Вы находитесь на странице: 1из 10

CHAPTER 3

Modules

3.1. Some generalities about modules


The concept of module generalizes the idea of both vector spaces and ideals. The
essential idea is simple: a module is like a vector space where we replace the base field by
an arbitrary ring.

3.1.1. Basic definitions.

D EFINITION 3.1.1. Let R be a ring (commutative, with unity). An R–module is an


abelian group (M, +) together with an action of R, i.e a map

R×M −→ M
(r, m) �−→ r ∗ m =: rm ∈ M

satisfying the following conditions:


M1 Distributivity: r(m + n) = rm + rn for all r ∈ R, m, n ∈ M .
M2 Distributivity: (r + s)m = rm + sm for all r, s ∈ R, m ∈ M .
M3 Pseudoassociativity: (rs)m = r(sm) for all r, s ∈ R, m ∈ M .
M4 Modularity: For all m ∈ M one has 1m = m.
If this is the case, we will also say that M is a module over R. When we want to emphasize
the R–module structure of M , we will write it as R M .

E XAMPLES 3.1.2.
(1) Let F be a field, then an F–module is precisely a vector space over F.
(2) Let (M, +) be an abelian group. Define the map

Z×M −→ M

(r)

 if r > 0,
m + m + · · · + m
(r, m) �−→ rm := (−r)
 (−m) + (−m) + · · · + (−m) if r < 0,


0 if r = 0.

From the group axioms it can be asily deduced that the previous map defines
a Z–action on M . In other words, every abelian group is a Z–module. The
converse is also true, if (M, +) is a Z–module the for any r > 0 in Z one has
(r)
r = 1 + 1 + · · · + 1, and thus
(r) (r)
rm = (1 + · · · + 1)m = 1m + · · · + 1m = m + · · · + m,

and for r < 0 one has r = (−1)+ thus Z–modules are precisely abelian groups.
(3) Let R be a ring. Define an action of R on (R, +) by r ∗ s := rs. By the ring laws
this action makes R into an R–module. This action is called the (left) regular
action of R on itself. When we want to refer to R as a module over itself rather
than as a ring we will write it as R R.
21
22 3. MODULES

(4) Let R, S be rings, ϕ : R → S a ring homomorphism, S M an S–module. Con-


sider the map
R×M −→ M
(r, m) �−→ r ∗ m := ϕ(r)m.
The ring morphism axioms, together with the fact that M is an S–module, ensure
that this action provides an R–module structure on M . This is called the R–
module structure induced by ϕ.
In particular, if R ≤ S subring, then every S–module is automatically an
R–module with the action given by restriction of scalars.
(5) Let V be a vector space over a field F, and α : V → V a linear map. Define the
map
F[x] × V −→ V
(x, v) �−→ x ∗ v := α(v)
(n) � �
extending to xn ∗ v := αn (v) = α(α(· · ·α(v))) and ( ai xi ) ∗ v := ai αi (v).
It is an easy exercise to check that under this action V becomes an F[x]–module.
Conversely, if V is a module over the polynomial ring F[x], since F ≤ F[x],
by the previous example V is automatically an F–module, thus a vector space
over F. Now, define α : V → V by α(v) := x ∗ v. By the module axioms, α
is an F–linear map, and thus there is a one to one correspondence between F[x]–
modules and vector spaces V over F endowed with a linear map α : V → V .
(6) Let R be any commutative ring. The set Mn (R) of all n × n matrices with
coefficients in R is an R–module under the action r ∗ (aij ) := (raij ).
R EMARK . This is a particular case of example 4, with M = S = Mn (R)
and ϕ : R → S given by α(r) := r · Id.
D EFINITION 3.1.3. Let RM be an R–module. A subset p ⊆ M is said to be a
submodule of M if:
(1) P is a subgroup of (M, +), i.e. P �= ∅, for all a, b ∈ P one has a + b in P and
−a ∈ P .
(2) For all r ∈ R, and for all m ∈ P one has rm ∈ P .
If P is a submodule of M we will denote that by P ≤ M . In this case, P is also an
R–module in its own right.
R EMARK 3.1.4. P is a submodule of M if and only if for all m, n ∈ P and for all
r, s ∈ R one has rm + sn ∈ P . The proof is left as an exercise.
E XAMPLES 3.1.5.
(1) If V is a vector space over a field F, then submodules are precisely vector sub-
spaces.
(2) If Z M is an abelian group (i.e. a Z–module), then the submodules of M are its
subgroups.
(3) If R R is a ring considered as a module over itself, then the submodules of R R
are the ideals of R.
(4) Let V be a module over the polynomial ring F[x], i.e. V a vector space endowed
with a linear map α : V → V . Assume that P ≤ V is a sumodule of V . Then,
as F ⊆ F[x], for all λ ∈ F and for all v ∈ P one has λv ∈ P , thus P is a
subspace of V . Also, for all v ∈ P one must have xp ∈ P , but xp = α(p), thus
P must satisfy α(P ) ⊆ P . In other words, P must be an α–invariant subspace
of V . Conversely, if P is an α–invariant subspace of V is is easy to see that P is
a submodule of V .
3.1. SOME GENERALITIES ABOUT MODULES 23

