<<

LINEAR A Pathway to Abstract 3rd Edition

G. Viglino

Ramapo College of New Jersey January 2018

Contents

CONTENTS

CHAPTER 1 MATRICES AND SYSTEMS OF LINEAR EQUATION 1.1 Systems of Linear Equations 1 1.2 Consistent and Inconsistent Systems of Equations 14 Chapter Summary 27

CHAPTER 2 VECTOR SPACES 2.1 Vectors in the , and Beyond 31 2.2 Abstract Vector Spaces 40 2.3 Properties of Vector Spaces 51 2.4 Subspaces 59 2.5 Lines and Planes 68 Chapter Summary 75

CHAPTER 3 BASES AND DIMENSION 3.1 Spanning Sets 77 3.2 86 3.3 Bases 94 Chapter Summary 109

CHAPTER4 LINEARITY 4.1 Linear Transformations 111 4.2 Image and 124 4.3 Isomorphisms 135 Chapter Summary 149

CHAPTER 5 MATRICES AND LINEAR MAPS 5.1 151 5.2 Invertible Matrices 165 5.3 Matrix Representation of Linear Maps 178 5.4 Change of 191 Chapter Summary 202 Contents

CHAPTER 6 AND EIGENVECTORS 6.1 Determinants 205 6.2 Eigenspaces 218 6.3 Diagonalization 233 6.4 Applications 246 6.5 Markov Chains 259 Chapter Summary 275

CHAPTER 7 INNER PRODUCT SPACES 7.1 279 7.2 Inner Product 292 7.3 301 7.4. The Spectral Theorem 315 Chapter Summary 307 Appendix A Principle of Mathematical Induction Appendix B Solutions to Check Your Understanding Boxes

Appendix C Answers to Selected Odd-Numbered Exercises Preface

PREFACE There is no mathematical ramp that will enable you to continuously inch your way higher and higher in mathematics. The climb calls for a ladder consisting of discrete steps designed to take you from one mathematical level to another. You are about to take an important step on that lad- der, one that will take you to a plateau where mathematical abstraction abounds. rests on a small number of axioms (accepted rules, or “laws”), upon which a beautiful and practi- cal theory emerges. Technology can be used to reduce the time needed to perform essential but routine tasks. We fea- ture the TI-84+ calculator, but any graphing utility or Algebraic System will do. The real value of whatever technological tool you use is that it will free you to spend more time on the development and comprehension of the theory and its applications. In any event, if you haven’t already discovered in other courses:

MATHEMATICS DOES NOT RUN ON BATTERIES

Systems of linear equations are introduced and analyzed in Chapter 1. Graphing utilities can be used to solve such systems, but understanding what those solutions represent plays a dominant role throughout the text. We begin Chapter 2 by sowing the seeds for vector spaces in the fertile , where they soon blossom into the concept of an abstract vector. The remainder of Chapter 2 and all of Chapter 3 are dedicated to a study of vector spaces in isolation. Functions from one to another which, in a sense, respect the algebraic structure of those spaces are investigated in Chapters 4 and 5. The sixth chapter focuses on Eigenvalues and Eigenvectors, along with some of their important applications. The first six chapters may provide a full plate for most one-semester courses. If not, then Chap- ter 7 (on inner product spaces) is offered for dessert.

We have made every effort to provide a leg-up for the step you are about to take. Our primary goal was to write a readable book, without compromising mathematical integrity. Along the way, you will encounter numerous Check Your Understanding boxes designed to challenge your under- standing of each newly introduced concept. Complete solutions to the problems in those boxes appear in Appendix B, but please don’t be in too much of a hurry to look at our solutions. You should TRY to solve the problems on your own, for it is only through ATTEMPTING to solve a problem that one grows mathematically. In the words of Descartes: We never understand a thing so well, and make it our own, when we learn it from another, as when we have discovered it for ourselves.

1.1 Systems of Linear Equations 1

1 CHAPTER 1 MATRICES AND SYSTEMS OF LINEAR EQUATIONS Much of the development of linear algebra calls for the solution and interpretation of systems of linear equations. While the “solution part” can be relegated to a calculator, the “interpretation part” cannot. We focus on the solution-part of the process in Section 1, and on the more important interpretation-part in Sections 2. §1. SYSTEMS OF LINEAR EQUATIONS To solve the system of equations: 2x +64y – 4z =   2x ++6y 4z = 0  xy++2z = – 2 is to determine values of the variables (or unknowns) x, y, and z for which each of the three equations is satisfied. You certainly solved such systems in earlier courses, and if you take the time to solve the above Brackets are used to denote sets. In particular, system, you will find that it has but one solution: –11– 1 x ===–11 y  z –1 . We can also say that the three-tuple denotes the set containing but one element—the ele- –11– 1 is a solution of the given system of equation, and that ment –11– 1 . –11– 1 is its solution set. In general: An (ordered) n-tuple is an expression of the form c1c2  cn , where each ci is a real number (written ci   ), for 1 in .

