1, do not depend on the chosen vertices u and v, but only on their distance k. In this case, we just write si1k and refer to them as the intersection numbers. (According to Damerell's notation [9], si1k would be just sik.) As happens in the undirected case, the notion of distance-regularity in digraphs can be seen as a generalization of the concept of distance-transitivity. The concept of distance-transitive digraphs was introduced by Lam in [19] as a natural generalization of distance-transitive graphs. A digraph G is said to be distance-transitive if for all vertices u,v,x,y such that dist (u,v)= dist (x,y) there is an automorphism p of G such that p(u)=x and p(v)=y. That paper provided some examples of distance-transitive digraphs with girth g and diameter D=g-1, as the directed cycle and the Paley tournament, and presented a method to construct more distance-transitive digraphs, with D=g, starting from a distance-transitive digraph with D=g-1. Damerell [9] proved that every distance-transitive digraph G with girth g, say, is stable; that is, dist (u,v)+ dist (v,u)=g for any pair of vertices u,v Î V(G) at distance 0 < dist (u,v) < g. (Thus, a stable digraph of girth two can be identified with the graph obtained by replacing each pair of opposite arcs, or `digon', by an undirected edge.) Consequently, every distance-transitive digraph has diameter D=g (`long' distance-transitive digraph ) or D=g-1 ( `short' distance-transitive digraph ), provided that g ³ 3. Moreover, Damerell showed that the classification of these digraphs can be reduced to the case g=D+1 (i.e., the short digraphs) since every long digraph is obtained from a short digraph, with the same girth, by the construction described by Lam. Using these results, Bannai, Cameron and Kahn [1] proved that the distance-transitive digraphs given by Lam were the only ones with odd girth.
The above results were extended by Damerell to the case of distance-regular digraphs. He also gave the classification in `short' and `long' digraphs, and the structure of the long digraphs. Taking into account that the case g=2 corresponds to the distance-regular graphs, the study of the distance-regular digraphs is reduced to the case g ³ 3 (and g=D+1). In particular, there is a correspondence between the distance-regular digraphs with girth g=3 and diameter D=2 and the so-called skew Hadamard matrices . For g=4,5,6, there are some works where the intersection numbers are computed by using the smallest number of defining parameters (see e.g. Enomoto and Mena [12] and Liebler and Mena [21]). Finally, Douglas and Nomura [10] showed that the girth g of a distance-regular digraph with degree D > 1 and diameter D always satisfies D £ g £ 8.
If we change G1+(v) to G1-(v) in the definition of distance-regularity, we get some new parameters pi1k, called again the intersection numbers. But now the stability property does not necessarily hold, and a class of digraphs with less structure appears. We focus our attention on such digraphs, here referred to as `weakly distance-regular', and study some of their properties, which are closely related to the properties enjoyed by the distance-regular digraphs. In fact, the diameter 2 weakly distance-regular digraphs are the same as the `directed strongly regular graphs' introduced by Duval [11], which have recently been studied by a number of authors and found several constructions, see [5]. In our more general context, and besides several characterizations given in the next section, it is shown that every (standard) distance-regular digraph is also weakly distance-regular, so justifying the name. In this case, the relationship between the defining parameters, si1k and pi1k, shows to be very simple. Subsequently, we devote Section 4 to the study of the spectra of weakly distance-regular digraphs, providing formulas for computing the eigenvalue multiplicities from the entries of some different `small' matrices. Finally, some constructions and specific families of weakly distance-regular digraphs, which are interesting for their applications in network theory, are analyzed.
Since most of our study is of an algebraic nature, we next recall some background from algebraic graph theory and matrix theory. The adjacency matrix A =(auv) of a digraph G (we assume4 that G contains neither loops nor multiple edges) is the N×N matrix indexed by the vertices of G, with entries auv = 1 if u is adjacent to v, and auv = 0 otherwise. The spectrum of the digraph G, denoted by sp G, consists of the eigenvalues of A, which might be not real since A is not symmetric, together with their (algebraic) multiplicities.
sp G={l0m(l0),l1m(l1),...,ldm(ld)}.
The distance-k matrix Ak of a digraph G with diameter D, where 0 £ k £ D, is defined by
(Ak)uv: = ì
í
î
1
if dist (u,v)=k,
0
otherwise.
(2)
In particular, A0=I and A1=A. As we will see, these matrices play a key role in the study of distance-regularity. For instance, it is known that a digraph G is distance-regular (in the sense of Damerell) if and only if the set of matrices {Ak}k=0D satisfies the axioms of a commutative association scheme; see Takahashi [26]. (The readers who are not familiar with association schemes can consult, for instance, Godsil's textbook [17].) This important result is going to be used later.
Some other basic results from algebraic graph theory are the following:
  1. By the Perron-Frobenius theorem, the maximum eigenvalue l0 of G is simple and has a positive eigenvector n. In particular, if G is D-regular, then n =j, where j denotes the all-1 vector, and l0=D.

  2. The number of walks of length l ³ 0 from vertex u to vertex v is equal to auv(l):=(Al)uv.

  3. The adjacency algebra of G (also called Bose-Mesner algebra when it is closed under the Hadamard-or componentwise-product) is defined by
    A(G):={p(A) : p Î \mathbbC[x]},
    where A is the adjacency matrix of G. The dimension of A(G), as a \mathbbC-vector space, equals the degree of the minimum polynomial mG of G, and it is at least D+1 since the powers I,A,¼,AD are linearly independent.

  4. A digraph G is (strongly) connected and regular if and only if there is a polynomial p Î \mathbbQ[x] such that p(A)=J, where J denotes the all-1 matrix (see [6], Th. 5.3.1). The polynomial HG of least degree satisfying this property is called the Hoffman polynomial of G and it is given by
    HG:=  N

    S(D)
    S,
    (3)
    where mG:=(x-D)S Î \mathbbZ[x] is the minimum polynomial of G.

