Claus Michael Ringel, Jan Schröer - Representation Theory of Algebras I - Modules (Preliminary Version Feb 10, 2009) (2009) PDF
Claus Michael Ringel, Jan Schröer - Representation Theory of Algebras I - Modules (Preliminary Version Feb 10, 2009) (2009) PDF
Claus Michael Ringel, Jan Schröer - Representation Theory of Algebras I - Modules (Preliminary Version Feb 10, 2009) (2009) PDF
Preliminary version
Contents
1. Introduction 10
1.1. About this course 10
1.2. Notation and conventions 11
1.3. Acknowledgements 11
2.3. Submodules 12
2.4. Factor modules 14
2.5. The lattice of submodules 15
2.6. Examples 16
2.10. Exercises 22
3. Homomorphisms between modules 23
3.1. Homomorphisms 23
3.2. Definition of a ring 24
3.3. Definition of an algebra 25
3.4. Homomorphism Theorems 25
3.9. Exercises 35
4. Digression: Categories 37
4.1. Categories 37
4.2. Functors 38
7.4. Exercises 61
8. Filtrations of modules 63
8.1. Schreier’s Theorem 63
8.2. The Jordan-Hölder Theorem 66
8.3. Exercises 67
10.4. Exercises 78
11. Direct summands of finite direct sums 79
11.1. The Exchange Theorem 79
11.2. Consequences of the Exchange Theorem 80
References 82
11.3. Examples 82
11.4. Exercises 83
4 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
12.13. Bimodules 97
12.14. Modules over tensor products of algebras 99
12.15. Exercises 99
13. Semisimple algebras 99
27.6. Short exact sequences and the first extension group 178
27.7. The vector space structure on the first extension group 181
***********************************************************************
10 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
1. Introduction
1.1. About this course. This is the first part of notes for a lecture course “In-
troduction to Representation Theory”. As a prerequisite only a good knowledge of
Linear Algebra is required. We will focus on the representation theory of quivers
and finite-dimensional algebras.
The intersection between the content of this course and a classical Algebra course
just consists of some elementary ring theory. We usually work over a fixed field K.
Field extensions and Galois theory do not play a role.
This part contains an introduction to general module theory. We prove the classical
theorems of Jordan-Hölder and Krull-Remak-Schmidt, and we develop the represen-
tation theory of semisimple algebras. (But let us stress that in this course, semisim-
ple representations carry the label “boring and not very interesting”.) We also start
to investigate short exact sequences of modules, pushouts, pullbacks and properties
of Auslander Reiten sequences. Some first results on the representation theory of
path algebras (or equivalently, the representation theory of quivers) are presented
towards the end of this first part. We study the Jacobson radical of an algebra,
decompositions of the regular representation of an algebra, and also describe the
structure of semisimple algebras (which is again regarded as boring). Furthermore,
we develop the theory of projective modules.
As you will notice, this first part of the script concentrates on modules and algebras.
But what we almost do not study yet are modules over algebras. (An exception are
semisimple modules and projective modules. Projective modules will be important
later on when we begin to study homological properties of algebras and modules.)
Here are some topics we will discuss in this series of lecture courses:
***********************************************************************
12 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
2. Basic terminology
2.1. J-modules. Our aim is to study modules over algebras. Before defining what
this means, we introduce a very straightforward notion of a module which does not
involve an algebra:
Let J be a set (finite or infinite). This set is our “index set”, and in fact only the
cardinality of J is of interest to us. If J is finite, then we often take J = {1, . . . , n}.
We also fix a field K.
A J-module is given by (V, φj )j∈J where V is a vector space and for each j ∈ J we
have a linear map φj : V → V .
Often we just say “module” instead of J-module, and we might say “Let V be a
module.” without explicitly mentioning the attached linear maps φj .
For a natural number m ≥ 0 an m-module is by definition a J-module where
J = {1, . . . , m}.
2.2. Isomorphisms of J-modules. Two J-modules (V, φj )j and (W, ψj )j are iso-
morphic if there exists a vector space isomorphism f : V → W such that
f φj = ψj f
for all j ∈ J.
f
V / W
φj ψj
f
V / W
The dimension of a J-module (V, φj )j is just the dimension of the vector space V .
Matrix version: If V and W are finite-dimensional, choose a basis v1 , . . . , vn of V and
a basis w1 , . . . , wn of W . Assume that the isomorphism f : V → W is represented by
a matrix F (with respect to the chosen bases), and let Φj and Ψj be a corresponding
matrices of φj and ψj , respectively. Then F Φj = Ψj F for all j, i.e. F −1 Ψj F = Φj
for all j.
If two modules V and W are isomorphic we write V ∼
= W.
Example: Let
0 1
φ= .
0 0
Then the 1-module (K 2 , φ) has exactly three submodules, two of them are proper
submodules.
Matrix version: If V is finite-dimensional, choose a basis v1 , . . . , vn of V such that
v1 , . . . , vs is a basis of U. Let φj,U : U → U be the linear map defined by φj,U (u) =
φj (u) for all u ∈ U. Observe that (U, φj,U )j is again a J-module. Then the matrix
Φj of φj (with respect to this basis) is of the form
Aj Bj
Φj = .
0 Cj
In this case Aj is the matrix of φj,U with respect to the basis v1 , . . . , vs .
Let V be a vector space and X a subset of V , then hXi denotes the subspace of
V generated by X. This is the smallest subspace of V containing X. Similarly, for
elements x1 , . . . , xn in V let hx1 , . . . , xn i be the subspace generated by the xi .
P
Let I be a set, and for each i ∈ I let Ui be S a subspace of V . Then the sum i∈I Ui
is defined as the subspace hXi where X = i∈I Ui .
Let V = (V, φj )j be a module, and let X be a subset of V . The intersection U(X)
of all submodules U of V with X ⊆ U is the submodule generated by X. We
call X a generating set of U(X). If U(X) = V , then we say that V is generated
by X.
Lemma 2.1. Let X be a subset of a module V . Define a sequence of subspaces Ui of
V as follows: Let U0 be the subspace of V which is generated by X. If Ui is defined,
let X
Ui+1 = φj (Ui ).
j∈J
Then X
U(X) = Ui .
i≥0
Proof. Set
X
UX = Ui .
i≥0
P
We have xj = i≥0 uij for some uij ∈ Ui and all but finitely many of the uij are
P
zero. Thus there exists some N ≥ 0 such that xj = N i=0 uij for all 1 ≤ j ≤ n. Each
element in Ui is a finite linear combination of elements of the form φji · · · φj1 (y) for
some j1 , . . . , ji ∈ J and y ∈ Y . This yields the result.
Warning: Finite minimal generating sets of a module V do not always have the
same cardinality: Let V = M2 (K) be the vector space of 2 × 2-matrices, and take
the module given by V together with all linear maps A : V → V , A ∈ M2 (K). Then
{( 10 01 )} and {( 10 00 ) , ( 00 10 )} are minimal generating sets of V .
Lemma 2.3. A module V is finitelyPgenerated if and only if for each family Ui ,
i ∈ I of submodules
P of V with V = i∈I Ui there exists a finite subset L ⊆ I such
that V = i∈L Ui .
Proof.
P Let x1 , . . . , xn be a generating set of V , and letPUi be submodules with V =
U . Then each element xl lies in a finite sum i∈I(l) Ui . This implies V =
Pni∈I Pi
l=1 i∈I(l) Ui .
2.5. The lattice of submodules. A partially ordered set (or poset) is given
by (S, ≤) where S is a set and ≤ is a relation on S, i.e. ≤ is transitive (s1 ≤ s2 ≤ s3
implies s1 ≤ s3 ), reflexive (s1 ≤ s1 ) and anti-symmetric (s1 ≤ s2 and s2 ≤ s1
implies s1 = s2 ).
One can try to visualize a partially ordered set (S, ≤) using its Hasse diagram:
This is an oriented graph with vertices the elements of S, and one draws an arrow
s → t if s < t and if s ≤ m ≤ t implies s = m or m = t. Ususally one tries to draw
the diagram with arrows pointing upwards and then one forgets the orientation of
the arrows and just uses unoriented edges.
For example, the following Hasse diagram describes the partially ordered set (S, ≤)
with three elements s1 , s2 , t with si < t for i = 1, 2, and s1 and s2 are not comparable
in (S, ≤).
t
@
@
s1 s2
For a subset T ⊆ S an upper bound for T is some s ∈ S such that t ≤ s for all
t ∈ T . A supremum s0 of T is a smallest upper bound, i.e. s0 is an upper bound
and if s is an upper bound then s0 ≤ s.
Proof.
P Straightforward: TheTsupremum of a set {Ui | i ∈ I} of submodules is
i∈I Ui , and the infimum is i∈I Ui .
Lemma 2.6 (Dedekind). Let U1 , U2 , W be submodules of a module V such that
U1 ⊆ U2 . Then we have
U1 + (W ∩ U2 ) = (U1 + W ) ∩ U2 .
Proof. It is sufficient to proof the statement for subspaces of vector spaces. The
inclusion ⊆ is obvious. For the other inclusion let u ∈ U1 , w ∈ W and assume
u + w ∈ U2 . Then w = (u + w) − u belongs to U2 and thus also to W ∩ U2 . Thus
u + w ∈ U1 + (W ∩ U2 ).
The non-zero proper submodules are he1 , e2 i, he3 , e4 i, he1 + e3 , e2 + e4 i and he1 +
2e3 , e2 + 2e4 i.
(d):
Let
c1 0 0 1
0 c2
φ= and ψ = 1 0
c3 0 0 1
0 c4 1 0
with pairwise different ci . Then the lattice of submodules of (K 4 , φ, ψ) looks like
this:
18 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
r
@
@@r
r
@
@
@r
• V 6= 0,
• Let U1 and U2 be submodules of V with U1 ∩ U2 = 0 and U1 + U2 = V , then
U1 = 0 or U2 = 0.
(i)
modules. This is defined in the obvious way. For modules (Vi , φj )j , 1 ≤ i ≤ t we
write
t
M
(1) (t) (i)
(V1 , φj )j ⊕ · · · ⊕ (Vt , φj )j = (Vi , φj )j .
i=1
2.8. Products of modules. Let I be a set, and for each i ∈ I let Vi be a vector
space. The product of the vector spaces Vi is by definition the set of all sequences
(vi )i∈I with vi ∈ Vi . We denote the product by
Y
Vi .
i∈I
With componentwise addition and scalar multiplication, this is again a vector space.
The Vi are called the factors of the product. For linear maps fi : Vi → Wi with i ∈ I
we define their product
Y Y Y
fi : Vi → Wi
i∈I i∈I i∈I
Q L Q
by i∈I fi ((vQ
i )i ) = (fi (v
Li ))i . Obviously, i∈I Vi is a subspace of i∈I Vi . If I is a
finite set, then i∈I Vi = i∈I Vi .
(i)
Now for each i ∈ I let Vi = (Vi , φj )j be a J-module. Then the product of the
modules Vi is defined as
!
Y Y (i)
Y Y (i)
(V, φj )j = Vi = (Vi , φj )j = Vi , φj .
i∈I i∈I i∈I i∈I j
Thus V is the product of the vector spaces Vi , and φj is the product of the linear
(i)
maps φj .
2.9. Examples: Nilpotent endomorphisms. Sometimes one does not study all
J-modules, but one assumes that the linear maps associated to the elements in J
satisfy certain relations. For example, if J just contains one element, we could
study all J-modules (V, f ) such that f n = 0 for some fixed n. Or, if J contains two
elements, then we can study all modules (V, f, g) such that f g = gf .
Assume |J| = 1. Thus a J-module is just (V, φ) with V a vector space and φ : V → V
a linear map. We additionally assume that φ is nilpotent, i.e. φm = 0 for some m
and that V is finite-dimensional. We denote this class of modules by N f.d..
We know from LA that there exists a basis v1 , . . . , vn of V such that the correspond-
ing matrix Φ of φ is of the form
J(λ )
1
J(λ2 )
Φ= ..
.
J(λt )
20 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
denote a basis of
K 10 = K 4 ⊕ K 2 ⊕ K 2 ⊕ K 1 ⊕ K 1 .
Let φλ : K 10 → K 10 be the linear map defined by φλ (eij ) = eij−1 for 2 ≤ j ≤ λi and
φλ (ei1 ) = 0. Thus N(λ) = (K 10 , φλ ).
So φλ operates on the basis vectors displayed in the boxes of the Young diagram
by mapping them to the vector in the box below if there is a box below, and by
mapping them to 0 if there is no box below.
The matrix of φλ with respect to the basis
e11 , e12 , e13 , e14 , e21 , e22 , e31 , e32 , e41 , e51
is 0 1
J(4) 0 1
0 1
J(2) 0
J(2) =
0 1
0
.
J(1) 0 1
J(1) 0
0
0
for some cij ∈ K. We want to compute the submodule U(x) ⊆ N(λ) generated by
x:
We get
!
X X X
φ(x) = φ cij eij = cij φ(eij ) = cij eij−1 .
i,j i,j i,j:j≥2
Similarly, we can easily write down φ2 (x), φ3 (x), etc. Now let r be maximal such
that cir 6= 0 for some i. This implies φr−1 (x) 6= 0 but φr (x) = 0. It follows that the
vectors x, φ(x), . . . , φr−1 (x) generate U(x) as a vector space, and we see that U(x)
is isomorphic to N(r).
22 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
For example, the submodule U(eij ) of N(λ) is isomorphic to N(j) and the corre-
sponding factor module N(λ)/U(eij ) is isomorphic to
M
N(λi − j) ⊕ N(λa ).
a6=i
e13
e12
e11 e21
We get
U(e21 ) ∼
= N(1), N(3, 1)/U(e21 ) ∼
= N(3),
U(e11 ) ∼
= N(1), N(3, 1)/U(e11 ) ∼
= N(2, 1),
U(e12 ) ∼
= N(2), N(3, 1)/U(e12 ) ∼
= N(1, 1),
U(e13 ) ∼
= N(3), N(3, 1)/U(e13 ) ∼
= N(1),
U(e12 + e21 ) ∼
= N(2), N(3, 1)/U(e12 + e21 ) ∼
= N(2).
Let us check the last of these isomorphisms: Let x = e12 + e21 ∈ N(3, 1) = (K 4 , φ).
We get φ(x) = e11 and φ2 (x) = 0. It follows that U(x) is isomorphic to N(2).
Now as a vector space, N(3, 1)/U(x) is generated by the residue classes e13 and
e12 . We have φ(e13 ) = e12 and φ(e12 ) = e11 . In particular, φ(e12 ) ∈ U(x). Thus
N(3, 1)/U(x) ∼= N(2).
and !
[ [
(W ∩ Ui ) = W ∩ Ui .
i∈I i∈I
Of course it is better if you find examples which are independent of the field, if this
is possible.
Let K = Fp with p a prime number, and let λ and µ be partitions. How many
= N(λ) and V /U ∼
submodules U of V with U ∼ = N(µ) are there?
5: Let U be a maximal submodule of a module V , and let W be an arbitrary
submodule of V . Show that either W ⊆ U or U + W = V .
————————————————————————————-
3.1. Homomorphisms. Let (V, φj )j and (W, ψj )j be two modules. A linear map
f : V → W is a homomorphism (or module homomorphism) if
f φj = ψj f
for all j ∈ J.
f
V / W
φj ψj
f
V / W
One can easily check that Ker(f ) is a submodule of V : For v ∈ Ker(f ) and j ∈ J
we have f φj (v) = ψj f (v) = ψj (0) = 0.
Similarly, Im(f ) is a submodule of W : For v ∈ V and j ∈ J we have ψj f (v) =
f φj (v), thus ψj f (v) is in Im(f ).
For a homomorphism f : V → W let f1 : V → Im(f ) defined by f1 (v) = f (v) (the
only difference between f and f1 is that we changed the target module of f from
W to Im(f )), and let f2 : Im(f ) → W be the canonical inclusion. Then f1 is an
epimorphism and f2 a monomorphism, and we get f = f2 f1 . In other words, every
homomorphism is the composition of an epimorphism followed by a monomorphism.
3.2. Definition of a ring. A ring is a set R together with two maps + : R×R → R,
(a, b) 7→ a + b (the addition) and · : R × R → R, (a, b) 7→ ab (the multiplication)
such that the following hold:
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 25
In contrast to the definition of a field, the definitions of a ring and an algebra do not
require that the element 0 is different from the element 1. Thus there is a ring and
an algebra which contains just one element, namely 0 = 1. If 0 = 1, then R = {0}.
Proof. One easily shows that f is well defined, and that it is a homomorphism.
Obviously f is injective and surjective, and thus an isomorphism.
26 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Remark: The above result is very easy to prove. Nevertheless we call it a Theorem,
because of its importance.
Proof. Note that U2 /U1 is a submodule of V /U1 . Thus we can build the factor
module (V /U1 )/(U2 /U1 ). Let
V → V /U1 → (V /U1 )/(U2 /U1 )
be the composition of the two canonical projections. This homomorphism is obvi-
ously surjective and its kernel is U2 . Now we use Theorem 3.1.
Corollary 3.3 (Second Isomorphism Theorem). If U1 and U2 are submodules of a
module V , then
U1 /(U1 ∩ U2 ) ∼
= (U1 + U2 )/U2 .
Proof. Let
U1 → U1 + U2 → (U1 + U2 )/U2
be the composition of the inclusion U1 → U1 + U2 and the projection U1 + U2 →
(U1 + U2 )/U2 . This homomorphism is surjective (If u1 ∈ U1 and u2 ∈ U2 , then
u1 +u2 +U2 = u1 +U2 is the image of u1 .) and its kernel is U1 ∩U2 (An element u1 ∈ U1
is mapped to 0 if and only if u1 + U2 = U2 , thus if and only if u1 ∈ U1 ∩ U2 .).
U1 + UH2
vv HH
vv HH
v HH
vv HH
vv
U1 HH U2
HH vv
HH vvv
HH vv
H vv
U1 ∩ U2
U2
U1 + U2 ∩OW U2 + WL
rr OOO o LLL
rrr OOO ooooo LLL
rrr OOO oo LLL
rrr
O ooo L
U1 LLL U2 ∩ W U1 + WO r
U2
LLL ooo OOO
OOO rrr
LLL
ooooo OOO rrrr
LL ooo O rrr
U1 ∩ W U1 + U2 ∩ W
Now let V and W be modules, which are a finite direct sum of certain submodules,
say
M n Mm
V = Vj and W = Wi .
j=1 i=1
If f : V → W is a homomorphism, define
fij = πW,i ◦ f ◦ ιV,j : Vj → Wi
We can write f : V → W in matrix form
f11 · · · f1n
f= .. ..
. .
fm1 · · · fmn
28 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
and we can use the usual matrix calculus: Let us write elements v ∈ V and w ∈ W
as columns
v1 w1
v= . .
. and w = ...
vn wm
with vj ∈ Vj and wi ∈ Wi . If f (v) = w we claim that
Pn
f11 · · · f1n v1 j=1 f1j (vj ) w1
.
.. . .
.. .. = . .
.. = .. .
Pn
fm1 · · · fmn vn j=1 fmj (vj ) wm
Namely, if v ∈ V we get for 1 ≤ i ≤ m
n
X n
X
fij (vj ) = (πW,i ◦ f ◦ ιV,j ) (vj )
j=1 j=1
n
!!
X
= πW,i ◦ f ◦ ιV,j ◦ πV,j (v)
j=1
= (πW,i ◦ f )(v) = wi .
The first term is the matrix product of the ith row of the matrix of f with the
column vector v, the last term is the ith entry in the column vector w.
and !
n
M n
M
HomJ Vj , X → HomJ (Vj , X).
j=1 j=1
The endomorphism ring End(V ) of a module V contains of course always the idem-
potents 0 and 1. Here 0 corresponds to the direct decomposition V = 0 ⊕ V , and 1
corresponds to V = V ⊕ 0.
If e is an idempotent in a ring, then 1 − e is also an idempotent. (Namely (1 − e)2 =
1 − e − e + e2 = 1 − e.)
30 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
• V is indecomposable;
• V =
6 0, and 0 and 1 are the only idempotents in End(V ).
Later we will study in more detail the relationship between idempotents in endo-
morphism rings and direct decompositions.
Proof. Assume first that f is a split monomorphism. Thus W = Im(f ) ⊕ C for some
submodule C of W . Let ι : Im(f ) → W be the inclusion homomorphism, and let
π : W → Im(f ) be the projection with kernel C. Let f0 : V → Im(f ) be defined by
f0 (v) = f (v) for all v ∈ V . Thus f = ιf0 . Of course, f0 is an isomorphism. Define
g = f0−1 π : W → V . Then we get
gf = (f0−1 π)(ιf0 ) = f0−1 (πι)f0 = f0−1 f0 = 1V .
It is easy to check that HomJ (f, h) is a linear map of vector spaces: For g, g1, g2 ∈
HomJ (W, X) and c ∈ K we have
h(g1 + g2 )f = hg1 f + hg2 f and h(cg)f = c(hgf ).
If V = W and f = 1V , then instead of HomJ (1V , h) we mostly write
HomJ (V, h) : HomJ (V, X) → HomJ (V, Y ),
thus by definition HomJ (V, h)(g) = hg for g ∈ HomJ (V, X). If X = Y and h = 1X ,
then instead of HomJ (f, 1X ) we write
HomJ (f, X) : HomJ (W, X) → HomJ (V, X),
thus HomJ (f, X)(g) = gf for g ∈ HomJ (W, X).
Let U, V, W be modules, and let f : U → V and g : V → W be homomorphisms. If
Im(f ) = Ker(g), then (f, g) is called an exact sequence. Mostly we denote such
an exact sequence in the form
f g
U−→V − → W.
We also say, the sequence is exact at V . Given such a sequence with U = 0,
exactness implies that g is injective. (For U = 0 we have Im(f ) = 0 = Ker(g), thus
g is injective.) Similarly, if W = 0, exactness yields that f is surjective. (For W = 0
we have Ker(g) = V , but Im(f ) = V means that f is surjective.)
Given modules Vi with 0 ≤ i ≤ t and homomorphisms fi : Vi−1 → Vi with 1 ≤ i ≤ t,
then the sequence (f1 , . . . , ft ) is an exact sequence if
Im(fi−1 ) = Ker(fi )
for all 2 ≤ i ≤ t. Also here we often write
f1 f2 ft
V0 −
→ V1 −
→ ··· −
→ Vt .
X
b′
~
~ b
~
~~ f g
0 / U / V / W
f g
U / V /W / 0
}
c′ }
c }
~}
Y
Proof. Exercise.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 33
Proof. Exercise.
Remark: The expression commutative diagram means the following: Given are
certain modules and between them certain homomorphisms. One assumes that for
any pair of paths which start at the same module and also end at the same module,
the compositions of the corresponding homomorphisms coincide. It is enough to
check that for the smallest subdiagrams. For example, in the diagram appearing in
the next lemma, commutativity means that bf = f ′ a and cg = g ′b. (And therefore
also cgf = g ′ f ′ a.) In the above diagram, commutativity just means hf = f ′ and
g = g ′ h. We used the homomorphisms 1U and 1W to obtain a nicer looking diagram.
Arranging such diagrams in square form has the advantage that we can speak about
rows and columns of a diagram. A frequent extra assumption is that certain columns
or rows are exact. In this lecture course, we will see many more commutative
diagrams.
Lemma 3.11. Let
f g
0 / U / V / W / 0
a b c
f′ g′
0 / U′ / V′ / W′ / 0
34 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Proof. First, we show that b is injective: If b(v) = 0 for some v ∈ V , then cg(v) =
g ′b(v) = 0. This implies g(v) = 0 since c is an isomorphism. Thus v belongs to
Ker(g) = Im(f ). So v = f (u) for some u ∈ U. We get f ′ a(u) = bf (u) = b(v) = 0.
Now f ′ a is injective, which implies u = 0 and therefore v = f (u) = 0.
Second, we prove that b is surjective: Let v ′ ∈ V ′ . Then c−1 g ′ (v ′ ) ∈ W . Since g
is surjective, there is some v ∈ V with g(v) = c−1 g ′ (v ′ ). Thus cg(v) = g ′ (v ′). This
implies
g ′ (v ′ − b(v)) = g ′(v ′ ) − g ′ b(v) = g ′ (v ′ ) − cg(v) = 0.
So v ′ − b(v) belongs to Ker(g ′ ) = Im(f ′ ). Therefore there exists some u′ ∈ U ′ with
f ′ (u′) = v ′ − b(v). Let u = a−1 (u′ ). Because f ′ (u′ ) = f ′ a(u) = bf (u), we get
v ′ = f ′ (u′ ) + b(v) = b(f (u) + v). Thus v ′ is in the image of b. So we proved that b
is an isomorphism.
The method used in the proof of the above lemma is called “Diagram chasing”.
Lemma 3.11 shows that equivalence of short exact sequences is indeed an equivalence
relation on the set of all short exact sequences starting in a fixed module U and
ending in a fixed module W :
Given two short exact exact sequences (f, g) and (f ′ , g ′) like in the assumption of
Lemma 3.11. If there exists a homomorphism h : V → V ′ such that hf = f ′ and
g = g ′ h, then h−1 satisfies h−1 f ′ = f and g ′ = gh−1 . This proves the symmetry of
the relation.
If there is another short exact sequence (f ′′ , g ′′ ) with f ′′ : U → V ′′ and g ′′ : V ′′ →
W and a homomorphism h′ : V ′ → V ′′ such that h′ f ′ = f ′′ and g ′ = g ′′ h′ , then
h′ h : V → V ′′ is a homomorphism with h′ hf = f ′′ and g = g ′′ h′ h. This shows our
relation is transitive.
Finally, (f, g) is equivalent to itself, just take h = 1V . Thus the relation is reflexive.
is a commutative diagram: If we write ι1 = t [1, 0] and π2 = [0, 1], then we see that
t ′
[f , g]f = t [1, 0] = ι1 and g = [0, 1] ◦ t [f ′ , g] = π2 ◦ t [f ′ , g]. Thus (f, g) is equivalent
to (ι1 , π2 ).
Vice versa, assume that (f, g) and (ι1 , π2 ) are equivalent. Thus there exists some
h : V → U ⊕W such that hf = ι1 and g = π2 h. Let π1 be the projection from U ⊕W
onto U with kernel W . Then π1 hf = π1 ι1 = 1U . Thus f is a split monomorphism.
3: Let
0 1 2
V = (K , ).
0 0
a b
Show: End(V ) is the set of matrices of the form with a, b ∈ K.
0 a
4: Let
0 0 0 0 0 0
V = (K 3 , 1 0 0 , 0 0 0).
0 0 0 1 0 0
with a, b, c ∈ K.
Use this to show that V is indecomposable.
6: Let
0 1 0
V = (K 3 , 0 0 0).
0 0 0
8: Let
f1 f2 f3 f4
V1 / V2 / V3 / V4 / V5
a1 a2 a3 a4 a5
g1 g2 g3 g4
W1 / W2 / W3 / W4 / W5
be a commutative diagramm of J-modules with exact rows.
Show: If a1 is an epimorphism, and if a2 and a4 are monomorphisms, then a3 is a
monomorphism.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 37
————————————————————————————-
4. Digression: Categories
Remark: Note that we speak of a “class” of objects, and not of sets of objects, since
we want to avoid set theoretic difficulties: For example the J-modules do not form
a set, otherwise we would run into contradictions. (Like: “The set of all sets.”)
If C ′ and C are categories with Ob(C ′ ) ⊆ Ob(C) and C ′ (X, Y ) ⊆ C(X, Y ) for all
objects X, Y ∈ C ′ such that the compositions of morphisms in C ′ coincide with the
compositions in C, then C ′ is called a subcategory of C. In case C ′ (X, Y ) = C(X, Y )
for all X, Y ∈ C ′ , one calls C ′ a full subcategory of C.
We only look at K-linear categories: We assume additionally that the morphism
sets C(X, Y ) are K-vector spaces, and that the composition maps
C(Y, Z) × C(X, Y ) → C(X, Z)
are K-bilinear. In K-linear categories we often write Hom(X, Y ) instead of C(X, Y ).
By Mod(K) we denote the category of K-vector spaces. Let mod(K) be the category
of finite-dimensional K-vector spaces.
Thus if we deal with contravariant functors, the order of the composition of mor-
phisms is reversed.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 39
————————————————————————————-
5.1. The module N(∞). Let V be a K-vector space with basis {ei | i ≥ 1}. Define
a K-linear endomorphism φ : V → V by φ(e1 ) = 0 and φ(ei ) = ei−1 for all i ≥ 2.
The following is clear and will be used in the proof of the next lemma: Every
submodule of a J-module is a sum of cyclic modules.
Lemma 5.1. The following hold:
Proof. First we determine the cyclic submodules: Let x ∈ V . Thus there exists
some n such that x ∈ N(n) and
Xn
x= ai ei .
i=1
5.2. Polynomial rings. This section is devoted to study some interesting and im-
portant examples of modules arising from the polynomial ring K[T ] in one variable
T.
Let
T · : K[T ] → K[T ]
be the K-linear map which maps a polynomial f to T f . In particular, T i is mapped
to T i+1 .
Another important K-linear map is
d
: K[T ] → K[T ]
dT
P
which maps a polynomial m i
i=0 ai T to its derivative
X m
d
(f ) = ai iT i−1 .
dT i=1
where c is a constant (degree 0) polynomial, and the pi are monic irreducible poly-
nomials. The polynomials pi and c are uniquely determined up to reordering.
d
5.3. The module (K[T ], dT ). We want to study the 1-module
d
V := (K[T ], ).
dT
The 1-module M
(W, φ) := N(p)
i∈N0
has as a basis {eij | i ∈ N0 , 1 ≤ j ≤ p} where
(
0 if j = 1,
φ(eij ) =
ei,j−1 otherwise.