We say that the n-tuple c1c2  cn is a solution of the system of m equations in n unknowns:

The x s denote variables a11x1 ++a12x2 a1nxn = b1  i  (or unknowns), while the a x ++a x a x = b 21 1 22 2 2n n 2  aij ’s and bi ’s are con- . ... stants (or scalars). . ... . ... am1x1 ++am2x2 amnxn = bm 

if each equation in the system is satisfied when ci is substi- tuted for xi , for 1 in . The set of all solutions of a system of equations is said to be the solution set of that system. 2 Chapter 1 Matrices and Systems of Linear Equations

EQUIVALENT SYSTEMS OF EQUATIONS Consider the system of equations: – 3x +2y =   x 1  ---2+ y = ---  2 2  As you know, you can perform certain operations on that system which will not alter its solution set. For example, you can: (1) Interchange the order of the equations: – 3x +2y =  x 1   ---2+ y = ---  x 1   2 2  ---2+ y = ---   2 2  – 3x +2y =  (2) Multiply both sides of the resulting top equation by 2: x 1  ---2+ y = ---  x +14y =  2 2     – 3x +2y =  – 3x +2y =  (3) Multiply the resulting top equation by 3 and add it to the bottom You used this third equation: maneuver a lot when eliminating a variable x +14y =  x +14y =  from a given system of    equations For example: – 3x +2y =  13y = 5  (i) x +13yz– =  multiply by 3  3x +312y = (2) 2x –35y +3z =   – 3x +2y = (3) – 3x ++y 2z = 2 add: 13y = 5 multiply (1) by -2 and add it to (2) The above three operations, are said to be elementary equation –511y +1z =   operations: 10yz–5=  multiply (1) by 3 and add it ELEMENTARY OPERATIONS ON to (3) SYSTEMS OF LINEAR EQUATIONS Interchange the order of any two equations in the system. Multiply both sides of an equation by a nonzero number. Add a multiple of one equation to another equation.

EQUIVALENT SYSTEM Two systems of equations sharing a common solution set are said to OF EQUATIONS be equivalent. As you may recall: THEOREM 1.1 Performing any of elementary opera- tions on a system of linear equations will result in an equivalent system of equations. 1.1 Systems of Linear Equations 3

AUGMENTED MATRICES Matrices are arrays of numbers (or expressions representing numbers) arranged in rows and columns:

47 17 0 10 2134 10 6 36 5 12 7 4 1 –739 –38 11 2– 12 9 (i) (ii) (iii) (iv) (v) Matrix (i) contains 2 rows and 3 columns and it is said to be a 23 (two-by-three) matrix. Similarly, (ii) is a 32 matrix, and (iii) is a 33 matrix (a matrix). In general, an mn matrix is a matrix consisting of m rows and n columns. In particular, (iv) is a 13 matrix (a row matrix), and (v) is a 31 matrix (a column matrix). It is often convenient to represent a system of equations in a more com- pact matrix form. The rows of the matrix in Figure 1.1(b), for example, concisely represents the equations in Figure 1.1(a). Note that the vari- ables x, y, and z are suppressed in the matrix form, and that the vertical recalls the equal in the equations. Such a matrix is said to be the augmented matrix associated with the given system of equations. 2x +64y – 4z =   24–6 4 2x ++6y 4z = 0 26 4 0 AUGMENTED MATRIX  xy++2z = – 2 11 2– 2 System of Equations Augmented Matrix (a) (b) Figure 1.1 Switching two equations in a system of equations results in the switching of the corresponding rows in the associated augmented matrix. Indeed, each of the three previously introduced elementary equation operations corresponds with one of the following elementary matrix row operations:

ELEMENTARY MATRIX ROW OPERATIONS Interchange the order of any two rows in the matrix. Multiply each element in a row of the matrix by a nonzero number. Add a multiple of one row of the matrix to another row of the matrix. The following terminology is motivated by Theorem 1.1: DEFINITION 1.1 Two matrices are equivalent if one can be EQUIVALENT derived from the other by performing elemen- MATRICES tary row operations. 4 Chapter 1 Matrices and Systems of Linear Equations

HERE IS WHERE WE ARE AT THIS : A system of linear equations can be represented by an aug- mented matrix, and every augmented matrix represents a sys- tem of linear equations. Moreover:

SYSTEMS OF EQUATIONS ASSOCIATED WITH EQUIVA- LENT AUGMENTED MATRICES ARE THEMSELVES EQUIVALENT (SAME SOLUTION SET).

