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.

is a subspace.

3. For any n ≥ 0, the set of polynomials f(X) ∈ R[X] of degree at most n is a subspace

of R[X].

4. The set of upper triangular n × n matrices is a subspace of the space of n × n matrices.

Proposition 2.5. Given any vector space E, if S is any nonempty subset of E, then the

smallest subspace S (or Span(S)) of E containing S is the set of all (finite) linear combi-

nations of elements from S.

Proof. We prove that the set Span(S) of all linear combinations of elements of S is a subspace

of E, leaving as an exercise the verification that every subspace containing S also contains

Span(S).

First, Span(S) is nonempty since it contains S (which is nonempty). If u =

λ

i∈I

iui

and v =

µ

j∈J

j vj are any two linear combinations in Span(S), for any two scalars λ, µ ∈ R,

λu + µv = λ

λiui + µ

µjvj

i∈I

j∈J

=

λλiui +

µµjvj

i∈I

j∈J

=

λλiui +

(λλi + µµi)ui +

µµjvj,

i∈I−J

i∈I∩J

j∈J−I

which is a linear combination with index set I ∪ J, and thus λu + µv ∈ Span(S), which

proves that Span(S) is a subspace.

2.4. BASES OF A VECTOR SPACE

29

One might wonder what happens if we add extra conditions to the coefficients involved

in forming linear combinations. Here are three natural restrictions which turn out to be

important (as usual, we assume that our index sets are finite):

(1) Consider combinations

λ

i∈I

iui for which

λi = 1.

i∈I

These are called affine combinations. One should realize that every linear combination

λ

i∈I

iui can be viewed as an affine combination. For example, if k is an index not

in I, if we let J = I ∪ {k}, uk = 0, and λk = 1 −

λ

λ

i∈I

i, then

j∈J

j uj is an affine

combination and

λiui =

λjuj.

i∈I

j∈J

However, we get new spaces. For example, in

3

R , the set of all affine combinations of

the three vectors e1 = (1, 0, 0), e2 = (0, 1, 0), and e3 = (0, 0, 1), is the plane passing

through these three points. Since it does not contain 0 = (0, 0, 0), it is not a linear

subspace.

(2) Consider combinations

λ

i∈I

iui for which

λi ≥ 0, for all i ∈ I.

These are called positive (or conic) combinations It turns out that positive combina-

tions of families of vectors are cones. They show naturally in convex optimization.

(3) Consider combinations

λ

i∈I

iui for which we require (1) and (2), that is

λi = 1,

and λi ≥ 0 for all i ∈ I.

i∈I

These are called convex combinations. Given any finite family of vectors, the set of all

convex combinations of these vectors is a convex polyhedron. Convex polyhedra play a

very important role in convex optimization.

2.4

Bases of a Vector Space

Given a vector space E, given a family (vi)i∈I, the subset V of E consisting of the null vector 0

and of all linear combinations of (vi)i∈I is easily seen to be a subspace of E. Subspaces having

such a “generating family” play an important role, and motivate the following definition.

30

CHAPTER 2. VECTOR SPACES, BASES, LINEAR MAPS

Definition 2.12. Given a vector space E and a subspace V of E, a family (vi)i∈I of vectors

vi ∈ V spans V or generates V if for every v ∈ V , there is some family (λi)i∈I of scalars in

K such that

v =

λivi.

i∈I

We also say that the elements of (vi)i∈I are generators of V and that V is spanned by (vi)i∈I,

or generated by (vi)i∈I. If a subspace V of E is generated by a finite family (vi)i∈I, we say

that V is finitely generated . A family (ui)i∈I that spans V and is linearly independent is

called a basis of V .

Example 2.9.

1. In

3

R , the vectors (1, 0, 0), (0, 1, 0), and (0, 0, 1) form a basis.

2. The vectors (1, 1, 1, 1), (1, 1, −1, −1), (1, −1, 0, 0), (0, 0, 1, −1) form a basis of 4

R known

as the Haar basis. This basis and its generalization to dimension 2n are crucial in

wavelet theory.

3. In the subspace of polynomials in R[X] of degree at most n, the polynomials 1, X, X2,

. . . , Xn form a basis.

n

4. The Bernstein polynomials

(1 − X)kXn−k for k = 0, . . . , n, also form a basis of

k

that space. These polynomials play a major role in the theory of spline curves.

It is a standard result of linear algebra that every vector space E has a basis, and that

for any two bases (ui)i∈I and (vj)j∈J, I and J have the same cardinality. In particular, if E

has a finite basis of n elements, every basis of E has n elements, and the integer n is called

the dimension of the vector space E. We begin with a crucial lemma.