Define a K-linear map
f : W → K[T ]
by
1
eij 7→ T ip+j−1.
(j − 1)!
Since j ≤ p we know that p does not divide (j − 1)!, thus (j − 1)! 6= 0 in K. One
easily checks that f defines a vector space isomorphism.
Exercise: Prove that
d
f (φ(eij )) = (f (eij ))
dT
and determine f −1 .
We get that f is an isomorphism of 1-modules.
5.4. The module (K[T ], T ·). Next, we want to study the 1-module
V := (K[T ], T ·).
Pn
Let a = i=0 ai T i be a polynomial in K[T ]. The submodule U(a) of V generated
by a is
(a) := U(a) = {ab | b ∈ K[T ]}.
We call (a) the principal ideal generated by a.
Proposition 5.3. All ideals in the ring K[T ] are principal ideals.
In other words: Each submodule of V is of the form (a) for some a ∈ K[T ].
Now it is easy to check that (a) = (b) if and only if a|b and b|a if and only if there
exists some c ∈ K ∗ with b = ca. (For polynomials p and q we write p|q if q = pf for
some f ∈ K[T ].)
It follows that for submodules (a) and (b) of V we have
(a) ∩ (b) = l.c.m.(a, b)
and
(a) + (b) = g.c.d.(a, b).
Here l.c.m.(a, b) denotes the lowest common multiple, and g.c.d.(a, b) is the greatest
common divisor.
44 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Let P be the set of monic irreducible polynomials in K[T ]. Recall that every poly-
nomial p 6= 0 in K[T ] can be written as
p = cpe11 pe22 · · · pet t
where c ∈ K ∗ , ei ≥ 1 and the pi are pairwise different polynomials in P. Further-
more, c, the ei and the pi are uniquely determined (up to reordering).
If b|a then there is an epimorphism
K[T ]/(a) → K[T ]/(b)
defined by p + (a) 7→ p + (b).
defined by π(a) = (π1 (a), . . . , πt (a)). Clearly, a ∈ Ker(π) if and only if πi (a) = 0 for
all i if and only if pei i |a for all i if and only if p|a. This implies that π is injective.
For a polynomial a of degree n we have dim K[T ]/(a) = n, and the residue classes
of 1, T, . . . , T n−1 form a basis of K[T ]/(a).
In particular, dim K[T ]/(pei i ) = deg(pei i ) and
t
Y
dim K[T ]/(pei i ) = deg(p).
i=1
Special case: The polynomial T is an irreducible polynomial in K[T ], and one easily
checks that the 1-modules (K[T ]/(T e ), T ·) and N(e) are isomorphic.
has as a basis the canonical basis vectors e1 , . . . , en . We have Φ(e1 ) = ce1 and
Φ(ei ) = cei + ei−1 if i ≥ 2. Then
f : (T − c)i 7→ en−i
for i ≥ 0 yields an isomorphism of 1-modules: One easily checks that
f (T · (T − c)i ) = Φ(f ((T − c)i ))
for all i ≥ 0.
Conclusion: If we can determine the set P of irreducible polynomials in K[T ], then
one has quite a good description of the submodules and also the factor modules of
(K[T ], T ·). But of course, describing P is very hard (or impossible) if the field K is
too complicated.
5.5. The module (K(T ), T ·). Let K(T ) be the ring of rational functions in one
variable T . The elements of K(T ) are of the form pq where p and q are polynomials
′
in K[T ] wit q 6= 0. Furthermore, pq = pq′ if and only if pq ′ = qp′ . Copying the
ususal rules for adding and multiplying fractions, K(T ) becomes a ring (it is even
a K-algebra). Clearly, all non-zero elements in K(T ) have an inverse, thus K(T ) is
also a field. It contains K[T ] as a subring, the embedding given by p 7→ 1p .
5.6. Exercises. 1: Show: The module K[T, T −1 ]/K[T ] is isomorphic to N(∞). Its
basis are the residue classes of T −1 , T −2, . . ..
d
Show that the 1-module (K[T ], dT ) is indecomposable if char(K) = 0.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 47
d
Write (K[T ], dT ) as a direct sum of indecomposable modules if char(K) > 0.
4: Let K be a field, and let P be the set of monic irreducible polynomials in K[T ].
For a ∈ K(T ) set
K[T ]a = {f a | f ∈ K[T ]} ⊂ K(T ).
For every p ∈ P let K[T, p−1 ] be the subalgebra of K(T ) generated by T and p−1 .
In other words,
q
K[T, p−1 ] = { | q ∈ K[T ], n ∈ N0 } ⊂ K(T ).
pn
a: Show: The modules K[T ]p−n /K[T ] and K[T ]/(pn ) are isomorphic. Use this to
determine the submodules of K[T ]p−n /K[T ].
b: If U is a proper submodule of K[T ]p−n /K[T ], then U = K[T ]p−n /K[T ] for some
n ∈ N0 .
c: We have
X
K(T ) = K[T, p−1 ].
p∈P
Let
ιp : K[T, p−1 ]/K[T ] → K(T )/K[T ]
be the inclusion.
Show: The homomorphism
M X
ι= ιp : (K[T, p−1 ]/K[T ]) → K(T )/K[T ]
p∈P p∈P
is an isomorphism.
d: Determine the submodules of K(T )/K[T ].
————————————————————————————-
48 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Some topics discussed in this section are also known as “Artin-Wedderburn Theory”.
Just open any book on Algebra.
(i) V is semisimple;
(ii) V is a sum of simple modules;
(iii) Every submodule of V is a direct summand.
This is not surprising: The implication (ii) =⇒ (i) yields the existence of a basis
of a vector space. (We just look at the special case J = ∅. Then J-modules are just
vector spaces. The simple J-modules are one-dimensional, and every vector space
is a sum of its one-dimensional subspaces, thus condition (ii) holds.)
Proof of Theorem 6.1. The implication (i) =⇒ (ii) is obvious. Let us show (ii) =⇒
(iii): Let V be a sum of simple submodules, and let U be a submodule of V . Let W
be the set of submodules W of V with U ∩ W = 0. Together with the inclusion ⊆,
the set W is a partially ordered set. Since 0 ∈ W, we know that W is non-empty.
If W ′ ⊆ W is a chain, then
[
W′ = W
W ∈W ′
belongs to W: If x ∈ U ∩ W ′, then x belongs to some W in W ′ , and therefore
x ∈ U ∩ W = 0.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 49
Now Zorn’s Lemma 6.3 says that W contains a maximal element. So let W ∈ W be
maximal. We know that U ∩ W = 0. On the other hand, we show that U + W = V :
Since V is a sum of simple submodules, it is enough to show that each simple
submodule of V is containd in U + W . Let S be a simple submodule of V . If we
assume that S is not contained in U + W , then (U + W ) ∩ S is a proper submodule
of S. Since S is simple, we get (U + W ) ∩ S = 0, and therefore U ∩ (W + S) = 0:
If u = w + s with u ∈ U, w ∈ W and s ∈ S, then u − w = s ∈ (U + W ) ∩ S = 0.
Thus s = 0 and u = w ∈ U ∩ W = 0.
This implies that W + S belongs to W. The maximality of W in W yields that
W = W +S and therefore we get S ⊆ W , which is a contradiction to our assumption
S 6⊆ U + W . Thus we see that U + W = V . So W is a direct complement of U in
V.
(iii) =⇒ (ii): Let S be the set of submodules of V , which are a sum of simple
submodules of V . We have 0 ∈ S. (We can think of 0 as the sum over an empty set
of simple submodules of V .)
Together with the inclusion ⊆, the set S forms a partially ordered set. Since 0
belongs to S, we know that S is non-empty.
If S ′ is a chain in S, then
[
U
U ∈S ′
W + C ′ = W + (C ∩ W ′ ) = (W + C) ∩ W ′ = V ∩ W ′ = W ′ .
This implies
W ′ /W = (W + C ′ )/W ∼
= C ′ /(W ∩ C ′ ) = C ′ .
50 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Therefore C ′ is simple.
V A
{ AA
{{ AA
{{{ A AA
{{
W′C C
{ CC ~
{{ CC ~~
{{{ CC ~~~
{{ C ~~
WC C′
C CC {{
CC {{{
CC {
{{
0
Because U is a sum of simple modules, we get that U +C ′ is a sum of simple modules,
thus it belongs to S. Now U ⊂ U + C ′ yields a contradiction to the maximality of
U in S.
(ii) =⇒ (i): We show the following stronger statement:
LemmaP6.4. Let V be a module, and let U be the set of simple submodules
L U of V .
If V = U ∈U U, then there exists a subset U ′ ⊆ U such that V = U ∈U ′ U.
P
Proof. A subset U ′ of U is called independent, if the sum U ∈U ′ U is a direct sum.
Let T be the set of independent subsets of U, together with the inclusion of sets ⊆
this is a partially ordered set. Since the empty set belongs to T we know that T is
non-empty. If T ′ is a chain in T , then
[
U′
U ′ ∈T ′
V @
~~ @ @@
~~~ @@
~~ @
U@ C
@ @@ ~~
@@ ~~~
@@ ~
~~
0
Let U be submodule of V . We check condition (iii) for U: Every submodule U ′ of
U is also a submodule of V . Thus there exists a direct complement C of U ′ in V .
Then C ∩ U is a direct complement of U ′ in U.
V GG
ww GG
www GG
ww GG
ww G
C FF U
x AA
FF xx AA
FF x
FF xxx AA
A
F xx
C ∩ UG U′
GG }
GG }}
GG
GG }}}
}}
0
Of course U ′ ∩ (C ∩ U) = 0, and the modularity yields U ′ + (C ∩ U) = (U ′ + C) ∩ U =
V ∩ U = U.
Let V be a semisimple module. For every simple module S let VS be the sum of
all submodules U of V such that U ∼= S. The submodule VS depends only on the
isomorphism class [S] of S. Thus we obtain a family (VS )[S] of submodules of V
which are indexed by the isomorphism classes of simple modules. The submodules
VS are called the isotypical components of V .
Proposition 6.6. Let V be a semisimple module. Then the following hold:
L
• V = [S] VS ;
• If V ′ is a submodule of V , then VS′ = V ′ ∩ VS ;
• If W is another semisimple module and f : V → W is a homomorphism,
then f (VS ) ⊆ WS .
Products of rings: Let I be an index set, and for each i ∈ I let Ri be a ring. By
Y
Ri
i∈I
we denote the product of the rings Ri . Its elements are the sequences (ri )i∈I
with ri ∈ Ri , and the addition and multiplication is defined componentwise, thus
(ri )i + (ri′ )i = (ri + ri′ )i and (ri )i · (ri′ )i = (ri ri′ )i .
The above isomorphism tells us, that to understand End(V ), we only have to un-
derstand the rings End(VS ). Thus assume V = VS . We have
M
V = S
i∈I
for some index set I. The structure of End(V ) only depends on the skew field
D = End(S) and the cardinality |I| of I.
If I is finite, then |I| = n and End(V ) is just the ring Mn (D) of n × n-matrices with
entries in D.
L L
Proof. Let ιj : W → i∈I W be the canonical inclusions, and let πj : i∈I W → W
be the canonical projections. We map
!
M
f ∈ End W
i∈I
54 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
6: Let V = (V, φj )j∈J be a finite-dimensional J-module such that all φj are diago-
nalizable.
Show: If [φi , φj ] = 0 for all i, j ∈ J and if V is simple, then V is 1-dimensional.
————————————————————————————-
Proof. Obvious.
Lemma 7.2. soc(soc(V )) = soc(V ).
Proof. Obvious.
Lemma 7.3. If f : V → W is a homomorphism, then f (soc(V )) ⊆ soc(W ).
L
Proof. Let V = i∈I Vi . Every submodule soc(Vi ) is semisimple, thus it is contained
in soc(V ). Vice versa, let U be a simple submodule of V , and let πi : V → Vi be
the canonical projections. ThenL πi (U) is either 0 or simple, thus it is contained
in soc(Vi ). This implies U ⊆ i∈I soc(V
L i ). The simple submodules of V generate
soc(V ), thus we also have soc(V ) ⊆ i∈I soc(Vi ).
For |J| = 1 the example V = (K[T ], T ·) shows that this is not the case: For every
irreducible polynomial p in K[T ], the module K[T ]/(p) is simple with basis the
residue classes of 1, T, T 2 , . . . , T m−1 where m is the degree of the polynomial p.
Now assume that W = K[T ]/U is a largest semisimple factor module of V . This
would imply U ⊆ (p) for every irreducible polynomial p. Since
\
(p) = 0,
p∈P
we get U = 0 and therefore W = K[T ]. Here P denotes the set of all irreducible
polynomials in K[T ]. But V is not at all semisimple. Namely V is indecomposable
and not simple. In fact, V does not contain any simple submodules.
Recall: A submodule U of a module V is called a maximal submodule if U ⊂ V
and if U ⊆ U ′ ⊂ V implies U = U ′ .
56 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Note that rad(V ) = V if V does not contain any maximal submodule. For example,
rad(N(∞)) = N(∞).
The factor module V / rad(V ) is called the top of V and is denoted by top(V ).
Lemma 7.6. Let V be a module. The radical of V is the intersection of all submod-
ules U of V such that V /U is semisimple.
Note that in general the module V / rad(V ) does not have to be semisimple: If
V = (K[T ], T ·), then from the above discussion we get V / rad(V ) = V and V is not
semisimple. However, if V is a “module of finite length”, then the factor module
V / rad(V ) is semisimple. This will be discussed in Part 2, see in particular Lemma
10.9.
Let us list some basic properties of the radical of a module:
Lemma 7.7. We have rad(V ) = 0 if and only if 0 can be written as an intersection
of maximal submodules of V .
Lemma 7.8. If U is a submodule of V with U ⊆ rad(V ), then rad(V /U) =
rad(V )/U. In particular, rad(V / rad(V )) = 0.
Proof. Exercise.
Lemma 7.9. If f : V → W is a homomorphism , then f (rad(V )) ⊆ rad(W ).
t
W LL
tt LLL
ttt LLL
LLL
t
tt
tt
U II f (V )
II s
II
II sssss
II s
sss
f (f −1 (U))
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 57
Thus
W/U ∼ = f (V )/f (f −1(U)) ∼
= V /f −1 (U)
is simple, and therefore f −1 (U) is a maximal submodule of V and contains rad(V ).
So we proved that f (rad(V )) ⊆ f (f −1 (U)) for all maximal submodules U of W .
Since rad(V ) ⊆ f −1 (U) for all such U, we get
f (rad(V )) ⊆ f f −1 (rad(W )) ⊆ rad(W ).
Lemma 7.10. If U is a submodule of V , then (U + rad(V ))/U ⊆ rad(V /U).
Proof. Exercise.
L
Proof. Let V = i∈I Vi , and let πi : V → Vi be the Lcanonical projections. We
have πi (rad(V )) ⊆ rad(Vi ), and therefore rad(V ) ⊆ i∈I rad(Vi ). Vice versa, let
U be a maximal submodule of V . Let Ui be the kernel of the composition Vi →
V → V /U of the obvious canonical homomorphisms. We get that either Ui is a
maximal submodule of Vi or Ui = Vi . In both cases we get rad(Vi ) ⊆ Ui . Thus
L
L i ) ⊆ U. Since rad(V ) is the intersection of all maximal submodules of V ,
i∈I rad(V
we get i∈I rad(Vi ) ⊆ rad(V ).
7.3. Large and small submodules. Let V be a module, and let U be a submodule
of V . The module U is called large in V if U ∩ U ′ 6= 0 for all non-zero submodules
U ′ of V . The module U is small in V if U + U ′ ⊂ V for all proper submodules U ′
of V .
Lemma 7.12. Let U1 and U2 be submodules of a module V . If U1 and U2 are large
in V , then U1 ∩ U2 is large in V . If U1 and U2 are small in V , then U1 + U2 is small
in V .
• U ∩ U ′ = 0;
• If U ′′ is a submodule with U ′ ⊂ U ′′ , then U ∩ U ′′ 6= 0.
Proof. Clearly, (i) implies (ii) and (iii). Let us prove (i): Assume V is a finitely
generated module, and let x1 , . . . , xn be a generating set of V . Furthermore, let W
be a proper submodule of V . We show that W + rad(V ) is a proper submodule: For
0 ≤ t ≤ n let Wt be the submodule of V which is generated by W and the elements
x1 , . . . , xt . Thus we obtain a chain of submodules
W = W0 ⊆ W1 ⊆ · · · ⊆ Wn = V.
Since W ⊂ V , there exists some t with Wt−1 ⊂ Wt = V . Let U be the (cyclic)
submodule generated by xt . We get
U + Wt−1 = Wt = V,
and U 6⊆ Wt−1 . By Corollary 7.16 this implies that there exists a maximal submodule
W ′ of V with Wt−1 ⊆ W ′ . Since W ′ is a maximal submodule of V , we get rad(V ) ⊆
W ′ . Thus
W + rad(V ) ⊆ W + W ′ = W ′ ⊂ V.
This shows that rad(V ) is small in V .
Note that a corresponding statement for the socle of a module is in general wrong:
For example, the 1-module V = (K[T ], T ·) is finitely generated, and we have
soc(V ) = 0. So the socle is not large in V in this case.
Corollary 7.18. Every proper submodule of a finitely generated module V is con-
tained in a maximal submodule of V .
Proof. Let U0 be the intersection of all large submodules of V . We want to show that
soc(V ) is contained in every large submodule of V . This implies then soc(V ) ⊆ U0 .
Let U be a large submodule of V . Assume soc(V ) is not contained in U. Then
U ∩ soc(V ) is a proper submodule of soc(V ). Since soc(V ) is generated by simple
submodules, there exists a simple submodule S of V which is not contained in U.
Now S is simple and therefore U ∩ S = 0. Since S 6= 0, this is a contradiction. This
implies soc(V ) ⊆ U0 .
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 61
3: For λ ∈ K let J(λ, n) be the Jordan block of size n × n with eigenvalue λ. For
λ1 6= λ2 in K, show that the 1-module (K n , J(λ1 , n)) ⊕ (K m , J(λ2 , m)) is cyclic.
4: Classify the small submodules of (K[T ], T ·) and N(∞).
***********************************************************************
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 63
8. Filtrations of modules
A filtration
0 = U0′ ⊆ U1′ ⊆ · · · ⊆ Ut′ = V
is a refinement of the filtration above if
Ui /Ui−1 ∼
= Vπ(i) /Vπ(i)−1
for 1 ≤ i ≤ s.
Theorem 8.1 (Schreier). Any two given filtrations of a module V have isomorphic
refinements.
(U1 + V2 ∩ U2 )/(U1 + V1 ∩ U2 ) ∼
= (U2 ∩ V2 )/((U1 ∩ V2 ) + (U2 ∩ V1 ))
∼
= (V1 + U2 ∩ V2 )/(V1 + U1 ∩ V2 ).
U1 + V2 ∩ UT2 V1 + U2 ∩ V2
TTTT jjj
TTT
TTTT jjjjjjj
TTT j
jjjj
U1 + V1 ∩ UT2 U2 ∩ V2 V1 + U1 ∩ V2
TTTT kkk
TTTT
TTTT kkkkkkk
k
TT kkkk
(U1 ∩ V2 ) + (U2 ∩ V1 )
64 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
◦A
} AA
}} AA
}}} AA
}} A
◦ ?? ◦@
~~ ?? @@
~~~ ?? @@
@@
~~ ??
U2 @ ◦ ?? V2
@@ ?? ~~
@@ ?? ~~
@@ ?? ~
~~
⋆ AA } ◦ AA } ⋆
AA }} AA }}
}A A}
}} AA }} AA
}} A }} A
⋆ ?? ⋆ ⋆@
~~ ?? @@
~~~ ?? @@
@@
~~ ??
U1 @ ⋆? V1
@@ ?? ~
@@ ??
?? ~~~
@@ ? ~
~~
◦A ◦
AA }}
AA }}
AA }}
A }}
◦
U/(U ∩ U ′ ) ∼
= (U + U ′ )/U ′ .
U + UH′
ww HH
ww HH
ww HH
w HH
ww
U GG v
U′
GG v
GG vv
GG vvv
G vv
U ∩ U′
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 65
A filtration
0 = U0 ⊆ U1 ⊆ · · · ⊆ Us = V.
of a module V with all factors Ui /Ui−1 (1 ≤ i ≤ s) being simple is called a com-
position series of V . In this case we call s (i.e. the number of simple factors) the
length of the composition series. We call the Ui /Ui−1 the composition factors of
V.
Proof. Let
0 = U0 ⊂ U1 ⊂ · · · ⊂ Ul = V
be a composition series, and let
0 = V0 ⊆ V1 ⊆ · · · ⊆ Vt = V
be a filtration. By Schreier’s Theorem 8.1 there exist isomorphic refinements of these
filtrations. Let Fi = Ui /Ui−1 be the factors of the filtration (Ui )i . Thus Fi is simple.
If (Ui′ )i is a refinement of (Ui )i , then its factors are F1 , . . . , Fl together with some
0-modules. The corresponding refinement of (Vj )j has exactly l + 1 submodules.
Thus (Vj )j has at most l different non-zero factors. In particular, if (Vj )j is already
a composition series, then t = l.
If V has a composition series of length l, then we say V has length l, and we write
l(V ) = l. Otherwise, V has infinite length and we write l(V ) = ∞.
Assume l(V ) < ∞ and let S be a simple module. Then [V : S] is the number
of composition factors in a (and thus in all) composition series of V which are
isomorphic to S. One calls [V : S] the Jordan-Hölder multiplicity of S in V .
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 67
Let l(V ) < ∞. Then ([V : S])S∈S is called the dimension vector of V , where S
is a complete set of representatives of isomorphism classes of the simple modules.
Note that only finitely many entries of the dimension vector are non-zero. We get
X
[V : S] = l(V ).
S∈S
————————————————————————————-
We need some basic notations from ring theory. This might seem a bit boring but
will be of great use later on.
68 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Proof. We have r ′ = 1r ′ = r ′′ rr ′ = r ′′ 1 = r ′′ .
Lemma 9.2. Assume that r is right-invertible. Then the following are equivalent:
• r is left-invertible;
• There exists only one right inverse of r.
• 1 6= 0;
• If r ∈ R, then r or 1 − r is invertible.
(Recall that the only ring with 1 = 0 is the 0-ring, which contains just one element.
Note that we do not exclude that for some elements r ∈ R both r and 1 − r are
invertible.)
Local rings occur in many different contexts. For example, they are important in
Algebraic Geometry: One studies the local ring associated to a point x of a curve
(or more generally of a variety or a scheme) and hopes to get a “local description”,
i.e. a description of the curve in a small neighbourhood of the point x.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 69
Examples: K[T ] is not local (T is not invertible, and 1 − T is also not invertible),
Z is not local, every field is a local ring.
Remark: Property (iv) implies that R contains exactly one maximal left ideal. Using
the Axiom of Choice, the converse is also true.
Proof. We first show that under the assumptions (i), (ii) and (iv) the elements 0
and 1 are the only idempotents in R, and therefore every left-invertible element and
every right-invertible element will be invertible.
70 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
I := {r ∈ R | r non-invertible }
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 71
Proof. Part (iii) of the above proposition tells us that I is indeed an ideal in R.
Assume now I ⊂ I ′ ⊆ R with I ′ a left (resp. right) ideal. Take r ∈ I ′ \ I. Then r is
invertible. Thus there exists some r ′ such that r ′ r = rr ′ = 1. This implies I ′ = R.
So we proved that I ′ is a maximal left and also a maximal right ideal. Furthermore,
the proof of (iii) =⇒ (iv) in the above proposition shows that I contains all proper
left ideals, and similarly one shows that I contains all proper right ideals of R.
If I is the Jacobson radical of a local ring R, then the radical factor ring R/I is
a ring without left ideals different from 0 and R/I. It is easy to check that R/I is
a skew field. (For r ∈ R/I with r 6= 0 and r ∈ R \ I there is some s ∈ R such that
sr = 1 = rs. In R/I we have s · r = sr = 1 = rs = r · s.)
Example: For c ∈ K set
R = {f /g | f, g ∈ K[T ], g(c) 6= 0},
m = {f /g ∈ R | f (c) = 0, g(c) 6= 0}.
Then m is an ideal in the ring R. In fact, m = (T − c)R: One inclusion is obtained
since
f (T − c)f
(T − c) = ,
g g
and the other inclusion follows since f (c) = 0 implies f = (T −c)h for some h ∈ K[T ]
and therefore
f h
= (T − c) .
g g
• R is not local;
• 0 and 1 are the only idempotents in R;
• The Jacobson radical of R is 0.
————————————————————————————-
Proof. Assume that V has length n. Thus every chain of submodules of V has length
at most n. In particular this is true for all chains of submodules of U. This implies
l(U) ≤ n. The same holds for chains of submodules which all contain U. Such
chains correspond under the projection homomorphism V → V /U to the chains of
submodules of V /U = W . Thus we get l(W ) ≤ n. So if V has finite length, then so
do U and W .
Vice versa, assume that U and W = V /U have finite length. Let
0 = U0 ⊂ U1 ⊂ · · · ⊂ Us = U and 0 = W0 ⊂ W1 ⊂ · · · ⊂ Wt = W
be composition series of U and W , respectively. We can write Wj in the form
Wj = Vj /U for some submodule U ⊆ Vj ⊆ V . We obtain a chain
0 = U0 ⊂ U1 ⊂ · · · ⊂ Us = U = V0 ⊂ V1 ⊂ · · · ⊂ Vt = V
of submodules of V such that
Vj /Vj−1 ∼
= Wj /Wj−1
for all 1 ≤ j ≤ t. This chain is a composition series of V , since the factors Ui /Ui−1
with 1 ≤ i ≤ s, and the factors Vj /Vj−1 with 1 ≤ j ≤ t are simple. If S is simple,
then the number of composition factors of V which are isomorphic to S is equal
to the number of indices i with Ui /Ui−1 ∼ = S plus the number of indices j with
Vj /Vj−1 ∼= S. In other words, [V : S] = [U : S] + [W : S].
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 73
Recall that for every homomorphism f : V → W there are short exact sequences
0 → Ker(f ) → V → Im(f ) → 0
and
0 → Im(f ) → W → Cok(f ) → 0.
74 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
0→U →V →W →0
(i) f is injective;
(ii) f is surjective;
(iii) f is an isomorphism;
(iv) l(Im(f )) = l(V ).
Recall that in Section 7.2 we studied the radical rad(V ) of a module V . The following
lemma shows that V / rad(V ) is well behaved if V is of finite length:
Proof. Assume that l(V / rad(V )) = n. Inductively T we look for maximal submodules
U1 , . . . , Un of V such that for 1 ≤ t ≤ n and Vt := ti=1 Ui we have
t
M
V /Vt ∼
= V /Ui
i=1
V FF
{{ FF
{{{ FF
FF
{ F
{{
Vt B Ut+1
B
BB x
BB xx
BB xxx
xx
Vt+1
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 75
Thus we obtain
V /Vt+1 = Vt /Vt+1 ⊕ Ut+1 /Vt+1
∼
= V /Ut+1 ⊕ V /Vt
t
M
∼
= V /Ut+1 ⊕ V /Ui .
i=1
The last of these isomorphisms comes from the induction assumption.
Proof. We have
0 = Ker(f 0 ) ⊆ Ker(f 1 ) ⊆ Ker(f 2 ) ⊆ · · · .
(For x ∈ Ker(f i ) we get f i (x) = 0 and therefore f i+1 (x) = 0.)
Assume that Ker(f i−1 ) = Ker(f i ) for some i. It follows that Ker(f i ) = Ker(f i+1 ).
(Assume f i+1 (x) = 0. Then f i (f (x)) = 0 and therefore f (x) ∈ Ker(f i ) = Ker(f i−1 ).
This implies f i(x) = f i−1 (f (x)) = 0. Thus Ker(f i+1 ) ⊆ Ker(f i).)
If
0 = Ker(f 0 ) ⊂ Ker(f 1 ) ⊂ · · · ⊂ Ker(f i ),
then l(Ker(f i )) ≥ i. This implies i ≤ n, and therefore Ker(f m ) = Ker(f n ) for all
m ≥ n.
We have
· · · ⊆ Im(f 2 ) ⊆ Im(f ) ⊆ Im(f 0 ) = V.
(For x ∈ Im(f i ) we get x = f i (y) = f i−1 (f (y)) for some y ∈ V . Thus x ∈ Im(f i−1 ).)
Assume that Im(f i−1 ) = Im(f i ). Then Im(f i ) = Im(f i+1 ). (For every y ∈ V
there exists some z with f i−1 (y) = f i (z). This implies f i(y) = f i+1 (z). Thus
Im(f i ) ⊆ Im(f i+1 ).)
If
Im(f i ) ⊂ · · · ⊂ Im(f 1 ) ⊂ Im(f 0 ) = V,
then l(Im(f i)) ≤ n − i, which implies i ≤ n. Thus Im(f m ) = Im(f n ) for all m ≥ n.
So we proved that
Ker(f n ) = Ker(f 2n ) and Im(f n ) = Im(f 2n ).
We claim that Im(f n ) ∩Ker(f n ) = 0: Let x ∈ Im(f n ) ∩Ker(f n ). Then x = f n (y) for
some y and also f n (x) = 0, which implies f 2n (y) = 0. Thus we get y ∈ Ker(f 2n ) =
Ker(f n ) and x = f n (y) = 0.
76 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Next, we show that Im(f n ) + Ker(f n ) = V : Let v ∈ V . Then there is some w with
f n (v) = f 2n (w). This is equivalent to f n (v −f n (w)) = 0. Thus v −f n (w) ∈ Ker(f n ).