AND HERE IS WHERE WE ARE GOING: Suppose you want to solve the system of equations [1] in Figure 1.2. Assume, for the time being, that you can go from its augmented matrix ([2]) to matrix [3], via elementary row operations. System [4], which is associated with the augmented matrix [3], is easily seen to have the solution: x ===–1 y 1 z –1 . But this must also be the solution of system [1], since the two systems of equations are also equivalent!

2x +64y – 4z =  24–6 4  augmented matrix [1] 2x ++6y 4z = 0 2640 [2]  xy++2z = – 2 112– 2

Same Solution Set Elementary Row Operations

x ++0y 0z = – 1 100– 1 [4]  system of equations 0xy++0z = 1 0101 [3]  0x ++0yz= – 1 001– 1 Figure 1.2 The remainder of this section is designed to illustrate a method which can be used to go from matrix [2] of Figure 1.2 to matrix [3], via ele- mentary row operations.

PIVOTING ABOUT A PIVOT POINT Capital letters are used to represent matrices, and double subscripted lower case letters for their entries; as in:

a11 a12 a13 a14

A = a21 a22 a23 a24

a31 a32 a33 a34 1.1 Systems of Linear Equations 5

Note that the first subscript of the element aij denotes its row: i, and the second subscript, its column: j. For example, if: 2762 Aa==ij 3631 –2231

then a12 ===7 a21 3 a34 3 , and so on. In the next example, we specify a location in a given matrix (called PIVOT POINT the pivot-point), which contains a non-zero entry. We then illustrate a PIVOTING process (called pivoting) designed to turn the given matrix into an equivalent matrix with a 1 in the pivot-point, and with each entry above or below the pivot-point equal to 0. It is a routine process that plays a dominant role in a number of matrix applications, so please make sure that you understand it fully. The following notation will be used to represent elementary row operations:

ELEMENTARY ROW NOTATION Switch row i with row j: Ri  Rj

Multiply each entry in row i by a nonzero cR  R number c: i i

Multiply each entry in row i by a number cR + R  R c, and add the resulting row to row j: i j j

EXAMPLE 1.1 Pivot the matrix: 2 64–2 Aa==ij 36315 –2231

about the pivot point a11 with pivot entry 2, and then again about the pivot point a22 of the resulting equivalent matrix.

SOLUTION: Step 1. Get a 1 in the pivot-point by multiplying each entry

in row 1 by --1- : 2 1 ---R1  R1 2 64–22 1 32–1 36315 36315 –2231 –2231 6 Chapter 1 Matrices and Systems of Linear Equations

Step 2.Get 0’s below (there is no above) the pivot point position. Multiply row 1 by –3 and add it to row 2 (see margin), and then multiply row 1 by 1 and add it to row 3: –3  1 3 – 2 1

–3R 1 32–1 1 32–1 1 32–1 1 –93 –63– – 3R + R  R 1 2 2 1R1 + R3  R3 R 2 36315 36315 0 –9123 0 –9123 – 3R +:R 1 2 0 –9123 –2231 –2231 0 504

13–1 2 –2231 0 504 Repeating the above two-step procedure, we now pivot about

a22 = –3 :

1 32–11 –---R  R 1 32–1 1 0 713 1 0 713 3 2 2 – 3R2 + R1  R1 – 5R + R  R 0 –9123 0 1 –43 – 0 1 –43 – 2 3 3 0 1 –43 – 0 504 0 50 4 0 50 4 0015 24

03–912 05–1520 13–1 2 0504 1 0 713 0015 24

GRAPHING CALCULATOR GLIMPSE 1.1 We utilize a graphing calculator to perform the first of the two pivot- ing processes in the above example, and invite you to use your calcu- lator to address the other pivoting process. The TI-84+ calculator is fea- tured throughout the text. 1.1 Systems of Linear Equations 7

CHECK YOUR UNDERSTANDING 1.1

10 7 13 ??0 ? Pivot about a33 = 15 to go from: 01–4 3– to ??0 ? Answer: See page B-1. 0015 24 ??1 ?

