Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Linear Algebra
Linear Algebra
Linear Algebra
Ebook667 pages9 hours

Linear Algebra

Rating: 1 out of 5 stars

1/5

()

Read preview

About this ebook

A thorough first course in linear algebra, this two-part treatment begins with the basic theory of vector spaces and linear maps, including dimension, determinants, eigenvalues, and eigenvectors. The second section addresses more advanced topics such as the study of canonical forms for matrices.
The treatment can be tailored to satisfy the requirements of both introductory and advanced courses. Introductory courses that also serve as an initiation into formal mathematics will focus on the first six chapters. Students already schooled in matrices and linear mappings as well as theorem-proving will quickly proceed to selected chapters from part two. The selection can emphasize algebra or analysis/geometry, as needed. Ample examples, applications, and exercises appear throughout the text, which is supplemented by three helpful Appendixes.
LanguageEnglish
Release dateJul 1, 2014
ISBN9780486795034
Linear Algebra

Related to Linear Algebra

Related ebooks

Mathematics For You

View More

Related articles

Reviews for Linear Algebra

Rating: 1 out of 5 stars
1/5

1 rating1 review

What did you think?

Tap to rate

Review must be at least 10 words

  • Rating: 1 out of 5 stars
    1/5
    This book really sucks to the very end. It has almost no useful examples, and though it has some exercise questions, the author hasn't put any answers of those question. Then how could you call the questions exercise questions? To let us practice using pens? I'd rather think the author is retarded.

Book preview

Linear Algebra - Sterling K. Berberian

Algebra

Linear Algebra

Sterling K. Berberian

Prof. Emer., Mathematics

The University of Texas at Austin

Dover Publications, Inc.

Mineola, NewYork

Copyright

Copyright © 1992, 2014 by S. K. Berberian

All rights reserved.

Bibliographical Note

This Dover edition, first published in 2014, is an unabridged republication of the work originally published in 1992 by Oxford University Press, New York. The author has added a new Errata and Comments section specially for this edition.

Library of Congress Cataloging-in-Publication Data

Berberian, Sterling K., 1926–

Linear algebra / Sterling K. Berberian.–Dover edition.

p. cm.

Originally published: Oxford : Oxford University Press, 1992.

Includes index.

eISBN-13: 978-0-486-79503-4

1. Algebras, Linear. I. Title.

QA184.B47 2014

512'.5—dc23

2014010527

Manufactured in the United States by Courier Corporation

78055401   2014

www.doverpublications.com

For Jim and Susan

Preface

This book grew out of my experiences in teaching two one-semester courses in linear algebra, the first at the immediate post-calculus level, the second at the upper-undergraduate level. The latter course takes up general determinants and standard forms for matrices, and thus requires some familiarity with permutation groups and the factorization of polynomials into irreducible polynomials; this material is normally covered in a one-semester abstract algebra course taken between the two linear algebra courses.

Part 1 of the book (Chapters 1–6) mirrors the first of the above-mentioned linear algebra courses, Part 2 the second (Chapters 7–13); the information on factorization needed from the transitional abstract algebra course is thoroughly reviewed in an Appendix.

The underlying plan of Part 1 is simple: first things first. For the benefit of the reader with little or no experience in formal mathematics (axioms, theorem-proving), the proofs in Part 1 are especially detailed, with frequent comments on the logical strategy of the proof. The more experienced reader can simply skip over superfluous explanations, but the theorems themselves are cast in the form needed for the more advanced chapters of the book; apart from Chapter 6 (an elementary treatment of 2 × 2 and 3 × 3 determinants), nothing in Part 1 has to be redone for Part 2.

Part 2 goes deeper, addressing topics that are more demanding, even difficult: general determinant theory, similarity and canonical forms for matrices, spectral theory in real and complex inner product spaces, tensor products. {Not to worry: in mathematics, ‘difficult’ means only that it takes more time to make it easy.}

Courses with different emphases can be based on various combinations of chapters:

(A) An introductory course that serves also as an initiation into formal mathematics: Chapter 1–6.