Now v = f n (w) + (v − f n (w)). This finishes the proof.
Corollary 10.11. Let V be an indecomposable module of finite length n, and let
f ∈ End(V ). Then either f is an isomorphism, or f is nilpotent (i.e. f n = 0).
Combining the above with Lemma 9.3 we obtain the following important result:
Corollary 10.12. Let V be an indecomposable module of finite length. Then End(V )
is a local ring.
It follows that the endomorphism ring of any decomposable module contains idem-
potent which are not 0 or 1.
Example: The 1-module V = (K[T ], T ·) is indecomposable, but its endomorphism
ring End(V ) ∼
= K[T ] is not local.
Lemma 10.13. Let V be a module. If End(V ) is a local ring, then V is indecom-
posable.
Proof. If a ring R is local, then its only idempotents are 0 and 1. Then the result
follows from the discussion above.
Since Im(f ) and Im(hgf ) have the same length, we get Im(f ) ∩ Ker(hg) = 0. Now
Im(f ) has length n − a, and Ker(hg) has length l(Vm ) − l(Im(hg)). This implies
l(Im(hg)) = n − a, because
l(Im(hgf )) ≤ l(Im(hg)) ≤ l(Im(h)).
So we see that Im(f ) + Ker(hg) = Vm . In this way, we obtained a direct decompo-
sition
Vm = Im(f ) ⊕ Ker(hg).
But Vm is indecomposable, and Im(f ) 6= 0. It follows that Ker(hg) = 0. In other
words, hg is injective, and so g is injective.
In a similar way, we can show that g is also surjective: Namely
Vm+1 = Im(gf ) ⊕ Ker(h) :
Since Im(gf ) and Im(hgf ) have the same length, we get
Im(gf ) ∩ Ker(h) = 0.
On the other hand, the length of Ker(h) is
l(Vm+1 ) − l(Im(h)) = l(Vm+1 ) − (n − a).
Since Vm+1 is indecomposable, Im(gf ) 6= 0 implies Vm+1 = Im(gf ). Thus gf is
surjective, which yields that g is surjective as well.
Thus we have shown that g = fm is an isomorphism.
Corollary 10.15. If V is an indecomposable module of finite length n, and if I
denotes the radical of End(V ), then I n = 0.
Proof. LetPS be a subset of End(V ), and let SV be the set of all (finite) sums of
the form i fi (vi ) with fi ∈ S and vi ∈ V . This is a submodule of V . (It follows
from the definition, that SV is closed under addition. Since all fi are linear maps,
78 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
References
[ED] D. Eisenbud, J.A. de la Peña, Chains of maps between indecomposable modules, Journal für
die Reine und Angewandte Mathematik 504 (1998), 29–35.
[F] H. Fitting, Über direkte Produktzerlegungen einer Gruppe in unzerlegbare, Mathematische
Zeitschrift 39 (1934), 16–30.
[HS] M. Harada, Y. Sai, On categories of indecomposable modules I, Osaka Journal of Mathematics
7 (1970), 323–344.
[Ri] C.M. Ringel, Report on the Brauer-Thrall conjecture, In: Representation Theory II, Springer
Lecture Notes in Mathematics 832 (1980), 104–135.
10.4. Exercises. 1: Find the original references for Schreier’s Theorem, the Jordan-
Hölder Theorem, the Fitting Lemma and the Harada-Sai Lemma.
In particular: The endomorphism ring of N(∞) is a local ring with radical factor
ring isomorphic to the ground field K.
3: Let V = (K[T ], T ·). Show that V is indecomposable and that End(V ) is not a
local ring.
————————————————————————————-
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 79
If End(V ) is a local ring, then there exists some t with 1 ≤ t ≤ m and a direct
decomposition Wt = V ′ ⊕ Wt′ with V ′ ∼
= V and
M
X∼= Wt′ ⊕ Wj .
j6=t
This directL
decomposition of W shows that the cokernel of f is also isomorphic to
Z = Wt ⊕ j6=t Wj . This implies X ∼
′
= Z. In particular, we have Wt = V ′ ⊕ Wt′ ∼=
′
V ⊕ Wt .
such that Vi ∼
= Wπ(i) for all 2 ≤ i ≤ n. Now just set π(1) = t.
References
[K] W. Krull, Über verallgemeinerte endliche Abelsche Gruppen, Mathematische Zeitschrift 23
(1925), 161–196.
[R] R. Remak, Über die Zerlegung der endlichen Gruppen in direkte unzerlegbare Faktoren, Journal
für die Reine und Angewandte Mathematik (1911), 293–308.
[S1] O. Schmidt, Sur les produits directs, Bulletin de la S.M.F. 41 (1913), 161–164.
[S2] O. Schmidt, Über unendliche Gruppen mit endlicher Kette, Mathematische Zeitschrift 29
(1928), 34–41.
There are two cases: In the first case there is a direct decomposition W1 = U =
= V and X ∼
V ′ ⊕ U ′ with V ′ ∼ = U ′ ⊕ W2 . Since U ′ is isomorphic toLa direct summand
′ ′ ∼
of X, induction yields a subset I ⊆ {2, . . . , n} such that U = i∈I ′ Vi . Thus with
I := I ′ ∪ {1} we get M
U = V ′ ⊕ U′ ∼
= V1 ⊕ U ′ ∼ = Vi .
i∈I
11.3. Examples. We present some examples which show what happens if we work
with indecomposable direct summands, whose endomorphism ring is not local.
Assume |J| = 2, thus M = (K[T1 , T2 ], T1 ·, T2 ·) is a J-module. Let U1 and U2 be
non-zero submodules of M. We claim that U1 ∩ U2 6= 0: Let u1 ∈ U1 and u2 ∈ U2
be non-zero elements. Then we get u1 u2 ∈ U1 ∩ U2 , and we have u1 u2 6= 0.
In other words, the module M is uniform. (Recall that a module V is called uniform
if for all non-zero submodules U1 and U2 of V we have U1 ∩ U2 6= 0.) This implies
that every submodule of M is indecomposable.
where f = t [ι, −ι] and g = [ι, ι]. Here we denote all inclusion homomorphisms just
by ι.
Example 3: Here is another (trivial) example for the failure of the cancellation
rule: Let J = ∅, and let V be an infinite dimensional K-vector space. Then we have
V ⊕K ∼ =V ∼ = V ⊕ 0.
Thus we cannot cancel V . On the other hand, in contrast to Example 2, V is not
an indecomposable module.
(a): The direct summands of V ⊕ V are 0, V ⊕ V and all the submodules of the
form
Uf,g := {(hf, hg) | h ∈ K[T ]}
where f and g are polynomials with greatest common divisor 1.
(b): There exist direct summands U of V ⊕ V such that none of the modules 0,
V ⊕ V , U1,0 = V ⊕ 0 and U0,1 = 0 ⊕ V are a direct complement of U in V ⊕ V .
2: Let M1 , . . . , Mt be pairwise non-isomorphic modules of finite length, and let
mi ≥ 1 for 1 ≤ i ≤ t. Define
Mt
V = Mimi ,
i=1
84 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
***********************************************************************
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 85
means that η is compatible with the multiplication, and (R6 ) shows that the unit
element of A is mapped to the unit element of EndK (V ).
12.3. Modules and representations. Let Abb(V, V ) be the set of all (set theo-
retic) maps V → V . If we have any (set theoretic) map η : A → Abb(V, V ), then
we can define a map η : A × V → V by
η(a, v) = η(a)(v).
This defines a bijection between the set of all maps A → Abb(V, V ) and the set of
maps A × V → V .
Lemma 12.1. Let A be a K-algebra, and let V be a K-vector space. If η : A →
EndK (V ) is a map, then η is a representation of A if and only if η : A × V → V is
an A-module structure on V .
By Mod(A) we denote the K-linear category with all A-modules as objects, and
with A-module homomorphisms as morphisms. We call Mod(A) the category of
(left) A-modules. By mod(A) we denote the category of all finite-dimensional
A-modules. This is a full subcategory of Mod(A).
88 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
(Compare the definition of a free generating set with the definition of a basis of a
vector space, and with the definition of a linearly independent set of vectors.)
Lemma 12.2. An A-module is free if and only if it has a free generating set.
L
Proof. Let W be a direct sum of copies of A A, say W = i∈I Wi with Wi = A A
for all i ∈ I. By ei we denote the 1-element of Wi . (In coordinate notation: All
coefficients of ei are 0, except the ith coefficient is the element 1A ∈ A A = Wi .)
Thus the set {ei | i ∈ I} is a free generating set of W .
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 89
Vice versa, we want to show that every A-module V with a free generating set X
is isomorphic to a free module. We take a direct
L sum of copies of A A, which are
indexed by the elements in X. Thus W = x∈X Wx where Wx = A A for all x.
As before, let ex be the 1-element
P of Wx = A A. Then every element in W can be
written as a (finite) sum x∈X ax ex with ax ∈ A for all x ∈ X, and ax = 0 for
almost all (i.e. all but finitely many) x ∈ X. We define a map f : W → V by
!
X X
f ax ex = ax x.
x∈X x∈X
Inside the category of all |A|-modules, we can now characterize the A-modules as
follows: They are exactly the modules which are isomorphic to some factor module of
some free A-module. Thus up to isomorphism one obtains all A-modules by starting
with A A, taking direct sums of copies of A A and then taking all factor modules of
these direct sums.
a1 ⋆ a2 = a2 · a1 = a2 a1
for all a1 , a2 ∈ A (where · is the multiplication of A). This defines again an algebra.
Of course we have (Aop )op = A.
ρ : Aop → EndA (A A)
We know that
λa ρb = ρb λa
for all a, b ∈ A. (This follows directly from the associativity of the multiplication in
A.) In other words, the vector space endomorphisms ρb are endomorphisms of the
A-module A A, and ρ yields an embedding of Aop into EndA (A A).
It remains to show that every endomorphism f of A A is a right multiplication: Let
f (1) = b. We claim that f = ρb : For a ∈ A we have
ρ: V × A → V
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 91
12.8. Examples. For A-modules V and W the homomorphism space HomA (V, W )
carries a (left) EndA (W )-module structure defined by
EndA (W ) × HomA (V, W ) → HomA (V, W ), (f, g) 7→ f g,
and HomA (V, W ) has a right EndA (V )-module structure given by
HomA (V, W ) × EndA (V ) → HomA (V, W ), (g, f ) 7→ gf.
One can also turn A V into a module over EndA (V ) by
EndA (V ) × V → V, (f, v) 7→ f (v).
with ei ∈ Pi .
Lemma 12.6. For all i ∈ I we have Pi = Aei .
Proof. Only finitely many of the ei are different from 0. If ei = 0, then Pi = Aei = 0,
a contradiction to our assumption.
Proof. We have X
ej = ej · 1 = ej ei .
i∈I
As in the proof of Lemma 12.6, the unicity of the decomposition of an element in a
direct sum yields that ej = ej ej and ej ei = 0 for all i 6= j.
L
Warning: Given a direct decomposition A A = i∈I Pi . If we choose idempotents
ei ∈ Pi with Pi = Aei , then these idempotents do not have to be orthogonal to each
other. For example, let A = M2 (K) be the algebra of 2 × 2-matrices with entries in
K. Take
1 0 0 0
e1 = and e2 = ,
1 0 0 1
and define Pi = Aei . We obtain A A = P1 ⊕ P2 . The elements e1 and e2 are
idempotents, but they are not orthogonal.
Lemma 12.8 shows that any direct decomposition of A A yields a complete set of
orthogonal idempotents in A. Vice versa, assume that fi , i ∈ I is a complete set of
orthogonal idempotents in an algebra A, then
M
A A = Afi
is a direct decomposition of A A.
AA = Aei .
i=1
Note also that the modules Aei are isomorphic to each other: We get an isomorphism
Aei → Aej via right multiplication with eij .
Instead of working with this isomorphism, we could also argue like this: Let X = B n
be the vector space of n-tupels with coefficients in B. We interpret these n-tupels
as n × 1-matrices. So matrix multiplication yields an A-module structure on X. It
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 93
is clear that X and Aei have to be isomorphic: X and Aei only differ by the fact
that Aei contains some additional 0-columns.
Warning: The above direct decomposition of Mn (B) is for n ≥ 2 of course not the
only possible decomposition. For example for n = 2 and any x ∈ B the matrices
1 x 0 −x
and
0 0 0 1
form also a complete set of orthogonal idempotents in M2 (B). In this case
1 x 0 −x
M2 (B) = M2 (B) ⊕ M2 (B) ,
0 0 0 1
where
1 x
M2 (B)
0 0
consists of the matrices of the form
b1 b1 x
b2 b2 x
with b1 , b2 ∈ B, and
0 −x
M2 (B)
0 1
consists of the matrices whose only non-zero entries are in the second column.
Proof. Let f ∈ EndA (Ae), and let a = f (e) ∈ Ae. Then a = ae because a belongs
to Ae. Since f is a homomorphism, and e is an idempotent we have a = f (e) =
f (e2 ) = ef (e) = ea. Thus a = eae ∈ eAe. Clearly, the map defined by η(f ) = f (e)
is K-linear.
The map η is also surjective: For every a ∈ A let ρeae : Ae → Ae be the right
multiplication with eae defined by ρeae (be) = beae where b ∈ A. This map is
obviously an endomorphism of the A-module Ae, and we have η(ρeae ) = ρeae (e) =
eae.
Thus we have shown that η is bijective. In particular, the inverse η −1 (eae) is the
right multiplication with eae.
Lemma 12.10. If X is an A-module, then HomA (Ae, X) ∼
= eX as vector spaces.
Proof. Let η : HomA (Ae, X) → eX be the map defined by η(f ) = f (e). Since
f (e) = f (e2 ) = ef (e), we have f (e) ∈ eX, thus this is well defined. It is also clear
that η is K-linear.
If f1 , f2 ∈ HomA (Ae, X), then η(f1 ) = η(f2 ) implies f1 (e) = f2 (e), and therefore
f1 (ae) = af1 (e) = af2 (e) = f2 (ae)
for all a ∈ A. So f1 = f2 . This proves that η is injective.
Proof. We can identify HomA (Ae, M) with eM: We just map f : Ae → M to f (e).
Since e = e2 we get f (e) = f (e2 ) = ef (e), thus f (e) ∈ eM. Thus the homomor-
phisms f ∈ HomA (Ae, Ae′ ) correspond to the elements in eAe′ .
(i) =⇒ (ii): If Ae and Ae′ are isomorphic, there exist homomorphisms f : Ae → Ae′
and g : Ae′ → Ae such that gf = 1Ae . Set x = f (e) and y = g(e′ ). Thus x ∈ eAe′ ,
y ∈ e′ Ae, xy = e and yx = e′ .
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 95
(ii) =⇒ (i): Assume there exist elements x ∈ eAe′ and y ∈ e′ Ae with xy = e and
yx = e′ . Let f : Ae → Ae′ be the right multiplication with x, and let g : Ae′ → Ae be
the right multiplication with y. Then f and g are A-module homomorphisms, and we
have gf = 1Ae and f g = 1Ae′ . Thus the A-modules Ae and Ae′ are isomorphic.
The statement (ii) in the above lemma is left-right symmetric. Thus (i) and (ii) are
also equivalent to
Proof. Let Ae be a local module, and let M be the maximal submodule of Ae. For
every element x ∈ Ae we have x = xe, thus M = Me. We have eM = eAe ∩ M.
(Clearly, eM ⊆ eAe ∩ M. The other inclusion follows from the fact that e is an
idempotent: If a ∈ A and eae ∈ M, then eae = e(eae) ∈ eM.)
In particular we have eM = eMe ⊆ eAe. We have e ∈ eAe, but e does not belong
to M or eM. Thus eMe ⊂ eAe.
We claim that eMe is an ideal in eAe: Clearly, eAe · eMe ⊆ eMe. Since the right
multiplications with the elements from eAe are the endomorphisms of Ae, we have
Me · eAe ⊆ Me. (Me is the radical of the module Ae.) Thus eMe · eAe ⊆ eMe.
If x ∈ eAe \ eMe, then x ∈ Ae and x ∈ / M. (Note that exe = x.) Thus x generates
the local module Ae. It follows that there exists some y ∈ A with yx = e. Because
x = ex, we have
eye · x = eyx = e2 = e.
Thus x is left-invertible in eAe, and eye is right-invertible in eAe.
The element eye does not belong to eM, since eM is closed under right multiplication
with elements from eAe, and e ∈/ eM. So we get eye ∈ eAe \ eMe.
96 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Thus also the element eye has a left inverse in eAe. This proves that eye is invertible
in eAe. It follows that exe is invertible in eAe: Namely, we have
(eye)−1 · eye · x = (eye)−1 e.
Multiplying both sides of this equation from the right with eye yields x · eye = e.
We have shown that all elements in eAe \ eMe are invertible, thus eAe is a local
ring.
Vice versa, assume that eAe is a local ring. Then Ae is a non-zero cyclic module,
thus it has maximal submodules. Let M1 be a maximal submodule, and let M2 be
any proper submodule of Ae. Suppose M2 is not contained in M1 . This implies
Ae = M1 + M2 , thus e = x1 + x2 with xi ∈ Mi . We have e = ee = ex1 + ex2 .
Since eAe is a local ring, one of the elements exi = exi e, i = 1, 2 is invertible in
eAe. If e = yexi , then e belongs to Axi ⊆ Mi , thus Ae = Mi . By assumption both
modules M1 and M2 are proper submodules of Ae. This contradiction shows that
M1 contains all proper submodules of Ae, thus Ae is a local module.
12.12. Modules over products of algebras. Let R and S be rings. Recall that
the product R × S of R and S is again a ring with componentwise addition and
multiplication. Thus (r, s) + (r ′, s′ ) = (r + r ′ , s + s′ ) and (r, s) · (r ′ , s′ ) = (rr ′ , ss′ ).
Similarly, if A and B are algebras, then A×B is again an algebra. In this case, define
eA = (1, 0) and eB = (0, 1). These form a complete set of orthogonal idempotents.
We have (A × B)eA = A × 0 and (A × B)eB = 0 × B. These are ideals in A × B, and
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 97
The next result shows that bimodule structures allow us to see homomorphism
spaces again as modules.
Lemma 12.17. Let M be an A-B-bimodule, and let N an A-C-bimodule. Then
HomA (M, N) is an B op -C-bimodule via
b(c(f (m))) = c(f (bm))
for all b ∈ B, c ∈ C, f ∈ HomA (M, N) and m ∈ M.
Proof. Let ⋆ be the multiplication in B op , and set H := HomA (M, N). It is clear
that the two maps B op × H → H, (b, f ) 7→ (bf : m 7→ f (bm)) and C × H → H,
(c, f ) 7→ (cf : m 7→ cf (m)) are bilinear. We also have 1B · f = f and 1C · f = f for
all f ∈ H.
For b1 , b2 ∈ B and f ∈ H we have
((b1 ⋆ b2 )f )(m) = f ((b1 ⋆ b2 )m)
= f ((b2 b1 )m)
= f (b2 (b1 m))
= (b2 f )(b1 m)
= (b1 (b2 f ))(m).
This shows that
(b1 ⋆ b2 )f = b1 (b2 (f )).
Similarly,
((c1 c2 )f )(m) = (c1 c2 )(f (m))
= c1 (c2 (f (m)))
= c1 ((c2 f )(m))
= (c1 (c2 f ))(m).
shows that (c1 c2 )f = c1 (c2 f ) for all c1 , c2 ∈ C and f ∈ H.
————————————————————————————-
The opposite algebra Aop of a semisimple algebra A is again semisimple. This follows
directly from Condition (iii): If D is a skew field, then D op is also a skew field. For
an arbitrary ring R there is an isomorphism
Mn (R)op → Mn (Rop )
which maps a matrix Φ to its transpose t Φ.
Let A = Mn (D) for some K-skew field D and some n ∈ N1 . Let S = D n be the set
of column vectors of length n with entries in D. It is easy to show that S is a simple
Mn (D)-module. (One only has to show that if x 6= 0 is some non-zero column vector
in S, then Mn (D)x = S.) On the other hand, we can write A A as a direct sum of n
copies of S. Thus A A is semisimple.
Let A = Mn (D) for some K-skew field D and some n ∈ N1 . We have shown that A A
is a direct sum of n copies of the simple module S consisting of column vectors of
length n with entries in D. It follows that every A-module is a direct sum of copies
of S. (Each free module is a direct sum of copies of S, and each module is a factor
module of a free module. If a simple A-module T is isomorphic to a factor module
of a free A-module, we obtain a non-zero homomorphism S → T . Thus T ∼ = S by
Schur’s Lemma.) If
Ys
A= ∼ Mni (Di ),
i=1
then there are exactly s isomorphism classes of simple A-modules.
Proposition 13.2. Let K be an algebraically closed field. If A is a finite-dimensional
semisimple K-algebra, then
Y s
A∼ = Mni (K)
i=1
for some natural numbers ni , 1 ≤ i ≤ s.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 101
Proof. First, we look at the special case A = D, where D is a K-skew field: Let
d ∈ D. Since D is finite-dimensional, the powers di with i ∈ N0 cannot be linearly
independent. Thus there exists a non-zero polynomial p in K[T ] such that p(d) = 0.
We can assume that p is monic. Since K is algebraically closed, we can write it as
a product of linear factors, say p = (T − c1 ) · · · (T − cn ) with ci ∈ K. Thus in D we
have (d − c1 ) · · · (d − cn ) = 0. Since D has no zero divisors, we get d − ci = 0 for
some i, and therefore d = ci ∈ K.
Now we investigate the general case: We know that A is isomorphic to a product
of matrix rings of the form Mni (Di ) with K-skew fields Di and ni ∈ N1 . Since A
is finite-dimensional, every K-skew field Di must be finite-dimensional over K. But
since K is algebraically closed, and the Di are finite-dimensional K-skew fields we
get Di = K.
Proof. It is easy to show that the centre of a matrix ring Mn (K) is just the set of
scalar matrices. Thus we get
s
! s s
Y Y Y
C Mni (K) = C(Mni (K)) = ∼ K.
i=1 i=1 i=1
13.2. Examples: Group algebras. Let G be a group, and let K[G] be a K-vector
space with basis {eg | g ∈ G}. Define
eg eh := egh .
Extending this linearly turns the vector space K[G] into a K-algebra. One calls
K[G] the group algebra of G over K. Clearly, K[G] is finite-dimensional if and
only if G is a finite group.
A representation of G over K is a group homomorphism
ρ : G → GL(V )
where V is a K-vector space. In the obvious way one can define homomorphisms
of representations. It turns out that the category of representations of G over K is
equivalent to the category Mod(K[G]) of modules over the group algebra K[G]. If
V is a K[G]-module, then for g ∈ G and v ∈ V we often write gv instead of eg v.
The representation theory of G depends very much on the field K, in particular, the
characteristic of K plays an important role.
Theorem 13.4 (Maschke). Let G be a finite group, and let K be a field such that
the characteristic of K does not divide the order of G. Then every K[G]-module is
semisimple.
102 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
1 X −1
f (hv) = hx θ(xg v) = hf (v).
|G| g∈G g
Thus f is an endomorphism of V .
We have Im(f ) = U: Namely, Im(f ) ⊆ U since each term in the sum is in U. If
u ∈ U, then
1 X −1 1 X −1 1 X
f (u) = g θ(gu) = g gu = u = u.
|G| g∈G |G| g∈G |G| g∈G
13.3. Remarks. Let G be a finite group, and let K be a field. If char(K) does not
divide the order of G, then K[G] is semisimple. In this case, from our point of view,
the representation theory of K[G] is very boring. (But be careful: If you say this at
the wrong place and wrong time, you will be crucified.)
More interesting is the modular representation theory of G, i.e. the study of
representations of G over K where char(K) does divide |G|.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 103
Q
13.4. Exercises. 1: Prove that the K-algebra A := i∈I K is semisimple if and
only if the index set I is finite. If I is infinite, construct a submodule U of the
regular representation A A which does not have a direct complement in A A.
2: Let G = Z2 be the group with two elements, and let K be a field.
(a) Assume char(K) 6= 2. Show: Up to isomorphism there are exactly two simple
K[G]-modules.
(b) Assume char(K) = 2. Show: Up to isomorphism there are exactly two
indecomposable K[G]-modules, and one of them is not simple.
(c) Assume that K is an infinite field with char(K) = 2. Construct an infinite
number of 2-dimensional pairwise non-isomorphic representations of K[G ×
G].
104 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
————————————————————————————-
Proof. As an intersection of left ideals, J(A) is a left ideal. It remains to show that
J(A) is closed with respect to right multiplication. Let a ∈ A, then the right multi-
plication with a is a homomorphism A A → A A, and it maps rad(A A) to rad(A A).
Lemma 14.2. If A is semisimple, then J(A) = 0.
(i) x ∈ J(A);
(ii) For all a1 , a2 ∈ A, the element 1 + a1 xa2 has an inverse;
(iii) For all a ∈ A, the element 1 + ax has a left inverse;
(iv) For all a ∈ A, the element 1 + xa has a right inverse.
Proof. (i) =⇒ (ii): Let x ∈ J(A). We have to show that 1 + x is invertible. Since
x ∈ J(A), we know that x belongs to all maximal left ideals. This implies that 1 + x
does not belong to any maximal left ideal (because 1 is not contained in any proper
ideal).
We claim that A(1 + x) = A: The module A A is finitely generated. Assume that
A(1 + x) is a proper submodule of A A. Then Corollary 7.18 implies that A(1 + x)
must be contained in a maximal submodule of A A, a contradiction.
Therefore there exists some a ∈ A with a(1 + x) = 1. Let y = a − 1. We have
a = 1 + y, thus (1 + y)(1 + x) = 1, which implies y + x + yx = 0. This implies
y = (−1 − y)x ∈ Ax ⊆ J(A). Thus also 1 + y has a left inverse. We see that 1 + y
is left invertible and also right invertible. Thus its right inverse 1 + x is also its left
inverse. Since J(A) is an ideal, also a1 xa2 belongs to J(A) for all a1 , a2 ∈ A. Thus
all elements of the form 1 + a1 xa2 are invertible.
(ii) =⇒ (iii): Obvious.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 105
(iii) =⇒ (i): If x ∈
/ J(A), then there exists a maximal left ideal M, which does not
contain x. This implies A = M + Ax, thus 1 = y − ax for some y ∈ M and a ∈ A.
We get 1 + ax = y, and since y belongs to the maximal left ideal M, y cannot have
a left inverse.
(iii) ⇐⇒ (iv): Condition (ii) is left-right symmetric.
Corollary 14.4. The radical J(A) of A is the intersection of all maximal right
ideals.
Proof. Let I be a left ideal of A, and assume all elements in I are nilpotent. It is
enough to show that for all x ∈ I the element 1 + x is left-invertible. (If a ∈ A, then
ax ∈ I.) Since x is nilpotent, we can define
X
z= (−1)i xi = 1 − x + x2 − x3 + · · · .
i≥0
We get (1 + x)z = 1 = z(1 + x). The left-right symmetry shows that every right
ideal, which consists only of nilpotent elements is contained in the radical.
Warning: Nilpotent elements do not have to belong to the radical, as the following
example shows: Let A = M2 (K). Then A is a semisimple algebra, thus J(A) = 0.
But of course A contains many nilpotent elements, for example
0 1
x= .
0 0
But observe that there are elements y in Ax which are not nilpotent. In other words
1 + y is not invertible. For example
0 0
e=
0 −1
is in Ax and 1 + e is not invertible. We can also construct maximal left ideals of A
which do not contain x.
Proposition 14.6. Let a ∈ A. Then a ∈ J(A) if and only if aS = 0 for all simple
A-modules S.
Proof. Let T be a simple A-module, and let x be a non-zero element in T . The map
f : A A → T defined by f (b) = bx is an A-module homomorphism. Since x 6= 0, we
have f 6= 0. Since T is simple, f is surjective and the kernel of f is a maximal left
ideal. It follows from the definition of J(A) that J(A) is contained in the kernel of
f . Thus J(A)x = 0, and therefore J(A)T = 0.
Vice versa, assume aS = 0 for all simple A-modules S. We assume that a does not
belong to J(A). Since J(A) is the intersection of all maximal left ideals, there exists
106 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
In other words, the radical J(A) is the intersection of the annihilators of the simple
A-modules. (Given an A-module M, the annihilator of M in A is the set of all
a ∈ A such that aM = 0.)
Corollary 14.7. For every A-module M we have J(A)M ⊆ rad(M).
Proof. If A is a local ring, then J(A) is a maximal left ideal. Thus A/J(A) is a ring
which contains only one proper left ideal, namely the zero ideal. Thus A/J(A) is a
skew field.
Vice versa, if A/J(A) is a skew field, then J(A) is a maximal left ideal. We have to
show that J(A) contains every proper left ideal: Let L be a left ideal, which is not
contained in J(A). Thus J(A) + L = A. Now J(A) = rad(A A) is a small submodule
of A A, since A A is finitely generated. Thus L = A.
Theorem 14.11. If A/J(A) is semisimple, then for all A-modules M we have
J(A)M = rad(M).
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 107
Proof. We have seen that J(A)M ⊆ rad(M). On the other hand, M/J(A)M is
annihilated by J(A), thus it is an A/J(A)-module. Since A/J(A) is semisimple, each
A/J(A)-module is a semisimple A/J(A)-module, thus also a semisimple A-module.
But if M/J(A)M is semisimple, then rad(M) has to be contained in J(A)M.
Proof. We have J(A) ∩ eAe ⊆ eJ(A)e, since x ∈ eAe implies x = exe. Thus, if
additionally x ∈ J(A), then x = exe belongs to eJ(A)e.