ROW-REDUCED-ECHELON FORM A matrix may have many different equivalent forms. Here is the nic- est of them all: DEFINITION 1.2 A matrix is in row-reduced-echelon form A matrix satisfying (i), (ii) when it satisfies the following three conditions: and a slightly weaker form ROW-REDUCED of (i): ECHELON FORM (i) The first non-zero entry in any row is 1 The first non-zero entry (called its leading-one), and all of the entries in any row is 1, and the above or below that leading-one are 0. entries below (only) that leading-one are 0 (ii) In any two successive rows, not consisting is said to be in row-echelon entirely of zeros, the leading-one in the form. lower row appears further to the right than the leading-one in the row above it. (iii) All of the rows that consist entirely of zeros are at the bottom of the matrix. These three matrices are in row-reduced-echelon form: The matrices 10012 1000 10012 19012 103 and A = 010 5 B = 0011 C = 010 5 014 5 010 001 0 001 0 001 0 0000 are in row-echelon form

CHECK YOUR UNDERSTANDING 1.2 Determine if the given matrix is in row-reduced-echelon form. If not, list the condition(s) of Definition 1.2 which are not satisfied.

0010 10–0 3 01–05 3 100 a 0000 b 0001 c d 00013 Answer: Yes: (a), (c), and (d). 014 No: (b) [fails (ii)] 0000 0100 00000 Though a bit tedious, reducing a matrix to its row-reduced-echelon form is a routine task. Just focus on getting those all-important leading- ones (which are to be positioned further to the right as you move down), with zeros above and below them. Consider the following example: 8 Chapter 1 Matrices and Systems of Linear Equations

EXAMPLE 1.2 Perform elementary row operations to obtain the row-reduced-echelon form for the matrix: 24–6 4 26 4 0 11 2– 2

SOLUTION: Leading-one Step. We could divide the first row by 2 to get a leading-one in that row, but choose to switch the first row and third row instead:

24–6 4 1 12– 2 R  R 26 4 0 1 3 26 4 0 11 2– 2 24–6 4 Zeros-above-and-below Step: –22 –4–4 –22 –4–4 2640 24–6 4 0 404 0 28–10

1 12– 2 – 2R + R  R 1 12– 2 1 12– 2 1 2 2 – 2R1 + R3  R3 26 4 0 0 40 4 0 40 4 24–6 4 24–6 4 0 28–10 Next leading-one Step:

1 12– 2 1 1 12– 2 ---R2  R2 0 40 4 4 0 1 01 0 28–10 0 28–10

Zeros-above-and-below Step: 01–01– 02–02– 112– 2 02–10 8 1 0 23– 00–88

1 12– 2 – 1R + R  R 1 0 23– 1 0 23– 2 1 1 – 2R2 + R3  R3 0 1 01 0 1 01 0 1 01 0 28–10 0 28–10 00–88 1.1 Systems of Linear Equations 9

Next leading-one Step:

1 0 23– 1 1 0 23– –---R3  R3 0 1 01 8 0 1 01 00–88 001 –1

Zeros-above-and-below Step:

00–2 2 102– 3 1 00–1

1 0 23– 1 00–1 – 2R3 + R1  R1 0 1 01 0 1 0 1 001 –1 001 –1 We are now at a row-reduced-echelon form, and so we stop.

While not difficult, the above example illustrates that obtaining the row-reduced-echelon form of a matrix can be a bit tedious. It’s a dirty job, but someone has to do it:

GRAPHING CALCULATOR GLIMPSE 1.2

Row-reduced-echelon form

In harmony with graphing calculators, we will adopt the notation rrefA to denote the row-reduced-echelon form of a matrix A.

EXAMPLE 1.3 Solve the system: 2x +64y – 4z =   2x ++6y 4z = 0  xy++2z = – 2 10 Chapter 1 Matrices and Systems of Linear Equations

SOLUTION: All the work has been done: Example 1.2 augmented matrix x y z x y z 2x +64y – 4z =  x = –1   24–6 4 100– 1  2x ++6y 4z = 0 26 4 0  010 1 y = 1    xy++2z = – 2 11 2– 2 001– 1 z = –1 

These two systems of equations are equivalent (same solution sets) From the above we can easily spot the solution of the given system:

x ===–1 y 1 z –1

CHECK YOUR UNDERSTANDING 1.3 Proceed as in Example 1.3 to solve the given system of equations. xyz++= 6  3x +42yz– =  Answer:  x ===1 y 2 z 3 3xy++2z = 11

A bit of human-intervention was used in the pivoting process of Example 1.2. If it is “freedom from choice” that you want, then you can use the following algorithm to reduce a given matrix to its row- reduced-echelon form: Gauss, Karl Friedrich (1777 -1855), the great German mathematician Gauss-Jordan Elimination Method and astronomer. Wilhelm Jordan (1842- Step 1. Locate the left-most column that does not consist 1899) German professor entirely of zeros, and pick a nonzero element in that of geodesy. column. Let the position of that chosen element be the pivot-point. Step 2. Pivot about the pivot-point of Step 1. Step 3. If necessary, switch the pivot-row with the furthest row above it (nearest the top) that does not already contain a leading-one, to the left of the pivot col- umn. Step 4. If the matrix is in row-reduced-echelon form, then you are done. If not, return to Step 1, but ignore all rows with established leading-ones for that step of the process. 1.1 Systems of Linear Equations 11

EXERCISES

Exercises 1-2. Write down the augmented matrix associated with the given system of equations.