(5) For any R–module R M there are two distinguished submodules: the trivial sub-
module 0 := {0} ≤ M and the total submodule M ≤ M .
P ROPOSITION 3.1.6. If R M is an R–module, and A, B ≤ M are submodules of M ,
then the intersection A ∩ B is also a submodule of M . Moreover, A ∩ B is the largest (with
respect to inclusion) sumbodule of M contained in�both A and B. More generally, for any
family {Pα } of sumbodules of M the intersection Pα is also a submodule of M .
P ROOF. The proof is left as an exercise. �

P ROPOSITION 3.1.7. Let R M be an R–module, A, B ≤ M submodules. Define the


sum of A and B as
A + B := {a + b | a ∈ A, b ∈ B} .
Then, A + B is a submodule of M . Moreover, it is the smallest submodule of M containing
both A and B. More generally, for any finite collection A1 , . . . , An ≤ M of submodules
of M their sum
n

Ai = A1 + · · · + An = {a1 + · · · + an | ai ∈ Ai } ≤ M
i=1

is a submodule of M .
P ROOF. The proof is left as an exercise. �

The inclusions of submodules given by the last two propositions can be encoded in the
following diagram:
A �+ � B

A� B

A∩B
3.1.2. Cyclic modules and finitely generated modules.
D EFINITION 3.1.8. Let R M be an R–module, x ∈ M . Define Rx := {rx | r ∈ R}.
It is easy to check that Rx ≤ M is a submodule. The submodule Rx is called the cyclic
submodule of M generated by x.
R EMARK 3.1.9. Rx is the smallest submodule of M containing x.
D EFINITION 3.1.10. If R M is an R–module, and there exist some x ∈ M such that
M = Rx, we say that M is a cyclic module.
D EFINITION 3.1.11. Let R M be an R–module, and let x1 , . . . , xn ∈ M ; the set
Rx1 + · · · + Rxn ≤ M is the smallest sumbodules of M containing all the xi . This
submodule is called the submodule generated by x1 , . . . , xn . If M = Rx1 + · · · + Rxn
we say that M is finitely generated.
D EFINITION 3.1.12. Let P ≤ M be a submodule of the R–module R M . Consider
the abelian group
M
= {m = m + P | m ∈ M } ,
P
where addition is defined in the usual way as m + n := m + n. For each r ∈ R and each
m ∈ M/P , define rm := rm. This operation yields a well defined action of R on M/P ,
thus making M/P into an R–module. This module is called the quotient of M b yP .
24 3. MODULES