(B) An advanced course, for students who have met matrices and linear mappings before (on an informal, relatively proofless level), and have had a theorem-proving type course in elementary abstract algebra (groups and rings): a rapid tour of Chapters 1–5, followed by a selection of chapters from Part 2 tailored to the needs of the class. The selection can be tilted towards algebra or towards analysis/geometry, as indicated in the flow chart following this preface.

It is generally agreed that a course in linear algebra should begin with a discussion of examples; I concur wholeheartedly (Chapter 1, §§1 and 2). Now comes the hard decision: which to take up first, (i) linear equations and matrices, or (ii) vector spaces and linear mappings? I have chosen the latter course, partly on grounds of efficiency, partly because matrices are addictive and linear mappings are not. My experience in introductory courses is that once a class has tasted the joy of matrix computation, it is hard to get anyone to focus on something so austere as a linear mapping on a vector space. Eventually (from Chapter 4 onward) the reader will, I believe, perceive the true relation between linear mappings and matrices to be symbiotic: each is indispensable for an understanding of the other. It is equally a joy to see computational aspects (matrices, determinants) fall out almost effortlessly when the proper conceptual foundation has been laid (linear mappings).

A word about the role of Appendix A (‘Foundations’). The book starts right off with vectors (I usually cover Section 1 of Chapter 1 on Day 1), but before taking up Section 2 it is a good idea to go over briefly Appendix A.2 on set notations. Similarly, before undertaking Chapter 2 (linear mappings), a brief discussion of Appendix A.3 on functions is advisable. Appendix A.1 (an informal discussion of the logical organization of proofs) is cited in the text wherever it can heighten our understanding of what’s going on in a proof. In short, Appendix A is mainly for reference; what it contains is important and needs to be talked about, often and in little bits, but not for a whole hour at a stretch.

The wide appeal of linear algebra lies in its importance for other branches of mathematics and its adaptability to concrete problems in mathematical sciences. Explicit applications are not abundant in the book but they are not entirely neglected, and applicability is an ever-present consideration in the choice of topics. For example, Hilbert space operator theory and the representation of groups by matrices (the applications with which I am most familiar) are not taken up explicitly in the text, but the chapters on inner product spaces consciously prepare the way for Hilbert space and the reader who studies group representations will find the chapters on similarity and tensor products helpful. Systems of linear equations and the reduction of matrices to standard forms are applications that belong to everyone; they are treated thoroughly. For the class that has the leisure to take them up, the applications to analytic geometry in Chapter 6 are instructive and rewarding.

In brief, linear algebra is a feast for all tastes. Bon appétit!