We end this introductory section by recalling some results from matrix theory (see e.g. [20]). Given a complex matrix A, we denote by A* the transpose of its conjugate. Thus, A is hermitian if A*=A, and it is unitary when AA*=I. Moreover, A is said to be normal if AA*=A*A. The following result summarizes several characterizations of normal matrices.
Theorem 1 Let A be an n×n complex matrix with eigenvalues q1,q2,¼,qn. Then A is normal if and only if any of the following assertions hold:
Note that, from (a), the eigenvectors of a normal n×n square matrix constitute an orthogonal basis of the vector space \mathbbCn, with the Hermitian scalar product áu,vñ*:=uT[`(v)]. Let us now assume that A has distinct eigenvalues l0,l1,¼,ld. For each li, let Ui be the matrix whose columns form an orthonormal basis of the eigenspace Ei:= Ker (A-liI). Then the orthogonal projection onto Ei is represented by the matrix Ei:=UiUi* or, alternatively,
Ei=  1

fi

Õ
j ¹ i 
(A-ljI)       (0 £ i £ d),
(4)
where fi:=Õj=0(j ¹ i)d(li-lj). These (hermitian) matrices are called the (principal ) idempotents of A, and satisfy the following properties: EiEj=dijEi; AEi=liEi; sp Ei={1m(li),0n-m(li)}; and, for every rational function f defined at each li, 0 £ i £ d,
f(A)= d
å
i=0 
f(li)Ei.
(5)
In particular, when f=x, the above equation becomes the so-called spectral decomposition theorem: A =åi=0d liEi.

2  Weak distance-regularity

We begin this section by formally introducing the concept of weakly distance-regular digraph, through the invariance of the number of walks between vertices at a given distance. (This approach has already been used to characterize distance-regular graphs, see e.g. Rowlinson [25].) Afterwards, and as already mentioned, we will show that this concept is equivalent to change G1+(v) to G1-(v) in the definition of a distance-regular digraph.
Definition 1 A digraph G of diameter D is weakly distance-regular if, for each non-negative integer l £ D, the number auv(l) of walks of length l from vertex u to vertex v only depends on their distance dist (u,v)=k, for any l = 0,1,..., D. In this case we write auv(l)=akl for some constants akl, 0 £ k, l £ D.
Of course akl=0 for any k > l, since there is no walk of length l between vertices u,v at distance dist (u,v) > l. Moreover, if the digraph is geodetic (that is, the shortest path between any two vertices is unique), then akk=1 for any 0 £ k £ D.
>From the definition of G as a weakly distance-regular digraph , it follows that each matrix power Al, 0 £ l £ D, can be expressed as a linear combination of the distance matrices A0,A1,...,AD which, as we will see next, belong to the adjacency algebra of G. The following theorem provides some characterizations of weakly distance-regular digraphs which are analogous to those applying for distance-regular graphs (for a survey of the latter, see e.g. [15]).
Theorem 2 Let G be a connected digraph with diameter D and distance matrices {Ak}k=0D. Then, the following are equivalent:
Proof.   (w)Û (a) If G is weakly distance-regular, the number of walks of length l £ D from vertices u,v at distance k is auv(l)=akl (a constant). Hence,
Al = a0lI+a1lA+a2lA2+...+allAl       (0 £ l £ D)
(7)
where, necessarily, all ¹ 0 and, as already mentioned, akl=0 for any k > l. In matrix form,
æ
ç
ç
ç
ç
ç
ç
è
I
A
A2
·
·
AD
ö
÷
÷
÷
÷
÷
÷
ø
= æ
ç
ç
ç
ç
ç
ç
è
a00
a01
a11
a02
a12
a22
·
·
·
·
·
·
·
·
·
a0D
a1D
·
·
·
aDD
ö
÷
÷
÷
÷
÷
÷
ø
æ
ç
ç
ç
ç
ç
ç
è
I
A
A2
·
·
AD
ö
÷
÷
÷
÷
÷
÷
ø
(8)
where C:=(akl) is a lower triangular matrix. Since all > 0 for any 0 £ l £ D, the matrix C is non-singular and its inverse C-1 is also a lower triangular matrix. Hence Ak is a polynomial of degree k in A:
Ak=pk(A)=a0kI+a1kA+a2kA2+...+akkAk       (0 £ k £ D),
(9)
where akk ¹ 0. Conversely, let us assume that there are constants alk, with akk ¹ 0, satisfying (9). This implies that Eqs. (7) and (8) also hold with C =(akl) being the inverse matrix of C-1=(alk) and, hence, the number of walks of length l £ D from one vertex to another only depends on their distance.
(a)Þ (b) We know that {I,A,...,AD} constitutes a set of linearly independent matrices of the adjacency algebra A(G). Moreover, since Ak=pk(A) and åk=0D Ak=J, it follows that H(A)=J, where H:=åk=0D pk. As a consequence, G is a regular digraph of degree D, say. Furthermore, since the minimum polynomial mG of G divides (x-D)H we have D+1 £ dimA(G) = dgr mG £ D+1. Hence, the dimension of A(G) attains the minimum value allowed by the diameter, dimA(G)=D+1, and both {Al}l=0D and {Ak}k=0D are bases of A(G).
(b)Þ (c) Let u and v be two vertices of the digraph G, such that dist (u,v)=k. Notice that the number pijk(u,v), representing the number of vertices at distance i from u and at distance j to v, coincides with the (u,v)-entry of the matrix AiAj which, because of (b), is a linear combination of the basis {Ak}k=0D, say,
AiAj= D
å
k=0 
gijk Ak.
(10)
Consequently, pijk(u,v)=gijk=pijk for any two vertices u,v at distance k.
(c)Þ (a) If (10) holds for j=1:
AiA1= D
å
k=0 
pi1k Ak       (0 £ i £ D),
(11)
we can use an inductive argument, starting from A0=I and A1=A, to deduce that the distance matrix Ak is indeed a polynomial of degree k in the adjacency matrix A.   [¯]
>From now on, the polynomials pk such that Ak=pk(A), 0 £ k £ D, will be referred to as the distance polynomials of G. Notice that, from (b), any matrix power Al is a linear combination of the distance matrices and, consequently, the number auv(l) of walks of any fixed length l ³ 0 between vertices u,v of G only depends on their distance. Other interesting consequences of the above theorem are the following.
Corollary 3 Let G be a weakly distance-regular digraph on N vertices, with diameter D, adjacency matrix A, and distance polynomials {pk}k=0D. Then,
Proof.   (i)-(ii) From the proof of Theorem 2.2(b), we see that HG=H=åk=0D pk, G is a D-regular digraph, and the minimum polynomial mG is as claimed. Moreover, we have that j =(1,1,... ,1)T is an eigenvector of A corresponding to the eigenvalue D. Consequently, Akj =pk(A)j =pk(D)j, which implies that HG(D)=åk=0Dpk(D)=N, in concordance with (3).
(iii) This follows from the fact that the characteristic polynomial of G is uniquely determined by the minimum polynomial mG and the spectral invariants tr Al=N a0l, 0 £ l £ D (see Section 3 for more details).   [¯]
As we will see, in the case of weakly distance-regularity, the average of the intersection numbers (1) also play a role. In a given digraph with diameter D, let Nk, 0 £ k £ D, denote the number of (ordered) vertex pairs u,v such that dist (u,v)=k; that is, Nk:=åu Î V nk(u). Thus, N0=N and N1=|E|. Then the mean intersection numbers [`s]i1k are defined by

s
 
k
i1 
:=  1

Nk

å
dist (u,v)=k 
si1k(u,v)
(12)
and, more generally, we define [`s]ijk as the average of the numbers sijk(u,v):=|Gi+(u)ÇGj+(v)|.
To see what is the significance of these parameters, let us consider, for a given digraph G with adjacency matrix A, the following scalar product in \mathbbC[x]:
áf,gñG:=  1

