Basics of Algebra, Topology, and Differential Calculus by Jean Gallier - HTML preview

PLEASE NOTE: This is an HTML preview only and some elements such as links or page numbers may be incorrect.
Download the book in PDF, ePub, Kindle for a complete version.

k−1

k

k−j−1

Xk1 − f = (X1 − f)

(X1)jf

,

j=0

24.5. THE TORSION MODULE ASSOCIATED WITH AN ENDOMORPHISM

689

and

k−1

k−j−1

z = (X1 − f)

(X1)jf

(1 ⊗ uk) ,

k

j=0

which shows that z = ψ(y) for some y ∈ E[X].

Finally, we prove that ψ is injective as follows. We have

ψ(z) = ψ

Xk ⊗ uk

k

= (X1 − f)

Xk ⊗ uk

k

=

Xk+1 ⊗ (uk − f(uk+1)),

k

where (uk) is a family of finite support of uk ∈ E. If ψ(z) = 0, then

Xk+1 ⊗ (uk − f(uk+1)) = 0,

k

and because the Xk form a basis of A[X], we must have

uk − f(uk+1) = 0, for all k.

Since (uk) has finite support, there is a largest k, say m + 1 so that um+1 = 0, and then from

uk = f (uk+1),

we deduce that uk = 0 for all k. Therefore, z = 0, and ψ is injective.

Remark: The exact sequence of Theorem 24.21 yields a presentation of Mf .

Since A[X] is a free A-module, A[X]⊗AM is a free A-module, but A[X]⊗AM is generally

not a free A[X]-module. However, if M is a free module, then M [X] is a free A[X]-module,

since if (ui)i∈I is a basis for M, then (1 ⊗ ui)i∈I is a basis for M[X]. This allows us to define

the characterisctic polynomial χf (X) of an endomorphism of a free module M as

χf (X) = det(X1 − f).

Note that to have a correct definition, we need to define the determinant of a linear map

allowing the indeterminate X as a scalar, and this is what the definition of M [X] achieves

(among other things). Theorem 24.21 can be used to quick a short proof of the Cayley-

Hamilton Theorem, see Bourbaki [12] (Chapter III, Section 8, Proposition 20). Proposition

5.10 is still the crucial ingredient of the proof.

We now develop the theory necessary to understand the structure of finitely generated

modules over a PID.

690

CHAPTER 24. INTRODUCTION TO MODULES; MODULES OVER A PID

24.6

Torsion Modules over a PID; The Primary

Decomposition

We begin by considering modules over a product ring obtained from a direct decomposition,

as in Definition 21.3. In this section and the next, we closely follow Bourbaki [13] (Chapter

VII). Let A be a commutative ring and let (b1, . . . , bn) be ideals in A such that there is

an isomorphism A ≈ A/b1 × · · · × A/bn. From Theorem 21.16 part (b), there exist some

elements e1, . . . , en of A such that

e2i = ei

eiej = 0,

i = j

e1 + · · · + en = 1A,

and bi = (1A − ei)A, for i, j = 1, . . . , n.

Given an A-module M with A ≈ A/b1×· · ·×A/bn, let Mi be the subset of M annihilated

by bi; that is,

Mi = {x ∈ M | bx = 0, for all b ∈ bi}.

Because bi is an ideal, each Mi is a submodule of M. Observe that if λ, µ ∈ A, b ∈ bi, and

if λ − µ = b, then for any x ∈ Mi, since bx = 0,

λx = (µ + b)x = µx + bx = µx,

so Mi can be viewed as a A/bi- module.

Proposition 24.22. Given a ring A ≈ A/b1 × · · · × A/bn as above, the A-module M is the

direct sum

M = M1 ⊕ · · · ⊕ Mn,

where Mi is the submodule of M annihilated by bi.

Proof. For i = 1, . . . , n, let pi : M → M be the map given by

pi(x) = eix,

x ∈ M.

The map pi is clearly linear, and because of the properties satisfied by the eis, we have

p2i = pi

pipj = 0,

i = j

p1 + · · · + pn = id.

This shows that the pi are projections, and by Proposition 4.6 (which also holds for modules),

we have a direct sum

M = p1(M) ⊕ · · · ⊕ pn(M) = e1M ⊕ · · · ⊕ enM.

