• No results found

Group representations and Maschke’s Theorem

N/A
N/A
Protected

Academic year: 2022

Share "Group representations and Maschke’s Theorem"

Copied!
47
0
0

Loading.... (view fulltext now)

Full text

(1)

U.U.D.M. Project Report 2019:21

Examensarbete i matematik, 15 hp Handledare: Martin Herschend

Examinator: Veronica Crispin Quinonez Juni 2019

Group representations and Maschke’s Theorem

Hannes Fors

Department of Mathematics

(2)
(3)

1 Introduction

This paper gives an introduction to group theory and representation theory.

Group theory concerns algebraic structures known as groups, which consist of a set together with a binary operation on the set which fulfils certain axioms, known as the group axioms. The concept of groups originally arose from the study of polynomials, and in particular of permutations of their roots. Later studies in geometry and linear algebra led to interest in groups of various types of transformations.

The introduction of the general, abstract definition of a group was the start of the field of abstract algebra, in which structures are defined by axioms rather than by which types of elements they involve. Groups are still fundamental to abstract algebra, since many other algebraic structures, such as rings and modules, can be viewed as groups that have been expanded with more operations and axioms.

Group representations give us a way of expressing more general groups as groups of linear transformations or matrices. Representation theory has several applications. For example, it may allow us to reduce some problems in group theory to equivalent problems in linear algebra. There may also be circumstances in which we need a more concrete description of a group, and a representation can then provide such a description.

The paper begins with an introduction to group theory, followed by a section on rings and modules. We then introduce group representations and the group algebra and describe how these relate to modules. In the last section we state and prove Maschke’s theorem, an important theorem in representation theory. Overall, we will mostly follow the approach taken in [2], with some modifications. Since [2] does not define modules in gen- eral, instead focusing on modules over the group algebra, we will use the more general definition, which is given in [1], and then rephrase some of the material in [2] to conform with this definition. This approach is preferable since it allows us describe modules as a particular type of group, and then vector spaces as a particular type of module, rather than treating these as completely separate structures.

This text assumes some familiarity with linear algebra and ring theory.

For this reason, rings and vector spaces will only be covered briefly. For more reading on these subjects, refer to [3] for linear algebra and [1] for ring theory.

(4)

2 Groups

In this section we give an introduction to group theory. We begin by stating the definition of a group and giving a few examples. This is followed by subgroups, which are groups contained within other groups, and group mor- phisms, which are maps between groups that preserve the group structure.

All the material in this section is taken from [1].

2.1 Groups

Definition. A group is a set G together with a binary operation

· : G × G −→ G, (a, b) 7−→ a · b := ab for which the following three axioms hold:

(i) ∀a, b, c ∈ G a(bc) = (ab)c (ii) ∃e ∈ G ∀a ∈ G ae = ea = a

(iii) ∀a ∈ G ∃a−1 ∈ G aa−1= a−1a = e

The first axiom states that the operation is associative. The second states the existence of an identity element, and the third states that every element has an inverse. We say that G is a group under · and denote this by (G, ·) or simply by G if the operation is clear. A group is abelian if the operation is commutative, i.e. if ab = ba ∀a, b ∈ G.

Definition. The order of a group G is the number |G| of elements in its underlying set.

Proposition 1.

(i) The identity element is unique.

(ii) The inverse of any a ∈ G is unique.

Proof.

(i) Let e, e0 ∈ G be two identity elements. By the second group axiom, e = ee0 = e0.

(ii) Let a ∈ G and let x, y ∈ G be inverses of a. We then have that x = xe = x(ay) = (xa)y = ey = y

The exact choice of notation for the identity element and inverses will vary depending on which operation is used. For multiplicative groups, we will usually denote the identity and the inverse of a respectively by 1 and a−1, whereas for an additive group we will typically use 0 and −a.

(5)

Example 2.1.

(a) The singleton {x} with x·x := x is a group of order 1, called the trivial group.

(b) (Z,+), (Q,+), (R,+), and (C,+) are all abelian groups of infinite order.

(c) (Z, ·), (Q, ·), (R, ·), and (C, ·) are not groups. All four sets contain 0, and since 0 has no multiplicative inverse, the third group axiom cannot hold.

(d) The set {1, −1, i, −i} is a group under multiplication. Multiplication of complex numbers is associative, the identity is 1 and every element is its own inverse.

(e) Let A be any set of n elements. The set of all permutations of A forms a group under composition called the symmetric group on A, denoted by Sn. The order of Sn is the number of possible permutations of an n-set, which is n!.

(f) The set GLn(R) of all invertible n × n matrices with real coefficients forms a group under matrix multiplication. Since matrix multiplica- tion is not commutative, this is a non-abelian group.

Proposition 2. For any group G and a, b, c, ai ∈ G, the following is true (i) The cancellation laws hold, i.e. ab = ac implies b = c.

(ii) The equations ax = b and ya = b have unique solutions x = a−1b and y = ba−1.

(iii) (a−1)−1 = a and more generally (a1a2· · · an)−1 = a−1n · · · a−12 a−11 for any n ≥ 1.

Proof.

(i) Suppose ab = ac. Then b = 1b = a−1ab = a−1ac = 1c = c.

(ii) Suppose ax = b. We have have at most one solution, and substituting x = a−1b gives ax = aa−1b = 1b = b, so this is the unique solution.

Similarly for ya = b.

(iii) By definition, a−1(a−1)−1 = 1 and a−1a = 1. So a−1(a−1)−1 = a−1a, and by (i) we get that (a−1)−1 = a. For the general case, suppose that

(a1a2· · · ak)−1= a−1k · · · a−12 a−11 for some k ≥ 1. Then, for k + 1 we have that

(a1a2· · · ak+1)−1 = ((a1a2· · · ak)ak+1)−1

= a−1k+1(a1a2· · · ak)−1= a−1k+1a−1k · · · a−12 a−11 and the result now follows by induction.

(6)

Non-negative powers of group elements are defined inductively by a0= 1 and an+1 = a · an for all non-negative integers n. Negative powers are defined by a−n= (a−1)n= (an)−1 where n is a positive integer. As with the identity and inverses, the notation may vary depending on the operation.

For example, in an additive group we may write na instead of an.

Powers of group elements behave in the expected way, as the following proposition shows.

Proposition 3. Let G be a group, a ∈ G and m, n ∈ Z. Then, (i) aman= am+n

(ii) (an)m = amn

(iii) If G is abelian, then (ab)n= anbn Proof.

(i) By definition, aman= aa · · · a

| {z }

m times

· aa · · · a

| {z }

n times

= aa · · · a

| {z }

m + n times

= am+n (ii) By applying (i), (an)m = anan· · · an

| {z }

m times

= an + n + · · · + n

| {z }

m times

= amn (iii) If G is abelian, we can use commutativity and associativity to get that

(ab)n= (ab)(ab) · · · (ab)

| {z }

n times

= aa · · · a

| {z }

n times

· bb · · · b

| {z }

n times

= anbn.