3x – 3y +2z =  2x +53y – 4w =    1. 5x +15y – 9z = –  x – 4z +1w = –   2.  –43x – y +0z =  x –04y =   – x – y ++z 4w = 9

Exercises 3-4. Write down the system of equations associated with the given augmented matrix.

514 3 2410 9 3. –32 –1 4 4. 0552 2 1 --- –01 0 21–8 3 11 2

Exercises 5-8. Perform elementary row operations to obtain the row-reduced echelon form for the given matrix.

024 1002 025 2301 5. 6. 102 2121 7.442 8. 1012 245 0224 103 1001 235 2201

Exercises 9-11. Solve the system of equations corresponding to the given row-reduced-echelon form matrix. x y z w x y z x1 x2 x3 x4 x5 10002 10 01 100001 9. 0100– 3 010002 01 00 10. 11. 00 12 00101 001002 00010 000102 00001– 1

Exercises 12-15. Proceed as in Example 1.2 to solve the given system of equations.

x – 2y +1z =  xy– –2z =    12.–53x +5y – 2z = –  13. 4x –52y –2z = –    4x –38y +6z =  –3x ++y 6z = 0 12 Chapter 1 Matrices and Systems of Linear Equations

2xy–2= zw++1 2x +15y – 2z = –   wx– = y    14. – 4x – y +4z = –  15. 1   4y + 3z = –---  x +42yz– =  2  x + 3y = wz– 

16. Construct a system of three equations in three unknowns, x, y, and z such that x ===1 y 2 z 3 is a solution of the system. 17. Construct a system of four equations in four unknowns, x, y, z, and w with solution set x ===1,2 y ,3 z ,4 w =. Exercises 18-20. (Row-Echelon Form) A matrix is said to be in row-echelon form if it satisfies all of the conditions of Definition 1.2, except that elements above a leading 1 need not be zero (the entries below leading ones must still be zero). Determine if the matrix is in row-echelon form. If not, indicate why not.

1231 1032 0012 18.0123 19.0120 20. 0100 0001 0002 0001

Exercises 21-22. Perform elementary row operations to transform the given matrix to the given row-echelon form (see Exercise 18-20).

24–6 4 12–3 2 13–1 2 26–2 4 21. 2640 01 4– 3 01–4 3– 22. 36315 112– 2 00 1– 1 8 –2231 001--- 5

Exercises 23-25. Determine the solution set of the system of equations corresponding to the given row-echelon form matrix (see Exercise 18-20). Note: If you are using a graphing calculator, then you might as well use the row-reduced- echelon command, for that is the most revealing form. If you are doing things by hand, however, you may be able to save some time by going with the row-echelon form. x y z x y z x y z w 12 01 11 02 12012 23.01 10 24. 01 12 25. 0123– 3 00 12 00 10 00111 00010 26. Offer an argument to justify the following claim: If the jth column of a matrix A consists entirely of zeros, and if the matrix B is equiva- lent to A, then the jth column of B also consists entirely of zeros. 1.1 Systems of Linear Equations 13

In the remaining exercises you are to decide whether the given statement is True or False. If True, then you are to present a general argument to establish the validity of the statement in its most general setting. If False, then you are to exhibit a concrete specific example, called a counterexample, showing that the given statement does not hold in general. To illustrate: Prove or Give a Counterexample: (a) The sum of any two even integers is again an even integer. (b) Every odd number is prime. (a) Yes, each time you add two even integers, out pops another even integer, suggesting that statement (a) is true. But you certainly can’t check to see if (a) holds for all even integers—case by case—as there are infinitely many such cases. A general argument is needed: If a and b are even integers, then a = 2n and b = 2m for some integers n and m. We then have: ab+2==n +22m nm+ . Since ab+ is itself a multiple of 2, it is even. (b) Surely (b) is false. Why? Because 9 is odd, but 9 is not prime, that’s why. To be sure, we could offer a different counterexample, say 15, or 55, but we did have to come up with a specific concrete counterexample to shoot down the claim.

PROVE OR GIVE A COUNTEREXAMPLE

ax+0 by =  27. The system of equations  has a solution for all abcd   . cx+0 dy = 

ax+1 by =  28. The system of equations  always has a solution for all abcd   . cx+1 dy = 

ax+0 by =  29. The system of equations  can never have more than one solution. cx+0 dy =  30. The systems of equations associated with the two augmented matrices:

ab 0 and a b 0 cd 0 c d 0 will have the same solution set only if aa===', b b', c c' , and dd= ' .