Lemma 2.6. Given a linearly independent family (ui)i∈I of elements of a vector space E, if

v ∈ E is not a linear combination of (ui)i∈I, then the family (ui)i∈I ∪k (v) obtained by adding

v to the family (ui)i∈I is linearly independent (where k /

∈ I).

Proof. Assume that µv +

λ

i∈I

iui = 0, for any family (λi)i∈I of scalars in K. If µ = 0, then

µ has an inverse (because K is a field), and thus we have v = −

(µ−1λ

i∈I

i)ui, showing

that v is a linear combination of (ui)i∈I and contradicting the hypothesis. Thus, µ = 0. But

then, we have

λ

i∈I

iui = 0, and since the family (ui)i∈I is linearly independent, we have

λi = 0 for all i ∈ I.

The next theorem holds in general, but the proof is more sophisticated for vector spaces

that do not have a finite set of generators. Thus, in this chapter, we only prove the theorem

for finitely generated vector spaces.

2.4. BASES OF A VECTOR SPACE

31

Theorem 2.7. Given any finite family S = (ui)i∈I generating a vector space E and any

linearly independent subfamily L = (uj)j∈J of S (where J ⊆ I), there is a basis B of E such

that L ⊆ B ⊆ S.

Proof. Consider the set of linearly independent families B such that L ⊆ B ⊆ S. Since this

set is nonempty and finite, it has some maximal element, say B = (uh)h∈H. We claim that

B generates E. Indeed, if B does not generate E, then there is some up ∈ S that is not a

linear combination of vectors in B (since S generates E), with p /

∈ H. Then, by Lemma

2.6, the family B = (uh)h∈H∪{p} is linearly independent, and since L ⊆ B ⊂ B ⊆ S, this

contradicts the maximality of B. Thus, B is a basis of E such that L ⊆ B ⊆ S.

Remark: Theorem 2.7 also holds for vector spaces that are not finitely generated. In this

case, the problem is to guarantee the existence of a maximal linearly independent family B

such that L ⊆ B ⊆ S. The existence of such a maximal family can be shown using Zorn’s

lemma, see Appendix 31 and the references given there.

The following proposition giving useful properties characterizing a basis is an immediate

consequence of Theorem 2.7.

Proposition 2.8. Given a vector space E, for any family B = (vi)i∈I of vectors of E, the

following properties are equivalent:

(1) B is a basis of E.

(2) B is a maximal linearly independent family of E.

(3) B is a minimal generating family of E.

The following replacement lemma due to Steinitz shows the relationship between finite

linearly independent families and finite families of generators of a vector space.

Proposition 2.9. (Replacement lemma) Given a vector space E, let (ui)i∈I be any finite

linearly independent family in E, where |I| = m, and let (vj)j∈J be any finite family such

that every ui is a linear combination of (vj)j∈J, where |J| = n. Then, there exists a set L and

an injection ρ : L → J such that L ∩ I = ∅, |L| = n − m, and the families (ui)i∈I ∪ (vρ(l))l∈L

and (vj)j∈J generate the same subspace of E. In particular, m ≤ n.

Proof. We proceed by induction on |I| = m. When m = 0, the family (ui)i∈I is empty, and

the proposition holds trivially with L = J (ρ is the identity). Assume |I| = m + 1. Consider

the linearly independent family (ui)i∈(I−{p}), where p is any member of I. By the induction

hypothesis, there exists a set L and an injection ρ : L → J such that L ∩ (I − {p}) = ∅,

|L| = n − m, and the families (ui)i∈(I−{p}) ∪ (vρ(l))l∈L and (vj)j∈J generate the same subspace

of E. If p ∈ L, we can replace L by (L − {p}) ∪ {p } where p does not belong to I ∪ L, and

replace ρ by the injection ρ which agrees with ρ on L − {p} and such that ρ (p ) = ρ(p).

Thus, we can always assume that L ∩ I = ∅. Since up is a linear combination of (vj)j∈J

32

CHAPTER 2. VECTOR SPACES, BASES, LINEAR MAPS

and the families (ui)i∈(I−{p}) ∪ (vρ(l))l∈L and (vj)j∈J generate the same subspace of E, up is

a linear combination of (ui)i∈(I−{p}) ∪ (vρ(l))l∈L. Let

up =

λiui +

λlvρ(l).

(1)

i∈(I−{p})

l∈L

If λl = 0 for all l ∈ L, we have

λiui − up = 0,

i∈(I−{p})

contradicting the fact that (ui)i∈I is linearly independent. Thus, λl = 0 for some l ∈ L, say

l = q. Since λq = 0, we have

vρ(q) =

(−λ−1

q λi)ui + λ−1

q up +

(−λ−1

q λl)vρ(l).