24.6. TORSION MODULES OVER A PID; PRIMARY DECOMPOSITION

691

It remains to show that Mi = eiM. Since (1 − ei)ei = ei − e2i = ei − ei = 0, we see that

eiM is annihilated by bi = (1 − ei)A. Furthermore, for i = j, for any x ∈ M, we have

(1 − ei)ejx = (ej − eiej)x = ejx, so no nonzero element of ejM is annihilated by 1 − ei, and

thus not annihilated by bi. It follows that eiM = Mi, as claimed.

Given an A-module M , for any nonzero α ∈ A, let

M (α) = {x ∈ M | αx = 0},

the submodule of M annihilated by α. If α divides β, then M (α) ⊆ M(β), so we can define

Mα =

M (αn) = {x ∈ M | (∃n ≥ 1)(αnx = 0)},

n≥1

the submodule of M consisting of all elements of M annihilated by some power of α. If N

is any submodule of M , it is clear that

Nα = M ∩ Mα.

Recall that in a PID, an irreducible element is also called a prime element.

Definition 24.8. If A is a PID and p is a prime element in A, we say that a module M is

p-primary if M = Mp.

Proposition 24.23. Let M be module over a PID A. For every nonzero α ∈ A, if

α = upn1

1 · · · pnr

r

is a factorization of α into prime factors (where u is a unit), then the module M (α) anni-

hilated by α is the direct sum

M (α) = M (pn1

1 ) ⊕ · · · ⊕ M (pnr

r ).

Furthermore, the projection from M (α) onto M (pni) is of the form x

i

→ γix, for some γi ∈ A,

and

M (pni) = M (α)

.

i

∩ Mpi

Proof. First, observe that since M (α) is annihilated by α, we can view M (α) as a A/(α)-

module. By the Chinese Remainder Theorem (Theorem 21.15) applied to the ideals (upn1

1 ) =

(pn1

1 ), (pn2

2 ), . . . , (pnr

r ), we have an isomorphism

A/(α) ≈ A/(pn1

1 ) × · · · × A/(pnr

r ).

Since we also have isomorphisms

A/(pni)

)/(α)),

i

≈ (A/(α))/((pni

i

692

CHAPTER 24. INTRODUCTION TO MODULES; MODULES OVER A PID

we can apply Proposition 24.22, and we get a direct sum

M (α) = N1 ⊕ · · · ⊕ Nr,

where Ni is the A/(α)-submodule of M(α) annihilated by (pni)/(α), and the projections

i

onto the Ni are of the form stated in the proposition. However, Ni is just the A-module

M (pni) annihilated by pni, because every nonzero element of (pni)/(α) is an equivalence class

i

i

i

modulo (α) of the form apni for some nonzero a

i

∈ A, and by definition, x ∈ Ni iff

0 = apni x = apnix,

for all a

i

i

∈ A − {0},

in particular for a = 1, which implies that x ∈ M(pni).

i

The inclusion M (pni)

is clear. Conversely, pick x

, which

i

⊆ M(α) ∩ Mp

∈ M(α) ∩ M

i

pi

means that αx = 0 and psix = 0 for some s ≥ 1. If s < ni, we are done, so assume s ≥ ni.

Since pni is a gcd of α and ps

i

i , by Bezout, we can write

pni = λps

i

i + µα

for some λ, µ ∈ A, and then pnix = λps

), as

i

i x + µαx = 0, which shows that x ∈ M (pni

i

desired.

Recall that if M is a torsion module over a ring A which is an integral domain, then

every finite set of elements x1, . . . , xn in M is annihilated by a = a1 · · · an, where each ai

annihilates xi.

Since A is a PID, we can pick a set P of irreducible elements of A such that every nonzero

nonunit of A has a unique factorization up to a unit. Then, we have the following structure

theorem for torsion modules which holds even for modules that are not finitely generated.

Theorem 24.24. (Primary Decomposition Theorem) Let M be a torsion-module over a

PID. For every irreducible element p ∈ P , let Mp be the submodule of M annihilated by

some power of p. Then, M is the (possibly infinite) direct sum

M =

Mp.

p∈P

Proof. Since M is a torsion-module, for every x ∈ M, there is some α ∈ A such that

x ∈ M(α). By Proposition 24.23, if α = upn1

1 · · · pnr

r

is a factorization of α into prime factors

(where u is a unit), then the module M (α) is the direct sum

M (α) = M (pn1

1 ) ⊕ · · · ⊕ M (pnr

r ).