Sterling Berberian

Austin, Texas

August 1990

Flow chart of chapters

Contents

PART I

1   Vector spaces

1.1  Motivation (vectors in 3-space)

1.2  Rn and Cn

1.3  Vector spaces: the axioms, some examples

1.4  Vector spaces: first consequences of the axioms

1.5  Linear combinations of vectors

1.6  Linear subspaces

2   Linear mappings

2.1  Linear mappings

2.2  Linear mappings and linear subspaces: kernel and range

2.4  Isomorphic vector spaces

2.5  Equivalence relations and quotient sets

2.6  Quotient vector spaces

2.7  The first isomorphism theorem

3   Structure of vector spaces

3.1  Linear subspace generated by a subset

3.2  Linear dependence

3.3  Linear independence

3.4  Finitely generated vector spaces

3.5  Basis, dimension

3.6  Rank + nullity = dimension

3.7  Applications of R + N = D

3.9  Duality in vector spaces

4   Matrices

4.1  Matrices

4.2  Matrices of linear mappings

4.3  Matrix multiplication

4.4  Algebra of matrices

4.5  A model for linear mappings

4.6  Transpose of a matrix

4.7  Calculating the rank

4.8  When is a linear system solvable?

4.9  An example

4.10  Change of basis, similar matrices

5   Inner product spaces

5.1  Inner product spaces, Euclidean spaces

5.2  Duality in inner product spaces

5.3  The adjoint of a linear mapping

5.4  Orthogonal mappings and matrices

6   Determinants (2 × 2 and 3 × 3)

6.1  Determinant of a 2 × 2 matrix

6.2  Cross product of vectors in R³

6.3  Determinant of a 3 × 3 matrix

6.4  Characteristic polynomial of a matrix (2 × 2 or 3 × 3)

6.5  Diagonalizing 2 × 2 symmetric real matrices

6.6  Diagonalizing 3 × 3 symmetric real matrices

6.7  A geometric application (conic sections)

PART II

7   Determinants (n × n)

7.1  Alternate multilinear forms

7.2  Determinant of a linear mapping

7.3  Determinant of a square matrix

7.4  Cofactors

8   Similarity (Act I)

8.1  Similarity

8.2  Eigenvalues and eigenvectors

8.3  Characteristic polynomial

9   Euclidean spaces (Spectral Theory)

9.1  Invariant and reducing subspaces

9.2  Bounds of a linear mapping

9.3  Bounds of a self-adjoint mapping, Spectral Theorem

9.4  Normal linear mappings in Euclidean spaces

10   Equivalence of matrices over a PIR

10.1  Unimodular matrices

10.2  Preview of the theory of equivalence

10.3  Equivalence: existence of a diagonal form

10.4  Equivalence: uniqueness of the diagonal form

11   Similarity (Act II)

11.1  Invariant factors, Fundamental theorem of similarity

11.2  Companion matrix, Rational canonical form

11.3  Hamilton–Cayley theorem, minimal polynomial

11.4  Elementary divisors, Jordan canonical form

11.5  Appendix: proof that Mn(F)[t] = Mn(F[t])

12   Unitary spaces

12.1  Complex inner product spaces, unitary spaces

12.2  Orthogonality

12.3  Orthonormal bases, isomorphism

12.4  Adjoint of a linear mapping

12.5  Invariant and reducing subspaces

12.6  Special linear mappings and matrices

12.7  Normal linear mappings, Spectral Theorem

12.8  The Spectral Theorem: another way

13   Tensor products

13.1  Tensor product V ⊗ W of vector spaces

13.2  Tensor product S T of linear mappings

13.3  Matrices of tensor products

Appendix A  Foundations

A.1  A dab of logic

A.2  Set notations

A.3  Functions

A.4  The axioms for a field

Appendix B  Integral domains, factorization theory

B.1  The field of fractions of an integral domain

B.2  Divisibility in an integral domain

B.3  Principal ideal rings

B.4  Euclidean integral domains

B.5  Factorization in overfields

Appendix C  Weierstrass–Bolzano theorem

Index of notations

Index

Errata and Comments

Part one

1

Vector spaces

1.1   Motivation (vectors in 3-space)

1.2   Rn and Cn

1.3   Vector spaces: the axioms, some examples

1.4   Vector spaces: first consequences of the axioms

1.5   Linear combinations of vectors

1.6   Linear subspaces

An arrow has direction and magnitude (its length) (Fig. 1). Quantities with magnitude and direction are familiar from physics (for example, force and velocity); such quantities are called vectorial, and it’s often useful to represent them by arrows—for instance, when determining the net effect of two forces F1 and F2 acting on a point (Fig. 2). (Viewing F1 and F2 as adjacent sides of a parallelogram, F is the diagonal they include.)

Fig. 1

Fig. 2

That’s an intuitive glimpse of vectors. The way mathematicians look at them (as a part of ‘abstract algebra’) leads to the idea of a ‘vector space’.¹ In between this austere view and the intuitive idea, the example of ‘geometric’ vectors in ‘ordinary 3-space’ is fun and instructive; let’s look at it.

1.1  Motivation (vectors in 3-space)

We assume the points of 3-space to have coordinates in the usual cartesian way: pick an origin, choose three mutually perpendicular axes, and assign coordinates to a point by measuring signed distances along the chosen axes (Fig. 3). If P1 and P2 are points of 3-space, Fig. 4 shows an arrow with initial point P1 and final point P2. The progress from P1 to P2 is indicated by measuring the change in each of the coordinates; thus, if P1(x1, y1, z1) and P2(x1, y2, z2) the coordinates, then the components of change in the coordinate directions are the numbers

Fig. 3

Fig. 4

One organizes this data (an ordered triple of real numbers) into the symbol

(Note the order of subtraction: ‘the twos minus the ones’.) For example,

For physical reasons, two arrows are thought of as representing ‘the same vector’ if they point in the same direction and have the same length (think of a vector as a force eligible to be applied at any point of the space). Thus (Fig. 5) the arrow from P1 to P2 represents the same vector as the arrow from P3 to P4 if and only if the figure P1P2P4P3 is a parallelogram (in particular, lies in a plane!). This will happen if and only if the midpoint of the segment P1P4 coincides with the midpoint of the segment P2P3, that is,

Fig. 5

comparing coordinates, we see that this means

in other words,

To summarize, the arrow from P1 to P2 represents the same vector as the arrow from P3 to P4 if and only if

the vector determined by the arrow from P1 to P2. The numbers

are called the components of the vector.

