Algebras, Rings and Modules Vol.2
Algebras, Rings and Modules Vol.2
Algebras, Rings and Modules Vol.2
Managing Editor:
M. HAZEWINKEL
Centre for Mathematics and Computer Science, Amsterdam, The Netherlands
Volume 586
Algebras, Rings and Modules
Volume 2
by
Michiel Hazewinkel
CWI,
Amsterdam, The Netherlands
Nadiya Gubareni
Technical University of Czestochowa,
Poland
and
V.V. Kirichenko
Kiev Taras Shevchenko University,
Kiev, Ukraine
A C.I.P. Catalogue record for this book is available from the Library of Congress.
Published by Springer,
P.O. Box 17, 3300 AA Dordrecht, The Netherlands.
www.springer.com
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix
v
vi TABLE OF CONTENTS
ix
x PREFACE
been obtained by A.V.Roiter for the case of finite dimensional algebra over an
arbitrary field.
Chapter 4 is devoted to study of Frobenius algebras and quasi-Frobenius rings.
The class of quasi-Frobenius rings was introduced by T.Nakayama in 1939 as a
generalization of Frobenius algebras. It is one of the most interesting and in-
tensively studied classes of Artinian rings. Frobenius algebras are determined
by the requirement that right and left regular modules are equivalent. And
quasi-Frobenius algebras are defined as algebras for which regular modules are
injective.
We start this chapter with a short study of duality properties for finite dimen-
sional algebras. In section 4.2 there are given equivalent definitions of Frobenius
algebras in terms of bilinear forms and linear functions. There is also a discussion
of symmetric algebras which are a special class of Frobenius algebras. The main
properties of quasi-Frobenius algebras are given in section 4.4.
The starting point in studying quasi-Frobenius rings in this chapter is the
Nakayama definition of them. The key concept in this definition is a permuta-
tion of indecomposable projective modules, which is naturally called Nakayama
permutation.
Quasi-Frobenius rings are also of interest because of the presence of a dual-
ity between the categories of left and right finitely generated modules over them.
The main properties of duality in Noetherian rings are considered in section 4.10.
Semiperfect rings with duality for simple modules are studied in section 4.11.
The equivalent definitions of quasi-Frobenius rings in terms of duality and semi-
injective rings are given 4.12. Quasi-Frobenius rings have many interesting equiv-
alent definitions, in particular, an Artinian ring A is quasi-Frobenius if and only
if A is a ring with duality for simple modules.
One of the most significant results in quasi-Frobenius ring theory is the theorem
of C.Faith and E.A.Walker. This theorem says that a ring A is quasi-Frobenius if
and only if every projective right A-module is injective and conversely.
Quivers of quasi-Frobenius rings are studied in section 4.13. The most impor-
tant result of this section is the Green theorem: the quiver of any quasi-Frobenius
ring is strongly connected. Conversely, for a given strongly connected quiver Q
there is a symmetric algebra A such that Q(A) = Q. Symmetric algebras with
given quivers are studied in section 4.14.
Chapter 5 is devoted to the study of the properties and structure of right serial
rings. Note that a module is called serial if it decomposes into a direct sum of
uniserial submodules, i.e., submodules with linear lattice of submodules. A ring is
called right serial if its right regular module is serial.
We start this chapter with a study of right Noetherian rings from the point of
view of some main properties of their homological dimensions.
In further sections we give the structure of right Artinian right serial rings in
terms of their quivers. We also describe the structure of particular classes of right
serial rings, suchas quasi-Frobenius rings, right hereditary rings, and semiprime
xii PREFACE
rings. In section 5.6 we introduce right serial quivers and trees and give their
description.
The last section of this chapter is devoted to the Cartan determinant conjecture
for right Artinian right serial rings. The main result of this section says that a
right Artinian right serial ring A has its Cartan determinant equal to 1 if and only
if the global dimension of A is finite.
In chapters 6 and 7 the theory of semiprime Noetherian semiperfect semidis-
tributive rings is developed (SP SD-rings). In view of the decomposition theorem
(see theorem 14.5.1, vol.I) it is sufficient to consider prime Noetherian SP SD-
rings, which are called tiled orders.
With any tiled order we can associate a reduced exponent matrix and its quiver.
This quiver Q is called the quiver of that tiled order. It is proved that Q is a simply
laced and strongly connected quiver. In chapter 6 a construction is given which
allows to form a countable set of Frobenius semidistributive rings from a tiled
order. Relations between finite posets and exponent (0,1)-matrices are described
and discussed. In particular, a finite ergodic Markov chain is associated with a
finite poset.
Chapter 7 is devoted to the study of Gorenstein matrices. We say that a
tiled order A is Gorenstein if r.inj.dimA A = 1. In this case r.inj.dimA A =
l.inj.dimA A = 1. Moreover, a tiled order is Gorenstein if and only if it is Morita
equivalent to a reduced tiled order with a Gorenstein exponent matrix.
Each chapter ends with a number of notes and references, some of which have
a bibliographical character and others are of a historical nature.
At the end of the book we give a literature list which can be considered as sug-
gestions for further reading to obtain fuller information concerning other aspects
of the theory of rings and algebras.
In closing, we would like to express our cordial thanks to a number of friends
and colleagues for reading preliminary versions of this text and offering valuable
suggestions which were taken into account in preparing the final version. We
are especially greatly indebted to Yu.A.Drozd, V.M.Bondarenko, S.A.Ovsienko,
M.Dokuchaev, V.Futorny, V.N.Zhuravlev, who made a large number of valu-
able comments, suggestions and corrections which have considerably improved the
book. Of course, any remaining errors are the sole responsibility of the authors.
Finally, we are most grateful to Marina Khibina for help in preparing the
manuscript. Her assistance has been extremely valuable to us.
1. Groups and group representations
Groups are a central subject in algebra. They embody the easiest concept of
symmetry. There are others: Lie algebras (for infinitesimal symmetry) and Hopf
algebras (quantum groups) who combine the two and more (see volume III). Fi-
nite groups, in particular permutation groups, are an increasingly important tool
in many areas of applied mathematics. Examples include coding theory, cryptog-
raphy, design theory, combinatorial optimization, quantum computing.
Representation theory, the art of realizing a group in a concrete way, usually
as a collection of matrices, is a fundamental tool for studying groups by means
of linear algebra. Its origins are mostly in the work of F.G.Frobenius, H.Weil,
I.Schur, A.Young, T.Molien about century ago. The results of the theory of repre-
sentations of finite groups play a fundamental role in many recent developments of
mathematics and theoretical physics. The physical aspects of this theory consist
in accounting for and using the concept of symmetry as present in various physical
processes – though not always obviously so. As understood at present, symme-
try rules physics and an elementary particle is the same thing as an irreducible
representation. This includes quantum physics. There is a seeming mystery here
which is explained by the fact that the representation theory of quantum groups
is virtually the same as that of their classical (Lie group) counterparts.
In this chapter we shall give a short introduction to the theory of groups
and their representations. We shall consider the representation theory of groups
from the module-theoretic point of view using the main results about rings and
modules as described in volume I of this book. This theoretical approach was
first used by E.Noether who established a close connection between the theory of
algebras and the theory of representations. From this point of view the study of
the representation theory of groups becomes a special case of the study of modules
over rings. At the end of this chapter we shall consider the characters of groups.
We shall give the basic properties of irreducible characters and their connection
with the ring structure of group algebras.
For the convenience of a reader in the beginning of this chapter we recall some
basic concepts and results of group theory which will be necessary for the next
chapters of the book.
1
2 ALGEBRAS, RINGS AND MODULES
From the axioms for a group G one can easily obtain the following properties:
(1) the identity element in G is unique;
(2) for each a ∈ G the element a−1 is uniquely determined;
(3) (a−1 )−1 = a for every a ∈ G;
(4) (a ∗ b)−1 = b−1 ∗ a−1 .
Examples 1.1.1.
1. The sets Z, Q, R and C are groups under the operation of addition + with
e = 0 and a−1 = −a for all a. They are additive Abelian groups.
2. The sets Q \ {0}, R \ {0} and C \ {0} are groups under the operation
of multiplication · with e = 1 and a−1 = 1/a for all a. They are multiplicative
Abelian groups. The set Z\{0} with the operation of multiplication · is not a group
because the inverse to n is 1/n, which is not integer if n = 1. The set R+ of all
positive rational numbers is a multiplicative Abelian group under multiplication.
3. The set of all invertible n × n matrices with entries from a field k forms a
group under matrix multiplication. This group is denoted by GLn (k) and called
GROUPS AND GROUP RINGS 3
the general linear group of order n (in dimension n). This group is finite if
and only if k is a finite field.
4. The set of all invertible linear transformations of a vector space V over a
field k forms a group under the operation of composition. This group is denoted
by GL(V, k). If V is an n-dimensional vector space over a field k, i.e., V k n , then
there is a one-to-one correspondence between invertible matrices of order n and
invertible linear transformations of the vector space k n . Thus the group GL(k n , k)
is isomorphic to the group GLn (k).
5. Suppose G is the set of all functions f : [0, 1] → R. Define an addition on
G by (f + g)(t) = f (t) + g(t) for all t ∈ R. Then G is an Abelian group under
(pointwise) addition.
Examples 1.1.2.
1. Z is a proper subgroup of Q and Q is a proper subgroup of R with the
operation of addition.
2. The set of all even integers is a subgroup of Z under addition.
3. If G = Z under addition, and n ∈ Z, then H = nZ is a subgroup of Z.
Moreover, every subgroup of Z is of this form.
4. Let k be a field. Define
which is called the special linear group or the unimodular group. This group
is a proper subgroup of GLn (k).
For finite groups of not to large order it can be convenient to represent the
operation on a group by means of a multiplication table, which is often called
4 ALGEBRAS, RINGS AND MODULES
its Cayley table. Such a table is a square array with the rows and columns
labelled by the elements of the group. In this table at the intersection of the i-th
row and the j-th column we write the product of the elements, which are in the i-th
row and the j-th column respectively. It is obvious, that this table is symmetric
with respect to the main diagonal if and only if the group is Abelian. For example,
consider for a group G = {e, a, b, c} the group table:
e a b c
e e a b c
a a e c b
b b c e a
c c b a e
This group is called the Klein 4-group.
In the general case for a group G one can write down a set of generators S
with the property that every element of G can be written as a finite product of
elements of S. Any equation in a group G that the generators satisfy is called a
relation in G. For example, in the previous example the Klein group G has the
relations
a2 = b2 = c2 = e, ab = c, ac = b, bc = a.
Example 1.1.3.
Symmetric groups. Let A be a nonempty set and let SA be the set of all
bijections from A to itself. If x, y ∈ SA , then their multiplication z = xy is defined
by z(a) = x(y(a)) for an arbitrary a ∈ A. It is easy to see that z ∈ SA , and that
the operation of multiplication of transformations is associative. The identity of
this operation is the identity transformation e of the set A, which is defined by
e(a) = a for all a ∈ A.
Obviously, ex = xe = x for all x ∈ SA . The inverse element to x is defined as
the transformation x−1 for which x−1 (x(a)) = a for all a ∈ A. Clearly, x−1 x =
xx−1 = e. Therefore SA is a group which is called the symmetric group on the
set A.
In the special case, when A = {1, 2, ..., n}, each transformation of A is called a
permutation and the symmetric group on A is called the permutation group
of A. It is also denoted by Sn and called the symmetric group of degree n.
The order of the group Sn is n! The group Sn is non-Abelian for all n ≥ 3.
GROUPS AND GROUP RINGS 5
Example 1.1.4.
Alternating group. Let Sn be a symmetric group, i.e., the group of all
permutations of {1, 2, ..., n}. Let x1 , x2 , ..., xn be independent variables. Consider
the polynomial
Δ= (xi − xk ), (i, k = 1, 2, ..., n).
i<k
For each σ ∈ Sn let σ act on Δ by permuting the variables in the same way; i.e.,
it permutes their indices:
σ(Δ) = (xσ(i) − xσ(k) ), (i, k = 1, 2, ..., n).
i<k
Then
σ(Δ) = ±Δ, for all σ ∈ Sn .
For each σ ∈ Sn let sign(σ) = ε(σ) be defined by
+1 for σ(Δ) = Δ
ε(σ) =
−1 for σ(Δ) = −Δ.
Example 1.1.5.
Let f be any polynomial in n independent variables x1 , x2 , ..., xn . Then
Example 1.1.6.
Groups in geometry. F.Klein was the first who wrote down the connection
between permutation groups and symmetries of convex polygons. He also posed
the idea that the background of all different geometries is the notion of a group of
6 ALGEBRAS, RINGS AND MODULES
Example 1.1.7.
Galois groups. In many examples groups appear in the form of automorphism
groups of various mathematical structures. This is one of the most important ways
of their appearance in algebra. In such a way we can consider Galois groups. Let K
be a finite, separable and normal extension of a field k. The automorphisms of K
leaving the elements of k fixed form a group Gal(K/k) with respect to composition,
called the Galois group of the extension K/k. Let f be a polynomial in x over k
and K be the splitting field of f . The group Gal(K/k) is called the Galois group
of f . One of the main applications of Galois theory is connected with the problem
of the solvability of equations in radicals. Indeed, the main theorem says that
the equation f (x) = 0 is solvable in radicals if and only if the group Gal(K/k) is
solvable (see section 1.5). This is where the terminology “solvable” (for groups)
comes from.
Example 1.1.8.
Homology groups. This kind of groups, considered in section 6.1 (vol.I),
occurs in many areas of mathematics and allow us to study non-algebraic objects
by means of algebraic methods. This is a fundamental method in algebraic
topology. To each topological space X there is associated a family of Abelian
groups H0 (X), H1 (X),..., called the homology groups, while each continuous
mapping f : X → Y defines a family of homomorphisms fn : Hn (X) → Hn (Y ),
n = 0, 1, 2, ....1
1 A homomorphism of groups f : G → H is a map that preserves the unit element and the
multiplication, i.e. f (eG ) = eH and f (xy) = f (x)f (y). It then also preserves inverses, i.e.,
f (x−1 ) = f (x)−1 . See section 1.3.
GROUPS AND GROUP RINGS 7
Example 1.2.1.
The symmetry group of an equilateral triangle is isomorphic to S3 . The struc-
ture of this group is completely determined by the relations σ 3 = τ 2 = 1 and
στ = τ σ −1 , where σ is the cyclic permutation (1, 2, 3) and τ is the reflection (1, 3).
Labeling the vertices of the triangle as 1, 2 and 3 permits us to identify the
symmetries with permutations of the vertices, and we see that there are three
rotation symmetries (through angles of 0, 2π/3 and 4π/3) corresponding to the
identity permutation, the cycles (1, 2, 3) and (1, 3, 2), and three reflection symme-
tries corresponding to the other three elements of S3 .
8 ALGEBRAS, RINGS AND MODULES
Example 1.2.2.
Dihedral groups. For each n ∈ Z+ , n ≥ 3, let Dn be a set of all symmetries
of an n-sided regular polygon. There are n rotation symmetries, through angles
2kπ/n, where k ∈ {0, 1, 2, ..., n − 1}, and there are n reflection symmetries, in the
n lines which are bisectors of the internal angles and/or perpendicular bisectors
of the sides. Therefore |Dn | = 2n. The binary operation on Dn is associative
since composition of functions is associative. The identity of Dn is the identity
symmetry, denoted by 1, and the inverse of s ∈ Dn is the symmetry which reverses
all rigid motions of s.
In Dn we have the relations: σ n = 1, τ 2 = 1 and στ = τ σ −1 , where σ is a
clockwise rotation through 2π n and τ is any reflection. Moreover, one can show
that any other relation between elements of the group Dn may be derived from
these three relations. Thus there is the following presentation of the group Dn :
Dn = {σ, τ : σ n = τ 2 = 1, στ = τ σ −1 }.
Dn is called the dihedral group of order n. Some authors denote this group by
D2n .
The rotation symmetries in the group Dn form a subgroup in it and this group
is called the rotation group of a given n-sided regular polygon. It is immediate
that this subgroup is isomorphic to Zn .
For n = 2 a degenerate “2-sided regular polygon” would be a line segment and
in this case we have the simplest dihedral group
D2 = {σ, τ : σ 2 = τ 2 = 1, στ = τ σ −1 },
Example 1.2.3.
Quasidihedral groups. Quasidihedral groups are groups with similar prop-
erties as dihedral groups. In particular, they often arise as symmetry groups of
regular polygons, such as an octagon. For each n ∈ Z+ , n ≥ 3, the group Q2n has
the following presentation:
n n−1
Qn = {σ, τ : σ 2 = τ 2 = 1, στ = τ σ 2 −1
}.
Example 1.2.4.
Generalized quaternion groups. A group is is called the generalized
quaternion group of order n if it has the following presentation
n n−1
Hn = {σ, τ : σ 2 = 1, σ 2 = τ 2 , στ = τ σ −1 }
GROUPS AND GROUP RINGS 9
H2 = {σ, τ : σ 4 = 1, σ 2 = τ 2 , στ = τ σ −1 }
which is the usual quaternion group H2 = {1, i, j, k, −1, −i, −j, −k} if one takes,
for instance, σ = i and τ = j (see example 1.1.12, vol.I).
Example 1.2.5.
Orthogonal groups. Let En be a Euclidean space, that is, a real n-
dimensional vector space Rn together with the scalar product (x, y) = x1 y1 +
x2 y2 + ... + xn yn in a given orthonormal basis e1 , e2 , ..., en . The linear transfor-
mations of En , which preserve the scalar product, are called orthogonal. They
form a group O(n) which is called the orthogonal group of En . The elements
of O(n) are orthogonal matrices, i.e.,
Example 1.2.6
Symmetry in physical laws. Group theory plays a similar role in physics.
The groups of transformations in physics describe symmetries of physical laws,
in particular, symmetry of space-time. Thus, the state of a physical system is
represented in quantum mechanics by a point in an infinite-dimensional vector
space. If the physical system passes from one state into another, its representative
point undergoes some linear transformation. The ideas of symmetry and the theory
of group representations are of prime importance here.
The laws of physics and invariants in mechanics must be preserved under trans-
formations from one inertial coordinate system to another. The corresponding
Galilean transformation of space-time coordinates in Newtonian mechanics has
the following form for (uniform) motion along the x-axis with velocity v:
x = x − vt, y = y; z = z; t = t,
and in the Einstein’s special theory of relativity the Lorentz transformation has
the form for motion along the x-axis with velocity v:
x − vt t − vx/c2
x = , y = y; z = z; t = ,
1 − (v/t)2 1 − (v/t)2
where c is the speed of light.
10 ALGEBRAS, RINGS AND MODULES
All Galilean transformations form a group which is called the Galilean group,
and all Lorentz transformations form the Lorentz group. The Lorentz transfor-
mations, named after its discoverer, the Dutch physicist and mathematician Hen-
rik Anton Lorentz (1853-1928), form the basis for the special theory of relativity,
which has been introduced to remove contradictions between the theory of elec-
tromagnetism and classical mechanics. The Lorentz group is the subgroup of the
Poincaré group consisting of all isometries that leave the origin fixed. This group
was been described in the work of H.A.Lorentz and H.Poincaré as the symmetry
group of the Maxwell equations:
For the proof of Lagrange’s theorem we can also introduce another relation
defined by a ∼ b if and only if b−1 a ∈ H. The resulting equivalence classes are
GROUPS AND GROUP RINGS 11
called left cosets. We can show that in this case each equivalence class has the
form aH for some a ∈ G and each set of the form bH is a left coset. The number
of left cosets is also equal to mn , i.e., the number of all right cosets in G. This
common number is called the index of H in G and denoted by |G : H|.
In the case of finite groups from the Lagrange theorem it follows that the index
|G|
of H in G is equal to , that is,
|H|
|G| = |G : H| · |H|.
If a group G is infinite and the number of left (or right) adjacent classes is infinite,
then we say that the index of H in G is infinite.
In general, the sets of right cosets and left cosets may be different. It is inter-
esting to know when these sets are the same. Suppose E is a right coset and a left
coset simultaneously. Then E = Ha = aH for all a ∈ E. If every right coset is a
left coset, then Ha = aH for all a ∈ G. Multiplying the last equality on a−1 we
obtain a−1 Ha = H for all a ∈ G. Subgroups with this property deserve special
attention.
Examples 1.3.1.
1. For any group G the group G itself and the unit subgroup are normal
subgroups.
2. If G is an Abelian group, then every subgroup of G is normal.
3. Let G = GLn (k) be the set of all square invertible matrices of order n
over a field k and let H = SLn (k) be the subset of elements from GLn (k) with
determinant equal to 1. Then H is a normal group in G.
If G and H are groups, then a map f : G → H such that f (ab) = f (a)f (b),
for all a, b ∈ G, is called a group homomorphism. The kernel of f is defined
by Kerf = {a ∈ G : f (a) = ē}, where ē is the identity of H. The image of f is
a set of elements of H of the form f (a), where a ∈ G, that is, Im(f ) = {h ∈ H :
∃a ∈ G, h = f (a)}. It is easy to show that Kerf is a subgroup of G and Imf is a
subgroup of H. If f is injective, i.e., Kerf = 1, f is called a monomorphism. If f
is surjective, i.e., Imf = H, f is called an epimorphism. If f is a bijection, then
f is called an isomorphism. In the case G = H, f is called an automorphism.
Quotient groups play an especially important role in the theory of groups ow-
ing to their connection with homomorphisms of groups. Namely, for any normal
subgroup N the quotient group G/N is an image of the group G. And conversely,
if G is a homomorphic image of a group G, then G is isomorphic to some quotient
group of G.
π
G G/Ker(ϕ) (1.3.1)
ϕ
ψ
G1
At one time groups of permutations were the only groups studied by mathe-
maticians. They are incredibly rich and complex, and they are especially impor-
tant because in fact they give all possible structures of finite groups as shown by
the famous Cayley theorem. This theorem establishes a relationship between the
subgroups of the symmetric group Sn and every finite group of order n.
Corollary 1.3.3. Any finite group G is a subgroup of GLn (k), where n = |G|
and k is a field.
Proof. This follows from the injection of Sn into GLn (k) given by the following
rule: σ → Aσ , where (Aσ )ij = 1 if σ(j) = i and (Aσ )ij = 0 if σ(j) = i for any
σ ∈ Sn .
The following statements are simple corollaries from the Lagrange theorem.
Corollary 1.3.4. The order of any element of a group G divides the order
of G.
The proof follows from the Lagrange theorem and the fact that the order of an
element is equal to the order of the cyclic subgroup generated by this element.
2 Mathematics has a dearth of words; so a word like ’order’ is used in many different meanings
14 ALGEBRAS, RINGS AND MODULES
Proof. By corollary 1.3.4, the orders of all elements of a finite group G must
be divisors of the order of this group. Therefore, for any non-trivial element g ∈ G
the cyclic subgroup Gp = {1, g, g 2, ..., g p−1 } coincides with G.
Examples 1.3.2.
1. An element of a group has order 1 if and only if it is the identity.
2. In the additive groups Z, Q, R and C every nonzero element has infinite
order.
3. The group Gn of all rotations of the plane which carry a regular n-sided
polygon to itself, is cyclic of order n with (as one possible) generator g which is a
rotation through 2π/n.
4. The additive group Zn , the ring of integers modulo n, is also a cyclic group,
with generator 1 ∈ Z and the identity element 0 ∈ Z.
5. The set Cn of all complex numbers satisfying the equality xn = 1 with
respect to the usual multiplication is a cyclic group.
6. Let p be a prime number. The set C∞ p of all complex roots of the equality
pn
x = 1 for some n ∈ N forms an infinite Abelian group. This group is isomorphic
to the quotient group Q/Zp , where Zp = { m n ∈ Q | (n, p) = 1}.
In terms of generators and relations this group can be defined by the the
countable set of generators a1 , a2 , . . . , an and relations:
ap1 = 1, apn+1 = an , n = 2, 3, . . .
Note that all proper subgroups of Cp∞ are finite.
7. Let F be a finite field. Then the multiplicative group F ∗ = F \ {0} is cyclic.
In general, the finite subgroups of the multiplicative group of a field are cyclic.
Gx = {y ∈ X : y = g · x, g ∈ G},
which is the equivalence class of x under the relation ∼ and it is called the orbit
of x under the action of G.
Obviously, the orbits of any two elements, being equivalence classes, either
coincide or are disjoint. So we have a partition of the set X into disjoint orbits.
If there is only one orbit, which then is the set X, one says that the group
G acts transitively on X. In other words, the group G acts transitively on the
set X if for each two elements x1 , x2 ∈ X there is an element g ∈ G such that
g · x1 = x2 .
Example 1.4.1.
1. For any nonempty set X the symmetric group SX acts on X by σ · x = σ(x),
for all σ ∈ SX , x ∈ X.
2. Let G be any group and let X = G. Define a map from G × X to X by
g · x = gx, for each g ∈ G and x ∈ X, where gx on the right hand side is the
product of g and x in the group G. This gives a group action of G on itself. This
action is called the left regular action of G on itself.
3. The additive group Z acts on itself by z · a = z + a, for all z, a ∈ Z.
16 ALGEBRAS, RINGS AND MODULES
The set
StG (x) = {g ∈ G : g · x = x}
is called the stabilizer of an element x ∈ X. It is easy to show that StG (x) is a
subgroup of G.
The set
Fix(g) = {x ∈ X : g · x = x}
is called the set of fixed points of an element g ∈ G.
Proof. Let A be a set of all left cosets of G by StG (x). Consider the map
ϕ : Gx → A defined by ϕ(g · x) = gStG (x). It is easy to see that ϕ is a bijection
and so we obtain the statement of the proposition.
In other words, two elements of G are conjugate in G if and only if they are in
the same orbit of G acting on itself by conjugation. An orbit of G under conjuga-
tion is called a conjugacy class C and (hence) defined as the set of elements of
a group G which can be obtained from a given element of the group G by conju-
gation. Obviously, distinct conjugacy classes are disjoint. Let C1 , C2 , ..., Ck be all
distinct conjugacy classes of a group G. Then we have a partition of the group G:
G = C1 ∪ C2 ∪ ... ∪ Ck
where Ci ∩ Cj = ∅ if i = j.
A group G acts on the set P(G) of all subsets of itself by defining g · S = gSg −1
for any g ∈ G and S ∈ P(G). As above, this defines a group action of G on P(G).
Note that if S is the one element set {s} then g · S is the one element set {gsg −1 }
GROUPS AND GROUP RINGS 17
Theorem 1.4.2.
1.The number of subgroups of a finite group G which are conjugate to a given
subgroup H is equal to |G : NG (H)|.
2. The number of elements of a group G which are conjugate to a given element
g ∈ G is equal to |G : CG (g)|.
Proof.
1. A group action of a group G on the set M = {xHx−1 : x ∈ G} is given
by conjugation: g · xHx−1 = gxHx−1 g −1 for any g ∈ G. This action is transitive
and StG (H) = NG (H). Then, by proposition 1.4.1, |M | = |G : NG (H)|.
2. This assertion of the proposition follows from the observation that
NG ({g}) = CG (g).
Theorem 1.4.3 (The class equation). Let G be a finite group and let
g1 , g2 , ..., gk be representatives of the distinct conjugacy classes of G not contained
18 ALGEBRAS, RINGS AND MODULES
k
|G| = |Z(G)| + |G : CG (gi )|.
i=1
Proof. Let G be a finite group and x ∈ G. Notice that the element {x} is a
conjugacy class of size 1 if and only if x ∈ Z(G). Let Z(G) = {z1 = 1, z2 , ..., zm },
and let C1 , C2 , ..., Ck be all distinct conjugacy classes of G not contained in the
center. Let gi be a representative of Ci for each i. Then the full set of conjugacy
classes of G is given by
k
|G| = m + |Ci |.
i=1
k
|G| = |Z(G)| + |G : CG (gi )|.
i=1
Proof. Let G be a finite group. Then, by theorem 1.4.3 (the class equation),
k
|G| = |Z(G)| + |G : CG (gi )|,
i=1
Proof. Let G be a group of order p2 , where p is prime, and let Z(G) be its
center. Suppose G is not Abelian, then Z(G) = G. From theorem 1.4.4 it follows
GROUPS AND GROUP RINGS 19
that |Z(G)| = p and |G/Z(G)| = p. Therefore the group G/Z(G) is cyclic. Let
x ∈ G/Z(G). Then xZ(G) is a generator of G/Z(G). So any element g ∈ G can
be written in the form g = xk z, where z ∈ Z(G). But any two elements of this
form commute, so we have a contradiction.
Proof. We shall prove this theorem by induction on the order of the group G.
If |G| = 1 then there is no p which divides its order, so the condition is trivial. If
G is an Abelian group, then this theorem immediately follows from the structure
theorem for finite Abelian groups (see, vol.I, theorem 7.8.6).
Suppose G is not Abelian, |G| = pn m > 1, where p is prime, (p, m) = 1, and
suppose the proposition holds for all groups of smaller order. Let Z(G) be the
center of G. Suppose the order of Z(G) is divisible by p. Let |Z(G)| = pr t, where
(p, t) = 1. Since G is not Abelian, |Z(G)| < |G| and, by the inductive hypothesis,
Z(G) has a subgroup H ⊂ Z(G) such that |H| = pr . As a subgroup of the centre
H is normal (or the fact that Z(G) is Abelian), so the quotient group G/H is well-
defined and of order pn−r m. By the induction hypothesis, G/H contains a Sylow
p-group K = P/H of order pn−k . Then the inverse image P = π −1 (K) ⊂ G under
the natural projection π : G → G/H is a subgroup of order |P | = |P : H|·|H| = pn ,
that is, P is a Sylow p-subgroup in G.
Now suppose s = |Z(G)| is not divisible by p. Let C1 , C2 , ..., Ck be all distinct
conjugacy classes of G not contained in the center, and let ni be a number of
elements of Ci , i = 1, 2, .., k. Then |G| = pn m = s + n1 + n2 + ... + nk . Therefore
there exists j such that nj is not divisible by p and |G : CG (gj )| = nj , where gj is
a representative of the class Cj , by theorem 1.4.3. Then |H| = pn t < |G|, where
H = CG (gj ). By the induction hypothesis, arguing as before, there is a subgroup
K ⊂ H ⊂ G such that |K| = pn , that is, K ∈ Sylp (G).
|R| · |P |
is a group and |RP | = , by the first isomorphism theorem (see theorem
|R ∩ P |
1.3.3, vol.I). Since P is a subgroup of RP , pn divides its order |RP |. But R is a
subgroup of Q, and |P | = pn , so |R| · |P | is a power of p. Then it must be that
|RP | = pn because RP ⊃ P , and therefore P = RP , and so R ⊆ P . Obviously,
R ⊂ Q, so R ⊆ Q ∩ P . Thus R = Q ∩ P .
The following construction will be used in the proof of the second and the third
Sylow theorem.
Given any Sylow p-subgroup P , consider the set of its conjugates Ω. Then
X ∈ Ω if and only if X = xP x−1 for some x ∈ G. Obviously, each X ∈ Ω is a
Sylow p-subgroup of G. Let Q be an arbitrary Sylow p-subgroup of G. We define
a group action of Q on Ω by:
g · X = g · xP x−1 = gxP x−1 g −1 = (gx)P (gx)−1 ,
for all g ∈ Q. Then we can write Ω as a disjoint union of orbits under the group
action of Q on the set Ω:
Ω = Ω1 ∪ Ω2 ∪ ... ∪ Ωr ,
where |Ω| = |Ω1 | + |Ω2 | + ... + |Ωk |. Let Pi be a representative of the orbit Ωi , for
i = 1, 2, ..., r. By proposition 1.4.1, |Ωi | = |Q : StQ (Pi )| = |Q : NQ (Pi )|. Using
proposition 1.4.7, we have
|Ωi | = |Q : Q ∩ Pi |.
If Q = Pi , then we obtain that |Ωi | = |Pi : Pi ∩ Pi | = 1. If Q = Pi , then we obtain
that |Ωi | = |Q : Q ∩ Pi | > 1, and since the index of any subgroup of Q divides Q,
p divides |Ωi |.
Theorem 1.4.9 (The second Sylow theorem). Any two Sylow p-subgroups
of a finite group G are conjugate.
Proof. Given a Sylow p-subgroup P and any other Sylow p-subgroup Q, con-
sider again the construction considered above. Suppose Q is not conjugate to P ,
Then Q = Pi for each i = 1, 2, ..., s. Therefore p divides |Ωi | for every orbit. If t is
the number of conjugates of P , then t ≡ 0(modp), which contradicts proposition
1.4.8.
Proof. Consider again the construction considered above. Since all Sylow p-
subgroups are conjugate, |Ω| is equal to the number np of all Sylow p-subgroups
of G. By proposition 1.4.8, np ≡ 1(modp).
Finally, since all Sylow p-subgroups are conjugate, theorem 1.4.2 shows that
np = |G : NG (P )| for any P ∈ Sylp (G). Since P is a subgroup of NG (P ), pn
divides |NG (P )|, hence np |m.
Example 1.5.1.
The subgroup H of all upper triangular matrices of the group GL(n, C), where
C is the field of complex numbers, is solvable.
Remark 1.5.1. Note that the term ’solvable’ arose in Galois theory and is
connected with the problem of solvability of algebraic equations in radicals. Let
f be a polynomial in x over a field k and K be the (minimal) splitting field of f .
The group Gal(K/k) is called the Galois group of f . The main result of Galois
theory says that the equation f (x) = 0 is solvable in radicals if and only if the
group Gal(K/k) is solvable.
22 ALGEBRAS, RINGS AND MODULES
Define also
G = {[x, y] : x, y ∈ G},
which is the subgroup of G generated by commutators of elements from G and
called the commutator subgroup of G.
The basic properties of commutators and the commutator subgroup are given
by the following statement.
Proof.
(1) This is immediate from the definition of [x, y].
(2) By definition, H G if and only if g −1 hg ∈ H for all g ∈ G and all h ∈ H.
For h ∈ H, we have that g −1 hg ∈ H if and only if h−1 g −1 hg ∈ H, so that H G
if and only if [h, g] ∈ H for all h ∈ H and all g ∈ G. Thus, H G if and only if
[H, G] ⊆ H.
(3) Let xG and yG be arbitrary elements of G/G . By the definition of the
group operation in G/G and since [x, y] ∈ G we have
Definition. For any group G define the following sequence of subgroups in-
ductively:
1 = H0 H1 H2 ... Hs = G
such that each quotient group Hi+1 /Hi is Abelian. We prove by induction that
G(i) ⊆ Hs−i . This is true for i = 0, so assume G(i) ⊆ Hs−i . Then
Proof.
1. This follows from the observation that since H ⊆ G, by definition of com-
mutator subgroups, [H, H] ⊆ [G, G], that is, H (1) ⊆ G(1) . Then, by induction,
H (i) ⊆ G(i) for all i ≥ 0. In particular, if G(n) = 1 for some n, then also H (n) = 1.
24 ALGEBRAS, RINGS AND MODULES
Proof. If G is a finite p-group, then, by theorem 1.4.4, its center Z(G) is not
trivial. Then the quotient group G/Z(G) is again a p-group, whose order is less
then the order of G. We prove this theorem by induction on the order of a group.
Assume that theorem is true for all p-groups with order less then pn . Then, by
induction hypothesis, Z(G) and G/Z(G) are solvable groups. Then, by theorem
1.5.4(3), G is also solvable.
Example 1.5.2.
The subgroup H of all upper triangular matrices of the group GL(n, C), where
C is the field of complex numbers, is not nilpotent. But the subgroup N of all
elements of H with 1 on the main diagonal is nilpotent.
In this case, however, Zm−2 would have index p2 in G, so G/Zm−2 (G) would be
Abelian, by corollary 1.4.5. But then G/Zm−2 (G) would equal its center and so
Zm−1 (G) would equal G, a contradiction. This proves that the nilpotency class of
G is ≤ m − 1.
We now give another equivalent definition of a nilpotent group using the notion
of a lower central series.
Recall that the commutator of two elements x, y in a group G is defined as
Proof. This follows immediately from theorem 1.5.7 taking into account that
G(i) ⊆ Gi for all i.
26 ALGEBRAS, RINGS AND MODULES
Thus, we can summarize the results obtained in this section as the following
chain of classes of groups:
(cyclic groups) ⊂ (Abelian groups) ⊂ (nilpotent groups) ⊂
⊂ (solvable groups) ⊂ (all groups)
Proof.
1. This follows from the observation that since H ⊆ G, by definition of commu-
tator subgroups, [H, H] ⊆ [G, G], that is, H 1 ⊆ G1 . Then, by induction, H i ⊆ Gi
for all i ≥ 0. In particular, if Gn = 1 for some n, then also H n = 1. And from
theorem 1.5.7 it follows that H is nilpotent.
2. Note that, by the definition of commutators, ϕ([x, y]) = [ϕ(x), ϕ(y)], so,
by induction, ϕ(Gi ) ⊆ K i . Since ϕ is surjective, every commutator in K is the
image of a commutator in G. Hence again, by induction, we obtain equality for
all i. Again, if Gn = 1 for some n then K n = 1.
(All sums in these formulas are finite.) This algebra is denoted by kG; the
elements of G form a basis of this algebra; multiplication of basis elements in the
group algebra is induced by the group multiplication. The algebra kG is isomorphic
to the algebra of functions defined on G with values in k which
assume only a finite
number of non-zero values. The function associated to αg g is f : g → αg . In
g∈G
this algebra multiplication is the convolution of such functions. Indeed if f1 , f2
are two functions G → k with finite support their product is given by
f1 f2 (g) = f1 (h)f2 (h−1 g).
h∈G
GROUPS AND GROUP RINGS 27
The same construction can also be considered for the case when k is an asso-
ciative ring. One thus arrives at the concept of the group ring of a group G over
a ring k; if k is commutative and has a unit element, the group algebra is often
called the group algebra of the group over the ring as well.
Note that, by definition of the multiplication, kG is a commutative ring if and
only if G is an Abelian group.
Examples 1.6.1.
1. If G = (g) is a cyclic group of order n and k is a field, then the elements of
kG are of the form
n−1
αi g i .
i=0
is commutative for all g ∈ G, that is, θϕ(g) = ψ(g)θ, for all g ∈ G, which is
equivalent to
ψ(g) = θϕ(g)θ−1 .
In the case where V is of finite dimension n it is common to choose a basis for V and
assign to each operator T (g) its matrix Tg in this basis. The correspondence g →
Tg defines a homomorphism of the group G into GL(n, k), the general linear group
of invertible n × n matrices over k, which is called the matrix representation of
the group G corresponding to the representation T . Thus we can define a matrix
representation of a group.
28 ALGEBRAS, RINGS AND MODULES
Example 1.6.2.
Consider the cyclic group C3 = {1, u, u2}, where u3 = 1. This group has a
two-dimensional representation ϕ over the field of complex numbers C:
1 0 1 0 2 1 0
ϕ(1) = , ϕ(u) = , ϕ(u ) =
0 1 0 ξ3 0 (ξ3 )2 ,
√
where ξ3 = − 21 + 12 i 3 is a primitive 3-rd root of unity. This representation is
faithful because ϕ is a one-to-one map.
Remark 1.6.1. The representations considered above are also called linear
representations. Other kinds of representations are permutation representa-
tions. A permutation representation of a group G on a set S is a homomor-
phism from G to the group of all permutations of S. In this book “representation”
usually means “linear representation”.
In this chapter we restrict our attention to finite groups and finite dimensional
representations over a field k.
Examples 1.6.4.
1. Let V be a one-dimensional vector space over a field k. Make V into a
kG-module by letting g · v = v for all g ∈ G and v ∈ V . This module corresponds
to the representation ϕ : G → GL(V ) defined by ϕ(g) = I, for all g ∈ G, where
I is the identity linear transformation. The corresponding matrix representation
is defined by ϕ(g) = 1. This representation of the group G is called the trivial
representation. Thus, the trivial representation has degree 1 and if |G| > 1, it
is not faithful.
GROUPS AND GROUP RINGS 29
√ √
−1 0 1
√−1 − 3 1 −1
√ 3
ϕ(c) = , ϕ(d) = , ϕ(f ) =
0 1 2 3 −1 2 − 3 −1.
The matrices ϕ(a), ϕ(b) and ϕ(c) correspond to reflections, while the matrices ϕ(d)
and ϕ(f ) correspond to rotations. These matrices form a faithful two-dimensional
representation of S3 . Take for example the equilateral triangle with corner points
√ √ 2√
(−1, − 31 3), (1, − 1
3 3), (0, 3 3) in E .
2
Dn = {σ, τ : σ n = τ 2 = 1, στ = τ σ −1 }.
compute that
cos 2π
n − sin 2π
n 0 1
S= and T =
sin 2π
n cos 2π
n
1 0
cos 2π
n − sin 2π
n 0 1
And so the map σ → and τ →
extends uniquely to
sin 2π
n cos 2π
n
1 0
a representation of Dn into GL2 (R). The matrices S and T have order n and 2
respectively. It is not difficult to check that S and T generate a matrix group of
order 2n so that this representation is faithful.
5. Consider the quaternion group which has the following presentation:
H2 = {i, j : i4 = 1, i2 = j 2 , ij = ji−1 }.
(i)
for all g ∈ G, where Sg are matrix representations of degree ni < n of G; otherwise
it is called reducible.
A matrix representation T of degree n of a group G is completely reducible
if and only if it is equivalent to a matrix representation S of the form:
⎛ (1) ⎞
Sg 0 ... 0
⎜ (2) ⎟
⎜ 0 Sg ... 0 ⎟
Sg = ⎜ ⎟
⎝ ... ... ... ... ⎠
(m)
0 0 . . . Sg
(i)
for all g ∈ G, where the Sg are irreducible matrix representations of degree ni < n
of G (i = 1, 2, ..., m).
Example 1.6.5.
The representations ϕ and ψ of the group C3 considered in examples 1.6.2
and 1.6.3 are completely reducible and are direct sums of two one-dimensional
representations.
We shall prove below a very important classical theorem. Taking into account
its great importance we give three different proofs of it.
Proof.
1. (A proof according to J.-P.Serre.3 Let M be any kG-module and let X be
an arbitrary kG-submodule of M . Since X is a vector subspace of M , there is a
vector subspace Y0 such that M is a direct sum of these vector subspaces:
M = X ⊕ Y0 .
3 See [Serre, 1967] .
32 ALGEBRAS, RINGS AND MODULES
1 1
= h(h−1 g −1 ) · π0 ((gh)m) = h(r−1 · π0 (rm)) =
n n
g∈G r=gh,g∈G
1 1 −1
= h(r−1 π0 r(m)) = h( r π0 r(m)) = hπ(m),
n n
r=gh,g∈G r∈G
where m = |G|.
Let R = rad(kG) be the Jacobson radical of kG. Since [kG : k] = m < ∞, i.e.,
kG is a finite dimensional k-algebra, kG is an Artinian algebra. Therefore R is
nilpotent, by proposition 3.5.1 (vol. I). Suppose R = 0, then there is an element
0 = x ∈ R. Let x = α1 g1 + α2 g2 + ... + αm gm .
Without lost of generality, we can assume that α1 = 0. Since R is an ideal in
kG, y = xg1−1 ∈ R and
y = α1 · 1 + α2 h2 + ... + αm hm , (1.6.1)
where hi ∈ G, y = 0, α1 = 0.
Since y ∈ R and R is nilpotent, the element y is nilpotent as well.
Therefore from any course in linear algebra it is well known that the corre-
sponding linear transformation Ty is nilpotent and Sp(Ty ) = 0.
On the other hand from (1.6.1), taking into account the linear properties of
trace, we have
(1) (2)
Sab = Sa Sb for all a, b ∈ G, we have Uab = Sa Ub + Ua Sb and hence
(2) (2)
Ua = Uab [Sb ]−1 − S(1)
a Ub [Sb ]
−1
.
and
(1)
−1 Sg 0
C Sg C = (2) ,
0 Sg
that is, S is a completely reducible representation.
Remark 1.6.2. This theorem was proved by H.Maschke in 1898 for finite
groups when the field k has characteristic 0. For fields whose characteristic does
not divide the order of a group G the result was pointed out by L.E.Dickson.
Note, that this theorem is also valid in a more general case, namely when k is a
commutative ring and |G| · 1 is a unit in k, which can be seen from the third proof
of this theorem.
Remark 1.6.3. The converse of the Maschke theorem is also true. Namely, if
the characteristic of a field k does divide |G|, then G possesses finitely generated
kG-modules which are not completely reducible. Specifically, the module kG itself
is not completely reducible. Indeed, let e = g. Since ge = eg = e for each
g∈G
g ∈ G, e spans a one-dimensional ideal I in kG. Since e2 = 0, this ideal is nilpotent.
Since kG is an Artinian ring (as a finite dimensional algebra), its radical rad(kG)
is nilpotent and contains all nilpotent ideals. So I ⊂ rad(kG). Hence rad(kG) = 0
and kG is not semisimple.
In this book we shall generally restrict our attention to finite groups in the
classical case, as this simplifies things and provides a more complete theory.
From the Maschke theorem and the theory of semisimple algebras and modules
one can easily obtain a number of important results describing the irreducible
representations of a finite group G over an algebraically closed field k, whose
characteristic does not divide |G|.
where n = |G|.
Proof. By corollary 1.7.2, kG Mn1 (k) × Mn2 (k) × ... × Mns (k), where
n1 , n2 , ..., ns are the dimensions of all the irreducible representations of the al-
gebra kG, and thus n = |G| = [kG : k] = n21 + n22 + ... + n2s .
Example 1.7.1.
For the symmetric group G = S3 we have that |G| = 6 and we have two
one-dimensional irreducible representations and one two-dimensional irreducible
representation, as described in example 1.6.2 (3). Thus, using (1.7.3), we have
6= n2i = 12 + 12 + 22 .
i
So, theorem 1.7.6 tells us that there are no additional distinct irreducible repre-
sentations of S3 over an algebraically closed field k whose characteristic does not
divide |G|.
right ideals Mi , and these right ideals give a complete set of isomorphism classes
of irreducible kG-modules, we have the corresponding decomposition of the regular
kG-module kG over an algebraically closed field k :
kG n1 M1 ⊕ n2 M2 ⊕ ... ⊕ ns Ms , (1.7.4)
This means that the coefficients of g and hgh−1 in the element x are equal. Since
h is arbitrary, every element in the same conjugacy class of a fixed group element
g has the same coefficient in x, hence x can be written as a linear combination of
the ci ’s. Since the ci ’s are linearly independent, they form a basis of the center
Z(kG), that is, dimk Z(kG) = s.
Example 1.7.2.
For the group S3 there are three conjugacy classes: {e}, {a, b, c} and {d, f }.
(See examples 1.6.4(3)). Thus, by theorem 1.7.8, there are three irreducible repre-
sentations which, as we have seen in example 1.7.1, consist of two one-dimensional
representations and one two-dimensional representation.
Corollary 1.7.10. If G and H are Abelian groups of the same order and k
is an algebraically closed field whose characteristic does not divide |G|, the group
algebras kG and kH are isomorphic.
1.8 CHARACTERS OF GROUPS. ORTHOGONALITY RELATIONS AND
THEIR APPLICATIONS
This section is an introduction to the theory of characters of groups which is one
of the important methods for the study of groups and their representations. We
shall consider the main properties of characters and their applications to obtain
some important results.
Let V be a vector space over a field k with a basis v1 , v2 , ..., vn , and let ϕ ∈
GL(V ) be a linear transformation with corresponding matrix A = (aij ) on this
basis. The trace of the transformation ϕ is the trace of the matrix A:
n
Sp(ϕ) = Sp(A) = aii .
i=1
From any course on linear algebra it is well known that the trace of the matrix
A does not depend on the choice of a basis v1 , v2 , ..., vn . Indeed, if B = PAP−1
then Sp(B) = Sp(PAP−1 ) = Sp(A). From the definition it follows immediately,
that a trace is a linear function, i.e.,
Sp(αϕ) = αSp(ϕ)
χσ (x) = Sp[σ(x)].
Example 1.8.1.
The character of the representation ϕ of the group C3 considered in example
1.6.2 is given by: χ(1) = 2, χ(u) = 1 + ξ3 , χ(u2 ) = 1 + (ξ3 )2 .
Notice that, by corollary 1.7.5, there is only a finite number of irreducible char-
acters of a finite group G over an algebraically closed field k whose characteristic
does not divide |G|.
The character of the regular representation is called the regular character
and denoted by χreg .
Examples 1.8.2.
1. The character of the trivial representation is the function χ(g) = 1 for all
g ∈ G. This character is called the principal character of G.
2. For representations of degree 1, the character and the representation are
usually identified. Thus for Abelian groups, irreducible representations over an
algebraically closed field and their characters are the same.
Proof.
1. Since dimk V = n, (ϕ(1)) = E is the n× n identity matrix. Thus Sp(ϕ(1)) =
Sp(E) = n, hence χϕ (1) = n.
2. It is well known that Sp(ab) = Sp(ba) for any a, b ∈ GL(V ). Then setting
a = v −1 , b = vu, we obtain that Sp(u) = Sp(vuv −1 ).
So equivalent representations have the same characters. Therefore
χϕ (gxg −1 ) = Sp[ϕ(gxg −1 )] = Sp[ϕ(g)ϕ(x)ϕ(g −1 )] = Sp[ϕ(x)] = χϕ (x) for all
g, x ∈ G.
n for g = 1
χreg (g) =
0 1
for g =
The proof follows immediately from proposition 1.8.1 and example 1.6.2(2).
40 ALGEBRAS, RINGS AND MODULES
Examples 1.8.3.
1. Let ϕ : D2n → GL2 (R) be the explicit matrix representation described in
example 1.6.2(3). If χ is the character of ϕ, then χ(σ) = 2 cos( 2π
n ) and χ(τ ) = 0.
Since ϕ takes the identity of D2n to the 2 × 2 identity matrix, χ(1) = 2.
2. Let ϕ : H2 = Q8 → GL2 (C) be the explicit matrix representation described
in example 1.6.4(5). If χ is the character of ϕ, then χ(i) = 0 and χ(j) = 0. Since
ϕ takes the identity of Q8 to the 2 × 2 identity matrix, χ(1) = 2.
the module M
M a1 M1 ⊕ a2 M2 ⊕ ... ⊕ as Ms , (1.8.1)
where the ai are nonnegative integers indicating the multiplicity of the irreducible
module Mi in the direct sum of the decomposition of M . If χM is the character
afforded by the module M , then, by proposition 1.8.3,
χM = a1 χ1 + a2 χ2 + ... + as χs (1.8.2)
χM = a1 χ1 + a2 χ2 + ... + as χs
and
χN = b1 χ1 + b2 χ2 + ... + bs χs ,
where the ai , bi are nonnegative integers indicating the multiplicity of the irre-
ducible module Mi in the decompositions of the modules M and N .