This means that x can be written as

x =

xp,

xp ∈ Mp,

p∈P

24.6. TORSION MODULES OVER A PID; PRIMARY DECOMPOSITION

693

with only finitely many xp nonzero. If

xp =

yp

p∈P

p∈P

for all p ∈ P , with only finitely many xp and yp nonzero, then xp and yp are annihilated by

some common nonzero element a ∈ A, so xp, yp ∈ M(a). By Proposition 24.23, we must

have xp = yp for all p, which proves that we have a direct sum.

It is clear that if p and p are two irreducible elements such that p = up for some unit u,

then Mp = Mp . Therefore, Mp only depends on the ideal (p).

Definition 24.9. Given a torsion-module M over a PID, the modules Mp associated with

irreducible elements in P are called the p-primary components of M .

The p-primary components of a torsion module uniquely determine the module, as shown

by the next proposition.

Proposition 24.25. Two torsion modules M and N over a PID are isomorphic iff for

every every irreducible element p ∈ P , the p-primary components Mp and Np of M and N

are isomorphic.

Proof. Let f : M → N be an isomorphism. For any p ∈ P , we have x ∈ Mp iff pkx = 0 for

some k ≥ 1, so

0 = f (pkx) = pkf (x),

which shows that f (x) ∈ Np. Therefore, f restricts to a linear map f | Mp from Mp to

Np. Since f is an isomorphism, we also have a linear map f −1 : M → N, and our previous

reasoning shows that f −1 restricts to a linear map f −1 | Np from Np to Mp. But, f | Mp and

f −1 | Np are mutual inverses, so Mp and Np are isomorphic.

Conversely, if Mp ≈ Np for all p ∈ P , by Theorem 24.24, we get an isomorphism between

M =

M

N

p∈P

p and N =

p∈P

p.

In view of Proposition 24.25, the direct sum of Theorem 24.24 in terms of its p-primary

components is called the canonical primary decomposition of M .

If M is a finitely generated torsion-module, then Theorem 24.24 takes the following form.

Theorem 24.26. (Primary Decomposition Theorem for finitely generated torsion modules)

Let M be a finitely generated torsion-module over a PID A. If Ann(M ) = (a) and if a =

upn1

1 · · · pnr

r

is a factorization of a into prime factors, then M is the finite direct sum

r

M =

M (pni).

i

i=1

Furthermore, the projection of M over M (pni) is of the form x

i

→ γix, for some γi ∈ A.

694

CHAPTER 24. INTRODUCTION TO MODULES; MODULES OVER A PID

Proof. This is an immediate consequence of Proposition 24.23.

In particular, Theorem 24.26 applies when A = Z. In this case, M is a finitely generated

torsion abelian group, and the theorem says that such a group is the direct sum of a finite

number of groups whose elements have order some power of a prime number p.

Theorem 24.24 has several useful corollaries.

Proposition 24.27. If M is a torsion module over a PID, for every submodule N of M ,

we have a direct sum

N =

N ∩ Mp.

p∈P

Proof. It is easily verified that N ∩ Mp is the p-primary component of N.

Proposition 24.28. If M is a torsion module over a PID, a submodule N of M is a direct

factor of M iff Np is a direct factor of Mp for every irreducible element p ∈ A.

Proof. This is because if N and N are two submodules of M , we have M = N ⊕ N iff, by

Proposition 24.27, Mp = Np ⊕ Np for every irreducible elements p ∈ A.

An A-module M is said to be semi-simple iff for every submodule N of M , there is some

submodule N of M such that M = N ⊕ N .

Proposition 24.29. Let A be a PID which is not a field, and let M be any A-module. Then,

M is semi-simple iff it is a torsion module and if Mp = M(p) for every irreducible element

p ∈ A (in other words, if x ∈ M is annihilated by a power of p, then it is already annihilated

by p).

Proof. Assume that M is semi-simple. Let x ∈ M and pick any irreducible element p ∈ A.