Every ordered triple u = [a, b, c(in many ways!): for instance, if P1 = (2, 3, 5) and P2 = (2 + a, 3 + b, 5 + c. Accordingly, we propose to call such triples vectors.

Real numbers will also be called scalars, to dramatize their difference from vectors (ordered triples of real numbers). {In physics, quantities having only magnitude are called ‘scalar’ quantities.}

Assuming coordinates P1(x1, y1, z1), etc., by definition

Fig. 6

, which means

in other words,

that is,

This shows that each component of is the sum of the corresponding components of and .

In general, the sum of two vectors u = [a, b, c], u′ = [a′, b′, c′] is defined by the formula

so to speak, the vectors are added ‘componentwise’. The message of the preceding paragraph is that

(provided that P1P2P4P3 is a parallelogram!).

is represented by its length, which is the distance |P1P2| from P1 to P2:

This suggests defining length for any vector u = [a, b, c]; it is usually called the norm of u, written ||u||, and it is defined by the formula

In particular,

but points in the opposite direction; since

. Accordingly, for any vector u = [a, b, c], the negative of u, written –u, is defined by the formula

In particular,

are all 0. The vector [0, 0, 0] is called the zero vector and will be denoted θ (to distinguish it from the number 0). Thus

for every point P. Also,

for every vector u; in particular,

for every pair of points P1, P2 of 3-space.

The negative –u of a vector u is obtained by ‘multiplying’ the vector u by –1 (in the sense that each component of u is multiplied by –1); the effect is to reverse the direction of u, while conserving its length. More generally, any vector u = [a, b, c] can be multiplied by any scalar r: one defines

called the scalar multiple of u by r. In particular, –u = (–1)u. Since

we have

In passing from u to ru, length is multiplied by a factor of |r|. If r = 0 then ru = θ; if r ) we say that ru has direction opposite to that of w, and if r > 0 we say that ru has the same direction as u. (However, if u = θ then ru = θ and statements about ‘direction’ are without content.)

To summarize, our study of arrows in 3-space leads us to consider the set of all ordered² triples u = [a, b, c], u′ = [a′, b′, c′], . . . of real numbers; we call such triples vectors (a, b, c being the components of u). There are three natural operations that we perform on vectors:

which we call the sum (of the vectors u and u′), scalar multiple (of the vector u by the real number r) and norm (of the vector u). Each of these operations has a natural geometric (or physical) interpretation when the vectors are thought of as arising from arrows (or, say, forces).

In this book, we undertake a systematic study of sets, called ‘vector spaces’, in which operations u + u′ and ru are defined, satisfying appropriate laws (for instance, u + u′ = u′ + u). We shall see in the following two sections that there is an immense variety of examples of such spaces, thus an immense economy in developing their common properties simultaneously. The appropriate strategy is to set down a list of axioms³ for a ‘vector space’, meaningful and verifiable for the examples we want to encompass, and to study the logical consequences of the axioms; what we learn will then be applicable to all of the relevant examples.

When a norm ||u|| for vectors u is also defined, one speaks of a ‘normed vector space’. In Chapter 5 we study a special type of normed vector space, called Euclidean space, whose ‘geometry’ accords with our intuition in ‘3-dimensional space’ but whose ‘dimension’ can be any positive integer n (goodbye intuition!).

Before setting down the axioms for an abstract vector space (in §1.3), let’s enlarge our list (presently of length 1) of concrete examples; this is the purpose of the next section.

  Exercises

1.  Given the point P(2, –1, 3) and the vector u .

2.  Given the points P(2, –1, 3), Q(3, 4, 1), R(4, –3, 4), S(5, 2, 2). True or false (explain): PQSR is a parallelogram.

3.  Let u = [2, 2, 1], υ = [–2, 1, 2], w = [1, –2, 2]. Show that

and interpret geometrically.