If i = j, then ej Mi = 0, i.e., ej acts on Mi trivially, hence χi (ej ) = 0. If
i = j, then ei Mi = Mi , i.e., ei acts on Mi as an identity, hence χi (ei ) = ni ,
where ni = dimk Mi . Therefore χM (ei ) = ai ni and χN (ei ) = bi ni for all i. Since
the characters χM and χN are equal, ai = bi for all i, i.e., M N , and the
representations T and R are equivalent.
Proof. This follows immediately from proposition 1.8.4 and theorem 1.7.7 (see
equality 1.7.4).
Note that the center Z(kG) has two different natural bases. Let 1 = e1 +
e2 + ... + es be a decomposition of the identity of the algebra kG into a sum of
primitive central idempotents. Since the idempotents e1 , e2 , ..., es are orthogonal
and central, they form a basis of the center Z(kG).
42 ALGEBRAS, RINGS AND MODULES
On
the other hand, as we have already pointed out in section 1.7, the elements
ci = g, for i = 1, 2, ..., s, also form a basis of Z(k). Consequently, there are
g∈Ci
s
s
elements αij , βij ∈ k such that ci = αij ej and ei = βij cj ; and the matrices
j=1 j=1
A = (αij ) and B = (βij ) are reciprocal (inverses of each other).
s
s
χj (ci ) = χj ( αik ek ) = αik χj (ek ) = nj αij .
k=1 k=1
On the other hand, χj (ci ) = hi χj (gi ), and the formula for the αij follows.
n
In order to compute βij we use corollary 1.8.5, which says that χreg = n i χi .
i=1
From proposition 1.8.1 it follows that χreg (ck g) = 0 if g −1 ∈ Ck and χreg (ck g) = n
if g −1 ∈ Ck , where n = |G|. Therefore, if gj ∈ Cj , then
s
χreg (ei gj−1 ) = χreg ( βik ck gj−1 ) = nβij .
k=1
s
χreg (ei gj−1 ) = ni χk (ei gj−1 ) = ni χi (gj−1 )
k=1
because χk (ei gj−1 ) = 0 for k = i and χi (ei gj−1 ) = χi (gj−1 ). The formula for the βij
follows.
Taking into account that the matrices A and B are reciprocal, we obtain
immediately the following ”orthogonal relations” for characters, obtained by
F.G.Frobenius.
1
s
0 for i = j,
χk (gi )χk (gj−1 ) =
n 1/hj for i = j.
k=1
1
s
hk χ(gk )χ(gk−1 ) = 1. (1.8.5)
n
k=1
1 1
s s s
hk χ(gk )χ(gk−1 ) = mi mj hk χi (gk )χj (gk−1 ) = m2i ,
n n i,j
k=1 k=1 k=1
and this sum is equal to 1 if and only if χ = χi for some i, i.e., in view of proposition
1.8.4, if and only if T is an irreducible representation.
The theory of characters has the most applications in the case when k = C is
the field of complex numbers. Therefore we restate the most important results in
this case.
Example 1.8.4.
Consider the representation ϕ of the group S3 considered in example 1.6.2(3).
Since we have three conjugacy classes of this group: {e}, {a, b, c} and {d, f }, so we
have h1 = 1, h2 = 3 and h3 = 2, respectively. The corresponding charactervalues
are χ(e) = 2, χ(a) = 0, χ(d) = −1.
Forming the sum in (1.8.6) and using the fact that |S3 | = 6, we obtain
3
1
hk χ(gk ) · χ(gk ) = 1 × 4 + 3 × 0 + 2 × 1 = 1,
6
k=1
Let k be an algebraically closed field and let χ1 , χ2 , ..., χs be all the irreducible
characters of a finite group G over the field k. Let C1 , C2 , ..., Cs be all the conjugacy
classes of the group G. Write χij = χi (gj ), where gj ∈ Cj . The square matrix
X = (χij ) is called the character table of G over k with conjugacy classes of
elements as the columns and characters as the rows. Character tables are central
to many applications of group theory to physical problems.
If gj ∈ Cj , then χi (gj−1 ) = χij . So using proposition 1.8.9 we can rewrite
theorem 1.8.7 for the case of the field of complex numbers in the following form:
Using this definition, theorem 1.8.11 can be rewritten in the following form:
The orthogonality relations (1.8.7) show that the rows and columns of a char-
acter table over the field C are orthogonal with respect to the Hermitian inner
products, which allows one to compute character tables more easily. Note, that the
first row of the character table always consists of 1’s, and corresponds to the trivial
representation. Certain properties of a finite group G can be deduced immediately
from its character table:
1. The order of G is given by the sum of (χ(1))2 over the characters in the
table.
2. G is Abelian if and only if χ(1) = 1 for all characters in the table.
3. G is not a simple group if and only if χ(1) = χ(g) for some non-trivial
character χ in the character table and some non-identity element g ∈ G.
Examples 1.8.5.
1. Consider the cyclic group Cn = (g), where g n = 1. This group may be
realized as the rotation group over the angles 2kπ/n over some line. This group
is commutative, therefore, by theorem 1.7.3, each irreducible representation over
the field of complex numbers is one-dimensional. Each such representation has a
character χ such that χ(g) = w ∈ C and χ(g k ) = wk . Since g n = 1, wn = 1, that
is, w = e2πk/n , where k = 0, 1, ..., n−1. Therefore, the n irreducible representations
of the group Cn have the characters χ0 , χ1 , χ2 , ..., χn−1 , defined by the formula
χh (g k ) = e2πihk/n .
C3 1 g g2
χ0 1 1 1
χ1 1 w w2
χ2 1 w2 w
46 ALGEBRAS, RINGS AND MODULES
1 + 3α + 2β = 0,
1 + 3α2 + 2β 2 = 6.
Using the orthogonality relations between the columns of the character table
we obtain:
1 + α + 2γ = 0
1 + β + 2δ = 0.
Using the relations between the elements of the group S3
a2 = e, b2 = e, c2 = e, d2 = f,
we obtain, that α2 = 1 and β 2 = β. Therefore taking into account all these
relations it follows that α = −1, β = 0, γ = 0 and δ = −1. Thus the complete
character table for S3 is given by
S3 e {a,b,c} {d,f}
χ0 1 1 1
χ1 1 -1 1
χ2 2 0 -1
Theorem 1.8.13. Let G be a finite group, and let the χi be all the irreducible
characters of the group G over the field of complex numbers C corresponding to
irreducible representations Mi , i = 1, 2, ..., s. Let M be a linear representation of
G with a character χ. Then
n
χ= (χ, χi )χi , (1.8.8)
i=1
GROUPS AND GROUP RINGS 47
n
(χ, χ) = a2i , (1.8.9.)
i=1
where ai is the number of representations Mi appearing in the direct sum of the
decomposition of M and ai = (χ, χi ).
Using properties 1 and 2 of the inner product and theorem 1.8.12, we obtain that
ai = (χ, χi ). Now computing the inner product (χ, χ) and using the orthogonality
relations (1.8.7) we obtain (1.8.9).
similarity. Later it became clear that the problem of the description contains, in
turn, the problem of the classification of a pair linear operators which act on a
finite dimensional vector space (these problems are called wild and the others are
called tame). In particular, it was proved that the problem of a description of
representations of a group (p, p) for p = 2 is wild. So, this problem is wild for
arbitrary non-cyclic groups. If p = 2, the problem is wild for the (2, 4) and (2, 2, 2)
groups. From these results it follows that one can hope to obtain a classification
of modular representations only for p = 2 and for some groups G such that its
quotient groups do not contain the groups (2, 4) and (2, 2, 2). These groups are
exhausted by the following three infinite sets of 2-groups:
1. Dihedral groups
m
Dm = {x, y : x2 = y 2 = 1, yx = xy −1 } (m 1);
2. Quasidihedral groups
m m−1
Qm = {x, y : x2 = y 2 = 1, yx = xy 2 −1
} (m 3);
We say that a group G is tame (wild) over a field k if the group algebra kG
is of tame (wild) representation type.7
V.Bondarenko and Yu.A.Drozd proved that a subgroup of finite index of a
tame group (over a field k) is tame (over k). It is easy to see that a generalized
quaternion group Hm is isomorphic to the subgroup of the quasidihedral group
Qm+1 , which is generated by xy and y 2 , so that generalized quaternion groups are
tame over a field with characteristic 2.
Denote by G the commutator subgroup of a group G. The above results can
be formulated in the following invariant form.
V.M.Bondarenko and Yu.A.Drozd also proved that a finite group is tame over
a field with a characteristic p > 0 if and only if its p-Sylow subgroup is tame.
This result can be formulated in the following invariant form.
and the notion of abstract groups was introduced by A.Cayley in three papers
starting in 1849 [Cayley, 1849], [Cayley, 1854a], [Cayley, 1854b], though these
papers received little attention at the time. This had certainly changed by the
1890’s and a discussion of the basic definitions and some basic properties of ab-
stract groups can be found in H.Weber’s influential treatise [Weber, 1899].
Sylow subgroups were introduced by Peter Ludwig Mejdeli Sylow (1833-
1918), a Norwegian mathematician. The Sylow theorems were proved also by
P.L.M.Sylow in 1872 [Sylow, 1872].
The founder of group representation theory was Ferdinand Georg Frobenius
(1849-1917). He discovered the amazing basic properties of irreducible group char-
acters and published them in the 1890’s. His student, Issai Schur, was another
who made many significant early contributions to the subject.
Group algebras were considered by F.G.Frobenius and I.Schur [Schur, 1905]
in connection with the study of group representations, since studying the repre-
sentations of G over a field k is equivalent to studying modules over the group
algebra kG.
H.Maschke is best known today for the Maschke theorem, which he published
in 1899. A special case of his theorem H.Maschke proved in the paper [Maschke,
1898]. The general result appeared in the following year [Maschke, 1899].
In its modern form the theory of group representations owes much to the con-
tributions of Emmy Noether (in the 1920’s), whose work forms the basis of what
is now called “modern algebra”.
Modular representation theory was developed by Richard Brauer.
The problem of dividing the class of group algebras into tame and wild ones
has been completely solved by V.M.Bondarenko and Yu.A.Drozd [Bondarenko,
Drozd, 1977].
[Hall, 1959] M.Hall, The theory of groups, Macmillan, New York, 1959.
[Herstein, 1968] I.N.Herstein, Noncommutative rings, Carus Mathematical Mono-
graphs, No.15, Mathematical Association of America, 1968.
[Higman, 1954] D.G.Higman, Indecomposable representations at characteristic p,
Duke Math. J., vol.21, 1954, p.377-381.
[Kronecker, 1870] L.Kronecker, Auseinandersetzung einiger Eigenschaften der
Klassenzahl idealer complexer Zahlen, Monatsber Kön. Preuss. Akad.
Wiss., Berlin (1870), p.881-889.
[Lagrange, 1771] J.L.Lagrange, Memoir on the algebraic solution of equations,
1771.
[Maschke, 1898] H.Maschke, Über den arithmetischen Charakter der Coefficienten
der Substitutionen endlicher linearer Substitutionsgruppen, 1898.
[Maschke, 1899] H.Maschke, Beweiss des Satzes, dass diejenigen endlichen lin-
earen Substitutionsgruppen, in welchen einige durchgehends verschwindende
Coefficienten auftreten intransitiv sind, 1899.
[Schoenflies, 1891] A.Schoenflies, Kristallsysteme und Kristallstruktur, Leipzig,
1891.
[Schur, 1905] I.Schur, Sitzungsber. Preuss. Akad. Wiss. 1905, p.406-432.
[Serre, 1967] J.-P.Serre, Represéntations linéaires des groups finis, Hermann,
Paris, 1967.
[Sylow, 1872] P.L.M.Sylow, Théorèmes sur les groupes de substitutions, Math.
Ann., v.5, N.54, 1872, p.584-594.
[Weber, 1899] H.Weber, Lehrbuch der Algebra, II, Vieweg, 1899, Buch 1, Ab-
schnitt 1.
2. Quivers and their representations
This chapter is devoted to the study of quivers and their representations. Quivers
were introduced by P.Gabriel in connection with problems of representations of
finite dimensional algebras in 1972. And from that time on the theory connected
with representations of quivers has developed enormously.
We start this chapter by considering some important algebras, including the
Grassmann algebra and the tensor algebra of a bimodule.
In sections 2.3 and 2.4 the main concepts are given pertaining to quivers, their
representations and path algebras of quivers. The main theorem in this section
says that the path algebra of a quiver over an arbitrary field is hereditary.
Dynkin diagrams and Euclidean diagrams (more often called extended Dynkin
diagrams) and the quadratic forms connected with them are studied in section 2.5.
The most important theorem, due to I.N.Berstein, I.M.Gel’fand, V.A.Ponomarev,
gives the classifications of all graphs from the point of view of their quadratic
forms.
The most remarkable result in the theory of representations of quivers is the
theorem classifying the quivers of finite representation type, which was obtained by
P.Gabriel in 1972. This theorem says that a quiver is of finite representation type
over an algebraically closed field if and only if the underlying diagram obtained
from the quiver by forgetting the orientations of all arrows is a disjoint union of
simple Dynkin diagrams. P.Gabriel also proved that there is a bijection between
the isomorphism classes of indecomposable representations of a quiver Q and the
set of positive roots of the Tits form corresponding to this quiver. The proof of
this theorem is given in section 2.6.
We end this chapter by studying representations of species, also a concept
introduced by P.Gabriel. The main results of this theory are stated here for
completeness without proofs.
53
54 ALGEBRAS, RINGS AND MODULES
M̄n Mn /Bn .
The tensor product of two graded A-modules L and M is again a graded
A-module where the grading is given by the formula:
(L ⊗ M )n = Lp ⊗ Mq .
p+q=n
Examples 2.1.1.
1. The polynomial ring A = K[x] over a field K is graded with An = Kxn .
The multiplication in A is defined by xp xq = xp+q , so that indeed deg (xp xq ) =
deg (xp ) + deg (xq ).
2. Let K be a ring, and let {x1 , x2 , . . . , xn } be a set of symbols. Consider
the free left K-module Λ = K(x1 , x2 , . . . , xn ) with as basis the elements 1 and
QUIVERS AND THEIR REPRESENTATIONS 55
all the elements (words) xi , xi1 xi2 , xi1 xi2 xi3 , . . ., xi1 xi2 . . . xin , . . ., where each
ij ∈ {1, 2, . . . , n}. Set Λn = {xi1 xi2 . . . xin }, the K-vectorspace spanned by all
∞
words of length n, then Λ = ⊕ Λn is a graded K-module. In this module we
n=0
introduce a multiplication of basis elements by the rule:
(αxi1 xi2 . . . xin )(βxj1 xj2 . . . xjm ) = αβxi1 xi2 . . . xin xj1 xj2 . . . xjm .
Examples 2.1.2.
Grassmann algebra. An important example of a graded associative algebra
is the Grassmann algebra. An associative algebra with 1 over a field K is called a
Grassmann algebra if it has a system of generators a1 , a2 , . . . , an satisfying the
following equalities:
a2i = 0, ai aj + aj ai = 0, i, j = 1, 2, . . . , n, (2.1.4)
and, moreover, any other identity for these elements is a corollary of the identities
(2.1.2). This algebra is denoted by Γn or Γ(a1 , a2 , . . . , an ). Note that the relations
(2.1.4) are homogeneous and thus generate a homogeneous ideal so that Γn inherits
a grading from K(x1 , . . . , xn ), see example 2.1.1(2) above. If the characteristic of
the field K is unequal to 2 the second group of the equalities (2.1.4) implies the first
froup. The equalities (2.1.4) are the anticommutators of the elements ai and aj .
56 ALGEBRAS, RINGS AND MODULES
From the definition of the Grassmann algebra it also follows that Γn has a basis
consisting of 1 and the following elements:
ai , i = 1, 2, . . . , n,
ai aj , i, j = 1, 2, . . . , n, i < j,
ai aj ak , i, j, k = 1, 2, . . . , n, i < j < k,
...............................
a 1 a 2 . . . an .
Tensor algebra. Another important example of graded algebras are the tensor
algebras. Let V be a fixed vector space over a field k, and let T 0 V = k, T 1 V = V ,
∞
T 2 V = V ⊗ V , . . . , T m V = V ⊗ . . . ⊗ V (m copies). Define T(V ) = ⊕ T i V .
i=0
Introduce an associative product on T(V ) defined on homogeneous generators of
T(V ) by the following rule: (v1 ⊗ . . . ⊗ vn )(w1 ⊗ . . . ⊗ wm ) = v1 ⊗ . . . ⊗ vn ⊗ w1 ⊗
. . . ⊗ wm ∈ T n+m V. This makes T(V ) an associative graded algebra with 1. If V is
a finite dimensional algebra, then T(V ) is generated by 1 together with any basis
of V .
The algebra T(V ) is the universal associative algebra in the following sense.
α
V T(V )
ψ
ϕ
We shall prove this proposition in a more general situation in the next section.
This proposition gives the motivation to introduce the following formal
definition:
θ
V T
ψ
ϕ
α
V T(V )
θ σ
is commutative.
Proof.
1. This follows immediately from proposition 2.1.1.
2. This uniqueness property of a tensor algebra is of the same type as proved
in proposition 4.5.2, vol.I, for the uniqueness of tensor products and therefore its
proof is left to the reader as a simple exercise.
58 ALGEBRAS, RINGS AND MODULES
Remark 2.1.1. Note that a tensor algebra over V is the same as a free
algebra on V , i.e., T(V ) kV . If V is a finite dimensional vector space with
basis (x1 , . . . , xn ), then T(V ) kx1 , . . . , xn .
Symmetric algebra. Many quotient algebras of the tensor algebra T(V ) are
of interest. First, there is the symmetric algebra on a given vector space V . Let I
be the two-sided ideal in T(V ) generated by all x ⊗ y − y ⊗ x , where x, y ∈ V . The
quotient algebra S(V ) = T(V )/I is called the symmetric algebra on V . We
write σ : T(V ) → S(V ) for the corresponding canonical map of k-algebras. Since
the generators of I lie in T 2 V , it obvious that I = (I ∩ T 2 V ) ⊕ (I ∩ T 3 V ) ⊕ . . ..
Therefore σ is injective on T 0 V = k and T 1 V = V . Moreover, S(V ) inherits a
∞
grading from T(V ): S(V ) = ⊕ S i V . The effect of factoring out I is universal (in
i=0
the sense of proposition 2.1.1) for linear maps from V into commutative associative
k-algebras with 1. Moreover, if dimk V = n and (x1 , . . . , xn ) is any fixed basis of
V , then S(V ) is canonically isomorphic to the polynomial algebra k[x1 , x2 , . . . , xn ]
over the field k in n variables.
u ∧ v = −v ∧ u
for all u, v ∈ V and
v1 ∧ v2 ∧ · · · ∧ vp = 0
whenever v1 , v2 , . . . , vp ∈ V are linearly dependent.
The exterior algebra is in fact the “most general” algebra with these properties.
This means that all equations that hold in the exterior algebra follow from the
above properties alone. This generality of Λ(V ) is formally expressed by a certain
universal property given below.
QUIVERS AND THEIR REPRESENTATIONS 59
is a basis for the p-th exterior power Λp (V ). The reason is the following: given
any wedge product of the form v1 ∧ v2 ∧ . . . ∧ vp , every vector vj can be writ-
ten as a linear combination of the basis vectors ej ; using the bilinearity of the
wedge product, this can be expanded to a linear combination of wedge prod-
ucts of those basis vectors. Any wedge product of elements from {e1 , . . . , en } in
which the same basis vector appears more than once is zero; any wedge product
in which the basis vectors don’t appear in the proper order can be reordered,
changing the sign whenever two basis vectors change places. The resulting co-
efficients of the basis k-vectors in a wedge product of v1 , . . . , vm can be com-
puted as the minors of the matrix that describes the vectors vj in terms of the
basis ej .
Counting the basis elements, we see that the dimension of Λp (V ) is equal to
n!
. In particular, Λp (V ) = {0} for p > n.
p!(n − p)!
So in this case the exterior algebra can be written in the following form:
Λ(V ) = Λ0 (V ) ⊕ Λ1 (V ) ⊕ Λ2 (V ) ⊕ · · · ⊕ Λn (V ),
Example 2.1.3.
Let V be a vector space with basis {e1 , e2 , e3 , e4 }. Then
Λ0 (V ) = {1}
Λ1 (V ) = {e1 , e2 , e3 , e4 }
Λ2 (V ) = {e1 ∧ e2 , e1 ∧ e3 , e1 ∧ e4 , e2 ∧ e3 , e2 ∧ e4 , e3 ∧ e4 }
Λ3 (V ) = {e1 ∧ e2 ∧ e3 , e1 ∧ e2 ∧ e4 , e1 ∧ e3 ∧ e4 , e2 ∧ e3 ∧ e4 }
Λ4 (V ) = {e1 ∧ e2 ∧ e3 ∧ e4 }
and Λk (V ) = {0}, where k > dimV .
f ⊗n : V ⊗n → A⊗n .
The next part of this section is devoted to studying the global dimension of
a tensor algebra TB (V ) of a B-bimodule V when V is a projective right (or left)
B-module.
We shall need some additional statements concerning projectivity and injec-
tivity of modules.
62 ALGEBRAS, RINGS AND MODULES
Proof. Let X be a right B-module. Since M and N are projective, the functor
HomB (N, X) is exact on X and HomA (M, HomB (N, X)) is exact on X, by propo-
sition 5.1.1, vol.I. Then, by the adjoint isomorphism (proposition 4.6.3, vol.I), the
functor HomB (M ⊗A N, X) is also exact. So, again by proposition 5.1.1, vol.I,
M ⊗A N is B-projective.
Proof. This follows from proposition 2.2.4 and corollary 2.2.5 taking into ac-
count that B Q HomS (S SB , S Q).
Proof. This follows from proposition 2.2.3 and corollary 2.2.5 taking into ac-
count that PB P ⊗S SB .
(2) inj. dimS HomB (S, M ) ≤ inj. dimB M ≤ r.gl. dim B for any right
B-module M .
Proof.
(1) Let M be a left B-module. Applying the functor S ⊗B ∗ to a projective
resolution of M and taking into account proposition 2.2.3, we obtain a projective
resolution of the S-module S ⊗B M .
(2) Let M be a right B-module. Applying the functor HomB (S, ∗) to an
injective resolution of M and taking into account proposition 2.2.4, we obtain an
injective resolution of the S-module HomB (S, M ).
Corollary 2.2.9. For any left S-module N and for any right S-module M we
have:
(1) inj. dimB N ≤ inj. dimS N ;
(2) proj. dimB M ≤ proj. dimS M .
m
m
xvi ⊗y ∈ V p+1 ⊗V q and the homogeneous component of xi ⊗vi yi of maximal
i=1 i=1
m
degree has the form x ⊗ vi y ∈ V p ⊗ V q+1 . Since (V p+1 ⊗ V q )∩(V p ⊗ V q+1 ) = 0,
i=1
m
m
m
m
we have xi vi ⊗ y i = 0 = xi ⊗ vi y i , i.e., xi vi ⊗ yi = 0 = xi ⊗ vi yi . But
i=1 i=1 i=1 i=1
m
the last equality is equivalent to the equality 0 = xi ⊗ vi ⊗ yi ∈ S ⊗B V ⊗B S.
i=1
Therefore Kerf = 0 and the lemma is proved.
0 → S ⊗B V ⊗B M −→ S ⊗B M −→ M → 0 (2.2.12)
0 → N ⊗B V ⊗B S −→ N ⊗B S −→ N → 0 (2.2.13)
Applying to the sequence (2.2.12) the functor Ext1B (X, ∗)and to the sequence
(2.2.13) the functor Ext1B (∗, X) and taking into account corollaries 2.2.8 and 2.2.9
we obtain that
2.2.2. Let I be a finite set and let Ki for i ∈ I be a set of skew fields.
Remark
Let B = Ki and and let V be a B-bimodule. In this case TB (V ) is called a
i∈I
special tensor algebra.
Example 2.2.1.
m
Let k be a field and B = k. For each n ≥ 1 let
i=1
⎛ ⎞
B 0 0 ... 0 0
⎜ . .. ... .. ⎟
⎜ V B .⎟
⎜ ⎟
⎜ .. .. .. .. .. .. ⎟
⎜ . . . . .⎟
Tn = ⎜ . ⎟
⎜ . . . .. ⎟
⎜ .. .. .. . 0⎟
⎜ ⎟
⎝V ⊗n−2 V ⊗n−3 . . . B 0⎠
V ⊗n−1 V ⊗n−2 . . . . . . V B
be the ring with addition and multiplication given by the matrix operations. Then
Tn is a special tensor algebra.
Example 2.2.2.
The special tensor algebra Tn from example 2.2.1 is a finite dimensional hered-
itary Artinian k-algebra.
Examples 2.2.3.
1. Let k be a field. The algebras
⎛ ⎞ ⎛ ⎞
k k k k 0 k
A = ⎝0 k k ⎠ and B = ⎝0 k k⎠
0 0 k 0 0 k
are finite dimensional special tensor algebras which correspond to the quivers
• •
• • • and respectively.
•
2. The algebra A = k[[α]], the ring of formal power series in one variable α
over a field k, is an infinite dimensional special tensor algebra which corresponds
to the quiver
α
t
Remark 2.2.5. Using the tensor algebra of a bimodule we can also define
the exterior algebra of a bimodule. Let B be a commutative ring with 1 and
let V be a B-bimodule. Then we can consider the two-sided ideal I which is
spanned by all elements of the form x1 ⊗ x2 ⊗ . . . ⊗ xr , where xi ∈ V and xi = xj
for some i = j. Then Λ(V ) = T(V )/I is called the exterior algebra of the
bimodule V .
• • •
α
t
1 2
• •
Examples 2.3.1.
1. For the quiver
1 2 3
• α
• •
β
1 2
• •
•
3
QUIVERS AND THEIR REPRESENTATIONS 69
and
1 2 3
• • •
3. A quiver may also have loops. For example:
t - t
For a quiver Q = (V Q, AQ, s, e) and a field k one defines the path algebra
kQ of Q over k. Recall that a path p of the quiver Q from the vertex i to the
vertex j is a sequence of r arrows σ1 σ2 ...σr such that the start vertex of each arrow
σm coincides with the end vertex of the previous one σm−1 for 1 < m ≤ r, and
moreover, the vertex i is the start vertex of σ1 , while the vertex j is the end vertex
of σr . The number r of arrows is called the length of the path p. For such a
path p we define s(p) = s(σ1 ) = i and e(p) = e(σk ) = j. By convention we also
include into the set of all paths the trivial path εi of length zero which connects
the vertex i with itself without any arrow and we set s(εi ) = e(εi ) = i for each
i ∈ V Q, and, also, for any arrow σ ∈ AQ with start at i and end at j we set
εi σ = σεj = σ. A path, connecting a vertex of a quiver with itself and of length
not equal to zero, is called an oriented cycle.
Remark 2.3.1. Note that if a quiver Q has an infinitely many vertices, then
kQ has no an identity element. If Q has infinitely many arrows, then kQ is not
finitely generated, and so it is not finite dimensional over k. In future we shall
always assume that V Q is finite and V Q = {1, 2, . . . , n}.
In the algebra kQ the set of trivial paths forms a set of pairwise orthogonal
idempotents, i.e.,
ε2i = εi for all i ∈ V Q
εi εj = 0 for all i, j ∈ V Q such that i = j.
If V Q = {1, 2, . . . , n}, the identity of kQ is the element which is equal to the sum
of all the trivial paths εi of length zero, that is, 1 = ε1 +ε2 +. . .+εn . The elements
70 ALGEBRAS, RINGS AND MODULES
The subspace εi A has as basis all paths starting at i, and the subspace Aεj has
as basis all paths ending at j. The subspace εi Aεj has as basis all paths starting
at i and ending at j.
Since {ε1 , ε2 , . . . , εn } is a set of pairwise orthogonal idempotents for A = kQ
with sum equal to 1, we have the following decomposition of A into a direct sum:
A = ε1 A ⊕ ε2 A ⊕ . . . ⊕ εn A.
Proof. Note that εi Aεi is spanned by all paths that start and end at vertex i.
In fact they form a basis. Also observe that if p is such a path of length r > 0 and
r1 + r2 = r, ri ∈ N ∪ {0} then there are unique paths x, y of lengths r1 and r2 such
that p = xy. They are of course not necessarily in εi Aεi . Let f be a nontrivial
idempotent of εi Aεi and g = e − f where e is the identity of EndA (εi A) = εi Aεi
and suppose f g = gf = 0. Take paths of maximal length x, y occurring with
nonzero coefficient in respectively f and g. Let at least one of them have length
> 0. Then we would have xy = 0 (because x ends at i and y starts at i) and
xy = x y for any other pair (x , y ) of such paths. This contradicts f g = 0 by the
uniqueness of factorization remark above. Hence f or g is a multiple of εi and the
result follows.
Examples 2.3.2.
1. Let Q be the quiver
1 2 3
• σ1
• σ2
•,
⎛ ⎞
k k k
Then kQ has a basis {ε1 , ε2 , ε3 , σ1 , σ2 , σ1 σ2 } and kQ T3 (k) = ⎝0 k k⎠ ⊂
0 0 k
M3 (k). So the algebra kQ is finite dimensional over k.
α • β
1 α 2
• •
β
i.e., V Q = {1, 2} and AQ = {α, β}. The algebra kQ hasa basis {ε
1 , ε2 , α, β}.
k k ⊕ k
This algebra is isomorphic to the Kronecker algebra1 A = , which is
0 k
four-dimensional over k.
contain oriented cycles, then we have only a finite number2 of paths in Q which
form a basis of kQ over k, so it is finite dimensional.
Remark 2.3.2. A path algebra kQis a special tensor algebra, with B the
commutative semisimple algebra B = k and V = ⊕ k, considered as a
i∈V Q σ∈AQ
k-bimodule via aσa = (ai σaj ), where σ is a path stating at i and ending at j,
that is, s(σ) = i and e(σ) = j and a, a ∈ B with components ai , aj . Therefore
corollary 2.2.13 gives the same statement for any path algebra:
Proposition 2.3.5. Let A = kQ. Then for any A-module X there is an exact
sequence:
f g
0→ ⊕ Xs(σ) ⊗ εe(σ) A −→ ⊕ Xi ⊗ εi A −→ X → 0 (2.3.1)
σ∈AQ i∈V Q
where the second sum is over all paths a with s(a) = e(σ) and the elements
xσ,a ∈ Xεs(σ) are almost all equal to 0. Then,
f (u) = (xσ,a ⊗ σa − xσ,a σ ⊗ a).
σ a
2 Because if there are no oriented cycles all vertices in a nontrivial path vi1 . . . vir must be
different.
QUIVERS AND THEIR REPRESENTATIONS 73
Let a be a longest path in u with xσ,a = 0 for some σ. Then f (u) involves xσ,a ⊗σa
and nothing can cancel this, so f (u) = 0. A contradiction (as in the proof of lemma
2.3.1).
2. We shall show that Kerg ⊆ Imf . Let u ∈ ⊕ Xi ⊗ εi A. Then one can
i∈V Q
represent it in the form
n
u= xa ⊗ a,
i=1 a
where the second sum is over all paths a starting at i and almost all the xa ∈ Xs(a)
are zero. Define deg(u) to be the length of the longest path a with xa = 0.
Now if a is a nontrivial path with s(a) = i, we can write it as a product a = σa ,
where σ is an arrow starting at i and a is some other path. Then
f (xa ⊗ a ) = xa ⊗ a − xa σ ⊗ a ,
viewing xa ⊗ a as an element of the σ-th component.
We now claim that u + Imf always contains an element of degree 0. Let
deg(u) = d > 0. Consider the element
n
ξ = u − f( xa ⊗ a ),
i=1 a
where the second sum is over all paths a staring at i, and having length equal to
d. Then deg(ξ) < d. Now the claim follows by induction.
We are now ready to prove that Kerg ⊆ Imf . Let u ∈ Kerg, and let u =
u + Imf with deg(u ) = 0. Then
n
n
0 = g(u) = g(u ) = g( xεi ⊗ εi ) = xεi ,
i=1 i=1
n
which belongs to ⊕ Xi . So each term in the final sum must be zero. Thus u = 0.
i=1
Hence u ∈ Imf .
Remark 2.3.3. The projective resolution (2.3.1) is often called the Ringel
resolution (see [Ringel, 1976]).
74 ALGEBRAS, RINGS AND MODULES
Examples 2.3.3.
1. Let G be the finite group consisting of two elements, and let k be a field.
Then we have kG kQ/(α2 ), where Q is the quiver:
α
t
α • β
⎛ ⎞
k k k
3. Let A = T3 (k) = ⎝0 k k ⎠ ⊂ M3 (k). Then A kQ, where Q is the
0 0 k
quiver:
1 2 3
• • •
Example 2.4.1.
Consider the quiver
1 2
σ
• •
For every matrix X ∈ Mn×m (k) we can define a representation VX by VX (1) = k m ,
VX (2) = k n and VX (σ) = X.
fs(σ)
Vs(σ) Ws(σ)
Vσ Wσ
fe(σ)
Ve(σ) We(σ)
Examples 2.4.2.
1. Consider the quiver considered in example 2.4.1. Suppose we have two
matrices X and Y. When are the quiver representations VX and VY isomorphic?
According to the definition, we must have invertible linear maps f1 : k n → k n and
f2 : k m → k m such that f2 X = Yf1 or equivalently f2 Xf1−1 = Y. In other words,
VX and VY are isomorphic if and only if Y can be obtained from X by changing
the basis in k n and changing the basis in k m .
This is a well known matter and X, Y ∈ Mn×m (k) are equivalent under this
equivalence relation if and only if they have the same rank.
2. Consider the one loop quiver
α
t
case the category Repk (Q) can be identified with the category of endomorphisms
of k-vector spaces.
Note that every quiver has a representation T with Ti = 0 and Tσ = 0, for all
i ∈ V Q and all σ ∈ AQ. This representation is called the zero representation.
The zero representation and a given representation V itself are called the trivial
subrepresentations of V .
Example 2.4.3.
Let Q be the quiver:
1 2
σ
• •
We have three irreducible representations: E1 : k 0 , E2 : 0 k and I:
id
k k , which correspond to the indecomposable modules: (k, 0), (0, k) and
(k, k). (Here Ei (j) = k if i = j and Ei (j) = 0 otherwise, Ei (σ) = 0.)
(V ⊕ W )i = Vi ⊕ Wi
Example 2.4.4.
Let Q be the quiver:
1 α 2
• •
β
V V1 ⊕ V2 ⊕ . . . ⊕ Vm W1 ⊕ W2 ⊕ . . . ⊕ Wn
Example 2.4.5.
Let Q be the quiver:
1 2
σ
• •
Then we have kQ = kε1 ⊕ kε2 ⊕ kσ and we have relations: σε1 = ε2 σ = σ, ε21 = ε1 ,
ε22 = ε2 , ε1 ε2 = ε2 ε1 = 0, ε1 σ = σε2 = 0.
For a simply laced quiver Q (without loops and multiply arrows) we define a
bilinear form by
n
< α, β > = αi βi − αs(σ) βe(σ) . (2.4.1)
i=1 σ∈AQ
dim HomA (V, U ) − dim Ext1A (V, U ) = < dim V, dim U > .
In particular,
0 → V1 → V2 → . . . → Vs → 0
s
(−1)i dim Vi = 0. (2.4.1)
i=1
0 → P2 → P1 → V → 0.
We apply to this exact sequence the functor HomA (∗, U ). Since P2 , P1 are projec-
tive, we obtain the following exact sequence:
n
< α, β > = αi βi − αs(σ) βe(σ) =
i=1 σ∈AQ
n
= dim ( ⊕ Hom(Vi , Ui )) − dim ( ⊕ Hom(Vs(σ) , Ue(σ) )) =
i=1 σ∈AQ
The main result in the theory of quiver representations is the famous Gabriel
theorem classifying the quivers of finite representation type. It turns out that such
quivers are closely connected with Dynkin diagrams. The next two sections are
devoted to a proof of the Gabriel theorem.
• • • ... • •
An : n≥1
1 2 3 n−1 n
1
•
Dn : • • ... • • n≥4
3 n
•
2
3
•
E6 :
• • • • •
1 2 4 5 6
3
•
E7 :
• • • • • •
1 2 4 5 6 7
3
•
E8 :
• • • • • • •
1 2 4 5 6 7 8
And the following list contains the simple Euclidean diagrams which are sim-
ply laced graphs. These diagrams are also called the extended simple Dynkin
diagrams. The corresponding simple Dynkin diagram is obtained from the ex-
tended Dynkin diagram by dropping the added vertex and associated edges.
•
n :
A n≥1
• • • ... • • •
QUIVERS AND THEIR REPRESENTATIONS 81
1 n+1
• •
n :
D • • ... • • n≥4
3 4 n−2 n−1
• •
2 n
6 :
E •
• • • • •
•
7 :
E
• • • • • • •
•
8 :
E
• • • • • • • •
Note that the diagram A 0 has one vertex and one loop, and A
1 has two vertices
joined by two edges.
Let Γ be a finite connected graph without loops with a set of vertices Γ0 =
{1, 2, . . . , n} and a set of natural numbers {tij }, where tij = tji is a number of
edges between the vertex i and the vertex j. Note that Γ may contain multiple
edges.
Define the quadratic form q : Zn → Z by
n
q(α) = α2i − tij αi αj (2.5.1)
i=1
If Q is a quiver and Γ is its underlying graph, then (·, ·) and q are the same
as before. However, the Euler bilinear form < ·, · > depends on the orientation
of Q. We say that a quadratic form q is positive definite if q(α) > 0 for all
0 = α ∈ Zn .
We say that a quadratic form q is positive semi-definite (or nonnegative
definite) if q(α) ≥ 0 for all 0 = α ∈ Zn .
Examples 2.5.1.
1. Let Γ be the graph: • •
Then we have
1 3
q(α) = q(α1 , α2 ) = α21 + α22 − α1 α2 = (α1 − α2 )2 + α22 > 0
2 4
for all α = 0. Hence q is positive definite.
2. Let Γ be the graph: • •
Then we have
for all α. Hence q is positive semi-definite, but it is not positive definite, since
q(α1 , α2 ) = 0 if α1 = α2 .
3. Let Γ be the graph: • •
Then we have
Since q(1, 1) = −1 and q(1, −1) = 1, q is not positive semi-definite (nor negative
semi-definite).
3) α ∈ rad(q).
βj 2
= tij α − tij (αi αj ) =
2βi i i<j
i=j
1 2
= (2 − 2tii )βi αi − tij (αi αj ) = α2i − tij (αi αj ) = q(α).
i
2βi i<j i i≤j
α
Hence q is positive semi-definite. If q(α) = 0, then αi = j whenever there is
βi βj
an edge connected the vertex i with the vertex j. Since Γ is connected, it follows
that α ∈ Q β. But that implies that α ∈ rad(q), since β ∈ rad(q), by assumption.
Finally, if α ∈ rad(q), then q(α) = 0. This completes the proof.
For all extended Dynkin diagrams we can give the corresponding graphs where
each vertex i is marked by the value δi , where δ = (δ1 , . . . , δn ) > 0 is strict. Here
δ is the unique vector such that radQ = Zδ, see lemma 2.5.1.
1
•
n :
A n≥1
• • • ... • • •
1 1 1 1 1 1
1 1
• •
n :
D • • ... • • n≥4
2 2 2 2
• •
1 1
84 ALGEBRAS, RINGS AND MODULES
6 :
E
2
• • • • •
1 2 3 2 1
2
•
7 :
E
• • • • • • •
1 2 3 4 3 2 1
3
•
8 :
E
• • • • • • • •
2 4 6 5 4 3 2 1
The following main theorem gives the classifications of all graphs from the
point of view of their quadratic forms.
Proof. The full proof of this theorem can be found in [Bourbaki, 1968]. We
give here only the proof of the necessary part of this theorem.
(2) Suppose that Γ is a simple Euclidean diagram. First we need to verify that
n
the vector δ = δi εi belongs to rad(q). We have (δ, εi ) = δj tij +2δi −2δi tii = 0
i=1 i=j
for each i. Then q is positive semi-definite and rad(q) = Q δ ∩ Zn , by lemma 2.5.1.
Since one of the δj ’s is 1, we have rad(q) = Q δ ∩ Zn = Zδ.
(1) Suppose Γ is a simple Dynkin diagram. We can add one more vertex to this
graph to obtain a simple Euclidean diagram Γ̃. Note that q|Γ is the restriction to
Zn of the quadratic form q|Γ̃ . By lemma 2.5.1, if q(α) = 0 for a nonzero α ∈ Zm ,
then α ∈ Zδ, hence α is strict. So q is strictly positive on all the α’s in Zn coming
from Γ, so q is positive definite on Γ.
QUIVERS AND THEIR REPRESENTATIONS 85
(3) Suppose that Γ is neither a simple Euclidean diagram nor a simple Dynkin
diagram. Then Γ has a subgraph Γ which is Euclidean, with a radical vector δ.
If all vertices of Γ are in Γ , we can take α = δ. If there is a vertex i of Γ which
is not in Γ and which is connected with Γ by any edge, we can take α = 2δ + εi .
It is now easy to check that q(α) < 0 and (α, εi ) ≤ 0 for all i.
Example 2.5.2.
Let Q be a quiver with underlying Dynkin diagram An . For all r, s ∈ N with
1 r < s n let Vr,s (i) = k for r i s and Vr,s (j) = 0 for j < r or j > s.
For σ ∈ AQ we set Vr,s (σ) = I if σ joints two points in {i : r i s} and
Vr,s (σ) = 0 otherwise. Then Vr,s is an indecomposable representation of Q and all
indecomposable representations of Q are isomorphic to one of these (see [Gabriel,
1972], [Dlab, Ringel, 1976]).
Example 2.5.3.
Consider the quiver Q5 :
1 2 3 4 5
• • • • •
•
0
with the vertices numbered as indicated. This quiver is wild. We show that any
finite dimensional k-algebra A is the algebra of all endomorphisms of some finite
dimensional representation V of the quiver Q5 . First consider the quiver Q4 ,
which is obtained from Q5 , by removing the vertex 5 and the arrow incident with
it. We now construct a representation U of Q4 with dim(U ) = 2n + 1, n = 1, 2, . . .
such that the endomorphism algebra of U is k. Let E be an n + 1-dimensional k-
vector space with a basis e1 , . . . , en+1 and let F be an n-dimensional vector space
with a basis f1 , . . . , fn . We set U (0) = E ⊕ F , U (1) = E ⊕ 0, U (2) = 0 ⊕ F ,
U (3) = {(λ(f ), f ) : f ∈ F }, U (4) = {(δ(f ), f ) : f ∈ F } , where λ, δ : F → E are
defined by λ(fi ) = ei , δ(fi ) = ei+1 . The maps associated to the arrows are the
natural inclusions. Then an endomorphism of U is given by an endomorphism α
of U (0) = E ⊕ F , which preserves the subspaces U (1), . . . , U (4). One can easily
check that this means that α is a multiplication by an element of k, i.e., one finds
End(U ) = k. Now let A be any finite dimensional algebra over k and let a1 , . . . , am
be a set of generators of A (as a k-module). Let a0 = 1 and assume that m is even,
m 2. Let U be the representation of Q4 constructed above with dim(U ) = m+1.
We now define a representation V of Q5 by V (0) = A ⊗k U (0), V (i) = A ⊗k U (i)
m
for i = 1, 2, 3, 4, V (5) = { aai ⊗ ei : a ∈ A}, where e0 , . . . , em is a basis of
i=0
U (0). An endomorphism of V is an endomorphism of V (0) which preserves the
five subspaces V (j) for j = 1, . . . , 5. Since End(U ) = k, an endomorphism of
V (0) which preserves V (1),. . .,V (4) is necessarily of the form φ ⊗ 1 where φ is a
86 ALGEBRAS, RINGS AND MODULES
m
m
k-vector space endomorphism of A. Now (φ ⊗ 1)( aai ⊗ ei ) = φ(aai ) ⊗ ei .
i=0 i=0
Therefore if φ ⊗ 1 also preserves V (5), then for all a ∈ A there exists b(a) such that
m m
φ(aai ) ⊗ ei = b(a)ai . Now 1 ⊗ e0 , . . . , 1 ⊗ em is a basis for A ⊗k U (0) as a
i=0 i=0
module over A, hence φ(aai ) = b(a)ai for all i. Taking i = 0 we find φ(a) = b(a).
Hence we have that φ(aai ) = φ(a)ai for all a ∈ A and all i .
Let c = φ(1), then φ(ai ) = cai for all i. So φ is given by multiplication by
c ∈ A. This shows that End(V ) = A.
Δ = {0 = α ∈ Zn : q(α) ≤ 1}
is called the set of roots of the quadratic form q of Γ. A root α is called real if
q(α) = 1 and it is called imaginary if q(α) = 0.
Note that each εi is a root. These roots are called the simple roots. If Γ is
a simple Dynkin diagram, then there are no imaginary roots for this graph. If Γ
is a Euclidean diagram, then its imaginary roots are the integer vectors which are
multiples of δ, by lemma 2.5.1.
Lemma 2.5.3.
1. If α ∈ Δ and β ∈ rad(q), then −α and α + β are roots.
2. Every root is positive or negative.
3. If Γ is a simple Euclidean diagram, then (Δ ∪ {0})/Zδ is a finite subset of
Zm /Zδ.
4. If Γ is a simple Dynkin diagram, then Δ is finite.
Proof.
1. We have q(β ± α) = q(β) + q(α) ± (β, α) = q(α) ≤ 1.
2. Write α = α+ − α− , where α+ , α− ≥ 0 and have disjoint support. Then
obviously, (α+ , α− ) ≤ 0, so
Since q is positive semi-definite, one of q(α+ ) or q(α− ) must be zero, i.e., one of
them is an imaginary root. Since each imaginary root is strict, the other of these
roots must be zero.
3. Choose an i with δi = 1. If α is a root with αi = 0, then δ − α and δ + α are
also roots by the condition (1). Since their i-th component is 1, they are positive
roots, so −δ ≤ α ≤ δ. Therefore
S = {α ∈ Δ : αi = 0} ⊆ {α ∈ Zn : −δ ≤ α ≤ δ}
In the general case when Q is not necessarily a simply laced quiver the right
corresponding quadratic form associated to the quiver is defined differently.
All Dynkin diagrams (in addition to the simple Dynkin diagrams drawn
above) are represented in the form of valued graphs by the following list:
(1,2)
Bn : • • • ... • • • n≥2
1 2 3 n−2 n−1 n
(2,1)
Cn : • • • ... • • • n≥3
1 2 3 n−2 n−1 n
(1,2)
F4 : • • • •
1 2 3 4
4 Associate to the valued graph (Γ, d) the generalized Cartan matrix A = (a ), a
ij ii = 2,
aij = −dij if i and j are connected by an edge (and are different) and aij = 0 if there is no edge
between i and j. Then condition 3 is equivalent to symmetrizability (on the right) of the GCM
A meaning that there is a diagonal integer matrix F such that AF is symmetric. As is easily
proved this is equivalent to symmetrizability on the left, meaning that there is a diagonal integer
matrix F such that F A is symmetric.
88 ALGEBRAS, RINGS AND MODULES
(1,3)
G2 : • •
1 2
And the valued graphs in the following list are all the Euclidean diagrams
(or extended Dynkin diagrams)5 , in addition to the simple Euclidean diagrams
which were presented above:
(1,2) (2,1)
n :
B • • • ... • • • n≥2
1 2 3 n−1 n n+1
(2,1) (1,2)
n :
C • • • ... • • • n≥3
1 2 3 n−1 n n+1
11 :
A (1,4)
• •
12 :
A (2,2)
• •
(1,2) (1,2)
n :
BC • • • ... • • • n≥2
1 2 3 n−1 n n+1
n+1
•
n : (1,2)
BD • • • ... • •
1 2 3 n−2 n−1
•
n
n+1
•
n : (2,1)
CD • • • ... • •
1 2 3 n−2 n−1
•
n
5 It is worth noting that the extra vertex in an extended Dynkin diagram is so placed that
omitting any vertex (and the edges involving this vertex) leaves as residue a disjoint union of
Dynkin diagrams; see e.g. [Wan, 1991] page 35.
QUIVERS AND THEIR REPRESENTATIONS 89
F41 : • • •
(1,2)
• •
F42 : • • •
(2,1)
• •
21 :
G (1,3)
• • •
22 :
G (3,1)
• • •
These are the notations from [Dlab-Ringel, 1976]. In books on Kac-Moody alge-
bras and quantum groups such as [Kac, 1985], [Wan, 1991], [Hong-Kang, 2002] a
l = A(1) , BC
different notation is used; e.g. A 2l = A(2) , CD
2l−1 = A(2)
l 2l 2l−1
For all these graphs here are the corresponding graphs where each vertex i is
marked by the value δi , where the strict vector δ = (δ1 , . . . , δn ) > 0 generates the
radical.
n : • • • ... • • •
B n≥2
1 1 1 1 1 1
n : • • • ... • • •
C n≥3
1 2 2 2 2 1
11 : • •
A
2 1
12 : • •
A
1 1
n : • • • ... • • 1
BC n≥2
2 2 2 2 2 1
1
•
n :
B • • • ... • •
2 2 2 2 2
•
1
90 ALGEBRAS, RINGS AND MODULES
1
•
n :
CD • • • ... • •
1 2 2 2 2
•
1
• • • • •
F41 :
1 2 3 2 1
• • • • •
F42 :
1 2 3 4 2
21 : • • •
G
1 2 1
22 : • • •
G
1 2 3
Examples 2.5.4.
1. Let (Γ, d) be the valued graph:
(1,2)
B3 : • • •
1 2 3
For this graph d12 = 1, d21 = 2, d23 = d32 = 1. Therefore from the requirement
dij fj = dji fi we obtain that f1 = 1, f2 = f3 = 2. Then the quadratic form has
the following form:
(2,1)
C3 : • • •
1 2 3
For this graph d12 = 2, d21 = 1, d23 = d32 = 1. Therefore from requirement
dij fj = dji fi we obtain that f1 = 2, f2 = f3 = 1. Then the quadratic form has
the following form:
1 1 1
q(x) = 2x21 + x22 + x23 − 2x1 x2 − x2 x3 = 2(x1 − x2 )2 + ( x2 − x3 )2 + x22
2 2 4
and so it is positive definite.