31. If the matrix A has n rows, and if rrefA contains less than n leading ones, then the last row of rrefA must consist entirely of zeros. 14 Chapter 1 Matrices and Systems of Linear Equations

1

§2. CONSISTENT AND INCONSISTENT SYSTEMS OF EQUATIONS A system of equations may have a unique solution, infinitely many solutions, or no solution whatsoever. If it has no solution, then the sys- CONSISTENT tem is said to be inconsistent, otherwise it is said to be consistent. As INCONSISTENT is illustrated in the following examples, the solution set of any system of equations can be spotted from the row-reduced-echelon form of its augmented matrix. EXAMPLE 1.4 Determine if the following system of equations is consistent. If so, find its solution set. 4x –72y –5z =   –56x ++y 10z = – 11  –23x ++y 5z = – 5

SOLUTION: Proceeding as in the previous section, we have: x y z x y z 4x –72y –5z =  x = 6  42–7–5 1 006 –56x ++y 10z = – 11  –6 5 10– 11 rref 0 1 01–  y = –1  –23x ++y 5z = – 5 –2553 – 001 3 z = 3 We see that the given system is consistent, and that it has but one solution: x ==6 y –1 z =3 .

EXAMPLE 1.5 Determine if the following system of equations is consistent. If so, find its solution set. 3x –72y –5z =   –56x ++y 10z = – 11  –32x ++y 4z = – 3  –23x ++y 5z = – 5 SOLUTION: Proceeding as in the previous section, we have: x y z x y z 3x –72y –5z =   32–7–5 1 0 0 0 –56x ++y 10z = – 11 –6 5 10– 11 0 1 0 0  rref –32x ++y 4z = – 3 –3432 – 00 1 0  –23x ++y 5z = – 5 –2553 – 00 0 1 0x ++0y 0z = 1 Since the equation represented by the last row in the above rref-matrix cannot be satisfied, the given system of equations is inconsistent. 1.2 Consistent and Inconsistent Systems of Equations 15

EXAMPLE 1.6 Determine the solution set of the system: 3x –36y +9w =   –42x ++y 2zw– = – 11  3x –68y ++z 7w = – 5 1 SOLUTION: x y z w x y z w 1 0002 3x –36y +9w =   36–03 9 1 1 2 rref 0 1 0 –--- –--- -1/2 –42x ++y 2zw– = – 11 –4212 – –11 2 2 -5/2  3x –68y ++z 7w = – 5 1 5  38–67–5 001 --- –--- 2 2 Figure 1.3 We know that the solution set of the above system of equations coin- cides with that of the one stemming from the row-reduced-echelon form of its augmented matrix; namely: x +++0y 0z 0w = 2 x = 2  1 1  1 1  0xy++0z – ---w = –---  y = – --- + ---w  2 2  or: 2 2    1 5 1 0x +++0yz---w = –--5-  z = – --- – ---w  2 2  2 2  Any variable that is not associated with a leading As you can see, that variable w, which we moved to the right of one in the row-reduced the equations, can be assigned any value whatsoever, after which echelon form of an aug- the values of x, y, and z (the variables associated with leading- mented matrix is said to be a free variable. In the cur- ones in Figure 1.3) are determined. For example, setting w = 0 rent setting, the variable w leads to the particular solution: is a free variable (see rref 1 5 in Figure 1.3). x ==2 y –--- z =–--- w =0 2 2 We can generate another solution by letting w = 1 : x ===2 y 0 z –3 w =1 Indeed, the solutions set of the system of equation is obtained by letting wc= , where c can be any real number whatsoever:

1 c 5 c x ==2 y – --- + --- z =–--- – --- w =c c   2 2 2 2 Read: such that We can arrive at a nicer representation of the solution set by replacing each c with 2c: 1 5 x ==2 y – --- + cz =–2--- – c w =c 2c   2 2 and then observing that as “2c runs through all of the numbers,” so does c: 1 5 x ==2 y – --- + cz =–2--- – c w =c c   2 2 16 Chapter 1 Matrices and Systems of Linear Equations

The system of equations of Example 1.4 has a unique solution, that of Example 1.5 has no solutions, and the one in Example 1.6 has infinitely many solutions. These examples cover all of the bases, for if a system of equations has more than one solution then it must have infinitely many solutions (Exercise 22).

CHECK YOUR UNDERSTANDING 1.4 Determine if the system associated with the given row-reduced-eche- lon augmented matrix is consistent. If it is, find its solution set.