N
tr (f(A)g(A)*).
(13)
In fact, notice that this product is well defined in the quotient ring \mathbbC[x]/Á, where Á:=(mG) is the ideal generated by the minimum polynomial of G. This is due to the additive property of the trace and tr (mG(A))= tr Ø = 0.
Then, if G is a weakly distance-regular digraph, its distance polynomials are orthogonal with respect to this product. Indeed,
ápk,plñG=  1

N
tr (AkAlT) =  1

N

å
u Î V 
|Gk+(u)ÇGl+(u) |=0        (k ¹ l),
(14)
and
||pk||G2=  1

N
tr (AkAkT)=nk=pk(D).
(15)
Consequently, the set {pk}k=0D is a basis of the quotient ring \mathbbC[x]/Á, which is isomorphic to A(G). The representation of pix in terms of such a basis must be of the form
pix = D
å
k=0 
gik pk       (0 £ i £ D)
where gik is the corresponding Fourier coefficient. Comparing with (11), we conclude that gik must be equal to the intersection number pi1k (a real number). Hence,
gik=

gik
 
=  ápix,pkñG

||pk||G2
=  ápk,pixñG

||pk||2G
       (0 £ i £ D)).
(16)
In the following result we show that these Fourier coefficients can also be computed from the average intersection numbers [`s]ijk.
Proposition 4 The coefficients of the above recurrence satisfied by the distance polynomials of a weakly distance-regular digraph are:
gik = pi1k=  Ni

Nk

s
 
i
k1 
.
Proof.   If we compute gik by using the first expression in (16), we get
gik
=
 ápix,pkñG

||pk||G2
=   tr (AiAAkT)

tr (AkAkT)
=

å
u Î V 

å
v Î V 
(Ai)uv(AAkT)vu


å
u Î V 

å
v Î V 
(Ak)uv(Ak)uv
=
 1

Nk

å
u Î V 

å
v Î Gi+(u) 
(AAkT)vu =  1

Nk

å
u Î V 

å
v Î Gi+(u) 
(AkAT)uv
=
 1

Nk

å
u Î V 

å
v Î Gi+(u) 
sk1i(u,v) =  Ni

Nk

s
 
i
k1 
.
This completes the proof, as we already know that gik=pi1k (in fact, this is what we get if we compute gik by using the second expression in (16)).   [¯]
In general, as a consequence of Theorem 2.2 and ápk,pipjñG=ápipj,pkñG, we reach the following conclusion:
Corollary 5 A distance-regular digraph G with intersection numbers sijk is also a weakly distance-regular digraph with intersection parameters pijk satisfying
Nk pijk=Ni skji.
(17)
Proof.  Since G is distance-regular, its adjacency algebra is a (P-polynomial) association scheme (that is, each distance matrix Ak, 0 £ k £ D, is a polynomial pk of degree k in A) and so Theorem 2.2(a) applies. This assures that the numbers pijk exist and the proof goes as before. (In fact, notice that (17) just counts in two ways the number of ordered triples (x,y,z) such that dist (x,y)=k, dist (x,z)=i, and dist (z,y)=j.)   [¯]
Thus, any distance-regular digraph G satisfies all the above properties of weakly distance-regular digraphs. Moreover, since G is stable and has girth g Î {2,D,D+1}, see [9], we have
AAT=AAg-1=Apg-1(A)=Ag-1A=ATA
(since 1 £ g-1 £ D) and A is a normal matrix.
As a consequence of the above, we now have the following characterization result, showing that the (standard) distance-regular digraphs are precisely those weakly distance-regular digraphs which are stable or have a normal adjacency matrix.
Proposition 6 Let G be a weakly distance-regular digraph with adjacency matrix A. Then G is distance-regular if and only if any of the following conditions hold:
Proof.   We only need to prove that each of the conditions is sufficient.
(a) Let G be a weakly distance-regular digraph with diameter D. From the proof of Theorem 2.2, we already know that AiAj=AjAi is a linear combination of the basis {Ak}k=0D, see (10). On the other hand, since G is stable, we have that AkT=Ag-k, 0 < k < g, where g is the girth of G. If g=2, then G is a symmetric digraph since AT = A and, consequently, AkT = Ak for each k=0,1,...,D. If g ³ 3, then g=D or g=D+1 (see [9,Theorem 2]), which implies that AkT = AD-k (0 < k < D) and ADT = AD, if g=D, and AkT = AD+1-k, if g=D+1 and 0 < k £ D. Thus, the adjacency algebra A(G) is closed under the "transposition operation". Hence, the set of matrices {Ak}k=0D satisfies the axioms of a (commutative) association scheme and G is a distance-regular digraph.
(b) If A is normal, then by Theorem 1.1(b) there exists a polynomial p Î \mathbbC[x] such that AT=p(A). Thus, AiAT Î A(G) and, hence, it admits a representation in terms of the basis {Ak}k=0D, which must be of the form
AiAT= D
å
k=0 
sikAk       (0 £ i £ D)
since (AiAT)uv=|Gi+(u)ÇG+(v)|. Therefore, as the Fourier coefficients sik only depend on k (and i), they coincide with the intersection numbers si1k and G is distance-regular.   [¯]
>From this result, we can see now that the scalar product associated to a distance-regular digraph has a simple expression in terms of its spectrum sp G ={l0m(l0),l1m(l1),...,ldm(ld)}. Indeed, using Theorem 1.1 and the properties of the idempotents, (13) becomes
áf,gñG=  1

N
d
å
i=0 
f(li)

g(li)
 
tr (EiEi*) =  1

N
d
å
i=0 
m(li) f(li)

g(li)
 
(18)
which, except for the conjugation, is like the scalar product used for distance-regular graphs (see [15]). In the next section, we will derive the corresponding expression for a weak distance-regular digraph.

3  The spectrum and the condensed matrices

In this section we study how to compute the spectrum of a weakly distance-regular digraph in terms of its defining parameters. We will see that the whole spectrum of such a digraph can be retrieved from the information given by some specific matrices characterizing it, such as the recurrence matrix or the multiplicity matrix, which have size much more smaller than the adjacency matrix. Thus, these matrices are here generically referred to as the `condensed matrices'.