11 :
3. Let (Γ, d) be the valued graph: A (1,4)
• •
For this graph d12 = 1, d21 = 4. Therefore from the requirement dij fj = dji fi
we obtain that f1 = 4, f2 = 1. Then the quadratic form has the following form:
The following theorem gives the characterization of all valued graphs from the
point of view their quadratic forms:
The full proof of this theorem can be found in [Dlab, Ringel, 1976].
Remark 2.5.1. The non-symmetric bilinear form q̃ given by (2.5.3) has the
following important property for the representations of a quiver Q (in the case of
Dynkin diagrams):
(dij ,dji )
• •
i j
we connect the vertex i with the vertex j by tij edges, where tij satisfies the
following conditions. If dij dji ≤ 4 then tij = max{dij , dji }. In this case and when
dij < dji , all these tij edges are marked together by one an arrow in the direction
from j to i8 . For example,
(3,1)
• • is drawn as • •
1 2 1 2
and
• • •
B3 :
1 2 3
If dij = dji = 2, we draw this graph as follows: Ã12 : • •
Conversely, each such a graph we can transform into a valued graph in the
following way. If we have a marked multiply arrow with tij edges in the direction
from j to i then we put dij = 1 and dji = tij . Otherwise we put dij = dji = tij .
8 This is the convention that makes things come out right when one starts with the lists from
[Dlab-Ringel, 1976]. In [Wan, 1991], [Kac, 1985] and [Hong-Kang, 2002] and other books on
Kac-Moody algebras and quantum groups another convention is used; viz. there is an arrow
towards i iff dij ≥ 2 Applied to the list above this switches the B and C diagrams.
QUIVERS AND THEIR REPRESENTATIONS 93
Remark 2.5.3. It is often convenient to describe a finite valued graph (Γ, d),
where Γ0 = {1, 2, . . . , n} in the terms of an associated matrix C = (cij ), called the
Cartan matrix. By definition, cii = 2, and cij = −dij . For example, if (Γ, d) is
(1,2)
F4 : • • • •
1 2 3 4
then the Cartan matrix of F4 is of the form
⎛ ⎞
2 −1 0 0
⎜−1 2 −2 0 ⎟
⎜ ⎟
⎝ 0 −1 2 −1⎠
0 0 −1 2
We shall prove this theorem for the case when k is an algebraically closed field.
To prove this theorem we need a little bit of algebraic geometry.
Let AN be affine N -space, whose points are just the N -tuples from the
field k. The coordinate ring k[AN ] of functions on AN is the polynomial ring
k[x1 , x2 , . . . , xn ], where xi is the i-th coordinate function. This space is endowed
with the Zariski topology, whose set of closed subsets coincides with the sets
of common zeros of the ideals in k[x1 , x2 , . . . , xn ]. A closed set in AN is called
an affine variety. A subset U in AN is called locally closed if it is open in
its closure U. A non-empty locally closed subset U is called irreducible if every
non-empty open subset of U is dense in U . The affine space AN itself is irreducible.
9 See [Gabriel, 1972], [Berstein, Gel’fand, Ponomarev, 1973].
94 ALGEBRAS, RINGS AND MODULES
Example 2.6.1.
The linear group GLn (k) of all n × n invertible matrices over k is an algebraic
group. This group is defined by the non-vanishing of the determinant and so it is
2
open in Mn (k) An and is an irreducible affine variety of dimension n2 . To see
2
this consider An +1 with coordinates xij , i, j = 1, . . . , n and z and consider the
zeros of the polynomial z det(xij ) − 1. These are in obvious bijection with GLn (k).
Lemma 2.6.2.
1. There is a bijection V → OV between the set of isomorphism classes of
indecomposable representations V of dimension α and the set of GL(α)-orbits on
Rep (α), where OV = {x ∈ Rep (α) : R(x) V }.
2. GL(α)x AutA (R(x)).
dim Rep (α) − dim OV = dim EndA (V ) − q(α) = dim Ext1A (V, V ).
Corollary 2.6.4. If α = 0 and q(α) < 0, then there are infinitely many orbits
in Rep (α), i.e., infinitely many non-isomorphic representations of dimension α.
Proof. It is known that any two non-empty open subvarieties inside an irre-
ducible variety intersect. Suppose OX = OY are open. Since orbits are disjoint,
OX ⊂ Rep (α)\OY , and so OX ⊂ Rep (α)\OY , which contradicts the irreducibility
of Rep (α).
Lemma 2.6.7. If
0→U →V →W →0 (2.6.1)
is a non-split exact sequence, then OU⊕W ⊂ OV \ OV .
To prove the sufficiency part of the Gabriel theorem we need the following
lemma.
0→K→V →I→0
0 → Kj → Y → I → 0.
If this splits, then Kj has a complement C in Y . But then the inverse image of C
is a complement to Kj in V , hence Kj is a summand of V . This contradicts the
assumption that V is indecomposable.
Proof. Apply lemma 2.6.10. If EndU = k, repeat. Since all modules are finite
dimensional, the process must terminate.
a contradiction.
Remark 2.6.1. The Gabriel theorem has also been proved in the more gen-
eral case when k is an arbitrary field. In this case I.N.Bernstein, I.M.Gel’fand,
V.A.Ponomarev proved that if Q is a Dynkin diagram then it has finite repre-
sentation type (see, [Berstein, Gel’fand, Ponomarev, 1973]). Their proof uses the
Weyl group and Coxeter functors.
Theorem 2.6.12. A quiver Q is of finite (resp. tame) type if and only its
quadratic form is positive definite (resp. semipositive definite).
In the general case, if k is an arbitrary field, the result was treated by V.Dlab
and C.M.Ringel (see [Dlab, Ringel, 1976]).
2.7 K-SPECIES
In this section we consider the notion of k-species introduced by P.Gabriel (see
[Gabriel, 1973]). Let k be a fixed field and let I be a finite set. A species
L = (Ki , i Mj )i,j∈I is a finite family (Ki )i∈I of skew fields together with a family
(i Mj )i,j∈I of (Ki , Kj )-bimodules. We say that (Ki , i Mj )i,j∈I is a k-species if
all Ki are finite dimensional and central over a common commutative subfield k
which acts centrally on i Mj , i.e., λm = mλ for all λ ∈ k and all m ∈ i Mj . We
also assume that each bimodule i Mj is a finite dimensional vector space over k.
It is a k-quiver if moreover Ki = k for each i. We shall consider species without
oriented cycles and loops, i.e., we shall consider the case where i Mi = 0, and if
i Mj = 0, then j Mi = 0.
The diagram of a species (Ki , i Mj )i,j∈I is defined in the following way:
1) the set of vertices is the finite set I;
2) the vertex j connects with the vertex i by tij arrows, where
If i Mj = 0 and dimKj (j Mi ) > dim(j Mi )Ki , then we denote this by the following
• •
arrow: (here tij = 3). (The number of horizontal lines in the “arrows”
j i
is tij .)
j ϕi : Vi ⊗Ki i Mj −→ Vj
j ψi
Wi ⊗Ki i Mj Wj
100 ALGEBRAS, RINGS AND MODULES
n+gl.dimR. In the paper [Hochschild, 1958] G.Hochschild proved that the global
dimension of a free ring over R on any set of letters is equal to 1+gl.dimR. This
theorem was generalized by Yu.V.Roganov for a tensor algebra of a bimodule which
is one-sided projective (see [Roganov, 1975]). In the proof of theorem 2.2.11 we
follow this paper.
In 1972 in the paper [Gabriel, 1972] P.Gabriel introduced quivers in connection
with the classification of finite dimensional algebras of finite type. In this paper he
gave a full description of quivers of finite representation type over an algebraically
closed field (theorem 2.6.1). P.Gabriel also proved that there is a bijection between
the isomorphism classes of indecomposable representations of a quiver Q and the
set of positive roots of the Tits form corresponding to this quiver.
Another proof of this theorem in the general case, for an arbitrary field, using
reflection functors and Coxeter functors has been given in the paper [Berstein,
Gel’fand, Ponomarev, 1973]. This paper also contains the connection between
indecomposable representations of a quiver of finite type and properties of the
Tits quadratic form (theorem 2.5.2).
The Gabriel theorem was generalized by V.Kac (see [Kac, 1980a], [Kac, 1980b]),
who proved that indecomposable representations occur in dimension vectors that
are roots of the so-called Kac-Moody Lie algebra associated to a given graph. In
particular, these dimension vectors do not depend on the orientation of the arrows
in the quiver Q.
The terms “tame type” and “wild type” were introduced by P.Donovan and
M.R.Freislich in their paper [Donovan, Freislich, 1972] in analogy with the separa-
tion of animals into tame and wild ones. In the same paper they first conjectured
that any finite dimensional algebra is either tame or wild.
Tame quivers in terms of extended Dynkin diagrams were classified by
L.A.Nazarova in [Nazarova, 1973] and by P.Donovan and M.R.Freislich in
[Donovan, Freislich, 1973].
The representation type of factor algebras of path algebras has been discussed
by E.Green, who gave some useful algorithms for studying representations of path
algebras (see [Green, 1975]).
The theory of species was first considered by P.Gabriel in [Gabriel, 1973]. Note
that in fact species and their connections with the representations of algebras were
already considered in the papers of T.Yoshii (see [Yoshii, 1956]; [Yoshii, 1957a];
[Yoshii, 1957b]. Later the results of P.Gabriel on representations of species were
generalized by V.Dlab and C.M.Ringel (see [Dlab, Ringel, 1974]; [Dlab, Ringel,
1975]; [Dlab, Ringel, 1976]; [Ringel, 1976].
Representations of quivers with relations were considered by P.Donovan
and M.Freislich [Donovan, Freislich, 1979], S.Ovsienko [Ovsienko, 1977],
C-M.Ringel [Ringel, 1975], [Ringel, 1980], V.Yu.Romanovskij, A.S.Shkabara, and
A.G.Zavadskij, (see [Romanovskij, Shkabara, 1976]; [Shkabara, 1978a], [Shkabara,
1978b]; [Zavadskij, Shkabara, 1976].
102 ALGEBRAS, RINGS AND MODULES
Recall (following [1]) that a valued graph is a graph with each edge marked by an
ordered pair of positive integers. In loc.cit. no loops are allowed, but actually for
the tame and finite representation type cases this makes little difference: just the
inclusion of one tame case, viz. the one vertex one loop graph.
By convention for an edge with value (1,1) the value indicator is omitted. Thus
for example one has the valued diagram
(1,2)
B4 : • • • •
1 2 3 4
which is in fact the diagram denoted by B4 in [1] and, using a convention explained
below, C4 in [12], [18] and [19] (and most other recent books and papers treating
of these matters (especially those dealing with Kac-Moody algebras and quantum
groups).
There is a further convention, induced, we feel, by both a sense of historical
continuity and nostalgia, that replaces labels with low numbers by an arrow no-
tation. It runs as follows. If the label (dij , dji ) of the edge between vertex i and
vertex j is such that dij dji ≤ 4 then there are max{(dij , dji )} edges between i and
j and there is an arrow towards i (resp. j) if and only if dij > 1 (resp. dji > 1).
Using this convention the valued graph above is denoted1
• • • •
C4 :
1 2 3 4
Here are some further examples
(2,2) • •
• • =
(3,1) = • •
• •
Edges with a label (dij , dji ) with dij dji > 4 are left as is. Note that from a dia-
gram with arrow notation the corresponding labeled graph is uniquely recoverable.
are said to be equivalent if one can be obtained from the other by simultaneous
permutation of rows and columns.
Here for a vector u = (u1 , u2 , . . . , un ), u > 0 means ui > 0 for all i and u ≥ 0
means ui ≥ 0 for all i.
A really deep and fascinating theory with applications to both other areas of
mathematics (e.g. the monstrous moonshine conjectures, algebraic combinatorics,
...) and physics (e.g. quantum field theory, string theory) has only been developed
(so far) for the integral case with moreover aii = 2 for all i.
There is a natural and rather obvious bijection between valued graphs and gener-
alized Cartan matrices. To a valued graph with edge labels (dij , dji ) associate the
generalized Cartan matrix A = aij with aii = 2, aij = −dij if i and j are connected
and aij = 0 if i and j are not connected. In this correspondence connectedness of
the valued graph is the same as indecomposability of the matrix.
APPENDIX. MORE ON DYNKIN DIAGRAMS 107
As it turns out the finite type generalized Cartan matrices A have as associated
Lie algebras the classical simple finite dimensional Lie algebras with the Dynkin
diagrams An , Bn , Cn , Dn , E6 , E7 , E8 , F4 , G2 as depicted just below lemma 2.5.1.
Nothing much new here. The affine case gives the infinite dimensional affine (or
Eyclidean) Lie algebras introduced independently and practically simultaneously
by Kac and Moody [11, 14, 15]. These are extraordinary rich in theory and
applications.
The corresponding extended Dynkin diagrams, also called Eyclidean diagrams
or affine diagrams, are listed in section 2.5 above just below the list of Dynkin
diagrams.
It may seem at first sight remarkable that finiteness of the Lie algebra asso-
ciated to a generalized Cartan matrix should correspond exactly to finite repre-
sentation type of the corresponding quiver, and that affineness of the Lie algebra
associated to a generalized Cartan matrix should correspond precisely to tame
representation type of the associated quiver 2
However, admitting some fairly deep theory in the two cases this is entirely
clear.
Whether a connected valued graph (quiver) Γ is finite, tame, wild represen-
tation type is ruled by the Dlab-Ringel-Ovsienko-Roiter quadratic form
of it. This quadratic form is defined as follows. It is assumed that the col-
lections of integers dij from the labels is (right) symmetrizable; meaning that
there are positive integers fi such that dij fj = dji fi . This is the same as saying
that the generalized Cartan matrix associated to the valued graph is (right) sym-
metrizable (which in turn is the same as being left symmetrizable (as is easy
to prove)). This symmetrizability condition is not much of a restriction be-
cause a generalized Cartan matrix of finite or affine type is always symmetrizable,
see [12, 18].
Then the Dlab-Ringel-Ovsienko-Roiter quadratic form of Γ is
n
qΓ (x) = 2fi x2i − dij fj xi xj
i=1 i=j
i.e. it is the quadratic form associated to the symmetric matrix C(Γ)F , where
C(Γ) is the generalized Cartan matrix associated to Γ and F is the diagonal
matrix F = diag(f1 , f2 , . . . , fn ). In terms of this quadratic form a quiver Γ is of
finite, tame, or wild representation type according to whether the quadratic form
is positive definite, positive semi-definite, or indefinite 3 4 .
2 The
diagrams in the lists correspond bijectively. However, the labels assigned to them in [1]
22 in [1] correspond respectively to A(2) , E (2) ,
n , F42 , G
and [12, 18, 19] differ. For instance, BC 6
(3)
D4 in [12, 18, 19]. It is the latter notation that is now mostly used.
3 It
cannot be negative semi-definite because the diagonal elements are positive.
4 Recall that the quadratic form of a symmetric matrix is positive definite (resp. positive
semi-definite) if and only if all principal minors of the matrix are positive (resp. all proper
principal minors are positive and matrix is singular). This does not necessarily hold for matrices
108 ALGEBRAS, RINGS AND MODULES
n
qΓ (x) = (2 − 2dii )fi x2i − dij fj xi xj
i=1 i=j
and this adds just one more semi-positive definite case, viz the graph of one vertex
with one loop which can be appropriately denoted A 0 (in the kind of notation
from [1]).
There is, or has been, a sort of rueful pessimism about doing something in
the case of wild quivers and wild representation type in general. This is not quite
justified as is evidenced for example by the remarkable results of Kac on dimensions
vectors, very nicely explained in [13].
Also, on a much more modest level, a great deal can sometimes be said in
specific cases. For instance the quivers
t - t
t -
t
t -
t - t
and their representations are of great importance in linear control and systems
theory and a great deal is known about (the moduli space of) their representations,
see [8,10].
⎛ ⎞
2 −1 0 0
⎜ −2 2 −1 0 ⎟
that are not symmetric. Consider for instance the matrix ⎜
⎝ 0
⎟ which is the
−1 2 −2⎠
0 0 −1 2
(1)
semi-positive definite Cartan matrix of C3 (in the notation of [12]). The quadratic form of this
matrix is 2x21 + 2x22 + 2x23 + 2x24 − 3x1 x2 − 2x2 x3 − 3x2 x4 which is indefinite.
APPENDIX. MORE ON DYNKIN DIAGRAMS 109
Let S be a finite set and let M = (mst )s,t∈S be a symmetric matrix such that
mss = 1 and mst ∈ {2, 3, 4, . . .}∪{∞} for s = t. Such a matrix is called a Coxeter
matrix. Now let W be a group with S as a subset. Then the pair (W, S) is called
a Coxeter system and W is called a Coxeter group if W has a presentation
with elements from S as generators and the relations
An , Bn , Dn , E6 , E7 , E8 , F4 , H3 , H4 , I2 (m) (2)
Most of these groups 6 turn up as the Weyl groups of the simple Lie algebras
with the corresponding labels. The correspondence between the Cartan matrix
defining the simple Lie algebra and the Coxeter matrix is according to the following
table
ast ats 0 1 2 3
mst 2 3 4 6
Note that G2 = I2 (6). Note also that the simple Lie algebras Bn , Cn give
the same (abstract) Coxeter group explaining that Cn is missing from the list (2)
above.
Similarly the extended Dynkin diagram give rise to affine Weyl groups which
are infinite Coxeter groups. This time there are no extra diagrams like the I2 (m),
m = 6 and H3 , H4 in the list (2).
See [5] and references quoted there for a great deal more on all this. The
subject of reflection groups is a very large one with applications to and/or links
with many parts of mathematics.
There are many more areas where Dynkin diagrams turn up, e.g. singularity
theory. This happens so often that V.I.Arnol’d posed it as a Hilbert type problem
(problem VIII in [16]). Here is the precise formulation:
The A-D-E classifications. The Coxeter-Dynkin graphs Ak , Dk , Ek appear in
many independent classification theorems. For instance
(a) classification of Platonic solids (or finite orthogonal groups) in Eyclidean
3-space.
5 As stated they are just elements of order 2 (involutions). It turns out it is indeed possible
(b) classification of the categories of linear spaces and maps7 ; Gabriel, Gel’fand-
Ponomarev, Roiter-Nazarova, see Séminaire Bourbaki, exposé 444, 1974.
(c) classification of the singularities of algebraic hypersurfaces with a definite
intersection form of the neighboring smooth fibre, Tjurina.
(d) classification of the critical points of functions having no moduli (see
Séminaire Bourbaki, exposé 443, 1974).
(e) classification of the Coxeter groups generated by reflections, or of Weyl
groups with roots of equal length.
The problem is to find the common origin of all the A − D − E classification
theorems and to substitute a priori proofs for the a posteriori verification of the
parallelism of the classifications.
The list of Arnol’d by no means exhaust the areas of mathematics where Dynkin
diagrams turn up. Some others areas are represented by the papers [2,3,20]. Here
one also finds some more different notations for Dynkin diagrams, both symbol-
ically and pictorially. Perhaps the most surprising of these papers (to us) is [3].
The numbers game is a one person game played on a finite simple graph with cer-
tain “amplitudes” assigned to its edges and an initial assignment of real numbers
to its nodes. The moves use the amplitudes to modify the node numbers. The
paper shows that if certain finiteness requirements are to be met one again finds
the Dynkin diagrams.
References
[8] Michiel Hazewinkel, (Fine) moduli spaces, for linear systems: what they are
and what they are good for. In: C.Byrnes and C.F.Martin (ed.), Geometric
methods for the theory of linear systems, Reidel, 1980, 125-193.
[9] M.Hazewinkel, W.Hesseling, D.Siersma, F.D.Veldkamp, The ubiquity of
Coxeter-Dynkin diagrams. An introduction to the A-D-E problem, Nieuw
Archief voor Wiskunde (3) 25 (1977), 257-307.
[10] Michiel Hazewinkel, Clyde F.Martin, Representations of the symmetric
groups, the specialization order, Schubert cells, and systems, Enseignement
Math. 29 (1983), 53-87.
[11] V.G.Kac, Simple irreducible graded Lie algebras of finite grouth, Izv. Acad.
Nauk SSSR (ser. mat) 32 (1968), 1923-1967.
[12] Victor G.Kac, Infinite dimensional Lie algebras, Cambridge Univ. Press,
1985.
[13] H.Kraft, Chr.Riedtmann, Geometry of representations of quivers. In:
P.Webb(ed.), Representations of algebras. Proceedings of the 1985 Durham
symposium, Cambridge Univ. Press, 1986, 109-145.
[14] R.V.Moody, A new class of Lie algebras, J. of Algebra 10(1968), 211-230.
[15] R.V.Moody, Euclidian Lie algebras, Canad. J. Math. 21(1969), 1432-1454.
[16] Problems of present day mathematics. In: F.E.Browder (ed.) Mathemati-
cal developments arising from Hilbert’s problems, American Mathematical
Society, 1976, 35-80.
[17] E.B.Vinberg, Discrete linear groups generated by reflections, Izv. Acad. Nauk
SSSR (ser.mat.) 35(1971), 1072-1112.
[18] Minoru Wakimoto, Lectures on infinite dimensional Lie algebra, World Sci-
entific, 2001
[19] Zhe-Xian Wan, Introduction to Kac-Moody algebra, World Scientific, 1991.
[20] Eric Zaslow, Dynkin diagrams of CP 1 orbifolds, arXiv: hep-th/9304130.
3. Representations of posets and of finite dimensional
algebras
In the theory of representations one tries to study a given object by means of
homomorphisms to another object which is in some way more concrete and easier
to understand. Such objects in the theory of finite dimensional algebras are endo-
morphisms of some finite dimensional vector space over a field k. Another, slightly
different, but related, point of view is that one attempts to realize an abstract ob-
ject in terms of more concrete things such as matrices. Whence the terminology
“representation”.
There exist different approaches to the representation theory of algebras. In
this chapter we consider only the part of representation theory connected with the
representations of partially ordered sets and Gabriel quivers. This chapter is more
of an informative character and it may be considered as a brief survey of those
well-known results of this theory which will be needed in this book. Therefore
most statements are made here for completeness without proofs.
Representations of finite partially ordered sets (posets, in short) play an im-
portant role in representation theory. They were first introduced by L.A.Nazarova
and A.V.Roiter. The first two sections of this chapter are devoted to partially
ordered sets and their representations. Here there are given the main results
of M.M.Kleiner on representations of posets of finite type and the results of
L.A.Nazarova on representations of posets of infinite type. The most important
result in this theory was been obtained by Yu.A.Drozd who showed that there is
a trichotomy between finite, tame and wild representation types for finite posets
over an algebraically closed field.
One of the main problems of representation theory is to obtain information
about the possible structure of indecomposable modules and to describe the iso-
morphism classes of all indecomposable modules. By the famous trichotomy the-
orem for finite dimensional algebras over an algebraically closed field, obtained by
Yu.A.Drozd, all such algebras are divided into three disjoint classes.
The main results on representations of finitely dimensional algebras are given
in section 3.4. Here we give the structure theorems for some special classes of finite
dimensional algebras of finite type, such as hereditary algebras and algebras with
zero square radical, obtained by P.Gabriel in terms of Dynkin diagrams. Section
3.5 is devoted to the first Brauer-Thrall conjecture, settled by A.V.Roiter for the
case of a finite dimensional algebra over an arbitrary field.
Unless otherwise stated, in this chapter we always suppose that all algebras
considered are associative finite dimensional with 1.
113
114 ALGEBRAS, RINGS AND MODULES
i ≺ j ⇒ i < j.
Definition. Let P = (S, ), where S = {1, 2, ..., m}, be a finite partially
ordered set (or poset, in short). A representation of P over a field k or a
P-space is a set of finite dimensional k-spaces V = (V0 ; Vi : i ∈ S) such that
Vi ⊂ V0 for all i ∈ S and Vi ⊂ Vj precisely when i j in P. Such an object
is called a P-space. Let V = (V0 ; Vi : i ∈ S) and W = (W0 ; Wi : i ∈ S) be
two P-spaces. A morphism f : V → W is a k-linear transformation V0 → W0
such that f (Vi ) ⊂ Wi for all i ∈ S. The direct sum of V and W is V ⊕ W with
(V ⊕ W )0 = V0 ⊕ W0 and (V ⊕ W )i = Vi ⊕ Wi for all i ∈ S. A nonzero P-space is
said to be indecomposable if it cannot be written as a direct sum of two nonzero
P-spaces.
Therefore P-spaces over a fixed field k form an additive category, which is
denoted by Rep (P, k). Every P-space is a uniquely determined as a direct sum of
indecomposable P-spaces, by a corollary of the Krull-Schmidt theorem (see, vol.I,
p. 242). This theorem can be written in the following form:
From this theorem it follows that all representations in Rep (P, k) are uniquely
determined by the indecomposable ones. The set of all isomorphism classes of
indecomposable representations in Rep (P, k) shall be written as Ind (P, k).
Definition. Let P = (S, ) be a finite poset, and let S = {1, 2, ..., m}. A
representation of P over a field k is an arbitrary matrix
REPRESENTATIONS OF POSETS AND ALGEBRAS 115
A= A1 A2 ... Am
B= B1 B2 ... Bm
Notice that by a sequence of transformations (b) and (c) we can add an arbi-
trary linear combination of columns of Ai to a column of Aj if i ≺ j in P.
..
A1 O A2 O . Am O
A ⊕ B=
..
O B1 O B2 . O Bm
In order to visualize a poset S we shall use its Hasse diagram: the elements of
S will be represented by points on the plane and the relation ≺ is always thought
of as going upwards along the edges drawn. Moreover, we join the point i with the
point j if and only if i ≺ j and there is no t ∈ S such that i ≺ t ≺ j. For example,
for the poset P = (S, ), with S = {a1 , a2 , a3 } and the sole relation a2 ≺ a3 we
have the picture:
• a3
a1 • • a2
(1, 1, 1, 1) : • • • •
• •
• • •
(2, 2, 2) : (1, 3, 3) : • •
• • •
• • •
•
• • •
•
• • •
(N, 4) : (1, 2, 5) : •
•
• •
•
• • •
REPRESENTATIONS OF POSETS AND ALGEBRAS 117
Here the n-element chain (1 ≺ 2 ≺ ... ≺ n) is denoted (n), and the disjoint
union (direct sum) of (n1 ), . . . , (ns ) is written (n1 , . . . , ns ). The symbol N denotes
the following 4-element poset {a1 ≺ a2 a3 ≺ a4 }, and (N, n) stands for the
disjoint union of N and the chain (n).
The posets in the list of the Kleiner theorem will be called the critical sub-
posets.
A characterization of all exact representations of a finite-representation poset
was obtained by M.M.Kleiner [Kleiner, 1972b]. We present here a complete list
of all 41 indecomposable exact matrix representations of Kleiner’s list with some
corrections. As was mentioned in the paper [Arnold, Richman, 1992] five represen-
tations from Kleiner’s list: (IX3 ), (IX9 ), (X2 ), (X8 ), and (XI) are decomposable.
We replaced these representations by the right indecomposable ones, and the nu-
meration remains unchanged.
Theorem 3.1.3 (M.Kleiner). A finite partially ordered set P of finite type
is exact if and only if it has one of the following forms:
• c1 a1 • • c1
a a b a b c
I ; II ; III ; IV a b ;V b ;
• • • • • •
• • • c2 a 2 • • • c2
a 1 b 1 c1 a1 • • c1 a1 • • c1
• • •
VI ; VII • c2 ; VIII a2 • • c2 ;
• • • b b
a 2 b 2 c2 a 2 • • • c3 • a3 • • c3
• c1
• c1
• c2
IX a1 • • b1 • c2 ; X ;
a1 • • c3
a2 • • b2 • c3 b
a2 • • • c4
a1 • • b 1 • c1
a1 • • c1
a2 • • b 2 • c2
XI a2 • • b1 • c2 ; XII ;
• c3
a3 • • b2 • c3
• c4
118 ALGEBRAS, RINGS AND MODULES
• c1
• c2
XIII .
b1 • • a 1 • c3
b2 • • a 2 • c4
If P = IV, then
1 1 1 0
A = ,B = , C1 = , C2 = . (IV)
0 1 0 1
If P = V, then
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 1 0 0
A1 = ⎝1⎠ , A2 = ⎝ 0 ⎠ B = ⎝1⎠ , C1 = ⎝ 1 ⎠ C2 = ⎝ 0 ⎠ (V1 )
0 0, 1 0, 1;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 1 1 0 0
A1 = ⎝ 1 ⎠ A2 = ⎝ 0 ⎠ B = ⎝0 1 ⎠ C1 = ⎝ 1 ⎠ C2 = ⎝ 0 ⎠ (V2 )
0, 0, 1 0, 0, 1;
0 1 1 1 0
A1 = A2 = ,B = C1 = C2 = (V3 )
1, 0 1, 0, 1.
If P = VI, then
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 0
A1 = ⎝ 1 ⎠ A2 = ⎝ 1 ⎠ B1 = ⎝ 1 ⎠ (VI)
0, 1, 0,
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0
B2 = ⎝ 0 ⎠ C1 = ⎝ 1 ⎠ C2 = ⎝ 0 ⎠
0, 0, 1.
REPRESENTATIONS OF POSETS AND ALGEBRAS 119
If P = VII, then
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 0 0 1
⎜1⎟ ⎜0 0 ⎟ ⎜1 1⎟
⎜ ⎟
A1 = ⎝ ⎠ A2 = ⎝ ⎜ ⎟ B = ⎝⎜ ⎟ (VII1 )
0 0 0⎠ 0 1⎠
0, 0 1, 1 0,
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠ C2 = ⎝ 0 ⎠ C3 = ⎝ 1 ⎠
0, 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 1 1
⎜0 1 ⎟ ⎜0⎟ ⎜1 0⎟
A1 = ⎜ ⎟ ⎜ ⎟
⎝0 0 ⎠ A2 = ⎝ 0 ⎠ B = ⎝0
⎜ ⎟, (VII2 )
1⎠
0 0, 1, 1 0
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠ C2 = ⎝ 0 ⎠ C3 = ⎝ 1 ⎠
0, 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 1
⎜0⎟ ⎜0⎟ ⎜1 1⎟
A1 = ⎜ ⎟ ⎜ ⎟ ⎜
⎝ 0 ⎠ A2 = ⎝ 0 ⎠ B = ⎝0 1⎠,
⎟ (VII3 )
0, 1, 1 0
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠ C2 = ⎝ 0 ⎠ C3 = ⎝ 1 ⎠
0, 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 1
A1 = ⎝ 1 ⎠ A2 = ⎝ 0 ⎠ B = ⎝1⎠, (VII4 )
0, 0, 1
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0
C1 = ⎝ 0 ⎠ C2 = ⎝ 1 ⎠ C3 = ⎝ 0 ⎠
0, 0, 1;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 1 1
A1 = ⎝ 1 ⎠ A2 = ⎝ 0 ⎠ B = ⎝0 1 ⎠ (VII5 )
0, 0, 1 0,
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0
C1 = ⎝ 0 ⎠ C2 = ⎝ 1 ⎠ C3 = ⎝ 0 ⎠
0, 0, 1.
120 ALGEBRAS, RINGS AND MODULES
If P = VIII, then
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 0 0 1
⎜1⎟ ⎜0⎟ ⎜0⎟ ⎜1 1 ⎟
A1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜
⎝ 0 ⎠ A2 = ⎝ 0 ⎠ A3 = ⎝ 0 ⎠ B = ⎝0 1 ⎠
⎟ (VIII)
0, 0, 1, 1 0,
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠ C2 = ⎝ 0 ⎠ C3 = ⎝ 1 ⎠
0, 0, 0.
If P =⎛IX,⎞then ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 1 0 1 0
⎜0⎟ ⎜1 0 ⎟ ⎜1 1 ⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
A1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 1 ⎟ A2 = ⎜0 0 ⎟ B1 = ⎜0 1 ⎟ B2 = ⎜ 0 ⎟ (IX1 )
⎝0⎠ ⎝0 0 ⎠ ⎝1 0 ⎠ ⎝0⎠
0, 1 0, 0 0, 1,
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1⎟ ⎜0 0 ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜
⎜ 0 ⎟ C2 = ⎜ 0 ⎟ C3 = ⎜1 0 ⎟
⎟
⎝0⎠ ⎝0⎠ ⎝0 1 ⎠
0, 0, 0 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 1 1 0 0
⎜0⎟ ⎜1 0 ⎟ ⎜0 1 ⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
A1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 1 ⎟ A2 = ⎜0 0 ⎟ B1 = ⎜1 1 ⎟ B2 = ⎜ 0 ⎟ (IX2 )
⎝0⎠ ⎝0 0 ⎠ ⎝0 1 ⎠ ⎝0⎠
0, 1 0, 0 0, 1,
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1 0 ⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ C2 = ⎜0 1 ⎟ C3 = ⎜ 0 ⎟
⎝0⎠ ⎝0 0 ⎠ ⎝1⎠
0, 0 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 0 1 0 0
⎜1⎟ ⎜0 0 ⎟ ⎜0 1 ⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
A1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ A2 = ⎜1 0 ⎟ B1 = ⎜1 0 ⎟ B2 = ⎜ 0 ⎟ (IX3 )
⎝0⎠ ⎝0 0 ⎠ ⎝1 1 ⎠ ⎝1⎠
0, 0 1, 0 0, 1,
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 0 1
⎜0 0 ⎟ ⎜1⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜1 0 ⎟ C2 = ⎜ 0 ⎟ C3 = ⎜ 0 ⎟
⎝0 1 ⎠ ⎝0⎠ ⎝0⎠
0 0, 0, 0;
REPRESENTATIONS OF POSETS AND ALGEBRAS 121
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 0 0 1 0
⎜1⎟ ⎜0 0 ⎟ ⎜1 1 ⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
A1 =⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ A2 = ⎜0 0 ⎟ B1 = ⎜0 1 ⎟ B2 = ⎜ 0 ⎟
⎜ ⎟ (IX4 )
⎝0⎠ ⎝0 1 ⎠ ⎝1 0 ⎠ ⎝0⎠
0, 1 0, 0 0, 1,
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
C1 =⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ C2 = ⎜ 0 ⎟ C3 = ⎜ 1 ⎟
⎝0⎠ ⎝0⎠ ⎝0⎠
0, 0, 0;
⎛ ⎞ ⎛ ⎞
0 0 1 ⎛ ⎞ ⎛ ⎞
⎜0⎟ ⎜1 0 ⎟ 1 1 0
⎜ ⎟ ⎜ ⎟ ⎜0 1 ⎟ ⎜0⎟
A1 =⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜1⎟ A2 = ⎜0 0⎟ B1 = ⎝1 0 ⎠ B2 = ⎝ 0 ⎠
⎜ ⎟ (IX5 )
⎝0⎠ ⎝1 0 ⎠
0 0, 1,
, ,
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟
C1 =⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝0⎠ , C2 = ⎝ 0 ⎠ C3 = ⎝ 1 ⎠
0 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 1 1 0
⎜0⎟ ⎜1 0 ⎟ ⎜1⎟ ⎜0⎟
A1 =⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 1 ⎠ A2 = ⎝0 0 ⎠ B1 = ⎝ 1 ⎠ B2 = ⎝ 0 ⎠ (IX6 )
0, 1 0, 0, 1,
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟
C1 =⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠ C2 = ⎝ 0 ⎠ C3 = ⎝ 1 ⎠
0, 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 0 1 0
⎜1⎟ ⎜0⎟ ⎜1 1⎟ ⎜0⎟
A1 =⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠ A2 = ⎝ 0 ⎠ B1 = ⎝1 0⎠ , B2 = ⎝ 0 ⎠
0, 1, 0 0 1,
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟
C1 =⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠ C2 = ⎝ 0 ⎠ C3 = ⎝ 1 ⎠ (IX7 )
0, 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 1 0
⎜1⎟ ⎜0⎟ ⎜1⎟ ⎜0⎟
A1 ⎜ ⎜
= ⎝ ⎠ A2 = ⎝ ⎠ B1 = ⎝ ⎠ B2 = ⎝ ⎟
⎟ ⎜ ⎟ ⎟ ⎜ , (IX8 )
0 0 1 0⎠
0, 1, 0, 1
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟
C1 =⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠ C2 = ⎝ 0 ⎠ C3 = ⎝ 1 ⎠
0, 0, 0;
122 ALGEBRAS, RINGS AND MODULES
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 1
10
A1 = A2 = ⎝ 0 ⎠ B1 = ⎝ 1 ⎠ B2 = ⎝ 1 ⎠ (IX9 )
0,
1, 0, 1,
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 1
C1 = ⎝ 0 ⎠ C2 = ⎝ 1 ⎠ C3 = ⎝ 0 ⎠
1, 0, 0.
If P = X, then
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 1 0 0 1 1
⎜1 0 ⎟ ⎜0 0 ⎟ ⎜0 0 1⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0 1 ⎟ ⎜0 0 ⎟ ⎜0 1 0⎟
⎜
A1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟,
0 0 ⎟ A2 = ⎜0 0 ⎟ B = ⎜1 0 1⎟
(X1 )
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝0 0 ⎠ ⎝0 0 ⎠ ⎝0 1 0⎠
0 0, 0 1, 1 0 0
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟ ⎜0 0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0⎟ ⎜0⎟ ⎜1⎟ ⎜0 0⎟
C1 = ⎜ ⎟ C2 = ⎜ ⎟ C3 = ⎜
⎜ ⎟ ⎜ ⎟ ⎟
⎜ 0 ⎟ C4 =⎜
⎜1
⎟
0
⎜ ⎟ 0
⎜ ⎟ ⎜ ⎟ ⎜ 0⎟⎟
⎝0⎠ ⎝0⎠ ⎝0⎠ ⎝0 1⎠
0, 0, 0, 0 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 1 0 0 1 0
⎜1 0 ⎟ ⎜0 0 ⎟ ⎜0 1 1⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0 1 ⎟ ⎜0 0 ⎟ ⎜0 0 1⎟
⎜
A1 = ⎜ ⎟ ⎜
A2 = ⎜ ⎟ B=⎜ ⎟, (X2 )
⎟ ⎟ ⎜1 0 1⎟
⎜0 0 ⎟ ⎜0 0 ⎟ ⎜ ⎟
⎝0 0 ⎠ ⎝0 0 ⎠ ⎝0 1 0⎠
0 0 0 1, 0 0 0
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0 0
⎜0⎟ ⎜1⎟ ⎜0 0 ⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0⎟ ⎜0⎟ ⎜ ⎟ ⎜0⎟
C1 = ⎜ ⎟ C2 = ⎜ ⎟ C3 = ⎜1 0 ⎟ C4 =⎜ ⎟
⎜0⎟ ⎜0⎟ ⎜0 1 ⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝0⎠ ⎝0⎠ ⎝0 0 ⎠ ⎝0⎠
0, 0, 0 0, 1;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 0 0 0 1 0
⎜0 0 ⎟ ⎜0 0⎟ ⎜0 0 1⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜1 0 ⎟ ⎜ 0⎟ ⎜ 1⎟
A1 = ⎜ ⎟ A2 = ⎜0 ⎟ B = ⎜1 0 ⎟, (X3 )
⎜0 1 ⎟ ⎜0 0⎟ ⎟ ⎜0 1 1⎟
⎜ ⎟ ⎜ ⎜ ⎟
⎝0 0 ⎠ ⎝1 0 ⎠ ⎝0 1 0⎠
0 0, 0 1, 1 0 0
REPRESENTATIONS OF POSETS AND ALGEBRAS 123
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 0 0 1
⎜0⎟ ⎜0 0 ⎟ ⎜1⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0⎟ ⎜1 0 ⎟ ⎜0⎟ ⎜0⎟
⎜ ⎟ ⎜
C1 = ⎜ ⎟ C2 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
0 0 1 ⎟ C3 = ⎜ 0 ⎟ C4 = ⎜ 0 ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝1⎠ ⎝0 0 ⎠ ⎝0⎠ ⎝0⎠
0, 0 0, 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 1 0 0 1 0
⎜1 0 ⎟ ⎜0 0 ⎟ ⎜0 1 1 ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0 1 ⎟ ⎜0 0 ⎟ ⎜0 0 1 ⎟
⎜
A1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎟ A2 = ⎜0 0 ⎟ B = ⎜1 0 1 ⎟ (X4 )
⎜0 0 ⎟ ⎜ ⎟ ⎜ ⎟
⎝0 0 ⎠ ⎝0 0 ⎠ ⎝0 1 0 ⎠
0 0, 0 1, 1 0 0,
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0 0
⎜0 1 ⎟ ⎜0⎟ ⎜0⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0 0 ⎟ ⎜1⎟ ⎜0⎟ ⎜0⎟
⎜
C1 = ⎜ ⎟ C2 = ⎜ ⎟ C3 = ⎜ ⎟ C4 = ⎜
⎜ ⎟ ⎜ ⎟ ⎟
⎟ ⎜0⎟
⎜0 0 ⎟ ⎜0⎟ ⎜1⎟ ⎜ ⎟
⎝0 0 ⎠ ⎝0⎠ ⎝0⎠ ⎝1⎠
0 0, 0, 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0 1 0 1
⎜0 1 ⎟ ⎜0 0 ⎟ ⎜0 1 1 ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0 0 ⎟ ⎜0 0 ⎟ ⎜0 0 1 ⎟
⎜
A1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎟ A2 = ⎜0 0 ⎟ B = ⎜0 1 0⎟, (X5 )
⎜0 0 ⎟ ⎜ ⎟ ⎜ ⎟
⎝0 0 ⎠ ⎝1 0 ⎠ ⎝0 1 0 ⎠
0 0, 0 1, 1 0 0
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0⎟ ⎜0⎟ ⎜1⎟ ⎜0⎟
C1 = ⎜ ⎟ C2 = ⎜ ⎟ C3 = ⎜ ⎟ C4 = ⎜
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜1⎟
⎟
⎜0⎟ ⎜0⎟ ⎜0⎟ ⎜ ⎟
⎝0⎠ ⎝0⎠ ⎝0⎠ ⎝0⎠
0, 0, 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 1 0 0 1 0
⎜1 0 ⎟ ⎜0 0 ⎟ ⎜0 1 1 ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜
A1 = ⎜0 1 ⎟ A2 = ⎜0 0 ⎟ B = ⎜
⎟ ⎜ ⎟
⎜0 0 1⎟,
⎟ (X6 )
⎝0 0 ⎠ ⎝0 0 ⎠ ⎝1 0 1 ⎠
0 0, 0 1, 1 0 0
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ C2 = ⎜ 0 ⎟ C3 = ⎜ 1 ⎟ C4 = ⎜ 0 ⎟
⎝0⎠ ⎝0⎠ ⎝0⎠ ⎝1⎠
0, 0, 0, 0;
124 ALGEBRAS, RINGS AND MODULES
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 0 0 1 0
⎜1⎟ ⎜0 0 ⎟ ⎜0 0 1⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
A1 = ⎜ 0
⎜ ⎟
⎟ A2 = ⎜0 0 ⎟ B = ⎜0 1 1⎟,
⎜ ⎟ ⎜ ⎟ (X7 )
⎝0⎠ ⎝0 0 ⎠ ⎝1 0 1⎠
0, 0 1, 1 0 0
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ C2 = ⎜ 0 ⎟ C3 = ⎜ 1 ⎟ C4 = ⎜ 0 ⎟
⎝0⎠ ⎝0⎠ ⎝0⎠ ⎝1⎠
0, 0, 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 0 0 1 1
⎜0 0 ⎟ ⎜0⎟ ⎜1 0 1⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ ⎟ ⎜ ⎟
A1 = ⎜1 0 ⎟ A2 = ⎜ 0 ⎟ B = ⎜ ⎜1 0 0⎟,
⎟ (X8 )
⎝0 1 ⎠ ⎝0⎠ ⎝0 1 0⎠
0 0, 1, 0 0 1
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 0 1
⎜0⎟ ⎜0⎟ ⎜1⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ C2 = ⎜ 1 ⎟ C3 = ⎜ 0 ⎟ C4 = ⎜ 0 ⎟
⎝1⎠ ⎝0⎠ ⎝0⎠ ⎝0⎠
0, 0, 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 1 0 0 1
⎜1 0 ⎟ ⎜0 0 ⎟ ⎜1 1 ⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 1 ⎟ ⎜ 0 0 ⎟ ⎜ ⎟
A1 = ⎜ ⎟ A2 = ⎜ ⎟ B = ⎜1 0 ⎟, (X9 )
⎝0 0 ⎠ ⎝0 0 ⎠ ⎝0 1 ⎠
0 0, 0 1, 1 0
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ C2 = ⎜ 0 ⎟ C3 = ⎜ 1 ⎟ C4 = ⎜ 0 ⎟
⎝0⎠ ⎝0⎠ ⎝0⎠ ⎝1⎠
0, 0, 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 0 0 1
⎜1⎟ ⎜0 0 ⎟ ⎜1 1⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
A1 = ⎜ 0
⎜ ⎟
⎟ A2 = ⎜0 0 ⎟ B = ⎜0 1⎟,
⎜ ⎟ ⎜ ⎟ (X10 )
⎝0⎠ ⎝0 0 ⎠ ⎝1 0⎠
0, 0 1, 1 0
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ C2 = ⎜ 0 ⎟ C3 = ⎜ 1 ⎟ C4 = ⎜ 0 ⎟
⎝0⎠ ⎝0⎠ ⎝0⎠ ⎝1⎠
0, 0, 0, 0;
REPRESENTATIONS OF POSETS AND ALGEBRAS 125
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0 1
⎜0 1 ⎟ ⎜0⎟ ⎜1 0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
A1 =⎜
⎜ 0 0 ⎟ A2 = ⎜ 0 ⎟ B = ⎜1 1⎟,
⎟ ⎜ ⎟ ⎜ ⎟ (X11 )
⎝0 0 ⎠ ⎝0⎠ ⎝0 1⎠
0 0, 1, 1 0
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
C1 =⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ C2 = ⎜ 0 ⎟ C3 = ⎜ 1 ⎟ C4 = ⎜ 0 ⎟
⎝0⎠ ⎝0⎠ ⎝0⎠ ⎝1⎠
0, 0, 0, 0;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 0 0 1
⎜0⎟ ⎜0 1 ⎟ ⎜1 0⎟
A1 =⎜ ⎟ ⎜ ⎟ ⎜
⎝ 1 ⎠ A2 = ⎝0 0 ⎠ B = ⎝1 1⎠,
⎟ (X12 )
0, 0 0, 1 0
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟ ⎜0⎟
C1 =⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠ C2 = ⎝ 0 ⎠ C3 = ⎝ 1 ⎠ C4 = ⎝ 0 ⎠
0, 0, 0, 1;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 1 0 1
⎜1 0 ⎟ ⎜0⎟ ⎜1 1⎟
A1 =⎜ ⎟ ⎜ ⎟ ⎜
⎝0 1 ⎠ A2 = ⎝ 0 ⎠ B = ⎝1 0⎠,
⎟ (X13 )
0 0, 0, 1 0
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟ ⎜0⎟
C1 =⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠ C2 = ⎝ 0 ⎠ C3 = ⎝ 1 ⎠ C4 = ⎝ 0 ⎠
0, 0, 0, 1;
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 1 0 1
⎜1⎟ ⎜0⎟ ⎜1 0 ⎟
A1 =⎜ ⎟ ⎜ ⎟ ⎜
⎝ 0 ⎠ A2 = ⎝ 0 ⎠ B = ⎝1 1⎠,
⎟ (X14 )
0, 0, 1 0
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟ ⎜0⎟
C1 =⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠ C2 = ⎝ 0 ⎠ C3 = ⎝ 1 ⎠ C4 = ⎝ 0 ⎠
0, 0, 0, 1.
If P = XI, then
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 1 0
⎜0⎟ ⎜0⎟ ⎜0⎟ ⎜1⎟ ⎜1⎟
A1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 0 ⎠ A2 = ⎝ 1 ⎠ A3 = ⎝ 0 ⎠ B1 = ⎝ 0 ⎠ B2 = ⎝1⎠, (XI)
0, 0, 1, 0, 1
126 ALGEBRAS, RINGS AND MODULES
⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 1
⎜0⎟ ⎜1⎟ ⎜0⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎝ 1 ⎠ C2 = ⎝ 0 ⎠ C3 = ⎝ 0 ⎠
0, 0, 0.
If P = XII, then
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 1 1 0 1 0
⎜0⎟ ⎜0 1 ⎟ ⎜1 0 ⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
A1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ A2 = ⎜1 0 ⎟ B1 = ⎜0 0 ⎟ B2 = ⎜0⎟,
⎜ ⎟ (XII)
⎝0⎠ ⎝0 0 ⎠ ⎝1 0 ⎠ ⎝0⎠
0, 0 1, 0 0, 1
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
C1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 0 ⎟ C2 = ⎜ 0 ⎟ C3 = ⎜ 1 ⎟ C3 = ⎜ 0 ⎟
⎝0⎠ ⎝0⎠ ⎝0⎠ ⎝1⎠
0, 0, 0, 0.
If P = XIII, then
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 0 1 1 0 0
⎜0⎟ ⎜1 0 ⎟ ⎜0 1 ⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
A1 = ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜ 1 ⎟ A2 = ⎜0 0 ⎟ B1 = ⎜1 1 ⎟ B2 = ⎜0⎟,
⎜ ⎟ (XIII)
⎝0⎠ ⎝0 0 ⎠ ⎝0 1 ⎠ ⎝0⎠
0, 1 0, 0 0, 1
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟ ⎜0⎟
⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟
C1 = ⎜ 0 ⎟ C2 = ⎜ 0 ⎟ C3 = ⎜ 1 ⎟ C3 = ⎜
⎜ ⎟ ⎜ ⎟ ⎜ ⎟
⎜0⎟
⎟
⎝0⎠ ⎝0⎠ ⎝0⎠ ⎝1⎠
0, 0, 0, 0.
Example 3.1.1.
In this example we show in what way one can obtain the vector representations
of posets from the list above.