Definition. Let G be a group and a ∈ G. The order of a is the smallest positive integer k such that xk = e. If no such k exists we say that a has infinite order. We denote the order of a by ord(a).

Example 2.2.

For all groups G it holds that ord(a) = 1 if and only if a = e.

Definition. Let (Gi, ·i)i∈I be a collection of groups. Their direct product is the group having as its underlying set the Cartesian product Q

i∈IGi and with the operation defined componentwise by

(ai)i∈I · (bi)i∈I = (ai·ibi)i∈I

If the index set is I = {1, 2, . . . , n} we may write the Cartesian product as G1× G2× · · · × Gn and the operation as

(a1, a2, . . . , an) · (b1, b2, . . . , bn) = (a1·1b1, a2·2b2, . . . , an·nbn)

(7)

2.2 Subgroups

Definition. Let G be a group and let H ⊆ G be a subset such that (i) ∀a, b ∈ H ab ∈ H

(ii) 1G∈ H

(iii) ∀a ∈ H a−1 ∈ H

Then H is a group under the same operation as G. We say that H is a subgroup of G and denote this by H ≤ G. In particular, if H ⊂ G then H is called a proper subgroup, and we write H < G.

The following proposition gives a simpler but equivalent definition of subgroups

Proposition 4. Let G be a group. A subset H ⊆ G is a subgroup of G if and only if H 6= ∅ and ab−1 ∈ H for all a, b ∈ H.

Proof.

(⇒) This follows immediately from the definition of a subgroup.

(⇐) H ⊆ G is given. Since H 6= ∅, there is some a ∈ H, and therefore aa−1 = 1 ∈ H and (ii) holds. Using that 1, a ∈ H, we get that 1 · a−1 = a−1∈ H and (iii) holds. Finally, if a, b ∈ H, then a, b−1∈ H and hence a(b−1)−1 = ab ∈ H and (i) holds.

It follows immediately from the definition that the subgroup relation is transitive, i.e. that if K ≤ H and H ≤ G, then K ≤ G also.

Example 2.3.

(a) Every group G has the trivial subgroup {1G} ≤ G and the entire subgroup G ≤ G.

(b) (Z, +) < (Q, +) < (R, +) < (C, +)

(c) 2Z is a subgroup of the additive group Z.

(d) The set {1, −1, i, −i} is a subgroup of the multiplicative group C \ {0}.

Proposition 5. Let G be a group. If (Hi)i∈I is a collection of subgroups of G, then T

i∈IHi is also a subgroup of G.

Proof. Since Hi ⊆ G for all i we have that T

i∈IHi ⊆ G. Every Hi by definition contains 1G, so the set T

i∈IHi also contains 1G and hence it is non-empty. Lastly, let a, b ∈T

i∈IHi. Then a, b ∈ Hi for all i. Since every Hi is a subgroup, ab−1 ∈ Hi for all i and therefore ab−1 ∈T

i∈IHi. It now follows from Proposition 4 thatT

i∈IHi is a subgroup of G.

(8)

Given a group G and a subset X ⊆ G we define hXi to be the intersection of all subgroups of G which contain X. By the previous proposition, this set is a subgroup, and we call hXi the subgroup generated by X. By definition, this is the smallest subgroup which contains X.

If G = hXi for some X ⊆ G, then we say that G is generated by X.

In particular, if X = {a1, . . . , an} then G is finitely generated and we may write G = ha1, . . . , ani. If G is generated by a single element a it is called a cyclic group and we write G = hai.

The following proposition gives another equivalent definition of gener- ated groups.

Proposition 6. Let G be a group and let X ⊆ G be a non-empty subset.

Define a new set X−1 := {a−1: a ∈ X} and use this to define

P (X ∪ X−1) := {a1· · · an: ai ∈ X ∪ X−1}. Then hXi = P (X ∪ X−1).

Proof. By definition, X ⊆ hXi. Since hXi is a group it follows immediately that X−1 ⊆ hXi and hence P (X ∪ X−1) ⊆ hXi.

Conversely, the set P (X ∪ X−1) is a subgroup of G since it is non-empty and a, b ∈ P (X ∪ X−1) implies ab−1∈ P (X ∪ X−1). It also contains X, so by definition hXi ⊆ P (X ∪ X−1).

Using this proposition, we see that the cyclic group hai has as its under- lying set {an : n ∈ Z}. The next proposition uses this fact to connect the order of an element to the order of a group.

Proposition 7. Let G be a group. For any a ∈ G we have ord(a) = |hai|, i.e. the order of a is the order of the subgroup generated by a.

Proof. Suppose ord(a) is finite and set k = ord(a). Then ak= 1, so the set {an: n ∈ Z} is equivalent to {an: n ∈ Zk}. Hence,

|hai| = |{an: n ∈ Zk}| = |Zk| = k = ord(a)

If ord(a) is infinite, then |hai| = |{an : n ∈ Z}| = |Z| which is also infinite.

Definition. Let G be a group and let H ≤ G be a subgroup. A left coset of H is a subset of G of the form xH = {xh : h ∈ H} for some x ∈ G. The right cosets Hx are defined similarly.

Proposition 8. Let G be a group and let H ≤ G be a subgroup. If a ∈ H, then aH = Ha = HH = H.

Proof. Since a ∈ H, it follows immediately that aH ⊆ H. Conversely, for any a, b ∈ H the equation ax = b has a unique solution x ∈ H, and this shows that H ⊆ aH. So aH = H, and similarly Ha = H. Finally, H = aH ⊆ HH ⊆ H shows that HH = H.

(9)

Proposition 9. Let G be a group and H ≤ G be a subgroup. The set of all left (or right) cosets of H is a partition of G.

Proof. Define a binary relation L on G by L(a, b) if and only a−1b ∈ H. We show that this is an equivalence relation.

• For any a ∈ G we have that a−1a = 1 ∈ H, since H is a subgroup.

Hence L(a, a) holds.

• Suppose L(a, b) holds. Then a−1b ∈ H, and from this it follows that (a−1b)−1 = b−1a ∈ H and L(b, a) holds.

• Suppose that L(a, b) and L(b, c) both hold. Then a−1b, b−1c ∈ H, implying that (a−1b)(b−1c) = a−1(bb−1)c = a−1c ∈ H and L(a, c) holds.

This shows that L is an equivalence relation.

If L(a, b) holds, then a−1b ∈ H which implies that b ∈ aH, so the equivalence class of any a ∈ G is the coset aH. By definition, the equivalence classes of L form a partition of G, and this shows that the set of left cosets of H is a partition of G.

For right cosets we define a relation R by R(a, b) if and only if ab−1 ∈ H.

The rest of the proof is similar.

Definition. Let G be a group and N ≤ G be a subgroup. N is a normal subgroup if if its left and right cosets are equal, that is if aN = N a for all a ∈ G. We denote this by N E G.