The multiplicities

First recall that, by Corollary 2.3(i), the distinct eigenvalues of a weakly distance-regular digraph G with distance polynomials {pk}k=0D are l0=D and the different zeros l1,l2,¼,ld (d £ D) of its Hoffman polynomial HG=åk=0D pk. Moreover, their respective multiplicities can be computed by solving the system of linear equations
tr Aj= d
å
i=0 
m(li)lij = N a0j       (0 £ j £ d),
(19)
where the independent terms a0j=auu(j) can be computed from the coefficients of the distance polynomials (see the proof of Theorem 2.2); and the coefficient matrix is a Vandermonde matrix formed from the distinct values l0,l1,¼,ld.
However, in our context, a more direct way of computing the multiplicities is to consider the traces of the distance matrices. Indeed, since
tr Ai= tr (pi(A))= d
å
j=0 
m(lj)pi(lj)=d0i N        (0 £ i £ d)
(20)
(notice that tr Ai=Nápi,p0ñG) we have the following result:
Proposition 1 Let G be a weakly distance-regular digraph with N vertices, distance polynomials {pk}k=0D, and distinct eigenvalues l0(=D),l1,¼,ld. Let ¶ ev be the matrix whose (i,j)-element is pi(lj), 0 £ i,j £ d. Then the multiplicities m(li) are given by
m(li)=N(¶ ev -1)i0       (0 £ i £ d).
(21)
Proof.   Since ¶ ev is clearly non-singular, (20) yields
æ
ç
ç
ç
è
m(l0)
m(l1)
:
m(ld)
ö
÷
÷
÷
ø
= æ
ç
ç
ç
è
p0(l0)
p0(l1)
...
p0(ld)
p1(l0)
p1(l1)
...
p1(ld)
:
:
:
:
pd(l0)
pd(l1)
...
pd(ld)
ö
÷
÷
÷
ø
-1





 
æ
ç
ç
ç
è
N
0
:
0
ö
÷
÷
÷
ø
,
(22)
which corresponds to (21).   [¯] In the following subsections we study other matrices related to G, which also contain the information about its multiplicities. We begin with a positive definite matrix which gives an alternative expression for the scalar product (13).

The multiplicity matrix

Let G be a weakly distance-regular digraph with diameter D, and adjacency matrix A with distinct eigenvalues l0(=D),l1,¼,ld, where d £ D. Let m0,m1,¼,md denote the multiplicities of such eigenvalues, as zeros of the minimum polynomial mG. Thus, m0=1, 1 £ mi £ m(li) for any 1 £ i £ d, and åi=0d mi=D+1. For every polynomial f Î \mathbbC[x], let us define the (row) vector f =(f0,f1,¼,fD) Î \mathbbCD+1 as
f:= æ
è
f(l0),f(l1),f¢(l1),  f"(l1)

2!
,¼,  f(m1-1)(l1)

(m1-1)!
,¼,f(ld),f¢(ld),¼,  f(md-1)(ld)

(md-1)!
ö
ø
where the superscripts stand for the orders of the derivatives. Then we claim that there exists a positive definite matrix M =(mrs), 0 £ r,s £ D, such that the scalar product (13) can be written as
áf,gñG=  1

N
fMg* =  1

N
D
å
r,s=0 
mrs fr

gs
 
.
(23)
Moreover, the row (or column) sums of a given submatrix M ev of M yield the multiplicities of G. More precisely let us consider the set of indexes J:={0,åj=0i-1 mj: 1 £ i £ d} Í \mathbbZD+1, and let M ev denote the principal submatrix of M whose row and column indexes are the (d+1) elements of J. Then
d
å
j=0 
(M ev )ij = m(li)      (0 £ i £ d).
Since the distance polynomials form a basis, to prove the first statement it is enough to show that there is a matrix M satisfying (14) and (15); that is, [ 1/N]pkMpl*=dklpk(D) or, in matrix form,
 1

N
PM*=D
where ¶ is the matrix with rows p0,p1,¼,pD and D:= diag (p0(D),p1(D),¼,pD(D)). Thus, since the vectors p0,p1,¼,pD are linearly independent (otherwise, Hermite interpolation at l0,l1,¼,ld would give a contradiction) the matrix ¶ has an inverse and so
M=N¶-1D-1)*=VV*
(24)
where we have introduced the new (nonsingular) matrix V:=Ö-1D1/2. Hence, M is a definite positive hermitian matrix (note that, for any f ¹ 0, fMf*=||f||G2 > 0).
Furthermore,
tr Ak= d
å
i=0 
pk(li)m(li)=pkm*       (0 £ k £ D),
(25)
where m is a (D+1)-vector whose ri-th entry is m(li), when ri Î J is defined as r0:=0 and ri:=m0+m1+¼+mi-1, 1 £ i £ d; and mr=0 if r Ï J. Also,
tr Ak=Nápk, p0ñG=pkMp0*       (0 £ k £ D).
(26)
Thus, by (25) and (26),
pk(Mp0*-m*)=0       (0 £ k £ D)
and, by the independence of p0,p1,¼,pD, we conclude that Mp0*=m*, and hence M ev j = m ev * where m ev :=(m(l0),m(l1),¼,m(ld)). Of course, this gives a way of computing the multiplicities from the entries of M which, in turn, can be computed from the distance polynomials using (24). This gives:
m*
=
Mp0*=N¶-1D-1)*p0* = N¶-1D(p0-1)*
=
-1e1*
(27)
where e1 is the 1-th coordinate vector. Consequently, m* is just N times the first column of ¶-1. (A result to be compared with (22) which, with the above notation, reads m ev *=N¶ ev -1e1*.) A more direct way of seeing this is to consider again (25). Indeed, as tr A0=N and tr Ak=0 for any 1 £ k £ D, Eqs. (25) yield, in matrix form, ¶m*=Ne1*, whence (27) follows.
Let us now prove that m0s = 0 for any s ¹ 0 and m00=1. To this end, we now compute the scalar product áZie,HGñG, where Zie: = mG/(x-li)e, 1 £ e £ mi, in two different ways:
áZie,HGñG=  1

N
tr (Zie(A)HG(A)*) =  1

N
tr (Zie HG(A)) = ì
í
î
0
if i ¹ 0
Z01(D)
if i=0
(28)
since HG=[ N/(Z01(D))] Z01 and hence Zie HG is a multiple of mG for any i ¹ 0. Also, from the vectors Zie = (0,0,¼,0,Z(mi-e)ie(li),¼,Zie(mi-1)(li),0,¼,0) and HG=(N,0,0,¼, 0) (derived from the respective polynomials), we have
áZie,HGñG=  1