Consider the exact indecomposable matrix representation (V2 ) of the poset
a1 • • c1
P =V: b from the list above:
a 2 • • • c2
⎛ ⎞
⎛ ⎞ 1 ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
0 ⎜0 ⎟ 1 1 0 0
A1 = ⎝ 1 ⎠ A2 = ⎜ ⎝0 ⎠
⎟ B = ⎝0 1 ⎠ C1 = ⎝ 1 ⎠ C2 = ⎝0⎠
0, 1 0, 0, 1
,
Then the corresponding matrix representation can be written as the following
matrix partitioned horizontally into 5 vertical blocks:
REPRESENTATIONS OF POSETS AND ALGEBRAS 127
0 1 1 1 0 0
A= 1 0 0 1 1 0
0 0 1 0 0 1
The corresponding vector representation V = {V0 , Va1 , Va2 , Vb , Vc1 , Vc2 } over
a field k can be obtained in the following way. Since the number of rows of the
matrix A is equal to 3, dimk V0 = 3. ⎛Therefore
⎞ V⎛
0 =⎞{e1 , e2 , ⎛
e3 }⎞is a vector space
1 0 0
spanned by the basis elements e1 = ⎝0⎠, e2 = ⎝1⎠, e3 = ⎝0⎠. Since a2 a1
0 0 1
in P, Va2 ⊂ Va1 and so Va2 = {e1 }, Va1 = {e1 , e2 }. Since c2 c1 in P and
a1 , c2 are incomparable in⎛P,⎞Vc2 ⊂ ⎛ V⎞c1 , and so Vc2 = {e3 }, Vc1 = {e2 , e3 },
1 1
Vb = {e1 + e3 , e1 + e2 } = k ⎝0⎠ + k ⎝1⎠.
1 0
For a finite poset P = (S, ) of size m one introduces the rational Tits
quadratic form qP : Qm+1 → Q defined by:
m+1 m
qP = x2i + xi xj − ( xi )xm+1 .
i=1 i≺j≤m i=1
The rational Tits quadratic form qP is called weakly positive if qP (x) > 0 for any
nonzero vector x = (x1 , x2 , . . . , xm+1 ) ∈ Qm+1 with x1 , x2 , . . . , xm+1 ≥ 0. And
this form is called weakly non-negative, if qP (x) ≥ 0 for any nonzero vector
x = (x1 , x2 , . . . , xm+1 ) ∈ Qm+1 with x1 , x2 , . . . , xm+1 ≥ 0.
Examples 3.1.2.
• 2
1. Let P be the poset . The corresponding rational Tits quadratic
• 1
form
qP = x21 + x22 + x23 + x1 x2 − (x1 + x2 )x3
is weakly positive since
1 1 1
qP = (x1 + x2 )2 + (x1 − x3 )2 + (x2 − x3 )2 > 0
2 2 2
for any nonzero vector (x1 , x2 , x3 ).
2. Let P be the poset • • • • . The corresponding rational Tits
quadratic form
1 1 1 1
qP = (x1 − x5 )2 + (x2 − x5 )2 + (x3 − x5 )2 + (x4 − x5 )2 ≥ 0
2 2 2 2
for any nonzero vector (x1 , x2 , x3 , x4 , x5 ).
As in the case of representations of quivers, for finite posets one can also
introduce the notions of tame representation type and wild representation type.
Let P = (S, ), where S = {1, 2, ..., m}, be a finite poset. And let V = (V0 ; Vi :
i ∈ S) be a representation of P over a field k. As already said, the dimension
vector d = dimV = (d0 , d1 , . . . , dm ) with coordinates d0 = dim V0 , di = dim Vi /V̄i
m
is called the dimension of the representation V . Write ω(V ) = d0 + di .
i=1
N1 = (1, 1, 1, 1, 1) : • • • • •
• • • •
N2 = (1, 1, 1, 2) :
•
• •
• • •
• •
N3 = (2, 2, 3) : • • • N4 = (1, 3, 4) :
• • •
•
•
•
• • •
•
• • •
•
N5 = (N, 5) : • N6 = (1, 2, 6) :
•
•
• •
•
• • •
130 ALGEBRAS, RINGS AND MODULES
Examples 3.2.1.
1. The posets (2,2,2), (1,3,3), (N,4), (1,2,5) in theorem 3.1.2 have width equal
to 3, whereas the poset (1,1,1,1) has width equal to 4.
2. A poset P is of width one if and only if P is a chain.
Lemma 3.2.1. If the width w(P) of a poset P is greater than or equal to four,
then P is of infinite representation type.
J(n, λ) E E O
A(n,λ) =
E E O E
From this lemma it follows that a poset of finite representation type has
width ≤ 3.
For a poset P = (S, ) and x ∈ S we define the upper and lower cone of x:
xΔ = {y ∈ S : x y}
x∇ = {y ∈ S : y x}
REPRESENTATIONS OF POSETS AND ALGEBRAS 131
Definition. Let P = (S, ) and T = (T, ) be any two (disjoint) posets. The
cardinal sum P ∪ T or (P, T ) of P and T is the set of all s ∈ S and t ∈ T with
a relation such that s s1 and t t1 (s, s1 ∈ S; t, t1 ∈ T ) have unchanged
meanings and there are no other relations in P ∪ T .
Pba = P \ (aΔ ∪ b∇ )
Remark 3.2.3. A suitable pair of elements does not always exist for any poset.
For example, the poset P : • • • • has no such suitable pair of elements.
If a poset contains no suitable pair of elements it is called non-differential.
Example 3.2.2.
Consider the poset P of the following form:
b • •
• • a
i = 1, 2, ..., n.
(3) If during the second step we get elements x, y with x ≺ y and y ≺ x, then
we identify them in δ(a,b) P.
REPRESENTATIONS OF POSETS AND ALGEBRAS 133
Example 3.2.3.
Let P be of the following form:
a2 • b • c •
a1 •
Then δ(a1 ,b) P has the form:
a2 b c+
a1 c−
Proposition 3.2.5. Any exact non-differential poset which does not contain
as a full subposet any poset of the form N1 = (1, 1, 1, 1, 1) and N2 = (1, 1, 1, 2) is
a sum of two garlands.
134 ALGEBRAS, RINGS AND MODULES
• •
NZ :
• • • •
Remarks 3.2.5. From results of the paper [Nazarova, Roiter, 1973] it follows
that all posets which are a sum of two garlands are of tame type.
T (a + b) = T (a) + T (b)
T (αa) = αT (a)
T (ab) = T (a)T (b)
T (1) = E (the identity operator)
for arbitrary a, b ∈ A, α ∈ k.
The action of the operators T (a) on V is written on the right, i.e. T (a) : V →
V , v → vT (a).
If the vector space V is finite dimensional over k, then its dimension is called
the dimension (or degree) of the representation T . Obviously, the image of a
representation T forms a subalgebra in Endk (V ). If T is a monomorphism, then
this subalgebra is isomorphic to the algebra A. In this case the representation T
is called faithful.
Proof. From the axioms for algebras it follows that for every a ∈ A the map
T (a) : x → xa, x ∈ A, is a linear operator on the space A. Moreover, T (a + b) =
T (a) + T (b), T (αa) = αT (a), T (ab) = T (a)T (b) and T (1) = E (the identity
operator). Thus, T is a representation of the algebra A. If a = b, then 1 · a = 1 · b.
This shows that the operators T (a) and T (b) are distinct and T is a faithful
representation, as required.
n
T ej = aij ei , aij ∈ k.
i=1
(1)
Tx Ux
Tx = (2)
0 Tx
(i)
for all x ∈ A, where the Tx are square matrices of degree ni < n, where n =
dimk V .
The Schur lemma (see proposition 2.2.1, vol.I) says that the endomorphism
ring of a simple representation is a division ring.
138 ALGEBRAS, RINGS AND MODULES
(i)
for all g ∈ G, where the Tx form irreducible square matrix representations of
degree ni < n, where n = dimk V , i = 1, 2, ..., m.
Any finite dimensional module over an algebra A can be uniquely written
(up to isomorphism) in the form of a direct sum of indecomposable modules by
the Krull-Remak-Schmidt theorem. This means that for many questions one can
restrict attention to the consideration of indecomposable modules.
Let A be a finite dimensional k-algebra, and let V be a finite dimensional right
A-module. We construct a category C(V ) whose objects are the submodules of
tensor products of the form U ⊗k V , the U being finite dimensional vector spaces;
a morphism from an object X ⊂ U ⊗k V to an object Y ⊂ W ⊗k V is a linear
mapping φ : U → W such that (φ ⊗ 1)(X) ⊂ Y . The problem of classifying the
objects of C(V ) up to isomorphism is, by definition, a linear matrix problem.
The following considerations and constructions serve to reformulate such a
linear matrix problem as a representation theoretic problem. Clearly, the module
V may be assumed faithful (by replacing, if needed, the algebra A with its quotient
algebra A/(Ker(A → Endk (V ))) ). Thus A is identified with a subalgebra of
E = Endk V . Let O be a discrete valuation ring with field of residues k, and let π
be a prime element of O. Consider an O-lattice L (i.e., a free O-module) of rank
n = dim V . If Γ = EndO L, then Γ/πΓ E, and L/πL V as an E-module. Let
Λ stand for the preimage of the subalgebra A ⊂ E in Γ, and consider the category
Rep(Λ) of representations of Λ (over O); that is, of Λ-modules that are O-lattices.
Every such module M can be naturally embedded into the Γ-module M Γ which is
a representation of Γ. But every Γ-module is of the form F ⊗O L, where F is some
O-lattice, and every Γ-homomorphism F ⊗O L → G ⊗O L is of the form f ⊗ 1,
REPRESENTATIONS OF POSETS AND ALGEBRAS 139
Remarks 3.3.1.
1) The classification of the representations of an arbitrary O-order Λ such that
Γ ⊃ Λ ⊃ πΓ for some maximal order Γ can be reduced in a similar fashion to a
140 ALGEBRAS, RINGS AND MODULES
Examples 3.4.1.
1. Every indecomposable representation of a finite dimensional semisimple
algebra is equivalent to a direct summand of the regular representation, by the
Wedderburn theorem. Hence, every finite dimensional semisimple algebra is an
algebra of finite representation type.
2. A k-algebra k[x]/(xn ) is an algebra of finite representation type.
3. The algebra A = {1, r, s : r2 = s2 = rs = sr = 0} is an algebra of infinite
representation type.
4. The k-algebra k[x, y]/(xn , y m ), for n, m ≥ 2, is an algebra of infinite repre-
sentation type.
5. The group algebra KG of a finite group G over a field K of characteristic
p > 0 has finite type if and only if the p-Sylow subgroup of G is cyclic.
6. Any serial algebra is of finite type.
This conjecture was proved by A.V.Roiter in 1968 [Roiter, 1968] for a finite
dimensional algebra A over an arbitrary field.
And it was generalized by M.Auslander for Artinian algebras [Auslander, 1971],
[Auslander, 1974a], [Auslander, 1974b].
So far this stronger conjecture has been proved for a finite dimensional algebra
over an algebraically closed field k by R.Bautista [Bautista, 1985] and K.Bongartz
[Bongartz, 1985]:
All k-algebras of infinite type are further divided into algebras of wild rep-
resentation type and algebras of tame representation type.
and free over kx, y and such that the functor M ⊗kx,y ∗ sends non-isomorphic
finite dimensional kx, y-modules to non-isomorphic A-modules. In this case the
category of all finite-dimensional A-modules includes the classification problem for
pairs of square matrices up to simultaneous equivalence.
Examples 3.4.2.
1. The Kronecker algebra
k k⊕k
A=
0 k
Taking into account theorems 2.4.1 and 2.3.4 we have the following statement.
With any finite dimensional algebra A over a field k one can associate its
Gabriel quiver Q(A) (see section 11.1, vol.I).
Let P1 , ..., Ps be all pairwise nonisomorphic principal right A-modules. Write
Ri = Pi R (i = 1, ..., s) and Vi = Ri /Ri R where R is the radical. Since Vi is a
s t
semisimple module, Vi = ⊕ Uj ij , where the Uj = Pj /Rj are simple modules. This
j=1
REPRESENTATIONS OF POSETS AND ALGEBRAS 143
s t
is equivalent to the isomorphism P (Ri ) ⊕ Pj ij . To each module Pi assign a
j=1
vertex i and join the vertex i with the vertex j by tij arrows. The thus constructed
graph is called the quiver of A in the sense of P.Gabriel and denoted by Q(A).
For any finite quiver Q = (V Q, AQ, s, e) we can construct a bipartite
quiver Qb = (V Qb , AQb , s1 , e1 ) in the following way. Let V Q = {1, 2, ..., s},
AQ = {σ1 , σ2 , ..., σk }. Then V Qb = {1, 2, ..., s, b(1), b(2), ..., b(s)} and AQb =
{τ1 , τ2 , ..., τk }, such that for any σj ∈ AQ we have s1 (τj ) = s(σj ) and e1 (τj ) =
b(e(σj )). In other words, in the quiver Qb from the vertex i to the vertex b(j) go
tij arrows if and only if in the quiver Q from the vertex i to the vertex j go tij
arrows. As before, denote by Q the undirected graph which is obtained from Q
by deleting the orientation of all arrows.
Example 3.4.3.
Let k be a field and ⎛ ⎞
k 0 k
A = ⎝0 k k⎠
0 0 k
Then ⎛ ⎞
0 0 k
R = rad A = ⎝0 0 k⎠
0 0 0
and so R2 = 0.The right principal modules are P1 = k 0 k , P2 = 0 k k ,
P3 = 0 0 k , while the simple
right modules
are the Ui Pi /Pi R for i = 1, 2, 3.
Therefore P1 R = 0 0 k U3 , P2 R = 0 0 k U3 , P3 R = 0. Thus the
quiver Qb (A) has the following form:
1 2 3
• •
• • •
b1 b2 b3
Proof. We shall give only a short sketch of the proof. Let A be a finite
dimensional algebra over an algebraically closed field k with Jacobson radical R
144 ALGEBRAS, RINGS AND MODULES
Taking into account theorem 3.4.6 and theorem 2.6.1, we obtain the following
theorem:
Remark 3.4.1. This theorem was proved by P.Gabriel in the case when the k-
species LA = (Ki , i Mj )i,j∈I has the property that all Ki are equal to a fixed skew
field F and F (i Mj )F = (F FF )nij for some natural number nij . Also P.Gabriel has
shown that the structure of a k-algebra A of finite type with (radA)2 = 0 can be
recovered from the known results in the case when k is a perfect field [Gabriel,
1972], [Gabriel, 1973]. In its general form for arbitrary fields k theorem 3.4.8 was
proved by V.Dlab and C.M.Ringel [Dlab, Ringel, 1973], [Dlab, Ringel, 1975].
With any species L = (Ki , i Mj )i,j∈I one can associate the special tensor alge-
bra of a bimodule
of the following form T(L) = TB (M ) = B ⊕ M ⊕ M ⊗B M ⊕ . . .,
where B = Ki and M = ⊕ i Mj .
i∈I i,j∈I
Remark 3.4.2. Theorem 3.4.9 was proved by V.Dlab and C.M.Ringel (see
[Dlab, Ringel, 1973], [Dlab, Ringel, 1975]. A different version of this theorem was
obtained by E.L.Green [Green, 1975].
Let L be a k-species. From corollary 2.2.13 it follows that the special tensor
algebra T(L) is hereditary. It was shown that the converse is also true for the case
of finite type.
Remark 3.4.3. This theorem first was proved by P.Gabriel2 for the case of
an algebraically closed field k and then was proved by V.Dlab, C.M.Ringel for an
arbitrary field k.3
Applying theorem 2.7.1, V.Dlab and C.M.Ringel proved the following theorems
which give a full description of hereditary Artinian algebras of finite type.
It is easy to see that if A and B are Noetherian modules then A|B if and only
if there is an integer n such that there is an exact sequence A(n) → B → 0.
for i < j. A module A which cannot be decomposed into a nontrivial normal direct
sum is called normally indecomposable.
It is obvious that any Noetherian module A can be decomposed into a direct
sum A = A ⊕ A , where A is normally indecomposable and A |A.
Proof. Let U = HomΛ (A, A), which is a finite dimensional k-algebra. Set
T = HomΛ (A, B)ϕ. Obviously, T is a left ideal of U and AT = A. Let R = rad Λ
be the radical of Λ. Write U = U/R and T = (T + R)/R. Then T is a left ideal
in U and T = eU , where e is an idempotent of T . Let e ∈ T be an idempotent
corresponding to the idempotent e. Then T ⊆ eU + R, A = AT = AeU + AR.
Therefore, by the Nakayama lemma, A = AeU . Since A = Ime ⊕ Kere, Ime|A.
By assumption, A is normally indecomposable, so Ime = A, i.e., e is an identity
of U = Hom(A, A). Since e ∈ T = Hom(A, B)ϕ, e = ψϕ, where ψ ∈ Hom(A, B).
ϕ
So, by proposition 4.2.1, vol.I, the exact sequence B → A → 0 splits.
ϕ
B ⊕ X −→ A ⊕ X → 0 (3.5.2)
which is also an exact sequence, where (b, x)ϕ = (bϕ, x). Then the sequence (3.5.1)
splits if and only the sequence (3.5.2) splits.
In what follows in this section we shall assume that all modules are finitely
generated over a finite dimensional k-algebra Λ. Each such module A obviously
has finite length which is denoted by l(A).
Then we set
Mk+1 = {A ∈ Mk : l(A) = max l(B)}.
B∈Mk
k
Since Mk ∩ ∪ Mi = ∅, and Mk+1 ⊂ Mk , it is obvious that Mi ∩ Mj = ∅ for
i=1
i = j.
Proof. Since the lengths of all modules from M are bounded it is sufficient to
show that modules with a fixed length k can be in only a finite number of Mi .
We shall prove this by induction on k. For k = 1 this statement is trivial because
∞ s
all modules of length 1 are in M1 and only in M1 . Set R = ∪ Mi , Rs = ∪ Mi .
i=1 i=1
Assume that for j ≤ k the modules of length j belong only to a finite number
of Mi . Then there is a number s such that for any A ∈ Rs its length l(A) ≥ k.
Suppose that A ∈ Rs and l(A) = k + 1. If C is a proper indecomposable quotient
module of A, then C ∈ R, since otherwise A ∈ Mi for any i, and so A ∈ R. Since
s
l(C) < l(A) = k + 1, C ∈ ∪ Mi . Therefore A ∈ Ms , and it is obvious that
i=1
A ∈ Mi for all i ≥ s. So if B is any other module such that B ∈ Mt for t > s and
l(b) = k + 1 then also A ∈ Mt , i.e., all modules from Rs , whose length is equal to
k + 1, belong to not more than one set Mi .
k
Lemma 3.5.5. If A ∈ M and A ∈ ∪ Mi , then there is a quotient module B
i=1
of A which is contained in Mk .
k
Proof. If all indecomposable quotient modules of A belong to ∪ Mi , then
i=1
we can take B = A. Otherwise we choose a quotient module A1 ∈ M such that
k
A1 ∈ ∪ Mi . If A1 ∈ Mk then take A1 = B. If A1 ∈ Mk , consider a quotient
i=1
k
module A2 ∈ M such that A2 ∈ ∪ Mi , and so on. From the ascending chain
i=1
condition it now follows that in a finite number of steps we find a quotient module
which belongs to Mk .
n
Lemma 3.5.6. There is an integer n such that ∪ Mi = M.
i=1
Proof. By lemma 3.5.4, there exists an integer n such that Mi = ∅ for i > n.
n
If ∪ Mi = M, then, by lemma 3.5.5, Mn = ∅, and so Mn+1 = ∅.
i=1
Proof. From the construction of the sets Mi and lemma 3.5.6 it follows that
for any indecomposable module A ∈ M there exists a unique integer i such that
A ∈ Mi . We set f (A) = i. From lemma 3.5.6 it follows that the function f satisfies
condition 3), the fulfillment of condition 2) follows from the construction of the
sets Mk+1 . So it remains to verify condition 1).
Let there be a non-split exit sequence
s
⊕ Bi → A → 0, (3.5.3)
i=1
B ⊕ A −→ A ⊕ X → 0,
Imϕij = C. Since C ∈ Mk−1 , l(Bi ) > l(C) for i = 1, . . . , q. Therefore for
i,j
1 ≤ i ≤ q and for any ϕij we have that either 1) Imϕij is a proper quotient
module or 2) Imϕij = C Bi .
If for all i (1 ≤ i ≤ q) we have the first case, then from the construction of
f it follows that the values of f on all direct summands of Imϕij are not more
than k − 1. Since f (Bi ) ≤ k − 1 for q < i ≤ r, we then obtain that D|C, where
r
D= ⊕Imϕij ⊕ ⊕Bi and the values of f on all indecomposable direct
i,j i=q+1
summands of D are not more than k−1. Taking an exact sequence D(n) → C → 0,
we see that either this sequence splits and then f (C) ≤ k −1, by the Krull-Schmidt
theorem, or this sequence is not split and then f (C) < k − 1, by the induction
assumption. In the both cases we have a contradiction with C ∈ Mk−1 .
Now consider the second case when C Bi for some i, where 1 ≤ i ≤ r. In
this case we consider the exact sequence:
A −→ C → 0. (3.5.5)
A ⊕ X −→ C ⊕ X → 0,
Proof. The group Ext(Mi , B) is a finite dimensional vector space over k. Sup-
pose its dimension is equal to si . If ni > si then there is a module Mi which is a
direct summand of X. Therefore it is sufficient to take
t
N= si l(Mi ) + l(B).
i=1
algebras (see [Roiter, 1968]) using a remarkably simple argument. The proof of
this theorem “marks the beginning of the new representation theory of finite di-
mensional algebras”, as C.M.Ringel wrote in his paper [Ringel, 2004].
The first Brauer-Thrall conjecture in a more general form (in particular, for
Artinian algebras and one-sided Artinian rings) was proved by M.Auslander (see
[Auslander, 1974a], [Auslander, 1974b]). In these papers M.Auslander proved
that if C is a skeletally small Abelian category with only a finite number of non-
isomorphic simple objects and such that each object has finite composition length,
then C has only a finite number of indecomposable objects if and only if C satisfies
the following conditions:
(1) C has a.c.c. on chains of indecomposable objects;
fi
(2) if {Mi −→ Mi−1 }i∈N is a sequence of epimorphisms, then there is an n
such that for i ≥ n, fi is an isomorphism.
M.Auslander also constructed a one-to-one correspondence between isomor-
phism classes of Artin algebras of finite type and isomorphism classes of Artin
algebras of global dimension at most two and dominant dimension at least two
(see [Auslander, 1971]). The generalizations of these results to Artinian rings were
given by M.Auslander and H.Tachikawa (see [Auslander, 1974b], [Auslander, 1975]
and [Tachikawa, 1973]).
The first Brauer-Thrall conjecture for arbitrary algebras (not necessarily finite
dimensional) was studied by A.D.Bell and K.R.Goodearl. We say that a k-algebra
A is of right bounded finite dimensional representation type if it has an
upper bound on the k-dimensions of the finite dimensional indecomposable right
A-modules. They showed that if a k-algebra A is either finitely generated as a
k-algebra, or Noetherian as a ring, then bounded finite dimensional type implies
that A has only finitely many isomorphism classes of finite dimensional indecom-
posable modules (see [Bell, Goodearl, 1995]).
The classification of hereditary finite dimensional k-algebras and algebras with
zero square radical of finite representation type were obtained by P.Gabriel in the
case when the corresponding k-species has the property that all Ki are equal to
a fixed skew field F and F (i Mj )F = (F FF )nij for some natural nij (see [Gabriel,
1972]; [Gabriel, 1972/1973], [Gabriel, 1973], [Gabriel, 1974]). In the general case
these theorems were proved by V.Dlab, C.M.Ringel (see [Dlab, Ringel, 1975] and
[Dlab, Ringel, 1976]).
The description of hereditary finite dimension algebras of tame type and wild
type was obtained by V.Dlab and C.M.Ringel (see [Dlab, Ringel, 1976], [Ringel,
1976], [Ringel, 1978]).
In 1980 Yu.A.Drozd proved his famous theorem which says that for an
algebraically closed field k there is a trichotomy between finite, tame and wild
representation type for finite dimensional algebras (see [Drozd, 1980]).
For more references see [Gustafson, 1982].
156 ALGEBRAS, RINGS AND MODULES
1975, p.607-615.
[Kleiner, 1972b] M.M.Kleiner, On the exact representations of partially ordered
sets of finite type, Zapiski Nauch. Semin. LOMI, vol.28, 1972, p.42-59;
English transl.: J. Soviet. Math., v.3, 1975, p.616-628.
[Kleiner, 1988] M.M.Kleiner, Pairs of partially ordered sets of tame representation
type, Linear Algebra Appl., v.104, 1988, p.103-115.
[Krugljak, 1972] S.A.Krugljak, Representations of algebras with zero square rad-
ical, Zap. Nauchn. Sem. LOMI, v.28, 1972, p.60-68; English transl. J.
Soviet Math., v.3, N.5, 1975.
[Nakayama, 1940] T.Nakayama, Note on uni-serial and generalized uni-serial
rings, Proc. Imperial Acad. Japan, v.16, 1940, p.285-289.
[Nazarova, Roiter, 1972] L.A.Nazarova, A.V.Roiter, Representations of partially
ordered sets, Zap. Nauchn. Sem. LOMI, v.28, 1972, p.5-31 (in Russian);
English transl. J. Soviet Math., vol.3, 1975, p.585-606.
[Nazarova, 1974] L.A.Nazarova, Partially ordered sets of infinite type, In: Lecture
Notes in Math., Springer-Verlag, v.488, 1974, p.244-252.
[Nazarova, 1975] L.A.Nazarova, Representations of partially ordered sets of infi-
nite type, Izv. Akad. Nauk SSR, Ser. Mat., v.39, 1975, p.963-991.
[Nazarova, Roiter, 1973] L.A.Nazarova, A.V.Roiter, On a problem of
I.M.Gel’fand, Funk. Anal i Prilozhen., v.7, No.2, 1973, p.54-69 (in Russian).
[Nazarova, Roiter, 1983] L.A.Nazarova, A.V.Roiter, Representations and forms of
weakly completed partially ordered sets, In: Linear algebra and the theory of
representations, Akad. Nauk Ukrain. SSR, Inst. Mat., Kiev, 1983, p.19–54
(in Russian).
[Nazarova, Zavadskij, 1977] L.A.Nazarova, A.G.Zavadskij, Partially ordered sets
of tame type, In: Matrix problems, Kiev, 1977, 122–143.
[Nazarova, Zavadskij, 1977] L.A.Nazarova, A.G.Zavadskij, Partially ordered sets
of finite growth, Funkts. Anal. Prilozh., v.16, 1982, p.72-73 (in Russian),
English transl.: Funct. Anal. Appl., v.16, 1982, p.135–137.
[Ringel, 2004] C.M.Ringel, Foundation of the representation theory of Artin alge-
bras, using the Gabriel-Roiter Measure. In: José A.de la Peña (ed.), Trends
in the representation theory of algebras and related topics, Contemporary
Mathematics 406, Americal Mathematical Society, 2006, 105–135.
And also in: Masahisa Sato (ed.), Proceedings of the 36th symposium on ring
theory, Vol.2, 2004, 1–19.
[Ringel, 1976] C.M.Ringel, Representations of K-species and bimodules, J. Alge-
bra, v.41, 1976, p.269–302.
[Ringel, 1980] C.Ringel, Tame algebras (On algorithms for solving vector space
problems II), Carleton Math. Lecture Notes, v.25, 1980, p.1–32.
[Ringel, 1978] C.M.Ringel, Finite dimensional hereditary algebras of wild repre-
sentation type, Math. Z., v.161, 1978, p.235–255.
[Roiter, 1968] A.V.Roiter, Unbounded dimensionality of indecomposable repre-
sentations of an algebra with an infinite number of indecomposable repre-
sentations, Math. USSR Izv., v.2, N.6, 1968, p.1223–1230.
REPRESENTATIONS OF POSETS AND ALGEBRAS 159
161
162 ALGEBRAS, RINGS AND MODULES
Remark 4.1.1. Taking into account the dual definitions of projective and
injective modules, it is easy to see that any indecomposable left injective A-module
Q is equal to a P ∗ = Homk (P, k), where P is a principal right A-module.
Remark 4.1.2. By the annihilation lemma (see vol.I, p.265), for any simple
right A-module U there exists a unique canonical idempotent f ∈ A such that
U f = U . By the definition of U ∗ , we have U f = U if and only if f U ∗ = U ∗ .
Proposition 4.1.3. For any finitely generated right A-module M we have that
(radM )⊥ is the socle of M ∗ and soc M ∗ (M/rad M )∗ .
Qi → soc Qi = Ui
and
Ui → E(Ui ),
where E(Ui ) is the injective hull of Ui .
164 ALGEBRAS, RINGS AND MODULES
If an algebra A is Frobenius then the left modules A A and (AA )∗ are also
isomorphic.
Taking remark 4.1.1 into account, it follows that if A is a Frobenius algebra,
then the right module (A A)∗ is injective. Analogously, the left module (AA )∗ is
injective.
Proof.
(1) ⇒ (2). Let the left A-modules A A and (AA )∗ be isomorphic and let
Θ : A A → (AA )∗
f (x, y) = xGy T ,
where ⎛ ⎞
y1
⎜ ⎟
yT = ⎝ ... ⎠ .
yn
The implication
G y T = 0 ⇒ y T = 0.
FROBENIUS ALGEBRAS AND QUASI-FROBENIUS RINGS 165
From theorem 4.2.1 there immediately follows the following equivalent defini-
tion of a symmetric algebra.
⎛ ⎞
I11 + J11 I12 + J12 ··· I1n + J1n
⎜ I21 + J21 I22 + J22 ··· I2n + J2n ⎟
⎜ ⎟
I+J =⎜ .. .. .. .. ⎟,
⎝ . . . . ⎠
In1 + Jn1 In2 + Jn2 ··· Inn + Jnn
n
(IJ )ij = Iik Jkj (i, j = 1, . . . , n),
k=1
Example 4.3.1.
Let A = Tn (D) be a ring of all upper-triangular matrices over a division ring
D. Let P1 = e11 A, . . . , Pn = enn A and A P1 = Ae11 , . . . , A Pn = Aenn . Let
Zr = soc AA and Zl = socA A. Obviously Zr = A Pn = Un ⊕ . . . ⊕ Un , where
n times
Un = Pn is a simple right A-module and Zl = P1 = V1 ⊕ . . . ⊕ V1 , where V1 = A P1
n times
is a simple left A-module. Thus Zr ∩ Zl = e11 Aenn is a two-sided ideal. Obviously,
(Zl ∩ Zr )A = Un and A (Zl ∩ Zr ) = V1 . Moreover, dimD Zr = dimD Zl = n.
Theorem 4.3.2. Let A be a semiperfect ring such that the socles of all principal
right A-modules and of all principal left A-modules are simple. Suppose further-
more, that if the socles of two principal right A-modules P and P are isomorphic
then P ∼ = P . Then A satisfies the following conditions:
(i) soc AA = soc A A = Z and Z is a monomial ideal;
(ii) the ring A admits a Nakayama permutation ν = ν(A) with ν(A) = ν(Z).
Recall the notions of right and left annihilators (see vol.I, p.219). Let S be a
subset in a ring A. Then r.annA (S) = {x ∈ A : Sx = 0} and l.annA (S) = {x ∈ A :
xS = 0}. We shall write r(S) instead of r.annA (S) and l(S) instead of l.annA (S).
Remark 4.4.1. From the duality properties it is easy to see that AA is injective
if and only if A A is injective. So, the definition of quasi-Frobenius algebras is right
and left symmetric.
Examples 4.4.1.
1. Any Frobenius algebra is quasi-Frobenius, since in this case AA (A A)∗ is
injective.
2. Any group algebra kG of a finite group G over a field k is quasi-Frobenius,
by theorem 4.2.3.
Recall that a finite dimensional algebra A over a field k is called basic if the
quotient algebra Ā = A/R is a direct product of division algebras. It is equivalent
to the fact that if AA = P1 ⊕ P2 ⊕ . . . ⊕ Ps is a decomposition of AA into a direct
sum of principal modules, then Pi Pj for i = j (i, j = 1, . . . , s).
The fact that the definition of quasi-Frobenius algebra has been given in terms
of module categories implies the following proposition.
s
s
s
dim A = dim Qi = dim E(Pi ) > dim Pi = dim A.
i=1 i=1 i=1
Proof. By theorem 4.4.2 and remark 4.4.2, every QF -algebra with identity
Nakayama permutation is automatically Frobenius. Clearly, every algebra which
is Morita equivalent to a Frobenius algebra with identity Nakayama permutation
is Frobenius.
Let A be a Frobenius algebra and let ν(A) be not the identity. Then we can
assume that soc P1 = top P2 . Let A = P1n1 ⊕ P2n2 ⊕ . . . ⊕ Psns be a canonical
decomposition of A into a direct sum of non-isomorphic principal A-modules. From
the definition of a Frobenius algebra it follows that n2 = n1 . Set P = P12 ⊕ P2 ⊕
. . . ⊕ Ps . Then C = EndA P is a QF -algebra, ν(A) = ν(C), the multiplicity of
the first principal C-module is 2 and does not coincide with the multiplicity of the
second principal C-module. Therefore, C is not Frobenius.
Examples 4.4.2.
(a) Let G = { g } be a cyclic group of order 4, and let k = F2 be the field of
two elements, A = kG.
Set r = 1 + g. Then R = rA is the Jacobson radical of A, and the elements
1, r, r2 , r3 form a basis of A, i.e., A is a Köthe algebra of length 4.
FROBENIUS ALGEBRAS AND QUASI-FROBENIUS RINGS 173
It is easy to see that the algebras given in examples 4.4.2(b) and 4.4.2(c) are
non-isomorphic. Indeed, in case (b) x2 = 0 for all x ∈ R, on the other hand, in
case (c) this does not hold.
We have the following strict inclusions:2
(group algebras)⊂ (symmetric algebras)⊂
⊂ (Frobenius algebras)⊂ (quasi-Frobenius algebras).
The algebra from example 4.4.2(c) is an example of a symmetric algebra which
is not a group algebra, and so the first inclusion is strict.
Example 4.4.2(d) shows that the second inclusion is strict. From theorem 4.4.4
it follows that the third inclusion is also strict.
Example 4.4.3.
In conclusion we give an example of a semidistributive weakly symmetric al-
gebra A over the field k = F2 = {0, 1}. This algebra is a quotient algebra of the
1 see
example 4.5.1.
2 see
C.W.Curtis, I.Reiner, Methods of Representation Theory I,II. John Wiley and Sons,
New York, 1990, §66.
174 ALGEBRAS, RINGS AND MODULES
O ⊃ M ⊃ M2 ⊃ . . . ⊃ Mn ⊃ . . . ,
and, moreover, this chain is also the unique chain of left ideals of A. Then,
obviously, O is Noetherian, but not Artinian, all powers of M are distinct and
! ∞
k=1 M = 0. Moreover, M is principal as a right (left) ideal.
k
FROBENIUS ALGEBRAS AND QUASI-FROBENIUS RINGS 175
Example 4.5.1.
Denote by Hs (O) the ring of all s × s matrices of the following form:
⎛ ⎞
O O ... O O
⎜ .. .. ⎟
⎜M O . .⎟
⎜ ⎟
⎜ .. ⎟ .
H = Hs (O) = ⎜ ... ..
.
..
.
..
. .⎟
⎜ ⎟
⎜ . .. ⎟
⎝ .. . O O⎠
M M ... M O
soc P1 ∼
= top P2 , soc P2 ∼
= top P3 , . . . , soc Ps ∼
= top P1
and
top Q1 ∼
= soc Q2 , top Q2 ∼
= soc Q3 , . . . , top Qs ∼
= soc Q1 .
Example 4.5.2.
Let B2,s be the ring of 2s × 2s matrices of the form:
H R
B2,s = ,
R H
If M = AA , then soc (AA ) is the sum of all minimal right ideals of A and it is
a right ideal of A. Analogously, soc (A A) is a left ideal in A.
For a semisimple module M we have soc (M ) = M .
178 ALGEBRAS, RINGS AND MODULES
Example 4.6.1.
Let A = Tn (D) be the ring of upper triangular matrices of degree n over a
division ring D. Obviously, soc A A is the “first row ideal” of Tn (D) and soc AA
is the “last column ideal” of Tn (D). Therefore soc A A = V1n and soc AA = Unn ,
where Ui (resp. Vi ) is a right (resp. left) simple A-module (i = 1, 2, . . . , n). For
n ≥ 2 soc A A = soc AA . Obviously, dim D (soc A A) = dim D (soc AA ) = n.
Example 4.6.2.
Let k be a field and B = Bn (k) be the subalgebra of Mn (k) of matrices of the
following form:
⎛ ⎞
a11 0 0 ... 0 0
⎜ a21 a 0 ... 0 0 ⎟
⎜ 22 ⎟
⎜ a31 0 a33 ... 0 0 ⎟
⎜ ⎟
⎜ .. .. .. .. .. ..⎟.
⎜ . . . . . .⎟
⎜ ⎟
⎝an−1,1 0 0 . . . an−1,n−1 0 ⎠
an1 0 0 ... 0 ann
FROBENIUS ALGEBRAS AND QUASI-FROBENIUS RINGS 179
Obviously,
⎛ ⎞
0 0 0 ... 0
"⎜ a21 0 0 . . . 0⎟ #
⎜ ⎟
⎜ . . . 0⎟
R = rad(B) = ⎜ a31 0 0 ⎟ .
⎜ .. .. .. .. .. ⎟
⎝ . . . . .⎠
an1 0 0 ... 0
It is easy to see that
⎛ ⎞
" a11 0 ... 0 #
⎜ a21 0 . . . 0⎟
⎜ ⎟
l(R) = soc BB = ⎜ . .. . . .. ⎟
⎝ .. . . .⎠
an1 0 ... 0
and
⎛ ⎞
0 0 0 ... 0
"⎜ a21 a22 0 ... 0 ⎟#
⎜ ⎟
⎜ ⎟
r(R) = soc B B = ⎜ a31 0 a33 ... 0 ⎟ .
⎜ .. .. .. .. .. ⎟
⎝ . . . . . ⎠
an1 0 0 . . . ann
Consequently, soc BB = U1n and soc B B = V22 ⊕. . .⊕Vn2 . We have dim k (soc BB ) =
n and dim k (soc B B) = 2n − 2, where Ui (resp. Vi ) is a right (resp. left) simple
B-module (i = 1, 2, . . . , n).
Example 4.6.3.
Here we give an example of a right Noetherian serial ring A for which soc AA =
U2 ⊕ U2 , and soc A A = 0.
Let Q be the field of rational numbers, p be a prime integer, Z(p) = {m/n ∈
Q : (n, p) = 1}. Set
Z(p) Q
A= .
0 Q
pZ(p) Q
It is clear that R = radA = and P1 = (Z(p) , Q), P2 = (0, Q).
0 0
The left principal A-modules are:
Z(p) Q
Q1 = , Q2 = .
0 Q
Since A is serial, if the socle of a principal left (or right) module P is nonzero,
then soc P has to be simple.
180 ALGEBRAS, RINGS AND MODULES
Example 4.6.4.
Let Q∞ be the following countable directed graph:
1 2 3 ... n n+1
• • • ... • • ...
So, the set of vertices V Q∞ is the set of all natural integers N and the set of
arrows AQ∞ = {σk,k+1 : k ∈ N}. For all i < j, where (i, j) ∈ N × N, there is a
path pij = σi,i+1 . . . σj−1,j . Let k be a field and consider kQ∞ . This is an infinite
dimensional k-algebra with the following basis:
Example 4.6.5.
Here we give an example of a commutative local semiprimary ring A whose
socle is simple, but A is not Artinian.
Let k[x1 , x2 , . . . , xn , . . .] be the polynomial ring over a field k in a countable
number of variables and let J be the ideal of k[x1 , x2 , . . . , xn , . . .] generated by
the elements: x21 , x22 , . . . , x2n , . . . and x1 x2 − xi xj for i = j. Consider the quotient
ring A = k[x1 , x2 , . . . , xn , . . .]/J. Let π : k[x1 , x2 , . . . , xn , . . .] → A be the natural
epimorphism. Denote π(x1 x2 ) by a0 . The images of 1, x1 , x2 , . . . , . . . , xn , . . . in A
will be denoted by the same symbols.
Now π(xi xj xk ) = 0 for all i, j, k. Indeed if i = j, then π(x2i xk ) = π(x2i )π(xk ) =
0. This follows from x2i = 0. And if i = j then π(xi xj xk ) = π(x1 x2 xk ) which is
zero if k ∈ {1, 2} and which is equal to π(x1 x1 x2 ) = 0 for k ≥ 3.
FROBENIUS ALGEBRAS AND QUASI-FROBENIUS RINGS 181
Obviously the square (resp. third power) of any element of the form r =
α1 x1 + . . . + αm xm + α0 a0 is a sum of monomials of degree 2 (resp. 3). So r3 = 0
in A for all such r. It immediately follows that the Jacobson radical R of A is the
infinite dimensional vector space with basis a0 , x1 , x2 , . . . , xn , . . .. Therefore the
Loewy series of A is:
A ⊃ R ⊃ R2 ⊃ 0,
Lemma 4.7.1. Let R be the Jacobson radical of a right (or left) perfect ring A.
Suppose, that X + R2 = R for some right (or left) ideal X in R. Then X = R.
Lemma 4.7.2. If the pair (({An }), F ) has arbitrarily long paths, then it has
a path of infinite length.
Theorem 4.7.3 (B.Osofsky). Let A be a right (or left) perfect ring with
Jacobson radical R. Then A is a right Artinian if and only if the right quiver
Q(A) of A is defined, i.e., the quotient ring A/R2 is right Artinian.
Consequently,
Rk = ri1 ri2 . . . rik A. (4.7.3)
Write S = (r1 , . . . , rn ). Let An consist of all products ri0 · · · rin = 0. Consider
A ⊃ R ⊃ R2 ⊃ . . . ⊃ Rt−1 ⊃ 0.
By remark 4.7.1, all inclusions in this Loewy series are strict and all Rk are
finitely generated A-modules. Obviously, the Rk /Rk+1 are semisimple and finitely
generated for k = 1, . . . , t − 1. Therefore A has a composition series and A is a
right Artinian ring. The theorem is proved.
Corollary 4.7.4. A semidistributive right (or left) perfect ring A is right (or
left) Artinian.
The proof follows from theorem 14.1.6, vol.I, and the Osofsky theorem.
Proposition 4.7.5. Let R be the Jacobson radical and let P r(A) be the prime
radical of a ring A. If A is a one-sided perfect ring then P r(A) = R.
The proof follows from the following assertion: any element r ∈ R of a one-sided
perfect ring is strongly nilpotent.
FROBENIUS ALGEBRAS AND QUASI-FROBENIUS RINGS 183
M0 ⊂ M1 ⊂ . . . ⊂ Mα ⊂ . . .
Proof.
(1) ⇒ (2). Let A be a right (or left) perfect ring with Jacobson radical R.
Then A is semilocal, by theorem 10.5.3, lemma 10.4.9 and theorem 10.4.8, vol.I.
Let M be a left A-module and suppose that soc M = 0. If RM = 0 then M
is semisimple, by theorem 2.2.5, vol.I. Therefore, there exists r1 ∈ R such that
r1 M = 0. Let r1 m = 0. Consider Am ⊃ Ar1 m. If Ar1 m is not semisimple,
then there exists r2 such that r2 r1 · m = 0. Continuing this process we obtain a
sequence {rk } of elements of R such that rn · rn−1 . . . r1 = 0 for any n ∈ N. This
contradiction proves (1) ⇒ (2).
(2) ⇒ (1). Let 0 ⊃ R1 ⊃ . . . ⊃ Rα ⊃ . . . ⊃ Rδ = R be the transfinite ascending
Loewy series of R. Therefore, if r ∈ R, we can define h(r) to be the least $ α such
that r ∈ Rα . Note that h(r) can never be a limiting ordinal since if r ∈ Rβ ,
β<α
then r ∈ Rβ for some β < α. Thus, if r ∈ R we can write h(r) = β + 1 for some
β. Since Rβ ⊃ R · Rβ+1 , we have h(b · r) < h(r) for any b ∈ R and h(r) = 0 for
r = 0.
Suppose we have a sequence {rn } of elements of R. Then if rn rn−1 . . . r1 = 0
for all n the sequence {h(rn rn−1 . . . r1 )} is a countable strongly decreasing chain
184 ALGEBRAS, RINGS AND MODULES
of ordinals from a totally ordered ascending set of ordinals with a unique least
element. This is impossible. The theorem is proved.
Piecewise domains coincide with l-hereditary rings A, that is, rings such
that every local one-sided ideal of A is projective. From theorem 10.7.9, vol.I, we
immediately obtain the following lemma:
Proof. Let e be a local idempotent. By proposition 4.9.2, the ring eAe is a do-
main. By proposition 11.2.9 and corollary 11.2.7, vol.I, we have that P r(eAe) = 0.
By theorem 11.4.1, vol.I, P r(A) is nilpotent.
From corollary 11.8.7, vol.I, and corollary 4.9.3 we obtain the following theo-
rem.
The set eAe coincides with the prime radical of the ring eAe, by proposition 11.2.9,
vol.I. But eAe is a domain, so P r(eAe) = 0 and ēĀē = eAe. By theorem 14.4.6,
vol.I, the ring Ā is a direct product of prime rings. So, all the rings Āi are prime.
We shall show that P r(A1 ) = 0, . . . , P r(At ) = 0.
First we assume that Ā = A/P r(A) is prime. Let 1 = e1 + . . . + es be a
decomposition of 1 ∈ A into a sum of mutually orthogonal local idempotents and
%
s
P r(A) = I = ep Ieq .
p,q=1
1 2 r−1 r
• • ... • •
Corollary 4.9.7. Let t be the number of vertices in the prime quiver P Q(A)
of a semiperfect piecewise domain A. Then (P r(A))t = 0.
Theorem 4.9.8. Let A be a right hereditary semiperfect ring. Then the fol-
lowing conditions are equivalent:
(a) A contains a monomial ideal;
(b) A is isomorphic to a finite direct product of rings of the form Mnk (B),
where all the Bk are local right hereditary domains.
Moreover, if the ring A is semidistributive, then all the rings Bi are either
division rings or discrete valuation rings.
Proof.
(b) ⇒ (a). Obvious.
(a) ⇒ (b). Let A be an indecomposable ring. From lemma 4.9.1 it fol-
lows that A is a piecewise domain. Every semiperfect piecewise domain has a
triangular Peirce decomposition (see formula (4.9.1)). So A is a prime right
hereditary semiperfect ring. Since A contains a monomial ideal, A = P n and
A Mn (EndA P ), where B = EndA P is a local right hereditary domain.
If a ring A is semidistributive, then B is a right Noetherian right hereditary
local semidistributive domain. By proposition 14.4.10 and corollary 14.4.11, vol.I,
B is either a division ring or a discrete valuation ring.
Corollary 4.9.9. Let A be a right Artinian and right hereditary ring with a
monomial ideal. Then A is semisimple.
Lemma 4.10.2. Let A be a right Noetherian ring. Then the dual to any
finitely generated left A-module is also finitely generated.
Let M be a right A-module with dual module M ∗ . Then M ∗ itself has a dual
module M ∗∗ . Suppose m ∈ M and f ∈ M ∗ = HomA (M, A). Define a mapping
ϕm : M ∗ → A
by ϕm (f ) = f (m). Obviously,
δM : M → M ∗∗ (4.10.2)
Note that any finite dimensional vector space as a module over its field k is
reflexive; on the other hand an infinite dimensional vector space over a field k is
never reflexive.3
Proof. Let F be a free module with finite free basis f1 , f2 , . . . , fn and let
ϕ1 , ϕ2 , . . . , ϕn be the free basis of F ∗ dual to f1 , f2 , . . . , fn . Let ψ1 , ψ2 , . . . , ψn be
a basis of F ∗∗ dual to the basis ϕ1 , ϕ2 , . . . , ϕn . Then δF (fi ) and ψi both belong
to HomA (F ∗ , A) and
theory. For free Abelian groups (as modules over Z) the answer is as follows: A free Abelian
group is reflexive if and only if its cardinality is non-ω-measurable. In particular a countable
free Abelian group is reflexive. See P.C.Eklof, A.H.Mekler, Almost free modules. Set-theoretic
methods, North Holland, 1990 for this and much more.
190 ALGEBRAS, RINGS AND MODULES
Lemma 4.10.6. Let A be a right Noetherian ring. Then any finitely generated
submodule of a free module with a finite basis is semi-reflexive.
Proposition 4.11.3. Let A be a semiperfect ring with duality for simple mod-
ules, and let P be a simple projective A-module. Then A = Mn1 (D) × A2 , where
D = EndA P . Conversely, if A = Mn1 (D), where D is a division ring, then
U ∗ = V and V ∗ = U , where U is the unique simple right A-module and V is the
unique simple left A-module.
Observe that the dual of a right A-module eAeJ is isomorphic to the left A-
module l(J)e.
Examples 4.12.1.
1. Each Frobenius algebra and each quasi-Frobenius algebra is a self-injective
ring.
2. The ring of integers Z is not self-injective.
3. The ring Z/nZ, for n > 0 is self-injective.
4. A semisimple Artinian ring is right and left self-injective.
Examples 4.12.2.
Let T = { ai xi : ai ∈ K} where α ⊂ Q≥0 and the exponent set α is such
i∈α
that each lower limit point is in α (i.e., every subset of α has a smallest element
(in α)). There is a (nondiscrete) valuation on T obtained by
v( ai xi ) = smallest j in α with aj = 0.
i∈α
The next statement gives connections between a self-injective ring and annihi-
lators of its ideals.
l(r(H)) = H. (4.12.2)
n
l(r(H)) = l(r(Ahi )).
i=1
Corollary 4.12.4. If a ring A is right (resp. left) Noetherian and right (resp.
left) self-injective then soc (A A) = 0 (resp. soc (AA ) = 0).
and
ϕ(xn+1 an+1 ) = b2 xn+1 an+1 .
Taking into account (4.12.1) we have
n
n
b1 − b2 ∈ l xi A ∩ xn+1 A = l xi A + l(xn+1 A).
i=1 i=1
n
So there are elements y ∈ l( xi A) and z ∈ l(xn+1 A) such that b1 − b2 = y − z.
i=1
Set b = b1 − y = b2 − z. Then
n+1
n
ϕ xi ai = ϕ xi ai + ϕ(xn+1 an+1 ) =
i=1 i=1
n
n+1
= (b1 − y) xi ai + (b2 − z)xn+1 an+1 = b xi ai ,
i=1 i=1
as required.
Proof. Since A is right Noetherian, each of its right ideals is finitely generated.
Taking into account the Baer criterion we obtain the statement as a corollary of
propositions 4.12.1 and 4.12.5.
Proof. Since A is right self-injective ring, l(r(L)) = L for all left finitely
generated ideals L of A, by proposition 4.12.1. This condition implies that if we
have some infinite descending sequence of left finitely generated ideals
L1 ⊃ L2 ⊃ . . . ⊃ Ln ⊃ . . .
then we have also an infinite ascending sequence of right ideals
r(L1 ) ⊂ r(L2 ) ⊂ . . . ⊂ r(Ln ) ⊂ . . .