Example 2.4.

(a) The trivial subgroup and the entire subgroup are normal for any group.

(b) Every subgroup of an abelian group is normal.

Proposition 10. Let G be a group and N E G be a normal subgroup. Let G/N = {aN : a ∈ G} be the set of all cosets of N in G. Define an operation on G/N by

(aN )(bN ) = (ab)N for all a, b ∈ G

Under this operation, G/N forms a group which we call the quotient group of G by N .

Proof. We need to show that the operation is well-defined and that it satisfies the group axioms.

First, let a, b, x, y ∈ G and assume that aN = xN and bN = yN . We need that (xy)N = (ab)N . By our assumption, there exists n, m ∈ N such that a = xn and b = ym, so we can write ab = (xn)(ym) = x(ny)m. Since

(10)

ny ∈ yN , we can rewrite it as ny = yk for some k ∈ N . Using this, we get that

ab = x(ny)m = x(yk)m = (xy)(km) and therefore

(ab)N = (xy)(km)N = (xy)N

and the operation is well-defined. Next, for any a, b, c ∈ G we have that aN (bN cN ) = aN (bc)N = a(bc)N = (ab)cN = (ab)N cN = (aN bN )cN so the operation is associative. The equation

aN 1GN = (a1G)N = aN = (1Ga)N = 1GN aN shows that 1GN = N is the identity element. Finally,

aN a−1N = (aa−1)N = 1GN = (a−1a)N = a−1N aN shows that (aN )−1= a−1N . This shows that G/N is a group.

Note in particular that if G is an abelian group, then so is G/N . By assumption, ab = ba for all a, b ∈ G and from this it follows immediately that aN bN = (ab)N = (ba)N = bN aN .

2.3 Group morphisms

Definition. Let (G, ·G) and (H, ·H) be groups. A group morphism is a map φ : G −→ H such that φ(a ·Gb) = φ(a) ·H φ(b) for all a, b ∈ G.

Example 2.5.

(a) For any group G, the identity map G −→ G, a 7−→ a is a group morphism.

(b) For any subgroup H ≤ G, the inclusion map H −→ G, a 7−→ a is a group morphism.

(c) For any normal subgroup N E G, the quotient map G −→ G/N , a 7−→ aN is a group morphism.

(d) Let G be a group and a ∈ G. Using the additive group Z, we can define a function φ : Z −→ G by setting φ(n) = an. This is a group morphism, since φ(m + n) = am+n = aman= φ(m)φ(n), and we call this the exponential map with basis a.

(e) Consider the group GLn(R). We can define a map φ : GLn−→ R\{0}

by setting φ(A) = det(A). Since R\{0} is a group under multiplication and since φ(AB) = det(AB) = det(A)det(B) = φ(A)φ(B), this map is a group morphism.

(11)

Definition. A bijective group morphism is called an isomorphism. If there exists an isomorphism φ : G −→ H we say that the groups G and H are isomorphic and denote this by G ∼= H.

Example 2.6.

Let A =a −b b a



: a, b ∈ R (a, b) 6= (0, 0)

 .

The set A forms a group under matrix multiplication. By observing that

a −b b a

 c −d

d c



=ac − bd −(ad + bc) ad + bc ac − bd



we see that A is closed under matrix multiplication. Associativity follows from the fact that matrix multiplication is associative, and A clearly contains the identity matrix. For any matrix in A we have that

deta −b b a



= aa − (−b)b = a2+ b2 6= 0 which shows that every element in A has an inverse. Finally,

a −b

b a

−1

= 1

a2+ b2

 a −(−b)

−b a



= 1

a2+ b2

 a b

−b a



and this shows that the inverse is also in A.

We will now show that this group is isomorphic to the group of C \ {0}

under multiplication. Define a map φ : C \ {0} −→ A by φ(a + bi) =a −b

b a



for all a + bi ∈ C \ {0}. This map is clearly a bijection, so all that remains is to show that it is a morphism. Let a + bi, c + di ∈ C \ {0}. Then,

φ((a + bi)(c + di)) = φ((ac − bd) + (ad + bc)i)

=ac − bd −(ad + bc) ad + bc ac − bd



=a −b b a

 c −d d c



= φ(a + bi)φ(c + di) so φ is an isomorphism and A ∼= C \ {0}.

(12)

Proposition 11. Let G and H be groups and let φ : G −→ H be a group morphism.

(i) φ(1G) = 1H

(ii) φ(a−1) = (φ(a))−1 ∀a ∈ G

(iii) φ(a1· · · an) = φ(a1) · · · φ(an) ∀a1, . . . , an∈ G (iv) φ(an) = (φ(a))n ∀a ∈ G

Proof. Let a, a1, . . . , an∈ G (i) The equation

1Hφ(a) = φ(a) = φ(1Ga) = φ(1G)φ(a) implies that φ(1G) = 1H, by Proposition 2(i).

(ii) The equation

φ(a−1)φ(a) = φ(a−1a) = φ(1G) = 1H = (φ(a))−1φ(a) implies that φ(a−1) = (φ(a))−1, again by Proposition 2(i).

(iii) The base case φ(a1a2) = φ(a1)φ(a2) holds by the definition of a group morphism. Suppose φ(a1· · · ak) = φ(a1) · · · φ(ak) holds for some posi- tive integer k. Then, for k + 1 we have that

φ(a1· · · ak+1) = φ((a1· · · ak)ak+1)

= φ(a1· · · ak)φ(ak+1)

= φ(a1) · · · φ(ak)φ(ak+1) The result now follows by induction.

(iv) Apply (iii) with a1 = · · · = an= a.

Proposition 12.

(i) If φ : G −→ H and ψ : H −→ K are group morphisms, then so is ψ ◦ φ : G −→ K.

(ii) If φ : G −→ H is an isomorphism, then so is φ−1 : H −→ G.

Proof.

(i) Let a, a ∈ G. Then,

(ψ ◦ φ)(ab) = ψ(φ(ab))

= ψ(φ(a)φ(b))

= ψ(φ(a))ψ(φ(b)) = (ψ ◦ φ)(a)(ψ ◦ φ)(b) and ψ ◦ φ is a morphism.

(13)

(ii) The inverse of a bijection is a bijection, so it remains to show that φ−1 is a morphism. Let a, b ∈ H. Then there exists unique a0, b0 ∈ G such that φ(a0) = a and φ(b0) = b. Hence

φ−1(ab) = φ−1(φ(a0)φ(b0)) = φ−1(φ(a0b0)) = a0b0 = φ−1(a)φ−1(b) and φ−1 is an isomorphism.

Definition. Let φ : G −→ H be a group morphism. The kernel of φ is the subset ker(φ) = {a ∈ G : φ(a) = eH} of G. The image of φ is the subset im(φ) = {φ(a) : a ∈ G} of H.