N
ri+mi-1
å
r=ri+mi-e 
mr0(Zie)r(HG)0 = ri+mi-1
å
r=ri+mi-e 
mr0Zie(r-ri)(li).
(29)
Thus, equating (28) and (29) for i=0 (whence r0=0 and e = 1) we get m00=1. Similarly, for i ¹ 0 and e = 1, we obtain mr0Zie(mi-1)(li)=0 where r=ri+mi-1. Hence, for such a value of r, we must have mr0=0, as Zie(mi-1)(li) ¹ 0. Recursively, for the same i and e = 2,¼, mi-1 we obtain mr0=0 for r=ri+mi-e.
As a consequence of the above, the scalar product (23) can be rewritten in a slightly more simplified form:
áf,gñG=  1

N
æ
è
f0

g0
 
+ D
å
r,s=1 
mrsfr

gs
 
ö
ø
.
(30)

The recurrence matrix

First, let us consider the (D+1)×(D+1) matrices Bi, 0 £ i £ D, whose entries are the intersection numbers of G; that is, (Bi)jk=(pijk). It is known that the character algebra, generated by the matrices B0,B1,¼,BD, is isomorphic to the adjacency algebra A(G) (this is a generalization of a result from [4]). Since such an isomorphism assigns the matrix B:=B1 to the matrix A =A1, it turns out that Bk=pk(B), 0 £ k £ D, and, consequently, it is enough to compute B1, called the intersection or recurrence matrix. In particular, A and B have the same minimum polynomial and, consequently, the same distinct eigenvalues. As we have already seen, the entries of B =(p1jk) give the following recurrence satisfied by the distance polynomials
xpj= D
å
k=0 
p1jk pk       (0 £ j £ D)
(31)
where p0=1 and p1=x. (Remember that polynomial equalities must be understood in the quotient ring \mathbbC[x]/Á.) Since the entries of B correspond to the cardinalities p1jk=|G1+(u)ÇG-j(v)|, provided that dist (u,v)=k, and G is D-regular, we infer that åj=0D p1jk=D, for any k=0,1,...,D; that is, j is a left eigenvector of B with eigenvalue D. In particular, the last row of B is uniquely determined for the other D. In matrix form, the D+1 equations in (31) correspond to
Bp(x)=xp(x)
(32)
where p(x):=(p0(x),p1(x),¼,pD(x))T. Note that, for a given value of x and starting from p0(x)=1 and p1(x)=x, the j-th equation, 2 £ j £ D-1, allows us to compute the value of pj+1(x) from those of pk(x), k=0,1,¼,j, through the formula
j-1
å
k=0 
p1jk pk(x) + (p1jj-x)pj(x)+p1jj+1pj+1(x)=0 ;
(33)
whereas the last equation (j=D), which we write in the form
pD+1(x):=(x-p1DD)pD(x)- D-1
å
k=0 
p1Dk pk(x)=0 ,
(34)
yields a condition for p(l) being a (right) eigenvector of B with eigenvalue l, namely that pD+1(l)=0. (Notice again that pD+1=0 must be understood in the quotient ring \mathbbC[x]/Á.) From this, we infer that the characteristic polynomial of B, which coincides with the minimum polynomial of B as both have the degree D+1 (and the later divides the former), is
y B (x):= det
(xI-B) =  1

aDD
pD+1(x).
(35)
Let li be an eigenvalue of B with eigenvector vi:=p(li), and assume that its multiplicity, as a zero of y B, is mi (1 £ mi £ m(li)). Then we can repeatedly take the derivative of (32) and evaluate at x=li to obtain
Bp(li)
=
lip(li),
Bp¢(li)
=
p(li) + lip¢(li),
Bp" (li)
=
2p¢(li) + lip"(li),
:
Bp(mi-1)(li)
=
(mi-1)p(mi-2)(li) +lip(mi-1)(li).
Then, dividing the k-th equation by (k-1)!, 1 £ k £ mi, and writing all the resulting equalities in matrix form we get
Bi = ¶i Ji       (0 £ i £ d),
(36)
where
i:= æ
ç
è
p(li)
 p¢(li)

1!
 p"(li)

2!
¼
 p(mi-1)(li)

(mi-1)!
ö
÷
ø
is the matrix formed by the mi columns of ¶ with indices ri, ri+1,¼, ri+mi-1 (see the previous subsection) and Ji is the Jordan block of size mi for the eigenvalue li:
Ji:= æ
ç
ç
ç
è
li
1
li
···
···
1
li
ö
÷
÷
÷
ø
.
Consequently, we have proved the following result:
Proposition 2 Let G be a weakly distance-regular digraph of degree D, diameter D, and distance-polynomials {pk}k=0D. Let A and B =(p1jk) be, respectively, its adjacency and intersection matrices. Then, the following statements hold:
Notice that the first component of each of the eigenvectors vi is 1 as p0=1. In this case we say that the eigenvector is standard . Furthermore, if all the eigenvalues of B are simple then {vi}i=0D is a basis for \mathbbCD+1 (equivalently, B diagonalizes on \mathbbC). The following result shows that, in this case, B has also a standard left eigenvector for every li, denoted by ui (as a column vector), and this allows us to compute the multiplicity of li as an eigenvalue of G.
Proposition 3 Let G be a weakly distance-regular digraph with order N, diameter D, and intersection matrix B. If each eigenvalue li of B is simple, then li has standard left and right eigenvectors, uiT and vi, and the multiplicity of li as an eigenvalue of G is
m(li)=  N

uiT vi
       (0 £ i £ d).
(37)
Proof.   Let p0=1,p1,¼,pD be the distance polynomials of G. Since all the zeros of mG are simple, then D=d and the (standard) right eigenvectors vj, 0 £ j £ d, are the columns of the matrix ¶ ev =(pi(lj)). Furthermore, since {vj}j=0d is a basis of \mathbbCd+1, ¶ ev is invertible, and the rows of ¶ ev -1 are left eigenvectors for B. So, from any set {ui}i=0d of such eigenvectors, we can write ¶ ev -1=D-1U , where D:= diag (u0Tv0,u1Tv1¼,udTvd) and U is the matrix with rows u0T,u1T,¼,udT. Thus, by Proposition 3.1,
m(li)=N (D-1U)i0 = N  ui0

uiTvi
       (0 £ i £ d).
This proves both statements since, as m(li) ¹ 0, it must be that ui0 ¹ 0 and hence we can chose every ui, 0 £ i £ d, in such a way that ui0=1. (As a by-product, notice also that uiTvi must be a real number.)   [¯]
Here it is worth mentioning the similarity between (37) and the known formula to compute the multiplicities of a distance-regular graph G. Thus, since the corresponding symmetric digraph G =G* (obtained by replacing every edge of G by a digon) is distance-regular, the above can be seen as an extension of such a formula to the (more general) directed case. In fact, if G is a distance-regular digraph we have, from the orthogonality of the distance polynomials and the simple form of the scalar product (18), that ui=([ 1/(n0)][`(p0(li))],[ 1/(n1)][`(p1(li))],¼, [ 1/(nd)][`(pd(li))]), where nj=nj(u)=pj(D), 0 £ j £ d; and the way of computing (37) from such polynomials is the same as for distance-regular graphs (see e.g. Bannai and Ito [2] or Biggs [4]).