Since A is right Noetherian the last sequence must stabilize and so A satisfies
the d.c.c. for all finitely generated left ideals and, in particular, for all principal
left ideals. From theorem 10.5.5, vol.I, it follows that A is a right perfect ring.
Therefore A/R is semisimple and R = radA is right T -nilpotent. By corollary
4.12.9 and theorem 4.12.7, A is a right Artinian ring and R is nilpotent. Now
we shall show that A is also a left Noetherian ring. As we have shown above A
satisfies the d.c.c. for all finitely generated left ideals. Suppose there is a left ideal
L in A which is not finitely generated. Then for every finitely generated left ideal
H ⊂ L there is a finitely generated ideal H1 such that H ⊂ H1 ⊂ L. Then we can
build by induction an infinite strictly ascending chain of finitely generated ideals
of A. This contradiction shows that A is a left Noetherian ring. Since A/R is
semisimple and R is nilpotent, A is left Artinian, by corollary 4.12.8. Thus, A is
a two-sided Artinian.
Proof. Since A is a right Noetherian ring, any of its right ideals is finitely
generated. Then from proposition 4.12.5 and the Baer criterion it follows that A
is a right self-injective ring. And so, by lemma 4.12.10, A is two-sided Artinian.
Proof. Suppose A is a right Noetherian ring and r(l(I)) = I for all two-sided
ideals I of A. Let R = radA. Consider the descending chain of ideals
R ⊇ R2 ⊇ R3 ⊇ . . .
Since A is right Noetherian, the last chain is finite, i.e., there is an integer n such
that l(Rn ) = l(Rn+1 ). Hence Rn = r(l(Rn )) = r(l(Rn+1 )) = Rn+1 . Since A is
right Noetherian, Rn is a right finitely generated ideal, and, by the Nakayama
lemma (lemma 3.4.11, vol.I), Rn = 0.
with the top row exact. Since A is self-injective, this diagram can be completed
to a commutative diagram
ψ
0 −→ U −→ M
h
↓ϕ
AA
Theorem 4.12.14. For any ring A the following conditions are equivalent:
(1) A is a right Noetherian and right self-injective ring.
(2) A is two-sided Artinian and satisfies the following double annihilator
conditions:
(2a) r(l(H)) = H for any right ideal H
(2b) l(r(L)) = L for any left ideal L.
Proof.
2) ⇒ 1) Since A is a two-sided Artinian ring, it is also two-sided Noetherian
by theorem 3.5.6, vol.I.
Let H1 , H2 be right ideals of A. Then from condition (2a) it follows that
Theorem 4.12.14*. For any ring A the following conditions are equivalent:
1) A is right Noetherian and right self-injective.
2) A is left Noetherian and left self-injective.
200 ALGEBRAS, RINGS AND MODULES
Remark 4.12.2. Note that in the case if A is neither left Noetherian nor right
Noetherian self-injectivity is not necessarily left-right symmetric. For example,
the endomorphism ring of any infinitely generated free left module over a quasi-
Frobenius k-algebra is left self-injective but not right self-injective (see [Osofsky,
1966], [Sandomerski, 1970]. On the other hand there are right and left self-injective
ring which are not Artinian (see [Goodearl, 1974]).
In section 4.11 we have introduced rings with duality for simple modules (DSM-
rings). By theorem 4.11.5, any quasi-Frobenius ring is a DSM-ring. Now we shall
show that any two-sided Artinian DSM-ring satisfies the two double annihilator
conditions.
0 = K0 ⊆ K1 ⊆ . . . ⊆ Kn = A (4.12.3)
By assumption, each (Ki+1 /Ki )∗ is simple. So, by lemma 4.12.16,
l(Ki )/l(Ki+1 ) is either zero or simple.
Thus
0 = l(Kn ) ⊆ . . . ⊆ l(Ki ) ⊆ l(K0 ) = A (4.12.4)
is a series of submodules with either simple or zero factors.
Consequently, length(A A) ≤ length(AA ). By symmetry we have length(AA ) ≤
length(A A). Therefore l(Ki ) = l(Ki+1 ) for i = 0, . . . , n. Applying the right
annihilator to (4.12.4) we obtain that 0 = r(l(K0 )) ⊆ . . . ⊆ r(l(Ki )) ⊆ . . . ⊆
r(l(Kn )) = A is also a composition series for AA . Obviously, Ki ⊆ r(l(Ki )) and
Ki = r(l(Ki )). Any right ideal H may be a member of a composition series so
we have proved the double annihilator property for right ideals. By symmetry the
same hold for left ideals. The proposition is proved.
FROBENIUS ALGEBRAS AND QUASI-FROBENIUS RINGS 201
Proof. The conditions (2), (3) and (4) are equivalent by theorem 4.12.14*. So
assume that A is two-sided Artinian and right and left self-injective. Let P be a
principal right A-module. Then it contains a simple right A-module U . Since P
is a direct summand of AA and A is self-injective, P is injective as well. So P is
an indecomposable injective A-module, and then from proposition 5.3.6, vol.I, it
follows that P E(U ). So U is an essential simple module in P , therefore U =
soc P . If P and P are principal indecomposable A-modules then their injectivity
implies that soc P soc P if and only if P P . Thus, every simple right A-
module is isomorphic to the socle of some right indecomposable injective module.
Analogously, if Q and Q are left principal indecomposable then soc Q and soc Q
are simple and A is quasi-Frobenius, by theorem 4.5.2.
Let A be a quasi-Frobenius ring. Then A is a DSM-ring and, by proposition
4.12.15, A satisfies the two double annihilator conditions, i.e., we proved (1) ⇒ (2).
Therefore the theorem is proved.
Lemma 4.12.18. Let A be a right Noetherian ring. Then any injective right
A-module M is a direct sum of indecomposable injective submodules.
Proof. Let A be a right Noetherian ring. We first show that any injective right
A-module Q contains an indecomposable injective submodule. Let x ∈ Q. Then
it suffices to consider the case when Q is an injective hull of xA, i.e., Q = E(xA).
Since A is right Noetherian, xA cannot contain an infinite direct sum. Since Q is
an essential extension of xA, Q also cannot contain an infinite direct sum as well.
Hence it follows that Q contains an indecomposable injective submodule.
Let M be an injective right A-module. Consider a set of all indecomposable
injective submodule of M whose sum is direct. By Zorn’s lemma, there exists such
a set {Mi : i ∈ I} which is maximal. From theorem 5.2.12, vol.I, it follows that
⊕ Mi is an injective module, and so M = X ⊕ ⊕ Mi , where X is some submodule
i i
of M , by propositions 5.2.2 and 5.3.6, vol.I. Since M is injective, X is injective as
well, and it contains an indecomposable injective submodule. From the maximality
of ⊕ Mi it follows that X = 0. So M = ⊕ Mi as required.
i i
202 ALGEBRAS, RINGS AND MODULES
Proof.
1) ⇒ 2) Let A be a quasi-Frobenius ring, then, by theorem 4.12.16, A is right
and left self-injective. So every free module is injective. Hence every projective
module is injective as well.
Conversely, suppose M is an injective module. Since A is a Noetherian ring, by
lemma 4.12.17, M ⊕ Mi , where the Mi are indecomposable injective modules.
i
By lemma 4.12.13, Mi may be embedded in the injective module AA . Therefore
Mi is projective, and so is M .
2) ⇒ 1) Conversely, since A itself is projective, we obtain that AA is injective.
Since any direct sum of projective modules is projective, by proposition 5.1.4,
vol.I, and any projective module is injective, any direct sum of injective module
is injective. Then, by theorem 5.2.12, vol.I, A is right Noetherian. Therefore, by
theorem 4.12.17, A is quasi-Frobenius.
In section 4.10 we have studied reflexive modules. The following theorem gives
a characterization of finitely generated modules over quasi-Frobenius rings.
0 −→ N −→ F −→ M −→ 0
↓ δN ↓ δF ↓ δM
0 −→ N ∗∗ −→ F ∗∗ −→ M ∗∗ −→ 0
Proof. Assume that Q is not strongly connected. Then there exist vertices k
and l such that there is no path from k to l. Hence l ∈ Q(k). Let T be the set of ver-
tices of Q that do not belong to Q(k). Since Q is connected, there exists i ∈ T and
FROBENIUS ALGEBRAS AND QUASI-FROBENIUS RINGS 205
an arrow σij such that j ∈ Q(k). Clearly, Q(j) ⊂ Q(k). Let Q(k) = {1, 2, . . . , m}.
Then from the property (β) it follows that Q(k) = {π(1), π(2), . . . , π(m)} and (α)
implies that π(i) ∈ Q(k). But i ∈ T and so i ∈ Q(k) and some vertex from Q(k)
is mapped to π(i). The obtained contradiction completes the proof of the lemma.
Proof. Since all rings Morita equivalent to a QF -ring are QF -rings, we can
assume that the quasi-Frobenius ring A is basic. Since a QF -ring is a two-sided
Artinian ring, we can consider its quiver Q. Then the proof of this theorem
follows from lemma 4.13.1.
Lemma 4.14.1. Let Q be a connected quiver with at least one arrow. The
following statements are equivalent:
(i) Q is strongly connected;
(ii) there is a cycle c = (σ1 , . . . , σm ) such that every arrow σ of Q occurs as
some σi .
A
where ϕ is the canonical ring surjection. Let C = ϕ−1 (b). Then C is a left ideal in
Ω and C ⊆ Ker μ. Obviously, C ⊃ I(μ). We shall show that CΩ ⊆ I(μ).
Let w ∈ C. Then Ωw ⊆ C ⊆ Ker μ. By definition of the Frobenius system
(S, μ), we have wΩ ⊆ Ker μ. Therefore, CΩ ⊆ I(μ) and ϕ(C) = 0 = ϕϕ−1 (b) = b.
A similar proof shows that if b is a right ideal in A and b ⊆ Ker μ̄ then b = 0.
Therefore, A is a Frobenius algebra, by theorem 4.2.1.
FROBENIUS ALGEBRAS AND QUASI-FROBENIUS RINGS 207
c, or
f ∈S ⇔ f =
(σj , . . . , σm , σ1 , . . . , σj−1 ) for some j = 2, . . . , m.
Note that S is a finite set.
Define a set map: μ : B → k by the formula
1 if and only if f ∈ S
μ(f ) = (4.14.1)
0 otherwise
Extend this map to a k-linear map Ω → k by “linearity”, i.e., for w ∈ Ω, w =
n
α1 b1 + . . . + αn bn we set μ(w) = αi μ(bi ).
i=1
Theorem 4.14.5. Let Q be a strongly connected quiver. Then for any field k
there is a symmetric k-algebra A such that Q = Q(A).
n
t = αi εi + ασ σ + βi h i ,
i=1 σ∈AQ i∈V Q
μ(tf − α1 ε1 f ) = 0.
μ(tf ) = α1= 0 and t ∈ I(μ). Consequently we can assume that
So,
t = ασ σ + βi hi . Let ασ = 0 for some σ. We can assume that
σ∈AQ i∈V Q
c = (σ, σ2 , . . . , σm ) is a cycle. Let f ∗ = σ2 . . . σm and f = σ1 . . . σm . Obviously,
μ(tf ∗ ) = ασ + α τ μ(τ σ2 . . . σm ),
τ =σ
Proof. Let us view M as a left End(M )-module. there exists an exact sequence
End(M )J End(M )I M 0.
Apply to this sequence the left exact functor HomEnd(M) (∗, M ). We obtain
0 A(M ) MI ;
FROBENIUS ALGEBRAS AND QUASI-FROBENIUS RINGS 209
consequently,
ϕ
0 AM MI .
We now show that AM can be embedded in a direct sum of a finite number of
copies of M . Let ϕi , i ∈ I, denote!the projection of ϕ on the i-th coordinate of
M I . Since ϕ is a monomorphism, Ker ϕi = 0. Since the ring AM is Artinian,
i∈I
!
n
there exist i1 , . . . , in such that Ker ϕik = 0. Therefore, AM can be embedded
i∈I
in M n .
i
0 N N
α
M1
M̄
ψ
π
N N 0
π1
0 M2 M M̄ 0
ψ
π
N N 0
To prove the reverse implication, we shall need the following simple lemma.
HomA (Pj , P1 ) = 0, then the composition series of P1 has a simple factor iso-
morphic to Uj , and this is a contradiction. Since A is indecomposable as a ring,
it follows that A = P1n1 . Arguing in just the same way for left modules, we find
that A is uniserial. This completes the proof of theorem 4.15.5.
Artinian ring A is quasi-Frobenius if and only if A is a ring with duality for simple
modules was proved by C.Curtis, J.Reiner [Curtis, Reiner, 1962].
B.Osofsky proved in the paper [Osofsky, 1966a] the following equivalent defi-
nition for QF-rings: A is a QF-ring if and only if A is right (left) perfect and right
and left semi-injective.
It is well known that if A is a QF-ring, then the ring Mn (A), the ring of n × n
matrices over A, and AG, the group ring over A for any finite group G, are both
QF-rings. The construction of QF-rings in the general case has been studied by
T.A.Hannula in the paper [Hannula, 1973].
In section 4.5, which is devoted to duality in Noetherian rings, we follow to a
considerable extent the book by D.G.Northcott [Northcott, 1973].
Rings with duality for simple modules were considered in [Dieudonné, 1958]
and in [Morita, Tachikawa, 1956]. These rings were also studied in [Curtis, Reiner,
1962]. M.A.Dokuchaev and V.V.Kirichenko studied semiperfect rings with duality
for simple modules (see [Dokuchaev, Kirichenko, 2002]. For an equivalent formula-
tion of the Osofsky theorem see the paper [Osofsky, 1966a]. In this paper she also
gave the following equivalent definition for QF-rings: a ring A is quasi-Frobenius
if and only if A is right (left) perfect and right and left semi-injective.
The notion of a symmetric algebra was introduced by R.Brauer and C.Nesbitt
(see [Brauer, Nesbitt, 1937], [Nesbitt, 1938]). The properties of symmetric alge-
bras were studied by T.Nakayama [Nakayama, 1939]. The structure of symmetric
algebras was studied by H.Kupisch (see [Kupisch, 1965] and [Kupisch, 1970]).
Cohomological dimension of Frobenius algebras and quasi-Frobenius rings
has been considered in the paper of S.Eilenberg and T.Nakayama [Eilenberg,
Nakayama, 1955]. In this paper they proved that for Frobenius algebras the
cohomological dimension is either 0 or ∞. They also proved that for right or
left Noetherian rings the notions “quasi-Frobenius ring” and “left self-injective
ring” are equivalent. In our proof of this theorem (theorem 4.12.17 in this chap-
ter) we follow to a considerable extent F.Kasch [Kasch, 1982]. For left Noetherian
rings which are left self-injective Eilenberg and Nakayama proved that their left
global dimension is either 0 or ∞. Corollary 4.12.20 in this chapter, which states
that the global dimension of a quasi-Frobenius ring is equal to 0 or ∞, was proved
by M.Auslander in his paper [Auslander, 1955].
Piecewise domains were first considered by R.Gordon, L.W.Small in [Gordon,
Small, 1972].
Theorem 4.9.5 was proved in [Kirichenko, 1993].
Theorem 4.13.2 and theorem 4.14.5 first were proved by E.Green in the paper
[Green, 1978].
The classification of quasi-Frobenius algebras of finite representation type in
terms of Dynkin diagrams was obtained by C.Riedtmann (see [Riedtmann, 1980a]):
214 ALGEBRAS, RINGS AND MODULES
Proof. The necessity of this statement follows from proposition 6.5.4, vol.I.
To prove sufficiency we consider an exact sequence
0 → M → P → X, (5.1.1)
219
220 ALGEBRAS, RINGS AND MODULES
Proof. F is a free module with a basis {ei : i ∈ I}, and P is flat. For
a given x ∈ X we have x = a1 ei1 + a2 ei2 + . . . + am eim , where ai ∈ A. Let
I = a1 A + a2 A + . . . + am A. Since P is flat, we have x ∈ X ∩ F I = XI, by
proposition5.4.12, vol.I. Therefore
x = xj ∈ X and cj ∈ I. Now
xj cj , where
each cj = ai bij , so that x = ai xi , where xi = xj bij . Define θ : F → X by
i i j
θ(eik ) = xk , while θ sends all the other basis elements of F into 0. Then
θ(x) = θ( ak eik ) = ak θ(eik ) = ak xk = x.
k k k
As was shown earlier (see corollary 5.4.5, vol.I), every projective module is flat.
The converse to this statement is not true in general. The following statement gives
an example of a case where a flat module is always projective.
0 −→ Fn −→ Pn−1 −→ . . . −→ P1 −→ P0 −→ X −→ 0, (5.1.2)
where Fn is flat, all Pi are projective and there is no shorter such sequence.
We set w.dimA M = ∞ if there is no n with w.dimA X ≤ n.
Proof. Since TorA n+1 (X, Y ) can be computed by using the projective
resolution (5.1.3) of X and TorA n (Ker d0 , Y ) can be computed by using the as-
sociated projective resolution of Ker d0 (where d0 = π):
. . . −→ Pk −→ Pk−1 −→ . . . −→ P1 −→ Ker d0 −→ 0,
n+1 (X, Y ) Torn (Ker d0 , Y ) Torn−1 (Ker d1 , Y ) ... Tor1 (Ker dn−1 , Y )
TorA A A A
Proof.
1) ⇒ 2) and 2) ⇒ 3) are trivial.
3) ⇒ 4). Consider a projective resolution of X
d
2 1 d π
. . . −→ P2 −→ P1 −→ P0 −→ X −→ 0
with projective modules P0 , P1 ,..., Pn−1 , and a flat module Ker dn−1 . Hence
w.dimA X ≤ n.
Proof. This follows immediately from corollary 5.1.6 and its analog for
l.w.gl.dim A.
Theorem 5.1.8. For any ring A w.gl.dim A ≤ min {r.gl.dim A, l.gl.dim A}.
We shall also show that there holds an analogous statement for projective
global dimension for an arbitrary ring.
The following statements are dual to corresponding statements on projective
modules over arbitrary rings as given in section 6.5, vol.I.
be an injective resolution of the right A-module M . Then for any right A-module N
1
A (N, M ) ExtA (N, Im dn−1 )
Extn+1
(3) inj.dimA M ≤ n if and only if ExtiA (N, M ) = 0 for all i > n and all right
A-modules N .
Remark 5.1.1. Thus, propositions 5.1.9 and 5.1.13 say that to compute the
global dimension and weak dimension of a ring, it suffices to know the dimensions
of the cyclic modules, and, in particular, the dimensions of finitely generated
modules. In the general case the weak dimension and global dimension of a ring
are distinct things, but for a right Noetherian ring they are the same, as we shall
show below.
0 → K1 → P1 → K0 → 0,
where P1 a finitely generated free module. The usual iteration gives a pro-
jective resolution in which each Pi is a finitely generated free module. Since
TorAn+1 (X, Y ) = 0 for all left A-modules Y , we have Tor1 (Ker dn−1 , Y ) =
TorAn+1 (X, Y ) = 0 for all B. Therefore, by proposition 6.3.7, vol.I, Ker dn−1 is
flat. Thus Ker dn−1 is a finitely generated flat right A-module. Then, by propo-
sition 5.1.3, Ker dn−1 is projective. Therefore Ext1A (Ker dn−1 , B) = 0 for all right
A-modules B, hence Extn+1 A (X, B) = 0, by lemma 6.5.3, vol.I.
w.gl.dim A = r.gl.dim A.
w.gl.dim A = l.gl.dim A.
r.gl.dim A = l.gl.dim A.
structure theorem for these rings and described most of them in terms of quasi-
matrix rings over division rings (see [Murase, 1963a], [Murase, 1963b], [Murase,
1964]). Serial non-Artinian rings were studied and described by R.B.Warfield and
V.V.Kirichenko. In particular, they gave a full description of the structure of serial
Noetherian rings. Most of these results are described in chapters 12, 13 of vol.I of
this book.
Let A be a right Artinian ring with Jacobson radical R. Then for this ring
we can define the right quiver Q(A) of A (see chapter 11, vol.I). Recall the defini-
tion. Let P1 , P2 , . . . , Pn be all pairwise non-isomorphic principal right A-modules.
Consider the projective cover of Ri = Pi R (i = 1, ..., s), which we shall denote by
s t
P (Ri ). Let P (Ri ) = ⊕ Pj ij . We assign to the principal modules P1 , . . . , Ps the
j=1
vertices 1, . . . , s in the plane and join the vertex i with the vertex j by tij arrows.
The so constructed graph is called the right quiver of the ring A and denoted by
Q(A). From the definition of a projective cover it follows that Q(A) = Q(A/R2 ).
Theorem 11.1.9, vol.I, gives us that if A is a right Artinian ring then the
following conditions are equivalent:
1) A is an indecomposable ring;
2) A/R2 is an indecomposable ring;
3) the quiver of A is connected.
Remark 5.2.2. If a ring A is left Artinian we can analogously construct the left
quiver Q (A). Note that there is a bijection between the right principal A-modules
and the left principal A-modules which is given by a mapping ϕ : Pi → Pi∗ , where
ϕ(Pi ) = HomA (Pi , A), and, moreover, HomA (Pj , Pi ) HomA (Pi∗ , Pj∗ ). Therefore
if we have an arrow from the vertex i to the vertex j in the quiver Q(A), then there
is an arrow from the vertex j to the vertex i in the quiver Q (A). In particular,
both quivers are connected or disconnected simultaneously.
If A is a finite dimensional algebra over an algebraically closed field k, then
the number of arrows from the vertex i to the vertex j in the quiver Q(A) is equal
to the number of arrows from the vertex j to the vertex i in the quiver Q (A).
However, this is not true in general.
In this section we shall study the structure of right Artinian right serial rings
in terms of quivers. If a right Artinian ring A is right serial, then every one of its
right principal module is uniserial, i.e., the lattice of its submodules is linear.
226 ALGEBRAS, RINGS AND MODULES
Theorem 5.2.1. A ring A is right serial if and only if each vertex of the
quiver Q(A) is the start of at most one arrow.
Proof. Let A be a right serial ring with Jacobson radical R, and let
P1 , P2 , . . . , Pn be all pairwise nonisomorphic principal right A-modules. Then
Ri = Pi R = 0 or Ri R is a unique maximal submodule in Ri . In the first case
there is no arrow which starts at the vertex i, and in the second case Ri /Ri R is a
simple module, i.e., there is exactly one arrow which starts at the vertex i.
Conversely, suppose that each vertex of a quiver Q(A) is the start of at most
one arrow. Then for any i ∈ V Q we have either Ri = 0 or Ri R is the unique
maximal submodule in Ri . We shall show by induction on k that if Pi Rk = 0,
then Pi Rk+1 is a unique maximal submodule in Pi Rk . For k = 1 the statement
is true. Suppose the statement is true for k − 1, i.e., Pi Rk is the unique maximal
submodule in Pi Rk−1 . Then Pi Rk is a quotient module of Ri and it has exactly
one maximal submodule which is Pi Rk+1 . This equivalent to the fact that Pi is a
uniserial module.
Example 5.2.1.
A ring with quiver
•
• • • •
or
•
• • •
•
is right serial.
Taking into account remark 5.2.2 we have the analogous theorem for left serial
rings:
Theorem 5.2.1*. A ring A is left serial if and only if each vertex of the quiver
Q(A) is the end of at most one arrow.
Example 5.2.2.
A ring with the quiver
• • • •
or
RIGHT SERIAL RINGS 227
•
• • • •
•
is left serial.
Corollary 5.2.2. A ring A is right (left) serial if and only if A/R2 is right
(left) serial.
Example 5.2.3.
A ring with quiver
•
•
•
• • • or • • •
•
•
Remark 5.2.3. From corollary 5.2.3 and corollary 5.2.3* it follows that the
quiver of an indecomposable serial ring is a cycle or a chain. In this case all
division rings entering into the Wedderburn-Artin decomposition of the ring A/R
are isomorphic. If A is a finite dimensional algebra over a field k these conditions
are also sufficient for A to be a serial ring. The following example shows that this
is not true in general.
Example 5.2.2.
Consider the set A of all matrices of the form
f (x2 ) g(x)
0 h(x),
where the f, g, h are arbitrary rational functions over a field k. Then A forms an
infinite dimensional algebra for which the conditions above hold, but this algebra
is not serial.
(Note that it is sufficient to verify this for minimal idempotents from a fixed
decomposition of the identity of B.)
Corollary 5.2.4. Let B = A/R and V = R/R2 . Then the ring A is right
serial if and only if the B-bimodule V is right serial.
Proof. The proof follows immediately from corollary 5.2.4 and theorem 2.2.2.
Suppose that A = T(V )/I T(V1 )/I1 , where V1 is a right serial bimodule
over a semisimple Artinian ring B1 , and I1 is an essential ideal in T(V1 ). Then
B A/R B1 and V R/R2 V1 (where R = radA). Moreover, if π :
T(V ) → A and π1 : T(V1 ) → A are the natural epimorphisms onto the quotient
rings, then there are unique isomorphisms ϕ : B → B1 and f : V → V1 such
that π(b) ≡ π1 ϕ(b)(mod R) and π(v) ≡ π1 f (v)(mod R2 ). Hence it follows that
ϕ and f are compatible, that is, f (bv) = ϕ(b)f (v), f (vb) = f (v)ϕ(b) for any
b ∈ B and v ∈ V . Besides, if U1 , . . . , Us are the simple B-modules, and U1 , . . . , Us
are the simple B1 -modules, while T1 , ..., Ts and T1 , . . . , Ts are the corresponding
components of T(V ) and T(V1 ); li = l(Ti /Ti I); li = l(Ti /Ti I); then li = li if
ϕ(Ui ) Ui . (Obviously, li = li = l(Pi ), where Pi = π(Ti ) π1 (Ti )).
Conversely, if ϕ : B → B1 and f : V → V1 are compatible isomorphisms, they
induce a ring isomorphism F : T(V ) → T(V1 ) and an isomorphism of quivers:
σ : Q → Q1 , where Q = Q(T(V )/J 2 ), Q1 = Q(T(V1 )/J12 ). Let I be an essential
ideal in T(V ) defined by the set of numbers {l1 , . . . , ls }. Then I1 = F (I) is an
essential ideal in T (V1 ) defined by the set of numbers {l1 , . . . , ls }, where li = lj
if σ(j) = i. This analysis together with theorem 5.2.5 gives a full description of
Wedderburn right serial rings.
Two tuples {Q; n1 , . . . , ns ; l1 , . . . , ls } and {Q1 ; n1 , . . . , ns ; l1 , . . . , ls } define
isomorphic algebras if and only if there is an isomorphism of quivers σ : Q → Q1
such that if σ(j) = i then li = lj and ni = nj .
Proof. This follows from the fact that in this case the functor HomA (∗, A) is
exact and it establishes a duality of categories of right and left finite generated
A-modules by theorem 4.12.21.
l1 ≤ l2 ≤ . . . ≤ ls ≤ l1 ,
Proposition 5.4.4. Let A be a right Artinian ring. Then the following con-
ditions are equivalent for the ring A:
232 ALGEBRAS, RINGS AND MODULES
Proof.
1) ⇐⇒ 2). This follows from theorem 5.5.6, vol.I.
2) ⇒ 3) is trivial.
3) ⇒ 4). Let A = P1 ⊕ P2 ⊕ . . . ⊕ Pn be the decomposition of the right regular
module AA into a direct sum of principal right A-modules. Then, by proposition
3.4.3, vol.I, rad A = R1 ⊕ R2 ⊕ . . . ⊕ Rn , where Ri is the radical of Pi . Since all
the Pi are projective, rad A is projective as well.
4) ⇒ 1). Let R = rad A be a projective right A-module, i.e., r.proj.dimA R = 0.
Consider the exact sequence 0 → R → A → A/R → 0. Then, by proposition
6.5.1, vol.I, r.proj.dimA (A/R) = r.proj.dimA (R) + 1 = 1. We shall show that
r.gl.dim A ≤ r.proj.dimA (A/R) = 1, which means, by proposition 6.6.6, vol.I,
that A is right hereditary. Since A/R is a semisimple A-module, it is a direct sum
of simple modules. So,
Proposition 5.4.5. If a right serial ring A is right hereditary, then the quiver
Q(A) does not contain oriented cycles.
Proof. Let Q = Q(A) be the quiver of a right hereditary and right serial ring A.
If there is an arrow σ ∈ AQ with s(σ) = i and e(σ) = j, then there is a non-zero
homomorphism fσ : Pi → Pj and Im (fσ ) ⊂ rad (Pj ). Suppose Q(A) contains an
oriented cycle and let p = σ1 σ2 . . . σm be a path of Q with start and end at the
vertex i. Then f = fσ1 fσ2 . . . fσm is a homomorphism from Pi to Pi . Since all
Pi are indecomposable and projective, by lemma 5.5.8, vol.I, all homomorphisms
fσk are monomorphisms. So f is also a monomorphism, that is, Pi Im (f ). But
Im (f ) ⊂ rad (Pi ) and rad (Pi ) = Pi , by proposition 5.1.8, vol.I. This contradiction
proves the proposition.
di = tij dj + 1. This function is well-defined. Denote by l(M ) the length of a
module M .
Theorem 5.4.6. A right Artinian right serial ring A is right hereditary if and
only if the quiver Q(A) does not contain oriented cycles and li = l(Pi ) = di for
each principal right A-module Pi .
Proof. Let A be a right Artinian right serial ring. If A is a hereditary ring, then,
by proposition 5.4.5, Q(A) does not contain oriented cycles. For any principal
right A-module Pi , Ri = rad (Pi ) is a projective A-module, as well. Therefore
t
Ri P (Ri ) = ⊕ Pj ij and li = tij lj + 1. If i ∈ Q(A) is an end vertex, then Pi
j
is a simple module, and so li = 1. Hence it follows that li = di .
Conversely, let A be a right Artinian right serial ring and let the
quiver Q(A)
have no oriented cycles and li = di for all i. Then l(Ri ) = li − 1 = tij lj + 1 =
l(P (Ri )), therefore Ri P (Ri ) is a projective module. Therefore R = rad A is
also a projective module, since it is a direct sum of projective modules Pi . Thus
A is a right hereditary ring, by proposition 5.4.4.
Corollary 5.4.7. A right Artinian right serial ring A is right hereditary if and
only if Q(A) is a disjoint union of trees with only one end vertex and li = l(Pi ) is
one more than the length of the (unique) path from the vertex i to the end vertex
(for all i).
Theorem 5.4.6 and corollary 5.4.7 give a full description of right Artinian right
serial hereditary rings. Moreover, theorem 5.4.6 and a simple calculation of the
length of the principal T(V )-modules give a description of all Wedderburn hered-
itary rings.
Proof. Let I1 and I2 be two nonzero ideals from the ring A such that I1 I2 = 0.
Since A is right serial ring, we can assume that I1 ⊃ I2 . But then I22 = 0 and
from the semiprimality of A it follows that I2 = 0. A contradiction.
234 ALGEBRAS, RINGS AND MODULES
Proof. Since a right serial ring is semiperfect, the proof immediately follows
from lemma 5.5.1, proposition 9.2.13, vol.I, and theorem 14.4.6, vol.I.
K ∩ (L + N ) = K ∩ L + K ∩ N
Proof. It is obvious, that in all cases we can consider that the ring A is reduced
and indecomposable.
(a) ⇒ (b). By theorem 14.5.1, vol.I, the ring A can be considered to be a
division ring or a ring of the form Λ = {O, E(Λ)}. Obviously, a division ring is
a hereditary ring. Since Λ is a reduced ring, the matrix E(Λ) has no symmetric
zeros, therefore the first row corresponding to the first indecomposable projective
Λ-module P1 can be made zero. The module P1 R is projective (it contains a
unique maximal submodule since the ring Λ is right serial) and so it coincides with
P2 = (α21 , 0, . . . , 0) = (1, 0, . . . , 0); the module P2 R = P3 = (1, 2, 0, . . . , 0) and
continuing this process we obtain that Ps−1 R = Ps = (1, 1, . . . , 1, 0). Therefore
the matrix E(Λ) coincides with a matrix of the following form:
⎛ ⎞
0 0 0 ... 0 0
⎜1 0 0 . . . 0 0⎟
⎜ ⎟
⎜1 1 0 . . . 0 0⎟
⎜ ⎟
⎜ .. .. .. . . .. .. ⎟
⎜. . . . . .⎟
⎜ ⎟
⎝1 1 1 . . . 0 0⎠
1 1 1 ... 1 0
Thus, A is two-sided hereditary, by corollary 12.3.7, vol.I.
(b) ⇒ (c) follows from theorem 14.5.1, vol.I, and theorem 5.5.3.
(c) ⇒ (d) follows from theorem 14.5.1, vol.I, and corollary 12.3.7, vol.I.
(d) ⇒ (a) follows from theorem 14.2.1, vol.I, and corollary 12.3.7, vol.I.
Example 5.5.1.
Let Z(p) be the ring of p-integral numbers (p is a prime number), i.e., Z(p) =
m
{ ∈ Q : (n, p) = 1}, and let Fp = Z(p) /pZ(p) be the field consisting of p
n
elements. Consider the ring A of 2 × 2-matrices of the following form:
Fp Fp
A =
0 Z(p)
(where Fp is considered as an Zp -module via the canonical quotient map Zp → Fp ).
It is easy to see that
0 Fp
R = rad A =
0 pZ(p)
236 ALGEBRAS, RINGS AND MODULES
and
0
2 0
R =
0 p2 Z(p) .
So
0 Fp
R/R2 =
0 pZ(p) /p2 Z(p)
and, by the Q-Lemma, the quiver Q(A) of A is the two-pointed quiver with the
adjacency matrix
0 1
[Q(A)] =
0 1.
Therefore, by theorem 14.2.1, vol.I, A is a right serial and left semidistributive
ring. So there is no analogue of the decomposition theorem for serial Noetherian
rings (see theorem 12.3.8, vol.I) even for Noetherian right serial and left semidis-
tributive rings with a two-pointed quiver.
Examples 5.6.1.
10
•
9 •
6 3 7 2 4 • • 5 11 • • 12
1 • • • • •
•
8
RIGHT SERIAL RINGS 237
Quivers which come from maps in this way have the property that each vertex
is the start of precisely one arrow. They are therefore right serial.
•
(a) • • ... • •
•
•
(b)
• • • • • • •
• •
• •
(c)
• • •
• • • • •
Theorem 5.6.1. A connected right serial quiver Q is either a tree with a single
root, or a quiver with a unique circuit which is a cycle such that after deleting all
arrows of this cycle, the remaining quiver is a disconnected union of trees whose
roots are the vertices of the cycle. Conversely, every such a quiver is right serial.
We shall use right serial quivers to prove the Cayley formula for the number of
different trees on a vertex set {1, 2, . . . , n}.
We shall use the main notions of graph theory (see [Harary, 1969]). In what
follows we consider undirected graphs without loops and multiple edges.
We shall assume that this numbering is fixed and say that Tn is a tree with
vertex set V Tn = {1, . . . , n}.
Examples 5.6.1.
1. n = 3.
y
2 •
1
• •
0 x
3 •
2 • 2 • 2 •
1 1 1
; ;
• • •
3 • 3 • 3 •
2. n = 4.
2 y
•
3 1
• • •
0 x
•
4
2 2 2
• • •
3 • • 1 3 • • 1 3 • • 1
• • •
4 4 4
2 2 2
• • •
3 • • 1 3 • • 1 3 • • 1
• • •
4 4 4
2 2 2
• • •
3 • • 1 3 • • 1 3 • • 1
• • •
4 4 4
2 2 2
• • •
3 • • 1 3 • • 1 3 • • 1
• • •
4 4 4
2 2 2
• • •
3 • • 1 3 • • 1 3 • • 1
• • •
4 4 4
RIGHT SERIAL RINGS 241
2
•
3 • • 1
•
4
Theorem 5.6.3. (The Cayley formula). For any n ≥ 2 there are nn−2
different trees with vertex set {1, . . . , n}.
Proof. Let T be a tree with vertex set V T = {1, . . . , n}. Fix two vertices x
and y in T .
Case I. x = y = k.
For any vertex i ∈ V T there is a single path that starts at i and ends at the
m
vertex k. Let i be the first edge in this path. Define ϕ(T, k, k) = ϕk in
• •
the following way: ϕk (k) = k and ϕk (i) = m for each i = k. Obviously, ϕk is a
well-defined map.
Case II. x = y (x = p, y = q).
Let M = {p, i1 , . . . , ik−2 , q} be the sequence of vertices of the unique path P
starting at the vertex p and ending at the vertex q, where i1 is the end of the edge
p i1 i i2
, i2 is the end of the edge 1 , etc., and q is the end of the edge
• • • •
ik−1 q
. Write
• •
a1 a2 . . . ak−1 ak
ϕ(T, x, y)|M =
p i1 . . . ik−2 q
such that the numbers a1 , a2 , . . . , ak−1 , ak in the first row are the numbers
p, i1 , . . . , ik−2 , q in their natural order. Let ϕ = ϕ(T, x, y). Consequently, write
ϕ(a1 ) = p, ϕ(a2 ) = i1 , . . ., ϕ(ak−1 ) = ik−2 , ϕ(ak ) = q. And for all remaining
m
vertices we define ϕ(i) = m, where i is the first edge in the unique path
• •
starting at the vertex i and ending at the vertex p.
a1 a2 ... ak−1 ak
f |M =
f (a1 ) f (a2 ) . . . f (ak−1 ) f (ak )
and draw
in this order as a path beginning at the vertex f (a1 ) and ending at the vertex
f (ak ). If the cycle is of length 1 (i.e., ak = a1 ), no edges are drawn. The remaining
vertices are connected as they were connected in Qf without orientation.
Now the proof follows from the fact that the set of all mappings from Nn into
Nn contains nn elements and the fact that a tree with vertex set {1, . . . , n} defines
n2 different mappings from Nn to Nn .
The algorithms used in the proof of theorem 5.6.3 are illustrated by the fol-
lowing examples.
Example 5.6.2.
⎧ ⎫
⎪
⎪ 2 • ⎪
⎪
⎪
⎨ ⎪
⎬
For n = 4 consider the tree T = T4 = 3 • • 1 .
⎪
⎪ ⎪
⎪
⎪
⎩ ⎪
⎭
4 •
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
⎜ 1 2 3 4 ⎟ ⎜ 1 2 3 4 ⎟ ⎜ 1 2 3 4 ⎟ ⎜ 1 2 3 4 ⎟
⎝ ⎠ ⎝ ⎠ ⎝ ⎠ ⎝ ⎠
1 1 1 1 1 2 1 1 1 1 3 1 1 1 1 4
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
⎜ 1 2 3 4 ⎟ ⎜ 1 2 3 4 ⎟ ⎜ 1 2 3 4 ⎟ ⎜ 1 2 3 4 ⎟
⎝ ⎠ ⎝ ⎠ ⎝ ⎠ ⎝ ⎠
2 1 1 1 2 2 1 1 2 1 3 1 2 1 1 4
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
⎜ 1 2 3 4 ⎟ ⎜ 1 2 3 4 ⎟ ⎜ 1 2 3 4 ⎟ ⎜ 1 2 3 4 ⎟
⎝ ⎠ ⎝ ⎠ ⎝ ⎠ ⎝ ⎠
3 1 1 1 3 1 2 1 3 1 3 1 4 1 1 3
⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
⎜ 1 2 3 4 ⎟ ⎜ 1 2 3 4 ⎟ ⎜ 1 2 3 4 ⎟ ⎜ 1 2 3 4 ⎟
⎝ ⎠ ⎝ ⎠ ⎝ ⎠ ⎝ ⎠
4 1 1 1 4 1 1 2 4 1 1 3 4 1 1 4
Example 5.6.3.
Consider the mapping ϕ : N12 → N12 for N12 = {1, 2, . . . , 12} as in example
5.6.1(2). The tree corresponding to this map ϕ has the following form:
4 y
5 • 3
• •
6 2
• •
7 1
• •
x
• •
8 12
• •
9 • 11
10
Note that this tree has leafs 6, 8, 10, 11 (where a leaf is defined as an end
vertex of a tree).
244 ALGEBRAS, RINGS AND MODULES
0 → M → M → M → 0
Clearly, the ring A has the same Cartan matrix as its basic ring. There is also
the following obvious proposition:
Therefore further on in this section we shall assume that A is basic and inde-
composable.
Examples 5.7.1.
1. Let A be a local right Artinian ring. Then C(A) = (m), where m = l(A A).
2. The Cartan matrix of a right Artinian ring is the identity matrix if and only
if A is semisimple.
3. Let A be the ring of upper triangular n × n matrices over a field k. Then
det C(A) = 1.
Lemma 5.7.4. Let A be a right Artinian ring with a set of pairwise orthogonal
primitive idempotents e1 , e2 , . . . , en and Cartan matrix C(A). Then
1. cij = l(ei Aei (ei Aej ));
n
2. l(ei A) = cij .
j=1
n
= hkj (−1)l [Pk ]
k=1
n
n
= cik hkj [Ui ].
i=1 k=1
Thus, writing H = (hij ) ∈ Mn (Z), we see that C(A)−1 = H ∈ Mn (Z), the ring of
n × n integral matrices. Therefore det C(R) = ±1.
Remark 5.7.1. The determinant of the right Cartan matrix is not necessarily
equal to the determinant of the left Cartan matrix even in the case when a ring
is two-sided Artinian (see [Fuller, 1992]). However they are equal to one-another
when A is an Artin algebra (see [Nakayama, 1938]).
dij = l(ei Aei (ei Aej )) and cij = l((ej Aei )ei Aei ).
Now if A is an Artin algebra over K and ti = l(K (ei Aei /ei Rei )) then
so that
In 1957 J.Jans and T.Nakayama2 proved that if a ring A has finite global
dimension and R2 = 0, where R = radA, or A is a quotient ring of a hereditary
ring, then the Cartan determinant is equal to 1. Since there are known examples
of Artinian rings of finite global dimension with Cartan determinant equal −1,
the problem was posed to settle what is now known as the Cartan determinant
conjecture:
If A is an Artinian ring of finite global dimension then its Cartan determinant
is equal to 1.
At the end of this section we shall prove this conjecture for right serial rings.
This conjecture was first proved in 1985 by W.D.Burgess, K.R.Fuller, E.R.Voss
and B.Zimmermann-Huisgen in their paper [Burgess, 1985].
It is easy to prove the following lemma.
Lemma 5.7.7. Let A be a right serial right Artinian ring with a set of primitive
pairwise orthogonal idempotents e1 , e2 , . . . , en and corresponding Cartan matrix
C(A) = (cij ). Then
n
l(ei A) = cij .
j=1
Lemma 5.7.8. Let A be a right serial ring, and suppose that A has a Kupisch
series e1 A, . . . , en A, all of whose members have the same composition length m.
Write m = an + r with 0 ≤ r < n. Then C has the following form:
a + 1 if 0 ≤ j − i < r or n − r < i − j ≤ n
cij =
a otherwise
(in other words, for r > 0, the matrix C has entries a on its first n−r subdiagonals
and the last n − r superdiagonals, and a + 1 on the remaining diagonals.)
Proof. Denote by [k] the least positive remainder of k modulo n. Then the
sequence of composition factors of ej A is Uj , U[j−1] , . . . , U[j−(n−1)] , Uj ,U[j−1] , . . . ;
it continues for m terms. Thus there are a + 1 copies of the first r candidates in
the list and a copies of the others.
Remark 5.7.2. Since for a quasi-Frobenius serial ring A all ei A have the same
composition length by theorem 5.3.2, the previous lemma gives the structure of
Cartan matrices for quasi-Frobenius serial rings.
Remark 5.7.3. The matrices that occur in lemma 5.7.8 are a special type of
circulant matrices.3 By a circulant matrix of order n is meant a square matrix
of the form
C = circ(c1 , c2 , . . . , cn ) =
⎛ ⎞
c1 c2 . . . cn
⎜cn c1 . . . cn−1 ⎟
⎜ ⎟
=⎜ . .. . . .. ⎟
⎝ .. . . . ⎠
c2 c3 ... c1
The elements of each row of C are identical to those of the previous row, but are
moved one position to the right and wrapped around. A circulant matrix can be
also written in the form:
C = (cjk ) = (c[k−j+1] ),
where [t] is the least positive remainder of t modulo n. Denote by (m, n) the
greatest common divisors of two integers m and n. We shall use the following
statement:
Taking this lemma into account there results the following as an immediately
corollary of lemma 5.7.8.
Lemma 5.7.11. If A is a right serial right Artinian ring with a simple module
of finite projective dimension, then A has a simple module of projective dimension
≤ 1.
e1 R ⊕ ei Ami
i=1
Lemma 5.7.13. Suppose that the projective dimension of the simple right
A-module U1 = e1 A/e1 R is ≤ 1 and set e = 1 − e1 . Then det C(eAe) = det C(A).
with mi ≥ 0.
Denoting the j-th column of the Cartan matrix C of A by
⎛ ⎞
c1j
⎜ ⎟
cj = ⎝ ... ⎠
cnj ,
we obtain ⎛ ⎞
1
⎜0⎟ n
⎜ ⎟
c1 = ⎜ . ⎟ + m j cj .
⎝ .. ⎠ j=2
0
Thus, subtraction of mj times the j-th column from the first column for j =
2, . . . , n yields the matrix
⎛ ⎞
1 c12 · · · c1n
⎜0 c22 · · · c2n ⎟
⎜ ⎟
⎜ .. .. .. .. ⎟
⎝. . . . ⎠
0 cn2 · · · cnn ,
RIGHT SERIAL RINGS 251
and consequently, ⎛ ⎞
c22 ··· c2n
⎜ .. .. .. ⎟
det C = det ⎝ . . . ⎠
cn2 ··· cnn .
But the latter matrix is the Cartan matrix of eAe.
Theorem 5.7.14. Given a right serial ring A, the determinant of its Cartan
matrix is 1 if and only if its right global dimension is finite. In any case the
determinant is nonnegative.
Corollary 5.7.15. Let A be an Artin algebra which is left or right serial, and
let C be its right Cartan matrix. Then det C = 1 if and only if gl.dim A < ∞. In
any case det C ≥ 0.
252 ALGEBRAS, RINGS AND MODULES
Proof. It is sufficient to note that in this case the left and right Cartan
matrices have the same determinant by proposition 5.7.6. So the right and left
global dimensions are equal.
Recall that a tiled order over a discrete valuation ring is a Noetherian prime
semiperfect semidistributive ring A with nonzero Jacobson radical. By theorem
5.1.1, vol.I, any tiled order A is of the form
⎛ ⎞
O π α12 O ... π α1n O
⎜ .. ⎟
⎜ α21 .. .. ⎟
⎜π O . . . ⎟
A=⎜
⎜ .
⎟,
⎟ (6.1.1)
⎜ .. .. ..
⎝ . . π αn1 O ⎟
⎠
π αn1 O ... π αn2 O O
The ring O is embedded into its classical division ring of fractions D, and so
that the tiled order A is the subset of all matrices (aij ) ∈ Mn (D) such that
where the e11 , . . . , enn are the matrix units of Mn (D). It is clear that Q = Mn (D)
is the classical ring of fractions of A.
Since eAe = O is a discrete valuation ring for any primitive idempotent e of a
tiled order A, we have the following simple lemma.
255
256 ALGEBRAS, RINGS AND MODULES
Throughout this section, unless specifically noted, A denotes a tiled order with
nonzero Jacobson radical R and classical ring of fractions Q.
We now recall the important notions of a band and a semilattice.
Let E be the set of idempotent elements of a semigroup S. For e, f ∈ E we
define e f iff ef = f e = e. In this case we say that e is under f and f is over
e. To see that is a partial ordering of E, let e, f, g ∈ E. Then
(1) e2 = e, and hence e e.
(2) If e f and f e then ef = f e = e and f e = ef = f , whence e = f .
(3) If e f and f g then ef = f e = e and f g = gf = f , whence eg =
(ef )g = e(f g) = ef = e, and ge = g(f e) = (gf )e = f e = e. Hence e g.
We shall call the natural partial ordering of E.
An element b of a partially ordered set X is called an upper bound of a subset
Y of X if y b for every y in Y . An upper bound b of Y is called a least upper
bound (or join) of Y if b c for every upper bound c of Y . If Y has a join in
X, it is clearly unique. Lower bound and greatest lower bound (or meet)
are defined dually. A partially ordered set X is called an upper (resp. lower)
semilattice if every two-element subset {a, b} of X has a join (resp. meet) in X;
it follows that every finite subset of X has a join (or meet). The join (resp. meet)
of {a, b} will be denoted by a ∪ b (resp. a ∩ b). A partially ordered set which is
both an upper and lower semilattice is called a lattice. A lattice X is said to be
complete if every subset of X has a join and a meet.
Lemma 6.1.3. Let O be a discrete valuation ring and D be its classical ring of
fractions, which is a division ring. Then D is a uniserial right and left O-module.
If X ⊂ D is a right O-module (X = D), then X = π t O = Oπ t .
⎛ ⎞ ⎛ ⎞
π α1 0 ... ... 0 π −α1 0 ... ... 0
⎜ ⎟ ⎜ ⎟
⎜ .. ⎟ ⎜
.. .. ⎟
..
⎜ 0 πα2 . ⎟ ⎜
. 0 π−α2 . ⎟
.
⎜ ⎟ ⎜ ⎟
⎜ .. ..⎟ ⎜ .. ..⎟
⎜ .. .. .. ⎟A⎜ .. .. .. ⎟
⎜ . . . . .⎟ ⎜ . . . . .⎟
⎜ ⎟ ⎜ ⎟
⎜ .. .. .. ⎟ ⎜ .. .. .. ⎟
⎜ . . . 0 ⎟ ⎜ . . . 0 ⎟
⎝ ⎠ ⎝ ⎠
0 ... ... 0 π αn 0 ... ... 0 π −αn
is an isomorphic tiled order with all exponents ≥ 0. Thus, if desired, this addi-
tional property can always be assumed.
For our purposes, it suffices to consider a reduced tiled order A. In this case,
the elements of Sr (A) (Sl (A)) are in a bijective correspondence with integer-valued
row vectors a = (α1 , . . . , αn ) (column vectors aT = (α1 , . . . , αn )T ), where a and
258 ALGEBRAS, RINGS AND MODULES
a b ⇐⇒ αi ≥ βi for i = 1, . . . , n.
Proof. Let A = {O, E(A)} be a tiled order with exponent matrix E(A) = (αij ).