Then, the submodule pAx has a supplement N such that

M = pAx ⊕ N,

so we can write x = pax + y, for some y ∈ N and some a ∈ A. But then,

y = (1 − pa)x,

and since p is irreducible, p is not a unit, so 1 − pa = 0. Observe that

p(1 − ap)x = py ∈ pAx ∩ N = (0).

Since p(1 − ap) = 0, x is a torsion element, and thus M is a torsion module. The above

argument shows that

p(1 − ap)x = 0,

24.7. FINITELY GENERATED MODULES OVER A PID

695

which implies that px = ap2x, and by induction,

px = anpn+1x,

for all n ≥ 1.

If we pick x in Mp, then there is some m ≥ 1 such that pmx = 0, and we conclude that

px = 0.

Therefore, Mp = M(p), as claimed.

Conversely, assume that M is a torsion-module and that Mp = M(p) for every irreducible

element p ∈ A. By Proposition 24.28, it is sufficient to prove that a module annihilated by a

an irreducible element is semi-simple. This is because such a module is a vector space over

the field A/(p) (recall that in a PID, an ideal (p) is maximal iff p is irreducible), and in a

vector space, every subspace has a supplement.

Theorem 24.26 shows that a finitely generated torsion module is a direct sum of p-primary

modules Mp. We can do better. In the next section, we show that each primary module Mp

is the direct sum of cyclic modules of the form A/(pn).

24.7

Finitely Generated Modules over a PID; Invariant

Factor Decomposition

There are several ways of obtaining the decomposition of a finitely generated module as a

direct sum of cyclic modules. One way to proceed is to first use the Primary Decomposition

Theorem and then to show how each primary module Mp is the direct sum of cyclic modules of

the form A/(pn). This is the approach followed by Lang [65] (Chapter III, section 7), among

others. We prefer to use a proposition that produces a particular basis for a submodule of

a finitely generated free module, because it yields more information. This is the approach

followed in Dummitt and Foote [30] (Chapter 12) and Bourbaki [13] (Chapter VII). The

proof that we present is due to Pierre Samuel.

Proposition 24.30. Let F be a finitely generated free module over a PID A, and let M be

any submodule of F . Then, M is a free module and there is a basis (e1, ..., en) of F , some

q ≤ n, and some nonzero elements a1, . . . , aq ∈ A, such that (a1e1, . . . , aqeq) is a basis of M

and ai divides ai+1 for all i, with 1 ≤ i ≤ q − 1.

Proof. The proposition is trivial when M = {0}, thus assume that M is nontrivial. Pick some

basis (u1, . . . , un) for F . Let L(F, A) be the set of linear forms on F . For any f ∈ L(F, A),

it is immediately verified that f (M ) is an ideal in A. Thus, f (M ) = ahA, for some ah ∈ A,

since every ideal in A is a principal ideal. Since A is a PID, any nonempty family of ideals

in A has a maximal element, so let f be a linear map such that ahA is a maximal ideal in A.

Let πi : F → A be the i-th projection, i.e., πi is defined such that πi(x1u1 + · · · + xnun) = xi.

696

CHAPTER 24. INTRODUCTION TO MODULES; MODULES OVER A PID

It is clear that πi is a linear map, and since M is nontrivial, one of the πi(M) is nontrivial,

and ah = 0. There is some e ∈ M such that f(e ) = ah.

We claim that, for every g ∈ L(F, A), the element ah ∈ A divides g(e ).

Indeed, if d is the gcd of ah and g(e ), by the Bézout identity, we can write

d = rah + sg(e ),

for some r, s ∈ A, and thus

d = rf (e ) + sg(e ) = (rf + sg)(e ).

However, rf + sg ∈ L(F, A), and thus,

ahA ⊆ dA ⊆ (rf + sg)(M),

since d divides ah, and by maximality of ahA, we must have ahA = dA, which implies that

d = ah, and thus, ah divides g(e ). In particular, ah divides each πi(e ) and let πi(e ) = ahbi,

with bi ∈ A.

Let e = b1u1 + · · · + bnun. Note that

e = π1(e )u1 + · · · + πn(e )un = ahb1u1 + · · · + ahbnun,

and thus, e = ahe. Since ah = f (e ) = f (ahe) = ahf (e), and since ah = 0, we must have

f (e) = 1.