Proposition 13. A group morphism φ : G −→ H is injective if and only if ker(φ) = {1G}

Proof.

(⇒) Suppose a ∈ ker(φ). Then φ(a) = 1H = φ(1G) and since φ is injective it follows immediately that a = 1G.

(⇐) Let a, b ∈ G and suppose φ(a) = φ(b). Then φ(a)(φ(b))−1 = φ(b)(φ(b))−1 which implies that φ(ab−1) = φ(bb−1) = φ(1G) = 1H. Because ker(φ) = {1G} we must have that ab−1 = 1G and thus a = b and φ is injective.

Proposition 14. For a group morphism φ : G −→ H, ker(φ) E G and im(φ) ≤ H.

Proof. By definition, ker(φ) ⊆ G and im(φ) ⊆ H. From Proposition 11(i) it follows that ker(φ) 6= ∅ and im(φ) 6= ∅.

Let x, y ∈ im(φ). Then there exists a, b ∈ G such that x = φ(a) and y = φ(b), so xy−1 = φ(a)φ(b)−1 = φ(ab−1). Hence xy−1 ∈ im(φ) and im(φ) ≤ H.

Let a, b ∈ ker(φ). Then φ(ab−1) = φ(a)φ(b)−1 = 1H1−1H = 1H, so ab−1 ∈ ker(φ) and ker(φ) ≤ G.

Finally, set K = ker(φ) and let a ∈ G. Suppose that x ∈ aKa−1, i.e.

x = aka−1 for some k ∈ K. Then,

φ(x) = φ(aka−1)

= φ(a)φ(k)φ(a)−1

= φ(a)1Hφ(a)−1

= φ(a)φ(a)−1

= 1H

(14)

so x ∈ K. This shows that aKa−1 ⊆ K, or equivalently that aK ⊆ Ka.

Since a ∈ G was arbitrary, we can replace it by a−1 ∈ G and get that a−1Ka ⊆ K, which is equivalent to Ka ⊆ aK. Hence aK = Ka and K = ker(φ) E G.

Using this proposition, we can now prove the following important theo- rem about group morphisms.

Theorem 15. Let G and H be groups and let φ : G −→ H be a group morphism. Then G/ker(φ) ∼= im(φ) with the isomorphism given by

φ : G/ker(φ) −→ im(φ), ker(φ) · x 7−→ φ(x).¯

Proof. We need to show that ¯φ is well defined, that it is a group morphism and that it is bijective. We set K := ker(φ).

First, suppose x, y ∈ G and Kx = Ky. Then xy−1∈ K, so we have that φ(x)φ(y)−1 = φ(xy−1) = 1H

and thus φ(x) = φ(y). Using this, we see that

φ(Kx) = φ(x) = φ(y) = ¯¯ φ(Ky)

so ¯φ is well defined. Next, for any Kx, Ky ∈ G/K we have that φ((Kx)(Ky)) = ¯¯ φ(K(xy)) = φ(xy) = φ(x)φ(y) = ¯φ(Kx) ¯φ(Ky) and so ¯φ is a morphism.

For any y ∈ im(φ) there exists x ∈ G such that y = φ(x). Hence there exists Kx ∈ G/K such that ¯φ(Kx) = φ(x) = y and ¯φ is surjective. Finally, suppose Kx ∈ ker( ¯φ). Then ¯φ(Kx) = 1H, or equivalently φ(x) = 1H. This means that x ∈ K, and therefore Kx = K. So ker( ¯φ) = {K} and ¯φ is injective.

3 Rings and modules

Having introduced groups, we can now build on this and define more complex algebraic structures. We first state the definition of a ring and then follow this by introducing modules, which join a group and a ring into a single structure. After this we cover the direct sum, an operation that can be used to combine multiple modules into one. Finally, we have a brief section on vector spaces.

Most of the material in this section is taken from [1], with some additional material on vector spaces coming from [2] and [3].

(15)

3.1 Rings and fields

Definition. A ring is a set R together with an addition + : R × R −→ R and a multiplication · : R × R −→ R satisfying the following axioms for all x, y, z ∈ R

(i) x + (y + z) = (x + y) + z (ii) 0 + x = x + 0 = x

(iii) ∃ − x ∈ R such that x + (−x) = (−x) + x = 0 (iv) x + y = y + x

(v) x(yz) = (xy)z for all x, y, z ∈ R (vi) ∃1 ∈ R such that 1x = x1 = x (vii) (x + y)z = xz + yz

(viii) x(y + z) = xy + xz

If multiplication is also commutative, we say that R is a commutative ring.

A commutative ring in which every non-zero element has a multiplicative inverse is called a field.

Note that axioms (i) to (iv) imply that every ring is also an abelian group under addition. Similarly, the axioms for a field imply that the set of non-zero elements of a field is a group under multiplication.

Example 3.1.

(a) Z, Q, R and C are all rings under the usual addition and multiplication.

In particular, Z is a commutative ring and all the others are fields.

(b) The set of all continuous functions f : R −→ R forms a ring if we define addition and multiplication pointwise, i.e. (f + g)(x) = f (x) + g(x) and (f · g)(x) = f (x) · g(x) for all x ∈ R.

(c) For any commutative ring R, the set

R[X] = {a0+ a1X + · · · + anXn : ai∈ R n ∈ N}

forms a ring under polynomial addition and multiplication, called the polynomial ring in X over R.

Definition. The characteristic of a ring R is the least positive integer n such that

r + r + · · · + r

| {z }

n times

= 0

for every r ∈ R. If no such n exists, the characteristic is 0. We denote the characteristic of R by char(R).

(16)

For any positive integer n, we can define n ∈ R by n := 1R+ 1R+ · · · + 1R

| {z }

n times

Using this, we can slightly rephrase the definition and say that the charac- teristic is the least positive integer n ∈ R such that nr = 0 for all r ∈ R, or 0 if no such n exists. Note that this is similar to the definition of order in an additive group.

With this definition, it is easy to see that if the characteristic of a ring R is 0, then R must infinite. By definition, if char(R) = 0, then for any positive integer n there exists at least one r ∈ R \ {0} having an additive order greater than n. This implies that the set {0, r, 2r, . . . , nr} must contain n + 1 distinct elements. Since this holds for arbitrarily large n, it follows that R has infinitely many elements.

The converse of this statement does not hold. There are many infinite rings which nonetheless have non-zero characteristic. For example, consider the polynomial ring Z3[X]. The characteristic of this is ring is clearly the same as for Z3, i.e. 3. However, since Xn∈ Z3[X] for all n ∈ N, and since m 6= n implies Xm6= Xn, the ring contains infinitely many polynomials.

3.2 Modules

Definition. Let R be a ring. An R-module is an abelian group (M, +) together with a map · : R × M −→ M called a scalar multiplication such that ∀r, s ∈ R ∀x, y ∈ M

(i) r · (x + y) = r · x + r · y (ii) (r + s) · x = r · x + s · x (iii) (rs) · x = r · (s · x) (iv) 1R· x = x

Remark. The above is more specifically the definition of a left R-module.