Let M = (α1 , . . . , αn ) ∈ Sr (A). Let a = min (α1 , . . . , αn ). Then M1 =
(α1 − a, . . . , αn − a) is an irreducible A-lattice and M1 M . Suppose that αi = a.
Then M1 = (β1 , . . . , βn ), where the β1 , . . . , βn are non-negative and βi = 0. Con-
sequently, every irreducible A-lattice M is isomorphic to a lattice M1 with at least
one zero coordinate. We obtain from (6.1.2) that 0 ≤ βj ≤ αij . So, the number of
irreducible A-lattices of the form M1 is finite. The proposition is proved.
Proposition 6.1.6. All overrings of a tiled order A are tiled orders. They
form a finite lower semilattice.
Theorem 6.1.8 (R.P.Dilworth).1 For a poset P with finite width w(P ) the
minimal number of disjoint chains that together contain all elements of P is equal
to w(P ).
Remark 6.1.1. This theorem holds for any poset P of finite width. Here is
a proof of this theorem for the countable case. To prove this theorem we use the
following lemma.
.
w−1
P \C = Ci
i=0
$
w−1
and P = ( Ci ) ∪ C is a union of the w disjoint chains.
i=0
Let w(P \ C) = w and let a1 , . . . , aw be a maximal antichain. Then every
set C ∪ {ai } is not strongly dependent for i = 1, . . . , w. It means that for every
i = 1, . . . , w there exists a finite subset Si ⊂ P such that w(Si ∩ (C ∪ {ai })) = 2.
Consequently, ai ∈ Si for i = 1, . . . , w.
$w
Consider the finite set S = Si . The set S may be presented as a union of w
i=1
$
w
disjoint chains S = Ki by the lemma in the finite case (which has been proved
i−1
above). Renumbering the Ki ’s if needed we can assume that S ∩ C ⊂ K1 . We can
TILED ORDERS OVER DISCRETE VALUATION RINGS 261
Proof of theorem 6.1.8. Let m be the minimal number of disjoint chains that
together contain all elements of P , and let w = w(P ). So, there exist pairwise
incomparable elements x1 , . . . , xw . Obviously, any chain cannot contain two of
these elements together. Consequently, m ≥ w.
Let P be an arbitrary partially ordered set. Then one can construct a new
whose elements are the nonempty subsets of P that consist
partially ordered set P,
of pairwise incomparable elements, including the subsets consisting of only one
element of P. We shall introduce an order in P in the following way. If A, B ∈ P,
then A B in P if and only if for any a ∈ A there exists b ∈ B such that a b
an element a ∈ P is mapped into
in P . The poset P is naturally embedded in P:
the singleton set {a}.
Example 6.1.1.
1 2
In particular if P : then
• •
{1, 2}
•
:
P
• •
{1} {2}
;
262 ALGEBRAS, RINGS AND MODULES
1 2 3
If P : then
• • •
{1, 2, 3}
•
• • •
{1} {2} {3}
Θ : (P1 , . . . , Pm ) → P1 + . . . + Pm
defines an epimorphism of the partially ordered set M /r (A) onto Sr (A). Let M =
P1 + . . . + Pm = Pi1 + . . . + Pit . Then P1 ⊆ Pi1 + . . . + Pit . We may assume that
P1 e11 A, i.e., P1 = (a, α12 + a, . . . , αm + a) for some a ∈ Z.
Obviously, if L1 = (α1 , . . . , αn ) and L2 = (β1 , . . . , βn ) are irreducible A-lattices
from Sr (A) then L1 + L2 = (min (α1 , β1 ), . . . , min (αn , βn )).
(i ) (i )
Let Pik = (a1 k , . . . , an k ). Then
(i )
Pi1 + . . . + Pit = ( min a1 k , . . . , min a(ik)
n ).
1≤k≤t 1≤k≤t
(i ) (i )
From the ordering relation in Sr (A) we obtain that min a1 k ≤ a, i.e., a1 k ≤
1≤k≤t
a for some k. Let Pik eik ik A and Pi1 = (αi1 + b, . . . , αin + b) for some b ∈ Z.
TILED ORDERS OVER DISCRETE VALUATION RINGS 263
Remark 6.1.2. The notions of a lattice and an A-lattice used in this proof are
different; the one is a partially ordered set, the other is a special kind of module.
Note that Mr (A) = Sr (A) if and only if the width of the set Mr (A) is equal
to 1.
Let a tiled order A be reduced. Then αij +αji > 0 for i = j, and the set Mr (A)
can be simply constructed from the matrix (αij ): let pki = (αi1 + k, . . . , αin + k),
where i = 1, . . . , n and k ∈ Z. Suppose pki = plj . This means that αim +k = αjm +l
for m = 1, . . . , n. In particular, αii + k = αji + l, i.e., k = αji + l. Analogously,
αij + k = αjj + l, i.e., l = αij + k. Therefore, k + l = αji + αij + k + l and we obtain
the contradiction: αji + αij = 0. Obviously, the partial ordering on Mr (A) can
be defined as follows: pki plj if and only if αjm + l ≤ αim + k for m = 1, . . . , n.
Therefore, αji + l ≤ k.
Conversely, let αji + l ≤ k. Adding αim to the both sides of this equality we
obtain αji +αim +l ≤ k+αim . But αji +αim ≥ αjm and αji +αim +l ≥ k+αjm +l.
Thus, αjm + l ≤ k + αim , i.e., pki plj .
Analogously, we can construct Ml (A). The element pki is in Ml (A) if pki =
(α1i + k, . . . , αni + k)T and, as above, all elements pki are different and pki plj if
and only if k ≥ l + αij .
p−l
j ≤ pi
−k
if and only if −l ≥ −k + αji . This proves the statement.
Remark 6.1.3. One should note that this anti-isomorphism cannot be ex-
tended to an anti-isomorphism of the lattices Sl (A) and Sr (A), since the anti-
isomorphism of lattices is given by the correspondence:
Proof. The equivalence of (a) and (b) follows from the fact that any irreducible
A-lattice M ∈ Sr (A) = M /r (A), is identified with a collection T = {t1 , . . . , tk } of
pairwise incomparable elements of the set M(A), has exactly k maximal submod-
ules identified by the sets T1 , . . . , Tk , where Ti is the union of the sets T \ {ti } with
the set of points strictly less than ti , but not comparable with any point of T \{ti }.
On the other hand, the number k is the number of indecomposable summands of
a projective cover P (M ) of M , which is equal to the length of the module M/M R
and this proves the equivalence of (a) and (c).
We shall use the following notation: A = {O, E(A)}, where E(A) = (αij ) is the
exponent matrix of the tiled order A, i.e.,
n
A= eij π αij O,
i,j=1
where the eij are the matrix units. If a tiled order is reduced, i.e., A/R is a
direct product of division rings, then αij + αji > 0 if i = j, i.e., E(A) is a reduced
exponent matrix.
Let E = (αij ) be a n × n reduced exponent matrix. Define n × n matrices
"
(1) αij if i = j,
E = (βij ), where βij =
1 if i = j,
and
E (2) = (γij ), where γij = min (βik + βkj ).
1≤k≤n
(2) (1)
Obviously, [Q] = E −E is a (0, 1)-matrix.
The following theorem is the same as theorem 14.7.1, vol.I, where it was proved
using the fact that a tiled order is a prime ring. Here we shall give a direct proof
of this theorem using only the matrix definitions.
Theorem 6.1.15. The matrix [Q] = E (2) − E (1) is the adjacency matrix of
the strongly connected simply laced quiver Q = Q(E).
Proof. Since [Q] is a (0, 1)-matrix, it is the adjacency matrix of a simply laced
quiver.
We shall show that [Q] is a strongly connected quiver. Suppose the contrary.
This means that there is no path from the vertex i to the vertex j in Q for some
i, j. Denote by V Q(i) = V1 a set of all vertices k ∈ Q such that there exists a
path beginning at the vertex i and ending at the vertex k. Then, by assumption,
V2 = V Q\V Q(i) = ∅ (because j ∈ V (Q)\V (Q)(i)). Consequently, V Q = V1 ∪V2
266 ALGEBRAS, RINGS AND MODULES
α1p = 0 p = 1, . . . , n.
Note that this is transitive because αij + αjk ≥ αik ≥ 0. We can assume that
m + 1 is a minimal element. It follows that
Also k ≤ m because α1,k = 0 = α1,m+1 while αk,m+1 > 0 for k > m + 1. Thus
there is a k, 2 ≤ k ≤ m + 1 with αk,m+1 = 0. Interchanging the 2-nd and k-
th columns and rows simultaneously we can assume that α2,m+1 = 0. Further
q2,m+1 = 0, and so (arguing as before)
Recall that the quiver of a reduced exponent matrix E is the quiver Q(E)
with adjacency matrix [Q]. A strongly connected simply laced quiver is called
admissible if it is a quiver of a reduced exponent matrix.
A reduced exponent matrix E = (αij ) ∈ Mn (Z) is Gorenstein if there exists
a permutation σ of {1, 2, . . . , n} such that αik + αkσ(i) = αiσ(i) for i, k = 1, . . . , n.
Proof. Consider the matrix E = (αij ), where αii = 0 and αij is equal to the
minimum length of a path from the vertex i to the vertex j for i = j. (Note that
a path of minimum length always exists because Q is a strongly connected quiver.
There may be more than one path of minimum length.)
Let us show that E is a reduced exponent matrix. Since the minimum length
of a path from i to k is less than or equal to the minimum length of a path from
268 ALGEBRAS, RINGS AND MODULES
i to k which passes through the vertex j, we have αik ≤ αij + αjk . By definition,
αij ≥ 1 if i = j, and so αij + αji > 0.
Let us show that Q(E) = Q. Since αij + αji ≥ 2 for j = i, qii = mink (βik +
βki − βii ≥ 1, and so there exists a loop at each vertex of Q(E). Assume there
exists an arrow from i to j. Then αij = 1. Since αtk ≥ 1 if t = k, there is no
k = i, j such that αik + αkj = 1 = αij . Therefore αik + αkj > αij for all k = i, j,
and
σ1 . . . σu σu+1 . . . σv : i → j
σ1 . . . σu : i → t, σu+1 . . . σv : t → j,
and αij ≥ 2. Then σ1 . . . σu and σu+1 . . . σv are paths of minimum length from i
to t and from t to j, respectively. Hence, αit + αtj = αij . Thus
Example 6.1.2.
It is easy to see that the quiver Q with the adjacency matrix
1 1
[Q] =
1 0
is not admissible.
Proof. We have ⎧
⎪
⎪ αij , if i = l, j = l,
⎨
0, if i = l, j = l,
θij =
⎪ αlj − t,
⎪ if i = l, j = l,
⎩
αil + t, if i = l, j = l,
where t is an integer. One can directly check that if αij + αjk = αik for some i, j, k,
then θij + θjk = θik . Since these transformations are invertible, the converse
transformations have a similar form. So the equality θij + θjk = θik implies
αij + αjk = αik . Therefore, θij + θjk = θik if and only if αij + αjk = αik .
Denote Θ(1) = (μij ) and Θ(2) = (νij ).
The equalities γij = βij , νij = μij or inequalities γij > βij , νij > μij hold
simultaneously for the entries of the matrices (βij ) = E1 , (μij ) = Θ(1) , (γij ) = E (2) ,
(νij ) = Θ(2) . Therefore, E (2) − E (1) = Θ(2) − Θ(1) and [Q(E)] = [Q(Θ)].
Suppose that E is a reduced Gorenstein exponent matrix with a permutation
σ(E), i.e., αij + αjσ(i) = αiσ(i) . Whence, θij + θjσ(i) = θiσ(i) . This means that the
matrix Θ is also Gorenstein with the same permutation σ(E).
μij = βτ (i)τ (j) , νij = min(μik + μkj ) = min(βτ (i)l + βlτ (j) ) = γτ (i)τ (j) ,
k l
it follows that,
Note that the type (= conjugacy class) of a permutation is not changed un-
der transformations of the second type. Therefore, in order to describe reduced
270 ALGEBRAS, RINGS AND MODULES
Gorenstein exponent matrices, one needs to examine matrices with different types
of permutations. Further, to simplify calculations we can assume that some row
or some column of E is zero. This can always be assured by transformations of
the first type and then the entries of a new exponent matrix will be non-negative
integers (see the proof of theorem 6.1.15).
Indeed, let E = (αij ) ∈ Mn (Z) be an exponent matrix. Subtracting α1i from
the entries of the i-th column and adding this number to the entries of the i-th
row, we obtain a new exponent matrix
⎛ ⎞
0 0 0 ... 0
⎜ θ21 0 θ23 . . . θ2n ⎟
⎜ ⎟
⎜ θ31 θ32 0 . . . θ3n ⎟
Θ = ⎜ ⎟
⎜ .. .. .. .. .. ⎟
⎝ . . . . . ⎠
θn1 θn2 θn3 . . . 0
The first row of Θ equals zero. Consequently, θ1i + θij ≥ θ1j = 0 and θij ≥ 0 for
i, j = 1, . . . , n.
Proof. The classical ring of fractions Q is the direct limit of flat submodules
π k A = Aπ k of A, for k ∈ Z (as k → −∞). Thus Q is flat, by proposition 5.4.6,
vol.I.
To prove the injectiveness of Q we use the Baer criterion (see proposition 5.2.4,
vol.I). Let I be a right ideal in A. Since A is a Noetherian ring, I is a finitely
generated ideal. Consider a diagram
i
0 IA AA ,
ϕ
i⊗1Q
0 IA ⊗ Q AA ⊗ Q
is exact. Then, by proposition 5.4.11, vol. I, we obtain the following diagram
TILED ORDERS OVER DISCRETE VALUATION RINGS 271
i
0 I Q,
ϕ
Let A be a tiled order of the form (6.1.1). Recall that an A module M is called
an A-lattice if it is a finitely generated free O-module, where O is the discrete
valuation ring of A (see vol.I, p.353). We shall denote by Latr (A) (resp. Latl (A))
the category of right (resp. left) A-lattices.
i p
0 L M N 0
We shall establish now the duality between the category Latr (A) and Latl (A).
Let M ∈ Latr (A) and let M # = HomO (M, O). For any f ∈ M # and a ∈ A we
can define af by the formula (af )(m) = f (ma) where m ∈ M . Then it is easy to
verify that M # is a left A-module.
Since M ∈ Latr (A), it is a free O-module with a finite O-basis e1 , e2 , . . . , en .
As in section 4.10, we can define an O-homomorphism ϕi : M → O by the formula
ϕi (ej ) = δij for i, j = 1, . . . , n, where δij is the Kronecker symbol. Then ϕi ∈ M # .
It is easy to see that M # is a free O-module with O-basis ϕ1 , . . . , ϕn . This O-basis
is called the dual O-basis of M # . Thus, M # ∈ Latl (A). If M ∈ Latl (A), then
M # ∈ Latr (A).
Let ϕ : M → N be a homomorphism of M, N ∈ Latr (A), i.e., ϕ ∈
HomA (M, N ). Then ϕ# : N # → M # can be defined by the formula (ϕ# f )(m) =
f ϕ(m), where f ∈ N # , is a homomorphism from N # to M # , i.e., ϕ# ∈
HomA (N # , M # ). Obviously, if we have homomorphisms ψ : L → M and
ϕ : M → N , then (ϕψ)# = ψ # ϕ# and 1# M = 1M # . Moreover, for any
M ∈ Latr (A) we have M ## = M and for any N ∈ Latl (A) it is true N ## = N .
Besides, for any ϕ : M → N we have ϕ## = ϕ. It is also obvious that
(M ⊕ N )# = M # ⊕ N # .
p#
0 N# M# L# 0
is exact.
Proof. Let
0 # 0
AA M N
be an exact sequence. By corollary 6.2.7, we obtain that
0 N# M# AA 0
AA = e11 A ⊕ . . . ⊕ enn A
and also a completely decomposable left A-lattice
AA = Ae11 ⊕ . . . ⊕ Aenn .
Every finite generated left projective A-module A P has the following form: AP =
(Ae11 )m1 ⊕ . . . ⊕ (Aenn )mn . Obviously, A P ∈ Latl (A) and
#
AP = (Ae11 )#m1 ⊕ . . . ⊕ (Aenn )#mn .
In what follows we assume that the tiled order A is reduced. In this case E(A)
is reduced, i.e., αij + αji > 0 for i = j. An A-lattice N ⊂ Mn (D) is said to be
2
complete if N (OO )n as a right O-module. If a complete A-lattice N is a left
A-module then eii N ejj ⊂ N . So eii N ejj = π γij O and
n
N = eij π γij O.
i,j=1
TILED ORDERS OVER DISCRETE VALUATION RINGS 275
Note that N is a right and left A-module if and only if γij + αik ≥ γik and
αik +γkj ≥ γij for all i, j, k. In this case the matrix (γij ) is said to be the exponent
matrix of the A-lattice N and we write it as E(N ). Complete A-lattices which
are left A-modules are said to be fractional ideals of the order A. Denote by Δ
the completely decomposable lattice A# A.
Proof. Let us show that the k-th row (−α1k , −α2k , . . . , −αnk ) of the matrix
E(Δ) defines an irreducible right A-lattice. Write βi = −αik . We can rewrite the
inequality αij + αjk ≥ αik in the form −αik + αij ≥ −αjk , i.e., βi + αij ≥ βj ,
which implies the assertion of the lemma.
Proof. The proof for the left case follows from the fact that eii R is the only
maximal submodule of eii A and from the duality properties and the annihilation
lemma. The proof for the right case is just the same.
276 ALGEBRAS, RINGS AND MODULES
Note once more, that the eii Δ (Δeii ) are all indecomposable relatively right
(left) injective A-lattices (up to isomorphism) and the eii X (Xeii ) are unique
minimal overmodules of eii Δ (Δeii ). Moreover, the notions of indecomposable
relatively injective A-lattice and irreducible relatively injective A-lattice coincide.
Let A1 and A2 be Morita equivalent tiled orders. Then the relatively injective
indecomposable A1 -lattices correspond to relatively injective indecomposable A2 -
lattices. Thus, from lemma 6.2.12 there follows the following lemma
Lemma 6.2.13. Every relatively injective irreducible A-lattice Q has only one
minimal overmodule. Let Q1 and Q2 be relatively injective irreducible A-lattices,
and let X1 ⊃ Q1 and X2 ⊃ Q2 be the unique minimal overmodules of Q1 and Q2 ,
respectively. Then the simple A-modules X1 /Q1 and X2 /Q2 are isomorphic if and
only if Q1 Q2 .
The dual statement to proposition 6.1.7 is the following proposition, the proof
of which can be simply obtained from duality properties:
The following proposition states some interesting fact about the injective di-
mension of the lattice A A# .
For any finite poset P = {p1 , . . . , pn } we can construct a reduced tiled (0, 1)-
order A(P) by setting
E(A(P)) = (αij ),
where αij = 0 ⇐⇒ pi pj and αij = 1, otherwise.
Then A(P) = {O, E(A(P))} is a reduced (0, 1)-order (see vol.I, §14.6).
Theorem 6.3.1. For any finite poset P there is a countable set of Frobe-
nius rings Fm (P) with identity Nakayama permutation such that Q(Fm (P)) =
Q(A(P)).
A ⊃ R ⊃ R2 ⊃ π m+t X ⊃ π m J.
The same relation holds for the left modules. Therefore, the Nakayama permuta-
tion of the ring Fm (P) is the identity permutation.
Theorem 6.3.2. For every reduced tiled order A over a discrete valuation
ring, there is a countable set of Frobenius rings Fm (A) with identity Nakayama
permutation such that Q(Fm (A)) = Q(A).
278 ALGEBRAS, RINGS AND MODULES
Proof. For the fractional ideal Δ, there is the least positive integer t such that
π t Δ ⊂ R2 . Then the quotient ring Q(Fm (A)) = A/π m+t Δ is a Frobenius ring
with identity Nakayama permutation.
Example 6.3.1.
Let k be a field, O = k[[x]] and π = x. Let
O O
A = ,
πα O O
where α ≥ 2. Obviously,
0 0 1 1
E(A) = and [Q(A)] = .
α 0 1 1
In this case
0 −α
E(Δ) = .
0 0
We have
2 1
E(R2 ) = .
α+1 2
Consequently, t = α + 1 and
α+1 1
E(π α+1
Δ) =
α+1 α+1
and the quotient ring Fm (A) = A/π m+t Δ is a Frobenius ring with the identity
Nakayama permutation.
Note that
m+α+1 m+1
E(π m+t Δ) = .
m+α+1 m+α+1
Let k be a finite field with q elements. Then Fm (A) is a finite Frobenius ring
and |Fm (A)| = q 4m+3α+4 .
Proof. Indeed, let O be a discrete valuation ring with unique maximal ideal
M, and let ⎛ ⎞
O M ... ... M
⎜ .. .. .. ⎟
⎜M . . . ⎟
⎜ ⎟
⎜ .. ⎟
Kn (O) = ⎜ ... ..
.
..
.
..
. . ⎟
⎜ ⎟
⎜ . .. .. ⎟
⎝ .. . . M⎠
M ... ... M O
be a tiled order.
TILED ORDERS OVER DISCRETE VALUATION RINGS 279
Let O = k[[t]] be the ring of formal power series over a field k, then
Fm (k[[t]]) = Kn (k[[t]])/Im is a countable set of Frobenius semidistributive al-
gebras Am = Fm (k[[t]]) such that ν(Am ) = σ. If k is finite, then all the algebras
Am are finite.
The original poset P and the [Q]-matrix of the exponent matrix EP have
a great deal to do with one another. Recall that the diagram of P is the
quiver Q(P) which has an arrow i → j if and only if i ≺ j and there is no
k ∈ P such that i ≺ k ≺ j; see the beginning of section 6.3 above. Let
Q(P) (the “extended diagram” of P) be obtained from Q(P) by adding an ar-
row i → j to each pair (i, j) such that i is maximal in P and j is minimal
in P.
For example, if P is the two component poset
4 • • 5
• 3 •
1
2 •
280 ALGEBRAS, RINGS AND MODULES
then Q(P) is
4 • • 5
• 3 •
1
2 •
Such extended diagrams have played a role before, see theorem 14.6.3, vol.I.
There is now the following relation between Q(P) and the “[Q]-matrix” of
the exponent matrix EP . Let (pij ) be the adjacency matrix of Q(P) and (qij ) =
[Q(EP )], then pij = qij , i, j = 1, . . . , n.
The proof of this is a rather simple matter of checking various cases, as follows.
There are 4 cases: i = j, i ≺ j, j ≺ i, i and j are incomparable. Let EP = (αij ),
where
0 if i j
αij =
1 otherwise
Then
qij = min(βik + βkj ) − βij
k
where
1 if i = j
βij =
αij j
if i =
Note that for all i, j always βii + βij − βij = βii = 1 and βik + βkj ≥ βij so
that 0 ≤ qij ≤ 1. Also, always βij + βjj − βij = βjj = 1. So to calculate qij it only
remains to deal with the k ∈ {i, j}.
Case 1: i = j. There are two subcases.
Case 1.1: i = j and i is not comparable to any other element of P so that {i}
is a one-element component of P. Then i is both maximal and minimal and there
is a loop at i in Q(P). Thus pii = 1. On the other hand, βii + βii − βii = βii = 1
and for all k = i (if any) βik + βki − βii = 1 + 1 − 1 = 1 because k and i are
incomparable. Thus qii = 1.
Case 1.2: i = j and i is comparable to some k = i. Then 1 = αik + αki =
βik + βki . And so βik + βki − βii = 0 and qii = 0. On the other hand, pii = 0
because i is not maximal and of course not i ≺ i.
Case 2: i ≺ j, so that αij = 0. There are again two subcases.
Case 2.1: i ≺ j and there is no k such that i ≺ k ≺ j, i.e. j covers i. Then
pij = 1. On the other hand, αij = 0. Note that always βii + βij − βij = 1,
βij + βjj − βij = 1 so that to calculate qij it remains to look at the k ∈ {i, j}. For
the location of k vis à vis i and j there are 4 subsubcases.
i ≺ k and k ≺ j. This is not possible under case 2.1.
i ≺ k and not (k ≺ j). Then βik + βkj − βij = αik + βkj − αij = 0 + 1 − 0 = 1.
TILED ORDERS OVER DISCRETE VALUATION RINGS 281
Definition. Two finite partially ordered sets S and T are called Q-equivalent
if the reduced exponent (0, 1)-matrices ES and ET are equivalent (meaning that
ET can be obtained from ES by repeated use of the transformations (1) and (2) in
the definition just above proposition 6.1.17 above).
4 2 3
• • •
S = 2 • • 3 and T = • 1
• •
1 4
Obviously,
⎛ ⎞ ⎛ ⎞
0 0 0 0 0 0 0 1
⎜1 0 1 0⎟ ⎜1 0 1 1⎟
ES = ⎜
⎝1
⎟ and ET = ⎜ ⎟.
1 0 0⎠ ⎝1 1 0 1⎠
1 1 1 0 0 0 0 0
1 2 • • 3
•
1 •
Q(S) = 2 • • 3 )=
Q(T
• •
4 4
⎛ ⎞
0 0 0 1
⎜1 0 0 0⎟
⎜
[Q(S)] = ⎝ ⎟ = [Q(ES )]
1 0 0 0⎠
0 1 1 0
⎛ ⎞
0 1 1 0
⎜0 0 0 1⎟
)] = ⎜
[Q(T ⎟ = [Q(ET )]
⎝0 0 0 1⎠
1 0 0 0
Note that the two adjacency matrices go into one another by a simultaneous
interchange of the first row and fourth row and first column and fourth column,
making Q(S) ) isomorphic quivers.
and Q(T
The matrix ET is obtained from ES by subtracting 1 from the last row of ES
with simultaneous adding 1 to the last column.
Proposition 6.4.1. For any two posets S and T , if the exponent matrices ES
and ET are equivalent then Q(ES ) and Q(ET ) are isomorphic.
TILED ORDERS OVER DISCRETE VALUATION RINGS 283
Example 6.4.2.
The following two posets satisfy the conditions of theorem 6.4.2.
(a) (b)
• • • •
• • • • •
• • • •
• • • • •
• •
• • •
and
• •
• •
and moreover in these posets each element of the one subset is not greater than
any element of the other subset.
If qij = 0, then the element αj does not cover the element αi and either αi
isn’t maximal or αj isn’t minimal (or both); and the element γj does not cover
the element γi and either γi isn’t maximal or γj isn’t minimal (or both).
If qij = 1, then either the element αj covers the element αi or αi is maximal
and αj is minimal; and either the element γj covers the element γi or γi is maximal
and γj is minimal.
Assume that Q(P1 ) = Q(P2 ). Then (because Q(P 1 ) = Q(P
2 )) there exist an i
and j such that αj covers αi , but γi is maximal and γj is minimal (or vice versa).
Let αj = αj1 cover the elements αi = αi1 , . . . , αis . Then γj = γj1 is a
minimal element, and γi1 , . . . , γis are maximal elements, moreover, there are no
other maximal elements in the poset P2 . Indeed, if there was a maximal element
γi0 ∈ {γi1 , . . . , γis } in P2 there would be an arrow from γj0 to γj in Q(P 2 ) and
hence an arrow from αi0 to αj in Q(P 1 ). But this would mean that αj covers αi0 .
Therefore αi0 ∈ {αi1 , . . . , αis }, this is a contradiction.
If P2 min = {γj1 , ..., γjr }, then all the αj1 , ..., αjr cover every element of
{αi1 , . . . , αis }.
Let P1 max = {αp1 , ..., αpm }, P1 min = {αl1 , ..., αlt }, P2 max = {γi1 , ..., γis }.
Then, if αp = αpv ∈ P1 max and αl = αlu ∈ P1 min , then γlu covers γpv .
Denote P1 = {αq ∈ P1 : αq > αi1 }, P1 = P1 \P1 ; P2 = {γq ∈ P2 : γq >
αp1 }, P2 = P2 \P2 .
Then Q(P1 ) = Q(P2 ) and Q(P1 ) = Q(P2 ). Indeed, since in each set
P1 , P2 , P1 , P2 there are no simultaneously elements of P1 max and P1 min ; P2 max
and P2 min , from the equality qij = 1 for, e.g., αi , αj ∈ P1 , it follows that αj covers
αi and γj covers γi , where γi , γj ∈ P2 .
Conversely, let the connected posets$P1 and P2 be either$isomorphic or there
exist partitions of these sets P1 = P1 P1 and P2 = P2 P2 such that each
element of P1 is not greater than any element of P1 , each element of P2 is not
greater than any element of P2 , and P1 P2 , P1 P2 . If P1 P2 then,
obviously, E(P1 ) ∼ E(P2 ) (renumbering of elements of the poset P2 corresponds
to the second type equivalent transformations of the matrix E(P2 )). If P1 P2
and there exists a suitable partition then
E(P1 ) U E(P2 ) U
E(P1 ) = , E(P2 ) = ,
0 E(P1 ) 0 E(P2 )
where U is a matrix with all entries equal to 1, E(P1 ) = E(P2 ), E(P1 ) = E(P2 )
(therefore P1 P2 , P1 P2 ).
Matrices E(P1 ) and E(P2 ) are equivalent. Really, at first, by means of trans-
formations of the second type from the matrix E(P2 ) we obtain a matrix
E(P 2 ) 0
E(P 2 ) = ,
U E(P 2 )
where E(P 2 ) = E(P2 ) = E(P1 ), E(P 2 ) = E(P2 ) = E(P1 ). Then, by means of
equivalent transformations of the first type from the matrix E(P 2 ) we obtain the
TILED ORDERS OVER DISCRETE VALUATION RINGS 285
matrix E(P1 ). Thus, the posets P1 and P2 are Q-equivalent. The theorem is
proved.
In the proof of this theorem, we have also proved the following proposition.
Proposition 6.4.3. For finite connected posets S and T , if Q(ES ) and Q(ET )
are isomorphic, then ES and ET are equivalent.
S (1) = {ϕ(αi1 ), ..., ϕ(αim )}; S (2) = {ϕ(αk1 ), ..., ϕ(αkt )}.
286 ALGEBRAS, RINGS AND MODULES
S (3) = {ϕ(αj1 ), ..., ϕ(αjs )}; S (4) = {ϕ(αl1 ), ..., ϕ(αlr )}.
The diagram Q(P) of a finite poset P is the union of its connected compo-
nents Q(P1 ), . . . , Q(Ps ). The subsets P1 , . . . , Ps will be called the connected
components of the poset P.
Theorem 6.4.6. For finite posets S and T the following conditions are
equivalent:
(a) Q(ES ) and Q(ET ) are isomorphic;
(b) ES and ET are equivalent.
n
Pτ = eiτ (i)
i=1
be the corresponding permutation matrix, where the eij are the matrix units.
Clearly, PτT Pτ = Pτ PτT = En is the identity matrix of Mn (R). In particular,
⎛ ⎞
0 0 ... 0 1
⎜ 0 0 ... 1 0 ⎟
⎜ ⎟
⎜ .. .. .. .. .. ⎟
Dn = ⎜ . . . . . ⎟
⎜ ⎟
⎝ 0 1 ... 0 0 ⎠
1 0 ... 0 0
1 2 ... n − 1 n
is a Pσ , where σ = , and DnT = Dn .
n n − 1 ... 2 1
Recall that a matrix B ∈ Mn (R) is called permutationally reducible if
there exists a permutation matrix Pτ such that
B1 B12
PτT BPτ = ,
0 B2
where B1 and B2 are square matrices of order less that n. Otherwise, the matrix
B is called permutationally irreducible.
288 ALGEBRAS, RINGS AND MODULES
(1) (2)
where B1 and B2 are square matrices of order less that n.
Recall that a matrix A = (aij ) ∈ Mn (R) is called positive if aij > 0 for
i, j = 1, . . . , n. If all aij ≥ 0, A is called non-negative.
λh − rh = 0. (6.5.1)
3 see [Perron, 1907].
4 see [Frobenius, 1912]
TILED ORDERS OVER DISCRETE VALUATION RINGS 289
n
si = aij (i = 1, 2, . . . , n), s = min si , S = max si .
1≤i≤n 1≤i≤n
j=1
s ≤ r ≤ S,
and the equality sign on the left or the right of r holds for s = S only; i.e., they
hold only when all the row-sums s1 , s2 , . . . , sn are all equal.
Let A be a semiperfect ring with Jacobson radical R. Suppose that the quotient
ring A/R2 is right Artinian. In this case the quiver Q(A) of the ring A is defined.
Write [Q(A)] = (tij ) for the adjacency matrix of Q(A). Recall that tij is the
number of arrows between i and j, where V Q(A) = {1, . . . , n}.
By theorem 6.5.3, there exists a non-negative characteristic value r such that
the moduli of all the characteristic values of A do not exceed r.
290 ALGEBRAS, RINGS AND MODULES
⎛ ⎞
0 B/J 0 ... ... 0
⎜ .. .. .. ⎟
⎜ 0 0 . . . ⎟
⎜ ⎟
⎜ .. .. .. .. .. .. ⎟
⎜ . . . . . . ⎟
and Rn /Rn = ⎜
2
⎜ ..
⎟.
⎟ (6.5.3)
⎜ .. .. ..
⎜ . . . . 0 ⎟ ⎟
⎜ .. ⎟
⎝ 0 ... ... . 0 B/J ⎠
J/J 2 0 ... ... 0 0
Consider the following two chains of inclusions:
(i) B ⊃ J ⊃ J 2;
Proof. The proof follows from (6.5.4). Indeed, in this case we have
χ[Q(Hn (B)] (x) = det(xn E − [Q(B)]), where χM (x) is the characteristic polyno-
mial of a square matrix M .
Example 6.5.1.
Let ⎛ ⎞
1 ... 1
⎜ ⎟
Un = ⎝ ... . . . ... ⎠
1 ... 1
292 ALGEBRAS, RINGS AND MODULES
be the square n × n-matrix all of whose entries are 1, and let En ∈ Mn (R) be the
identity matrix.
Let Kn (O) be the following tiled order: Kn (O) = {O, E(Kn (O))}, where
E(Kn (O)) = Un − En . Obviously, the corresponding matrix (βij ) = Un and √
so [Q(Kn (O))] = Un and inx Kn (O) = n. If n → ∞ then inx Hn (Kn (O)) = n n
goes to 1.
Proof. We can assume that A is reduced. Suppose that inx A = 1 and in the
i-th row of [Q(A)] there are two elements 1. By theorem 14.6.1, vol.I, Q(A) is
strongly connected and, by remark 6.5.1, inx A > 1. So in each row of A there
is only one 1. The matrices [Q(A)]T and [Q(A)] are permutationally irreducible
simultaneously and their maximal real eigenvalues coincide. Consequently, in each
column of A there is only one 1. Thus [Q(A)] = Pσ for some σ ∈ Sn . From the
indecomposability of A and theorem 11.1.9, vol.I, σ is a cycle. By corollary 12.3.7
and theorem 12.3.8, vol.I, A is Morita-equivalent to Hn (O). It is obvious that
inx O equals 1 for a discrete valuation ring O. It follows from proposition 6.5.6
that inx Hn (O) = 1. The theorem is proved.
Theorem 6.5.7. Let A be a tiled order and suppose Q(A) contains n vertices.
then 1 ≤ inx A ≤ n and for any integer k (1 ≤ k ≤ n) there exists a tiled order Ak
with inx Ak = k.
S 1 , S 2 , . . . , Sn (6.6.1)
and a sequence of instants
t 0 , t1 , t2 , . . . .
TILED ORDERS OVER DISCRETE VALUATION RINGS 293
Suppose that at each of these instants the system is in one and only one of the
states (6.6.1) and that pij denotes the probability of finding the system in the state
Sj at the instant tk if it is known that at the preceding instant tk−1 the system
was in the state Si (i, j = 1, 2, . . . , n; k = 1, 2, . . .). If the transition probabilities
pij (i, j = 1, 2, . . . , n) do not depend on the index k (of the instant tk ), then the
process is called a homogeneous Markov chain with a finite number of states.
The matrix
P = (pij ) ∈ Mn (R)
is called the transition matrix for the Markov chain. From the above assump-
tions it is obvious that
n
pij ≥ 0, and pij = 1 (i, j = 1, 2, . . . , n). (6.6.2)
j=1
Thus, every stochastic matrix can be regarded as the transition matrix for a
finite (homogeneous) Markov chain and, conversely, the transition matrix for such
a Markov chain is stochastic.
Proof. Let S ∈ Mn (R) be a doubly stochastic matrix. Suppose that the quiver
Q(S) is connected but non-strongly
connected.
Then there exists a permutation
T S1 X
matrix Pτ such that Pτ SPτ = .
0 S2
294 ALGEBRAS, RINGS AND MODULES
The matrix PτT SPτ is also doubly stochastic as a product of the doubly stochas-
tic matrices. Therefore S1T and S2 are stochastic matrices (or because, obviously,
if S is doubly stochastic so are SP and P S for any permutation matrix P ).
Let S1 ∈ Mm (R) and S2 ∈ Mn−m (R), and m ≥ 1. Denote by Σ(Y ) the
sum of all elements of an arbitrary matrix Y ∈ Mn (R). Obviously, Σ(PτT SPτ ) =
Σ(S1 ) + Σ(S2 ) + Σ(X). For any stochastic matrix S ∈ Mn (R), the equality
Σ(S) = Σ(S T ) = n holds. This sum does not change under a simultaneous
transposition of rows and columns. Hence, Σ(PτT SPτ ) = n. Clearly, S1T and
S2 are stochastic matrices. Consequently, n = m + n − m + Σ(X). Whence,
Σ(X) = 0 and X = 0. Thus, the doubly stochastic matrix S is permutationally
decomposable. This completes the proof.
Let Q be a quiver with adjacency matrix [Q] = (qij ). We shall refer to the
eigenvectors (resp. eigenvalues) of [Q] as the eigenvectors (resp. eigenvalues) of
the quiver Q. If Q is strongly connected, then the index of Q (written inx Q) is
the maximal real eigenvalue of [Q]; its eigenvector
f = (f1 , . . . , fn )T
Theorem 6.6.2.5 For any Frobenius quiver Q there exists a stochastic matrix
P such that Q(P ) = Q.
n
n
pij = λ−1 zi−1 qij zj = λ−1 zi−1 λzi = 1. Obviously, [Q(P )] = [Q].
j=1 j=1
The Markov chain with this stochastic matrix is called the Markov chain of the
Frobenius quiver Q.
It follows from the Perron-Frobenius theorem6 and corollary 11.3.3, vol.I that
every strongly connected quiver is Frobenius.
Example 6.6.1.
Let
0 1
P = .
0 1
Then
Q(P ) = • •
1 2
is a Frobenius quiver.
Example 6.6.2.
⎛ ⎞ ⎡ ⎤
1 0 0 0 1 0 0 0
⎜ 0 1/2 1/2 0 ⎟ ⎢ 0 1 1 0 ⎥
Let P = ⎜⎝ 0 1/2
⎟. Then [Q(P )] = ⎢
⎠ ⎣
⎥.
1/2 0 0 1 1 0 ⎦
1/4 1/4 1/4 1/4 1 1 1 1
Obviously, χ[Q(P )] = x(x − 1)2 (x − 2) and we have
⎛ ⎞⎛ ⎞ ⎛ ⎞
1 0 0 0 0 0
⎜0 1 1 0⎟ ⎜1⎟ ⎜1⎟
⎜ ⎟⎜ ⎟ ⎜ ⎟
⎝0 1 1 0⎠ ⎝1⎠ = 2 ⎝1⎠ .
1 1 1 1 2 2
Consequently, the quiver of a Markov chain is not necessarily Frobenius.
Example 6.7.1.
The index of a finite linearly ordered set CHn is 1.
Example 6.7.2.
Let
" #
1 2 3 ... n−1 n
ACHn =
• • • ... • •
be an antichain of width n. Clearly, Q(ACH n ) is a complete simply laced quiver
with n vertices. Thus inxACHn = n.
Example 6.7.3.
Let Pm,n = (m, m, . . . , m) be a primitive poset formed by n linearly
√ ordered
disjoint sets each of length m. It is easy to verify that inx Pm,n = m n.
TILED ORDERS OVER DISCRETE VALUATION RINGS 297
Example 6.7.4.
Consider ⎧ ⎫
⎪ • •⎪
⎨ ⎬
P4 = .
⎪
⎩ ⎪
⎭
• •
Denote by ⎛ ⎞
1 ... 1
⎜ .. .. .⎟
Un = ⎝ . . .. ⎠
1 ... 1
the square n × n-matrix whose all entries are 1. Obviously, the adjacency matrix
4 ) is
Q(P
0 U2
[Q(P4 )] =
U2 0
and inx P4 = 2
Example 6.7.5.
Let
⎧ ⎫
⎪
⎪ 1 3 5 2n − 3 2n − 1 ⎪
⎪
⎪
⎪ ⎪
⎪
⎪
⎪ ⎪
⎪
⎪
⎪ • • • ... • • ⎪ ⎪
⎪
⎨ ⎪
⎬
P2n = .
⎪
⎪ ⎪
⎪
⎪
⎪ ⎪
⎪
⎪
⎪ • • • ... • • ⎪
⎪
⎪
⎪ ⎪
⎪
⎪
⎩ ⎪
⎭
2 4 6 2n − 2 2n
Obviously,
⎡ ⎤
0 U2 0 ... ... 0
⎢ .. .. ⎥
⎢ 0 0 U2 . . ⎥
⎢ ⎥
⎢ .. .. .. .. .. .. ⎥
2n )] = ⎢
[Q(P ⎢ . . . . . . ⎥
⎥
⎢ .. .. .. .. ⎥
⎢ . . . 0 ⎥
⎢ . ⎥
⎣ 0 ... 0 U2 ⎦
U2 0 ... ... 0 0
and inx P2n = 2.
298 ALGEBRAS, RINGS AND MODULES
Example 6.7.6.
Let
2 3
• •
N: (6.7.1)
• •
4 1
be a partially ordered set with 4 elements. Obviously,
⎛ ⎞
0 1 1 0
⎜1 0 0 1⎟
[Q(EN )] = ⎜
⎝1
⎟
0 0 1⎠
0 1 0 0
and
χB (x) = x2 (x2 − 3).
√
So inx (N ) = 3 and the Frobenius eigenvector is
√ √
f = (2, 3, 3, 1)T .
The√ numeration (6.7.1) of N is standard with Frobenius eigenvector
√
(2, 3, 3, 1)T . The transition matrix TN of the Markov chain associated with
N is
⎛ ⎞⎛ ⎞⎛ ⎞
1/2 0√ 0 0 0 1 1 0 2 √0 0 0
√ ⎜ 0 1/ 3 0√ 0⎟ ⎜ ⎟⎜ 3 √0 0⎟
TN = 1/ 3 ⎜ ⎟ ⎜1 0 0 1⎟ ⎜0 ⎟=
⎝ 0 0 1/ 3 0⎠ ⎝1 0 0 1⎠ ⎝0 0 3 0⎠
0 0 0 1 0 1 0 0 0 0 0 1
⎛ ⎞
0 1/2 1/2 0
⎜2/3 0 0 1/3⎟
=⎜⎝2/3 0
⎟.
0 1/3⎠
0 1 0 0
Evidently, TN defines an ergodic cyclic Markov chain. So, the poset N is cyclic.
The numeration of N :
4 3 (6.7.2)
• •
• •
2 1
TILED ORDERS OVER DISCRETE VALUATION RINGS 299
is nonstandard and this is in accordance with the Frobenius theorem (see, [Gant-
makher, 1998], Section XIII, § 2).
Indeed, let EN be an exponent matrix of N corresponding the numeration
(6.7.2). We have ⎛ ⎞
0 0 1 1
⎜0 0 0 1⎟
[Q(EN )] = ⎜
⎝1
⎟
1 0 0⎠
1 1 0 0
and √ the √Frobenius eigenvector corresponding the numeration (6.7.2) is
(2, 1, 3, 3)T . So, the numeration (6.7.2) is not standard.
Below there is the list of indexes and Frobenius vectors of posets with at most
four elements.
Remark 6.7.1. Obviously, inx CHn = 1 and inx ACHn = n. The vector
(1, . . . , 1)T is the Frobenius vector of CHn and ACHn .
⎧ ⎫
⎧ ⎫ ⎪
⎪ 4 ⎪
⎪
⎪ •⎪
⎪ ⎪
⎪ ⎪
⎪
⎪
⎪ ⎪ ⎪
⎪ • ⎪
⎪
⎪
⎨•⎪⎬ ⎪
⎨ ⎪
⎬
IV. (1) = , inx (IV, 1) = 1; (2) = 2 • • 3 , inx (IV, 2) =
⎪
⎪ •⎪⎪ ⎪
⎪ ⎪
⎪
⎪
⎪ ⎪ ⎪ ⎪
⎩ ⎪ ⎭ ⎪
⎪
⎪ • ⎪
⎪
⎪
• ⎪
⎩ ⎪
⎭
1
√
3
2;
⎧ ⎫ ⎧ ⎫
⎪
⎪2 3⎪
⎪ ⎪
⎪ 1 ⎪
⎪
⎪
⎪ • •⎪
⎪ ⎪
⎪ • ⎪
⎪
⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪
⎪
⎨ ⎪
⎬ ⎪
⎨ ⎪
⎬
1
• √
3
(3) = • , (4) = , inx (IV, 3) = inx (IV, 4) = 2;
⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪
⎪
⎪ ⎪
⎪ ⎪
⎪
4 ⎪
⎪
⎪
⎪ • ⎪
⎪ ⎪
⎪ • • ⎪
⎪
⎪
⎩ ⎪
⎭ ⎪
⎩ ⎪
⎭
4 2 3
√ √
f(IV, 2) = f(IV, 3) = f(IV, 4) = ( 3 4, 1, 1, 3 2)T
⎧ ⎫ ⎧ ⎫
⎪
⎪ • 3⎪⎪ ⎪
⎪ • 1⎪⎪
⎪
⎨ ⎪
⎬ ⎪
⎨ ⎪
⎬
(5) = 4 • • 2 , (6) = 4 • • 3 ; χ5,6 (x) = x(x3 − x − 1) and
⎪
⎪ ⎪
⎪ ⎪
⎪ ⎪
⎪
⎪
⎩ ⎪
⎭ ⎪
⎩ ⎪
⎭
• 1 • 2
1.32 < inx (IV, 5) = inx (IV, 6) < 1.33; f(IV, 5) = f(IV, 6) = (λ2 , 1, λ, λ)T ,
where λ3 − λ − 1 = 0.
⎧ ⎫
⎪
⎪ 3 4⎪
⎪
⎪
⎨• •⎪⎬ √ √
√ 2 2
(7) =
, inx (IV, 7) = 2; f (IV, 7) = ( , , 1, 1)T ;
⎪
⎪ • • ⎪
⎪ 2 2
⎪
⎩ ⎪
⎭
⎧ 1 2 ⎫
⎪
⎪ • 4⎪ ⎪
⎪
⎨ ⎪
⎬
(8) = • 3 , χ8 (x) = x(x3 − x2 − 1) and 1.46 < inx (IV, 8) < 1.47;
⎪
⎪ ⎪
⎪
⎪
⎩ ⎪
⎭
1 • • 2
f(IV, 8) = (1, λ − 1, λ2 − λ, 1)T , where λ3 − λ2 − 1 = 0.
⎧ ⎫
⎪
⎪ 4 2⎪
⎪
⎪
⎨• •⎪⎬ √ √ √
(9) = , inx (IV, 9) = 3; f(IV, 9) = (2, 3, 1, 3)T .
⎪
⎪ • •⎪⎪
⎪
⎩ ⎪
⎭
3 1
TILED ORDERS OVER DISCRETE VALUATION RINGS 301
⎧ ⎫ ⎧ ⎫
⎪
⎪ 4 ⎪
⎪ ⎪
⎪ 1 2 3⎪⎪
⎪
⎨ • ⎪
⎬ ⎪
⎨• • •⎪ ⎬ √
(10) = , (11) = , inx (IV, 10) = inx (IV, 11) = 3;
⎪
⎪ • • •⎪ ⎪ ⎪
⎪ • ⎪
⎪
⎪
⎩ ⎪
⎭ ⎪
⎩ ⎪
⎭
1 2 3 4
√ √ √
f(IV, 10) = f(IV, 11) = ( 3, 3, 3, 3)T .
⎧ ⎫ ⎧ ⎫
⎪
⎪ 4 ⎪
⎪ ⎪
⎪1 2 3⎪ ⎪
⎪
⎨ • ⎪
⎬ ⎪
⎨• • •⎪ ⎬
(12) = , (13) = ,
⎪
⎪ • • •⎪ ⎪ ⎪
⎪ • ⎪
⎪
⎪
⎩ ⎪
⎭ ⎪
⎩ ⎪
⎭
1 2 3 4
inx (IV, 12) = inx (IV, 13) = 2; f(IV, 12) = (2, 1, 1, 2)T ; f(IV, 13) =
(1, 1, 1, 1)T .
⎧ ⎫
⎪
⎪ 3 4⎪⎪
⎪
⎨• •⎪ ⎬
(14) = , inx(IV, 14) = 2; f(IV, 14) = (1, 1, 1, 1)T ;
⎪
⎪ • • ⎪
⎪
⎪
⎩ ⎪
⎭
1 2
⎧ ⎫
⎪
⎪ 4⎪ ⎪
⎪
⎨ •⎪ ⎬ √
(15) = , χ15 (x) = x2 (x2 − 2x − 1) and inx (IV, 15) = 1 + 2;
⎪
⎪ • • •⎪ ⎪
⎪
⎩ ⎪
⎭
1 2 3
√ √ √
f(IV ), 15) = (1 + 2, 1 + 2, 1, 1 + 2)T ;
(16) = { • • • • }, inx (IV, 16) = 4.
Note that the posets (IV, 2), (IV, 3) and (IV, 4) are Q-equivalent. The other
non-singleton Q-equivalence classes are {(III, 2), (III, 3)}, {(IV, 10), (IV, 11)}.
√
For the posets N = (IV, 9) and F4 = (IV, 11) we have inx N = inx F4 = 3,
but N and F4 are not Q-equivalent.
The posets (IV, 12) and (IV, 13) are antiisomorphic, and w(IV, 12) =
w(IV, 13) = 3 but (IV, 12) and (IV, 13) are not Q-equivalent, because
f(IV, 12) = (2, 1, 1, 2)T and f(IV, 13) = (1, 1, 1, 1)T . The posets (IV, 13)
and (IV, 14) have both index and Frobenius vector equal but they are not
Q-equivalent because Q(IV,
13) has a loop and Q(IV, 14) does not.
Proof. In what follows we assume that exponent matrices are reduced and their
first rows are zero, which can be done because we are after admissible quivers,
which are quivers coming from reduced exponent
matrices..