R EMARK 3.1.13. For the previous definition to make sense, one needs to check that
the action of R on M/P is well defined. This is the case as m = n if and only if m−n ∈ P ,
and thus for all r ∈ R one has r(m − n) ∈ P , yielding rm = rn.
E XAMPLES 3.1.14.
(1) For each R–module R M , the submodule 0 = R0 is a cyclic submodule.
(2) If V is a vector space over F, 0 �= x ∈ V , then the cyclic submodule Fx is the
1–dimensional subspace of V generated by x.
(3) Let M be a Z–module (i.e. an abelian group), x ∈ M , then Zx is the cyclic
subgroup of M generated by x.
(4) Let x ∈ R R, then the cyclic submodule Rx is precisely the principal ideal (x)
generated by x. In particular R R = R1 is a cyclic R–module.
(5) The quotient of any R–module R M by the trivial submodule 0 is M/0 = M .
(6) The quotient of any R–module R M by the total submodule M is M/M = 0.
(7) The quotient of the Z–module Z Z by the cyclic submodule Zn = (n) is the
cyclic group of order n, Zn = Z/(n).
P ROPOSITION 3.1.15. Let R M be an R–module. If P ≤ M is a submodule and
{x1 , . . . , xn } ⊆ M generate M , then {x1 , . . . , xn } ⊆ M/P generate the quotient module
M/P .
P ROOF. Since
� x1 , . . . , xn generate M , for each m ∈ M there exist r1 , . . . , rn ∈ R
such that m = ri xi , and hence
� �
m= ri x i = ri x i ,
therefore {x1 , . . . , xn } generate M/P . �
C OROLLARY 3.1.16. If M is a cyclic R–module, then for any submodule P ≤ M the
quotient M/P is also cyclic.
P ROOF. By the previous proposition, if M = Rx then M/P = Rx. �
R EMARK 3.1.17. Unlike it happens for groups, a submodule of a cyclic module
doesn’t need to be cyclic. For instance, one can take the R–module R R = R1. Its sub-
modules are the ideals of R, and in general ideals do not have to be principal, as long as
we take R to be a ring which is not a principal ideal domain.
E XAMPLE 3.1.18. Let us consider the R–module R R = R1, which is a cyclic module,
and let I � R be an ideal of R, i.e. a submodule of R R, then the quotient R/I is also a
cyclic R–module. Indeed, it is generated by 1 = 1 + I.
3.1.3. Module homomorphisms.
D EFINITION 3.1.19. Let R be a ring, R M and R N two R–modules; a map α : M →
N is said to be an R–module homomorphism (also called R–homomorphism or R–linear
map) if it satisfies the following conditions:
(1) α is an additive group homomorphism, i.e. α(0) = 0 and α(m1 + m2 ) =
α(m1 ) + α(m2 ) for all m1 , m2 ∈ M .
(2) For all r ∈ R and all m ∈ M one has α(rm) = rα(m)
E XAMPLES 3.1.20.
(1) For any two R–modules M and N , there is always a trivial module homomor-
phism from M to N , the zero homomorphism, defined by 0(m) = 0N for all
m ∈ M . This homomorphism is simply denoted by 0.
(2) For any module M , the identity map Id : M → M is a module homomorphism.
Similarly, if P ≤ M is a submodule, the inclusion map ı : P → M is also a
module homomorphism.
3.1. SOME GENERALITIES ABOUT MODULES 25

(3) If V, W are vector spaces over a field F, a map α : V → W is an F–module


homomorphism if and only if it is a linear map.
(4) If M , N are Z–modules, i.e. abelian groups, a map α : M → N is a module
homomorphism if and only if it is a group homomorphism.
D EFINITION 3.1.21. Let R be a ring, and let R M and R N be R–modules. We will
denote by HomR (M, N ) the set of all R–module homomorphisms from M to N , i.e.
HomR (M, N ) = Hom(M, N ) = {α : M → N | α is a module homomorphism} .
If α ∈ HomR (M, N ), it is easy to see that −α ∈ HomR (M, N ), and that for any r ∈ R
the map rα defined by (rα)(m) = rα(m) is also a module homomorphism. Similarly it
is immediate to prove that for any two module homomorphisms α, β ∈ HomR (M, N ) the
sum map α + β is also a module homomorphism. As a consequence, we obtain that the set
HomR (M, N ) is also an R–module.
P ROPOSITION 3.1.22. Let M ,N and P be R–modules, α ∈ HomR (M, N ) and
β ∈ HomR (N, P ), then βα := β ◦ α ∈ HomR (M, P ).
P ROOF. The proof is completely straightforward, and thus left as an exercise. �

D EFINITION 3.1.23. Let α ∈ HomR (M, N ). If α is injective, we say that it is a


monomorphism, if α is surjective it is called an epimorphism, and if it is bijective it is
called an isomorphism. If there exists an isomorphism α : M → N , we will say that M
and N are isomorphic, and write M ∼ = N.
D EFINITION 3.1.24. Let R be a ring, M and N two R–modules, α ∈ HomR (M, N )
a module homomorphism. The image of α is defined as
Im(α) := {α(m) | m ∈ M } = α(M ),
the kernel of α is defined as
Ker(α) := {m ∈ M | α(m) = 0} = α−1 ({0}).
E XAMPLES 3.1.25.
(1) Let R M be an R–module, P ≤ M a submodule, and consider the canonical
projection ΠP : M → M/P defined by πP (m) = m = m + P . The map πP
is a module homomorphism and we have Ker πP = P and Im πP = M/P , i.e.
πP is an epimorphism.
(2) The zero map 0 : M → N , defined as 0(m) = 0N for all m ∈ M , is a module
homomorphism. One has Ker 0 = M , Im 0 = 0.
(3) α : M → N is a monomorphism if and only if Ker α = 0

P ROOF. One has α(m) = α(n) if and only if 0 = α(m) − α(n) = α(m −
n), i.e. if and only if m − n ∈ Ker α. Thus if Ker α = 0 we have α(m) = α(n)
if and only if m = n, i.e. α is injective. Conversely, if α is injective, then if
0 = α(m) we have α(m) = α(0) and by injectivity we get m = 0, and hence
Ker α = 0. �

(4) A module homomorphism α : M → N is an isomorphism if and only if Ker α =


0 and Im α = N .
L EMMA 3.1.26. Let M, N be R–modules, α ∈ HomR (M, N ), then we have:
(1) Ker(α) ≤ M is a submodule of M .
(2) Im α ≤ N is a submodule of N .
P ROOF.
26 3. MODULES

(1) Let m1 , m2 ∈ Ker α, r1 , r2 ∈ R, then, using the fact that α is a homomorphism,


we have
α(r1 m1 + r2 m2 ) = r1 α(m1 ) + r2 α(m2 ) = r1 0 + r2 0 = 0,
thus r1 m1 + r2 m2 ∈ Ker α, and hence Ker α is a submodule of M .
(2) Let r1 , r2 ∈ R, and let n1 , n2 ∈ Im α,.then n1 = α(m1 ) and n2 = α(m2 ) for
some m1 , m2 ∈ M . Thus, we have
r1 n1 + r2 n2 = r1 α(m1 ) + r2 α(m2 ) = α(r1 m1 + r2 m2 ),
hence r1 n1 + r2 n2 ∈ Im α and thus Im α is a submodule of N .

T HEOREM 3.1.27 (First Isomorphism Theorem for Modules). Let M, N be R–modules,


and α ∈ HomR (M, N ) a module homomorphism, then
M ∼
= Im α.
Ker α
P ROOF. The isomorphism is given by m + Ker α ↔ α(m). The proof is almost
identical, mutatis mutandi, to the one made for ring homomorphisms or for group homo-
morphisms. �

E XAMPLES 3.1.28.
(1) 0 : M → M the zero homomorphism, Ker 0 = M , Im 0 = 0, so the first
isomorphism theorem tells us that M/M ∼= 0.
(2) Id : M → M , the identity map, Ker Id = 0, Im Id = M , thus M/0 ∼= M.
(3) Let R M be an R–module, P ≤ M a submodule, πP : M → M/P the canonical
projection, so Ker πP = P , Im πP = M/P , and the first isomorphism theorem
tells us that M/P ∼
= M/P .
T HEOREM 3.1.29 (Classification of cyclic modules). Let R M be an R–module, then
M is cyclic if and only if M ∼
= R/I for some ideal I � R. Moreover, the ideal I is unique.
R EMARK . This result only holds for commutative rings.
P ROOF.
⇐ The module R = R1 is cyclic, so for any ideal I � R the quotient R/I is also cyclic
(cf. corollary 3.1.16).
⇐ Let M be a cyclic R–module, then there is some x ∈ M such that M = Rx. Define
α : R → Rx = M by α(r) := rx. For any a, b ∈ R R, r, s ∈ R we have
α(ra + sb) = (ra + sb)x = rax + sbx = r(ax) + s(bx) = rα(a) + sα(b),
thus α ∈ HomR (R R, Rx) is a module homomorphism. As M is cyclic, we get Im α = M ,
i.e. α is surjective. By the first isomorphism theorem for modules, we get M ∼
= R/ Ker α,
where Ker α ≤ R R is a submodule of R. But we know that submodules of R are precisely
ideals, thus Ker α = I � R, and thus M ∼ = R/I.
It remains to prove the uniqueness of the ideal I. Suppose M ∼ = R/I ∼= R/J for two