(2)

i∈(I−{p})

l∈(L−{q})

We claim that the families (ui)i∈(I−{p}) ∪ (vρ(l))l∈L and (ui)i∈I ∪ (vρ(l))l∈(L−{q}) generate the

same subset of E. Indeed, the second family is obtained from the first by replacing vρ(q) by up,

and vice-versa, and up is a linear combination of (ui)i∈(I−{p}) ∪ (vρ(l))l∈L, by (1), and vρ(q) is a

linear combination of (ui)i∈I ∪(vρ(l))l∈(L−{q}), by (2). Thus, the families (ui)i∈I ∪(vρ(l))l∈(L−{q})

and (vj)j∈J generate the same subspace of E, and the proposition holds for L − {q} and the

restriction of the injection ρ : L → J to L − {q}, since L ∩ I = ∅ and |L| = n − m imply that

(L − {q}) ∩ I = ∅ and |L − {q}| = n − (m + 1).

The idea is that m of the vectors vj can be replaced by the linearly independent ui’s in

such a way that the same subspace is still generated. The purpose of the function ρ : L → J

is to pick n − m elements j1, . . . , jn−m of J and to relabel them l1, . . . , ln−m in such a way

that these new indices do not clash with the indices in I; this way, the vectors vj , . . . , v

1

jn−m

who “survive” (i.e. are not replaced) are relabeled vl , . . . , v

, and the other m vectors v

1

ln−m

j

with j ∈ J − {j1, . . . , jn−m} are replaced by the ui. The index set of this new family is I ∪ L.

Actually, one can prove that Proposition 2.9 implies Theorem 2.7 when the vector space

is finitely generated. Putting Theorem 2.7 and Proposition 2.9 together, we obtain the

following fundamental theorem.

Theorem 2.10. Let E be a finitely generated vector space. Any family (ui)i∈I generating E

contains a subfamily (uj)j∈J which is a basis of E. Furthermore, for every two bases (ui)i∈I

and (vj)j∈J of E, we have |I| = |J| = n for some fixed integer n ≥ 0.

Proof. The first part follows immediately by applying Theorem 2.7 with L = ∅ and S =

(ui)i∈I. Assume that (ui)i∈I and (vj)j∈J are bases of E. Since (ui)i∈I is linearly independent

and (vj)j∈J spans E, proposition 2.9 implies that |I| ≤ |J|. A symmetric argument yields

|J| ≤ |I|.

2.4. BASES OF A VECTOR SPACE

33

Remark: Theorem 2.10 also holds for vector spaces that are not finitely generated. This

can be shown as follows. Let (ui)i∈I be a basis of E, let (vj)j∈J be a generating family of E,

and assume that I is infinite. For every j ∈ J, let Lj ⊆ I be the finite set

Lj = {i ∈ I | vj =

λiui, λi = 0}.

i∈I

Let L =

L

j∈J

j . By definition L ⊆ I , and since (ui)i∈I is a basis of E, we must have I = L,

since otherwise (ui)i∈L would be another basis of E, and this would contradict the fact that

(ui)i∈I is linearly independent. Furthermore, J must be infinite, since otherwise, because

the Lj are finite, I would be finite. But then, since I =

L

j∈J

j with J infinite and the Lj

finite, by a standard result of set theory, |I| ≤ |J|. If (vj)j∈J is also a basis, by a symmetric

argument, we obtain |J| ≤ |I|, and thus, |I| = |J| for any two bases (ui)i∈I and (vj)j∈J of E.

When E is not finitely generated, we say that E is of infinite dimension. The dimension

of a vector space E is the common cardinality of all of its bases and is denoted by dim(E).

Clearly, if the field K itself is viewed as a vector space, then every family (a) where a ∈ K

and a = 0 is a basis. Thus dim(K) = 1. Note that dim({0}) = 0.

If E is a vector space, for any subspace U of E, if dim(U ) = 1, then U is called a line; if

dim(U ) = 2, then U is called a plane. If dim(U ) = k, then U is sometimes called a k-plane.

Let (ui)i∈I be a basis of a vector space E. For any vector v ∈ E, since the family (ui)i∈I

generates E, there is a family (λi)i∈I of scalars in K, such that

v =

λiui.

i∈I

A very important fact is that the family (λi)i∈I is unique.

Proposition 2.11. Given a vector space E, let (ui)i∈I be a family of vectors in E. Let v ∈ E,

and assume that v =

λ

λ

i∈I

iui. Then, the family (λi)i∈I of scalars such that v =

i∈I

iui

is unique iff (ui)i∈I is linearly independent.