4.  Show that for every triple of points P, Q, R,

{Method 1: Calculate components. Method 2: Draw a picture.}

5.  Given the points P1(x1, y1, z1) and P2(x2, y2, z2), let M be the point such that

where O(0, 0, 0) is the origin. Find the coordinates of M and interpret geometrically.

1.2  Rn and Cn

Rn and Cn are the most important examples of ‘vector spaces’ (defined officially in the next section). The essential idea is that what was done in the preceding section for ordered triples [a, b, c] can be done equally well for ordered n-ples [x1, x2, . . ., xn].

1.2.1  Definition

Fix a positive integer n and let Rn be the set of all ordered n-ples [x1, x2, . . ., xn] of real numbers. {For example, [3, –1/2, √2, π] is an element of R⁴.}

The elements of Rn will be called vectors and the elements of R (that is, real numbers) will be called scalars.

Vectors [x1, x2, . . ., xn] and [y1, y2, . . ., yn] are said to be equal if xi = yi for i = 1, . . ., n; this is expressed by writing [x1, x2, . . ., xn] = [y1, y2, . . ., yn]. If x = [x1, x2, . . ., xn] then xi is called the ith component of the vector x. {Thus, two vectors are equal if and only if they are ‘componentwise equal’.}

If x = [x1, x2, . . ., xn] and y = [y1, y2, . . ., yn], the sum of x and y, denoted x + y is the vector defined by the formula

and if c is a scalar, then the multiple of x by c, denoted cx, is the vector defined by the formula

{Briefly, sums and scalar multiples of vectors are defined ‘componentwise’.}

Similarly, Cn denotes the set of all ordered n-ples of complex numbers, with x + y and cx defined by the above formulas; in this context, the ‘scalars’ are the elements of C (that is, complex numbers).

In the arguments of this section, it doesn’t matter whether the set of scalars is R or C. To cover both cases simultaneously, we let F stand for either R or C and we write Fn for either Rn or Cn. (More generally, F can be any field.⁴)

1.2.2  Definition

The element of Fn all of whose components are zero is denoted θ and is called the zero vector; thus θ = [0, . . ., 0]. If x = [x1, x2, . . ., xn] is any vector, the vector [–x1, . . ., –xn] is called the negative of x and is denoted –x.

The key properties of sums and scalar multiples are collected in the following theorem:

1.2.3  Theorem

Let F be a field (for example, R or C), let n be a positive integer; and let Fn be defined as in 1.2.1.

(1)  If x, y ∈ Fn then x + y ∈ Fn. (Fn is said to be ‘closed under addition’.)

(2)  If c ∈ F and x ∈ Fn then cx ∈ Fn. (Fn is said to be ‘closed under multiplication by scalars’.)

(3)  x + y = y + x for all x, y ∈ Fn. (The addition of vectors is ‘commutative’.)

(4)  (x + y) + z = x + (y + z) for all x, y, z in Fn. (The addition of vectors is ‘associative’.)

(5)  x + θ = x = θ + x for every vector x. (The zero vector is ‘neutral’ for addition.)

(6)  x + (–x) = θ = (–x) + x for every vector x.

(7)  c(x + y) = cx + cy and (c + d)x = cx + dx for all vectors x, y and all scalars c, d. (Multiplication by scalars is ‘distributive’.)

(8)  (cd)x = c(dx) for all scalars c, d and all vectors x. (Multiplication by scalars is ‘associative’.)

(9)  1x = x for every vector x. (’Neutrality’ of 1 for scalar multiplication.)

Proof. Sums and scalar multiples are defined to be elements of Fn, so (1) and (2) are obvious.

(3) Say x = [x1, x2, . . ., xn], y = [y1, y2, . . ., yn]. To show that the vectors x + y and y + x are equal, let’s apply the criterion of 1.2.1: we have to show that for each index i (i = 1, . . ., n), the ith component of x + y is equal to the ith component of y + x. These components are, respectively, xi + yi and yi + xi (by the definition of sum for vectors), and xi + yi = yi + xi by the commutativity of addition in F.