A right R-module is defined similarly, but with scalar multiplication on the right. For simplicity’s sake, we will work only with left modules and refer to these simply as modules.

Proposition 16. Let M be an R-module. For all r, s ∈ R and x, y ∈ M (i) 0Rx = 0M

(ii) r0M = 0M

(iii) (r − s)x = rx − sx (iv) r(x − y) = rx − ry

(17)

Proof. Let r, s ∈ R and x, y ∈ M .

(i) 0Rx + 0M = 0Rx = (0R+ 0R)x = 0Rx + 0Rx and this implies that 0Rx = 0M.

(ii) r0M + 0M = r0M = r(0M + 0M) = r0M + r0M and this implies that r0M = 0M.

(iii) rx − sx + sx = rx = (r − s + s)x = (r − s)x + sx and this implies that (r − s)x = rx − sx.

(iv) rx − ry + ry = rx = r(x − y + y) = r(x − y) + ry and this implies that r(x − y) = rx − ry.

Example 3.2.

(a) For any ring R, the trivial additive group {0} is an R-module under the scalar multiplication defined by r · 0 = 0 for all r ∈ R. We call this the trivial module.

(b) Any ring R is also an R-module, with scalar multiplication interpreted as the multiplicative ring operation. More generally, Rn with scalar multiplication defined by

r · (x1, . . . , xn) := (rx1, . . . , rxn) is an R-module for any n.

3.3 Submodules

Definition. Let M be an R-module. An R-submodule of M is a subgroup N ≤ M such that x ∈ N implies rx ∈ N for all r ∈ R. In particular, if N < M then we say that N is a proper submodule.

As with subgroups, it follows immediately from the definition that the submodule relation is transitive.

Example 3.3.

(a) Every module M has the trivial submodule {0}.

(b) Let R3 be an R-module. It is easy to see that the subset X = {(x1, x2, x3) ∈ R3 : x1+ x2+ x3 = 0}

is a subgroup of R3. For any r ∈ R and x ∈ R3 we have that rx1+ rx2+ rx3= r(x1+ x2+ x3) = r0 = 0 so rx ∈ X and X is a submodule of R3.

(18)

Definition. A non-trivial module M is simple if it has no submodules other than {0} and M itself.

Proposition 17. Let M be an R-module. If (Ni)i∈I is a collection of sub- modules of M , thenT

i∈INi is also a submodule of M .

Proof. Every submodule of the module M is a subgroup of the group M , so it follows from Proposition 5 that T

i∈INi is also subgroup of M . Let x ∈ T

i∈INi. Then, for all i ∈ I, we have that x ∈ Ni. Since these Ni are submodules of M we have rx ∈ Ni for all r ∈ R and i ∈ I, and thus rx ∈T

i∈INi.

Since every module is by definition an abelian group, all submodules are normal subgroups. Just as we used normal subgroups to define quotient groups, we can use submodules to define quotient modules.

Proposition 18. Let M be an R-module and let N ≤ M be a submodule.

Let M/N be the quotient group. We define a scalar multiplication

· : R × M/N −→ M/N by

r · (x + N ) = rx + N ∀ r ∈ R ∀ x ∈ N

Under this operation, M/N forms an R-module, and we call this the quotient module of M by N .

Proof. We know from before that if M is an abelian group, then so is M/N . All that remains to verify that the scalar multiplication is well-defined and that it satisfies the module axioms.

Let r ∈ R and x, y ∈ M . Suppose that x + N = y + N . Equivalently, x−y ∈ N , and since N is a submodule, we have that r(x−y) = rx−ry ∈ N . This shows that rx + N = ry + N and using this, we see that

r · (x + N ) = rx + N = ry + N = r · (y + N ) and so the scalar multiplication is well-defined.

Next, we check the module axioms. let r, s ∈ R and x+N, y +N ∈ M/N . Then, we see that

(i)

r · ((x + N ) + (y + N )) = r · ((x + y) + N )

= r · (x + y) + N

= (rx + ry) + N

= rx + N + ry + N

= r · (x + N ) + r · (y + N )

(19)

(ii)

(r + s) · (x + N ) = (r + s)x + N

= (rx + sx) + N

= rx + N + sx + N

= r · (x + N ) + s · (x + N ) (iii)

(rs) · (x + N ) = (rs)x + N

= r(sx) + N

= r · (sx + N )

= r · (s · (x + N )) (iv)

1R· (x + N ) = 1Rx + N = x + N and this shows that M/N is an R-module.

3.4 Morphisms of modules

Definition. Let M and N be R-modules. An R-module morphism is a function f : M −→ N such that ∀r ∈ R and ∀x, y ∈ M

(i) f (x + y) = f (x) + f (y) (ii) f (r · x) = r · f (x)

Rephrasing the definition, we could say that a module morphism is sim- ply a group morphism that is compatible with scalar multiplication. As such, the kernel and image of module morphism are defined in the same way as for group morphisms. As with group morphisms, a bijective R-module mor- phism is called an isomorphism. If we have an isomorphism f : M −→ N we say that the modules M and N are isomorphic and denote this by M ∼= N .

Example 3.4.

(a) For any module M , the identity map M −→ M , x 7−→ x is a mor- phism.

(b) For any submodule N of M , the inclusion map N −→ M , x 7−→ x and the quotient map N −→ M/N , x 7−→ x + N are morphisms.

(c) Consider the R-module R[X]. If we define d(f (X)) = f0(X) for all f (X) ∈ R[X], then d : R[X] −→ R[X] is a morphism.

(20)

Proposition 19.

(i) If f : M −→ L and g : L −→ N are R-module morphisms, then so is g ◦ f : M −→ N .

(ii) If f : M −→ N is an isomorphism, then so is f−1 : N −→ M . Proof.

(i) We showed in Proposition 12 that the composition of two group mor- phisms is a group morphism. It remains is to show that g ◦ f is compatible with the scalar multiplications of M and N .

Let r ∈ R and x ∈ M . Then,

(g ◦ f )(r · x) = g(f (r · x)) = g(r · f (x)) = r · g(f (x)) = r · (g ◦ f )(x) and hence g ◦ f is an R-module morphism.

(ii) The same proposition also showed that the inverse of a group isomor- phism is a group isomorphism. Again, it remains to show that f−1 is compatible with the scalar multiplication.

Let r ∈ R and y ∈ N . Since f is bijective, there exists a unique x ∈ M such that y = f (x), or equivalently f−1(y) = x. Thus,

f−1(r · y) = f−1(r · f (x)) = f−1(f (r · x)) = r · x = r · f−1(y) and f−1 is an R-module morphism.

Proposition 20. Let M and N be R-modules and let f : M −→ N be an R-module morphism. Then ker(f ) and im(f ) are submodules of M and N , respectively.