Next we show that eJ(A)e ⊆ J(eAe): Let x ∈ J(A). If a ∈ A, then 1 + eae · x · e
is invertible, thus there exists some y ∈ A with y(1 + eaexe) = 1. This implies
eye(e + eaexe) = ey(1 + eaexe)e = e. Thus all elements in e + eAe(exe) are left-
invertible. This shows that exe belongs to J(eAe).
Finally, we show that J(eAe) ⊆ J(A) ∩ eAe: Clearly, J(eAe) ⊆ eAe, thus we have
to show J(eAe) ⊆ J(A). Let S be a simple A-module. Then eS = 0, or eS is a
simple eAe-module. Thus J(eAe)eS = 0, and therefore J(eAe)S = 0, which implies
J(eAe) ⊆ J(A).
14.2. Exercises. 1: Let Q be a quiver. Show that the radical J(KQ) has as a basis
the set of all paths from i to j such that there is no path from j to i, where i and j
run through the set of vertices of Q.
————————————————————————————-
Path algebras are an extremely important class of algebras. In fact, one of our main
aims is to obtain a better understanding of their beautiful representation theory and
also of the numerous links between representation theory of path algebras and other
areas of mathematics.
Several parts of this section are taken from Crawley-Boevey’s excellent lecture notes
on representation theory of quivers.
Thus we can think of Q as a finite directed graph. But note that multiple arrows
and loops (a loop is an arrow a with s(a) = t(a)) are allowed.
a = (a1 , a2 , . . . , am )
The path algebra KQ of Q over K is the K-algebra with basis the set of all paths
in Q. The multiplication of paths a and b is defined as follows:
If a = ei is of length 0, then
(
b if t(b) = i,
ab := a · b :=
0 otherwise.
If b = ei , then
(
a if s(a) = i,
ab := a · b :=
0 otherwise.
a b
2o 4 / 5
c d e
f
1 / 3
Clearly, each Aei is a left A-module. So this is a direct decomposition of the regular
representation A A.
Lemma 15.1. If 0 6= x ∈ Aei and 0 6= y ∈ ei A, then xy 6= 0.
Proof. Look at the longest paths p and q involved in x and y, respectively. In the
product xy the coefficient of pq cannot be zero.
Lemma 15.2. The ei are primitive idempotents.
Proof. The vector space Aej A has as a basis the paths passing through the vertex
j.
Lemma 15.5. If i 6= j, then Aei ∼
6= Aej .
Proof. Assume i 6= j and that there exists an isomorphism f : Aei → Aej . Set
y = f (ei ). It follows from Lemma 12.10 that y ∈ ei Aej . Let g = f −1 , and let
x = g(ej ). This implies
(gf )(ei) = g(y) = g(yej ) = yg(ej ) = yx = ei .
A similar calculation shows that xy = ej . But y ∈ ei Aej and x ∈ Aei . Thus
y = ei yej and x = xei . This implies ej = xy = xei yej ∈ Aei A, a contradiction to
Lemma 15.4.
A morphism
θ: V → W
between representations V = (Vi , Va )i,a and W = (Wi , Wa )i,a is given by linear maps
θi : Vi → Wi , i ∈ Q0 such that the diagram
θs(a)
Vs(a) / Ws(a)
Va Wa
θt(a)
Vt(a) / Wt(a)
commutes for each a ∈ Q1 . The vector space of homomorphisms from V to W is
denoted by Hom(V, W ), or more precisely by HomQ (V, W ).
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 111
So from now on we can use all the terminology and the results we obtained for
modules over algebras also for representations of quivers. In particular, we get a
Jordan-Hölder and a Krull-Remak-Schmidt Theorem, we can ask for Auslander-
Reiten sequences of quiver representations, etc. We will often not distinguish any
longer between a representation of Q and a module over KQ.
If V = (Vi , Va )i,a is a representation of Q let
d = dim(V ) = (dim Vi )i∈Q0
be its dimension vector. If V is a finite-dimensional indecomposable representa-
tion, then dim(V ) ∈ NQ0 is called a root of Q. A root d is a Schur root if there
exists a representation V with EndQ (V ) ∼ = K and dim(V ) = d. Assume that d is
a root. If there exists a unique (up to isomorphism) indecomposable representation
V with dim(V ) = d, then d is called a real root. Otherwise, d is an imaginary
root.
A representation V of Q is rigid (or exceptional) if each short exact sequence
0→V →W →V →0
splits, i.e. if W ∼
= V ⊕V.
Here are some typical problems appearing in representation theory of quivers:
4: Let Q be any quiver. Determine the centre of KQ. (Reminder: The centre
C(A) of an algebra A is defined as C(A) = {a ∈ A | ab = ba for all b ∈ A}.)
5: Let Q be a quiver with n vertices. Show that there are n isomorphism classes of
simple KQ-modules if and only if Q has no oriented cycles.
6: Let Q be a quiver. Show that the categories RepK (Q) and Mod(KQ) are equiv-
alent.
◦ / ◦ / ◦o ◦o ◦
with dimension vector
1
1 2 3 2 1
◦ / ◦ / ◦ / ◦ / ◦ / ◦
Let A = KQ. Write A A as a direct sum of indecomposable representations and
compute the dimension of the indecomposable direct summands.
10: Let
K[T ]/(T 2 ) 0
A= .
K[T ]/(T 2 ) K
This gives a K-algebra via the usual matrix multiplication. (The elements of A are
of the form
a 0
b c
where a, b ∈ K[T ]/(T 2 ) and c ∈ K.) Show that A is isomorphic to KQ/I where Q
is the quiver
9◦
α /◦
————————————————————————————
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 115
Many problems in Linear Algebra can be reformulated using quivers. In this section,
we give some examples of this kind.
Answer: Of course we can, since we paid attention in our Linear Algebra lectures.
For n = 0 everything is trivial, there is just one orbit containing only the zero
map. Thus assume n ≥ 1. Fix a basis B of V . Now each map f ∈ EndK (V ) is
(with respect to B) given by a particular matrix, which (via conjugation) can be
transformed to a Jordan normal form. It follows that each orbit Gf is uniquely
determined by a set
{(n1 , λ1 ), (n2 , λ2 ), . . . , (nt , λt )}
where the ni are positive integers with n1 + · · · + nt = n, and the λi are elements in
K. Here (ni , λi ) stands for a Jordan block of size ni with Eigenvalue λi .
Answer: Of course we can. This is even easier than the previous problem: Fix
bases B1 and B2 of V1 and V2 , respectively. Then each f ∈ HomK (V1 , V2 ) is given
by a matrix with respect to B1 and B2 . Now using row and column transformations
116 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
(which can be expressed in terms of matrix multiplication from the left and right)
we can transform the matrix of f to a matrix of the form
Er 0
0 0
where Er is the r × r-unit matrix and the zeros are matrices with only zero entries.
Here r is the rank of the matrix of f .
It turns out that there are 1+min{n1 , n2 } different G-orbits where ni is the dimension
of Vi .
Answer: Yes, we can do that, by we will need a bit of theory here. As you can
see, the problem became more complicated, because we simultaneously transform
the matrices of f1 and f2 with respect to some fixed bases of V1 and V2 .
1 λ 1 µ
Example: The orbits G(K −
→ K, K −
→ K) and G(K −
→ K, K −
→ K) are equal if
and only if λ = µ.
• f (Vi ) ⊆ Wi ;
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 117
We will see for which n there are only finitely many indecomposable n-subspace
configurations.
Instead of asking for the classification of all n-subspace configurations, we might ask
the following easier question:
Problem 16.5. Classify the dimension vectors of the indecomposable n-subspace
configurations.
118 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
***********************************************************************
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 119
In this section let A be a K-algebra. As before, let Mod(A) be the category of left
A-modules, and let mod(A) be the full subcategory of finitely generated A-modules.
The map h′ constructed in the proof of the above lemma is in general not uniquely
determined. There can be many different maps h′ with the property gh′ = h:
For example, for F = A A the set {1A } is a free generating set. Let M be an
A-module and let U be a submodule of M. By g : M → M/U we denote the
corresponding projection map. This is a typical epimorphism. For x ∈ M let x =
x+U be the corresponding residue class in M/U. Let h : A A → M/U be an arbitrary
homomorphism. Then h(1A ) = x for some x ∈ M. Now the homomorphisms
h′ : A A → M such that gh′ = h correspond to the elements in x = x + U = {x + u |
120 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
(i) P is projective;
(ii) Every epimorphism M → P splits;
(iii) P is isomorphic to a direct summand of a free module.
(iii) =⇒ (i): The class of projective modules contains all free modules and is closed
under direct summands.
Now we proof the last statement of the Lemma: If P has a generating set X of
cardinality c, then let F be a free module of rank c. Thus F has a generating set of
cardinality c. We get an epimorphism F → P which has to split.
Vice versa, if P is a direct summand of a free module F of rank c, then P has a
generating set of cardinality c: We choose an epimorphism f : F → P , and if X is
a generating set of F , then f (X) is a generating set of P .
Proof. By definition J(A) = rad(A A) and J(A)A = J(A). This shows that the
statement is true for P = A A. Now let Mi , i ∈ I be a family of modules. We have
! !
M M M M
J(A) Mi = J(A)Mi and rad Mi = rad(Mi ).
i∈I i∈I i∈I i∈I
122 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
L i ⊂ rad(M
This is a proper inclusion only if there existsLsome i with J(A)M i ). Thus,
if J(A)Mi = rad(Mi ) for all i, then J(A) i∈I M i = rad i∈I Mi . This shows
that the statement is true for free modules.
L L
Vice versa, if J(A) i∈I Mi = rad i∈I M i , then J(A)Mi = rad(Mi ) for all i.
Since projective modules are direct summands of free modules, and since we proved
the statement already for free modules, we obtain it for all projective modules.
Lemma 17.8. Let U be a submodule of a projective module P such that for every
endomorphism f of P we have f (U) ⊆ U. Define
f∗ : P/U → P/U
by f∗ (x + U) := f (x) + U. Then the following hold:
Let M be an A-module. For all f ∈ EndA (M) we have f (rad(M)) ⊆ rad(M). Thus
the above lemma implies that for any projective module P there is a surjective
algebra homomorphism EndA (P ) → EndA (P/ rad(P )), and the kernel is the set of
all endomorphisms of P whose image is contained in rad(P ).
We have shown already: If rad(P ) is a small submodule of P , then the set of all
endomorphisms of P whose image is contained in rad(P ) is exactly the radical of
EndA (P ). Thus, we proved the following:
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 123
The converse of Lemma 17.11 is also true. We will not use this result, so we skip
the proof.
• P is projective;
• p is an epimorphism;
• Ker(p) is a small submodule of P .
In this situation one often calls the module P itself a projective cover of M and
writes P = P (M).
Lemma 17.12. Let P be a finitely generated projective module. Then the projection
map P → P/ rad(P ) is a projective cover.
Proof. The projection map is surjective and its kernel is rad(P ). By assumption P
is projective. For every finitely generated module M the radical rad(M) is a small
submodule of M. Thus rad(P ) is small in P .
124 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Proof. Since P and Q are finitely generated projective modules, the projections
p : P → P/ rad(P ) and q : Q → Q/ rad(Q) are projective covers. If f : P/ rad(P ) →
Q/ rad(Q) is an isomorphism, then f ◦ p : P → Q/ rad(Q) is a projective cover. The
uniqueness of projective covers yields P ∼
= Q. The other direction is obvious.
Corollary 17.15. Let P be a direct sum of local projective modules. If U is a
submodule of P which is not contained in rad(P ), then there exists an indecomposable
direct summand P ′ of P which is contained in U.
L
Proof. Let P = i∈I Pi with local projective modules PiL . Let U be a submodule of
P which is not contained in rad(P ). We have rad(P ) = i∈I rad(Pi ) and therefore
M
P/ rad(P ) = Pi / rad(Pi ).
i∈I
U
C
u
P
{
f { p
{
{
{
{ P/ rad(P )
{ g
{ { πi
{
}{ pi
Pi / Pi / rad(Pi )
Thus we have
pi guf = πi puf = pi .
Proof. (i) =⇒ (ii): Small submodules of a module are always contained in the
radical.
(ii) =⇒ (iii): Since Ker(p) ⊆ rad(P ) we have rad(P/ Ker(p)) = rad(P )/ Ker(p).
Now p induces an isomorphism P/ Ker(p) → M which maps rad(P/ Ker(p)) onto
rad(M) and induces an isomorphism P/ rad(P ) → M/ rad(M).
(iii) =⇒ (i): We assume that p : P → M induces an isomorphism p∗ : P/ rad(P ) →
M/ rad(M). This implies rad(M) + Im(p) = M. Since M is a finitely generated
module, its radical is a small submodule. Thus Im(p) = M. We see that p is an
epimorphism. Since p∗ is injective, the kernel of p must be contained in rad(P ). The
radical rad(P ) is small in P because P is finitely generated. Now Ker(p) ⊆ rad(P )
implies that Ker(p) is small in P .
————————————————————————————-
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 127
(i) I is injective;
(ii) The functor HomA (−, I) is exact;
(iii) Every monomorphism I → N splits;
(i) A is semisimple;
(ii) Every A-module is projective;
(iii) Every A-module is injective.
Proof. Recall that A is semisimple if and only if all A-modules are semisimple. A
module M is semisimple if and only if every submodule of M is a direct summand.
128 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
For any left A-module A M let D(A M) = HomK (A M, K) be the dual module of
op
A M. This is a right A-module, or equivalently, a left A -module: For α ∈ D(A M),
a ∈ Aop and x ∈ A M define (aα)(x) := α(ax). It follows that ((ab)α)(x) = α(abx) =
(aα)(bx) = (b(aα))(x). Thus (b ⋆ a)α = (ab)α = b(aα) for all x ∈ M and a, b ∈ A.
and !
M Y
HomA Mi , − ∼
= HomA (Mi , −).
i∈I i∈I
Proof. Exercise.
Lemma 18.5. The following hold:
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 129
Warning: Infinite direct sums of injective modules are often not injective. The
reason is that in general we have
!
M M M
HomA (−, Mi ) ∼
6= HomA (Mi , −) 6∼
= HomA Mi , − .
i∈I i∈I i∈I
Now Lemma 18.5 (ii) implies that D(PA ) is projective. Any projective module is
a direct summand of a free module. Thus Lemma 18.5 (i) yields that D(PA ) is an
injective A-module for all projective Aop -module PA .
One can now define injective resolutions, and develop Homological Algebra with
injective instead of projective modules.
Recall that a submodule U of a module M is called large if for any non-zero sub-
module V of M the intersection U ∩ V is non-zero.
A homomorphism f : M → I is called an injective envelope if the following hold:
(i) I is injective;
(ii) f is a monomorphism;
(iii) f (M) is a large submodule of I.
Lemma 18.8. Let I be an injective module, and let U and V be submodules of I
such that U ∩ V = 0. Assume that U and V are maximal with this property (i.e.
if U ⊆ U ′ with U ′ ∩ V = 0, then U = U ′ , and if V ⊆ V ′ with U ∩ V ′ = 0, then
V = V ′ ). Then I = U ⊕ V .
Proof. Let
V := {W ⊆ M | U ∩ W = 0}.
Take S
a chain (Vi )i∈J in V. (Thus for all Vi and Vj we have Vi ⊆ Vj or Vj ⊆ Vi .) Set
V = i Vi . We get
!
[ [
U ∩V =U ∩ Vi = (U ∩ Vi ) = 0.
i i
Now the claim follows from Zorn’s Lemma.
Next, let
U := {U ⊆ I | U ∩ V = 0 and X ⊆ U}.
Again, by Zorn’s Lemma we obtain a submodule U of I which is maximal with
U ∩ V = 0 and X ⊆ U.
Thus, U and V are as in the assumptions of the previous lemma, and we obtain
I = U ⊕ V and X ⊆ U. We know that U is injective, and we have our embedding
X → U.
Proof. Exercise.
————————————————————————————-
If I is an ideal in C we can define the factor category C/I as follows: The objects
are the same as in C and
HomC/I (C, C ′) := HomC (C, C ′)/I(C, C ′ ).
The composition of morphisms is defined in the obvious way.
If X is a class of objects in C which is closed under finite direct sums, then we say
that f : C → C ′ factors through X , if f = f2 ◦ f1 with f1 : C → X, f2 : X → C ′
and X ∈ X . Let I(X )(C, C ′ ) be the set of morphisms C → C ′ which factor through
X . In this way we obtain an ideal I(X ) in C.
Now let A be an arbitrary K-algebra, and as before let Proj(A) be the full subcate-
gory of projective A-modules. The stable module category of Mod(A) is defined
as
Mod(A) = Mod(A)/I(Proj(A)).
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 133
Define
Hom(M, N) := HomMod(A)/Proj(A) (M, N) = HomA (M, N)/I(Proj(A))(M, N).
Similarly, we define mod(A) = mod(A)/I(proj(A)).
Thus the objects of Mod(A) are the same ones as in Mod(A), namely just the
A-modules. But it follows that all projective A-modules become zero objects in
Mod(A): If P is a projective A-module, then 1P lies in I(Proj(A))(P, P ). Thus 1P
becomes zero in Mod(A). Vice versa, if a module M is a zero object in Mod(A),
then M is a projective A-module: If 1M factors through a projective A-module, then
M is a direct summand of a projective module and therefore also projective.
Now Schanuel’s Lemma implies the following: If M is an arbitrary module, and if
p : P → M and p′ : P ′ → M are epimorphisms with P and P ′ projective, then the
kernels Ker(p) and Ker(p′ ) are isomorphic in the category Mod(A).
If we now choose for every module M an epimorphism p : P → M with P projective,
then M 7→ Ker(p) yields a functor Mod(A) → Mod(A). If we change the choice of
P and p, then the isomorphism class of Ker(p) in Mod(A) does not change.
So it makes sense to work with an explicit construction of a projective module P
and an epimorphism p : P → M. Let M be a module, and let F (M) be the free
module with free generating set |M| (the underlying set of the vector space M).
Define
p(M) : F (M) → M
by m 7→ m. In this way we obtain a functor F : Mod(A) → Mod(A)
————————————————————————————-
There is a beautiful general theory on projective modules, however one can cut
this short and concentrate on finite-dimensional projective modules over finite-
dimensional algebras. The results in this section can be generalized considerably.
The general theory is developed in Sections 22 and 23.
Theorem 20.1 (Special case of Theorem 23.1). Let A be a finite-dimensional alge-
bra. Then A/J(A) is semisimple.
Proof. The module A A is a finite direct sum of local modules, thus A A/ rad(A A) is
a finite sum of simple modules and therefore semisimple.
Theorem 20.2 (Special case of Theorem 23.2). Let A be a finite-dimensional alge-
bra. If
A A = P1 ⊕ · · · ⊕ Pn
Proof. If P and Q are isomorphic modules, then P/ rad(P ) and Q/ rad(Q) are also
isomorphic. Therefore P 7→ P/ rad(P ) yields a well defined map.
Ln
The map is surjective: Let S be a simple module. We write A A = i=1 Pi with
indecomposable modules Pi . Since Pi is of finite length and indecomposable, we
know that EndA (Pi ) is a local ring. Furthermore, Pi = Aei for some idempotent ei ∈
A. Since EndA (Pi ) ∼
= (ei Aei )op we know that ei Aei is also a local ring. Therefore,
Lemma 12.14 implies that Aei is a local module.
There exists a non-zero homomorphism A A → S, and therefore for at least one index
i we get a non-zero map f : Pi → S. Since S is simple, we know that f is surjective.
Furthermore, the kernel of f is rad(Pi ) because Pi is local. Thus S is isomorphic to
Pi / rad(Pi ).
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 135
(i)P is local;
(ii)EndA (P ) is a local ring;
(iii)J(EndA (P )) = {f ∈ EndA (P ) | Im(f ) ⊆ rad(P )};
(iv) Each endomorphism of P induces an endomorphism of S, and we obtain an
algebra isomorphism EndA (P )/J(EndA (P )) → EndA (S); Lm
(v) The multiplicity of P in a direct sum decomposition A A = i=1 Pi with
indecomposable modules Pi is exactly the dimension of S as an EndA (S)-
module.
The multiplicity of P in this decomposition (in other words, the number of direct
summands Pi which are isomorphicLto P ) is equal to the multiplicity of S in the
m
direct decomposition A A/J(A) = i=1 Pi / rad(Pi ). But this multiplicity of S is
the dimension of S as an EndA (S)-module. (Recall that A/J(A) is a semisimple
algebra, and that EndA (S) is a skew field.)
Theorem 20.5. Let A be a finite-dimensional algebra. Then every finitely generated
module has a projective cover.
s9 M
p s
s ε
s
s
s
P q
/ M/ rad(M)
Since P and M are finitely generated we see that p is a projective cover.
————————————————————————————-
• A is finite-dimensional;
• A/J(A) ∼ =K| ×K× {z· · · × K}.
n times
In this case, there are n isomorphism classes of simple A-modules, and each simple
A-module is 1-dimensional.
Proof. Later.
Theorem 21.2 (Gabriel). Let A be a finite-dimensional K-algebra with K alge-
braically closed. Then there exists a uniquely determined quiver Q and an admissible
ideal I in KQ such that the categories mod(A) and mod(KQ/I) are equivalent.
Proof. Later.
We will actually not use Theorems 21.1 and 21.2 very often. But of course these
results are still of central importance, because they tell us that path algebras and
their quotients by admissible ideals are not at all exotic. They are hidden behind
every finite-dimensional algebra over an algebraically closed field.
Assume now that I is an admissible ideal in a path algebra KQ and set A := KQ/I.
Proof. We know that J(A) consists of all elements x ∈ A such that xS = 0 for all
6 0 for all i ∈ Q0 .
simple A-modules S, see Proposition 14.6. By definition ei Si =
Thus none of the ei belongs to J(A).
The rest of the theorem follows from results we proved before for arbitrary finite-
dimensional algebras.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 139
————————————————————————————-
Let
m
M n
M
M =U⊕ Mi = U ⊕ Nj ⊕ V
i=1 j=1
be direct decompositions of M, and assume that the endomorphism rings of the
modules Nj are local. Take
M = U ⊕ N ′ ⊕ (Nn ⊕ V )
Ln−1
where N ′ = j=1 Nj . By the induction assumption there are direct decompositions
Mi = Xi ⊕ Yi such that
Mm Mm
M = U ⊕ N′ ⊕ Xi and N ′ ∼ = Yi .
i=1 i=1
and the inclusion homomorphism from Nn into M. Then we apply the Exchange
Theorem (see Skript 1) to this situation: We use that Nn ⊕ (U ⊕ N ′ ) is a direct
summand of M. For 1 ≤ i ≤ m we obtain a direct decomposition Xi = Mi′ ⊕ Xi′
such that
Mm
′
M = (U ⊕ N ) ⊕ Nn ⊕ Mi′
i=1
Lm ′ ∼
with i=1 Xi = Nn . Note that only one of the modules Xi′ is non-zero. Define
Mi′′ = ′
Xi ⊕ Yi . This implies
Mi = Xi ⊕ Yi = Mi′ ⊕ Xi′ ⊕ Yi = Mi′ ⊕ Mi′′
140 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
and m m m
M M M
Mi′′ = Xi′ ⊕ Yi ∼
= N ′ ⊕ Nn = N.
i=1 i=1 i=1
This finishes the proof.
L
If M = i∈I Mi is a direct sum of modules Mi , and J is a subset of the index set
I, we define M
MJ := Mi .
i∈J
We want to study modules which can be written as direct sums of modules with
local endomorphism ring. The key result in this situation is the following:
L
Theorem 22.2. Let M = i∈I Mi be a direct sum of modules Mi with local endo-
morphism rings, and let U be a direct summand of M. Then the following hold:
Since M/C is indecomposable, we know that exactly one of the modules Mi′ , say
Mi′0 , is non-zero. On the other hand, Mi0 = Mi′0 ⊕ Mi′′0 is indecomposable, and
therefore Mi′0 = Mi0 . Thus M = C ⊕ Mi0 . This proves part (b) for the direct
summand C of M.
L
Corollary 22.3. Let M = i∈I Mi be a direct sum of modules Mi with local endo-
morphism rings. Then every non-zero direct summand of M has a direct summand
which is isomorphic to one of the Mi ’s.
Proof. First, let J(N) be finite and non-empty, and let j0 ∈ J(N). Corollary 22.4
yields that there exists some i0 ∈ I with Mi0 ∼= Uj0 . The Cancellation Theorem
implies that M M
Mi ∼
= Uj .
i∈I\{i0 } j∈J\{j0 }
142 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
L
Proof. Let U = j∈J Uj with indecomposable module Uj . We choose a direct
complement C of U, thus
M
M = U ⊕C = Uj ⊕ C.
j∈J
To this decomposition we apply the above Theorem 22.5. Thus for any indecompos-
able module N we have |J(N)| ≤ |I(N)|. Thus there is an injective map σ : J → I
such that Uj ∼
= Mσ(j) for all j. We can identify J with a subset I ′ of I, and we
∼
obtain U = MI ′ .
L
Corollary 22.7 (Krull-Remak-Schmidt-Azumaya). Assume that M = L i∈I Mi is
a direct sum of modules Mi with local endomorphism rings. Let M = j∈J Uj
with indecomposable modules Uj . Then there exists a bijection σ : I → J such that
Mi ∼
= Uσ(i) .
————————————————————————————-
146 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
If A is semiperfect, then Aop is semiperfect as well. This follows since condition (iii)
is left-right symmetric.
Examples:
Proof. The module A A is a finite direct sum of local modules, thus A A/ rad(A A) is
a finite sum of simple modules and therefore semisimple.
Warning: The converse of the above theorem does not hold: By K(T ) we denote
the field of rational functions in one variable T . Thus K(T ) consists of fractions
f /g of polynomials f, g ∈ K[T ] where g 6= 0. Now let A be the subring of K(T )
consisting of all rational functions f /g such that neither T nor T − 1 divide g. The
radical J(A) of A is the ideal generated by T (T − 1), and the corresponding factor
ring A/J(A) is isomorphic to K × K, in particular it is semisimple. Note that A
has no zero divisors, but A/J(A) contains the two orthogonal idempotents −T + 1
and T . For example
(−T + 1)2 = T 2 − 2T + 1 = (T 2 − T ) − T + 1,
and modulo T 2 − T this is equal to −T + 1.
Theorem 23.2. Let A be a semiperfect algebra. Then the following hold:
AA = Qi
i=1
with local modules Qi . As a direct summand of A A each Qi is of the form Aei
for some idempotent ei . In particular Qi is cyclic. The endomorphism ring of an
A-module of the form Ae (where e is an idempotent) is (eAe)op , and if Ae is local,
then so is eAe. Thus also (eAe)op is a local ring.
Let P be a projective A-module. Thus P is a direct summand of a free A-module F .
We know that F is a direct sum of modules with local endomorphism ring, namely
of copies of the Qi . Kaplansky’s Theorem implies that
M
P = Pj
j∈J
(i)P is local;
(ii)EndA (P ) is a local ring;
(iii)J(EndA (P )) = {f ∈ EndA (P ) | Im(f ) ⊆ rad(P )};
(iv) Each endomorphism of P induces an endomorphism of S, and we obtain an
algebra isomorphism EndA (P )/J(EndA (P )) → EndA (S); Lm
(v) The multiplicity of P in a direct sum decomposition A A = i=1 Pi with
indecomposable modules Pi is exactly the dimension of S as an EndA (S)-
module.
————————————————————————————-
...
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 149
>
V1 ATTTT
f1 ~~ AA TTTTT h
~ AA TTT1 T
~~ g1 AAA TTTT
~~ TTT
U@ W _ _ _h _ _ _Tj*/4 X
@ > } j jjj
@@ g2 }} jjjj
@@ }}} jjjjjj
f2 @ }}jjjjj h2
j
V2
150 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
• g1 f1 = g2 f2 ;
• For all homomorphisms h1 : V1 → X, h2 : V2 → X such that h1 f1 = h2 f2
there exists a unique(!) homomorphism h : W → X such that hg1 = h1 and
hg2 = h2 .
Lemma 25.2. Let f1 : U → V1 , f2 : U → V2 be homomorphisms, and assume that
(g1 : V1 → W, g2 : V2 → W ) and also (g1′ : V1 → W ′ , g2′ : V2 → W ′ ) are pushouts of
(f1 , f2 ). Then there exists an isomorphism h : W → W ′ such that hg1 = g1′ and
hg2 = g2′ . In particular, W ∼= W ′.
Proof. Exercise.
Of course, also
t [f
1 ,−f2 ] [g1 ,g2 ]
0 → U −−−−−→ V1 ⊕ V2 −−−→ W.
is exact.
Proposition 25.3 (Universal property of the pullback). For the module U and
the homomorphisms f1 : U → V1 and f2 : U → V2 as defined above the following
hold: We have g1 f1 = g2 f2 , and for every module Y together with a pair (h1 : Y →
V1 , h2 : Y → V2 ) of homomorphisms such that g1 h1 = g2 h2 there exists a uniquely
determined homomorphism h : Y → U such that h1 = f1 h and h2 = f2 h.
j5 V1 A
h1 jjjjj
jjjj~~> AA g1
~ AA
jjj ~
~ f1 AA
j jjjjj ~~ A
j j h
Y T_TT_T _ _ _ _/ U @ W
TTTT
TTTT @@ f }}>
2
TTTT @@@ }}
}}
h2 TTTT @
T) }} g2
V2
Proof. Exercise.
• g1 f1 = g2 f2 ;
• For all homomorphisms h1 : Y → V1 , h2 : Y → V2 such that g1 h1 = g2 h2
there exists a unique(!) homomorphism h : Y → U such that f1 h = h1 and
f2 h = h2 .