Next, we claim that

F = Ae ⊕ f−1(0)

and

M = Ae ⊕ (M ∩ f−1(0)),

with e = ahe.

Indeed, every x ∈ F can be written as

x = f (x)e + (x − f(x)e),

and since f (e) = 1, we have f (x − f(x)e) = f(x) − f(x)f(e) = f(x) − f(x) = 0. Thus,

F = Ae + f −1(0). Similarly, for any x ∈ M, we have f(x) = rah, for some r ∈ A, and thus,

x = f (x)e + (x − f(x)e) = rahe + (x − f(x)e) = re + (x − f(x)e),

we still have x − f(x)e ∈ f−1(0), and clearly, x − f(x)e = x − rahe = x − re ∈ M, since

e ∈ M. Thus, M = Ae + (M ∩ f−1(0)).

To prove that we have a direct sum, it is enough to prove that Ae ∩ f−1(0) = {0}. For

any x = re ∈ Ae, if f(x) = 0, then f(re) = rf(e) = r = 0, since f(e) = 1 and, thus, x = 0.

Therefore, the sums are direct sums.

24.7. FINITELY GENERATED MODULES OVER A PID

697

We can now prove that M is a free module by induction on the size, q, of a maximal

linearly independent family for M .

If q = 0, the result is trivial. Otherwise, since

M = Ae ⊕ (M ∩ f−1(0)),

it is clear that M ∩ f−1(0) is a submodule of F and that every maximal linearly independent

family in M ∩ f−1(0) has at most q − 1 elements. By the induction hypothesis, M ∩ f−1(0)

is a free module, and by adding e to a basis of M ∩ f−1(0), we obtain a basis for M, since

the sum is direct.

The second part is shown by induction on the dimension n of F .

The case n = 0 is trivial. Otherwise, since

F = Ae ⊕ f−1(0),

and since, by the previous argument, f −1(0) is also free, f −1(0) has dimension n − 1. By

the induction hypothesis applied to its submodule M ∩ f−1(0), there is a basis (e2, . . . , en)

of f −1(0), some q ≤ n, and some nonzero elements a2, . . . , aq ∈ A, such that, (a2e2, . . . , aqeq)

is a basis of M ∩ f−1(0), and ai divides ai+1 for all i, with 2 ≤ i ≤ q − 1. Let e1 = e, and

a1 = ah, as above. It is clear that (e1, . . . , en) is a basis of F , and that that (a1e1, . . . , aqeq)

is a basis of M , since the sums are direct, and e = a1e1 = ahe. It remains to show that a1

divides a2. Consider the linear map g : F → A such that g(e1) = g(e2) = 1, and g(ei) = 0,

for all i, with 3 ≤ i ≤ n. We have ah = a1 = g(a1e1) = g(e ) ∈ g(M), and thus ahA ⊆ g(M).

Since ahA is maximal, we must have g(M) = ahA = a1A. Since a2 = g(a2e2) ∈ g(M), we

have a2 ∈ a1A, which shows that a1 divides a2.

We need the following basic proposition.

Proposition 24.31. For any commutative ring A, if F is a free A-module and if (e1, . . . , en)

is a basis of F , for any elements a1, . . . , an ∈ A, there is an isomorphism

F/(Aa1e1 ⊕ · · · ⊕ Aanen) ≈ (A/a1A) ⊕ · · · ⊕ (A/anA).

Proof. Let σ : F → A/(a1A) ⊕ · · · ⊕ A/(anA) be the linear map given by

σ(x1e1 + · · · + xnen) = (x1, . . . , xn),

where xi is the equivalence class of xi in A/aiA. The map σ is clearly surjective, and its

kernel consists of all vectors x1e1 + · · · + xnen such that xi ∈ aiA, for i = 1, . . . , n, which

means that

Ker (σ) = Aa1e1 ⊕ · · · ⊕ Aanen.

Since M/Ker (σ) is isomorphic to Im(σ), we get the desired isomorphism.

698

CHAPTER 24. INTRODUCTION TO MODULES; MODULES OVER A PID

We can now prove the existence part of the structure theorem for finitely generated

modules over a PID.