0 0 (1) 1 0
Let s = 2. Then E = , E = and E (2) =
α 0 α 1
(2, α) (1, α)
, where (α1 , . . . , αt ) = min(α1 , . . . , αt ). Here, quite generally, as
α + 1 (2, α)
before, see just below corollary 6.1.14 in section 6.1, the following notation is used.
If E is an (reduced) exponent matrix, E = (αij ), then E (1) = (βij ) with βii = 1 and
(2) (2) (1)
βij = αij if i =
j, E = (γij ), γij =minK (βik + βkj ), so that [Q(E)] = E − E .
(1, α − 1) 1
So [Q(E)] = and Q(E) is either C2 for α = 1 or LC2 for
1 (1, α − 1)
α ≥ 2 (Cn is a simple cycle with n vertices, [LCn ] = [Cn ] + En , En is the identity
n × n matrix). ⎛ ⎞ ⎛ ⎞
0 0 0 1 0 0
Let s = 3. Then E = ⎝α 0 δ ⎠ , E (1) = ⎝α 1 δ ⎠ and E (2) =
β γ 0 β γ 1
TILED ORDERS OVER DISCRETE VALUATION RINGS 303
⎛ ⎞
(2, α, β) (1, γ) (1, δ)
⎝ (α + 1, β + δ) (2, α, γ + δ) (α, δ + 1) ⎠.
(β + 1, α + γ) (β, γ + 1) (2, β, γ + δ)
Obviously, one can suppose 1 ≤ α ≤ β. Then α = 1. Indeed, if α ≥ 2 we have a
loop in the first vertex. If β = 1 we have either E ∼ H3 or E ∼ F3 . Obviously,
Ω3 ∼ F3 . If β = 2 then E ∼ Ω3 .
Let s = 4. As above we obtain the admissible quivers without loops, listed be-
low. The notation E ∼ Θ means equivalence of these matrices by transformations
of the first type.
n
We put d = d(E) = αij for an exponent matrix E = (αij ). Obviously,
i,j=1
d(E) = d(Θ) for equivalent reduced exponent matrices E and Θ.
It is convenient to place the first six exponent matrices in the following se-
quence: ⎛ ⎞ ⎛ ⎞
0 0 0 0 0 1 0 0
⎜1 0 0 0 ⎟ ⎜0 0 1 0⎟
(1) d = 6, E1 = H4 = ⎜ ⎟ ⎜
⎝1 1 0 0⎠, [Q(E1 )] = ⎝0 0 0 1⎠;
⎟
1 1 1 0 1 0 0 0
⎛ ⎞ ⎛ ⎞
0 0 0 0 0 1 0 0
⎜1 0 0 0⎟ ⎜ ⎟
(2) d = 7, E2 = ⎜ ⎟, [Q(E2 )] = ⎜0 0 1 0⎟
⎝1 1 0 0⎠ ⎝1 0 0 1 ⎠
2 1 1 0 0 1 0 0
⎛ ⎞
0 1 1 1
⎜0 0 0 0 ⎟
and E2 ∼ Θ2 , where Θ2 = ⎜ ⎝0 1 0 0⎠;
⎟
1 1 1 0
⎛ ⎞ ⎛ ⎞
0 0 0 0 0 1 0 0
⎜1 0 0 0⎟ ⎜1 0 1 0 ⎟
(3) d = 8, E3 = ⎜ ⎟ ⎜
⎝2 1 0 0⎠, [Q(E3 )] = ⎝0 0 0 1⎠
⎟
2 1 1 0 0 1 0 0
⎛ ⎞
0 1 1 1
⎜0 0 0 0 ⎟
and E3 ∼ Θ3 , where Θ3 = ⎜ ⎝0 1 0 0⎠;
⎟
1 1 1 0
⎛ ⎞ ⎛ ⎞
0 0 0 0 0 1 0 0
⎜1 0 0 0⎟ ⎜1 0 1 0 ⎟
(4) d = 9, E4 = ⎜ ⎟ ⎜
⎝2 1 0 0⎠, [Q(E4 )] = ⎝0 1 0 1⎠
⎟
2 2 1 0 1 1 1 0
304 ALGEBRAS, RINGS AND MODULES
⎛ ⎞
0 0 1 1
⎜1 0 1 1 ⎟
and E4 ∼ Θ4 , where Θ4 = ⎜
⎝1 0 0 0⎠;
⎟
1 1 1 0
⎛ ⎞ ⎛ ⎞
0 0 0 0 0 1 0 0
⎜1 0 0 0⎟ ⎜1 0 1 0⎟
(5) d = 10, E5 = Ω4 = ⎜ ⎟ ⎜
⎝2 1 0 0⎠, [Q(Ω4 )] = ⎝0 1 0 1⎠;
⎟
3 2 1 0 0 0 1 0
⎛ ⎞ ⎛ ⎞
0 0 0 0 0 1 1 1
⎜1 0 1 1⎟ ⎜1 0 0 0⎟
(6) d = 9, E6 = F4 = ⎜ ⎟ ⎜
⎝1 1 0 1⎠, [Q(F4 )] = ⎝1 0 0 0⎠;
⎟
1 1 1 0 1 0 0 0
⎛ ⎞ ⎛ ⎞
0 0 0 0 0 1 1 0
⎜1 0 1 0⎟ ⎜1 0 0 1 ⎟
(7) d = 8, E7 = ⎜ ⎟ ⎜
⎝1 1 0 0⎠, [Q(E7 )] = ⎝1 0 0 1⎠ ,
⎟
2 1 1 0 0 1 1 0
⎛ ⎞
0 0 0 1
⎜1 0 1 1 ⎟
and E7 ∼ Θ7 , where Θ7 = ⎜
⎝1 1 0 1⎠;
⎟
1 0 0 0
⎛ ⎞ ⎛ ⎞
0 0 0 0 0 1 0 0
⎜1 0 0 0 ⎟ ⎜1 0 1 0 ⎟
(8) d = 10, E8 = ⎜ ⎟ ⎜
⎝2 2 0 0⎠, [Q(E8 )] = ⎝0 0 0 1⎠;
⎟
2 2 1 0 1 0 1 0
⎛ ⎞ ⎛ ⎞
0 0 0 0 0 1 1 0
⎜1 0 1 0⎟ ⎜1 0 0 1 ⎟
(9) d = 10, E9 = ⎜ ⎟ ⎜ ⎟
⎝2 1 0 0⎠, [Q(E9 )] = ⎝0 1 0 1⎠, σ(E9 ) =
2 2 1 0 1 0 1 0
(1423);
⎛ ⎞ ⎛ ⎞
0 0 0 0 0 1 1 0
⎜1 0 1 0⎟ ⎜1 0 0 1⎟
(10) d = 11, E10 = ⎜ ⎟ ⎜
⎝2 2 0 0⎠, [Q(E10 )] = ⎝0 0 0 1⎠;
⎟
2 2 1 0 1 0 1 0
⎛ ⎞ ⎛ ⎞
0 0 0 0 0 1 1 0
⎜1 0 1 1⎟ ⎜1 0 0 0⎟
(11) d = 11, E11 = ⎜ ⎟ ⎜
⎝2 2 0 0⎠, [Q(E11 )] = ⎝0 0 0 1⎠.
⎟
2 2 1 0 1 0 1 0
TILED ORDERS OVER DISCRETE VALUATION RINGS 305
Clearly, any prime ring is weakly prime and a weakly prime ring is indecom-
posable.
Proof. Let I and J be two-sided ideals in eAe which are not contained in
the Jacobson radical eRe of the ring eAe (see proposition 3.4.8, vol.I), and let
1 = e + f . Consider the following two-sided ideals
Proof. Assume that the ring A is weakly prime and Apq = 0 for some p = q.
Consider the ring C = App +Aqq +Aqp . Then, Z = App +Aqp and N = Aqq +Aqp
are two-sided ideals of the ring C that do not belong to the Jacobson radical of C
and such that ZN = 0. But this contradicts lemma 6.9.1.
We shall show that, if all Aij = 0, then A is weakly prime. There exists a
decomposition of the identity of the ring A into the sum of mutually orthogonal
idempotents 1 = f1 + . . . + fs such that fi Afi = Mni (Oi ), with local rings Oi ,
i = 1, . . . , s, and fi Afj = fi Rfj (i = j, i, j = 1, . . . , s) (see proposition 11.1.1,
vol.I).
306 ALGEBRAS, RINGS AND MODULES
The proof follows from theorem 11.6.3, vol.I and theorem 6.9.2.
Proposition 6.9.5. Let A be a tiled order and B = A/πA. The quiver Q(B)
of the ring B is obtained from the quiver Q(A) of the tiled order A by deleting all
loops.
Recall that Q(A) is simply laced and strongly connected. Let there exist an
arrow from i to j in Q(A) (i = j). This means that eii Rejj = π αij O and eii R2 ejj =
π αij +1 O, where R is the Jacobson radical of A. Therefore in Q(B) there exists an
arrow from i to j (i = j), and so in Q(A) there exists an arrow from i to j.
Let O = k[[x]] be the ring of formal power series over a field k. We know that
O is a discrete valuation ring with a prime element x. Let A = {k[[x]], E(A) =
(αij )} be a reduced tiled order. Then A/xA is an n2 -dimensional weakly prime
semidistributive algebra over a field k, where n is the number of vertices in Q(A).
Proof. Consider the quiver Q1 with the following adjacency matrix: [Q1 ] =
[Q] + En . By theorem 6.1.16, every strongly connected simply laced quiver Q1
with a loop in each vertex is admissible. So there exists a reduced exponent
matrix E = (αij ) such that Q(E) = Q1 . Let O be a discrete valuation ring with a
prime element π and let A = {O, E} be a tiled order. Obviously, by proposition
6.9.5, the quiver Q(B) coincides with Q, where B = A/πA.
Corollary 6.9.7. For any simply laced quiver Q with n vertices and for any
field k there exists a weakly prime semidistributive n2 -dimensional algebra B over
k such that Q(B) = Q.
Proof. By lemma 6.9.8 we can assume that A = P1n ⊕ . . . ⊕ Psn . But then
A Mn (End (P1 ⊕ . . . ⊕ Ps )) and the ring B = End A (P1 ⊕ . . .⊕ Ps ) is a basic ring,
i.e., the quotient ring of B by its Jacobson radical is a direct product of division
rings.
TILED ORDERS OVER DISCRETE VALUATION RINGS 309
(a) 1 • • 2
(b) 1 2
• •
TILED ORDERS OVER DISCRETE VALUATION RINGS 311
(c) 1 2
• •
Case (a). In this case the ring B is serial, by theorem 12.3.11, vol.I. We shall
study in which cases B is not Artinian. Let 1 ∈ B and let 1 = e1 + e2 be a
decomposition of 1 into a sum of local pairwise orthogonal idempotents. Write
R = rad B, Bi = ei Bei and Ri = rad (Bi ), X = e1 Be2 , Y = e2 Be1 , where B1
and B2 are uniserial rings, X is an uniserial right B2 -module and an uniserial left
B1 -module; Y is an uniserial right B1 -module and an uniserial left B2 -module, by
theorem 14.2.1, vol.I. If B1 and B2 are Artinian, then B is Artinian. Let R be the
Jacobson radical of B. Then, as usual,
2
R1 X R1 + XY R1 X + XR2
R = and R2 = .
Y R2 Y R1 + R2 Y Y X + R22
In this case there are the following countable descending Loewy series for P1 =
e1 B and P2 = e2 B:
and
Proof. If Q(B) is strongly connected and B is weakly prime then the proof
follows from the consideration of the cases: (a), (b) and (c) above. Let B be
not weakly prime. Up to renumbering we may assume that Y = 0. If X = 0,
then B = B 1 × B2 , where B is serial (B1 and B2 are uniserial). If X = 0, then
0 X
L= is a two-sided ideal of B, L2 = 0 and B/L B1 × B2 is serial.
0 0
Let A be a ring. Denote μ∗r (A) = max μA (I), where I is a right ideal A.
I⊆A
Analogously, one can define μ∗l (A). By definition, A is a right principal ideal ring
if and only if μ∗r (A) = 1.
O K
B= .
K K
We describe the multiplication and addition in B. Denote by e11 , e12 , e21 , e22
the matrix units of B: e12 e21 = 0 and e21 e12 = 0. Let ϕ : O → K be the
canonical epimorphism. If α ∈ O, then (αe11 )e12 = ϕ(α)e12 = e12 ϕ(α) and
e21 (αe11 ) = e21 ϕ(α) = ϕ(α)e21 . Further, αe11 = e11 α for α ∈ O and βe22 = e22 β
for β ∈ K. The multiplication in K is defined as multiplication of 2 × 2-matrices
and the addition is defined elementwise. It is easy to see that μ∗r (B) = 2 and
μ∗l (B) = ∞.
Proof. Let J1 , J2 ∈ I(A). Then (J1 +J2 )2 = J12 +J1 J2 +J2 J1 +J22 = J1 +J2 .
Obviously, for any idempotent e ∈ A the ideal AeA is idempotent. This follows
from the following inclusions: AeAAeA ⊂ AeA and AeA ⊂ AeAAeA.
0 = J0 ⊂ J1 ⊂ . . . ⊂ Jt−1 ⊂ Jt ⊂ . . . ⊂ Jm = A
Let Λ = {O, E(Λ)} be a tiled order over a discrete valuation ring O and let
Mn (D) be its classical ring of fractions, where D is the classical division ring of frac-
tions of O, and write E(Λ) = (αij ). Let E(Λ)T = (αji ) and ΛT = {O, E(Λ)T }.
Then the following proposition is obvious.
314 ALGEBRAS, RINGS AND MODULES
Proof. The proof follows from the equality gl.dim ΛT = l.gl.dim Λ and from
Auslander’s theorem 5.1.16, which asserts that l.gl.dim Λ = r.gl.dim Λ if Λ is
two-sided Noetherian.
We now consider the global dimension of tiled orders having finite global di-
mension and width at most 2.
Note that if two tiled orders A1 and A2 are Morita equivalent, then gl. dim A1 =
gl. dim A2 . So, we can assume (for global dimension considerations) that a tiled
order Λ with finite global dimension is reduced.
Let Λ be a reduced tiled order of finite global dimension with w(Λ) = 2. The
following theorem is stated without proof.
Since, by this theorem, any tiled order of width at most 2 is a direct sum
of irreducible lattices, to determine the global dimension of such a tiled order it
is sufficient to check the projective dimensions of irreducible Λ-lattices, by the
Auslander theorem (theorem 5.1.13).
It is obvious that the notion of the diagram of a finite poset (see vol. I, p.279
and section 6.3 above) may be extended in the same way to M(Λ), where M(Λ) is
the partially ordered set, which is formed by all irreducible projective A-lattices.
We shall denote this infinite quiver by Q(M(Λ)).
Lemma 6.10.5. Let Λ be a tiled order of width at most two. If gl. dim Λ < ∞,
then there exists an element P ∈ M(Λ) such that only one arrow in Q(M(Λ))
starts from P .
Q1 • • Q2 R1 • • R2
• •
P1 P2
Proof. By lemma 6.10.5, there exists an element P ∈ M(Λ) such that only one
arrow in Q(M(Λ)) issues from P . Let P = eΛ, and 1 = e + f . If af, bf ∈ M(f Λf )
is a pair of non-comparable elements in M(f Λf ), then a, b ∈ M(Λ) is a pair of
non-comparable elements in M(Λ). Consider a projective resolution of a + b as
a Λ-module. To a projective resolution of a + b we can assign a sequence of non-
comparable pairs (a1 , b1 ), . . . , (ak , bk ), where aj ∈ L1 , bj ∈ L2 for j = 1, . . . , k and
from ak and bk there start two arrows. Let ak ∩ bk be projective. Obviously, two
arrows start from ak ∩ bk . Consequently, a, b, a1 , b1 , . . . , ak , bk ∈ M(Λ) \ {Pn }. By
lemma 6.10.6, (af, bf ), (a1 f, b1 f ), . . . , (ak f, bk f ) corresponds to a finite projective
resolution of the f Λf -module af + bf . So, gl. dim f Λf ≤ gl. dim Λ. Conversely, in
(1) (1) (k) (k) (k) (k)
a sequence (P1 , P2 ), (P1 , P2 ), . . . , (P1 , P2 ), where P1 ∩ P2 is projective,
(1) (1) (k) (k) (k) (k)
all the modules P1 , P2 , . . . , P1 , P2 , P1 ∩ P2 ∈ M(Λ) \ {Pn }. Therefore,
gl. dim f Λf ≥ gl. dim Λ − 1.
The proof of this theorem follows from lemma 6.10.7 and the fact that all
tiled orders in M2 (D) of finite global dimension are isomorphic to H2 (O) with
gl. dim H2 (O) = 1.
Example 6.10.1.
The tiled order Λn = {O, E(Λn )}, where
⎛ ⎞
0 0 ... ... ... 0
⎜ .. .. .. ⎟
⎜1 . . .⎟
⎜ ⎟
⎜ .. .. .. .. ⎟
⎜2 . . . .⎟
E(Λn ) = ⎜
⎜. .⎟
⎟
⎜ .. .. . . .
⎜ . . . . . . . .. ⎟ ⎟
⎜. .. .. .. ⎟
⎝ .. . . . 0⎠
2 ... ... 2 1 0
is an n × n-matrix, is a triangular tiled order of width 2 and gl.dim Λn = n − 1.
Example 6.10.2.
The tiled order Ωn = {O, E(Ωn )}, where
⎛ ⎞
0 0 ... ... ... ... ... 0
⎜ .. ⎟
⎜ .. .. ⎟
⎜ 1 . . .⎟
⎜ ⎟
⎜ .. .. .. .. ⎟
⎜ 2 . . . .⎟
⎜ ⎟
⎜ .. .. .. .. .. ⎟
⎜ 3 . . . . .⎟
⎜ ⎟
E(Ωn ) = ⎜ . .. ⎟
⎜ . . .. . .. .. .. .. ⎟
⎜ . . . . .⎟
⎜ ⎟
⎜ .. .. .. .. .. .⎟
⎜n − 3 ... . . . . . .. ⎟
⎜ ⎟
⎜ .. .. .. .. ⎟
⎜n − 2 n − 3 ... . . . . 0⎟
⎝ ⎠
n−1 n−2 n−3 ... 3 2 1 0
The following proposition is very useful. The proof of it can be found in the
paper [Kirkman, Kuzmanovich, 1989].
and gl.dim Δ5 ≥ gl.dim Λ/I = 4. From theorem 5.10.4 it follows that gl.dim Δ5 =
4. Let f = e11 + e22 + e33 . Then
⎛ ⎞
0 0 0 1 0
⎜ 1 0 0 1 1 ⎟
⎜ ⎟
E(J ) = E(Δ5 f Δ5 ) = ⎜
⎜ 1 1 0 1 1 ⎟.
⎟
⎝ 1 1 0 1 1 ⎠
2 1 1 2 2
From this example it follows that both equalities in proposition 6.10.10 may
hold.
Theorem 6.10.11.11 If Λ is a tiled order and gl.dim Λ < ∞, then Q(Λ) has
no loops.
10 see [Dlab, Ringel, 1989]
11 see [Weidemann, Roggenkamp, 1983]
TILED ORDERS OVER DISCRETE VALUATION RINGS 319
Theorem 6.10.12. If Λ is a tiled order and Q(Λ) has at most 3 vertices, then
gl.dim Λ is finite if and only if Q(Λ) has no loops. In this case w(Λ) ≤ 2.
A proof follows from theorem 6.10.11 and proposition 6.7.1. If Q(Λ) is a cycle,
then Λ is hereditary and gl.dimΛ = 1. If Ω3 (O) = {O, Ω3 }, then gl.dim Ω3 (O) = 2.
Remark 6.10.2. This theorem was first proved by R.B.Tarsy (see [Tarsy,
1970]).
The list of the orders Λ with gl.dim Λ < ∞ and such that Q(Λ) has 4 vertices
is given in the papers [Fujita, 1990], [Fujita, 1991]. The first six exponent matrices
(1)-(6) from section 6.8 exhaust this list.
⎛ ⎞
0 0 0 0
⎜ 1 0 1 1 ⎟
Recall that E(F4 ) = ⎜ ⎟
⎝ 1 1 0 1 ⎠ . Obviously, w(F4 ) = 3.
1 1 1 0
Note, that all tiled orders of finite global dimension, whose quivers have at
most four vertices, are isomorphic to (0, 1)-orders, except Ω4 . Now we give a list
of the associated posets PΛ , where gl.dim Λ < ∞ and Λ is a (0, 1)-order.
List of posets:
n = 1, P1 = {•}, gl.dim ΛP1 = 1;
⎧ ⎫
⎨ • ⎬
n = 2, P2 = | , gl.dim ΛP2 = 1;
⎩ ⎭
•
⎧ ⎫
⎪
⎪ • ⎪
⎪
⎪
⎪ ⎪
⎪
⎨ | ⎬
n = 3, P3 = • , gl.dim ΛP3 = 1;
⎪
⎪ ⎪
⎪
⎪ |
⎪ ⎪
⎪
⎩ ⎭
•
⎧ ⎫
⎨ • • ⎬
n = 3, P4 = , gl.dim ΛP4 = 2;
⎩ ⎭
•
⎧ ⎫
⎪
⎪ • ⎪
⎪
⎪
⎪ ⎪
⎪
⎪ | ⎪
⎪
⎪
⎪
⎪ ⎪
⎨ • ⎪
⎬
n = 4, P5 = | , gl.dim ΛP5 = 1;
⎪
⎪ ⎪
⎪
⎪
⎪ • ⎪
⎪
⎪
⎪ ⎪
⎪
⎪ | ⎪
⎪
⎩ ⎪ ⎭
•
320 ALGEBRAS, RINGS AND MODULES
⎧ ⎫
⎪
⎪ • ⎪
⎪
⎪
⎪ ⎪
⎪
⎨ | ⎬
n = 4, P6 = • • , gl.dim ΛP6 = 2;
⎪
⎪ ⎪
⎪
⎪
⎪ | ⎪
⎪
⎩ ⎭
•
⎧ ⎫
⎪
⎪ • • ⎪
⎪
⎪
⎪ ⎪
⎪
⎨ ⎬
n = 4, P7 = • , gl.dim ΛP7 = 2;
⎪
⎪ ⎪
⎪
⎪
⎪ | ⎪
⎪
⎩ ⎭
•
⎧ ⎫
⎨ • • ⎬
n = 4, P8 = | | , gl.dim ΛP8 = 3;
⎩ ⎭
• •
⎧ ⎫
⎨ • • • ⎬
n = 4, P9 = | , gl.dim ΛP9 = 2.
⎩ ⎭
•
It follows from proposition 6.10.2 that if the finite posets PΛ1 and PΛ2 , which
are associated with (0, 1)-orders Λ1 and Λ2 , are anti-isomorphic, then gl.dim Λ1 =
gl.dim Λ2 .
Proof. By the Michler theorem (see theorem 12.3.4, vol.I), gl.dim Λ = 1 if and
only if M(Λ) is a chain. In this case M(Λ) is a lower semilattice. If M(Λ) is not
a chain, let Pi and Pj be non-comparable elements of M(Λ). Then Pi + Pj = M
and the projective cover P (M ) of M is Pi ⊕ Pj . Let ϕ : P (M ) −→ M . Then
Ker ϕ Pi ∩ Pj is projective.
I1 ⊂ I2 . . . ⊂ In−1 ⊂ Ωn
with ⎛ ⎞
n−1 n−2 ... 1 0
⎜n − 1 n − 2 . . . 1 0⎟
⎜ ⎟
E(I1 ) = ⎜ . .. .. .. .. ⎟ ,
⎝ .. . . . .⎠
n−1 n−2 ... 1 0
⎛ ⎞
n−2 n−3 ... 0 0
⎜n − 2 n − 3 . . . 0 0⎟
⎜ ⎟
⎜ .. .. ⎟ ,
E(I2 ) = ⎜ ... ..
.
..
. . .⎟
⎜ ⎟
⎝n − 2 n − 3 . . . 0 0⎠
n−1 n−2 ... 1 0
··········································
⎛ ⎞
1 0 ... ... 0
⎜ .. .. ⎟
⎜ 1 0 . .⎟
⎜ ⎟
⎜ .. . .. .. .. ⎟ ,
E(In−1 ) = ⎜ . .. . . .⎟
⎜ ⎟
⎜ .. .. .. ⎟
⎝n − 2 . . . 0⎠
n−1 n−2 ... 1 0
⎛ ⎞
0 0 ... ... ... 0
⎜ ⎟
⎜ .. .. .. ⎟
⎜ 1 . . .⎟
⎜ ⎟
⎜ . . . .. .. ⎟
⎜ .. .. .. . .⎟
⎜ ⎟
E(Ωn ) = ⎜ .. ⎟
⎜ .. .. .. ⎟
⎜n − 3 ... . . . .⎟
⎜ ⎟
⎜ .. .. ⎟
⎜n − 2 n − 3 ... . . 0⎟
⎝ ⎠
n−1 n−2 n−3 ... 1 0
J1 ⊂ J2 ⊂ . . . ⊂ Jn−1 ⊂ Λn
322 ALGEBRAS, RINGS AND MODULES
with
⎛ ⎞ ⎛ ⎞
2 2 ... 1 0 2 ... ... 2 1 0 0
⎜2 2 ... 1 0⎟ ⎜2 . . . ... 2 1 0 0⎟
⎜ ⎟ ⎜ ⎟
⎜ .. ⎟ , E(J ) = ⎜ .. .. ⎟
E(J1 ) = ⎜ ... ..
.
..
.
..
. .⎟ 2 ⎜.
..
.
..
.
..
.
.. ..
. . .⎟
⎜ ⎟ ⎜ ⎟
⎝2 2 ... 1 0⎠ ⎝2 . . . ... 2 1 0 0⎠
2 2 ... 1 0 2 ... ... 2 2 1 0
⎛ ⎞ ⎛ ⎞
1 0 ... ... ... 0 0 0 ... ... ... 0
⎜ .. .. ⎟ ⎜ .. .. .. ⎟
⎜ . ⎟ ⎜ . . ⎟
⎜1 0 .⎟ ⎜1 .⎟
⎜ ⎟ ⎜ ⎟
⎜ .. .. .. .. ⎟ ⎜ .. .. .. .. ⎟
⎜2 . . . .⎟ ⎜ . . . .⎟
. . . , E (Jn−1 ) = ⎜ ⎟ E (Λn ) = ⎜2 ⎟
⎜. .. .. .. .. .. ⎟ ⎜. .. .. .. .. .. ⎟
⎜ .. . . . . .⎟ ⎜ .. . . . . .⎟
⎜ ⎟ ⎜ ⎟
⎜. ⎟ ⎜. ⎟
⎜. .. .. .. ⎟ ⎜. .. .. .. ⎟
⎝. . . . 0⎠ ⎝. . . . 0⎠
2 ... ... 2 1 0 2 ... ... 2 1 0
Now we shall compute the quiver Q(Ωn ) and its transition matrix for the re-
(2) (1)
duced exponent matrix Ωn . We use the formula [Q(Ω)n ] = Ωn −Ωn . Obviously,
⎛ ⎞
1 1 0
... ... ... ... 0
⎜ .. .. .. .. ⎟
⎜ 2 . . . .⎟
⎜ ⎟
⎜ .. .. .. .. .. ⎟
⎜ 2 . . . . .⎟
⎜ ⎟
⎜ .. .. .. .. .. .. ⎟
⎜ 3 . . . . . .⎟
Ω(2)
n = ⎜
⎜ . .. ⎟
⎟
⎜ .. .. .. .. .. .. ..
⎜ . . . . . . .⎟ ⎟
⎜ . .. .. .. .. .. .. ⎟
⎜ .. . . . . . . 0⎟
⎜ ⎟
⎜ .. .. .. .. ⎟
⎝n − 2 n−3 ... . . . . 1⎠
n−1 n − 2 n − 3 ... 3 2 2 1
and [Q(Ωn )] = Jn− (0) + Jn+ (0) = Yn , where Jn+ (0) = e12 + e23 + . . . + en−1n
π
and Jn− (0) = e21 + e32 + . . . + enn−1 . We have that inx Ωn = 2 cos and
n+1
π
f = (a1 , a2 , . . . , an ) is a positive eigenvector of Y with eigenvalue λ = 2 cos ,
n+1
iπ
Z = diag (a1 , . . . , an ), where ai = sin for i = 1, 2, . . . , n.
n+1
TILED ORDERS OVER DISCRETE VALUATION RINGS 323
Sn = λ−1 Z −1 Yn Z =
1 1 1 1
= π · diag( , . . . , ) · Yn · diag(a1 , . . . , an ) = π · C,
2 cos n+1 a1 an 2 cos n+1
iπ
where Yn = (yij ), C = (cij ) and ai = sin for i = 1, 2, . . . , n,
n+1
1, if i = j − 1 or i = j + 1,
yij =
0, otherwise,
⎧ ai+1
⎪
⎪ , if i = j − 1,
⎪
⎪ ai
⎪
⎪
⎨
cij = ai
⎪ , if i = j + 1,
⎪
⎪ ai+1
⎪
⎪
⎪
⎩
0, otherwise,
for i, j = 1, 2, . . . , n.
Then gl. dimA = 6 and this is also a counterexample to the Tarsy conjecture (see
[Tarsy, 1970]) saying that the maximal possible finite global dimension of a tiled
order in (K)n is n − 1.
The following example of W.Rump shows that the global dimension of an order
is determined not only by its exponent matrix even for the case of (0,1)-orders.
Consider the tiled O-order Λ in the matrix algebra M14 (K), where O is a
discrete valuation domain with quotient field K with exponent matrix:
⎛ ⎞
0 1 1 1 1 1 1 1 1 1 1 1 1 1
⎜0 0 1 1 1 1 1 1 1 1 1 1 1 1 ⎟
⎜ ⎟
⎜0 1 0 1 1 1 1 1 1 1 1 1 1 1 ⎟
⎜ ⎟
⎜0 1 1 0 1 1 1 1 1 1 1 1 1 1 ⎟
⎜ ⎟
⎜0 0 1 0 0 1 1 1 1 1 1 1 1 1 ⎟
⎜ ⎟
⎜0 0 0 1 1 0 1 1 1 1 1 1 1 1 ⎟
⎜ ⎟
⎜0 0 0 1 1 1 0 1 1 1 1 1 1 1 ⎟
E =⎜ ⎜0 1 0 0 1 1 1 0 1 1 1 1 1 1 ⎟
⎟
⎜ ⎟
⎜0 1 0 0 1 1 1 1 0 1 1 1 1 1 ⎟
⎜ ⎟
⎜0 0 1 0 1 1 1 1 1 0 1 1 1 1 ⎟
⎜ ⎟
⎜0 0 0 0 0 0 1 1 0 1 0 1 1 1 ⎟
⎜ ⎟
⎜0 0 0 0 0 1 0 0 1 1 1 0 1 1 ⎟
⎜ ⎟
⎝0 0 0 0 1 0 1 0 1 0 1 1 0 1 ⎠
0 0 0 0 1 1 0 1 0 0 1 1 1 0
If the characteristic of the residue class field k of O is 2, then gl.dim (Λ) = 4,
otherwise gl.dim (Λ) = 3.
Details can be found in example 1 of the paper [Rump, 1996].
The authors thank W.Rump for helpful discussions on the presentation of this
chapter.
Theorem 7.1.1. The following conditions are equivalent for a tiled order A:
(i) inj. dimA AA = 1;
(ii) inj. dimA A A = 1;
(iii) A∗A is projective left A-module;
(iv) A A∗ is projective right A-module.
0 → AA → Q0 → Q0 /AA → 0.
By proposition 6.5.5, vol. I, the module Q0 /AA is injective. Obviously, every inde-
composable direct summand of Q0 /AA has the form eii Q0 /eii A. Since eii Q0 /eii A
is indecomposable injective soc (eii Q0 /eii A) is simple. Therefore every eii A is a
relatively injective irreducible A-lattice by proposition 6.2.14, and AA is a rela-
tively injective right A-module. By definition, AA A P ∗ . By duality properties,
A P =A P1 ⊕ . . . ⊕A Ps ⊕ P , where A P1 , . . .A Ps are all pairwise non-isomorphic left
principal A-modules and every indecomposable direct summand of P is isomorphic
to some A Pi . Therefore, A A∗ is a projective right A-module.
From corollary 6.2.11 we obtain (iii) ⇔ (iv). In conclusion, we obtain that
(iv) ⇒ (i), by corollary 6.2.15 and the fact, that A A∗ and AA contain the same
indecomposable summands if A A∗ is projective. The equivalence (ii) ⇔ (iii) for
left modules is proved just like (i) ⇔ (iv) for right modules. The theorem is
proved.
1 Recallthat a commutative ring is called Gorenstein if its injective dimension is finite. These
rings were first considered by H.Bass (see [Bass, 1963])
327
328 ALGEBRAS, RINGS AND MODULES
Proposition 7.1.2. Let A = {O, E(A)} be a reduced tiled order with exponent
matrix E(A) = (αij ) ∈ Mn (Z). A is Gorenstein if and only if the matrix E(A)
is Gorenstein, i.e., there exists a permutation σ of the set {1, . . . , n} such that
αik + αkσ(i) = αiσ(i) for i, k = 1, . . . , n.
and ⎛ ⎞
⎜ 0 −α21 ... ... −αn1 ⎟
⎜ ⎟
⎜ .. .. ⎟
⎜ −α12 0 . . ⎟
⎜ ⎟
⎜ . .. .. .. .. ⎟
E (A A ) = ⎜
∗
⎜ .
. . . . . ⎟.
⎟
⎜ ⎟
⎜ .. .. .. ⎟
⎜ . . . −αnn−1 ⎟
⎜ ⎟
⎝ ⎠
−α1n −α2n . . . −αn−1n 0
Example 7.1.1.
Let ⎧ ⎛ ⎞⎫
⎨
0 0 0 ⎬
A = O, ⎝1 0 1⎠
⎩ ⎭
1 1 0
and P1 = (0, 0, 0); P2 = (1, 0, 1); P3 = (1, 1, 0) be projective A-modules. Obvi-
ously, P1∗ = (0, 0, 0)T ; P2∗ = (−1, 0, −1)T (0, 1, 0)T and P3∗ = (−1, −1, 0)T
(0, 0, 1)T . Therefore, the modules P2 and P3 are relatively injective, but A is not
Gorenstein. It is well-known that gl.dim A = 2.
GORENSTEIN MATRICES 329
In this section we shall assume that the first row of a Gorenstein matrix E is
zero.
Let ⎛ ⎞
0 0 ... ... 0
⎜ .. ⎟
⎜α 0 . . . .⎟
⎜ ⎟
⎜ .. . . . . . . .. ⎟
.
Tn,α = ⎜ . . . . ⎟,
⎜ ⎟
⎜. .. ⎟
⎝ .. . 0 0⎠
α ... ... α 0
where α is a natural integer, Tn,α ∈ Mn (Z). Obviously, Tn,α is a cyclic Gorenstein
matrix with σ = σ(Tn,α ) = (n n−1 . . . 2 1). Write Jn+ (0) = e12 +e23 +. . .+en−1n .
It is easy to see that [Q(Tn,1 )] = Jn+ (0) + en1 and [Q(Tn,α )] = En + Jn+ (0) + en1 ,
where En is the identity n × n-matrix and α ≥ 2.
We shall write Tn,1 = Hn . Therefore, Q(Hn ) is a simple cycle Cn . For the
quiver Q(Tn , α) (α ≥ 2) we use the notation LCn , i.e., LCn is a simple cycle with
a loop in each vertex.
Obviously, inx Hn = 1 and inx Tn,α = 2. In this case [Q(Hn )] = Pσn−1 and
[Q(Tn,α )] = Pid + Pσn−1 .
Below the Gorenstein tiled orders and hence Gorenstein matrices are described
in detail for n ≤ 6.
To do this efficiently it is important to recall that up to equivalence it can be
assumed that the first row of an exponent matrix E = (αij ) consists of zeroes (and
then αij ≥ 0 for all i, j). Up to equivalence that still leaves the freedom to permute
columns and rows simultaneously by a permutation τ which leaves 1 fixed. Such
an operation turns a Gorenstein matrix E with first row zero and permutation σ
into a new Gorenstein matrix E with first row zero and permutation σ = τ −1 στ .
A second important point to note is that the permutation of a reduced Goren-
stein matrix cannot have a fixed point. Indeed, suppose that σ(i) = i. Then for
all k, αik + αkσ(i) = αiσ(i) and so αik + αki = αii = 0 contradicting that E is
reduced.
We shall use the decomposition of a permutation σ ∈ Sn into a product of
1 2 3 4 5
independent cycles. For example, the permutation σ = is the
3 5 4 1 2
2 see [Roggenkamp, 2001]
330 ALGEBRAS, RINGS AND MODULES
For n = 4 there are two possibilities for σ(E): a simple cycle and a product of
two transpositions. By proposition 4.1.12, we may assume that
1 2 3 4 1 2 3 4
σ(E) = and σ(E) = .
4 1 2 3 2 1 4 3
Case (a): σ = σ(E) = (4 3 2 1).
We shall use notations like: (α1 , . . . , αk ) = min (α1 , . . . , αk ) and δ = (2, α, β).
It is easy to see that
⎛ ⎞
0 0 0 0
⎜α 0 α − β 0⎟
E = Eα,β = ⎜⎝β β
⎟ (α ≥ β > 0).
0 0⎠
α β α 0
We have
⎛ ⎞
1 0 0 0
⎜α 1 α−β 0⎟
= ⎜ ⎟
(1)
Eα,β ⎝β β 1 0⎠
α β α 1
GORENSTEIN MATRICES 331
and
⎛ ⎞
δ 1 (1, α − β) 0
⎜ α δ (α, α − β + 1) (1, α − β)⎟
= ⎜ ⎟.
(2)
Eα,β ⎝(α, β + 1) ⎠
β δ 1
α+1 (α, β + 1) α δ
Therefore, E1,1 = H4 and Q(E1,1 ) = C4 . If α ≥ 2, then Eα,α = T4,α and
Q(Eα,α ) = LC4 .
If β = 1 we have α ≥ 2 and
⎛ ⎞
0 1 1 0
⎜0 0 1 1⎟
[Q(Eα,1 )] = ⎜
⎝1
⎟.
0 0 1⎠
1 1 0 0
1
•
4 • • 2
•
3
In this case, [Q(Eα,1 )] = Pσ3 + Pσ2 and inx Eα,1 = 2.
If α > β > 1, then
⎛ ⎞
1 1 1 0
⎜0 1 1 1 ⎟
[Q(Eα,β )] = ⎜ ⎟
⎝1 0 1 1⎠ = Pσ2 + Pσ3 + Pσ4
1 1 0 1
and inx Eα,β = 3.
If Δ = 1 we have
⎛ ⎞
0 0 1 1
⎜0 0 1 1⎟
B = [Q(E1,δ )] = [Q(Eγ,1 )] = ⎜
⎝1
⎟
1 0 0⎠
1 1 0 0
[Q(Eγ,δ )] = E + B = Pτ + Pτ 2 + Pw
or σ = (1 2)(3 4 5).
Case (a):
n = 5, σ = (5 4 3 2 1). It is easy to see that
⎛ ⎞
0 0 0 0 0
⎜μ 0 μ − ν μ − ν 0⎟
⎜ ⎟
Eμ,ν = ⎜⎜ ν ν 0 μ − ν 0⎟
⎟ (μ ≥ ν > 0)
⎝ν 2ν − μ ν 0 0⎠
μ ν ν μ 0
⎛ ⎞
1 0 0 0 0
⎜ μ 1 μ−ν μ−ν 0 ⎟
⎜ ⎟
(1)
Eμ,ν =⎜
⎜ ν ν 1 μ−ν 0 ⎟.
⎟
⎝ ν 2ν − μ ν 1 0 ⎠
μ ν ν μ 1
GORENSTEIN MATRICES 333
(2)
Let δ = min (2, μ, ν). The matrix Eμ,ν is:
⎛ ⎞
δ (1, 2ν − μ) (1, μ − ν) (1, μ − ν) 0
⎜ μ δ (ν, μ − ν + 1) (μ, μ − ν + 1) (1, μ − ν) ⎟
⎜ ⎟
⎜ (μ, ν + 1) ν δ (ν, μ − ν + 1) (1, μ − ν) ⎟
⎜ ⎟
⎝ (μ, ν + 1) (ν, 2ν − μ + 1) ν δ (1, 2ν − μ) ⎠
(μ + 1, 2ν) (μ, ν + 1) (μ, ν + 1) μ δ
and
(2) (1)
[Q(Eμ,ν )] = Eμ,ν − Eμ,ν =
⎛ ⎞
δ−1 (1, 2ν − μ) (1, μ − ν) (1, μ − ν) 0
⎜ 0 δ−1 (1, 2ν − μ) (1, μ − ν) (1, μ − ν) ⎟
⎜ ⎟
= ⎜
⎜ (1, μ − ν) 0 δ−1 (1, 2ν − μ) (1, μ − ν) ⎟⎟.
⎝ (1, μ − ν) (1, μ − ν) 0 δ−1 (1, 2ν − μ)⎠
(1, 2ν − μ) (1, μ − ν) (1, μ − ν) 0 δ−1
If μ = ν we obtain that E1,1 = H5 and Eμ,μ = T5,μ for μ ≥ 2.
If μ = 2ν and ν = 1, we have
⎛ ⎞
0 0 1 1 0
⎜0 0 0 1 1⎟
⎜ ⎟
C = [Q(E2,1 )] = ⎜
⎜1 0 0 0 1⎟⎟.
⎝1 1 0 0 0⎠
0 1 1 0 0
and
1 3
Q(E5 ) =
5 4
334 ALGEBRAS, RINGS AND MODULES
If μ = 2ν > 2, then
⎛ ⎞
1 0 1 1 0
⎜0 1 0 1 1⎟
⎜ ⎟
[Q(E2ν,ν )] = ⎜
⎜1 0 1 0 1⎟⎟ = E + C = Pid + Pσ2 + Pσ3
⎝1 1 0 1 0⎠
0 1 1 0 1
where α ≥ 0, β ≥ 0, γ ≥ 0, β + γ ≥ 1, α + γ ≥ 1. Then
⎛ ⎞
δ (1, γ) (1, α, γ) (1, β) (1, α, β) 0
⎜ 0 δ (1, γ) (1, α, γ) (1, β) (1, α, β)⎟
⎜ ⎟
⎜(1, α, β) 0 δ (1, γ) (1, α, γ) (1, β) ⎟
[Q(Eα,β,γ )] = ⎜
⎜ (1, β)
⎟,
⎜ (1, α, β) 0 δ (1, γ) (1, α, γ)⎟
⎟
⎝(1, α, γ) (1, β) (1, α, β) 0 δ (1, γ) ⎠
(1, γ) (1, α, γ) (1, β) (1, α, β) 0 δ
where δ = (2, β + γ, γ + α) − 1.
Let α = β = 0, then E0,0,γ = T6,γ . In this case, E0,0,1 = H6 and inx E0,0,γ = 2
for γ ≥ 2. In the case α = δ = 0 we have γ = 1 and
⎛ ⎞
0 1 0 (1, β) 0 0
⎜ 0 0 1 0 (1, β) 0 ⎟
⎜ ⎟
⎜ 0 0 0 1 0 (1, β)⎟
[Q(E0,β,1 )] = ⎜
⎜(1, β)
⎟.
⎜ 0 0 0 1 0 ⎟ ⎟
⎝ 0 (1, β) 0 0 0 1 ⎠
1 0 (1, β) 0 0 0
and
6 2
Q(Γα ) =
5 3
for any n and inx Eα,β,γ = 5. Obviously, U6 − Pσ = Pid + Pσ2 + Pσ3 + Pσ4 + Pσ5 .
GORENSTEIN MATRICES 337
n
Un = Pσi ,
i=1
where σ = (n n − 1 . . . 2 1).
Note that ⎛ ⎞
1 0 0 1 1 1
⎜1 1 0 0 1 1⎟
⎜ ⎟
⎜0 1 1 0 1 1⎟
[Q(Γ6 )] = ⎜
⎜0
⎟
⎜ 0 1 1 1 1⎟
⎟
⎝1 1 1 1 1 0⎠
1 1 1 1 0 1
is not a multiple of a doubly stochastic matrix. We have that
1 2
Q(Γ6 ) = 5 6
4 3
Example 7.1.2.
338 ALGEBRAS, RINGS AND MODULES
3 3 2 0
⎡ ⎤ ⎡ ⎤
1 0 0 0 2 1 1 0
⎢ 2 1 1 1 ⎥ ⎢ 3 2 2 1 ⎥
Δ(1) =⎢
⎣ 2
⎥, Δ(2) =⎢ ⎥,
2 1 0 ⎦ ⎣ 3 2 2 1 ⎦
3 3 2 1 4 3 3 2
⎡ ⎤
1 1 1 0
⎢ 1 1 1 0 ⎥
B = [Q(Δ)] = ⎢
⎣ 1 0
⎥
1 1 ⎦
1 0 1 1
Corollary 7.2.4. If for any j = i, j = k, αij + αjk > αik for some i, k (i = k),
then
ασm (i)σm (j) + ασm (j)σm (k) > ασm (i)σm (k) .
ασn−m (σm (i))σn−m (σm (j)) + ασn−m (σm (j))σn−m (σm (k)) = ασn−m (σm (i))σn−m (σm (k)) .
So ασn (i)σn (j) + ασn (j)σn (k) = ασn (i)σn (k) , i.e., αij + αjk = αik . The corollary is
proved.
Let [Q] = (tij ) be the adjacency matrix of the quiver Q(E) of a Gorenstein
matrix E.
Proof. Let tij = 0 for i = j. By the definition of [Q(E)], there exists an integer
k such that βik + βkj = βij . If k = i or k = j then βik + βkj > βij . Consequently,
i, j, k are distinct and βik = αik , βkj = αkj , βij = αij . By corollary 7.2.3, we
have
βσm (i)σm (k) + βσm (k)σm (j) = βσm (i)σm (j) .
Then tσm (i)σm (j) = min (βσm (i)k + βkσm (j) ) − βσm (i)σm (j) = 0 = tij .
If tij = 1 for i = j, then βik + βkj > βij for all k. We shall prove that
βσm (i)σm (k) + βσm (k)σm (j) > βσm (i)σm (j)
for k = 1, . . . , n.
It is obvious for i = k or j = k. Therefore we can consider that i, j, k are
distinct. And so the inequality βik + βkj > βij is the same thing as the inequality
αik + αkj > αij . By corollary 7.2.4, we have
ασm (i)σm (k) + ασm (k)σm (j) > ασm (i)σm (j)
and βσm (i)σm (k) + βσm (k)σm (j) > βσm (i)σm (j) . As k ranges over 1, . . . , n so does
σ m (k) and hence tσm (i)σm (j) > 0 and hence tij = 1 as [Q(E)] is a (0, 1)-matrix.
So, tij = tσm (i)σm (j) for i = j.
Let tii = 0. Then αik + αki = 1 for some k = i. By corollary 7.2.3,
Hence, γσm (i)σm (i) = 1, and tσm (i)σm (i) = 0. If tii = 1 then γii = 2. Since
βii + βii = 1 + 1 = 2 for i = 1, . . . , n it follows that βij + βji ≥ 2 for i = j. But
βij = αij for i = j and, by corollary 7.2.3, we have
So γσm (i)σm (i) ≥ 2 and tσm (i)σm (i) = 1. The lemma is proved.
n
n
belong to different rows and columns. Let Ci = tij and Dj = tij . Then
j=1 i=1
n
n
Ci = tij = tσm (i)σm (j) = Cσm (i)
j=1 j=1
GORENSTEIN MATRICES 341
and
n
n
Dj = tij = tσm (i)σm (j) = Dσm (j)
i=1 i=1
n
n
n
Ci = tij = Dj .
i=1 i,j=1 j=1
The proof of this theorem is based on the Frobenius-König lemma for which
we shall need to introduce some new notions.
Proof. We shall prove the lemma by induction on the degree of the matrix.
We assume that the statement is true for all matrices of degree < n.
The case n = 1 is trivial: t11 = 0 and 1 + 1 = 2.
342 ALGEBRAS, RINGS AND MODULES
The matrix T1 satisfies the conditions of the lemma, since otherwise there is a
normal set of positive elements of T . So, by the induction hypothesis, there is a
zero (p1 × q1 )-submatrix N of T1 such that p1 + q1 = n; p1 , q1 ≤ n − 1. We may
assume that N = N1,...,p1 ;1,...q1 and T has the following form
p1 0 T2
q1 T3
q1 p1
p2 0 *
* *
q2
p2 0 0 *
p1 − p2 0 * *
* * *
q1 q2
And thus there is a zero (p2 ×(q1 +q2 ))-submatrix of T . Therefore, p2 +q1 +q2 =
p1 + q1 + 1 = n + 1. The lemma is proved.
GORENSTEIN MATRICES 343
n
n
bij = bij = ω.
i=1 j=1
Then there exists a normal set b1i1 , . . . , bnin of strictly positive elements of B.
Proof. Suppose that the statement is not true. Then, by the previous lemma,
we can assume that B has the following form:
p 0
n−p
q n−q
and let P1 be a permutation matrix which has nonzero elements on the places
corresponding to the set (7.2.4). Consider the matrix
B1 = T − τ1 P1 .
The elements of B1 are non-negative, and the sum of elements in each row and
each column of B1 is equal to 1 − τ1 = ω1 ≥ 0. The number of zero elements in
B1 is at least one more than the number of zero elements in T . If 1 − τ1 = 0 we
are through. Continuing this process after k steps we obtain:
Bk = T − τ1 P1 − τ2 P2 − . . . − τk Pk .
As the number of zeros in Bk is at least one more than the number of of zeros
in Bk−1 this procedure terminates, proving that a doubly stochastic matrix is a
linear sum of permutation matrices whose coefficients are positive and sum to 1.
344 ALGEBRAS, RINGS AND MODULES
Conversely, the right side of the equality T = τσ Pσ , where τσ ≥ 0 and
σ∈Sn
τσ = 1, is obviously a doubly stochastic matrix. The theorem is proved.
σ∈Sn
n
n
n
[Q] = tij eij = tiσk (i) eiσk (i) = t1σk (1) Pσk .
i,j=1 i,k=1 k=1
Taking into account that Pτ m = (Pτ )m for any permutation matrix Pτ , we obtain
n
[Q] = t1σk (1) Pσk , where t1σk (1) equals either 0 or 1.
k=1
Example 7.2.1.