(4)–(9) are proved using the same strategy: the equality of two vectors is established by showing that their corresponding components are equal (by a known property of the scalar field F).

For example, let’s prove the first equality in (7). The ith components of cx and cy are cxi and cyi, so the ith component of cx + cy is cxi + cyi; but this is equal to c(xi + yi) (distributive law in F), which is the ith component of c(x + y). Conclusion: cx + cy = c(x + y).

The elements of Fn are also denoted x = (x1, x1, . . ., xn) (parentheses instead of brackets). For instance, denotes 3-space coordinatized in the usual way, consisting of points (x, y, z); it also denotes the set of vectors [a, b, c] obtained from pairs of points (as in the preceding section). Thus, the symbol is used equivocally, for ordered triples as points (’geometric’ objects) and for ordered triples as vectors (’algebraic’ objects). The use of parentheses or brackets is a way of distinguishing between these two conceptions of .

In practice, the distinction is frequently blurred: one also writes (x1, x2, . . ., xn) for vectors and one speaks of xi as the ith coordinate of [x1, x2, . . ., xn]. Confusion is easily avoided by a liberal use of the words ‘point’ or ‘vector’ as appropriate.

  Exercises

1.  For the vectors u = (1, –2, 7), υ = (2, –1, 3), w = (4, 1, –5) in , compute 2u – 3υ + w.

2.  Consider the vectors u = (2, 1), υ = (–5, 3), w = (3, 4) in . Do there exist real numbers a, b such that au + = w? What if υ = (6, 3)?

3.  Give a detailed proof of (8) of Theorem 1.2.3 (just checking!). {Write out all steps of the proof and give a reason for each step.}

4.  In Definition 1.2.1, n = 1 is not ruled out. What do the elements of look like? Describe sums and scalar multiples in .

1.3  Vector spaces: the axioms, some examples

The properties of Fn listed in Theorem 1.2.3 were not chosen at random; they are just what it takes to be a ‘vector space’ in the sense of the following definition:

1.3.1  Definition

Let F be a field (for example, R or C). A vector space over F is a set V admitting two ‘operations’, called addition and multiplication by scalars, subject to the set of rules given below. The elements of V are called vectors and the elements of F are called scalars. For each pair of vectors x, y, there is determined a vector, denoted x + y, called the sum of x and y. For each vector x and each scalar c, there is determined a vector, denoted cx, called the scalar multiple of x by c. Thus we have two ways of ‘combining’ vectors and scalars:

(1)  if x ∈ V and y ∈ V then x + y ∈ V (V is closed under addition);

(2)  if x ∈ V and c ∈ F then cx ∈ V (V is closed under multiplication by scalars).

The following rules are assumed for sums and scalar multiples (they are also called the axioms, or postulates, or laws of a vector space):

(3)  x + y y + x for all x, y in V (commutative law for addition);

(4)  (x + y) + z = x + (y + z) for all x, y, z in V (associative law for addition);

(5)  there exists a vector θ in V such that x + θ = x = θ + x for all x in V (existence of a zero vector);

(6)  for each x in V there exists a vector –x in V such that x + (–x) = θ = (–x) + x (existence of negatives);

(7)  c(x + y) = cx + cy, (c + d)x = cx + dx for all vectors x, y and all scalars c, d (distributive laws);

(8)  (cd)x = c(dx) for all vectors x and all scalars c, d (associative law for scalar multiplication);

(9)  1x = x for all vectors x (unity law for scalar multiplication).

{The reader who doesn’t like the looks of the next couple of paragraphs is forgiven for jumping ahead to 1.3.2.}

The above definition bristles with the terms ‘for all’, ‘there exists’, ‘such that’, ‘if . . . then’, terms for which there are efficient symbolic abbreviations (Appendix A.1.5, A.2.11): ∀, ∃, ∋, ⇒. When these logical symbols are used, the statements become both compact and unreadable; writing out the symbolic form is still good practice in grasping the logical structure of the statements. That’s not the way to write a book, but it’s useful for taking notes and for recording thought-experiments (for example, when struggling with the exercises); with a pledge not to make an obsession of it, let’s practice a little by rewriting the axioms (1)–(9) for a vector space in symbolic form:

(1)  x, y ∈ V ⇒ x + y ∈ V;