ideals I, J � R. Then there is some R–module isomorphism β : R/I −→ R/J (note that
β does not need to be a ring homomorphism). As β is surjective, there must exist some
r = r + I such that β(r + I) = 1 + J ∈ R/J. For any i ∈ I, on ehas ir = ir = 0, as
ir ∈ I, thus
0 + J = 0 = β(0 + I) = β(ir) = iβ(r) = i(1 + J) = i + J,
and hence i ∈ J, so we obtain I ⊆ J. Doing the same reasoning with the inverse isomor-
phism β −1 : R/J → R/I one gets J ⊆ I, and therefore I = J. �
3.1. SOME GENERALITIES ABOUT MODULES 27

D EFINITION 3.1.30. Let R M be an R–module, X ⊆ M a nonempty subset of M .


The annihilator of X is defined by
ann(X) := {r ∈ R | rx = 0 ∀x ∈ X} .
P ROPOSITION 3.1.31. For R M R–module and X ⊆ M , the annihilator ann(X) is
an ideal of R.
P ROOF. Proof is straightforward and left as an exercise. �
R EMARK 3.1.32. If Rx ⊆ M is a cyclic submodule of an R–module M , and we
define the map α : R → Rx by α(r) := rx, as in the proof of the previous theorem, then
Ker α is precisely the annihilator ann(x). In other words, we have
R
Rx ∼
= .
ann(x)

P ROPOSITION 3.1.33. For any subset X ⊆ M , one has ann(X) = x∈X ann(x).
P ROOF. Exercise. �
E XAMPLES 3.1.34.
(1) Let (M, +) be an abelian group, 0 �= x ∈ M an element of order n, i.e. n is
the smallest positive integer such that nx = 0. Then ann(x) = (n) � Z, and
Zx = Z/(n) = Zn .
(2) Let F M be a module over a field F (i.e. M is a vector space over F), and let
0 �= x ∈ M be a nonzero vector. Suppose that there is some scalar λ ∈ F
such that λ ∈ ann(x), i.e. λx = 0. As x is a nonzero vector, the set {x} is
linearly independent, so if λx = 0 it must be λ = 0, hence ann(x) = 0, and
Fx ∼= F/0 ∼= F as a vector space.
(3) Let R be an ID, 0 �= x ∈ R, then ann(x) = 0 (prove it as an exercise!), and
hence (x) = Rx ∼ = R/ =∼ = R as R–modules.
T HEOREM 3.1.35 (Second Isomorphism Theorem for modules). Let R M be an R–
module, A, B ≤ M submodules, then
A+B ∼ B
= .
A A∩B
P ROOF. The proof is similar to the one for rings. �
T HEOREM 3.1.36 (Third Isomorphism Theorem for modules). Let R M be an R–
module, P ≤ M a submodule. There is a bijection between the set {Q ≤ M | P ⊆ Q} of
submodules of M containing P , and the set of submodules of M/P , given by Q �→ Q/P .
Moreover, one has
M/P ∼
= M/Q.
Q/P
3.1.4. Direct sum of modules. We have already classified cyclic modules, which will
be our basic “building blocks”. Now, we shall study the construction that will allow us to
glue these building block together.
D EFINITION 3.1.37. Let M1 , . . . , Mn be R–modules. Let M be the set
M = {(m1 , . . . , mn ) | mi ∈ Mi }
of ordered n–tuples in M1 × · · · × Mn . In the set M , define
(m1 , . . . , mn ) + (m�1 , . . . , m�n ) := (m1 + m�1 , . . . , mn + m�n ),
0 = (0M1 , . . . , 0Mn ),
−(m1 , . . . , mn ) := (−m1 , . . . , −mn ),
r(m1 , . . . , mn ) := (rm1 , . . . , rmn ).
28 3. MODULES

Endowed with these operations, M becomes an R–module, called the external direct sum
of the modules Mi . We will represent this as
n