Lemma 25.4. Let g1 : V1 → W , g2 : V2 → W be homomorphisms, and assume that
(f1 : U → V1 , f2 : U → V2 ) and also (f1′ : U ′ → V1 , f2′ : U ′ → V2 ) are pullbacks of
(g1 , g2 ). Then there exists an isomorphism h : U ′ → U such that f1 h = f1′ and
f2 h = f2′ . In particular, U ∼
= U ′.
Proof. Exercise.
j5 V1 A
f1jjj jjjj}}> AA g1
j jj }} AA
jjjj } f′ AA
jj j }} 1
A
j jj h
U T_TT_T _ _ _ _/ U ′ >W
TTTT AA ′
TTTT AAf2 }}}
TTTT AA }
}}
f2 TTTT A
T) }} g2
V2
Proof. Exercise.
Lemma 25.6. Let (f1 : U → V1 , f2 : U → V2 ) be the pullback of homomorphisms
(g1 : V1 → W, g2 : V2 → W ), and let (g1′ : V1 → W ′ , g2′ : V2 → W ′ ) be the pushout of
(f1 , f2 ). Then the uniquely determined homomorphism h : W ′ → W with g1 = hg1′
and g2 = hg2′ is injective. If [g1 , g2 ] is surjective, then h is an isomorphism, and
(g1 , g2 ) is the pushout of (f1 , f2 ).
V1 TBTTT
f1 ~? BB TTTTT g1
~~~ BB
B
TTTT
TTTT
~ g1′ BB!
~~ TTTT
T
′ _ _ _h _ _ _/*4
U@ >W jjjj
W
@ | j
@@ g2′ || jjjj
@@ ||| jjjjjgjj
f2 @ || jj 2
jjjj
V2
Proof. Exercise.
Lemma 25.7. Let (g1 : V1 → W, g2 : V2 → W ) be the pushout of a pair (f1 : U →
V1 , f2 : U → V2 ). If f1 is injective, then g2 is also injective.
Proof. Assume g2 (v2 ) = 0 for some v2 ∈ V2 . By definition g2 (v2 ) is the residue class
of (0, v2 ) in W , thus there exists some u ∈ U with (0, v2 ) = (f1 (u), −f2 (u)). If we
assume that f1 is injective, then 0 = f1 (u) implies u = 0. Thus v2 = −f2 (u) = 0.
Lemma 25.8. Let (f1 : U → V1 , f2 : U → V2 ) be the pullback of a pair (g1 : V1 →
W, g2 : V2 → W ). If g1 is surjective, then f2 is also surjective.
Pushouts are often used to construct bigger modules from given modules. If V1 , V2
are modules, and if U is a submodule of V1 and of V2 , then we can construct the
pushout of the inclusions f1 : U → V1 , f2 : U → V2 . We obtain a module W and
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 153
homomorphisms g1 : V1 → W , g2 : V2 → W with g1 f1 = g2 f2 .
>
V1 A
~~
f1 AA g1
~ AA
~~ AA
~~ A
U@ >W
@ @@ }}
@@ }}
}}
f2 @ }} g2
V2
Since f1 and f2 are both injective, also g1 and g2 are injective. Also (up to canonical
isomorphism) (f1 , f2 ) is the pullback of (g1 , g2 ).
Since (f, g) is an exact sequence, there is some u ∈ U with f (u) = v. This implies
and similarly as before we can show that (f ′′ , g ′′ ) is again a short exact sequence.
We write b∗ (f, g) = (f ′′ , g ′′), and call this the (short exact) sequence induced
by b.
a′′ : V → X
b′′ : Y → V
Y
}
b′′ }
} b
f ~} g
0 / U /V / W / 0
~
a ~
~ a′′
~~
X
> V @T@TTTTTT
f ~~~ @@ TTTTa′′
~~ @@ TTTT
~~ a′ @ TTTT
h_ _ T_
T/)
U @@ > P _ _ _
jj j
j 5 X
@@ f ′ ~~~ jjj
jjj
@
a @@ ~ ~~ jjjjj1jX
~jjj jj
X
In particular, f ′ is a split monomorphism. Thus the sequence (f ′ , g ′ ) = a∗ (f, g)
splits.
The second part of the lemma is proved dually.
Lemma 25.10. Let
f g
0 / U / V / W / 0
a a′′
f ′′ g ′′
0 / X / P′ / W / 0
be a commutative diagram with exact rows. Then the pair (a′′ , f ′′ ) is a pushout of
(f, a).
Proof. We construct the induced exact sequence a∗ (f, g) = (f ′ , g ′): Let (a′ : V →
P, f ′ : X → P ) be the pushout of (f, a). For g ′ : P → W we have g = g ′ a′ .
Since a′′ f = f ′′ a there exists some homomorphism h : P → P ′ with a′′ = ha′
and f ′′ = hf ′ . We claim that g ′ = g ′′ h: This follows from the uniqueness of the
factorization through a pushout, because we know that
g ′ a′ = g = g ′′ a′′ = g ′′ ha′
and
g ′ f ′ = 0 = g ′′ f ′′ = g ′′ hf ′ .
Thus we have seen that h yields an equivalence of the short exact sequences (f ′ , g ′)
and (f ′′ , g ′′ ), In particular, h has to be an isomorphism. But if h is an isomorphism,
then the pair (a′′ = ha′ , f ′′ = hf ′ ) is a pushout of (f, a), since by assumption (a′ , f ′ )
is a pushout of (f, a).
Proof. Let M := P/U, and let p : P → M be the projection map. Similarly, let
q : Q → M be the epimorphism with kernel V (since Q/V is isomorphic to M such
a q exists).
We construct the pullback of (p, q) and obtain a commutative diagram
p′
E / Q
q′ q
p
P / M
where p is an epimorphism with kernel isomorphic to U = Ker(p), and q ′ is an
′
iV
V ′ _ _ _/ V
p′
0 / U ′ / E / Q / 0
iU
q′ q
p
0 / U / P / M / 0
0 0
We can assume E = {(v, w) ∈ P ⊕ Q | p(v) = q(w)}, q ′ (v, w) = v and p′ (v, w) = w
for all (v, w) ∈ E. Now it is easy to define homomorphisms iU and iV such that
everything commutes, and then ones shows that iU and iV are in fact isomorphisms.
f
25.7. Short exact sequences with projective middle term. Let η : 0 → X − →
g ′ ′
′ ′ f ′ g
Y −→ Z → 0 and η : 0 → X − →Y − → Z → 0 be short exact sequences. We say
that η induces η if there exist homomorphisms h and h′ such that the diagram
′
f g
0 / X / Y / Z / 0
h h′
f′ g′
0 / X′ / Y′ / Z / 0
158 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
commutes.
Lemma 25.13. Let
u p
0→U − →P − →W →0
be a short exact sequence of A-modules with P a projective module. Then this se-
quence induces every short exact sequence which ends in W .
Proof. Let
f g
0 → U′ −
→ V′ −
→W →0
be an arbitrary short exact sequence of A-modules. Since g is an epimorphism, the
lifting property of the projective module P yields a homomorphism p′ : P → V ′ such
that gp′ = p. This implies gp′ u = pu = 0. Thus p′ u can be factorized through the
kernel of g. So there exists some h : U → U ′ such that p′ u = f h. Thus we obtain
the following commutative diagram with exact rows:
u p
0 / U / P / W / 0
h p′
f g
0 / U′ / V′ / W / 0
This shows that (f, g) is the short exact sequence induced by h.
• The induced sequences (a′ a)∗ (f, g) and a′∗ (a∗ (f, g)) are equivalent;
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 159
• The induced sequences (bb′ )∗ (f, g) and (b′ )∗ (b∗ (f, g)) are equivalent;
• The induced sequences a∗ (b∗ (f, g)) and b∗ (a∗ (f, g)) are equivalent.
————————————————————————————-
f0 g0
U0 / V0 / W0
a0 b0 c0
f1 g1
U1 / V1 / W1 / 0
a b c
f2 g2
0 / U2 / V2 / W2
a2 b2 c2
f3 g3
U3 / V3 / W3
0 0 0
with exact rows and columns. Then
δ(x) := (a2 ◦ f2−1 ◦ b ◦ g1−1 ◦ c0 )(x)
defines a homomorphism (the “connecting homomorphism”)
δ : Ker(c) → Cok(a)
such that the sequence
f0 g0 δ f3 g3
Ker(a) −
→ Ker(b) −
→ Ker(c) −
→ Cok(a) −
→ Cok(b) −
→ Cok(c)
is exact.
160 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Proof. The proof is divided into two steps: First, we define the map δ, second we
verify the exactness.
Relations
δ is a homomorphism
w_0
c0
g1
v_1 / w1
b
f2
u_2 / v2
a2
u3
We have to show that for every w0 ∈ W0 there exists a tuple
(w0 , w1 , v1 , v2 , u2 , u3)
in S, and that for two tuples (w0 , w1 , v1 , v2 , u2 , u3 ) and (w0′ , w1′ , v1′ , v2′ , u′2 , u′3) with
w0 = w0′ we always have u3 = u′3 .
Thus, let w ∈ W0 . Since g1 is surjective, there exists some v ∈ V1 with g1 (v) = c0 (w).
We have
g2 b(v) = cg1 (v) = cc0 (w) = 0.
Therefore b(v) belongs to the kernel of g2 and also to the image of f2 . Thus there
exists some u ∈ U2 with f2 (u) = b(v). So we see that
(w, c0 (w), v, b(v), u, a2(u)) ∈ S.
Now let (w, c0(w), v ′, b(v ′ ), u′ , x) also be in S. We get
g1 (v − v ′ ) = c0 (w) − c0 (w) = 0.
Thus v − v ′ belongs to the kernel of g1 , and therefore to the image of f1 . So there
exists some y ∈ U1 with f1 (y) = v − v ′ . This implies
f2 (u − u′ ) = b(v − v ′ ) = bf1 (y) = f2 a(y).
Since f2 is injective, we get u − u′ = a(y). But this yields
a2 (u) − x = a2 (u − u′) = a2 a(y) = 0.
Thus we see that a2 (u) = x, and this implies that δ is a homomorphism.
Exactness
Next, we want to show that Ker(δ) = Im(g0 ): Let x ∈ V0 . To compute δg0 (x)
we need a tuple (g0 (x), w1 , v1 , v2 , u2 , u3) ∈ S. Since g1 b0 = c0 g0 and bb0 = 0 we
can choose (g0 (x), c0 g0 (x), b0 (x), 0, 0, 0). This implies δg0 (x) = 0. Vice versa, let
w ∈ Ker(δ). So there exists some (w, w1, v1 , v2 , u2, 0) ∈ S. Since u2 belongs to the
kernel of a2 and therefore to the image of a, there exists some y ∈ U1 with a(y) = u2 .
We have
bf1 (y) = f2 a(y) = f2 (u2 ) = b(v1 ).
Thus v1 − f1 (y) is contained in Ker(b). This implies that there exists some x ∈ V0
with b0 (x) = v1 − f1 (y). We get
c0 g0 (x) = g1 b0 (x) = g1 (v1 − f1 (y)) = g1 (v1 ) = c0 (w).
162 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Finally, we want to show that Ker(f3 ) = Im(δ): Let (w0 , w1, v1 , v2 , u2 , u3) ∈ S, in
other words δ(w0 ) = u3 . We have
f3 (u3 ) = f3 a2 (u2 ) = b2 f2 (u2 ).
Since f2 (u2) = v2 = b(v1 ), we get b2 f2 (u2) = b2 b(v1 ) = 0. This shows that the
image of δ is contained in the kernel of f3 . Vice versa, let u3 be an element in U3 ,
which belongs to the kernel of f3 . Since a2 is surjective, there exists some u2 ∈ U2
with a2 (u2 ) = u3 . We have b2 f2 (u2 ) = f3 a2 (u2 ) = f3 (u3 ) = 0, and therefore f2 (u2 )
belongs to the kernel of b2 and also to the image of b. Let f2 (u2 ) = b(v1 ) =: v2 .
This implies cg1 (v1 ) = g2 b(v1 ) = g2 f2 (u2 ) = 0. We see that g1 (v1 ) is in the kernel
of c and therefore in the image of c0 . So there exists some w0 ∈ W0 with c0 (w0 ) =
g1 (v1 ). Altogether, we constructed a tuple (w0 , w1 , v1 , v2 , u2 , u3 ) in S. This implies
u3 = δ(w0 ). This finishes the proof of the Snake Lemma.
f1′ g1′
U1′ / V1′ / W1′ / 0
a′ b′ c′
f2′ g2′
0 / U2′ / V2′ / W2′ .
Let δ : Ker(c) → Cok(a) and δ ′ : Ker(c′ ) → Cok(a′ ) be the corresponding connecting
homomorphisms.
Remark: We will mainly formulate results and definitions by using complexes, but
there are always corresponding results and definitions for cocomplexes. We leave it
to the reader to perform the necessary reformulations.
In this lecture course we will deal only with (co)complexes of modules over a K-
algebra A and with (co)complexes of vector spaces over the field K.
A complex C• = (Cn , dn )n∈Z is an exact sequence of A-modules if
Im(dn ) = Ker(dn−1)
164 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Given two complexes C• = (Cn , dn )n∈Z and C•′ = (Cn′ , d′n )n∈Z , a homomorphism
of complexes (or just map of complexes) is given by f• = (fn )n∈Z : C• → C•′
where the fn : Cn → Cn′ are homomorphisms with d′n fn = fn−1 dn for all n.
dn+1 dn
··· / Cn+1 / Cn / Cn−1 / ···
fn+1 fn fn−1
d′n+1 d′n
′ ′
··· / Cn+1 / Cn′ / Cn−1 / ···
The maps C• → C•′ of complexes form a vector space: Let f• , g• : C• → C•′ be such
maps, and let λ ∈ K. Define f• + g• := (fn + gn )n∈Z , and let λf• := (λfn )n∈Z .
If f• = (fn )n : C• → C•′ and g• = (gn )n : C•′ → C•′′ are maps of complexes, then the
composition
g• f• = g• ◦ f• : C• → C•′′
is defined by g• f• := (gn fn )n .
Let C• = (Cn , dn )n be a complex. A subcomplex C•′ = (Cn′ , d′n )n of C• is given by
submodules Cn′ ⊆ Cn such that d′n is obtain via the restriction of dn to Cn′ . (Thus we
require that dn (Cn′ ) ⊆ Cn−1
′
for all n.) The corresponding factor complex C• /C•′
is of the form (Cn /Cn′ , d′′n )n where d′′n is the homomorphism Cn /Cn′ → Cn−1 /Cn−1
′
induced by dn .
Let f• = (fn )n : C•′ → C• and g• = (gn )n : C• → C•′′ be homomorphisms of com-
plexes. Then
f• g•
0 → C•′ − → C•′′ → 0
→ C• −
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 165
the nth homology module (or homology group) of C• . Set H• (C• ) = (Hn (C• ))n .
H n (C • ) = Ker(dn )/ Im(dn−1 )
(One has to check that fn (Im(dn+1 )) ⊆ Im(d′n+1 ) and fn (Ker(dn )) ⊆ Ker(d′n ).)
fn+1 ′
Cn+1 / Cn+1
dn+1 d′n+1
fn
Cn / Cn′
dn d′n
fn−1
′
Cn−1 / Cn−1
dn+1 n d dn−1
Cn+1 −−−
→ Cn −→ Cn−1 −−−→ Cn−2 .
The following picture illustrates the situation. Observe that the homology groups
are highlighted by the thick vertical lines. The marked regions indicate which parts
of Ci and Ci−1 get identified by the map di . Namely di induces an isomorphism
Ci / Ker(di) → Im(di ).
dn+1 - Cn
dn -
dn−1 -
Cn+1 Cn−1 Cn−2
Ci rH rH rH r
HH HH HH
HH HH HH
Ker(di ) r
HH HH
r
HH HH
r
HH HH r
HH HH HH HH HH HH
HH HH HH HH HH HH
Im(di+1 ) r
HH
r
HH
r
HH
r
HH HH HH
0 r Hr Hr Hr
The map dn factors through Ker(dn−1 ) and the map Cn → Ker(dn−1 ) factors through
Cok(dn+1 ). Thus we get an induced homomorphism dn : Cok(dn+1 ) → Ker(dn−1 ).
The following picture describes the situation:
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 167
dn - Ker(dn−1 )
Cok(dn+1)
HH
r
HH
rH HH
HH
r
HH
HH HH
r HH Hr
HH
HH
r
dn
Cok(dn+1) / Ker(dn−1 )
The kernel of dn is just Hn (C• ) and its cokernel is Hn−1 (C• ). Thus we obtain an
exact sequence
n iC n d pC
n−1
0 → Hn (C• ) −→ Cok(dn+1 ) −→ Ker(dn−1 ) −−−→ Hn−1 (C• ) → 0
where iC C
n and pn−1 denote the inclusion and the projection, respectively. The inclu-
sion Ker(dC C
n ) → Cn is denoted by un .
hn := fn − gn = d′n+1sn + sn−1 dn .
Proof. Let C• = (Cn , dn ) and C•′ = (Cn′ , d′n ), and let x ∈ Ker(dn ). We get
since dn (x) = 0. This shows that fn (x) and gn (x) only differ by an element in
Im(d′n+1 ). Thus they belong to the same residue class modulo Im(d′n+1 ).
Proof. As in the proof of Proposition 26.3 we show that fn (x) ∈ Im(d′n+1). This
implies (i). We have Hn (g• )Hn (f• ) = Hn (g• f• ) = Hn (1C• ) and Hn (f• )Hn (g• ) =
Hn (f• g• ) = Hn (1′C• ). Thus Hn (f• ) is an isomorphism.
Recall that the elements in Hn (C• ) are residue classes of the form x + Im(dC
n+1 ) with
x ∈ Ker(dn ). Here we write A• = (An , dn ), B• = (Bn , dn ) and C• = (Cn , dC
C A B
n ).
For x ∈ Ker(dC
n ) set
δn (x + Im(dC A
n+1 )) := z + Im(dn )
−1 B −1
where z ∈ (fn−1 dn gn )(x).
Theorem 26.5 (Long Exact Homology Sequence). With the notation above, we
obtain a well defined homomorphism
is exact.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 169
0 0 0
fn′
′
gn
0 / Ker(dA
n)
/ Ker(dB
n)
/ Ker(dC
n)
fn gn
0 / An / Bn / Cn / 0
dA
n dB
n dC
n
fn−1 gn−1
0 / An−1 / Bn−1 / Cn−1 / 0
′′
fn−1 ′′
gn−1
Cok(dA
n)
/ Cok(dB
n)
/ Cok(dC
n)
/ 0
0 0 0
(The arrows without label are just the canonical inclusions and projections, re-
spectively. By fn′ , gn′ and fn−1
′′ ′′
, gn−1 we denote the induced homomorphisms on the
kernels and cokernels of the maps dA B C
n , dn and dn , respectively.)
The map fn′ is a restriction of the monomorphism fn , thus fn′ is also a monomor-
phism. Since gn−1 is an epimorphism and gn−1 (Im(dB C
n )) ⊆ Im(dn ), we know that
′′
gn−1 is an epimorphism as well.
A A
a = dA
n : Cok(dn+1 ) → Ker(dn−1 ).
170 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Similarly, we obtain b = dB C
n and c = dn . The kernels and cokernels of these homo-
morphisms are homology groups. We obtain the following commutative diagram:
0 0 0
Hn (f• ) Hn (g• )
Hn (A• ) / Hn (B• ) / Hn (C• )
iA
n iB
n iC
n
fn′′
′′
gn
Cok(dA
n+1 )
/ Cok(dB
n+1 )
/ Cok(dC
n+1 )
/ 0
a b c
′
fn−1 ′
gn−1
0 / Ker(dA
n−1 )
/ Ker(dB
n−1 )
/ Ker(dC
n−1 )
pA
n−1 pB
n−1 pC
n−1
Hn−1 (f• ) Hn−1 (g• )
Hn−1 (A• ) / Hn−1 (B• ) / Hn−1 (C• )
0 0 0
Now we can apply the Snake Lemma: For our n we obtain a connecting homomor-
phism
δ : Hn (C• ) → Hn−1 (A• )
which yields the required exact sequence. It remains to show that δ = δn .
Let x = dCn+1 (r) for some r ∈ Cn+1 . Since gn+1 is surjective there exists some
s ∈ Bn+1 with gn+1 (s) = r. We have
gn (y) = x = dC C B
n+1 (r) = dn+1 gn+1 (s) = gn dn+1 (s).
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 171
Therefore y − dBn+1 (s) is an element in Ker(gn ) and thus also in the image of fn . Let
y − dB
n+1 (s) = fn (t) for some t ∈ An . We get
fn−1 dA B B B B B
n (t) = dn fn (t) = dn (y) − dn dn+1 (s) = dn (y) = fn−1 (z).
We have
fn−2 dA B B B
n−1 (z) = dn−1 fn−1 (z) = dn−1 dn (y) = 0.
Since fn−2 is injective, we get dA n−1 (z) = 0.
Now it is not difficult to show that this relation coincides with the relation
Γ(pA ′
n−1 ) ◦ Γ(fn−1 )
−1
◦ Γ(b) ◦ Γ(gn′′ )−1 ◦ Γ(iC
n)
b
′
fn−1
Ker(dA
n−1 )
/ Ker(dB
n−1 )
pA
n−1
Hn−1 (A• )
This implies δ = δn .
172 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
The exact sequence in the above theorem is called the long exact homology
sequence associated to the given short exact sequence of complexes. The homo-
morphisms δn are called connecting homomorphisms.
The connecting homomorphisms are “natural”: Let
f• g•
0 / A• / B• / C• / 0
p• q• r•
f•′ g•′
0 / A′• / B•′ / C•′ / 0
be a commutative diagram with exact rows. Then the diagram
δn
Hn (C• ) / Hn−1 (A• )
Hn (r• ) Hn−1 (p• )
′
δn
Hn (C•′ ) / Hn−1 (A′• )
commutes, where δn and δn′ are the connecting homomorphisms coming from the
two exact rows.
————————————————————————————-
the nth syzygy module of M. This does not depend on the chosen minimal
projective resolution.
Lemma 27.1. If
0→U →P →M →0
is a short exact sequence of A-modules with P projective, then U ∼
= Ω(M) ⊕ P ′ for
′
some projective module P .
Proof. Exercise.
Sometimes we are a bit sloppy when we deal with syzygy modules: If there exists
a short exact sequence 0 → U → P → M → 0 with P projective, we just write
Ω(M) = U, knowing that this is not at all well defined and depends on the choice
of P .
Lemma 27.2. For every module M there is a projective resolution.
(i) There exists a “lifting” of f , i.e. there are homomorphisms fi : Pi → Pi′ such
that
p′i fi = fi−1 pi and ε′ f0 = f ε
for all i;
(ii) Any two liftings f• = (fi )i≥0 and f•′ = (fi′ )i≥0 are homotopic.
174 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
(ii): Assume we have two liftings, say f• = (fi )i≥0 and f•′ = (fi′ )i≥0 . This implies
f ε = ε′ f0 = ε′ f0′
and therefore ε′ (f0 − f0′ ) = 0.
Let ιi : Im(p′i ) → Pi−1
′
be the inclusion and let πi : Pi′ → Im(p′i ) be the obvious
projection. Thus p′i = ιi ◦ πi .
The image of f0 − f0′ clearly is contained in Ker(ε′ ) = Im(p′1 ). Now let s′0 : P0 →
Im(p′1 ) be the map defined by s′0 (m) = (f0 −f0′ )(m). The map π1 is an epimorphism,
and s′0 is a map from a projective module to Im(p′1 ). Thus by the projectivity of P0
there exists a homomorphism s0 : P0 → P1′ such that π1 ◦ s0 = s′0 .
We obtain the following commutative diagram:
ε /
k kk P0 M
k x
s0 kkkkk xxx
kk x f0 −f0′ 0
kkkk xxx s′
kkk {
x 0
u
k
P1′ / Im(p′ )
π1
/ P′
1
/N
ι1 0 ε′
Therefore
Im(fi − fi′ − si−1 pi ) ⊆ Ker(p′i ) = Im(p′i+1 ).
Let s′i : Pi → Im(p′i+1 ) be defined by s′i (m) = (fi − fi′ − si−1 pi )(m).
′
Since Pi is projective there exists a homomorphism si : Pi → Pi+1 such that πi+1 ◦si =
′
si . Thus we get a commutative diagram
j Pi
jjjjjvjvv
j si
jjj vvv fi −fi′ −si−1 pi
jjjjjjj {
v vv s′
u jj /
j i
′
Pi+1 πi+1
Im(p′ ) / P′
i+1 ιi+1 i
Thus p′i+1 si = fi − fi′ − si−1 pi and therefore fi − fi′ = p′i+1 si + si−1 pi , as required.
This shows that f• − f•′ is zero homotopic. Therefore f• = (fi )i and f•′ = (fi′ )i are
homotopic.
By Theorem 27.3 these liftings exist and we have g• f• ∼ 1P• and f• g• ∼ 1P•′ . Thus,
we get a diagram
p3 p 2 p 1
··· /P
2
/P
1
/P
0
~~ ~~ ~~
~ g2 f2 −1P2 ~ g1 f1 −1P1 ~ g0 f0 −1P0
~~ ~~ ~~
~~~ ~~~ ~~~
~ ~ ~
~~~ s2 ~~~ s1 ~~~ s0
~ ~~ ~~
~~ ~~ ~~
~~~ p3 p2 p1
··· / P2 / P1 / P0
such that gi fi − 1Pi = pi+1 si + si−1 pi for all i. (Again we think of P• as a complex
with Pi = 0 for all i < 0.)
Next we apply HomA (−, N) to all maps in the previous diagram and get
HomA (g• f• , N) ∼ HomA (1P• , N).
Similarly, one can show that HomA (f• g• , N) ∼ HomA (1P•′ , N). Now Corollary 26.4
tells us that H n (HomA (g• f• , N)) = H n (HomA (1P• , N)) and H n (HomA (f• g• , N)) =
H n (HomA (1P•′ , N)). Thus
H n (HomA (f• , N)) : H n (HomA (P•′ , N)) → H n (HomA (P• , N))
is an isomorphism.
p3 p2
Proof. If P• = (Pi , pi)i≥0 is a projective resolution of M, then · · · P3 −
→ P2 −
→ P1 is
a projective resolution of Ω(M).
Proof. ...
Let N be any A-module. In the situation of the above lemma, we can apply
HomA (−, N) to the exact sequence η. Since η splits, we obtain an exact sequence
of cocomplexes
0 → HomA (P•′′ , N) → HomA (P• , N) → HomA (P•′ , N) → 0.
Thus we get an exact sequence
27.6. Short exact sequences and the first extension group. Let M and N be
modules, and let
pn+1 pn p2 p1
P• = (· · · −→ Pn −→ · · · −→ P1 −→ P0 )
ε
be a projective resolution of M = Cok(p1 ). Let P0 −→ M be the cokernel map of
p1 , i.e.
p1 ε
P1 −→ P0 −→ M −→ 0
is an exact sequence.
We have
H n (HomA (P• , N)) := Ker(HomA (pn+1 , N))/ Im(HomA (pn , N)).
Let [α] := α + Im(HomA (pn , N)) be the residue class of some homomorphism
α : Pn → N with α ◦ pn+1 = 0.
Clearly, we have
Im(HomA (pn , N)) = {s ◦ pn | s : Pn−1 → N} ⊆ HomA (Pn , N).
with exact rows and where all squares made from solid arrows commute.
f g
Let 0 → X −→Y −→ Z → 0 be a short exact sequence, and let M and N be modules.
Then the connecting homomorphism
is given by h 7→ [η] where η is the short exact sequence h∗ (f, g) induced by h via a
pullback.
η: 0 /X /⋆
/M /0
h
f g
0 / X / Y / Z / 0
If (f, g) is a split short exact sequence, then ψ(f, g) = 0 + Im(HomA (p1 , N)) is the
zero element in Ext1A (M, N): Obviously, the diagram
p1 ε
P 1 / P 0 / M / 0
0
[ 0ε ]
[ 10 ] [0 1]
0 / N / N ⊕M / M / 0
is commutative. This implies
ψ([ 10 ] , [ 0 1 ]) = 0 + Im(HomA (p1 , N)).
In fact, Ext1A (M, N) is a K-vector space and ψ is an isomorphism of K-vector spaces.
So we obtain the following fact:
Lemma 27.11. For an A-module M we have Ext1A (M, M) = 0 if and only if each
short exact sequence
0→M →E→M →0
splits. In other words, E ∼
= M ⊕ M.
27.7. The vector space structure on the first extension group. Let
ηM : 0 → Ω(M) → P0 → M → 0
be a short exact sequence with P0 projective. For i = 1, 2 let
fi gi
ηi : 0 → N −
→ Ei −
→M →0
be short exact sequences.
Take the direct sum η1 ⊕η2 and construct the pullback along the diagonal embedding
M → M ⊕ M. This yields a short exact sequence η ′ .
182 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
u ε
ηM : 0 / Ω(M) / P0 / M / 0
[ αα12 ]
η′ : 0 / N ⊕N / E′ / M / 0
[ 11 ]
η1 ⊕ η2 : 0 / N ⊕ N hf / iE1 ⊕ E2 h g /
iM ⊕M / 0
1 0 1 0
0 f2 0 g2
u ε
ηM : 0 / Ω(M) / P0 / M / 0
[ αα12 ]
η′ : 0 / N ⊕N / E′ / M / 0
[1,1]
η ′′ : 0 / N / E ′′ / M / 0
In other words,
η ′ = [ αα12 ]∗ (ηM ),
η ′′ = [1, 1]∗ (η ′ ).