Answer: (a) Inconsistent 10– 2 1 10– 2 1 1 2 01 1 (b) 12+ r 45– rr r   (a) 01 5 4 (b) 01 5 4 (c) –2 (c): 0 014 12– r – sr –42 – ssrs   R 00 0 2 00 0 0 0000 0

Our concern thus far has been with systems of equations with fixed con- stants on the right side of the equations. We now turn to the question of whether or not a system of equations has a solution for all such constants: EXAMPLE 1.7 Determine if the following system of equations is consistent for all a, b, and c. 2xz+ = a   3xy+ = b   –5x – y – z = c 

SOLUTION: x y z 1 a 2xz+ = a  201a 1 1 0 ------ ---R1  R1 2 2 3xy+ = b  310b 2  310b –5x – y – z = c –51 –1– c  –51 –1– c

–3R1 + R2  R2 1R1 + R3  R3

1 a 1 0 ------1 a 2 2 1 0 ------2 2 3 3a 5R2 + R3  R3 Unlike the TI-84+, the TI-89 0 1 –--- –------+ b 3 3a and above have symbolic capa- 2 2 0 1 –--- – ------+ b bilities. In particular: 2 2 1 a 0 – 5–------+ c 00 –8 – 7a ++5bc 2 2 1 –---R  R 8 3 3 x y z 1 a a ++5bc 1 0 ------1 0 0 ------2 2 16 1 3 3a –---R + R  R – 3a + b – 3c 0 1 –--- – ------+ b 2 3 1 1 0 1 0 ------2 2 3 16 ---R + R  R 7a – 5b – c 2 3 2 2 7a – 5b – c 00 1 ------001 ------8 8 1.2 Consistent and Inconsistent Systems of Equations 17

We see that the given system of equations has a solution for all a, b, a ++5bc – 3a + b – 3c 7a – 5b – c and c; namely: x = ------ y = ------ z = ------16 16 8 EXAMPLE 1.8 Determine if the following system of equations is consistent for all possible values of a, b, c, and d. 2xy++2zw + = a   x –32y ++z 2w = b   2yz++4w = c   xz–4– w = d  SOLUTION: If you go through the Gauss-Jordon elimination method without making a mistakes you will find that: x y z w 2xy++2zw + = a  2121a  x –32y ++z 2w = b  12–32b  0214c 0x +++2yz4w = c   10–4 1– d x + 0yz–4– w = d  x y z w 8a –73b – c 1 0 0 – 2 ------13 Here, unlike with the a –42b + c 0 1 0 1 ------smaller system of equa- 13 tions in Example 1.8, the rref –42a ++b 5c TI-89 (or higher) is of lit- 0 0 1 2 ------tle help: 13 –710a ++b 12c +13d 0 0 0 0 ------13 Figure 1.4 The last row of the above rref-matrix represents the equation: 0x +++0y 0z 0w = ------–710a ++b 12c +13d 13 As such, that matrix tells us that the given system of equation will have a solution if and only if: –710a +b +12c +13d = 0 (*) The last row of the above In particular, if you choose random numbers for a, b, c, and d, then it rref matrix tec ++3ab= 0 lls us that is very unlikely that the system will have a solution, for what are the there is no solution to the odds that those four numbers will satisfy (*)? system, but it “lies,” for solutions do exist for cer- tain values of a, b, c, and d CHECK YOUR UNDERSTANDING 1.5 [see (*)]. Determine if the given system of equations has a solution for all a, b, and c. If not, find some specific values of a, b, and c for which a solu- tion does not exist. 4x – 2y + z = a  x –44y – z = a  Answer: (a) It is consistent   for all a, b, and c. (a) –42x ++y 2z = b  (b) 2x + 8y – 12z = b  (b) Consistent if and only if   c ++3ab= 0 5xy–4+ z = c  –12x ++y 2z = c  18 Chapter 1 Matrices and Systems of Linear Equations