Proof. Since f is a group morphism, it follows by Proposition 14 that ker(f ) E M and im(f ) ≤ N . Suppose x ∈ ker(f ). Then f (x) = 0N

and so

f (rx) = rf (x) = r0N = 0N

Hence rx ∈ ker(f ) and ker(f ) is a submodule of M . Suppose y ∈ im(f ).

Then y = f (x) for some x ∈ M and thus ry = rf (x) = f (rx)

where rx ∈ M . Hence ry ∈ im(f ) and im(f ) is a submodule of N . The next result is the analogue of Theorem 15 for modules.

Theorem 21. Let f : M −→ N be a morphism of R-modules. Then M/ker(f ) ∼= im(f ), with the isomorphism given by

f : M/ker(f ) −→ im(f )¯ x + ker(f ) 7−→ f (x)

(21)

Proof. The proof of Theorem 15 shows that ¯f is a well-defined group iso- morphism, so all that remains is to show that it is compatible with the scalar multiplications of M/ker(f ) and N .

Let x ∈ M and r ∈ R. Then, using the fact that f is a module morphism, we get that

r · ¯f (x + N ) = r · f (x) = f (r · x) = ¯f (r · x + N ) and so ¯f is a module isomorphism.

3.5 Direct products and direct sums

Definition. Let (Mi)i∈I be a collection of R-modules. LetQ

i∈IMi be the direct product of (Mi)i∈I as groups. The direct product of (Mi)i∈I as R- modules is the R-module with Q

i∈IMi as its underlying group and with scalar multiplication defined componentwise by

r · (xi)i∈I = (r ·ixi)i∈I ∀r ∈ R

As with direct products of groups, if I = {1, 2, . . . , n} we may instead write Y

i∈I

Mi = M1× M2× · · · × Mn If I = ∅ we set Q

i∈IMi= {0}.

Given a direct product of R-modules Q

i∈IMi, we can for every j ∈ I define a projection πj by

πj :Y

i∈I

Mi−→ Mj (xi)i∈I 7−→ xj

It is easy to see that this is a morphism. Using these projections, we obtain the following property of the direct product.

Proposition 22. Let M be an R-module and let (Ni)i∈I be a collection of R-modules. For any collection (φi)i∈I of morphisms φi : M −→ Ni there exists a unique morphism φ : M −→Q

i∈INi such that πi ◦ φ = φi for all i ∈ I.

Proof. Define a map φ : M −→Q

i∈INi by φ(x) = (φi(x))i∈I for all x ∈ M . We need to show that φ is a morphism, that πi◦ φ = φi for all i ∈ I and that φ is the only morphism with this property. Let x, y ∈ M and let r ∈ R.

Then,

φ(x + y) = (φi(x + y))i∈I

= (φi(x) + φi(y))i∈I

= (φi(x))i∈I + (φi(y))i∈I

= φ(x) + φ(y)

φ(rx) = (φi(rx))i∈I

= (rφi(x))i∈I

= r(φi(x))i∈I

= rφ(x)

(22)

and this shows that φ is a morphism. For any j ∈ I we have that (πj◦ φ)(x) = πj(φ(x)) = πj((φi(x))i∈I) = φj(x)

so πj ◦ φ = φj for all i ∈ I. Finally, suppose ψ : M −→ Ni is another morphism such that πj ◦ ψ = φi for all i ∈ I. Then, for any j ∈ I we have that

φj(x) = (πj◦ ψ)(x) = πj(ψ(x)) = ψ(x)j

which implies that

ψ(x) = (ψ(x)i)i∈I = (φi(x))i∈I = φ(x) and therefore ψ = φ.

Definition. Let M be an R-module and let (Ni)i∈I be a collection of sub- modules of M . The sum of modules P

i∈INi is the submodule of M having as its underlying group

X

i∈I

Ni= (

X

i∈I

xi: xi∈ Ni for all i and xi = 0 for all but finitely many i )

Similarly to the direct product, if I = {1, 2, . . . , n} we may instead write X

i∈I

Ni = N1+ N2+ · · · Nn

To see that that the sum N := P

i∈INi is a subgroup of M , we first note that it is non-empty, sinceP

i∈I0i ∈ N . Next, consider two elements x =P

i∈Ixi and y =P

i∈Iyi in N . We then have that x − y =X

i∈I

xi−X

i∈I

yi=X

i∈I

xi− yi

Since the sums x and y are both in N , their terms must fulfil the conditions in the definition of N , i.e. that xi, yi ∈ Ni for all i and that xi = 0 and yi = 0 for all but finitely many i. Therefore, the same is true of the terms xi− yi and so x − y ∈ N . This shows that N ≤ M .

To see that N is a submodule, first note that since Ni is a submodule we have that rxi ∈ Ni for any xi ∈ Ni and any r ∈ R. For any x ∈ N

rx = rX

i∈I

xi=X

i∈I

rxi

and since the terms of x fulfil the conditions in the definition of N , so do the terms rxi. Hence rx ∈ N and N is a submodule.

Now that we have the idea of a direct product and a sum of modules, we can begin to define the direct sum.

(23)

Definition. Let (Mi)i∈I be a collection of R-modules. The external direct sum of the modules (Mi)i∈I is the R-module

M

i∈I

Mi = (

(xi)i∈I ∈Y

i∈I

Mi : xi = 0 for all but finitely many i ∈ I )

This is a submodule of the direct productQ

i∈IMi. In particular, if we have I = {1, 2, . . . , n} then this definition is equivalent to the definition of the direct product, and we may write

M

i∈I

Mi= M1⊕ M2⊕ · · · ⊕ Mn= M1× M2× · · · × Mn

Similarly, if I = ∅ we setL

i∈IMi= {0}.

Given the external direct sum L

i∈IMi we can for every j ∈ I define an injection ιj : Mj −→L

i∈IMi. We do this by taking x ∈ Mj and setting ιj(x)j = x ∈ Mj and ιj(x)i = 0 ∈ Mi for i 6= j

It is easy to see that this is a morphism. Similarly to how we used the projections of a direct product to prove Proposition 22, we can use these injections to obtain a similar property of the external direct sum. Before we begin, we need the following lemma.

Lemma 23. LetL

i∈IMi be a direct sum of R-modules. Any x ∈L

i∈IMi

can be written uniquely as a sum of the form x =P

i∈Iιi(yi), where yi∈ Mi for all i and yi= 0 for all but finitely many i.

Proof. First, x ∈ L

i∈IMi implies that x = (xi)i∈I with xi ∈ Mi for all i and xi = 0 for all but finitely many i. It now follows immediately from the definition of ιi that x =P

i∈Iιi(xi), and thus we have a sum of the desired form. It remains to show that the sum is unique.

Suppose we have another sum with the same property, i.e.that we have (yi)i∈I such that x = P

i∈Iιi(yi), where yi ∈ Mi for all i and yi = 0 for all but finitely many i. Then y = (yi)i∈I ∈L

i∈IMi and by the same logic as above, y = P

i∈Iιi(yi) = x. Hence yi = xi for all i and the sums are identical.