Proof. First, assume that (ui)i∈I is linearly independent. If (µi)i∈I is another family of scalars

in K such that v =

µ

i∈I

iui, then we have

(λi − µi)ui = 0,

i∈I

and since (ui)i∈I is linearly independent, we must have λi−µi = 0 for all i ∈ I, that is, λi = µi

for all i ∈ I. The converse is shown by contradiction. If (ui)i∈I was linearly dependent, there

would be a family (µi)i∈I of scalars not all null such that

µiui = 0

i∈I

34

CHAPTER 2. VECTOR SPACES, BASES, LINEAR MAPS

and µj = 0 for some j ∈ I. But then,

v =

λiui + 0 =

λiui +

µiui =

(λi + µi)ui,

i∈I

i∈I

i∈I

i∈I

with λj = λj +µj since µj = 0, contradicting the assumption that (λi)i∈I is the unique family

such that v =

λ

i∈I

iui.

If (ui)i∈I is a basis of a vector space E, for any vector v ∈ E, if (xi)i∈I is the unique

family of scalars in K such that

v =

xiui,

i∈I

each xi is called the component (or coordinate) of index i of v with respect to the basis (ui)i∈I.

Given a field K and any (nonempty) set I, we can form a vector space K(I) which, in

some sense, is the standard vector space of dimension |I|.

Definition 2.13. Given a field K and any (nonempty) set I, let K(I) be the subset of the

cartesian product KI consisting of all families (λi)i∈I with finite support of scalars in K.4

We define addition and multiplication by a scalar as follows:

(λi)i∈I + (µi)i∈I = (λi + µi)i∈I,

and

λ · (µi)i∈I = (λµi)i∈I.

It is immediately verified that addition and multiplication by a scalar are well defined.

Thus, K(I) is a vector space. Furthermore, because families with finite support are consid-

ered, the family (ei)i∈I of vectors ei, defined such that (ei)j = 0 if j = i and (ei)i = 1, is

clearly a basis of the vector space K(I). When I = {1, . . . , n}, we denote K(I) by Kn. The

function ι : I → K(I), such that ι(i) = ei for every i ∈ I, is clearly an injection.

When I is a finite set, K(I) = KI, but this is false when I is infinite. In fact, dim(K(I)) =

|I|, but dim(KI) is strictly greater when I is infinite.

Many interesting mathematical structures are vector spaces. A very important example

is the set of linear maps between two vector spaces to be defined in the next section. Here

is an example that will prepare us for the vector space of linear maps.

Example 2.10. Let X be any nonempty set and let E be a vector space. The set of all

functions f : X → E can be made into a vector space as follows: Given any two functions

f : X → E and g : X → E, let (f + g): X → E be defined such that

(f + g)(x) = f (x) + g(x)

4Where KI denotes the set of all functions from I to K.

2.5. LINEAR MAPS

35

for all x ∈ X, and for every λ ∈ K, let λf : X → E be defined such that

(λf )(x) = λf (x)

for all x ∈ X. The axioms of a vector space are easily verified. Now, let E = K, and let I

be the set of all nonempty subsets of X. For every S ∈ I, let fS : X → E be the function

such that fS(x) = 1 iff x ∈ S, and fS(x) = 0 iff x /

∈ S. We leave as an exercise to show that

(fS)S∈I is linearly independent.

2.5

Linear Maps

A function between two vector spaces that preserves the vector space structure is called

a homomorphism of vector spaces, or linear map. Linear maps formalize the concept of

linearity of a function. In the rest of this section, we assume that all vector spaces are over

a given field K (say R).

Definition 2.14. Given two vector spaces E and F , a linear map between E and F is a

function f : E → F satisfying the following two conditions:

f (x + y) = f (x) + f (y)

for all x, y ∈ E;

f (λx) = λf (x)

for all λ ∈ K, x ∈ E.

Setting x = y = 0 in the first identity, we get f (0) = 0. The basic property of linear

maps is that they transform linear combinations into linear combinations. Given a family

(ui)i∈I of vectors in E, given any family (λi)i∈I of scalars in K, we have

f (

λiui) =

λif (ui).

i∈I

i∈I

The above identity is shown by induction on the size of the support of the family (λiui)i∈I,

using the properties of Definition 2.14.

Example 2.11.

1. The map f :

2

2

R → R defined such that

x

= x − y

y

= x + y

is a linear map. The reader should check that it is the composition of a rotation by

π/4 with a magnification of ratio

2.