Theorem 24.32. Let M be a finitely generated nontrivial A-module, where A a PID. Then,

M is isomorphic to a direct sum of cyclic modules

M ≈ A/a1 ⊕ · · · ⊕ A/am,

where the ai are proper ideals of A (possibly zero) such that

a1 ⊆ a2 ⊆ · · · ⊆ am = A.

More precisely, if a1 = · · · = ar = (0) and (0) = ar+1 ⊆ · · · ⊆ am = A, then

M ≈ Ar ⊕ (A/ar+1 ⊕ · · · ⊕ A/am),

where A/ar+1 ⊕ · · · ⊕ A/am is the torsion submodule of M. The module M is free iff r = m,

and a torsion-module iff r = 0. In the latter case, the annihilator of M is a1.

Proof. Since M is finitely generated and nontrivial, there is a surjective homomorphism

ϕ : An → M for some n ≥ 1, and M is isomorphic to An/Ker (ϕ). Since Ker (ϕ) is a submod-

ule of the free module An, by Proposition 24.30, Ker (ϕ) is a free module and there is a basis

(e1, . . . , en) of An and some nonzero elements a1, . . . , aq (q ≤ n) such that (a1e1, . . . , aqeq) is

a basis of Ker (ϕ) and a1 | a2 | · · · | aq. Let aq+1 = . . . = an = 0.

By Proposition 24.31, we have an isomorphism

An/Ker (ϕ) ≈ A/a1A ⊕ · · · ⊕ A/anA.

Whenever ai is unit, the factor A/aiA = (0), so we can weed out the units. Let r = n − q,

and let s ∈ N be the smallest index such that as+1 is not a unit. Note that s = 0 means that

there are no units. Also, as M = (0), s < n. Then,

M ≈ An/Ker (ϕ) ≈ A/as+1A ⊕ · · · ⊕ A/anA.

Let m = r + q − s = n − s. Then, we have the sequence

as+1, . . . , aq, aq+1, . . . , an,

q−s

r=n−q

where as+1 | as+2 | · · · | aq are nonzero and nonunits and aq+1 = · · · = an = 0, so we define

the m ideals ai as follows:

(0)

if 1 ≤ i ≤ r

ai =

ar+q+1−iA if r + 1 ≤ i ≤ m.

With these definitions, the ideals ai are proper ideals and we have

ai ⊆ ai+1,

i = 1, . . . , m − 1.

When r = 0, since as+1 | as+2 | · · · | an, it is clear that a1 = anA is the annihilator of M.

The other statements of the theorem are clear.

24.7. FINITELY GENERATED MODULES OVER A PID

699

The natural number r is called the free rank or Betti number of the module M . The

generators α1, . . . , αm of the ideals a1, . . . , am (defined up to a unit) are often called the

invariant factors of M (in the notation of Theorem 24.32, the generators of the ideals

a1, . . . , am are denoted by aq, . . . , as+1, s ≤ q).

As corollaries of Theorem 24.32, we obtain again the following facts established in Section

24.1:

1. A finitely generated module over a PID is the direct sum of its torsion module and a

free module.

2. A finitely generated torsion-free module over a PID is free.

It turns out that the ideals a1 ⊆ a2 ⊆ · · · ⊆ am = A are uniquely determined by the

module M . Uniqueness proofs found in most books tend to be intricate and not very intuitive.

The shortest proof that we are aware of is from Bourbaki [13] (Chapter VII, Section 4), and

uses wedge products.

The following preliminary results are needed.

Proposition 24.33. If A is a commutative ring and if a1, . . . , am are ideals of A, then there

is an isomorphism

A/a1 ⊗ · · · ⊗ A/am ≈ A/(a1 + · · · + am).

Sketch of proof. We proceed by induction on m. For m = 2, we define the map

ϕ : A/a1 × A/a2 → A/(a1 + a2) by

ϕ(a, b) = ab (mod a1 + a2).

It is well-defined because if a = a + a1 and b = b + a2 with a1 ∈ a1 and a2 ∈ a2, then

a b = (a + a1)(b + a2) = ab + ba1 + aa2 + a1a2,

and so

a b ≡ ab (mod a1 + a2).

It is also clear that this map is bilinear, so it induces a linear map ϕ : A/a1 ⊗ A/a2 →