M = M1 ⊕ · · · ⊕ M n = Mi .
i=1

Given an external direct sum M = Mi , we will denote by
Mi� := {(0, . . . , mi , 0, . . . , 0) | mi ∈ Mi } ,
where the only possibly nonzero coordinate is in position i. Obviously Mi� is a submodule
of M , and the maps mi ←→ (0, . . . , mi , 0, . . . , 0) provide a module isomorphism Mi ∼
=
Mi� . Henceforth, we can identify the modules Mi with the Mi� and thus think of Mi as
submodules of M .
Question: Given an R–module R M and submodules Mi ≤ M , when can we ensure that
M∼ = M1 ⊕ · · · ⊕ Mn ? What properties must the submodules Mi satisfy?
If it is the case that we can rewrite a module M as a direct sum of certain submodules
Mi ≤ M , we say that M is� an internal direct sum of the Mi ’s.
Firstly, let m ∈ M = Mi , then
m = (m1 , . . . , mn ) = (m1 , 0, . . . , 0) + · · · + (0, . . . , 0, mn ),

and hence we get M = M1 + · · · + Mn = Mi . Secondly, if mi ∈ Mi then mi =
(0, . . . , 0, mi , 0, . . . , 0), and m1 +· · ·+mn = 0 if and only if (m1 , . . . , mn ) = (0, . . . , 0),
i.e if and only if mi = 0 for all i. This condition leads us to the following notion:
D EFINITION 3.1.38. Let R M be an R–module, M1 , . . . , Mn ≤ M submodules of M .
We say that {Mi }ni=1 is an independent set of submodules if whenever m1 +· · ·+mn = 0
for some mi ∈ Mi , then one must have mi = 0 for all i = 1, . . . , n.

R EMARK 3.1.39. From what we have stated above, if M = Mi , then {Mi } is
always an independent set of submodules. This also means that “being independent” is not
an intrinsic property of the submodules, but rather of the way they are included inside M .
P ROPOSITION 3.1.40. Let R M be an R–module, Mi ≤ M for i = 1, . . . , n submod-
ules. The following are equivalent:
(1) {M1 , . . . , Mn } is an independent set of modules.
(2) Any m ∈ M1 + · · · + Mn can be written as m = m1 + · · · + mn for unique
mi ∈ Mi .
(3) For each i = 1, . . . , n one has Mi ∩ M �i = � i �= jMi .
�i = 0, where M

P ROOF.
1 ⇒ 2 If m = m1 +· · ·+mn = m�1 +· · ·+m�n m then 0 = (m1 −m�1 )+· · ·+(mn −m�n ),
where m − i − m�i ∈ Mi , thus mi − m�i = 0 and thus mi = m�i for all i = 1, . . . , mn , i.e.
the mi are unique.
2 ⇒ 3 Let m ∈ Mi ∩ M �i , then m = mi = m1 + · · · + mi−1 + mi+1 + · · · + mn , thus
m1 + · · · + mi−1 − mi + mi+1 + · · · + mn = 0, and hence m1 = · · · = mn = 0, implying
m = 0.
3⇒1 �

E XAMPLE 3.1.41. Let R M be an R–module, A, B ≤ M submodules, then {A, B} is


an independent set of modules if and only if A ∩ B = 0.
P ROPOSITION 3.1.42. Let R M be an R–module, M1 , . . . , Mn ≤ M submodule. The
following are equivalent:
�n
(1) M = � i=1 Mi .
n
(2) M = i=1 Mi and {Mi } is an independent set of modules.
3.1. SOME GENERALITIES ABOUT MODULES 29

P ROOF.
1 ⇒ 2 Already shown above. �
1 ⇒ 1 Define α : M → Mi by m �→ (m1 , . . . , mn ), where m1 , . . . , mn are the
unique elements with mi ∈ Mi such that m = m1 + · · · + mn . Clearly, α is a surjective
module homomorphism. Let m ∈ Ker α, then α(m) = 0 = (0, . . . , 0) and thus m =
0 + 0 + · · · + 0 = 0, hence α is injective and therefore an isomorphism.