We are now ready to prove the following general property of the external direct sum.

Proposition 24. Let M be an R-module and let (Ni)i∈I be a collection of R-modules. For any collection (φi)i∈I of morphisms φi : Ni −→ M there exists a unique morphism φ : L

i∈INi −→ M such that φ ◦ ιi = φi for all i ∈ I.

(24)

Proof. The sum P

i∈Iφi(xi) is defined for all x = (xi)i∈I ∈L

i∈INi, hence we can define a map

φ :M

i∈I

Ni −→ M x 7−→X

i∈I

φi(xi)

We need to show that this is a morphism, that φ ◦ ιi = φi for all i ∈ I and that φ is the unique morphism with this property.

To see that φ is a morphism, let x = (xi)i∈I and y = (yi)i∈I be in L

i∈INi and let r ∈ R. Then,

φ(x + y) = φ((xi)i∈I + (yi)i∈I)

= φ((xi+ yi)i∈I)

=X

i∈I

φi(xi+ yi)

=X

i∈I

φi(xi) + φi(yi)

=X

i∈I

φi(xi) +X

i∈I

φi(yi)

= φ(x) + φ(y)

φ(rx) = φ((rxi)i∈I)

=X

i∈I

φi(rxi)

= rX

i∈I

φi(xi)

= rφ(x)

and this shows that φ is a morphism. For any j ∈ I we have that (φ ◦ ιj)(x) = φ(ιj(x)) = φ(xj) = xj = φj(x) so φ ◦ ιi = φi for all i ∈ I. Lastly, suppose ψ :L

i∈INi −→ M is another morphism such that ψ ◦ ιi = φi for all i ∈ I. By using the lemma we just

(25)

proved, we see that

ψ(x) = ψ((xi)i∈I)

= ψ X

i∈I

ιi(xi)

!

=X

i∈I

ψ(ιi(xi))

=X

i∈I

(ψ ◦ ιi)(xi)

=X

i∈I

φi(xi)

= φ(x) and thus ψ = φ.

Proposition 25. Let (Mi)i∈I be a collection of R-modules. If M is also an R-module, then the following are equivalent:

(i) M ∼=L

i∈IMi

(ii) M contains submodules (Ni)i∈I such that Ni ∼= Mi for all i and such that every x ∈ M can be written uniquely as a sum of the form x = P

i∈Iyi, with yi ∈ Ni for all i and yi = 0 for all but finitely many i.

(iii) M contains submodules (Ni)i∈I such that Ni ∼= Mi for all i, M = P

i∈INi and Nj∩ (P

i6=jNi) = {0} for all j ∈ I.

Proof.

(i) ⇒ (ii) Using the previously defined injective morphisms ιj : Mj −→L

i∈IMi we see that for any i, the image im(ιi) = ιi(Mi) is by definition a submodule of L

i∈IMi such that ιi(Mi) ∼= Mi. By slightly rephrasing the statements of Lemma 23 we find that any x ∈ L

i∈IMi can be written uniquely as a sum of the form x =P

i∈Iιi(xi), where we have ιi(xi) ∈ ιi(Mi) for all i and ιi(xi) = 0 for all but finitely many i.

(ii) ⇒ (i) Using the same idea as in the proof of Proposition 24, the collection (ιi)i∈I of injections ιi : Ni −→ M can be used to define a morphism ι :L

i∈INi −→ M . We define this morphism by setting ι((xi)i∈I) =X

i∈I

ιi(xi) =X

i∈I

xi

for all (xi)i∈I ∈L

i∈INi. By assumption, any y ∈ M can be written uniquely as a sum y =P

i∈Iyi, where yi∈ Nifor all i and yi= 0 for all but finitely many i. Hence we have an element (yi)i∈I ∈L

i∈INiwhich maps onto y and so ι is surjective. Moreover, since the sum is unique it

(26)

follows that the element which maps onto y is also unique, and hence ι is also injective and therefore an isomorphism. Using the assumption that Ni ∼= Mi for all i, we get thatL

i∈IMi∼=L

i∈INi ∼= M .

(ii) ⇒ (iii) By (ii), any x ∈ M can be written (uniquely) as a sum of elements from (Ni)i∈I and therefore M ⊆P

i∈INi. Conversely, since every Ni

is a submodule of M it follows immediately that P

i∈INi ⊆ M and therefore M =P

i∈INi. If x ∈ Nj ∩ (P

i6=jNi) then x can be written as a sum x = P

i∈Iyi

where yi ∈ Ni for all i, with yj = x and yi = 0 for all i 6= j. We can also write x as a sum x =P

i∈Izi where zi ∈ Ni for all i, zi = 0 for all but finitely many i and in particular zj = 0. Since there is a unique way of writing x as a sum of this form, the two sums must be identical and thus x = yj = zj = 0.

(iii) ⇒ (ii) Since M =P

i∈INi, it follows immediately any x ∈ M can be written as a sum of the desired form. All that remains is to show that the sum is unique. Suppose x =P

i∈Iyi and x =P

i∈Izi are two sums of this form. Then, for any j ∈ I we can write

zj− yj =X

i6=j

(yi− zi) ∈ Nj∩ (X

i6=j

Ni)

implying that yj = zj. Since this holds for every j ∈ I, the sums are identical.

Definition. Let M be an R-module and let (Ni)i∈I be a collection of sub- modules of M . M is the internal direct sum of the submodules (Ni)i∈I if every x ∈ M can be written uniquely as a sum of the form x = P

i∈Iyi, with yi ∈ Ni for all i and yi= 0 for all but finitely many i. We denote this internal direct sum by M =L

i∈INi.

The use of the same notation for both the external and internal direct sum is motivated by the fact that the two definitions are equivalent up to isomorphism. This is easy to see using Proposition 25:

If M is an external direct sum of a collection of modules (Mi)i∈I, then condition (i) holds, so we can consider the equivalent condition (ii). It states that there exists a collection of submodules (Ni)i∈I such that M is the internal direct sum L

i∈INi. Moreover, Mi ∼= Ni for all i, so M is isomorphic to the internal direct sumL

i∈IMi.

Conversely, suppose M is an internal direct sum of submodules (Ni)i∈I. Then, if we let (Mi)i∈I be a collection of R-modules such that Mi ∼= Ni

for all i, condition (ii) holds. The equivalent condition (i) then states that M is isomorphic to the external direct sumL

i∈IMi and hence also to the external direct sumL

i∈INi.

(27)

Because of this fact, we usually only speak of direct sums, without spec- ifying which type is being used.

Definition. Let M be a module. A submodule N of M is a direct summand of M if there exists another submodule L of M such that M = N ⊕ L.

Proposition 26. Let M be an R-module. A submodule N of M is a direct summand of M if and only if there exists a morphism π : M −→ M with im(π) = N and π2= π.