4  Some constructions

In this last section we give several constructions and examples of weakly distance-regular digraphs. (For the case of diameter two, we refer the reader to the known constructions of directed strongly regular graphs [5].) For instance, it is shown that every line digraph of a (standard) distance-regular digraph is weakly distance-regular. Moreover two families of digraphs, well-known in the context of network theory, are shown to be weakly-distance regular and so they are analyzed in the light of the present theory.
Of course, (trivial) examples of weakly distance-regular digraphs are the known distance-regular digraphs but, for degree D > 1, they can only exist for diameter D and girth g satisfying D £ g £ 8, if g > 2; see [10]. Other easy examples are obtained from the distance-regular (undirected) graphs, which can be seen as distance-regular digraphs with girth two. Indeed, given a distance-regular graph G, its symmetric digraph G = G* is clearly a weakly distance-regular digraph.
Let us now see that, through the line digraph technique [16], every distance-regular digraph gives rise to a weakly distance-regular digraph. For a given digraph G, its line digraph LG has vertices representing the arcs of G, and vertex uv is adjacent to vertex wz when the corresponding arcs are adjacent in G; that is, when v=w.
Proposition 1 Let G be a distance-regular digraph different from a cycle (D > 1), with girth g and diameter D. Then its line digraph LG is a weakly distance-regular digraph. Moreover, if 1,x,p2,...,pD are the distance polynomials of G, then the distance polynomials of LG are
1,x,xp1,...,xpg-2,xpg-1-1,xpg,¼, xpD.
Proof.   Let A and [(A)\tilde] be the adjacency matrices of G and LG, respectively. Then, the number of walks of length l+1 from vertices uv ¹ wz of LG, such that dist (uv,wz)=k+1, 0 £ k £ D, is
~
a
 
(l+1)
uv,wz 
=av,w(l)=akl
since dist (v,w)=k; whereas in the case uv=wz (that is, dist (uv,wz)=0) we have
~
a
 
(l+1)
uv,uv 
=av,u(l)=ag-1l
as dist (u,v)=1 implies dist (v,u)=g-1. Thus the number of such walks only depend on the distance between vertices, and LG is weakly distance-regular.
Moreover, if pk is the k-th distance polynomial of G, 0 £ k £ D, then
(
~
A
 
pk(
~
A
 
))uv,wz=(pk(A))vw=(Ak)vw.
Therefore, taking into account that ([(A)\tilde]k+1)uv,wz=(Ak)vw, if uv ¹ wz, and
(
~
A
 

k+1 
)uv,uv=0= ì
í
î
(Ak)vu
if
k ¹ g-1
(Ak)vu -1
if
k = g-1,
we conclude that the (k+1)-th distance polynomial of LG is xpk, if k ¹ g-1, and xpg-1-1, if k=g-1.   [¯]
In fact, note that the above result still holds under some slightly less restricted condition on G. For instance, it suffices to require G to be weakly distance-regular, and with every arc (u,v) contained in a cycle of smallest length g (for then dist (v,u)=g-1). For example, this is the case when G =G*, G being a distance-regular graph, as mentioned above. More generally, we can consider a weakly distance-regular digraph G with every arc a contained in the same number of closed walks of length l ³ 0. (This is equivalent to require that LG must be a `walk regular digraph', a concept defined by Godsil [17] in the undirected case). As we will see, an example of digraphs satisfying these properties are the so-called butterfly networks, which are the topic of the next subsection.

The butterfly networks

Butterfly networks have been extensively studied in the literature because of their multiple applications in computer architecture (see e.g. [3,13,18]). In particular, the wrapped butterfly digraph G =BD(n), of degree D and dimension n, has vertices labeled by the pairs (l,x), where l Î \mathbbZn and x Î \mathbb ZDn; and vertex (l, x0x1...xn-1) is adjacent to the vertices ( l+1, x0 ...xl-1yl xl+1... xn-1) for any yl Î \mathbb ZD. Thus G is a regular digraph with degree D, order N=nDn, girth g=n, and diameter D=2n-1.
Proposition 2 The wrapped butterfly digraph G =BD(n) is a weakly distance-regular digraph with distance polynomials
p0 = 1,    p1=x,    ¼,   pn-1=xn-1,    pn = xn-1, pn+1=  1

D
xn+1-x,   ¼,    p2n-1=  1

Dn-1
x2n-1-xn-1.
Furthermore, the eigenvalues of BD(n) are
li=Dwi    (0 £ i £ n-1)      and       ln=0,
where w is any primitive n-th root of unity, say, w:=ej[(2p)/n], with respective multiplicities
m(li)=1    (0 £ i £ n-1)      and       m(ln)=N-n.
Proof.   The invariance of the number of walks, implying the weak distance-regularity, and the expressions for the distance polynomials follow immediately from the following well-known properties of BD(n): (i) The lengths of all walks from vertex (l,x) to vertex (l¢,y) are congruent to l¢-l modulo n. (ii) For 0 £ k £ n, the number of vertices at distance k from a given vertex (l,x) is the maximum possible for a D-regular graph with girth g=n, that is, Dk if 0 £ k £ n-1 and Dn-1 when k=n. This means that the (shortest) path from (l,x) to each of these vertices is unique. (iii) For n+1 £ k £ 2n-1, the number of walks of length k from (l,x) to every vertex of the form (l+k,y) (including those previously reached) is Dk-n. (To see this, use induction on k.)
Now, from the distance polynomials and Proposition 3.2(i), we know that, apart from l0=D, the distinct eigenvalues of G are the zeros of the polynomial
S:= D
å
k=0 
pk=xn æ
è
1+  x