E XAMPLE 3.1.43. Let R M be an R–module, A, B submodules, then M = A ⊕ B if
and only if M = A + B and A ∩ B = 0. In this case, A and B are called direct summands
of M and B is called a complement of B in M .
R EMARK 3.1.44. If M = A ⊕ B, then using the second isomorphism theorem we
have
M A⊕B A+B ∼ B B ∼
= = = = = B.
A A A A∩B 0
Similarly, we obtain (A ⊕ B)/B ∼
= A.
D EFINITION 3.1.45. If M1 = M2 = · · · = Mn = M , we will represent the external
direct sum M1 ⊕ · · · ⊕ Mn = M ⊕ · · · ⊕ M simply by M n .
3.1.5. Free modules.
(n)
D EFINITION 3.1.46. Let R be a ring. A module of the form R Rn = R R ⊕ · · · ⊕ R R
is called a free module over R.
D EFINITION 3.1.47. Let {e1 , . . . , en } ⊆ M be a subset of an R–module M . We say
that the set {e1 , . . . , en } is a basis for M if for all m ∈ M there exist unique ri ∈ R such
that m = r1 e1 + · · · + rn en .
P ROPOSITION 3.1.48. Let R M be an R–module, the following are equivalent:
(1) M is a free module.
(2) M has a basis.
P ROOF.
1 ⇒ 2: Suppose M = Rn is a free module, let ei = (0, . . . , 0, 1, 0, . . . , 0), where the 1
is in the i–th position. If m ∈ M = Rn , then m = (r1 , . . . , rn ) with ri ∈ R, and hence
m = r1 e1 + · · · + rn en for unique ri ∈ R.
2 ⇒ 1: Let {e1 , . . . , en } be a basis of M , then for each m ∈ M there exist unique ri ∈ R
such that m = r1 e1 + · · · + rn en . Define the map α : M → Rn by α(m) := (r1 , . . . , rn ).
It is easy to check that this is an R–module isomorphism, showing that M ∼ = Rn . �
R EMARK 3.1.49. If {ei } is a basis of M and r ∈ R is such that rei = 0, then
0 = 0e1 + · · · + 0ei−1 + rei + 0en+1 + · · · + 0en , and hence r = 0, thus ann(ei ) = 0 and
Rei ∼
= R/(0) = R. Henceforth, for any module with a basis we have M ∼ = Re1 ⊕ · · · ⊕
(n)
∼ R ⊕ ··· ⊕ R = R .
Ren = n

P ROPOSITION 3.1.50. Let F = Rn be a free module with basis {e1 , . . . , en }. Let


R M be a module over R, and let m1 , . . . , mn ∈ M . Then, there is a unique R–module
homomorphism α : F → M such that α(ei ) = mi .
R EMARK . In other words, the proposition states that any module homomorphism
having as a source a free module is completely determined by its effect on the elements of
a basis.
P ROOF. Let u = r1 e1 + · · · + rn en . Assume that α is a module homomorphism such
that α(ei ) = mi ; then, necessarily we must have
α(u) = α(r1 e1 + · · · + rn en ) = r1 α(e1 ) + · · · + rn α(en ) = r1 m1 + · · · + rn mn ,
30 3. MODULES

and hence α is uniquely defined. Thus, define α : F → M by α(r1 e1 + · · · + rn en ) :=


r1 m1 + · · · + rn mn ; we only need to show that α is a module homomorphism.
Let r, s ∈ R, u = u1 e1 + · · · + un en , v = v1 e1 + · · · + vn en ∈ F , then
�� �
α(ru + sv) = α (rui + svi )ei

= (rui + svi )mi
� �
= rui mi + svi mi
� �
= r u i mi + s vi mi
= rα(u) + sα(v),
therefor α is an R–module homomorphism, and clearly it verifies α(ei ) = mi . �
P ROPOSITION 3.1.51. Let R M be a finitely generated R–module, so that M =
Rm1 + · · · Rmn for some mi ∈ M ; then, there exists a free module F and a submod-
ule P ≤ F such that M ∼= F/P .
P ROOF. Leet F = Rn , consider the unique module homomorphism α : F → M
satisfying α(ei ) = mi . If m ∈ M , then there exist ri ∈ R such that m = r1 m1 + · · · +
rn mn , and hence
� �
α(r1 e1 + · · · rn en ) = ri α(ei ) = ri mi = m,
thus α is surjective, i.e. Im α = M . Applying the first isomorphism theorem for modules,
we conclude that M ∼ = F/ Ker α, where Ker α is a submodule of F . �

Вам также может понравиться