η1 + η2 := η ′′ .
u ε
ηM : 0 / Ω(M) / P0 / M / 0
αi βi
fi gi
ηi : 0 / N / Ei / M / 0
...
minimal injective resolution
...
Theorem 27.12. Let I • be an injective resolution of an A-module N. Then for any
A-module M we have an isomorphism
Extn (M, N) ∼
A = H n (HomA (M, I • )).
which is “natural in M and N”.
Proof. Exercise.
————————————————————————————-
It often happens that the global dimension of an algebra A is infinite, for example
if we take A = K[X]/(X 2 ). One proves this by constructing the minimal projective
resolution of the simple A-module S. Inductively one shows that Ωi (S) ∼ = S for all
i ≥ 1.
Proposition 28.1. Assume that A is finite-dimensional. Then we have
gl. dim(A) = max{proj. dim(S) | S a simple A-module}.
Similarly, let inj. dim(M) be the minimal j ≥ 0 such that there exists an injec-
tive resolution (Ii , di )i of M with Ij = 0, if such a minimum exists, and define
inj. dim(M) = ∞, otherwise.
We call inj. dim(M) the injective dimension of M.
Theorem 28.2 (No loop conjecture). Let A be a finite-dimensional K-algebra. If
Ext1A (S, S) 6= 0 for some simple A-module S, then gl. dim(A) = ∞.
Conjecture 28.3 (Strong no loop conjecture). Let A be a finite-dimensional K-
algebra. If Ext1A (S, S) 6= 0 for some simple A-module S, then proj. dim(S) = ∞.
————————————————————————————-
Let R(X, Y ) be the subspace of V (X, Y ) which is generated by all vectors of the
form
where x ⊗ y := (x, y) + R(X, Y ). But note that this expression of z is in general not
unique.
Warning
From here on there are only fragments, incomplete proofs or no proofs
at all, typos, wrong statements and other horrible things...
186 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
as A-modules.
Proof. We just prove (i) and leave (ii) as an exercise. Clearly, X ⊗A − is a functor:
We have X ⊗A (g ◦ f ) = (X ⊗A g) ◦ (X ⊗A f ). In particular, X ⊗A 1Y = 1X⊗A Y .
Additivity:
(X ⊗A (f + g))(x ⊗ y) = x ⊗ (f + g)(y)
= x ⊗ (f (y) + g(y))
= (x ⊗ f (y)) + (x ⊗ g(y))
= (X ⊗ f )(x ⊗ y) + (X ⊗ g)(x ⊗ y).
Right exactness:
...
188 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Proof. Exercise.
Proof. ...
29.3. Tor. We will not need any Tor-functors, but at least we will define them and
acknowledge their existence.
Let P• be a projective resolution of A Y , and let XA be a right A-module. This yields
a complex
· · · → X ⊗A P1 → X ⊗A P0 → X ⊗A 0 → · · ·
For n ∈ Z define
TorA
n (X, Y ) := Hn (X ⊗A P• ).
Let P• be a projective resolution of a right A-module XA . Then one can show that
for all A-modules A Y we have
TorA ∼
n (X, Y ) = Hn (P• ⊗A Y ).
(i) Let
η : 0 → XA′ → XA → XA′′ → 0
be a short exact sequence of right A-modules. For every A-module A Y this induces
an exact sequence
··· / TorA (X ′′ , Y )
g 2
ggggggg
ggggg
gg gg ggggg
sgg
TorA
1 (X ′
, Y ) / TorA (X, Y )
1 g
/ TorA (X ′′ , Y )
1
gggg
ggggggg
gggg
ggggg
sggggg
X′ ⊗ Y A
/X⊗ Y
A
/ X ′′ ⊗ Y
A
/ 0
(ii) Let
η : 0 → A Y ′ → A Y → A Y ′′ → 0
190 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
be a short exact sequence of A-modules. For every right A-module XA this induces
an exact sequence
··· / TorA (X, Y ′′ )
g 2
g gggggg
ggggg
gg gg ggggg
sgg
TorA
1 (X, Y ′
) / TorA (X, Y )
1 g
/ TorA (X, Y ′′ )
1
gggg
gggg ggg
gggg
ggggg
sggggg
X ⊗ Y′ A
/ X ⊗A Y / X ⊗ Y ′′
A
/ 0
***********************************************************************
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 191
(We will see that for many module cateogories assumption (ii) is not necessary: If M
is closed under kernels of surjective homomorphisms and f : U → V is an injective
homomorphism which is irreducible in M, then Cok(f ) is indecomposable. Similarly,
if M is closed under cokernels of injective homomorphisms and g : V → W is a
surjective homomorphism which is irreducible in M, then Ker(g) is indecomposable.)
Let (Γ0 , Γ1 ) and (Γ0 , Γ2 ) be two “quivers” with the same set Γ0 of vertices, but with
disjoint sets Γ1 and Γ2 of arrows. Then (Γ0 , Γ1 , Γ2 ) is called a biquiver. The arrows
in Γ1 are the 1-arrows and the arrows in Γ2 the 2-arrows. To distinguish these two
types of arrows, we usually draw dotted arrows for the 2-arrows. (Thus a biquiver
Γ is just an oriented graph with two types of arrows: The set of vertices is denoted
by Γ0 , the “1-arrows” are denoted by Γ1 and the “2-arrows” by Γ2 .)
Let M be a module category, which is closed under direct summands. Then the
Auslander-Reiten quiver of M is a biquiver ΓM which is defined as follows: The
vertices are the isomorphism classes of indecomposable modules in M. For a module
V we often write [V ] for its isomorphism class. There is a 1-arrow [V ] → [W ] if
and only if there exists an irreducible homomorphism V → W in M, and there is a
2-arrow from [W ] to [U] if and only if there exists an Auslander-Reiten sequence
0 → U → V → W → 0.
The Auslander-Reiten quiver is an important tool which helps to understand the
structure of a given module category.
Later we will modify the above definition of an Auslander-Reiten quiver and also
allow more than one arrow between two given vertices.
V′ or V ′
b1 } b2 }>
} g′ } g′
} }
f ~} g f } g
0 / U / V / W / 0 0 / U / V / W / 0
Now we prove the converse of the Bottleneck Lemma (but note the different assump-
tion on M):
f g
0→U −
→V −
→W →0
be a non-split short exact sequence in M such that the following hold: For every
homomorphism g ′ : V ′ → W in M there exists a homomorphism b1 : V ′ → V with
gb1 = g ′ , or there exists a homomorphism b2 : V → V ′ with g ′b2 = g. Then it follows
that f is irreducible in M.
f g
0 / U / VO / WO / 0
f2 g′
f1 g1
0 / U / V′ / W′ / 0
It follows that the pair (f2 , g1 ) is the pullback of (g, g ′). Our assumption implies that
for g ′ there exists a homomorphism b1 : W ′ → V with gb1 = g ′ or a homomorphism
b2 : V → W ′ with g ′ b2 = g.
If b1 exists with gb1 = g ′ = g ′1W ′ , then the pullback property yields a homomorphism
h : W ′ → V ′ with b1 = f2 h and 1W ′ = g1 h. In particular we see that g1 is a split
196 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
f g f g
0 / U /V / W / 0 or 0 / U /V / W / 0
} }>
f′
} f′
}
} a1 } a2
~} }
′ ′
V V
Corollary 30.8. Let M be a module category which is closed under cokernels of
injective homomorphisms. If
f g
0→U −
→V −
→W →0
is a short exact sequence in M, and if g is irreducible in M, then U is indecompos-
able.
Corollary 30.9. Let M be a module category which is closed under cokernels of
injective homomorphisms. If
f1 g1
0→U −
→ V1 −
→ W1 → 0
f2 g2
0→U −
→ V2 −
→ W2 → 0
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 197
and
t
M
g = [g1 , . . . , gt ] : N(λi ) → N(n + 1)
i=1
with fi : N(n) → N(λi ) and gi : N(λi ) → N(n + 1) homomorphisms and
t
X
h = gf = gi fi .
i=1
Since h is injective, we have h(e1 ) 6= 0. Thus there exists some i with gi fi (e1 ) 6= 0.
This implies that gi fi is injective, and therefore fi is injective.
If λi > n + 1, then gi (e1 ) = 0, a contradiction. (Note that gi fi (e1 ) 6= 0 implies
gi (e1 ) 6= 0).) Thus λi is either n or n + 1. If λi = n, then fi is an isomorphism, if
λi = n + 1, then gi is an isomorphism. In the first case, set
f ′ = [0, . . . , 0, fi−1, 0, . . . , 0] : N(λ) → N(n).
We get f ′ f = 1N (n) , thus f is a split monomorphism.
In the second case, set
g ′ = t [0, . . . , 0, gi−1, 0, . . . , 0] : N(n + 1) → N(λ).
It follows that gg ′ = 1N (n+1) , so g is a split epimorphism. This proves part (i). Part
(ii) is proved similarly.
Next, let f : N(n) → N(m) be an irreducible homomorphism. We proved already
before that every irreducible homomorphism has to be either injective or surjective.
If m ≥ n + 2, then f factors through N(n + 1) as f = f2 f1 where f1 is injective but
not split, and f2 is not surjective, a contradiction. Similarly, if m ≤ n − 2, then f
factors through N(n − 1) as f = f2 f1 where f1 is not injective, and f2 is surjective
but not split, again a contradiction. This proves (iii).
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 199
dim Im(f − at g) ≤ t − 1.
30.6. Exercises. Use the Converse Bottleneck Lemma to show that for n ≥ 1 the
short exact sequence
————————————————————————————-
Proof. ...
Lemma 31.3. Let f : X → Y be a homomorphism, and let u : Y → Z be a
monomorphism. Then
Ker(HomA (N, f )) = Ker(HomA (N, u ◦ f )).
Proof. We have
n
! n
X X
ηXY αi ⊗ yi = ηXY (αi ⊗ yi)
i=1 i=1
Xn
= ρyi ◦ αi .
i=1
2 α1 3
.. 5
4
. n
M [ρy1 ,...,ρyn ]
αn
X −−−→ A A −−−−−−→ Y
i=1
To prove the other direction, let P be a finitely generated projective module, and
assume h = g ◦ f for some homomorphisms h : X → Y , f : X → P and g : P → Y .
There exists a module C such that P ⊕ C is a free module of finite rank. Thus
without loss of generality we can assume that PP is free of finite rank. Let e1 , . . . , en
be a free generating set of P . Then f (x) = i αi (x)ei for some αi (x) ∈ A. This
defines some homomorphisms αi : X → A A. Set yi := g(ei). It follows that
!
X X
ηXY αi ⊗ yi (x) = αi (x)yi
i i
X
= αi (x)g(ei )
i
!
X
=g αi (x)ei
i
= (g ◦ f )(x) = h(x).
This finishes the proof.
202 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Proof. Exercise.
= A Y and HomA (A A, A Y ) ∼
= AA , AA ⊗A Y ∼
Recall that (A A)∗ = HomA (A A, A A) ∼ =
A Y .
Thus we have isomorphisms AA ⊗A Y → Y , α ⊗ y 7→ α(1)y and Y → HomA (A A, Y ),
y 7→ ρy . Composing these yields a map α ⊗ y 7→ ρα(1)y = ρy ◦ α. We have
ρα(1)y (a) = aα(1)y = α(a)y = (ρy ◦ α)(a).
Proof. We know that for all modules N the functor HomA (−, N) is left exact. It is
also clear that D is contravariant and exact. Thus ν is right exact.
Now let P be finitely generated projective. It follows that D(P ∗ ) is injective: With-
out loss of generality assume P = A A. Then P ∗ = AA and HomK (AA , K) is injec-
tive.
where i := Adj is the isomorphism given by the adjunction formula Theorem 29.7.
We know by Lemma 31.6 that ξXY is bijective, provided X is finitely generated
projective.
Using this, we obtain a commutative diagram
D HomA (P1 , N) / D HomA (P0 , N) / D HomA (M, N) / 0
ξP1 N ξP0 N ξM N
µ: HomA (N, ν(P1 )) / HomA (N, ν(P0 )) / HomA (N, ν(M))
whose first row is exact and whose second row is a complex. This is based on the
facts that the functor D is exact, and the functor HomA (−, N) is left exact.
Thus we can apply Lemma 31.2 to this situation and obtain
H(µ) = Ker(ξM N )
= Ker(D(ηM N ))
= {α ∈ D HomA (M, N) | α(Im(ηM N )) = 0}.
(If f : V → W is a K-linear map, then the kernel of f ∗ : DW → DV consists of all
g : W → K such that g ◦ f = 0. This is equivalent to g(Im(f )) = 0.)
Recall that
ξM N = Adj ◦ D(ηM N ).
If M is finitely generated, then Lemma 31.4 and Lemma 31.5 yield that
Im(ηM N ) = HomA (M, N)P .
This implies
{α ∈ D HomA (M, N) | α(Im(ηM N )) = 0} = DHomA (M, N).
This finishes the proof of Theorem 31.1.
The isomorphism
DHomA (M, N) → Ext1A (N, τ (M))
is “natural in M and N”:
Let M be a finitely presented A-module, and let f : M → M ′ be a homomorphism.
This yields a map
D HomA (f, N) : D HomA (M, N) → D HomA (M ′ , N)
and a homomorphism τ (f ) : τ (M) → τ (M ′ ). Now one easily checks that the diagram
...
Let
f g
0→X−
→Y −
→Z→0
206 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Now let
f g
η : 0 → τ (M) −
→Y −
→ M → 0.
be the short exact sequence we constructed above.
Lemma 31.8. g is a right almost split homomorphism.
where φM M (ε) = η and DHomA (M, h)(ε) = 0. This implies Ext1A (h, τ (M))(η) = 0.
Note that the map Ext1A (h, τ (M)) sends a short exact sequence ψ to the short exact
sequence h∗ (ψ) induced by h via a pullback.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 207
So we get h∗ (η) = 0 for all h : N → M which are not split epimorphisms. In other
words, g is a right almost split morphism.
f g
Lemma 31.9. Let 0 → X − → Y − → Z → 0 be a non-split short exact sequence.
Assume that g is right almost split and that EndA (X) is a local ring. Then f is left
almost split.
Proof. Use Skript 1, Cor. 11.5 and the dual statement Cor. 11.10 and
Skript 1, Lemma 11.6 (Converse Bottleneck Lemma) and the dual state-
ment Lemma 11.11. Furthermore, we need Skript 1, Cor. 11.3 and Cor.
11.8.
Proof. Clearly, proj. dim(M) ≤ 1 if and only if Ω2 (M) = 0. By the Lemma above
this is equivalent to HomA (D(AA ), τ (M)) = 0. But we know that each indecom-
posable injective A-module is isomorphic to a direct summand of D(AA ). (Let I be
an indecomposable injective A-module. Then D(I) is an indecomposable projective
right A-module. It follows that D(I) is isomorphic to a direct summand of AA . Thus
I∼= DD(I) is a direct summand of D(AA ).) This finishes the proof.
• If
f g
0→X− →Y − →Z→0
is an Auslander-Reiten sequence, then f is a source map for X, and g is a
sink map for Z.
• If X is an indecomposable module which is not injective, then there exists a
source map for X.
• If Z is an indecomposable module which is not projective, then there exists
a sink map for Z.
Lemma 31.18. (i) If f : X → Y is a source map, then X is indecomposable;
(ii) If g : Y → Z is a sink map, then Z is indecomposable.
(i) For each arrow [X] → [Y ] in ΓA with {[X], [Y ]}∩Γ0 6= ∅ we have {[X], [Y ]} ⊆
Γ0 ;
(ii) If [X] and [Y ] are vertices in Γ, then there exists a sequence
([X1 ], [X2 ], . . . , [Xt ])
of vertices in Γ with [X] = [X1 ], [Y ] = [Xt ], and for each 1 ≤ i ≤ t − 1 there
is an arrow [Xi ] → [Xi+1 ] or an arrow [Xi+1 ] → [Xi ].
L
Corollary 31.26. Let X → Y be a source map, and let Y = ti=1 Yini where Yi is
indecomposable, ni ≥ 1 and Yi 6∼
= Yj for all i 6= j. Then there are precisely t arrows
in ΓA starting at [X], namely [X] → [Yi ], 1 ≤ i ≤ t.
Lemma 31.27. A vertex [X] is a source in ΓA if and only if X is simple projective.
f g
Let 0 → X − → Y − → Z → 0 be an Auslander-Reiten sequence in mod(A). Thus,
by definition f and g are irreducible. We proved already that X and Z have to be
indecomposable (Skript 1). It follows that we get a commutative diagram
f′ g′
0 / X / E / τ −1 (X) / 0
h h′
f g
0 / X / Y / Z / 0
′
where h and h are isomorphisms.
Here τ −1 (X) := TrD(X).
Source maps are unique in the following sense: Let X be an indecomposable A-
module which is not injective, and let f : X → Y and f : X → Y ′ be source maps.
By g : Y → Z and g ′ : Y ′ → Z ′ we denote the projections onto the cokernel of f and
f ′ , respectively. Then we get a cimmutative diagram
f′ g′
0 / X / Y′ / Z′ / 0
h h′
f g
0 / X / Y / Z / 0
where h and h′ are isomorphisms.
Dually, sink maps are unique as well.
(i) A is connected;
(ii) For any indecomposable projective A-modules P 6∼ = P ′ there exists a tu-
ple (P1 , P2 , . . . , Pm ) of indecomposable projective modules such that P1 =
P , Pm = P ′ and for each 1 ≤ i ≤ m − 1 we have HomA (Pi , Pi+1 ) ⊕
HomA (Pi+1 , Pi) 6= 0;
(iii) For any simple A-modules S and S ′ there exists a tuple (S1 , S2 , . . . , Sm ) of
simple modules such that S1 = S, Sm = S ′ and for each 1 ≤ i ≤ m − 1 we
have Ext1A (Si , Si+1 ) ⊕ Ext1A (Si+1 , Si ) 6= 0;
(iv) If A = A1 × A2 then A1 = 0 or A2 = 0;
(v) 0 and 1 are the only central idempotents in A.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 215
Proof. Exercise. Hint: If ExtA (Si , Sj ) 6= 0, then there exists a non-split short exact
sequence
f g
0 → Sj −
→E− → Si → 0.
Then there exists an epimorphism pi : Pi → Si . This yields a homomorphism
p′i : Pi → E such that gp′i = pi . Clearly, h′ has to be an epimorphism. (Why?)
Let pj : Pj → Sj be the obvious epimorphism. Then there exists an epimorphism
p′j : Pj → E such that f pj = p′j . Next, there exists a non-zero homomorphism
q : Pj → Pi such that pi q = f pj .
Theorem 31.31 (Auslander). Let A be a finite-dimensional connected K-algebra,
and let C be a component of the Auslander-Reiten quiver of A. Assume that there
exists some b such that all indecomposable modules in C have length at most b. Then
C is a finite component and it contains all indecomposable A-modules. In particular,
A is representation-finite.
Proof. (a): Let X be an indecomposable A-module such that there exists a non-zero
homomorphism h : X → Y for some [Y ] ∈ C. We claim that [X] ∈ C: Let
t1
M
(1) (1) (1) (1)
g = [g1 , . . . , gt1 ] : Yi →Y
i=1
(1)
be the sink map ending in Y , where is indecomposable for all 1 ≤ i ≤ t1 . If h
Yi
is a split epimorphism, then h is an isomorphism and we are done. Thus, assume
h0 := h is not a split epimorphism. It follows that there exists a homomorphism
(1)
f1 Mt1
(1) .. (1)
f = . :X→ Yi
(1) i=1
ft1
such that
t1
X
(1) (1) (1) (1)
h0 = g f = gi fi : X → Y.
i=1
(1) (1) (1)
Since h0 6= 0, there exists some 1 ≤ i1 ≤ t1 such that gi1 ◦ fi1 6= 0. Set h1 := fi1
(1)
and h′1 := gi1 . Next, assume that for each 1 ≤ k ≤ n − 1 we already constructed a
non-invertible homomorphism
(k) (k−1)
h′k : Yik → Yik−1 ,
(k) (0)
where [Yik ] ∈ C and Yi0 := Y , and a homomorphism
(k)
hk : X → Yik
such that h′1 ◦ · · · ◦ h′k ◦ hk 6= 0. So we get the following diagram:
X X ··· X X
hn−1 hn−2 h1 h0
(n−1) (n−2) (1)
Yin−1 / Yin−2 / ··· / Yi1 / Y
h′n−1 h′n−2 h′2 h′1
216 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Thus the 1st Brauer-Thrall Conjecture says that bounded representation type implies
finite representation type. There exists a completely different proof of the 1st Brauer-
Thrall conjecture due to Roiter, using the Gabriel-Roiter measure.
Conjecture 31.33 (2nd Brauer-Thrall Conjeture). Let A be a finite-dimensional
algebra over an infinite field K. If A is representation-infinite, then there exists
some d ∈ N such that the following hold: For each n ≥ 1 there are infinitely many
isomorphism classes of indecomposable A-modules of dimension nd.
Theorem 31.34 (Smalø). Let A be a finite-dimensional algebra over an infinite field
K. Assume there exists some d ∈ N such that there are infinitely many isomorphism
classes of indecomposable A-modules of dimension d. Then for each n ≥ 1 there are
infinitely many isomorphism classes of indecomposable A-modules of dimension nd.
Thus to prove Conjecture 31.33, the induction step is already known by Theorem
31.34. Just the beginning of the induction is missing...
Conjecture 31.33 is true if K is algebraically closed. This was proved by Bautista
using the well developed theory of representation-finite algebras over algebraically
closed fields.
Proof. Exercise.
(i) f is irreducible;
(ii) f ∈ radA (X, Y ) \ rad2A (X, Y ).
Proof. Exercise.
Lemma 31.39. Assume K is algebraically closed. If X is an indecomposable A-
module, then F (X) ∼
= K.
Proof. Exercise.
Theorem 31.40. Let M and N be indecomposable A-modules. Let g : Y → N be a
sink map for N. Write
Y = Mt ⊕ Y ′
with t maximal. Thus g = [g1 , . . . , gt , g ′] where gi : M → N, 1 ≤ i ≤ t and g ′ : Y ′ →
N are homomorphisms. Then the following hold:
(i) The residue classes of g1 , . . . , gt in IrrA (M, N) form a basis of the F (M)op -
vector space IrrA (M, N);
(ii) We have
dimK (IrrA (M, N))
t = dimF (M )op (IrrA (M, N)) = .
dimK (F (M))
(iii) The residue classes of f1 , . . . , fs in IrrA (M, N) form a basis of the F (N)-
vector space IrrA (M, N);
(iv) We have
dimK (IrrA (M, N))
s = dimF (N ) (IrrA (M, N)) = .
dimK (F (N))
Proof. (a): First we show that the set {g1 , . . . , gt } is linearly independent in the
F (M)op -vector space IrrA (M, N):
Assume
t
X
(2) λi ⋆ g i = 0
i=1
By Equation (2) we know that this map is contained in rad2A (M, N).
" λ1 #
.
Clearly, .. is a split monomorphism, since
λt
0
" λ1 #
..
[λ−1
1 , 0, . . . , 0] ◦ . = 1M .
λt
0
P
Using Lemma 31.24 this implies that ti=1 giλi is irreducible and can therefore not
be contained in rad2A (M, N), a contradiction.
(b): Next, we show that {g1 , . . . , gt } generates the F (M)op -vector space IrrA (M, N):
Let u : M → N be a homomorphism with u ∈ radA (M, N). We have to show that
u := u + rad2A (M, N) is a linear combination of g1 , . . . , gt .
Since g is a sink map and u is not a split epimorphism, we get a commutative
diagram
2 u1 3
6 .. 7
4 .mM
5
ut
m m m
u′
m u
vm m
Mt ⊕Y′ /N
[g1 ,...,gt ,g ′ ]
Pt
such that u = i=1 gi u i + g ′ u ′ .
We know that g ′ ∈ radA (Y ′ , N), since g ′ is just the restriction of the sink map g
to a direct summand Y ′ of Y . Thus g ′ is irreducible or g ′ = 0. Furthermore, M is
indecomposable and Y ′ does not contain any direct summand isomorphic to M. So
u′ ∈ radA (M, Y ′ ). Thus implies g ′ u′ ∈ rad2A and therefore g ′u′ = 0. It follows that
t
X t
X
u= u i ⋆ gi + g ′ u′ = u i ⋆ gi .
i=1 i=1
Proof. Let t be maximal such that Y = M t ⊕ Y ′ for some module Y ′ . Then we get
dimK IrrA (M, Z) dimK IrrA (τ (Z), M)
t= = .
dimK F (M) dimK F (M)
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 221
Proof. ...
∅ = −1 Γ ⊆ 0 Γ ⊆ · · · ⊆ n−1 Γ ⊆ n Γ ⊆ · · ·
of subsets of Γ0 .
By definition −1 Γ = ∅. For n ≥ 0, if n−1 Γ is already defined, then let n Γ be the set
of all vertices z of Γ such that z − ⊆ n−1 Γ.
222 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Note that a translation quiver can have multiple arrows between two vertices.
If Γ = (Γ0 , Γ1 , s, t, τ, σ) is a translation quiver, then τ is called the translation of Γ.
The vertices in Γ0 \Γ′0 are the projective vertices, and Γ0 \τ (Γ′0 ) are the injective
vertices. If Γ does not have any projective or injective vertices, then Γ is stable.
A translation quiver Γ is preprojective if the following hold:
the mesh relation associated to z, where the sum runs over all arrows ending in
z. This is an element in the path category KΓ.
The mesh category KhΓi of the translation quiver Γ is by definition the factor
category of KΓ modulo the ideal generated by all mesh relations ρz where z runs
through the set Γ′0 of all non-projective vertices of Γ.
Example: Let Γ be the following translation quiver:
w
β ~~? @@@
~ @@δ
~~~ @@
~~ @
o_ _ _ _ _ _ _ y
? v @@ ~? ??? ξ
α @@ γ ǫ ~~~ ??
@@ ~
@@ ~~ ??
~
u o
_ _ _ _ _ _ _ x o
_ _ _ _ _ _ _z
This is a translation quiver without multiple arrows. The dashed arrows describe τ ,
they start in some z and end in τ (z). Thus we have three projective vertices u, v, w
224 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
For example, in the path category KΓ we have dim HomKΓ(u, y) = 2. But in the
mesh category KhΓi, we obtain HomKhΓi (u, y) = 0.
Assume that Γ = (Γ0 , Γ1 , s, t, τ, σ) is a translation quiver without multiple arrows.
A function
d : Γ0 ∪ Γ1 → N 1
is a valuation for Γ if the following hold:
Proof. We have IrrA (X, X) = 0 for every indecomposable A-module X. (Recall that
every irreducible map between indecomposable modules is either a monomorphism
or an epimorphism.) Thus the quiver ΓA does not have any loops. If Z is an
indecomposable non-projective module, then the skew fields F (τA (Z)) and F (Z) are
isomorphic, and dimK IrrA (τA (Z), Y ) = dimK IrrA (Y, Z) for each indecomposable
module Y . This shows that ΓA is locally finite, and that the conditions (T1), (T2),
(T3) and (V2) are satisfied. Since IrrA (X, Y ) is an F (X)op -F (Y )-bimodule, also
(V1) holds.
...
The quiver (Γ, d)e has the same vertices as (Γ, d), and also the same translation τ .
For every arrow α : x → y in Γ, we get a sequence of d(x → y) arrows αi : x → y
where 1 ≤ i ≤ d(α). (Thus the arrows in (Γ, d)e starting in x and ending in y are
enumerated, there is a first arrow, a second arrow, etc.) Now σ sends the ith arrow
y → z to the ith arrow τ (z) → y provided z is a non-projective vertex.
Next, we look for the source map starting in N: We have dimK IrrA (N, Q) =
dimK IrrA (M, N) = 2 and dimK F (Q) = 1. Thus N → Q ⊕ Q is a source map.
We obtain an Auslander-Reiten sequence 0 → N → Q ⊕ Q → R → 0 where
R = [ C0 ].
The modules τ −1 (M) and τ −1 (N) are injective, thus the following is the Auslander-
Reiten quiver of A:
dN = 2 [ C ] o_ _ _ _ _ _ _ _ _ [ C0 ]
>} C DD z<
2 }}} DD 2 2 zzz
DDD z
}} D" zz
}} C/R z
dM = 1 [ R ] o_ _ _ _ _ _ _ _
0 C
Note that the valuation of the vertices remains constant on τ -orbits (and τ −1 -orbits),
so it is enough to display them only once per orbit.
(b): Next, let
k K
A= ⊂ M2 (K)
0 K
√
where k ⊂ K is a field extension of dimension three, e.g. k = Q and K = Q( 3 2).
The indecomposable projective A-modules are
M = Ae11 = [ k0 ] and N = Ae22 = [ K
K].
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 227
In this case there are 6 indecomposable A-modules, and the Auslander-Reiten quiver
ΓA looks like this:
3 o_ _ _ _ _ _ _ 3 0
dN = 3 @ 1 << @ 2
o_ _ _ _ _ _ _
<< @ 1
<< 3 << 3
3 << 3 << 3
<< <<
1 o_ _ _ _ _ _ _ 2 o_ _ _ _ _ _ _ 1
dM = 1 0 1 1
(c): Here is the Auslander-Reiten quiver of the algebra A = KQ/I where Q is the
quiver
1=
==
a
==c
==
2 == 3
==
=
b ==
d
4
0
1 0 o_ _ _ _ _ _ _ _ 0 0 1 o_ _ _ _ _ _ _ _ 1 1 0
> 1 BB > 0 BB > 0 BB
||| BB ||| BB ||| BB
|| BB || BB || BB
| B B | BB | BB
|| || ||
0 0 1 1
0 0 o_ _ _ _ _ _ _ _ 1 1 o_ _ _ _ _ _ _ _ 1 1 o_ _ _ _ _ _ _ _ 0 0
1 1
F 00 BB 0 0
00 |> F 00 F
00 00 BBB |
|
| 00
00 00 BBB || 00
00 0 || 00
00 000 1 1 1 00
00 00 1 00
00 00 00
00 00 00
0 0 0
0 0 1
0 1 o_ _ _ _ _ _ _ _ 1 0 o_ _ _ _ _ _ _ _ 0 1
1 0 0
Here the Yi are non-injective modules, the Ii are injective, and the Pi are projective.