Proof.

(⇒) If N is a direct summand of M then there exists another submodule L of M such that M = N ⊕ L. By the definition of the (internal) direct sum, this means that every x ∈ M can be written uniquely on the form x = xn+ xl, with xn ∈ N and xl ∈ L. Using this, we can easily define a map π : M −→ M by setting π(x) = xn. Let x, y ∈ M and r ∈ R. Then,

π(x + y) = π(xn+ xl+ yn+ yl)

= π(xn+ yn+ xl+ yl)

= xn+ yn

= π(x) + π(y)

π(rx) = π(r(xn+ xl))

= π(rxn+ rxl))

= rxn

= rπ(x) and this shows that π is an R-module morphism.

By definition, im(π) ⊆ N . If y ∈ N , then we also have that y ∈ M and π(y) = y, so N ⊆ im(π). This shows that im(π) = N .

Finally, for any x = xn+ xl∈ M we have that

π2(x) = π(π(x)) = π(xn) = xn= π(xn) and this shows that π2= π.

(⇐) Suppose π : M −→ M is an R-module morphism such that π2 = π.

By Proposition 20 both im(π) and ker(π) are submodules of M . We will show that im(π) is a direct summand of M by using condition (iii) in Proposition 25 to show that M = ker(π) ⊕ im(π).

First, let x ∈ M . Then π2(x) = π(x), or equivalently π(x − π(x)) = 0.

This means that x − π(x) = y for some y ∈ ker(π) and so every x ∈ M can be written as a sum x = y + π(x), with y ∈ ker(π) and π(x) ∈ im(π). This shows that M = ker(π) + im(π).

Next, suppose that y ∈ ker(π) ∩ im(π). Then y = π(x) for some x ∈ M , or equivalently π(y) = π2(x). Since y ∈ ker(π) also, it now follows that

0 = π(y) = π2(x) = π(x) = y

So ker(π) ∩ im(π) = {0}, and therefore M = ker(π) ⊕ im(π).

(28)

3.6 Vector spaces

Definition. Let M be an R-module.

• A linear combination is a finite sum of the form r1x1+ · · · + rnxn where ri ∈ R and xi∈ M .

• A subset {x1, . . . , xn} ⊆ M is R-linearly independent if r1x1+ · · · + rnxn= 0

implies r1= · · · = rn= 0. Otherwise it is R-linearly dependent.

• The span of the set {x1, . . . , xn} is the set of all linear combinations of x1, . . . , xn. For the empty set, the span consists only of the empty sum, which is equal to {0}.

• A basis of M is a subset B ⊆ M which is linearly independent and spans M .

Note that if we have a basis B = {x1, . . . , xn} of an R-module M , then every x ∈ M can be written as a unique linear combination of elements in B. This follows immediately from the definition of a basis. For if x could be written in two different ways, say x = r1x1 + · · · + rnxn and x = s1x1+ · · · + snxn, where ri, si ∈ R and ri 6= si for at least one i, then we would have that

(r1− s1)x1+ · · · + (rn− sn)xn= x − x = 0

with at least one (ri − si) 6= 0, contradicting the fact that B is linearly independent.

It should be noted that not every module has a basis. For example, consider the group of Z5 under addition. This is a Z-module under the scalar multiplication defined by

n · x = x + x + · · · + x

| {z }

n times

∈ Z5

for all n ∈ Z and x ∈ Z5. Every element in the group is of order 5, so for any non-empty subset B ⊆ Z5 there is an x ∈ B such that 5 · x = 0. Since 5 6= 0 in Z, this means that B is linearly dependent, and hence not a basis.

The empty subset can also not be a basis, since its span is the set {0}. We therefore conclude that this module has no basis.

Modules which do have a basis are called free modules. Free modules will not be dealt with in general in this text, but we will cover the following special case.

(29)

Definition. A K-module V where K is a field is called a K-vector space.

The elements of K are called scalars and the elements of V are called vectors.

Submodules of vector spaces are called subspaces, and morphisms of K- vector spaces are called K-linear maps. Every vector space has a basis and the dimension of V is the cardinality of its basis (which may be infinite).

Theorem 27. Let V be a vector space.

(i) Every finite subset X ⊆ V which spans V can be reduced to a basis of V .

(ii) If dim(V ) is finite, every finite linearly independent subset X ⊆ V can be extended to a basis of V .

Proof.

(i) Suppose X = {x1, . . . , xn} spans V . We check each element in order from x1 to xn. At each step, if xi ∈ span{x1, . . . , xi−1} we delete it.

Note that this includes the case x1 ∈ span(∅) = {0}. This process will not change the span of X, since we only delete vectors that are already in the span of previous vectors, but will reduce X to the set {xi1, . . . , xim}, where m ≤ n. Suppose this set is linearly dependent.

We then have

m

X

j=1

λijxij = 0

for some λij ∈ K not all equal to 0. This implies that

xim = −1 λim

m−1

X

j=1

λijxij

meaning that xim is in the span of the previous vectors, which is not possible due to the way we reduced the set X. Hence X is now linearly independent and therefore a basis of V .

(ii) Suppose the set X = {x1, . . . , xn} is linearly independent. Let B = {b1, . . . , bm} be a basis of B. Then, X ∪ B = {x1, . . . , xn, b1, . . . , bm} spans V , meaning that we can apply the process in (i) in order to reduce it to a basis of V . Since X is linearly independent, this will not delete any xi, meaning that the resulting basis will contain all of X and possibly some elements of B. We have thus extended X to a basis of V .

References

Related documents

Re-examination of the actual 2 ♀♀ (ZML) revealed that they are Andrena labialis (det.. Andrena jacobi Perkins: Paxton &amp; al. -Species synonymy- Schwarz &amp; al. scotica while

Industrial Emissions Directive, supplemented by horizontal legislation (e.g., Framework Directives on Waste and Water, Emissions Trading System, etc) and guidance on operating

Stöden omfattar statliga lån och kreditgarantier; anstånd med skatter och avgifter; tillfälligt sänkta arbetsgivaravgifter under pandemins första fas; ökat statligt ansvar

46 Konkreta exempel skulle kunna vara främjandeinsatser för affärsänglar/affärsängelnätverk, skapa arenor där aktörer från utbuds- och efterfrågesidan kan mötas eller

För att uppskatta den totala effekten av reformerna måste dock hänsyn tas till såväl samt- liga priseffekter som sammansättningseffekter, till följd av ökad försäljningsandel

The increasing availability of data and attention to services has increased the understanding of the contribution of services to innovation and productivity in

Syftet eller förväntan med denna rapport är inte heller att kunna ”mäta” effekter kvantita- tivt, utan att med huvudsakligt fokus på output och resultat i eller från

Generella styrmedel kan ha varit mindre verksamma än man har trott De generella styrmedlen, till skillnad från de specifika styrmedlen, har kommit att användas i större