D
+(  x

D
)2+¼+(  x

D
)n-1 ö
ø
;
that is, li=Dwi, 1 £ i £ n-1, and ln=0. Thus,
ev = æ
ç
ç
ç
ç
ç
ç
è
1
1
¼
1
1
D
Dw
¼
Dwn-1
0
D2
D2w2
¼
D2w(n-1)2
0
:
···
:
Dn-1
Dnwn-1
¼
Dnw(n-1)n-1
-1
ö
÷
÷
÷
÷
÷
÷
ø
has an inverse matrix with first column [ 1/N](1,1,¼, 1,N-n)T since n + (N-n) = N, Dkåi=0n-1wik = 0 for any 1 £ k £ n-1 and, using that N=nDn, åi=0n-1(Dnwin-1)+(N-n)(-1) = n(Dn-1)-N+n = 0. (Of course, this corresponds to verifying the equations (20).) Consequently, the multiplicities are as claimed.   [¯]
If we allow the presence of multiple arcs, we can reach to the same result in the following way. Let CnD be the directed cycle on n vertices and D parallel arcs between any adjacent vertices. So, CnD is a weakly distance-regular digraph with girth g=n, diameter D=n-1, and distance polynomials pk=[ 1/(Dk)]xk, 0 £ k £ n-1. Moreover, for any given 1 £ l £ n, every walk of length l is contained in a (smallest) cycle of length n. Thus, Proposition 4.1 can be repeatedly applied to prove that the iterated line digraph LnCnD is a weakly distance-regular digraph with the same parameters as BD(n). Then, as both digraphs can be shown to be isomorphic, we get again Proposition 4.2. (To see more details about the relationship between the spectra of a digraph and its line digraph, see [24,22].)

The cycle prefix digraphs

The cycle prefix digraphs were introduced by Faber and Moore as a Cayley coset digraphs, although to study some of their properties it is more useful to consider them as digraphs on alphabets (see [14,7]). The cycle prefix digraph G =GD(D) has vertices of the form x1x2¼xD where the xi¢s are distinct elements from the set {1,2,¼,D+1}, and vertex x1x2¼xD is adjacent to the vertices x2¼xDxD+1 and x1x2¼xk-1xk+1¼xDxk, 2 £ k £ D-1. Thus, G is a D-regular digraph, with order N=(D+1)D:=[((D+1)!)/((D+1-D)!)] and diameter D £ D. Moreover, in [23,8] the first and last authors showed that the eigenvalues of G are D,D-2,D-1,¼,0,-1. In fact they obtained this result by proving the existence of the distance polynomials, so that the cycle prefix digraph is an example of weakly distance-regular digraph. To show that this is a `genuine' example, and GD(D) is not distance-regular, notice that its girth is g=2 since the digraph has digons. Let us consider the vertices u=x1x2 ... xD and v=x2... xDx1. Clearly dist (u,v)=1 but dist (v,u) > 1 if D > 1. Hence, the digraph is not stable. (The above can also be proved directly from the definition of strongly distance-regular digraph given in [9].)
Given an integer x ³ 2, let us denote by Px the set of (ordered) t-tuples (x1,x2,¼,xt), which are a partition of x, x1+x2+¼+xt=x, with xi ³ 2 for any 1 £ i £ t. (Thus, 1 £ t £ ëx/2û.) Now, let Y be the function defined on the nonnegative integers as
Y(0):=1,   Y(1):=0,   and    Y(x):=
å
(x1,x2,¼,xt) Î Px 
t
Õ
i=1 
 (xi-2)!

xi
       (x ³ 2).
For example, Y(5)=[ 3!/5]+[ 0!/2][ 1!/3]+[ 1!/3][ 0!/2] = [ 23/15]. By looking at the possible values of the first term x1, we get the following useful recurrence formula for computing Y:
Y(x+1)= x-1
å
y=0 
 (x-y-1)!

x-y+1
Y(y).
(38)
In the next result we also use the following notation. For any given integer k ³ 1, let (x)k denote the polynomial in x obtained as the product x(x-1)(x-2)¼(x-k+1) (note that this definition is consequent with the meaning of (D+1)D).
Theorem 3 The cycle prefix digraph G =GD(D) is a weakly distance-regular digraph with distance polynomials
p0=1, p1=x, p2=  (x+1)3

x
, p3 =  (x+1)4

x-1
¼, pD=  (x+1)D+1

x-D+2
.
Moreover, the eigenvalues of G are
l0=D    and    li=D-i-1        (1 £ i £ D),
with respective multiplicities
m(l0) = 1, m(l1)=  1

(D-1)!
 (D+1)D+1

D-D+2
,  m(l2)=  1

(D-2)!
 (D+1)D

D-D+3
,
(39)
m(li) =  1

(D-i)!
i-1
å
x=0,x ¹ i-2 
Y(i-1-x)  (D+1)D+1-x

D-D+2+x
      (3 £ i £ D-1),
(40)
m(lD) = (D+1)D- D-1
å
i=0 
m(li).
(41)
Proof.   As already mentioned, the distance polynomials and the eigenvalues were given in [8]. Indeed, from such polynomials, the eigenvalues are l0=D and the zeros of
S= D
å
k=0 
pk=1+ D+1
å
k=2 
 (x+1)k

(x-k+3)
=(x+1)D
(to prove the last equality, use induction on D), which gives the above values for li, 1 £ i £ D. Then, notice that pD(li)=0 for any i=2,3,¼,D; and, when 1 £ k £ D-1, pD-k(li)=0 for any i=k,k+2,k+3,¼,D. Consequently, the matrix ¶ ev has the following quasi-triangular form
ev = æ
ç
ç
ç
ç
ç
ç
ç
è
1
1
¼
1
1
1
D
l1
¼
lD-2
0
-1
p2(D)
p2(l1)
¼
0
p2(lD-1)
p3(D)
p3(l1)
¼
p3(lD-2)
:
:
pD-1(D)
0
pD-1(l2)
pD(D)
pD(l1)
ö
÷
÷
÷
÷
÷
÷
÷
ø
which, from (20), allows us to obtain the claimed multiplicities recursively from m(D)=1,
m(l1)=-  pD(D)

pD(l1)
,      m(l2)=-  pD-1(D)

pD-1(l2)
,      
and
m(li)=  -1

pD-i+1(li)
i-1
å
j=0 
pD-i+1(lj)m(lj)      (3 £ i £ D-1)
by using (38) (we skip the cumbersome details). Finally, m(lD) is obtained from the first equation expressing that all multiplicities add up to N=(D+1)D.   [¯]