(2)  x ∈ V, c ∈ F ⇒ cx ∈ V;

(3)  x + y = y + x (∀x, y ∈ V);

(4)  (x + y) + z = x + (y + z) (∀x, y, z ∈ V);

(5)  ∃ θ ∈ V ∋ x + θ = x = θ + x (∀x ∈ V);

(6)  (∀x ∈ V) ∃ –x ∈ V ∋ x + (–x) = θ = (–x) + x;

(7)  c(x + y) = cx + cy, (c + d)x = cx + dx (∀ x, y ∈ V and ∀ c, d ∈ F);

(8)  (cd)x = c(dx) (∀ c, d ∈ F and ∀ x ∈ V);

(9)  1x = x (∀ x ∈ V).

The ‘compact symbolic form’ is not unique; for instance, (1) is expressed equally well by

and (3) by

Exercise. Cover up the verbal definitions in 1.3.1 and try to re-create them by decoding the above symbolic forms.

Back to business:

1.3.2  Definition

When F = R in Definition 1.3.1, V is called a real vector space; when F = C, V is called a complex vector space.

When several vector spaces appear in the same context, it is assumed (unless signalled otherwise) that they are vector spaces over the same field (for example, all real or all complex).

Now let’s explore the wealth of examples of vector spaces.

1.3.3  Example

For each positive integer n, Rn is a real vector space and Cn is a complex vector space for the operations defined in the preceding section (Theorem 1.2.3).

1.3.4  Example

Let F be a field, let T be a nonempty set, and let V be the set of all functionsx:T → F. For x, y in V, x = y means that x(t) = y(t) for all t ∈ T. If x, y ∈ V and c ∈ F, define functions x + y and cx by the formulas

for all t ∈ T. (So to speak, sums and scalar multiples in V are defined ‘pointwise’.) Let θ be the function defined by θ(t) = 0 for all t ∈ T, and, for x ∈ V, let –x be the function defined by (–x)(t) = –x(t) for all t ∈ T. It is straightforward to check that V is a vector space over F; it is denotedand is called the space of F-valued functions on T.

Exercise. In the case that T = {1, 2, . . ., n) for some positive integer n, do you see any similarity between V and Fn?

1.3.5  Example

Let F = R or C. If p is a polynomial with coefficients in F, regard p as a function on F in the natural way: if

where the coefficients a0, . . ., an are in F and t is an ‘indeterminate’⁸, define the function p:F → F by the formula

for all c ∈ F (i.e., by substituting c for t). {For example, if p = t² + 5 then pof is also a vector space over F.

1.3.6  Example

If m is a fixed positive integer and if, in the preceding example, one limits attention to the polynomials p = a0 + a1t + a1t² + . . . + amtm (that is, either p = 0 or p is a nonzero polynomial of degree ≤m)of functions is also a vector space (the set of polynomials in question is closed under addition and under multiplication by scalars).

1.3.7  Example

Let F be a field and let V be the set of all ‘infinite sequences’ x = (a1, a2, a3, . . .) of elements of F. If also y = (b1, b2, b3, . . .), x = y means that ai = bi for all i. Define sums and scalar multiples in V ‘term-by-term’, that is, by the formulas

obtained by setting T = P in Example 1.3.4? (P is the set of positive integers.¹⁰)

1.3.8  Example

A minor variation on the preceding example: consider sequences indexed by the set N = {0, 1, 2, 3, . . .} of nonnegative integers, that is, sequences x = (a0, a1, a2, . . .) with ai ∈ F for all i N. {This notation is better adapted to Example 1.3.5.}

1.3.9  Example

Let V be the space described in the preceding example (1.3.8). Consider the set W of all x ∈ V such that, from some index onward, the ai are all 0. {So to speak, W is the set of all ‘finitely nonzero’ sequences, or sequences that are ‘ultimately zero’. For example, x = (0, –3, 1, 5, 0, 0, 0, . . .).} Since W ⊂ V and since W is closed under sums and scalar multiples, it is easy to see that W is also a vector space over F. Assuming F = R or Cof polynomial functions (1.3.5)?