The sink map ending in X is of the form Y → X where
r
M s
M
dY X /dYi dIi X /dIi
Y = Yi i ⊕ Ii .
i=1 i=1
We get
r
M t
M
dXτ −1 (Y ) /dτ −1 (Y dXPi /dPi
Z= τA−1 (Yi) A i A i) ⊕ Pi .
i=1 i=1
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 229
Knitting preparations
where rij ≥ 1, and the Rij are indecomposable such that Rik ∼
= Ril if and
only if k = l.
(iii) For each 1 ≤ i ≤ n determine dPi = dimK F (Pi ).
Note that
dRij Pi = dimK IrrA (Rij , Pi ) = rij · dRij
where rij = rRij Pi . Furthermore, we know that
= EndA (Pi / rad(Pi )) ∼
F (Pi ) = EndA (Pi )/ rad(EndA (Pi )) ∼ = EndA (Si )
where Si is the simple A-module with Si ∼
= Pi / rad(Pi ).
(a0 ) Define 0 ∆: Let 0 ∆ be the quiver (without arrows) with vertices [S] where
S is simple projective.
(b0 ) Add projectives: For each [S] ∈ 0 ∆, if S ∼ = Rij for some i, j, then (if it
wasn’t added already) add the vertex [Pi ] with valuation dPi , and add an
arrow [S] → [Pi ] with valuation dSPi = rSPi · dS . We denote the resulting
quiver by 0 ∆′ .
(c0 ) Translate the non-injectives in 0 ∆: For each [S] ∈ 0 ∆ with S non-
injective, add the vertex [τA−1 (S)] to 0 ∆′ with valuation dτ −1 (S) = dS , and for
A
each arrow [S] → [Y ] constructed so far add an arrow [Y ] → [τA−1 (S)] to 0 ∆′
with valuation dY τ −1 (S) = dSY . We denote the resulting quiver by 0 ∆′′ .
A
Note that any source map starting in a simple projective module S is of the form
S → P where P is projective. (Proof: Assume there is an indecomposable non-
projective module X and an arrow [S] → [X]. Then there has to be an arrow
[τA (X)] → [S], a contradiction since [S] is a source in (ΓA , dA ).) Thus we get P
from the data collected in (i), (ii) and (iii). More precisely, we have
n
M dSPi /dPi
P = Pi ,
i=1
(an ) Define n ∆: Let n ∆ be the full subquiver of n−1 ∆′′ with vertices [X] such
that all direct predecessors of [X] in n−1 ∆′′ are contained in n−1 ∆, and if
[X] is a vertex with X ∼ = Pi projective, then we require additionally that
[Rij ] ∈ n−1 ∆ for all j.
(bn ) Add projectives: For each [X] ∈ n ∆, if X ∼ = Rij for some i, j, then (if
it wasn’t added already) add the vertex [Pi ] to n−1 ∆′′ with valuation dPi ,
and add an arrow [X] → [Pi ] to n−1 ∆′′ with valuation dXPi = rXP · dX . We
denote the resulting quiver by n ∆′ .
(cn ) Translate the non-injectives in n ∆\ n−1 ∆: For each [X] ∈ n ∆\ n−1 ∆ with
X non-injective, add the vertex [τA−1 (X)] to n ∆′ with valuation dτ −1 (X) = dX ,
A
and for each arrow [X] → [Y ] constructed to far add an arrow [Y ] → [τA−1 (X)]
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 231
The algorithm stops if n ∆ \ n−1 ∆ is empty for some n. It can happen that the
algorithm never stops.
Define [ [
∞∆ = n∆ and ∞∆ = n ∆.
n≥0 n≥0
Let [X] ∈ n ∆, and let [X] → [Zi ], 1 ≤ i ≤ t be the arrows in n ∆′ starting in [X].
Then the corresponding homomorphism
t
M dXZi /dZi
X→ Zi
i=1
C ⊆ ∞ ∆,
Proof. (a): By construction, for each [X] ∈ n ∆′′ we have τAn (X) is projective for
some n ≥ 0.
(b): The quiver n ∆ has no oriented cycles: One shows by induction on n that if
[X] → [Y ] is an arrow in n ∆, then there exists a unique t ≤ n such that [Y ] ∈
t ∆ \ t−1 ∆ and [X] ∈ t−1 ∆. The result follows.
(c): Let [X] ∈ n ∆. Then [X] has a direct predecessor in n ∆ if and only if X is not
in 0 ∆.
Often knitting does not work. For example, we cannot even start with the knitting
procedure, if there is no simple projective module. Furthermore, if an indecompos-
able projective module Pi is inserted such that an indecomposable direct summand
of rad(Pi ) does not show up in some step of the knitting prodedure, then we are
doomed and cannot continue.
But the good news is that in many interesting situations knitting does work. Here
are the two most important situations: Path algebras and directed algebras. In fact,
using covering theory, one can use knitting to construct the Auslander-Reiten quiver
of any representation-finite algebra (provided the characteristic of the ground field
is not two).
The dual situation: Obviously, there is also a “dual knitting algorithm” by starting
with the simple injective A-modules. As a knitting preparation one needs to de-
compose Ii / soc(Ii ) into a direct sum of indecomposables, and one needs the values
dIi = dimK F (Ii). If C is a component of Γ(A) which is obtained by the dual knitting
algorithm, then C is a preinjective component.
Lemma 31.47. Let Q be a finite connected quiver without oriented cycles. Then
the following hold:
Proof. Exercise.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 233
1 0 0 o_ _ _ _ _ _ _ _ 0 1 1 o_ _ _ _ _ _ _ _ 1 0 0
<x 1 FF < 1 FF x<
0
x x FF x xx FF xx
x FFF x FFF x
xx F" xx F" xx
xx xx xx
0 0 0 o_ _ _ _ _ _ _ _ 1 1 1 o_ _ _ _ _ _ _ _ 1 1 1
1 33FF < E 2 3F < E 1 3F
33 FFF x xx 33 FFF x xx 33 FFF
33 FFF x 33 FF x 33 FF
xx 33 FF" xx 33 FF"
33 " xx xx
3
33 0 1 0 o_ _ _ _ _ _ _33 _ 1 0 1 o_ _ _ _ _ _ _333 _ 0
1 0
33 1 33 1 33 0
33 33 33
33 33 33
0 0 1 o_ _ _ _ _ _ _ _ 1 1 0 o_ _ _ _ _ _ _ _ 0 0 1
1 1 0
3 3 3
33
3 33
3
0 1 1 o_ 3_3 _ _ _ _ _ _ 1 0 0 o_ 3_3 _ _ _ _ _ _ 0 1 0
1 F 3 1 F 3 0
xx< FF 333 xx< FF 333 xx<
x FFF 33 x FFF 33 xx
xxx x
FF 3 xxx FF 3 xxx
x x " x " x
0 0 1 o_ _ _ _ _ _ _ _ 1 1 1 o_ _ _ _ _ _ _ _ 1 1 0
< 1 FF < 2 FF < 1
xxx FF x xx FF x xx
x FF x FF x
xx FF xx FF xx
xx " xx " xx
0 0 1 o_ _ _ _ _ _ _ _ 0 0 0 o_ _ _ _ _ _ _ _ 1 1 1
0 1 1
and let A = KQ/I where I is generated by the path aa. Clearly, A is finite-
dimensional, and has two simple modules, whch we denote by 1 and 2. The
234 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
o_ _ _ _ _ _ _ _ _ 2
z= 2 GGG > 1 <<
zzz GG | || <<
G ||
z zz GG
G | <<
<<
zz G# ||
z
2 o_ _ _ _ _ _ _ _ _ 2 o_ _ _ _ _ _ _ 2
1 BB < 1 2?? A
BB yy ??
BB y y ??
BB yy ??
yy ?
2
1 2 o_ _ _ _ _ _ _ _ 2
2
? 1D
D @
D
DD
DD
D
DD
!
1 o_ _ _ _ _ _ _ _ _ _ 22
1
Note that this time, we did not display the dimension vectors of each indecomposable
module. Instead we used the composition factors 1 and 2 to indicate how the modules
look like. For example, the 4-dimensional A-module
2
1 2
1
has a simple top 2, its socle is isomorphic to 1 ⊕ 1. Note also that one has to identify
the two vertices on the upper left with the two vertices on the upper right. Thus ΓA
has in fact just 7 vertices. Sometimes one displays certain vertices more than once,
in order to obtain a nicer and easier to understand picture.
Clearly, ΓA does not contain a preprojective component. We have a simple projective
module, namely 1. So 0 ∆ = {1}. But then we see that 1 ∆ \ 0 ∆ = ∅. So there is just
one reachable vertex in ΓA .
We constructed ΓA “by hand”. In other words, our methods are not yet developed
enough to prove that this is really the Auslander-Reiten quiver of A.
(c): Let A be the path algebra of the quiver
3
2
1
1
1 o_ _ _ _ _ _ _ _ 3 2 o_ _ _ _ _ _ _ _ 4 4 o_ _ _ _ _ _ _ · · ·
2 3 5 A
@ == @ == @ <<
=== === <
== == <<
= = <<
<<
0 2 3
1 o
_ _ _ _ _ _ _ _ 2 o
_ _ _ _ _ _ _ _ 4 o
_ _ _ _ _ _ _ _ ·@ · ·<
@ 1 == @ 3 == @ 4 ==
<<
== == ==
<<
<<
== == ==
== == == <<
<<
0
0 o_ _ _ _ _ _ _ _ 2 1 o_ _ _ _ _ _ _ _ 3 3 o_ _ _ _ _ _ _ _ 5 4 o_ _ _ _ _ _ _ · · ·
1 2 4 5 A
== @ == @ == @ <<
== === === <
== == == <<
== = = <<
<<
1
1 o_ _ _ _ _ _ _ _ 3 2 o_ _ _ _ _ _ _ _ 4 4 o_ _ _ _ _ _ _ _ · · ·
2 3 5
2 o_ _ _ _ _ _ _ 4 o_ _ _ _ _ _ _ 6 o_ _ _ _ _ _ _ · · ·
1 @ 1 << @ 3 << @ 5 >> >
<< 2 << 2 >> 2 ~~
2 << 2 << 2 >> 2 ~~
<< << >> ~~
> ~~~
1 o_ _ _ _ _ _ _ 3 o_ _ _ _ _ _ _ 5 o_ _ _ _ _ _ _ ···
1 0 2 4
(e): Let
R C C
A = 0 C C ⊂ M2 (C).
0 0 R
Again using the dimension vector notation we get an infinite preprojective compo-
nent:
2
1 1 o_ _ _ _ _ 22 o_ _ _ _ _ 43 o_ _ _ _ _ · · ·
1 B
2
C 77
772 2
C 1 777 2 2
C 3 777 2 2
77 777 777
2 4 6 8
2 1 o_ _ _ _ _ 3 o_ _ _ _ _ 5 o_ _ _ _ _ 7
0 7
2
C 77 C 2 777 2 C 4 777 2 C 6 999 2
772 2
777 2
777 2
99
7 99
1
1 0 o_ _ _ _ _ 11 o_ _ _ _ _ 32 o_ _ _ _ _ 33 o_ _ _ _ _ · · ·
0 0 2 2
236 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
237
238 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
3
a
2
b c
1
and I is the ideal generated by ba. The indecomposable projective A-modules are
3
of the form P1 = 1 , P2 = 1 2 1 , P3 = 2 . Then ∞ ∆ consists of a preprojective
1
component
0
1 o_ _ _ _ _ _ _ 03 o_ _ _ _ _ _ _ 05 o_ _ _ _ _ _ _ · · ·
2 B 4 A
B 88
88
88
88 B 6 :::
2 2 2 2 :::2
88 882
88 88 ::
::
0
0 o_ _ _ _ _ _ _ 02 o_ _ _ _ _ _ _ 04 o_ _ _ _ _ _ _ · · ·
1 3 5
3
a
2 d
b c
1
and I is the ideal generated by ba. The indecomposable projective A-modules are of
3
the form P1 = 1 , P2 = 1 2 1 , P3 = 2 1 . Then ∞ ∆ consists of two points, namely
1
P1 and P2 :
0 4
1 o_ _ _ _ _ _ _ 7
B 2 88 B 12
88
2 2
88 2
88
0 2
0 4
1 88 B 7
88
882 2
88
1
1
2
Note that one of the direct summands of the radical of P3 does not show up in the
course of the knitting algorithm. So we get 2 ∆ \ 1 ∆ = ∅.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 239
3=
==
a ==d
==
2 4
b c
1
and I is the ideal generated by ba. The indecomposable projective A-modules are
3
of the form P1 = 1 , P2 = 1 2 1 , P3 = 2 4 , P4 = 4 . Then ∞ ∆ has two connected
1
components, one is an (infinite) preprojective component, and the other one consists
just of the vertex P4 :
0
1 0 o_ _ _ _ _ _ _ _ 03 0 o_ _ _ _ _ _ _ _ 05 0 o_ _ _ _ _ _ _ · · ·
2 @ 4 @ 6 A
@ ==
== 2
==
== 2
<<
2 == 2 == 2 << 2 2
<<
== == <<
<
0
0 0 o_ _ _ _ _ _ _ _ 02 0 o_ _ _ _ _ _ _ _ 04 0 o_ _ _ _ _ _ _ _ ···
1 3 5
0 1
0 1 o_ _ _ _ _ _ _ _ 1 0
0 1
==
== @
==
==
1
1 1
1
3
b a
2
c
1
and I is the ideal generated by ca and cb. The indecomposable projective A-modules
are of the form P1 = 1 , P2 = 1 2 1 , P3 = 2 3 2 . Then ∞ ∆ consists of an infinite
240 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
0
1
B 1 88
88
88
88
0 0
0 o_ _ _ _ _ _ _ 1 o_ _ _ _ _ _ _ 23 o_ _ _ _ _ _ _ 45 o_ _ _ _ _ _ _ · · ·
1 0 88 B 0 88 B 0 88 A <<<
88 88 88 <<
882 2 882 2 882 2
<<2
88 88 88 <<
<<
1
2 o_ _ _ _ _ _ _ 34 o_ _ _ _ 5
_ _ _ 6 o_ _ _ _ _ _ _ · · ·
0 0 0
(l): Let A = K[T ]/(T 4). There is just one simple A-modules S, and all indecom-
posable A-modules are uniserial. The Auslander-Reiten quiver looks like this:
S
S
S
S
O
S W
S hg
S
O
W
S
S
O hg
W
S hg
The only indecomposable projective A-module has length 4. For the other three
indecomposables we have τA (X) ∼
= X. For example, the obvious sequence of the
form
S S S
0→ S → S ⊕ S → S →0
S
is an Auslander-Reiten sequence.
(m): Let Q be the quiver
4=
==
==a
==
2 == 3
==
==
= b
1
Using the socle series notation the Auslander-Reiten quiver of A looks as follows:
3 o_ _ _ _ _ _ _ _ _ _ _ 2 o_ _ _ _ _ _ _ _ _ _ _ 4
? 1 FF tt: JJ
JJ x< 3 ??
FF t
FF tt
tt JJ
JJ xx ??
F tt JJ xxx ??
F" tt J% xx ??
1 o_ _ _ _ _ _ _ _ 2 3 o_ _ _ _ _ _ _ _ _ _ _ 4 o_ _ _ _ _ _ _ _ 4
== 1 3 2 @
== zz= HH
HH v vv: DD
D
== z HH v D DD
== zz HH vv
= zzz H$ vvv DD
D
z !
2 o_ _ _ _ _ _ _ _ _ 3 4 2 3 o_ _ _ _ _ _ _ _ _ 4
1 B 1 G > 2
BB
BB ww; GG |||
BB ww GG ||
BB ww GG ||
ww G# ||
4 4
2 3 o_ _ _ _ _ _ _ _ _ _ 2 3
>1 GG ; 1 B
||| GG www BB
BB
|| GG w w BB
| | GG w w BB
| G# ww B
| |
4
3 o_ _ _ _ _ _ _ _ _ _ 2 o_ _ _ _ _ _ _ _ _ _ 3
1
and let A = KQ/I where I is generated by ba, cd, ac − db. The (ΓA , dA ) looks as
follows (one has to identify the three modules on the left with the three modules on
the right):
3
2
? 3 ??
??
??
??
1 o_ _ _ _ _ _ _ _ _ 2 o_ _ _ _ _ _ _ _ 3 o_ _ _ _ _ _ _ _ _ 1
DD > 3 ? ?2 BB =
DD || ?? BB zz
DD | ?? BB zzz
DD || ?? BB z
DD || ?? B zz
! || zz
2 2
1 3 / 2 o_ _ _ _ _ _ _ _ 2 o_ _ _ _ _ _ _ _ 1 3 / 1 3
2 = 1 3 BB ? ?? > 2 DD 2
zzz BB ?? || DDD
zz BB ?? || DD
zz BB ?? ||| DD
zz B ? || D!
z
3 o_ _ _ _ _ _ _ _ _ 2 o_ _ _ _ _ _ _ _ 1 o_ _ _ _ _ _ _ _ _ 3
1 ? ? 2
??
??
??
?
1
2
1
and let A = KQ/I where I is generated by cba. Then (ΓA , dA ) looks as follows:
6 o_ _ _ _ _ _ _ 4 o_ _ _ _ _ _ _ 3 Ho_ _ _ _ _ _ _ 1 2
? 5 DD v; FFF xx
; HH z< ??
DD v vv FF xx HH
HH zz ??
D" v FF x z ??
vv # x x $ z
5 o_ _ _ _ _ _ 4 5 6 o_ _ _ _ _ _ _ 3 4 o_ _ _ _ _ _ _ 1 2 3 o_ _ _ _ _ _ 2
== = F < DD ;x BB @
==
| || FFF
zzz DD xx BB
== | F F# D" x B
|| zz xx B!
3 2
4
5
o
_ _ _ _ _ _ 4 6 1 3 o_ _ _ _ _ _ 2 3
@@ < 5 BB 4 G
@@
@@ z zz BB
F 222
BB 22
zz
3
B!
22
4 o
_ _ _ _ _ _ _ 6
22
5
22
22
2
1 o_ _ _ _ _ _ _ 3
4
? 5 DD u uu: MMMM
M qq qq8 IIII z< 1 ??
DD uu MMM qq II zz ??
D" uu MM& qqq I$ zz ??
q
5 o_ _ _ _ _ _ 4 5 6 o_ _ _ _ _ _ _ _ _ 3 4 o_ _ _ _ _ _ _ _ _ 1 2 3 o_ _ _ _ _ _ 2
== 9 K @
== ||= HH
HHH ssss KKK
K vvv
; BB
BB
== | s KKK% v BB
||
| H# sss vv !
3 2
4
5
o
_ _ _ _ _ _ 4 6 o
_ _ _ _ _ _ _ _ _ 1 3 o
_ _ _ _ _ _ 2
>> < 5 HH : 4
DD @ 3
>> zz HH v vv D
>> zz HH v DDD
zz $ vv "
3 2
4 o_ _ _ _ _ _ _ 1 3 4 6 o_ _ _ _ _ _ _ 2 3
5 4
BB
BB ww ; 5
GGG ||=
BB w GGG ||
! www # ||
2 2
1 o_ _ _ _ _ _ _ _ 3 4 6
3
4
5 5
=| GG ; BB
|| GG www BB
| GG# w BB
||| ww BB
| 2 !
1 o_ _ _ _ _ _ _ _ _ 3 4 o_ _ _ _ _ _ _ _ _ 6
5
————————————————————————————-
Note that n
X
[M : Si ] = l(X).
i=1
Furthermore, dim is additive on short exact sequences , i.e. if 0 → X → Y →
Z → 0 is a short exact sequence, then dim(Y ) = dim(X) + dim(Z).
Lemma 32.1. If
f : {A-modules}/ ∼
= −→ H
is a map which is additive on short exact sequences and H is an abelian group, then
there exists a unique group homomorphism f ′ : G(A) → H such that the diagram
dim
{A-modules}/ ∼
= / G(A)
j j j
j
f
j j j′
uj j j f
H
commutes.
Here is an alternative construction of G(A): Let F (A) be the free abelian group
with generators the isomorphism classes of finite-dimensional A-modules. Let U(A)
be the subgroup of F (A) which is generated by the elements of the form
[X] − [Y ] + [Z]
if there is a short exact sequence 0 → X → Y → Z → 0. Define
G(A) := F (A)/U(A).
244 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
For an A-module M set [M] := [M] + U(A). It follows that G(A) is isomorphic to
Zn with generators [Si ], 1 ≤ i ≤ n. By induction on l(M) one shows that
n
X
[M] = [M : Si ] · [Si ].
i=1
Define
d
X
hX, Y iA := (−1)t dim ExttA (X, Y ).
t=0
(If gl. dim(A) = ∞,
P but proj. dim(X) < ∞ or inj. dim(Y ) < ∞, then we can still
define hX, Y iA := t≥0 (−1)t dim ExttA (X, Y ).)
Recall that Ext0A (X, Y ) = HomA (X, Y ). We know that for each short exact sequence
0 → X ′ → X → X ′′ → 0
and an A-module Y we get a long exact sequence
0 / Ext0A (X ′′ , Y ) / Ext0A (X, Y ) g
/ Ext0A (X ′ , Y )
gg
ggggg
ggg gggggg
gg
sggggg
Ext1A (X ′′ , Y ) / Ext1 (X, Y )
A
/ Ext1A (X ′ , Y )
gggggg
ggggg
gg ggggggg
sgggg
Ext2A (X ′′ , Y ) / Ext2 (X, Y )
A
/ Ext2A (X ′ , Y )
gggggg
ggggg
gg ggggggg
sgggg
Ext3A (X ′′ , Y ) / ···
In other words,
hX ′′ , Y iA − hX, Y iA + hX, Y iA = 0.
It follows that
h−, Y iA : {A-modules}/ ∼
=→Z
is a map which is additive (on short exact sequences). Thus hdim(X), Y iA :=
hX, Y iA is well defined.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 245
————————————————————————————-
Examples: (a): Let A = K[T ]/(T m ) for some m ≥ 2. Then none of the indecom-
posable A-modules is directing.
We define classes
∅ = −1 M ⊆ 0 M ⊆ · · · ⊆ n−1 M ⊆ n M ⊆ · · ·
of indecomposable modules as follows: Set −1 M = ∅. Let n ≥ 0 and assume
that n−1 M is already defined. Then let n M be the subcategory of all indecompos-
able modules M in M with the following property: If N is indecomposable with
radA (N, M) 6= 0, then N ∈ n−1 M.
Let [
∞M = nM
n≥0
be the full subcategory of M containing all M ∈ n M, n ≥ 0.
Let
E(A)
be the full subquiver of all vertices [X] of ΓA such that X is a reachable module.
One easily checks that E(A) is again a valued translation quiver.
Summarizing our results and notation, we obtain
E(A) = ∞ (ΓA ) = ∞ ∆, and E(A) = ∞ M.
Furthermore, E(A) is the full subcategory of all A-modules X such that [X] ∈ E(A).
We say that K is a splitting field for A if EndA (S) ∼
= K for all simple A-modules
S.
Examples: If K is algebraically closed, then K is a splitting field for K. Also, if
A = KQ is a finite-dimensional path algebra, then K is a splitting field for A.
Roughly speaking, if K is a splitting field for A, then there are more combinatorial
tools available, which help to understand (parts of) mod(A). The most common
tools are mesh categories and integral quadratic forms.
Theorem 33.4. Let A be a finite-dimensional K-algebra, and assume that K is a
splitting field for A. Then the valuation for E(A) splits, and there is an equivalence
of categories
η : KhE(A)e i → E(A).
aiY Z aiXY = 0.
Y ∈n I i=1
we know that τAn (X) is projective for some n ≥ 1. Thus by induction we get
F (X) ∼
= EndA (X) ∼= K.)
Surjectivity of η: Let h : M → Z be a homomorphism in ∞ I, and let Z ∈ n I. We
use induction on n. If M = Z, then h = c · 1M for some c ∈ K. Thus h = η(c · 1[M ] ).
Assume now that M 6= Z. This implies that h is not an isomorphism. The sink
map ending in Z is M
g = (aiY Z )Y,i : Y dY Z → Z.
Y ∈n−1 I
We get X
h= aiY Z hY,i.
Y,i
By induction the homomorphisms hY,i : M → Y are in the image of η, and by the
construction of η also the homomorphisms aiY Z are contained in the image of η.
Thus h lies in the image of η
Injectivity of η: Let R be the mesh ideal in the path category KΓ. We investigate
the kernel K of
η : KΓ → ∞ I.
Clearly, R ⊆ K. Next, let ω ∈ K. Thus ω ∈ HomKΓ([M], [Z]) for some [M] and
[Z]. We have to show that ω ∈ R. Assume [Z] ∈ n I. We use induction on n.
Additionally, we can assume that ω 6= 0. Thus there exists a path from [M] to [Z].
If [M] = [Z], then ω = c · 1[M ] and η(ω) = c · 1M = 0. This implies c = 0 and
therefore ω = 0.
Thus we assume that [M] 6= [Z]. Now ω is a linear combination of paths from [M]
to [Z], i.e. ω is of the form X
ω= αYi Z ωY,i
Y,i
where the ωY,i are elements in HomKΓ([M], [Y ]). Note that [Y ] ∈ n−1 I. Applying η
we obtain X
0 = η(ω) = aiY Z η(ωY,i).
Y,i
Thus assume Z is not projective. Then we know the kernel of the map
M
g = (aiY Z )Y,i : Y dY Z → Z
Y ∈n−1 I
namely M
f = (aiXY )Y,i : X → Y dY Z .
Y ∈n−1 I
250 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
4o 6
and let A = KQ. Here is the Auslander-Reiten quiver of A, using the dimension
vector notation:
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 251
11 00 10 11 00 00
10 11 10 11 00 00
00 10 11 00 10 01
@ == @ == @ == @ == @ == @
== == == == ==
10 11 10 21 11 0 0
10 21 21 21 11 0 0
00 10 21 11 10 1 1
@ == @ == @ == @ == @ == @
== == == == ==
00 00 10 10 21 11 21 10 21 11 11 0 0
10 / 11 / 21 / 10 / 31 / 21 / 32 / 11 / 21 / 10 / 11 / 0 1
00 00 10 10 21 11 21 10 21 11 11 0 0
== @ == @ == @ == @ == @ ==
== == == == == ==
00 10 21 11 10 1 1
10 21 21 21 11 0 0
10 11 10 21 11 0 0
== @ == @ == @ == @ == @ ==
== == == == == ==
00 10 11 00 10 01
10 11 10 11 00 00
11 00 10 11 00 00
Here we display the locations of the indecomposable projective and the indecom-
posable injective A-modules:
P5 > ◦ ◦ ◦ ◦ I6
> >> A ;;; A ;;; A ;;; @ ???? ?
}} >> ;; ;; ;; ?
}}
2 P A ? ◦ ;; A◦; A◦; A◦= I4
}}> AA ;; ;;; ;;; ==== ?
}} AA ; ; ;
P1 /P /◦ /◦ /◦ /◦ /◦ /◦ /◦ /◦ /I
1 ? I3
/
3
AA }}> >>
>> A ;;
;; A ;;
;; A ;;
;; @ ?
AA ??
}}
} > ; ; ;
P4 A ? ◦ ;; A ◦ ;; A ◦ ;; A ◦ == I2
AA ;; ;; ;; == ? ??
??
A ; ; ; =
P6 ◦ ◦ ◦ ◦ I5
The following pictures show how to compute dim HomA (Pi , −) for all indecom-
posable projective A-modules Pi . Note that the cases P2 and P4 , and also P5
and P6 are dual to each other. We marked the vertices [Z] by a where a =
dim HomA (Pi , Z), provided none of the paths in E(A) starting in [Pi ] and ending
in [Z] contains a subpath of the form [Y ] → [E] → [τA−1 (Y )]. Of course, we can
compute dim HomA (X, −) for any indecomposable A-module.
252 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
1 1 1 1 0 0
@ ;; B 888 B 888 B 888 B 888 B
;;
;;
88
88
88
88
88
88
88
88
1 2 2 2 1 0
@ >> A 888 B 888 B 888 B 888 B
>> 8 8 8 8
>> 88 88 88 88
> 8 8 8 8
1 / 1 /2 /1 /3 /2 /3 /1 /2 /1 /1 /0
>> @ ;;
;; B 88
8 B 88
8 B 88
8 B 88
8
>>
;; 88 88 88 88
>>
; 88 88 88 88
1 2
A 8 2
B 8 2
B 8 1
B 8 B08
>> 888 888 888 888 888
>> 88 88 88 88 88
>>
8 8 8 8 8
1 1 1 1 0 0
1 0 1 1 0 0
@ >> @ ;;; B 888 B 888 B 888 B
>> ;; 88 88 88
>>
> ;; 88 88 88
1 1 A 1 88 B 2 88 B 1 88 0
A >> @ >>
>> 8 8 8 B
>>
>> 88 88 88
>>
>> 88 88 88
0; /0 / 1 / 1 /2 /1 /2 /1 /2 /1 /1 /0
@ ;; B 88 B 88 B 88
;; @ >>
>>
;; 88 88 88
;;
;; >> ;; 88 88 88
; 8 8 8
0 >> 1 2 1 1 1
>> @ >> A 888 B 888 B 888 B 888
>> 88 88 88 88
>> >> 88
88
88
88
>>
0 1 1 0 1 0
@0 >> @
1 B08 B18 B08 B0
>>
;;
;; 888 888 888
>> 88 88 88
>
;; 8
8
8
A0 >> @
1
>> A18 B18 B18 B0
>>
>> >> 888 888 888
88 88 88
>
>> 8
8
8
>
0 / 1 / 1 / 0 /1 /1 /2 /1 /1 /0 /1 /1
;;
;; @ >> @ ;;
;; B 888 B 888 B 888
>> 88 88 88
;;
>> ;; 88
88
88
; ;
0 >> 1 A 1 88 B 1 88 B 1 88 B08
>> @ >>
>> 88 88 88 888
>> >> 88 88 88 88
>>
8 8 8 8
0 1 0 1 0 0
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 253
Proof. Since rad(EndA (X)) = 0, we know that EndA (X) is a skew-field. It is also
clear that Ext1A (X, X) = 0: If 0 → X → M → X → 0 is a short exact sequence
which does not split, then we immediately get a cycle (X, Mi , X) where Mi is an
indecomposable direct summand of M.