2. For any vector space E, the identity map id : E → E given by

id(u) = u for all u ∈ E

is a linear map. When we want to be more precise, we write idE instead of id.

36

CHAPTER 2. VECTOR SPACES, BASES, LINEAR MAPS

3. The map D : R[X] → R[X] defined such that

D(f (X)) = f (X),

where f (X) is the derivative of the polynomial f (X), is a linear map.

4. The map Φ : C([a, b]) → R given by

b

Φ(f ) =

f (t)dt,

a

where C([a, b]) is the set of continuous functions defined on the interval [a, b], is a linear

map.

5. The function −, − : C([a, b]) × C([a, b]) → R given by

b

f, g =

f (t)g(t)dt,

a

is linear in each of the variable f , g. It also satisfies the properties f, g = g, f and

f, f = 0 iff f = 0. It is an example of an inner product.

Definition 2.15. Given a linear map f : E → F , we define its image (or range) Im f = f(E),

as the set

Im f = {y ∈ F | (∃x ∈ E)(y = f(x))},

and its Kernel (or nullspace) Ker f = f −1(0), as the set

Ker f = {x ∈ E | f(x) = 0}.

Proposition 2.12. Given a linear map f : E → F , the set Im f is a subspace of F and the

set Ker f is a subspace of E. The linear map f : E → F is injective iff Ker f = 0 (where 0

is the trivial subspace {0}).

Proof. Given any x, y ∈ Im f, there are some u, v ∈ E such that x = f(u) and y = f(v),

and for all λ, µ ∈ K, we have

f (λu + µv) = λf (u) + µf (v) = λx + µy,

and thus, λx + µy ∈ Im f, showing that Im f is a subspace of F .

Given any x, y ∈ Ker f, we have f(x) = 0 and f(y) = 0, and thus,

f (λx + µy) = λf (x) + µf (y) = 0,

that is, λx + µy ∈ Ker f, showing that Ker f is a subspace of E.

First, assume that Ker f = 0. We need to prove that f (x) = f (y) implies that x = y.

However, if f (x) = f (y), then f (x) − f(y) = 0, and by linearity of f we get f(x − y) = 0.

Because Ker f = 0, we must have x − y = 0, that is x = y, so f is injective. Conversely,

assume that f is injective. If x ∈ Ker f, that is f(x) = 0, since f(0) = 0 we have f(x) =

f (0), and by injectivity, x = 0, which proves that Ker f = 0. Therefore, f is injective iff

Ker f = 0.

2.5. LINEAR MAPS

37

Since by Proposition 2.12, the image Im f of a linear map f is a subspace of F , we can

define the rank rk(f ) of f as the dimension of Im f .

A fundamental property of bases in a vector space is that they allow the definition of

linear maps as unique homomorphic extensions, as shown in the following proposition.

Proposition 2.13. Given any two vector spaces E and F , given any basis (ui)i∈I of E,

given any other family of vectors (vi)i∈I in F , there is a unique linear map f : E → F such

that f (ui) = vi for all i ∈ I. Furthermore, f is injective iff (vi)i∈I is linearly independent,

and f is surjective iff (vi)i∈I generates F .

Proof. If such a linear map f : E → F exists, since (ui)i∈I is a basis of E, every vector x ∈ E

can written uniquely as a linear combination

x =

xiui,

i∈I

and by linearity, we must have

f (x) =

xif (ui) =

xivi.

i∈I

i∈I

Define the function f : E → F , by letting

f (x) =

xivi

i∈I

for every x =

x

i∈I

iui.

It is easy to verify that f is indeed linear, it is unique by the

previous reasoning, and obviously, f (ui) = vi.

Now, assume that f is injective. Let (λi)i∈I be any family of scalars, and assume that

λivi = 0.

i∈I

Since vi = f (ui) for every i ∈ I, we have

f (

λiui) =

λif (ui) =

λivi = 0.

i∈I

i∈I

i∈I

Since f is injective iff Ker f = 0, we have

λiui = 0,

i∈I

and since (ui)i∈I is a basis, we have λi = 0 for all i ∈ I, which shows that (vi)i∈I is linearly

independent. Conversely, assume that (vi)i∈I is linearly independent. Since (ui)i∈I is a basis

of E, every vector x ∈ E is a linear combination x =

λ

i∈I

iui of (ui)i∈I . If

f (x) = f (

λiui) = 0,

i∈I

38

CHAPTER 2. VECTOR SPACES, BASES, LINEAR MAPS

then

λivi =

λif (ui) = f (

λiui) = 0,

i∈I

i∈I

i∈I

and λi = 0 for all i ∈ I