1.3.10  Example

Let V1, . . ., Vn be vector spaces over F and let V = V1 ×. . . × Vn be their cartesian product¹¹, that is, the set of all n-ples (x1, . . ., xn) with xi ∈ Vi for i = 1, . . ., n. Write (x1, . . ., xn) = (y1, . . ., yn) if xi = yi for all i. For x = (x1, . . ., xn), y = (y1, . . ., yn) in V and for c ∈ F, define

Arguments formally the same as for Fn (1.2.3) show that V is a vector space over F for these operations; the only difference is that, instead of citing algebraic properties of the scalar field F, we cite the algebraic laws in the coordinate spaces Vi. This example is important enough to merit an official definition:

1.3.11  Definition

With notations as in Example 1.3.10, V is called the product vector space (or ‘direct sum’) of the vector spaces V1, . . ., Vn. For n = 2, we write simply V = V1 × V2; for n = 3, V = V1 × V2 × V3.

These examples¹² give us an idea of the mathematical objects that can be put into the context of the axioms for a vector space. In the next section we turn to the question of what can be gotten out of the axioms, that is, what properties can be inferred from the properties (1)–(9) postulated in the definition of a vector space.

  Exercises

1.be the set of all functions x:T → V. Show that W can be made into a vector space over F in a natural way. {Hint: Use the definitions in Example 1.3.4 as a guide.}

2.  The definition of a vector space (1.3.1) can be formulated as follows. A vector space over F is a nonempty set V together with a pair of mappings

(σ suggests ‘sum’ and μ suggests ‘multiple’) having the following properties: σ{x, y) = σ(y, x) for all x, y in V; σ(σ(x, y), z) = σ(x, σ(y, z)) for all x, y, z in V; etc. The exercise: Write out the ‘etc.’ in detail.

3.  Let V be a complex vector space (1.3.2). Show that V is also a real vector space (sums as usual, scalar multiplication restricted to real scalars). These two ways of looking at V may be indicated by writing VC and VR.

4.  Every real vector space V can be ‘embedded’ in a complex vector space W in the following way: let W = V × V be the real vector space constructed as in Example 1.3.10 and define multiplication by complex scalars by the formula

for a, b R and (x, y) ∈ W. {In particular, i(x, y) = (–y, x). Think of (x, y) as ‘x + iy’.}

Show that W satisfies the axioms for a complex vector space. (W is called the complexification of V.)

1.4  Vector spaces: first consequences of the axioms

Our first deductions from the axioms (1.3.1) have to do with the uniqueness of certain vectors mentioned in the axioms:

1.4.1  Theorem

Let V be a vector space over a field F, with notations as in 1.3.1.

  (i)  If θ′ is a vector such that θ′ + x = x for all x ∈ V, then θ′ = θ; thus, the vector θ whose existence is postulated in 1.3.1 is unique.

 (ii)  If x + y = θ, necessarily y = –x; thus, the vector –x whose existence is postulated in 1.3.1 is uniquely determined by x.

(iii)  θ + θ = θ; and if z is a vector such that z + z = z, necessarily z = θ.

Proof.

  (i)  We have θ′ = θ′ + θ = θ, where the first equality holds by the property of θ postulated in (5) of 1.3.1, and the second equality holds by the assumption on θ′. (See also Exercise 5.)

 (ii)  Suppose x + y = θ. Adding –x (the vector provided by (6) of 1.3.1) to both sides of the equation, we have, successively, –x + (x + y) = –x + θ, (–x + x) + y = –x, θ + y = –x, y = –x.

(iii)  If z + z = z, then θ = z + (–z) = (z + z) + (–z) = z + (z + (–z)) = z + θ = z. On the other hand, θ + θ = θ by (5) of

1.4.2  Corollary

For every vector x, –(–x) = x.

Proof. Since –x + x = θ by (6) of Definition 1.3.1, we have x = –(–x

1.4.3  Corollary

For every vector x, 0x = θ; for every scalar c, = θ.

Proof. Let z = 0x. Using one of the distributive laws (1.3.1) at the appropriate step, we have

therefore z = θ by (iii) of the theorem. The equality = θ

1.4.4  Corollary

For every vector x and every scalar c, c(–x) = –(cx) = (–c)x.

Proof. θ = = c(x + (–x)) = cx + c(–x), therefore c(–x)

Enjoying the preview?
Page 1 of 1