Let C be the class of indecomposable A-modules M with M X. We will show by
induction that ExtjA (M, X) = 0 for all M ∈ C and all j ≥ 1:
The statement is clear for j = 1. Namely, if Ext1A (M, X) 6= 0, then any non-split
short exact sequence M
0→X→ Yi → M → 0
i
yields X ≺ M X, a contradiction.
Next, assume j > 1. Without loss of generality assume M is not projective. Let
ε
0 → Ω(M) → P0 − → M → 0 be a short exact sequence where ε : P0 → M is a
projective cover of M. We get
Extj (M, X) ∼ j−1
= Ext (Ω(M), X). A A
(i) M is sincere;
(ii) For each simple A-module S we have [M : S] 6= 0;
(iii) If e is a non-zero idempotent in A, then eM 6= 0;
(iv) For each indecomposable projective A-module P we have HomA (P, M) 6= 0;
(v) For each indecomposable injective A-module I we have HomA (M, I) 6= 0
Proof. Exercise.
Theorem 33.8. Let M be a sincere directing A-module. Then the following hold:
Proof. (i): We can assume that M is not projective. Assume there exists an inde-
composable injective A-module I with HomA (I, τ (M)) 6= 0. Since M is sincere, we
have HomA (M, I) 6= 0. This yields M I ≺ τ (M) ≺ M, a contradiction. Thus
proj. dim(M) ≤ 1.
Proof. (a): Without loss of generality we can assume that X and Y are sincere:
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 255
Assume X is not sincere. Then let R be the two-sided ideal in A which is gen-
erated by all primitive idempotents e ∈ A such that eX = 0. It follows that
R ⊆ AnnA (X) := {a ∈ A | aX = 0} and R ⊆ AnnA (Y ) := {a ∈ A | aY = 0}.
Clearly, eX = 0 if and only eY = 0, since dim(X) = dim(Y ). We also know that
AnnA (X) is a two-sided ideal: If a1 X = 0 and a2 X = 0, then (a1 + a2 )X = 0.
Furthermore, if aX = 0, then a′ aX = 0 and also aa′′ X ⊆ aX = 0 for all a′ , a′′ ∈ A.
It follows that X and Y are indecomposable sincere A/R-modules. Furthermore, X
is also directing as an A/R-module, since a path in mod(A/R) can also be seen as
a path in mod(A). Thus from now on assume that X and Y are sincere.
(b): Since X is directing, we get proj. dim(X) ≤ 1, inj. dim(X) ≤ 1 and gl. dim(A) ≤
2. Furthermore, we know that hdim(X), dim(X)iA = dimK EndA (X) > 0, and
therefore
Summary
Note that we do not specify what “classify” and “describe” should exactly mean.
(a) Let E(A) be the subcategory of ind(A) containing all reachable A-modules.
For all X ∈ E(A) and all Y ∈ ind(A) we have dim(X) = dim(Y ) if and only
if X ∼
=Y.
(b) The knitting algorithm gives ∞ ∆ = ∞ (ΓA ) = E(A), and for each [X] ∈ E(A)
we can compute dim(X).
(c) For X ∈ ind(A) we have [X] ∈ E(A) if and only if X ∈ E(A).
256 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
We cannot hope to solve Problems 33.10 and 33.11 in general, but for the subcat-
egory E(A) ⊆ ind(A) of reachable A-modules, we get a complete classification of
reachable A-modules (the isomorphism classes of reachable modules are in bijection
with the dimension vectors obtained by the knitting algorithm), and we know a lot
of things about the morphism spaces between them.
Keep in mind that there is also a dual theory, using “coreachable modules” etc.
Furthermore, for some classes of algebras we have E(A) = ind(A), for example
if A is a representation-finite path algebra, or more generally if ΓA is a union of
preprojective components.
Note that QA and Qop A can be seen as valued translation quivers, where all vertices
are projective and injective.
Special case: Assume that A is hereditary. Then we have
dPj Pi = dij and d P i = d Si = d i .
Thus, the subquiver PA of preprojective components of (ΓA , dA ) is (as a valued
translation quiver) isomorphic to NQop
A.
The list of Dynkin graphs can be found in Skript 3. Note that non-isomorphic
hereditary algebras can have the same valued graph.
33.5. Exercises. 1: Let A be an algebra with gl. dim(A) ≥ d. Show that there
exist indecomposable A-modules X and Y with ExtdA (X, Y ) 6= 0.
————————————————————————————-
(Of course, the modules Pi , Ii and Si are just sets of representatives of isomorphism
classes of projective, injective and simple A-modules, respectively.)
Let X and Y be A-modules.
is the Ringel form of A. This defines a (not necessarily symmetric) bilinear form
h−, −iA : Zn × Zn → Z.
If proj. dim(X) < ∞ or inj. dim(X) < ∞, then set
X
χA (X) := χA (dim(X)) := hX, XiA = (−1)t dimK ExtiA (X, X).
t≥0
We did all the missing proofs in this section in the lectures. But you
also find them in Ringel’s book.
If dim(P1 ), . . . , dim(Pn ) are linearly independent, then define the Coxeter matrix
ΦA of A by
dim(Pi )ΦA = −dim(Ii )
258 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Proof. We know that gl. dim(A) < ∞ if and only if proj. dim(S) < ∞ for all simple
A-modules S. Furthermore {dim(Si ) | 1 ≤ i ≤ n} are a free generating set of the
Grothendieck group G(A). Let
0 → P (d) → · · · → P (1) → P (0) → S → 0
be a minimal projective resolution of a simple A-module S. This implies
d
X
(−1)i dim(P (i) ) = dim(S).
i=0
Dually, if gl. dim(A) < ∞, then dim(I1 ), . . . , dim(In ) are also linearly independent.
So ΦA is invertible in this case.
By the definition of ΦA , for each P ∈ proj(A) we have
(3) dim(P )ΦA = −dim(ν(P )).
p
Let M be an A-module, and let P (1) −
→ P (0) → M → 0 be a minimal projective
presentation of M. Thus we obtain an exact sequence
(4) 0 → M ′′ → P (1) → P (0) → M → 0
where M ′′ = Ker(p) = Ω2 (M). We also get an exact sequence
νA (p)
(5) 0 → τA (M) → νA (P (1) ) −−−→ νA (P (0) ) → νA (M) → 0
since the Nakajama functor νA is right exact.
There is the dual construction of τA−1 : For an A-module N let
q
(6) 0 → N → I (0) −
→ I (1) → N ′′ → 0
q
be an exact sequence where 0 → N → I (0) −
→ I (1) is a minimal injective presentation
of N.
Applying νA−1 yields an exact sequence
ν −1 (q)
(7) 0 → νA−1 /N) → νA−1 (I (0) ) −−
A
−→ νA−1 (I (1) ) → τA−1 (N) → 0
Lemma 34.2. We have
(8) dim(τA (M)) = dim(M)ΦA − dim(M ′′ )ΦA + dim(νA (M)).
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 259
Note that Equation (11) has many consequences and applications. For example, if
A is a hereditary algebra, then each A-module M satisfies proj. dim(M) ≤ 1, and if
M is non-projective, then HomA (M, A A) = 0.
Lemma 34.5. Assume proj. dim(M) ≤ 2. If dim(τA (M)) = dim(M)ΦA , then
proj. dim(M) ≤ 1 and HomA (M, A A) = 0.
Using the notations from Equation (6) and (7) we obtain the following dual state-
ments:
(i) We have
dim(τA−1 (N)) = dim(N)φ−1 ′′ −1 −1
A − dim(N )ΦA + dim(νA (N)).
Proof. Applying HomA (−, U), HomA (−, X) and HomA (−, V ) we obtain the com-
mutative diagram
0 0 0
δ
0 / HomA (V, U) / HomA (X, U) / HomA (U, U) / Ext1A (V, U)
0 / HomA (V, X) / HomA (X, X) / HomA (U, X)
0 / HomA (V, V ) / HomA (X, V ) / HomA (U, V )
with exact rows and columns. Since η does not split, we know that the connecting
homomorphism δ is non-zero. This implies
dim HomA (X, U) ≤ dim HomA (V, U) + dim HomA (U, U) − 1.
Thus we get
dim HomA (X, X) ≤ dim HomA (X, U) + dim HomA (X, V )
≤ dim HomA (V, U) + dim HomA (U, U) − 1
+ dim HomA (V, V ) + dim HomA (U, V )
= dim EndA (U ⊕ V ) − 1.
This finishes the proof.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 261
Proof. (i): Let x > 0 in G(A) = Zn . Thus x = dim(X) for some non-zero A-module
X. We choose X such that dim EndA (X) is minimal. In other words, if Y is another
module with dim(Y ) = x, then dim EndA (X) ≤ dim EndA (Y ).
and Lemma 34.6 implies that dim EndA (Y ) < dim EndA (X), a contradiction.) Fur-
thermore, since Xi is directing, we have Ext1A (Xi , Xi) = 0 for all i. Thus we get
Ext1A (X, X) = 0. Since gl. dim(A) ≤ 2, we have
χA (x) = χA (dim(X)) = dim EndA (X) + dim Ext2A (X, X) > 0.
Thus χA is weakly positive.
(ii): If Y is an indecomposable A-module, then we know that
χA (Y ) = dim EndA (Y ),
since Y is directing. We also know that EndA (Y ) is a skew field, which implies
F (Y ) ∼
= EndA (Y ). Thus, χA (Y ) = 1 in case F (Y ) ∼
= K.
Furthermore, we know that any two non-isomorphic indecomposable A-modules Y
and Z satisfy dim(Y ) 6= dim(Z). So the map dim is injective.
Assume additionally that x is a root of χA . Now
1 = χA (x) = dim EndA (X) + dim Ext2A (X, X)
262 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Proof. Use partial derivations of χ and some standard results from Analysis. For
details we refer to [Ri1].
The Cartan matrix CA = (cij )ij of A is the n × n-matrix with ijth entry equal to
cij := [Pj : Si ] = dim(Pj )i .
Thus the jth column of CA is given by dim(Pj )T .
Recall that the Nakayama functor ν = νA = D HomA (−, A A) induces an equivalence
ν : proj(A) → inj(A)
where ν(Pi ) = Ii . It follows that
dim(Ii )j = dim HomA (Ii , Ij ) = dim HomA (Pi , Pj ) = cij .
(Here we used our assumption that K is a splitting field for A.)
Thus the ith row of CA is equal to dim(Ii ). So we get
(12) dim(Pi ) = ei CAT and dim(Ii ) = ei CA .
Lemma 34.10. If gl. dim(A) < ∞, then CA is invertible over Z.
But note that there are algebras A where CA is invertible over Q, but not over Z,
for example if A is a local algebra with non-zero radical.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 263
Assume now that the Cartan matrix CA of A is invertible. We get a (not necesssarily
symmetric) bilinear form
h−, −i′A : Qn × Qn → Q
defined by
hx, yi′A := xCA−T y T .
Here CA−T denote the inverse of the transpose CAT of C. Furthermore, we define a
symmetric bilinear form
(−, −)′A : Qn × Qn → Q
by
(x, y)′A := hx, yi′A + hy, xi′A = x(CA−1 + CA−T )y T .
Set χ′A (x) := hx, xi′A . This defines a quadratic form
χ′A : Qn → Q.
It follows that
(x, y)′A = χ′A (x + y) − χ′A (x) − χ′A (y).
The following lemma shows that the form h−, −i′A we just defined using the Cartan
matrix, coincides with the Ringel form we defined earlier:
Lemma 34.11. Assume that CA is invertible. If X and Y are A-modules with
proj. dim(X) < ∞ or inj. dim(Y ) < ∞, then
X
hdim(X), dim(Y )i′A = hX, Y iA = (−1)t dim ExttA (X, Y ).
t≥0
Proof. Assume proj. dim(X) = d < ∞. (The case inj. dim(Y ) < ∞ is done dually.)
We use induction on d.
If d = 0, then X is projective. Without loss of generality we assume that X is
indecomposable. Thus X = Pi for some i. Let y = dim(Y ). We get
hdim(X), dim(Y )i′A = hdim(Pi ), yi′A = dim(Pi )CA−T y T = ei y T = dim HomA (Pi , Y ).
Furthermore, we have ExttA (Pi , Y ) = 0 for all t > 0.
Next, let d > 0. Let P → X be a projective cover of X and let X ′ be its kernel. It
follows that proj. dim(X ′ ) = d − 1. We apply HomA (−, Y ) to the exact sequence
0 → X ′ → P → X → 0.
264 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Proof. This holds since gl. dim(A) ≤ 1 and since K is a splitting field for A.
Lemma 34.13. Let A = KQ be a finite-dimensional path algebra. Then for any
simple A-module Si and Sj we have dim Ext1A (Si , Sj ) is equal to the number of
arrows i → j in Q.
and
n
X X
χKQ (X) = hα, αiKQ = αi2 − qij αi αj
i=1 i<j
where qij is the number of arrows a ∈ Q1 with {s(a), t(a)} = {i, j}.
Lemma 34.15. Assume that CA is invertible. Then
ΦA = −CA−T CA .
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 265
Proof. We have
hx, yi′A = xCA−T y T = (xCA−T y T )T = yCA−1 xT
= yCA−T CAT CA−1 xT = −yCA−T ΦTA xT = −hy, xΦA i′A .
This proves the first equality. Repeating this calculation we obtain the second
equality.
Lemma 34.17. If there exists some x > 0 such that xΦA = x, then χA is not weakly
positive.
Proof. We have (x, y)′A = 0 for all y if and only if x(CA−1 + CA−T ) = 0 if and only if
xCA−1 = −xCA−T if and only if xΦA = x.
Corollary 34.18. If there exists some x > 0 such that xΦA = x, then χ′A is not
weakly positive.
m
Assume there exists an indecomposable KQ-module X with τKQ (X) ∼
= X and
assume m ≥ 1 is minimal with this property. Set
m
M
i
Y = τKQ (X).
i=1
Then τKQ (Y ) ∼
= Y which implies
dim(Y ) = dim(Y )ΦKQ .
We get
(Y, Z)KQ = hY, ZiKQ + hZ, Y iKQ
= −hdim(Z), dim(Y )ΦKQ i − hdim(Y )Φ−1
KQ , dim(Z)i
= −(hY, ZiKQ + hZ, Y iKQ ).
This implies dim(Y ) ∈ rad(χKQ ).
Lemma 34.19. For an A-module M the following hold:
————————————————————————————-
Parts of this section are copied from Crawley-Boevey’s lecture notes “Lectures on
representations of quivers”, which you can find on his homepage.
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 267
For vertices i, j ∈ Q0 let qij = qji be the number of arrows a ∈ Q1 with {s(a), t(a)} =
{i, j}. Note that the numbers qij do not depend on the orientation of Q.
For α = (α1 , . . . , αn ) ∈ Zn define
n
X X
qQ (α) := αi2 − qij αi αj .
i=1 i≤j
n
We call the quadratic form qQ : Z → Z the Tits form of Q.
The symmetric bilinear form (−, −)Q : Zn × Zn → Z of Q is defined by
(
−qij if i 6= j,
(ei , ej )Q :=
2 − 2qii otherwise.
As before, ei denotes the canonical basis vector of Zn with ith entry 1 and all other
entries 0.
We have
(α, α)Q = 2qQ (α),
(α, β)Q = qQ (α + β) − qQ (α) − qQ (β).
Note that qQ and (−, −)Q do not depend on the orientation of the quiver Q.
For α, β ∈ Zn define
X X
hα, β, iQ := αi βi − αs(a) βt(a) .
i∈Q0 a∈Q1
(i) β is sincere;
(ii) qQ is positive semi-definite;
(iii) For α ∈ Zn the following are equivalent:
(a) qQ (α) = 0;
(b) α ∈ Qβ;
(c) α ∈ rad(qQ ).
If βi = 0, then X
qij βj = 0,
j6=i
and since qij ≥ 0 for all i, j and β ≥ 0, we get βj = 0 whenever qij > 0. Since Q is
connected, we get β = 0, a contradiction. Thus we proved that β is sincere.
X βj 2 X
= qij α − qij αi αj
i6=j
2βi i i<j
X 1 2 X
= (2 − 2qii )βi αi − qij αi αj = qQ (α).
i
2β i i<j
(c): If qQ (α) = 0, then the calculation above shows that αi /βi = αj /βj whenever
qij > 0. Since Q is connected it follows that α ∈ Qβ.
(d): If α ∈ Qβ, then α ∈ rad(qQ ), since β ∈ rad(qQ ).
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 269
Proof. (ii): It is easy to check that δ ∈ rad(qQ ): If there are no loops or multiple
edges we have to check that for all vertices i we have
X
2δi = δj
j
where j runs over the set of neighbours of i in Q. By Lemma 35.1 this implies that
qQ is positive semi-definite.
In each case there exists some vertex i such that δi = 1. Thus rad(qQ ) = Qδ ∩ Zn =
Zδ.
(i): Any Dynkin quiver Q with n vertices can be seen as a full subquiver of some
Euclidean quiver Q e with n + 1 vertices. We have q e (x) > 0 for all non-sincere
Q
elements in Zn+1 , since the x with qQe (x) = 0 are all multiples of the sincere element
δ. So qQ is positive definite. (The form qQ is obtained from qQe via restriction to the
e
subquiver Q of Q.)
(iii): Let Q be a quiver which is not Dynkin and not Euclidean. Then Q contains
a (not necessarily full) subquiver Q′ such that Q′ is a Euclidean quiver. Note that
any dimension vector of Q′ can be seen as a dimension vector of Q by just adding
some zeros in case Q has more vertices than Q′ .
Let δ be the radical vector associated to Q′ . If the vertex sets of Q′ and Q coincide,
then α := δ satisfies qQ (α) < 0.
Otherwise, if i is a vertex of Q which is not a vertex of Q′ but which is connected
to a vertex in Q′ by an edge, then α := 2δ + ei satisfies qQ (α) < 0.
35.2. Gabriel’s Theorem. Combining our results obtained so far, we obtain the
following famous theorem:
Theorem 35.4 (Gabriel). Let Q be a connected quiver. Then KQ is representation-
finite if and only if Q is a Dynkin quiver. In this case dim yields a bijection between
the set of isomorphism classes of indecomposable KQ-modules and the set of positive
roots of qQ .
Proof. (a): We know that there is a unique preprojective component PKQ of the
Auslander-Reiten quiver ΓKQ .
(b): We have χKQ (X) = qQ (dim(X)) for all KQ-modules X.
(c): Assume KQ is representation-finite. This is the case if and only if PKQ =
ΓKQ. Since all indecomposable preprojective modules are directed, we know that
KQ is a directed algebra. Furthermore, we have gl. dim(KQ) ≤ 1 ≤ 2. So we
can apply Theorem xx and obtain a bijection between the isomorphism classes of
indecomposable KQ-modules and the set of positive roots of χKQ . Furthermore, an
element α ∈ Nn is a positive root of χKQ if and only if α ∈ ∆Q . We also know that
χKQ = qQ is weakly positive. But this implies that Q has to be a Dynkin quiver.
(For all quivers Q which are not Dynkin we found some α > 0 with qQ (α) ≤ 0.)
(d): If KQ is representation-infinite, the component PKQ is infinite. Each indecom-
posable module X in PKQ is directed, and K is a splitting field for KQ. Thus
χKQ (X) = qQ (dim(X)) = 1.
Furthermore, we know that there is no other indecomposable KQ-module Y with
dim(X) = dim(Y ). So we found infinitely many α ∈ Zn with qQ (α) = 1.
Suppose that Q is a Dynkin quiver. Then
∆Q = {α ∈ Zn | qQ (α) = 1}
is a finite set, a contradiction.
————————————————————————————-
An · · ··· · · n≥1
(1,2)
Bn · · ··· · · n≥2
(2,1)
Cn · · ··· · · n≥3
Dn · · · ··· · n≥4
E6 · · · · ·
E7 · · · · · ·
E8 · · · · · · ·
(1,2)
F4 · · · ·
(1,,3)
G2 · ·
***********************************************************************
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 273
en
A 1 1 1 ··· 1 1 1
(1,2) (2,1)
en
B 1 1 1 ··· 1 1 1
(2,1) (1,2)
en
C 1 2 2 ··· 2 2 1
1 1
en
D 1 2 2 ··· 2 2 1
e6
E 1 2 3 2 1
e7
E 1 2 3 4 3 2 1
e8
E 2 4 6 5 4 3 2 1
(1,4) (2,2)
e11
A 2 1 e12
A 1 1
(1,2) (1,2)
gn
BC 2 2 2 ··· 2 2 1
(2,1)
gn
BD 1 2 2 ··· 2 2 2
(1,2)
gn
CD 1 2 2 ··· 2 2 1
(1,2) (2,1)
Fe41 1 2 3 2 1 Fe42 1 2 3 4 2
(1,3) (3,1)
e21
G 1 2 1 e22
G 1 2 3
Part 7. Extras
simple modules
serial modules
uniserial modules
cyclic modules
cocyclic modules
indecomposable modules
projective modules
injective modules
preprojective modules (which should really be called postprojective modules)
preinjective modules
regular modules
bricks
stones
exceptional modules
Schur modules
Prüfer modules
p-adic modules
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 275
generic modules
pure-injective modules
Classifications of modules
Solved:
tame hereditary algebras
tubular algebras
Gelfand-Ponomarev algebras
multicoil algebras
Open:
biserial algebras
However, one still has to be careful what it means to have a classification of all
indecomposable modules over an algebra. For example for tubular algebras, one can
parametrize all indecomposable modules by roots of a quadratic form. But given a
root, it is still very difficult to write down explicitely the corresponding indecom-
posable module(s). In fact, for tubular algebras this remains an open problem.
We list some names of classes of mostly finite-dimensional algebras which were stud-
ied in the literature:
Basic algebras
Biserial algebras
Canonical algebras
276 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Clannish algebras
Cluster-tilted algebra
Directed algebras
Dynkin algebras
Euclidean algebras
Gentle algebras
Group algebras
Hereditary algebras
Multicoil algebras
Nakayama algebras
Path algebras
Poset algebras
Preprojective algebras
Quasi-hereditary algebras
Quasi-tilted algebras
Representation-finite algebras
Selfinjective algebras
Semisimple algebras
Symmetric algebras
Tame algebras
Tilted algebras
Tree algebras
Triangular algebras
REPRESENTATION THEORY OF ALGEBRAS I: MODULES 277
Wild algebras
Here are some classes of algebras, which are not finite-dimensional, but linked to
the finite-dimensional world:
Repetitive algebras
Hecke algebras
39. Dimensions
The concept of “dimension” occurs frequently and with different meanings in the
representation theory of algebras. Here just some of the most common dimensions:
dimension of a module as a vector space
projective dimension of a module
***********************************************************************
278 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
References
[ASS] I. Assem, D. Simson, A. Skowroński, Elements of the Representation Theory of Associative
Algebras. LMS Student Texts 65.
[ARS] M. Auslander, I. Reiten, S.Smalø, Representation theory of Artin algebras. Corrected
reprint of the 1995 original. Cambridge Studies in Advanced Mathematics, 36. Cambridge
University Press, Cambridge, 1997. xiv+425 pp.
[B1] D.J. Benson, Representations and cohomology. I. Basic representation theory of finite groups
and associative algebras. Second edition. Cambridge Studies in Advanced Mathematics, 30.
Cambridge University Press, Cambridge, 1998. xii+246 pp.
[B2] D.J. Benson, Representations and cohomology. II. Cohomology of groups and modules. Sec-
ond edition. Cambridge Studies in Advanced Mathematics, 31. Cambridge University Press,
Cambridge, 1998. xii+279 pp.
[CE] H. Cartan, S. Eilenberg, Homological algebra. With an appendix by David A. Buchsbaum.
Reprint of the 1956 original. Princeton Landmarks in Mathematics. Princeton University
Press, Princeton, NJ, 1999. xvi+390 pp.
[DK] Y.A. Drozd, V. Kirichenko, Finite-dimensional algebras. Translated from the 1980 Russian
original and with an appendix by Vlastimil Dlab. Springer-Verlag, Berlin, 1994. xiv+249 pp.
[G] P. Gabriel, Auslander-Reiten sequences and representation-finite algebras. Representation
theory, I (Proc. Workshop, Carleton Univ., Ottawa, Ont., 1979), pp. 1–71, Lecture Notes in
Math., 831, Springer, Berlin, 1980.
[GR] P. Gabriel, A.V. Roiter, Representations of finite-dimensional algebras. Translated from the
Russian. With a chapter by B. Keller. Reprint of the 1992 English translation. Springer-
Verlag, Berlin, 1997. iv+177 pp.
[HSt] Hilton, Stammbach
[P] R.S. Pierce, Associative algebras. Graduate Texts in Mathematics, 88. Studies in the History
of Modern Science, 9. Springer-Verlag, New York-Berlin, 1982. xii+436 pp.
[ML] S. Mac Lane, Homology. Reprint of the 1975 edition. Classics in Mathematics. Springer-
Verlag, Berlin, 1995. x+422 pp.
[Ri1] C.M. Ringel, Tame algebras and integral quadratic forms. Lecture Notes in Mathematics,
1099. Springer-Verlag, Berlin, 1984. xiii+376 pp.
[Ri2] C.M. Ringel, Representation theory of finite-dimensional algebras. Representations of alge-
bras (Durham, 1985), 7–79, London Math. Soc. Lecture Note Ser., 116, Cambridge Univ.
Press, Cambridge, 1986.
Index
A-module, 86 Category, 37
A-module homomorphism, 87 Centre of a ring, 101
J-module, 12 Chinese Reminder Theorem, 44
J-module homomorphism, 23 Cocomplex of A-modules, 164
J(A), 121 Cohomology group, 166
K-algebra, 25 Cokernel of a homomorphism, 24
K-algebra homomorphism, 85 Column finite matrix, 53
K-linear category, 38 Complete lattice, 15
K-linear functor, 39 Complete set of pairwise orthogonal idempo-
N (∞), 40 tents, 92
N (λ), 20 Complex of A-modules, 164
U (X), 13 Component of an Auslander-Reiten quiver,
IrrA (X, Y ), 218 213
mod(A), 87 Composition factors, 66
radA (X, Y ), 217 Composition of relations, 161
rep(Q) = repK (Q), 111 Composition series, 66
Hom(M, N ), 135 Connected algebra, 103, 214
Mod(A), 135 Connecting homomorphism, 160, 172
mod(A), 135 Contravariant functor, 38
a∗ (f, g), 155 Converse Bottleneck Lemma, 195
b∗ (f, g), 156 Countably generated, 146
c-generated module, 13 Countably generated module, 14
f ⊗ g, 187 Covariant functor, 38
n-subspace problem, 116 Coxeter matrix, 257
Mod(A), 87 Crawley-Jønsson-Warfield Theorem, 145
Rep(Q) = RepK (Q), 111 Cycle (in a module category), 245
Cycle in a quiver, 221
Additive function on a translation quiver, 223 Cyclic module, 14
Additive on short exact sequences, 243
Adjoint functors, 188 Decomposable module, 18
Adjunction formula, 188 Dedekind Lemma, 16
Admissible ideal, 139 Degree of a polynomial, 41
Algebra, 25 Dense functor, 39
Algebra homomorphism, 85 Determined by composition factors, 255
Annihilator of a module, 106 Dimension of a J-module, 12
Auslander algebra, 184 Dimension of a representation, 110
Auslander-Reiten formula, 200 Dimension vector of a finite length module,
Auslander-Reiten quiver, 213, 224 67
Auslander-Reiten quiver (of M), 192 Dimension vector of a quiver representation,
Auslander-Reiten sequence (in M), 192 113
Direct complement of a submodule, 18
Balanced map, 185 Direct decomposition of a module, 18
Basic algebra, 139 Direct sum of modules, 18
Bimodule, 97 Direct sum of quiver representations, 111
Bimodule of irreducible maps, 218 Direct summand of a module, 18
Biquiver, 192 directed algebra, 261
Block of a group algebra, 103 Directing module, 245
Bottleneck Lemma, 192 Dominant dimension of an algebra, 184
Butterfly Lemma, 63 Dynkin quiver, 272
Bézout’s Theorem, 44
Endomorphism, 24
Cancellation Theorem, 81 Endomorphism algebra, 24
Cartan matrix, 262 Epimorphism, 23
279
280 CLAUS MICHAEL RINGEL AND JAN SCHRÖER
Young diagram, 20
Jan Schröer
Mathematisches Institut
Universität Bonn
Beringstr. 1
D-53115 Bonn
Germany