Let E2m ∈ M2m (Z) be the following exponent matrix:
⎛ ⎞
A C
... ... C
⎜ .. ⎟
⎜ .. ⎟
⎜B A . .⎟
⎜ ⎟
⎜. .. .. .. .. ⎟
E2m = ⎜ .. . . . .⎟ ,
⎜ ⎟
⎜. .. . ⎟
⎜ .. . . . C⎟
⎝ ⎠
B ... ... B A
0 0 1 1 0 0
where A = ,B = ,C = .
2 0 2 1 1 0
GORENSTEIN MATRICES 345
It is easy to see that E2m is a cyclic Gorenstein matrix and [Q(E2m )] = Pσn−2 +
Pσn−1 , where n = 2m with σ = (n n − 1 . . . 2 1).
Q(E2 ) = 1 2
Q(E4 ) = 4 2
6 2
Q(E6 ) =
5 3
8 1
7 2
Q(E8 ) =
6 3
5 4
346 ALGEBRAS, RINGS AND MODULES
If m = 1 then
1 1
[Q(G2 )] = = E + Pτ ,
1 1
where τ is the transposition (1 2).
In the general case, [Q(G2m )] = Pτ m−1 + Pτ 2m−1 , where
1 2 ... 2m
τ =
2m 1 . . . 2m − 1
PE = {1, . . . , n}
with the relation defined by the formula i j ⇔ αij = 0. It is easy to see that
(, PE ) is a poset. Conversely, with any finite poset
P = {1, . . . , n}
there is associated, as before (see just above theorem 6.3.1), the reduced exponent
(0, 1)-matrix EP = (αij ) defined by: αij = 0 if and only if i j in P, otherwise
αij = 1.
P1 P2 P3 ... Ps−1 Ps
P2s = (7.3.1)
Ps+1 Ps+2 Ps+3 ... P2s−1 P2s
conditions and lemma 7.3.2 say that α1i + αiσ(1) + α1σ(1) = α14 = 1. Then, taking
i = 2, 3, we obtain α24 = α34 = 0. Similarly α2i + αiσ(i) = α25 . By lemma
7.3.2, α25 = 1. For i = 3 we have α23 + α35 = 1 and α35 = 0. Analogously,
α21 + α15 = 1 and α15 = 0 (recall that we are dealing with a (0, 1)-matrix).
Similarly α16 = α26 = 0. Next the αij for 4 ≤ i, j ≤ 6 are calculated. For instance
α34 + α46 = α36 = 1, so that α46 = 1. Further α24 + α45 = α25 and so α45 = 1.
As a result E looks like
⎛ ⎞
0 1 1 1 0 0
⎜1 0 1 0 1 0 ∗ ⎟
⎜ ⎟
⎜1 1 0 0 0 1 ⎟
⎜ ⎟
⎜ ⎟
⎜ ⎟
E = ⎜
⎜ 0 1 1 ⎟.
⎟
⎜ ∗ 1 0 1 ∗ ⎟
⎜ ⎟
⎜ 1 1 0 ⎟
⎜ ⎟
⎝ ⎠
∗ ∗ ∗
As αij + αji > 0 (i = j) it follows that α51 = α61 = α42 = α62 = α43 = α53 = 1.
Further α24 +α41 ≥ α21 = 1 and as α24 = 0 this gives α41 = 1; α15 +α52 ≥ α12 = 1
and α15 = 0 gives α52 = 1; α16 + α63 ≥ α13 = 1 and α16 = 0 gives α63 = 1. Thus
the matrix E is of the form
⎛ ⎞
0 1 1 1 0 0
⎜1 0 1 0 1 0 ∗ ⎟
⎜ ⎟
⎜1 1 0 0 0 1 ⎟
⎜ ⎟
⎜ ⎟
⎜ ⎟
⎜
E = ⎜1 1 1 0 1 1 ⎟.
⎟
⎜1 1 1 1 0 1 ∗ ⎟
⎜ ⎟
⎜1 1 1 1 1 0 ⎟
⎜ ⎟
⎝ ⎠
∗ ∗ ∗
Now observe that σ(i) > 6 for i = 4, 5, 6. Indeed, if σ(4) = 1, then α4σ(4) = 1 =
α42 + α21 = 2, a contradiction. Analogously, the cases σ(4) = 2 and σ(4) = 3
are impossible. Recall that σ(1) = 4, σ(2) = 5 and σ(3) = 6. Therefore, σ(i) ∈
{4, 5, 6} for i = 4, 5, 6. So it can be assumed that σ(4) = 7, σ(5) = 8, σ(6) = 9.
Again, of course, the Gorenstein condition is used in the form
α4i + αi7 = α47 , α5i + αi8 = α58 , α6i + αi9 = α69 .
As α41 = α51 = α61 = 1, this gives α47 = α58 = α69 = 1. Put i = 4, 6 in
α5i + αi8 = 1 to find α48 = α68 = 0; put i = 5, 6 in α4i + αi7 = 1 to find
α57 = α67 = 0; put i = 4, 5 in α6i + αi9 = 1 to find α49 = α59 = 0.
Further put i = 8, 9 in α4i + αi7 = 1 to find α87 = 1, α97 = 1; put i = 7, 9 in
α5i + αi8 = α58 = 1 to find α78 = α98 = 1, put i = 7, 8 in α6i + αi9 = 1 to find
α79 = α89 = 1.
350 ALGEBRAS, RINGS AND MODULES
Next put i = 1, 2, 3 in α4i + αi7 = 1, α5i + αi8 = 1, α6i + αi9 = 1 to find αij = 0
for i ∈ {1, 2, 3}, j ∈ {7, 8, 9}.
Thus the matrix E is of the form
⎛ ⎞
0 1 1 1 0 0 0 0 0
⎜ 1 0 1 0 1 0 0 0 0 ∗⎟
⎜ ⎟
⎜ 1 1 0 0 0 1 0 0 0 ⎟
⎜ ⎟
⎜ ⎟
⎜ ⎟
⎜ 1 1 1 0 1 1 1 0 0 ⎟
⎜ ⎟
⎜ 1 1 1 1 0 1 0 1 0 ∗⎟
⎜ ⎟
E = ⎜
⎜ 1 1 1 1 1 0 0 0 1 ⎟.
⎟
⎜ ⎟
⎜ ⎟
⎜ 0 1 1 ⎟
⎜ ⎟
⎜ ∗ ∗ 1 0 1 ∗⎟
⎜ ⎟
⎜ 1 1 0 ⎟
⎜ ⎟
⎝ ⎠
∗ ∗ ∗ ∗
Because αij + αji > 0, (i = j), it now follows that αij = 1 for i ∈ {7, 8, 9},
j ∈ {1, 2, 3} and α75 = α76 = α84 = α86 = α94 = α95 = 1. Next
and α57 = α48 = α49 = 0, so that α74 = α85 = α96 = 1. So the exponent matrix
E looks like
⎛ ⎞
0 1 1 1 0 0 0 0 0
⎜ 1 0 1 0 1 0 0 0 0 ∗⎟
⎜ ⎟
⎜ 1 1 0 0 0 1 0 0 0 ⎟
⎜ ⎟
⎜ ⎟
⎜ ⎟
⎜ 1 1 1 0 1 1 1 0 0 ⎟
⎜ ⎟
⎜ 1 1 1 1 0 1 0 1 0 ∗⎟
⎜ ⎟
E = ⎜
⎜ 1 1 1 1 1 0 0 0 1 ⎟.
⎟
⎜ ⎟
⎜ ⎟
⎜ 1 1 1 1 1 1 0 1 1 ⎟
⎜ ⎟
⎜ 1 1 1 1 1 1 1 0 1 ∗⎟
⎜ ⎟
⎜ 1 1 1 1 1 1 1 1 0 ⎟
⎜ ⎟
⎝ ⎠
∗ ∗ ∗ ∗
Again, σ(i) > 6 for i = 7, 8, 9. Continuing this process we finally obtain after m
GORENSTEIN MATRICES 351
steps that the Gorenstein matrix E has the following block form:
⎛ ⎞
A E O ... O ∗
⎜ .. .. .. . .⎟
⎜U . . . .. .. ⎟
⎜ ⎟
⎜ .. .. .. .. .⎟
⎜ . . . O .. ⎟
E = ⎜. ⎟,
⎜. .. .. .. ⎟
⎜ .. . . E .⎟
⎜ ⎟
⎝U . . . . . . U A ∗⎠
∗ ... ... ∗ ∗ ∗
where n = 3m + r, 0 ≤ r ≤ 2, the bottom ∗’s are matrices with r rows. The ∗’s
on the right are have r columns, and where
⎛ ⎞ ⎛ ⎞
0 1 1 1 1 1
A = ⎝1 0 1⎠ , U = ⎝1 1 1⎠ ,
1 1 0 1 1 1
⎛ ⎞ ⎛ ⎞
0 0 0 1 0 0
O = ⎝0 0 0⎠ , E = ⎝0 1 0⎠ .
1 1 0 0 0 1
But then the Gorenstein condition forces (as above) that σ(i) > 3m for i ∈
{1, 2, . . . , 3m} for which there is no room. Thus the hypothesis w(E) ≥ 3 leads to
a contradiction proving that w(E) ≤ 2.
Consider the case w(PE ) = 2, that means PE has two non-comparable ele-
ments. Let they be P1 and P2 . Then α12 = α21 = 1, and the exponent matrix
has the following form:
⎛ ⎞
0 1
∗
E = ⎝ 1 0 ⎠.
∗ ∗
Suppose, σ(1), σ(2) > 2. One may then assume that σ(1) = 3 and σ(2) = 4.
Then, in view of the Gorenstein condition and lemma 7.3.2, we obtain α1j + αj3 =
α13 = 1 and α2j + αj4 = α24 = 1.
As α12 = 1 it follows that α13 = 1 and α23 = 0; and as α21 = 1 it follows that
α24 = 1 and α14 = 0.
Further α23 + α34 = α24 = 1, so α34 = 1; α14 + α43 = α13 = 1, so α43 = 1.
Thus the exponent matrix E looks like
⎛ ⎞
0 1 1 0
⎜ 1 0 ∗ ⎟
⎜ 0 1 ⎟
⎜ ⎟
⎜ ⎟
E = ⎜⎜ 0 1 ⎟.
∗ ∗ ⎟
⎜ 1 0 ⎟
⎜ ⎟
⎝ ⎠
∗ ∗ ∗
352 ALGEBRAS, RINGS AND MODULES
Next as αij + αji > 0 for i = j, it follows that α32 = α41 = 1. Also α14 + α42 ≥
α12 = 1 and as α14 = 0, α42 = 1; α23 + α31 ≥ α21 = 1, α23 = 0, so that α31 = 1.
So E is of the form
⎛ ⎞
0 1 1 0
⎜ 1 0 ∗⎟
⎜ 0 1 ⎟
⎜ ⎟
⎜ ⎟
⎜
E = ⎜ 1 1 0 1 ⎟.
⎜ 1 1 ∗⎟⎟
⎜ 1 0 ⎟
⎝ ⎠
∗ ∗ ∗
Now observe that it must be the case that σ(3), σ(4) > 4. Indeed, suppose
σ(3) ∈ {1, 2, 4} (a fixed point being impossible). The Gorenstein condition says
⎛ ⎞
0 1 1 0 0 0
⎜ 1 ∗⎟
⎜ 0 0 1 0 0 ⎟
⎜ ⎟
⎜ ⎟
⎜ 1 1 0 1 1 0 ⎟
⎜ ∗⎟
E = ⎜
⎜ 1 1 1 0 0 1 ⎟.
⎟
⎜ ⎟
⎜ ⎟
⎜ 1 1 1 1 0 1 ⎟
⎜ ∗⎟
⎝ 1 1 1 1 1 0 ⎠
∗ ∗ ∗ ∗
Now observe that the Gorenstein condition forces σ(i) > 6 for i ∈ {1, 2, 3, 4, 5, 6}
and continue in the same manner. The result, after m steps, is that E must be of
the form
⎛A E O ... ... O ∗⎞
⎜U .. .. .. .. . .. ⎟
⎜ . . . . .. .⎟
⎜. .. .. .. .. . .. ⎟
⎜ .. . . . . .. .⎟
⎜ ⎟
E = ⎜
⎜ ... .. .. .. .. .. ⎟ ,
⎜ . . . . O .⎟ ⎟
⎜. .. .. .. .. .. ⎟
⎜ .. . . . . E .⎟
⎝ ⎠
U ... ... ... U A ∗
∗ ... ... ... ∗ ∗ ∗
where n = 2m + r, 0 ≤ r ≤ 1, where the bottom ∗’s are matrices with r rows and
the ∗’s on the right are blocks with r columns and where
0 1 1 1 0 0 1 0
A = ,U = ,O = ,E = .
1 0 1 1 0 0 0 1
The Gorenstein condition forces σ(i) > 2m for i ∈ {1, 2, . . . , 2m} for which
there is no room, giving a contradiction with the hypothesis σ(1), σ(2) > 2.
Hence, at least one of the numbers σ(1) or σ(2) is less than 3. Suppose σ(1) = 2,
but σ(2) = 1. Let σ(2) = 3 and α2σ(2) = 1 = α23 = α2i + αi3 . We assume that
α12 = α21 = 1. Then αi3 = 1 − α2i . Therefore α13 = 0 and α23 = 1. Consequently
⎛ ⎞
0 1 0
∗⎠
E = ⎝ 1 0 1 .
∗ ∗ ∗
Since E is reduced, αij +αji > 0 for i = j, and α13 +α32 ≥ α12 . So α31 = α32 = 1.
Hence
⎛ ⎞
0 1 0
⎜ 1 0 1 ∗⎟
E = ⎜⎝ 1 1 0
⎟.
⎠
∗ ∗
354 ALGEBRAS, RINGS AND MODULES
Now consider σ(3). Since σ(1) = 2 and σ(2) = 3, σ(3) = 2 cannot hold. If
σ(3) = 1 then α32 + α21 = 2. So σ(3) ∈ {1, 2} and σ(3) > 3. It can be assumed
that σ(3) = 4. The Gorenstein condition gives α3i + αi4 = α34 . As α31 = 1,
α34 = 1 and α14 = 0, and then also α24 = 0 because α32 = 1. So E looks like
⎛ ⎞
0 1 0 0
⎜ 1 0 1 0 ∗⎟
E = ⎜
⎝ 1 1
⎟.
⎠
0 1
∗ ∗
Further α41 = α42 = 1 because α14 = α24 = 0 and α43 = 1 because α24 + α43 =
α23 = 1 and α24 = 0. Thus E has the form
⎛ ⎞
0 1 0 0
⎜ 1 0 1 0 ⎟
⎜ ∗⎟
E = ⎜⎜ 1 1 0 1 ⎟.
⎟
⎝ 1 1 1 0 ⎠
∗ ∗
Continuing this process it follows that E is equal to
⎛ ⎞
0 1 0
... ... 0
⎜ .. ..
.. .. ⎟
⎜1 . .. .⎟
⎜ ⎟
⎜ .. . . . .. .. .. ⎟
⎜. . .. . . .⎟
E = ⎜
⎜.
⎟.
⎟
⎜ .. . . .
⎜
.. .. .. 0⎟ ⎟
⎜. .. .. ⎟
⎝ .. . . 1⎠
1 ... ... ... 1 0
and that σ(i) = i + 1, i = 1, . . . , n − 1. But then σ(n) has to be 1 and αn2 + α21 =
αn1 which is not true because αn2 = 1 = α21 = αn1 (if n ≥ 3). Thus σ(2) > 2
cannot hold and we must have σ(1) = 2, σ(2) = 1.
Quite generally this means that if Pi and Pj are two incomparable elements
then σ(i) = j, σ(j) = i. And further, if Pk (k = j) would also be incomparable
with Pi , σ(i) = k, σ(k) = i which cannot be. So for any Pk , k = i, j, Pk is
comparable with both Pi and Pj . This means that P splits up in a number s of
pairs of incomparable elements which after renumbering can be labelled
But nothing can be done by renumbering about the pattern odd singletons and
pairs in the poset PE . For instance PE could look like
1 2 3
• • •
7 • • • • 10
8 9
• • •
4 5 6
Now we show that if such an E is to be Gorenstein there can not be both pairs
and singletons.
Indeed suppose that there are both pairs and singletons. Now σ takes each
pair {Pi , Pj } into itself. It follows that σ takes singletons into singletons. Thus
there must be at least one pair and 2 singletons. This means that PE must have
one of the 3 following subposets.
j
•
i • • l (a)
•
k
j
•
i • • (b)
l
•
k
j
•
i • • l (c)
•
k
σ(j) = k, σ(k) = j, σ(i) = l.
Indeed take for i the minimal singleton (among the singletons) and for {j, k}
the minimal pair (among the pairs). Suppose i ≺ j, k and σ(i) ≺ j, k then we
have case (b) with i = i and l = σ(i). Next suppose i ≺ j, k and σ(i) j, k. Let
A = {c : c is a singleton and c ≺ j, k}, B = {c : c is a singleton and c j, k}.
Then σ takes A∪B into itself and as σ(i ) ∈ B, i ∈ A there must be an l ∈ B with
σ(l) ∈ A. Take i = σ(l). This gives subposet (a). Finally if i j, k, σ(i ) i (as
i is minimal), so take i = i , l = σ(i) to find case (c).
Case (a). The Gorenstein condition says
r holds only for s = S, i.e., it holds only when all the “row-sums” s1 , s2 , . . . , sn
are equal.
Definition. Let X and Y be any two (disjoint) posets. The ordinal sum
X ⊕Y of X and Y is the set of all x ∈ X and y ∈ Y (as a set, i.e. the disjoint union
of the sets X and Y ). The ordering relation on it is defined as following: x ≺ y
for all x ∈ X and y ∈ Y ; the relations x x1 and y y1 (x, x1 ∈ X; y, y1 ∈ Y )
are as before.
For instance
(• •) ⊕ (• •) = • • • •
and
•
(• •) ⊕ (• •) = • •
•
The ordinal sum is associative (but not commutative), and we can consider the
ordinal power X ⊕n = X ⊕ . . . ⊕ X for any poset X.
n
In particular, CHn = CH1⊕n and P2n = ACH2⊕n .
If X and Y are finite posets, then
EX 0m×n
EX⊕Y = ,
Un×m EY
Recall that the index of a poset P is the largest eigenvalue of the adjacency
)].
matrix [Q(P
inx P ≤ w(P).
Example 7.4.1.
The quiver Q with adjacency matrix
⎡ ⎤
0 1 1
[Q] = ⎣ 1 0 1 ⎦
1 1 0
Proof. The equalities inx P2n = w(P2n ) = 2 follow from proposition 7.4.1.
Let P = {p1 , . . . , pn }, n ≥ 3 and inx P = 2. We shall show first that Q(P) has
no loops when n ≥ 3. For n = 2, w(P) = 2 implies directly that P = ACH2 . Let
pn be an isolated element. Then, as w(P) = 2, {p1 , . . . , pn−1 } is a chain CHn−1 .
One can suppose that
p1 ≺ p2 ≺ . . . ≺ pn−1 .
Thus, for n ≥ 3
⎛ ⎞
0 1 0 ... ... 0
⎜ .. .. .. .. .. ⎟
⎜. . . . .⎟
⎜ ⎟
⎜ .. .. .. .. .. ⎟
⎜ . . . .⎟
[Q(P)] = ⎜ . ⎟.
⎜ .. .. ⎟
⎜0 . . 0⎟
⎜ ⎟
⎝1 0 ... ... 0 1⎠
1 0 ... ... 0 1
We have s1 = 1 and sn = 2. By corollary 7.4.2, 1 < inx P < 2 contradicting
inx (P) = 2. So Q(P) has no loops as required. Consequently, the (0, 1)-matrix
[Q(P)] with inx P = 2 has the zero main diagonal and exactly two 1’s in each
GORENSTEIN MATRICES 359
row. Thus, Pmin consists of two elements as there is an arrow from each maximal
element to all minimal elements.
Denote by P T the poset anti-isomorphic to P. Obviously, inx P = inx P T .
Then inx P T = 2 and P T has exactly two minimal elements. Hence P also has
precisely two maximal elements, say pn−1 and pn . Thus, one can assume that
Pmin = {p1 , p2 }, Pmax = {pn−1 , pn }. The (0, 1)-matrix [Q(P)] has zero main
diagonal and exactly two 1’s in each row and in each column.
Every partial order can be refined to a total order. Number the elements of
the poset P \ {1, 2, n − 1, n} according to such a total order with the numbers
3, 4, . . . , n − 2. Then in P
pi ≺ pj ⇒ i < j (7.4.1)
From everyi, 1 ≤ i ≤ n − 2 there issue precisely two arrows of Q(P) and in every i,
3 ≤ i ≤ n there arrive precisely 2 arrows (because each column and row of [Q(P)]
has precisely two entries equal to 1 and is otherwise zero).
Now consider the element 3. Two arrows must terminate in p3 and as there
are no loops these must come from p1 and p2 because of (7.4.1). Thus
p1 • • p3
P = ...? ...
p2 •
Next consider the element p4 . Two arrows must terminate in p4 . These must
come from {p1 , p2 , p3 }. Suppose one of them comes from p3 . Let the other come
from p1 . This gives p1 p3 p4 and p1 p4 which is a contradiction
as p4 then does not cover p1 . Similarly, if the second arrow arriving in 4 comes
from p2 we would have p2 p3 p4 , p2 p4 also a contradiction. Thus
the two arrows arriving in p4 come from p1 and p2 and P looks like
p1 • • p3
...
p2 • • p4
Now consider p5 . There are two arrows terminating in p5 . These must come
from {p1 , p2 , p3 , p4 }. But there are already 2 arrows starting in p1 , p2 . Thus these
two arrows ending in p5 must come from p3 , p4 and P is of the form
p1 p3 p5
• • •
...
• •
p2 p4
360 ALGEBRAS, RINGS AND MODULES
p1 p3 p5
• • •
...
• • •
p2 p4 p6
Continuing in this way the partially ordered set P would look like
p1 p3 pn−4 pn−2 pn
• • • • •
...
• • • •
p2 p4 pn−3 pn−1
if n is odd. But that contradicts that there are two maximal elements. Thus n is
even and
p1 p3 pn−3 pn−1
• • • •
P = ... = (ACH2 )⊕n/2
• • • •
p2 p4 pn−2 pn
Remark 7.4.2. Similarly, one can show that if inx P = w(P) = m and Q(P)
⊕n
has no loops, then P = ACHm .
Denote by Mn (Z) the ring of all square n × n matrices over the integers Z. Let
A = (aij ) ∈ Mn (Z).
GORENSTEIN MATRICES 361
Examples 7.4.2.
I. Let ⎛ ⎞
0 0 0
E3 = ⎝ 2 0 0⎠
2 2 0
be the Gorenstein matrix with the permutation
1 2 3
σ = .
3 1 2
Straight forward calculation gives
⎛ ⎞
1 1 0
(2)
E3 = ⎝ 2 1 1⎠ and [Q(E3 )] = E + Pσ2 .
3 2 1
3 2
362 ALGEBRAS, RINGS AND MODULES
II. Let ⎛ ⎞
0 0 0 0
⎜2 0 1 0⎟
E4 = ⎜
⎝1
⎟
1 0 0⎠
2 1 2 0
be the Gorenstein matrix with permutation
1 2 3 4
σ = .
4 1 2 3
One calculates
⎛ ⎞ ⎛ ⎞
1 1 1 0 0 1 1 0
⎜2 1 2 1⎟ ⎜0 0 1 1⎟
= ⎜ ⎟ and [Q(E4 )] = Pσ2 + Pσ3 = ⎜ ⎟.
(2)
E4 ⎝2 1 1 1⎠ ⎝1 0 0 1⎠
3 2 2 1 1 1 0 0
Hence, Q(E4 ) has the following form:
4 2
III. Let ⎛ ⎞
0 0 0 0 0
⎜2 0 1 1 0⎟
⎜ ⎟
E5 = ⎜
⎜1 1 0 1 0⎟⎟
⎝1 0 1 0 0⎠
2 1 1 2 0
be the Gorenstein matrix with permutation
1 2 3 4 5
σ = .
5 1 2 3 4
By straight forward calculation
⎛ ⎞ ⎛ ⎞
1 0 1 1 0 0 0 1 1 0
⎜2 1 1 2 1⎟ ⎜0 0 0 1 1⎟
⎜ ⎟ ⎜ ⎟
E5 = ⎜ 1⎟ ⎜ 1⎟
(2)
⎜2 1 1 1 ⎟ and [Q(E5 )] = ⎜1 0 0 0 ⎟.
⎝2 1 1 1 0⎠ ⎝1 1 0 0 0⎠
2 2 2 2 1 0 1 1 0 0
So, Q(E5 ) has the following form
GORENSTEIN MATRICES 363
1 3
5 4
IV. Let ⎛ ⎞
0 0 0 0 0 0
⎜2 0 1 1 1 0⎟
⎜ ⎟
⎜1 1 0 1 1 0⎟
E6 = ⎜
⎜1
⎟
⎜ 0 1 0 1 0⎟
⎟
⎝1 0 0 1 0 0⎠
2 1 1 1 2 0
be the Gorenstein matrix with permutation
1 2 3 4 5 6
σ = .
6 1 2 3 4 5
Then
⎛ ⎞ ⎛ ⎞
1 0 0 1 1 0 0 0 0 1 1 0
⎜2 1 1 1 2 1⎟ ⎜0 0 0 0 1 1⎟
⎜ ⎟ ⎜ ⎟
⎜2 1 1 1 1 1⎟ ⎜ 1⎟
E6
(2)
= ⎜ ⎟ [Q(E6 )] = Pσ2 + Pσ3 = ⎜1 0 0 0 0 ⎟.
⎜2 1 1 1 1 0⎟⎟ ⎜1 1 0 0 0 0⎟
⎜ ⎜ ⎟
⎝1 1 1 1 1 0 ⎠ ⎝0 1 1 0 0 0⎠
2 1 2 2 2 1 0 0 1 1 0 0
And Q(E6 ) looks as follows
6 2
5 3
4
364 ALGEBRAS, RINGS AND MODULES
7.5 D-MATRICES
Definition. Let A ∈ Mm×n (R) and A ≥ 0, i.e., if A = (aij ), then aij ≥ 0.
n m
We say that A is a d-matrix for some d > 0, if aij = d and aij = d
j=1 i=1
for all i, j.
Remark 7.5.1. Let A be a d-matrix and let there exists a permutation matrix
Pτ such that
⎛ ⎞
0 A12 0 ... 0
⎜ 0 0 A23 . . . 0 ⎟
⎜ ⎟
T ⎜ .. . . . .. ⎟
B = Pτ APτ = ⎜ . .. .. .. . ⎟
⎜ ⎟
⎝ 0 0 0 . . . Ah−1h ⎠
Ah1 0 0 ... 0
By lemma 7.5.2, PτT APτ = B is a d-matrix. Therefore the matrices
A12 , A23 , . . . , Ah1 are d-matrices. By lemma 7.5.1, all these matrices are square.
Example 7.5.1.
By theorem 7.3.3, any Gorenstein (0, 1)-matrix E is equivalent to either Hs
or G2s . In the case Hs we obtain the adjacency matrix [Q(Hs )] of Q(Hs ) in the
following form
[Q(Hs )] = Pσ ,
where σ = (12 . . . s). The exponent matrix G2s is equivalent to the matrix E2s
366 ALGEBRAS, RINGS AND MODULES
⎛ ⎞
A O O ... O O
⎜U A O ... O O⎟
⎜ ⎟
⎜ .. ⎟ ,
E2s = ⎜ ... ..
.
..
.
.. .
. .. .⎟
⎜ ⎟
⎝U U U ... A O⎠
U U U ... U A
where
0 1 1 1 0 0
A = , U = , O = .
1 0 1 1 0 0
Obviously,
⎡ ⎤
0 U2 0 ... 0
⎢ 0 0 U2 ... ⎥
0
⎢ ⎥
⎢ .. .. .. .. ⎥
..
[Q(E2s )] = ⎢ . . . . ⎥.
.
⎢ ⎥
⎣ 0 0 0 . . . U2 ⎦
U2 0 0 ... 0
We recall some facts which can be found in the book [Gantmakher, 1959]
(Chapter III, §5 Primitive and imprimitive matrices).
Remark 7.5.2. An algebraic proof of this theorem was given by I.N. Herstein
in the paper [Herstein, 1954].
In the formulation of the next corollary we use the notation of theorem 6.5.2.
Ah = diag {A12 . . . Ah−1n Ah1 , A23 . . . Ah1 A12 , . . . , Ah1 A12 . . . Ah−1h },
where each block is permutationally irreducible, and each block has the same max-
imal characteristic number.
Lemma 7.5.9. If Q is a quiver and [Q]m = (tij ), then tij is the number of all
paths from a vertex i to a vertex j of length m.
Proof. Let Q be regular. By theorem 7.5.5, there exists an integer m such that
[Q]m > 0. Therefore [Q]m+1 > 0 and there exist two cycles from 1 to 1. The
first cycle has the length m, the second m + 1. So, the greatest common divisor
of all cycles equals 1.
If Q is not regular then, by theorem 6.5.2, there exists a permutation matrix
368 ALGEBRAS, RINGS AND MODULES
Pτ such that
⎛ ⎞
0 A12 0 ... 0
⎜ 0 0 A23 ... 0 ⎟
⎜ ⎟
⎜ .. .. .. .. .. ⎟
PτT [Q]Pτ =⎜ . . . . . ⎟,
⎜ ⎟
⎝ 0 0 0 . . . Ah−1,h ⎠
Ah1 0 0 ... 0
where there are square zero blocks along the main diagonal.
Denote by mi the order of the i-th zero square block. The set V Q of vertices
of Q may be numerated in such a way that
V Q = V1 ∪ V2 ∪ . . . ∪ Vh , Vi ∩ Vj = 0,
for i = j and | Vi | = mi for i = 1, . . . , h. Obviously, in Q there exists an arrow
k → l only in the case if k ∈ Vi and l ∈ Vi+1 for i = 1, . . . , h − 1 and k ∈ Vh for
i = h, l ∈ V1 . So, in Q the greatest common divisor d of the lengths of all cycles is,
at least, h ≥ 2. We shall show that if d = 1, then [Q] is primitive. The matrix [Q]
is permutationally irreducible, therefore, by definition, [Q] is either primitive or
imprimitive. If [Q] is imprimitive, then d ≥ h ≥ 2 and we obtain a contradiction.
Therefore, [Q] is primitive and Q is regular. The theorem is proved.
This theorem was first proved in the book [Dulmage, Mendelsohn, 1967].
Remark 7.5.3. Let Q be the quiver with adjacency matrix
⎛ ⎞
0 0 0 1
⎜0 0 1 0⎟
[Q] = ⎜⎝1 0 0 0⎠ .
⎟
0 1 0 0
Obviously, Q is the simple cycle
1 2
• •
• •
3 4
and h = 2 is not the greatest common divisor of all cycles.
Assume we have a homogeneous Markov chain with a finite number of states
and transition matrix P = (pij ) ∈ Mn (R). Let Q(P ) be the simply laced quiver
associated with P (see vol.I, pp. 275-276 and section 6.6 above).
Corollary 7.5.11. A Markov chain is regular if and only if the quiver Q(P )
of its transition matrix is regular, i.e., the greatest common divisor of all cycles of
Q(P ) equals 1.
GORENSTEIN MATRICES 369
Here the elements of Z/(2) × Z/(2) are numbered as follows: (0, 0) has the label
0, and (1, 0), (0, 1), (1, 1) have respectively the labels 1, 2, 3.
Consider
Γk−1 Γk−1 + Xk−1
Γk = .
Γk−1 + Xk−1 Γk−1
Proof. The proof goes by induction on k. The basis of the induction has already
been done. If Γk−1 is the Cayley table of Gk−1 , then, by proposition 7.6.2, Γk is
the Cayley table of Gk .
αk+1
2k +i,j
= αk+1
i,2k +j
= αkij + 2k , αk+1
2k +i,2k +j
= αk+1
ij = αkij for all i, j = 1, 2, . . . , 2k
and (αkij + 2k ) + αkjσk (i) = (αkij + αkjσk (i) ) + 2k = αkiσk (i) + 2k , we obtain that
αk+1
ij + αk+1
j,2k +σk (i)
= αk+1
i,2k +σk (i)
, αk+1
i,2k +j
+ αk+1
2k +j,2k +σk (i)
= αk+1
i,2k +σk (i)
,
αk+1
2k +i,2k +j
+ αk+1
2k +j,σk (i)
= αk+1
2k +i,σk (i)
, αk+1
2k +i,j
+ αk+1 k+1
jσk (i) = α2k +i,σk (i) ,
(2) (1)
(2) Γk−1 Γk−1 + Xk−1
Γk = (1) (2) .
Γk−1 + Xk−1 Γk−1
Hence, 7 8
[Q(Γk−1 )] E
[Q(Γk )] = .
E [Q(Γk−1 )]
Here is the characteristic polynomial χk+1 (x) = χ[Q(Γk+1 ] (x).
9 9
9 xE − [Q(Γk )] −E 9
χk+1 (x) = |xE − [Q(Γk+1 )]| = 99 9=
9
−E xE − [Q(Γk )]
9 9
9 xE − [Q(Γk )] − E 0 9
= 99 9=
−E xE − [Q(Γk )] + E 9
= |(x − 1)E − [Q(Γk )]| · |(x + 1)E − [Q(Γk )]|
Therefore,
χk+1 (x) = χk (x − 1) · χk (x + 1). (7.6.1)
Since 9 9
9 x−1 −1 99
χ1 (x) = 99 = x(x − 2),
−1 x−1 9
we obtain χ2 (x) = (x − 3)(x − 1)(x − 1)(x + 1) = (x − 3)(x − 1)2 (x + 1),
χ3 (x) = (x − 4)(x − 2)2 x(x − 2)x2 (x + 2) = (x − 4)(x − 2)3 x3 (x + 2).
m i
m!
Proposition 7.6.5. χm (x) = (x − m − 1 + 2i)Cm , where Cm
i
= .
i=0 (m − i)!i!
k
i
k−1
j
= (x − k − 2) (x − k − 2 + 2i)Ck · (x − k + 2j)Ck (x + k) =
i=1 j=0
k−1
i+1
k−1
j
= (x − k − 2) (x − k + 2i)Ck · (x − k + 2j)Ck (x + k) =
i=0 j=0
k−1
i i+1
= (x − k − 2) (x − k + 2i)Ck +Ck (x + k).
i=0
372 ALGEBRAS, RINGS AND MODULES
k−1 i+1
Since Cki +Cki+1 = Ck+1
i+1
, we obtain χk+1 (x) = (x−k−2) (x−k+2i)Ck+1 (x+k) =
i=0
k j
k+1 j
(x − k − 2) (x − k + 2(j − 1))
Ck+1
(x + k) = (x − (k + 1) − 1 + 2j)Ck+1 .
j=1 j=0
k k
2
2
By induction on k, it is easy to prove that qij (Γk ) = k+1, qij (Γk ) = k+1.
i=1 j=1
Thus, [Q(Γk )] = (k + 1)Pk , where Pk is a doubly stochastic matrix.
Examples 7.6.1.
I. The Latin square
⎛ ⎞
0 1 2 3
⎜1 0 3 2⎟
L4 = ⎜
⎝3
⎟
2 0 1⎠
2 3 1 0
is a Gorenstein matrix with permutation
1 2 3 4
σ = σ(L4 ) =
4 3 1 2
and ⎛ ⎞
1 1 1 0
⎜ 1 1 0 1⎟
[Q(L4 )] = ⎜
⎝0
⎟.
1 1 1⎠
1 0 1 1
So [Q(L4 )] = E + Pσ2 + Pσ3 and
1 2
Q(L4 ) =
3 4
⎛ ⎞ ⎛ ⎞
1 1 2 3 2 2 3 3
⎜1 1 3 2⎟ ⎜2 2 3 3⎟
= ⎜ ⎟; = ⎜ ⎟.
(1) (2)
Γ2 ⎝2 Γ2
3 1 1⎠ ⎝3 3 2 2⎠
3 2 1 1 3 3 2 2
⎛ ⎞
1 1 1 0
⎜1 1 0 1⎟
[Q(Γ2 )] = ⎜
⎝1
⎟
0 1 1⎠
0 1 1 1
1 2
Q(Γ2 ) =
3 4
Theorem 7.6.6. Suppose that a Latin square Ln with first row and first
column (0 1 . . . n − 1) is an exponent matrix. Then n = 2m and Ln = Γm is the
Cayley table of the direct product of m copies of the cyclic group of order 2.
Conversely, the Cayley table Γm of the elementary Abelian group Gm =
Z/(2) × . . . × Z/(2) = (2) × . . . × (2) (m factors) of order 2m is a Latin square and
a Gorenstein symmetric matrix with first row (0, 1, . . . , 2m − 1) and permutation
1 2 3 . . . 2m − 1 2m
σ(Γm ) = .
2m 2m − 1 2m − 2 ... 2 1
| i − j | ≤ αij ≤ i + j − 2.
|i + j − (n + 1)| ≤ |n − 1 − αij |.
Proof. By lemma 7.6.8 and corollary 7.6.9, we have αij ≤ αin + αnj = (n − i) +
(n − j) = 2n − (i + j). By lemma 7.6.7, αij ≤ i + j − 2. From the first inequality
we have
αij − (n − 1) ≤ n + 1 − (i + j).
From the second inequality we have
αij − (n − 1) ≤ (i + j) − (n + 1).
So
|(i + j) − (n + 1)| ≤ |αij − (n − 1)|.
⎛ ⎞
0 1
⎜ 1 0 ∗ ... ∗ ⎟ ⎛ ⎞
⎜ ⎟ Γ1 ∗ ... ∗
⎜ 0 1 ⎟
⎜ ∗ ... ∗ ⎟ ⎜ ∗ Γ1 ... ∗ ⎟
⎜ 1 0 ⎟ ⎜ ⎟
Ln = ⎜ ⎟=⎜ .. .. .. .. ⎟.
⎜ .. .. .. .. ⎟ ⎝ . . . . ⎠
⎜ . . . . ⎟
⎜ ⎟ ∗ ∗ . . . Γ1
⎝ 0 1 ⎠
∗ ∗ ...
1 0
⎛ ⎞
0 1 2
⎜ ⎟
⎜ 1 0 2 ⎟
⎜ * ⎟
⎜ 2 0 1 ⎟
⎜ ⎟
⎜ ⎟
⎜ 2 1 0 ⎟
⎜ ⎟
⎜ 0 1 2 ⎟
⎜ * ⎟
⎜ 2 ⎟
⎜ 1 0 ⎟
⎜ ⎟
⎝ 2 0 1 ⎠
2 1 0
The next and final step is to prove by induction the following two statements.
αi,i+2k = 2k for i = 1, . . . , 2k .
It helps to realize that this is a statement about the 2k × 2k block indicated
by ∗∗ above. The claim is proved by induction. For i = 1 it is true by the
assumptions of theorem 7.6.6. Now let i > 1. The number 2k must occur at one
of the places (i, 1), (i, 2), . . . , (i, i − 1), (i, i), (i, i + 1), . . . , (i, i + 2k ). The places
(i, 1), . . . , (i, 2k ) are already filled. Further the columns 2k + 1, . . . , 2k + i − 1
already contain a 2k by the induction hypothesis. So it must be the case that
αi,i+2k = 2k .
Similarly, switching rows and columns in the argument, one finds
αi+2k ,i = 2k for i = 1, 2, . . . , 2k .
Moreover, this pattern persists in that
2k in them. So it must be that 2k occurs at position (t2k+1 +i, t2k+1 +i+2k ). This
proves the first half of (7.6.2). The second half is handled by the same argument
(switching rows and columns).
Now suppose that the number of blocks Γk is odd. Then by what has just been
proved the matrix Ln looks like
⎛ ⎞
2k
⎜ .. ⎟
⎜ Γk . ⎟
⎜ ⎟
⎜ ⎟
⎜ 2k ⎟
⎜ ⎟
⎜ 2k ⎟
⎜ ⎟
⎜ .. ⎟
⎜ . Γk ⎟
⎜ ⎟
⎜ ⎟
⎜ 2k ⎟
⎜ ⎟
⎜ .. ⎟
⎜ . ⎟
⎜ ⎟
⎜ 2k ⎟
⎜ ⎟
⎜ ⎟
⎜ Γk .. ⎟
⎜ . ⎟
⎜ ⎟
⎜ 2k ⎟
⎜ ⎟
⎜ 2k ⎟
⎜ ⎟
⎜ ⎟
⎜ .. Γk ** ⎟
⎜ . ⎟
⎜ ⎟
⎜ 2 k ⎟
⎜ ⎟
⎜ ⎟
⎜ ⎟
⎜ ⎟
⎝ ** Γk ⎠
But then there is no 2k in the blocks indicated by ∗∗’s and hence no 2k in the
last 2k rows and columns because αij = 2k implies | i − j | ≤ 2k .
Let Yt denote the t-th 2k+1 × 2k+1 diagonal block of Ln , t = 1, 2, . . . , 2−1 nk ,
and write
Γk E12
Yt = .
E21 Γk
As 0, 1, 2, . . . , 2k −1 occur in each row (and column) of Γk and Ln is a Latin square
all elements of E12 must be ≥ 2k . On the other hand for any element of an E12
Since 2k U2k ≤ E12 < 2k+1 , we obtain 0 ≤ E12 − 2k U2k < 2k , and E12 − 2k U2k
is a Latin square on the set {0, 1, 2, . . . , 2k − 1}.
Since 2k U2k ≤ E21 < 2k+1 , we obtain 0 ≤ E21 − 2k U2k < 2k , and E21 − 2k U2k
is a Latin square on the set {0, 1, 2, . . . , 2k − 1}.
urthermore
αt·2k+1 +i,t·2k+1 +2k +1 ≤ αt·2k+1 +i,t·2k+1 +1 +αt·2k+1 +1,t·2k+1 +2k +1 = (i−1)+2k = 2k +i−1
αt·2k+1 +2k +i,t·2k+1 +1 ≤ αt·2k+1 +2k +i,t·2k+1 +i +αt·2k+1 +i,t·2k+1 +1 = 2k +(i−1) = 2k +i−1
for all i = 1, 2, . . . , 2k . Thus for i = 1, 2, . . . , 2k we have
αt·2k+1 +i,t·2k+1 +2k +1 = 2k + i − 1
Γm−1 Γm−1 + Xm−1
Γm = , Xm−1 = 2m−1 U2m−1 .
Γm−1 + Xm−1 Γm−1
By proposition 7.6.3, Γm is the Cayley table of the elementary Abelian group
Gk of order 2k . Theorem 7.6.2 is proved.
Remark 7.6.1. Example 7.6.1(I) shows, that for theorem 7.6.2 the condi-
tion for a Latin square having the first row and the first column of the form
(0 1 . . . n − 1) is essential.
Corollary 7.7.2. For any fixed point free permutation σ there exists a semidis-
tributive weakly prime Artinian Frobenius ring B with ν(B) = σ.
Lemma 7.7.3. Let e and f be non-zero idempotents of a ring A such that the
modules eA and f A are indecomposable. In this case eA and f A are isomorphic
if and only if the following equality holds: f = f λ0 eλ1 , where λ0 , λ1 ∈ A.
Theorem 7.7.4. Let A = {O, E(A)} be a reduced tiled Gorenstein order with
Jacobson radical R and let J be a two-sided ideal of A such that A ⊃ R2 ⊃ J ⊃ Rn
(n ≥ 2). The quotient ring A/J is quasi-Frobenius if and only if there exists a
p ∈ R2 such that J = pA = Ap.
s s
Proof. Let A = i,j=1 eij π αij O, and let Ms (D) = i,j=1 eij D be the ring of
fractions of A, where D is the division ring of fractions of O, and the eij are the
matrix units (i, j = 1, . . . , s). Let J = pA = Ap be a two-sided ideal of A and A ⊃
R2 ⊃ J ⊃ Rn . Obviously, Ms (D)J Ms (D) = Ms (D)pAMs (D) = Ms (D)pMs (D)
is a non-zero two-sided ideal of Ms (D). Therefore p ∈ Ms (D) has an inverse p−1
in Ms (D). Since the quotient ring A/Rn is Artinian, the quotient ring Ā = A/J
is also Artinian. We now show that the quotient ring Ā is quasi-Frobenius. Let
GORENSTEIN MATRICES 381
In [Nishida, 1988] there is the example of the (0, 1)-order Λ(P5 ) associated with
the finite poset
• •
P5 = •
• •
Let Λ be an O-order such that μΛ (I) ≤ 2 for each left ideal I of Λ. Then Λ is
a Bass order.
385
386 ALGEBRAS, RINGS AND MODULES
42. D.Simson, Linear Representation of Partially Ordered Sets and Vector Space
Categories, Algebra, Logic and Appl. v.4, Gordon and Breach Science Pub-
lishers, 1992.
43. B.Stenström, Rings of quotients: An introduction to methods of ring theory,
Springer-Verlag, 1975.
44. H.Tachikawa, Quasi-Frobenius Rings and Generalizations: QF-3 and QF-1
Rings, Lecture Notes in Math., v. 345, Springer-Verlag, Berlin-Heidelberg-
New York, 1973.
45. W.T.Trotler, Combinatorics and Partially Ordered Sets: Dimension Theory,
The John Hopkins Univ. Press, 1992.
46. A.A.Tuganbaev, Semidistributive Modules and Rings, Kluwer Academic
Publishers, Dordrecht-Boston-London, 1998.
47. J.Waschbüsch, On selfinjective algebras of finite representation type, Univer-
sidad Nacional Autonoma de Mexico, 1983.
Subject Index
A canonical two-sided Peirce decompo-
Abelian group, 2 sition of a ring, 166
additive group, 2 canonical two-sided Peirce decompo-
adjacency matrix, 68 sition of an ideal, 166, 167
admissible ideal, 74 cardinal sum of posets, 131
admissible relation, 73 Cartan determinant, 246
affine group, 6 Cartan determinant conjecture, 247
affine variety, 93 Cartan invariants, 245
algebra of finite type, 140 Cartan matrix, 93, 245
algebra of finite representation type, category of representations, 99
140 Cayley formula, 241
algebra of infinite representation Cayley table, 4
type, 140 Cayley theorem, 12
algebra of right bounded finite center, 17
dimensional representation centralizer, 17
type, 155 chain, 130, 259
algebra of a quiver with relations, 74 character of a group, 38
algebra of tame representation type, character of a representation, 39
141 character table, 44
algebra of wild representation type, circuit of a quiver, 239
141, 142 circulant matrix, 248
algebraic group, 94 class equation, 17
alternating group, 5 commutative group, 2
antichain, 259 commutative relation, 74
anticommutator, 55 commutator, 22
anti-ismorphism of lattices, 162 commutator series of a group, 23
arrow of a quiver, 67 commutator subgroup, 22
Artinian algebra, 147 complete lattice, 256, 274
associated poset of a Markov chain, completely decomposable lattice, 274
295 completely reducible representation,
automorphism, 12 31, 138
composition series of a group, 21
B conjugacy class, 16
band, 256 conjugate elements, 16
bijective module, 209 conjugating element, 16
bound quiver algebra, 74 connected components, 286
Brauer-Thrall conjectures, 154 connected poset, 282
coordinate vector, 115
contragredient representation, 136
C
Coxeter group, 109
canonical decomposition, 166
Coxeter matrix, 109
canonical epimorphism, 12
389
390 SUBJECT INDEX
V Weierstrass-Dedekind theorem, 36
valued graph, 87 width of a poset, 130, 259
vertex of a quiver, 67 width of a reduced exponent
(0,1)-matrix, 347
W width of a tiled order, 261
weak dimension, 222 wild algebra, 142
weakly non-negative quadratic wild group, 48
form, 127
weakly positive quadratic form, 127 Z
weakly prime ring, 305 Zariski topology, 93
weakly symmetric algebra, 172, 279 zero, 2
Wedderburn-Artin theorem, 35 zero-relation, 73
Wedderburn ring, 61 zero representation, 76
Name Index
A D
Abel N.H., 49 Danlyev Kh.M., 317, 324, 325
Albamowicz R., 385 Davis P.J., 248, 253
Alperin J.L., 385 Dickson L.E., 34, 35
Armstrong M.A., 385 Dieudonné J., 213, 214
Arnold D.M., 117, 154, 156, 385 Dilworth R.P., 259, 323, 325
Arnold V.I., 109 Dinh H., 214, 215
Assen I., 214 Dlab V., 89,90, 92, 99–102, 110,
Auslander M., 141, 155, 156, 202, 145–147, 155,
213, 214, 223, 252, 385 156, 318, 325, 385
Dokuchaev M.A., 213, 215, 323–325
B Donovan P., 98, 101, 102, 140,
Bass H., 327, 381, 383 153, 156, 252, 253
Bautista R., 141, 156 Donnelly R.G., 110
Bell A.D., 155, 156 Drozd Yu.A., 48–50, 113, 128,
Berstein I.N., 53, 93, 101, 102 140, 142, 153–156, 252,
Birkhoff G.D., 341 253, 382, 383, 386
Blyth T.S., 385 Dugas M., 154, 156
Bondarenko V.M., 48–50, 154, 156 Dulmage A.L., 368, 383
Bongartz K., 141, 155
Borel A., 110
Bourbaki N., 102 E
Brauer R., 47, 50, 154, 155, 212–214, Eilenberg S., 100, 102, 201, 213,
Brenner S., 154 215, 246, 252, 253, 385
Bretscher O., 214 Eklof P.C., 189
Browder F.E., 111 Erdmann K., 214, 215
Burgess W.D., 247, 252 Euler L., 49
Burnside W., 2, 49, 50
Butler M.C.R., 154
Byrnes C., 111 F
Facchini A., 385
Faddeev D.K., 385
C Faith C., 161, 202, 212, 215, 385
Cartan E., 100 Fedorov E.S., 7, 50, 51
Cartan H., 385 Freislich M.R., 98, 101, 102, 140,
Cayley A.L., 2, 12, 49, 50, 136 153, 156, 252, 253
Chatters A.W., 382, 383 Friedom R., 110
Chernousova Zh.T., 323–325 Frobenius F.G., 1, 42, 50, 215,
Crawley-Boevey W., 93, 102 288, 325
Curtis C.W., 173, 213, 214, Fujita H., 308, 323, 325
381–383, 385 Fuller K.R., 246, 247, 252, 253
397
398 NAME INDEX
J M
Jacobson N., 386 MacLane S., 386
James G., 386 Malle G., 111
Maschke H., 31, 34, 50, 51
NAME INDEX 399