A possible generalization

We end this section with a digression about a natural generalization of weak distance-regularity and its relation with some of our results. In the last section of Damerell's paper [9], entitled Related problems , the author introduce the notion of the so-called weakly distance-transitive digraph. According to him, a digraph G is weakly distance-transitive if G is connected and for any pairs of vertices, (u,v) and (w,z), such that
dist (u,v)= dist (w,z)        and        dist (v,u)= dist (z,w),
there exists an automorphism p Î Aut G such that p(u)=w and p(v)=z. (Damerell argues that this would be a definition closer to the corresponding concept for graphs). Inspired by this, we could consider an alternative (non-equivalent) definition of a weakly distance-regular digraph G as the digraph satisfying the following condition: Or, in terms of the adjacency matrix A of G, Observe that condition (W*) is stronger than the property characterizing walk regular digraphs, but `weaker' that our condition characterizing weakly distance-regularity. Namely, In this context we have the following result (we omit the proof, as some of the equivalences have been already given):
Theorem 4 Let G =(V,E) be a digraph with adjacency matrix A and girth g. Then the following statements are equivalent
One of the referees pointed out that, recently, H. Suzuki and K. Wang [27] suggested that a `weakly distance-regular digraph' is a digraph with the following property: for vertices u and v such that dist (u,v)=k1 and dist (v,u)=k2, the number of vertices z satisfying dist (u,z)=i1, dist (z,u)=i2, dist (v,z)=j1, and dist (z,v)=j2 depend only on the values k1,k2,i1,i2,j1,j2. (This is equivalent to require that the Hadamard products Ai°AjT form an association scheme.) Weakly distance-transitive digraphs satisfy this property, but it is not clear wheter these digraphs are easily related with property (W*).

Acknowledgment. The authors are indebted to the referees for helpful comments and suggestions which lead to numerous improvements of the manuscript.

References

[1]
E. Bannai, P. J. Cameron and J. Kahn, Nonexistence of certain distance-transitive digraphs, J. Combin. Theory Ser. B 31 (1981), 105-110.
[2]
E. Bannai and T. Ito, Algebraic Combinatorics I: Association Schemes. Benjamin/Cummings, Menlo Park, CA, 1984.
[3]
J.-C. Bermond, E. Darrot, O. Delmas, and S. Perennes, Hamilton circuits in the directed wrapped butterfly network, Discrete Appl. Math. 84 (1998), no. 1-3, 21-42.
[4]
N. Biggs, Algebraic Graph Theory, Cambridge University Press, Cambridge, 1974; second edition: 1993.
[5]
A. Brouwer, personal homepage: http://www.cwi.nl/~aeb/math/dsrg/dsrg.html
[6]
R. A. Brualdi and H. J. Ryser, Combinatorial Matrix Theory, Cambridge University Press, Cambridge, 1991.
[7]
F. Comellas and M.A. Fiol, Vertex-symmetric digraphs with small diameter, Discrete Appl. Math. 58 (1995), no. 1, 1-12.
[8]
F. Comellas and M. Mitjana, The spectra of cycle prefix digraphs, SIAM J. Discrete Math. 16 (2003), no. 3, 418-421.
[9]
R.M. Damerell, Distance-transitive and distance-regular digraphs, J. Combin. Theory Ser. B 31 (1981), no. 1, 46-53.
[10]
D. A. Douglas and K. Nomura, The girth of a directed distance-regular graph, J. Combin. Theory Ser. B 58 (1993), no. 1, 34-39.
[11]
Art M. Duval, A directed graph version of strongly regular graphs, J. Combin. Theory Ser. A 47 (1988), no. 1, 71-100.
[12]
H. Enomoto and R.A. Mena, Distance-regular digraphs of girth 4, J. Combin. Theory Ser. B 43 (1987), no. 3, 293-302.
[13]
M. Espona and O. Serra, Cayley digraphs based on the de Bruijn networks, SIAM J. Discrete Math. 11 (1998), no. 2, 305-317.
[14]
V. Faber, J. W. Moore, and W. Y. C. Chen, Cycle prefix digraphs for symmetric interconnection networks, Networks 23 (1993), 641-649.
[15]
M. A. Fiol, Algebraic characterizations of distance-regular graphs, Discrete Math. 246 (2002), no. 1-3, 111-129.
[16]
M. A. Fiol, J. L. A. Yebra, and I. Alegre, Line digraphs iterations and the (d,k) digraph problem, IEEE Trans. Comput. C-32 (1984), 400-403.
[17]
C. D. Godsil, Algebraic Combinatorics, Chapman and Hall, New York, 1993.
[18]
R. Klasing, R. Lüling, and B. Monien, Compressing cube-connected cycles and butterfly networks, Networks 32 (1998), no. 1, 47-65.
[19]
C. W. H. Lam, Distance-transitive digraphs, Discrete Math. 29 (1980), no. 3, 265-274.
[20]
P. Lancaster and M. Tismenetsky, The Theory of Matrices, Academic Press, Orlando, FL; second edition: 1985.
[21]
R. A. Liebler and R. A. Mena, Certain distance-regular digraphs and related rings of characteristic 4, J. Combin. Theory Ser. A 47 (1988), no. 1, 111-123.
[22]
X. Marcote, C. Balbuena, and I. Pelayo, The relationship between the Jordan normal forms of a digraph and its line digraph, submitted.
[23]
M. Mitjana, Propagació d'informació en grafs i digrafs que modelen xarxes d'interconnexió simètriques (Catalan), Ph.D. Thesis, Universitat Politècnica de Catalunya, 1999.
[24]
J. C. Montserrat, Propiedades matriciales de los grafos dirigidos (Spanish), Master Thesis, Universitat Politècnica de Catalunya, 1986.
[25]
P. Rowlinson, Linear algebra, in: Graph Connections (ed. L.W. Beineke and R.J. Wilson), Oxford Lecture Ser. Math. Appl., Vol. 5, 86-99, Oxford Univ. Press, New York, 1997.
[26]
T. Takahashi, Distance-regular digraphs of girth 6, Memoirs of the Faculty of Science, Kyushu University, Ser. A 45 (1991), no. 2, 155-166.
[27]
K. Wang and H. Suzuki, Weakly distance-regular digraphs, Discrete Math. 264 (2003), no. 1-3, 225-263.

Footnotes:

1Departament de Matemàtica Aplicada IV, Universitat Politècnica de Catalunya, Jordi Girona 1-3, Mòdul C3, Campus Nord, 08034 Barcelona, Spain.
2Departament de Matemàtica, Universitat de Lleida, Jaume II 69, 25005 Lleida, Spain.
3Departament de Matemàtica Aplicada I, Universitat Politècnica de Catalunya, Gregorio Marañon 44, 08028 Barcelona, Spain.
4If a weakly distance-regular digraph G contains a loop, then each vertex has the same number of loops. Moreover, if G contains a multiple edge, then all edges of G are multiple (with equal "multiplicity"). Hence, the deletion of all such repetitions and loops provides another weakly distance-regular digraph.


File translated from TEX by TTH, version 3.40.
On 26 Jun 2003, 19:19.