COEFFICIENT MATRIX The matrix of an mn system of equation is the mn matrix obtained by eliminating the last column of the augmented matrix of the system. For example, referring to system of equations of Examples 1.8, we have: augmented 2121a matrix 12–32b 2xy++2zw + = a  0214c  x –32y ++z 2w = b  10–4 1– d  2yz++4w = c  21 2 1  xz–4– w = d coefficient  matrix 12–32 0214 10–4 1– At this point, it behooves us to introduce a bit of notation. To begin with, we will use S to represent a general system of linear equations. We will then let aug(S) and coef (S) denote the augmented and coeffi- cient matrices of S, respectively. To illustrate: 2x +13yz– =   23–1 1 23– 1 For S: 3xy–2+5z =  augS ==31–25 and coefS 31–2  –2x +2y – 3z =  –231 –2 –231 – The following theorem will enable us to invoke a graphing calculator to resolve the issues of Examples 1.7 and 1.8: THEOREM 1.2 The system of equations: Let P and Q be two proposi- SPANNING a11x1 ++a12x2 a1nxn = b1  tions (a proposition is a math- THEOREM  ematical statement that is a x ++a x a x = b S: 21 1 22 2 2n n 2  either true or false). To say “P . ... if and only if Q,” (also writ- ......  ten in the form PQ ) is to  a x ++a x a x = b say that if P is true then so is m1 1 m2 2  mn n m  Q (also written PQ ), and has a solution for all values of b1b2  bm if if Q is true then so is P (also written QP ). and only if rref coefS does not contain a row consisting entirely of zeros.

PROOF: If rref coefS does not contain a row consisting entirely of free: set to 0 zeros, then each row of rref coefS will have a leading one, as will x x x x x x 1 2 3 4 5 6 every row of rref augS . For any given values of b1b2  bm , a 120003– 9 solution for S can then be obtained by setting each of the nm– free 0010012 variables in rref aug S to zero, and letting the variable associated 0001001  th 0000155 with a leading one in the i row of rref augS equal the last entry in that row (see margin for an illustration). a solution:–902150 1.2 Consistent and Inconsistent Systems of Equations 19

For the converse, assume that the last row of rref coefS consists entirely of zeros. The only difference between rref coefS and rref augS is that the latter has an additional column, the last entry of which (as was the case in Figure 1.6) must be a linear expression involving b1b2  bm , say Fb1b2  bm :

x1 x2 x3 . . . xm x1 x2 x3 . . . xm same

0 0 0 . . . 0 0 0 0 . . . 0 Fb1b2  bm rref coefS rref augS (see Example 1.8)

It follows that for any values of b1b2  bm for which Fb1b2  bm  0 , the resulting system of equations will not have a solution, for here is its last equation:

0x1 +++0x2  0xn = Fb1b2  bm

EXAMPLE 1.9 Use the spanning theorem to determine if the given system of equations has a solution for all values of the constants on the right side of the equations. 2xy++2zw + = a  2xz+ = a    x –32y ++z 2w = b  (a) 3xy+ = b  (b)   2yz++4w = c  –5x – y – z = c   xz–4– w = d  See Example 1.7 See Example 1.8

SOLUTION: 2xz+ = a   201 1 00 (a) S: 3xy+ = b  coef(S) 310 rref  coef  S 0 1 0  (see margin) –5x – y – z = c  –51 –1– 001

does not contain a row of zeros: system has a solution for all values of a, b, and c

2xy++2zw + = a   2121 1 00– 2 x –32y ++z 2w = b  coef(S) 12–32rref coefS 0 1 01 (b) S:  2yz++4w = c  0214(see margin) 001 2  xz–4– w = d  10–4 1– 000 0

contain a row of zeros: system does not have a solution for all values of a, b, c, and d 20 Chapter 1 Matrices and Systems of Linear Equations

Note that while the matrix rref coefS in (b) shows that the sys- tem S is not consistent for all values of a, b, c, and d, it does not reveal the specific values of a, b, c, and d for which a solution does exist. That information can be derived from the matrix rref augS (see Exam- ple 1.8).

CHECK YOUR UNDERSTANDING 1.6 Use the spanning theorem to determine if the given system of equa- tions has a solution for all values of the constants on the right side of the equations. x – 3y + w = a  3x + 7yz– = a    3xy–2+ z – 3w = b  (a) 13x –24y + z = b  (b)   xz+ – 5w = c  2x –24y + z = c   Answer: (a) Yes (b) No 2xy– + z – 2w = d 

HOMOGENEOUS SYSTEMS OF EQUATIONS A system of linear equations is said to be homogeneous if all of the A system with fewer equa- constants on the right side of the equations are zero: tions than unknowns  (“wide”) is said to be a11x1 + a12x2 +  + a1nxn =0 underdetermined.  H: a x + a x +  + a x =0  A system with more equa- 21 1 22 . 2 2n. n .  tions than unknowns . . . . (“tall”) is said to be overde- . . . .  termined.  am1x1 + am2x2 +  + amnxn =0  A square system is a system which contains as many (A homogeneous system of m equations in n unknowns) equations as unknowns. It is easy to see that every homogeneous system is consistent, with trivial solution: x1 ==0 x2 0 xn =0 . In the event that the homogeneous system is “wide”, then it has more than one solution:

THEOREM 1.3 Any homogeneous system S of m linear equa- FUNDAMENTAL THEO- tions in n unknowns with nm has nontrivial REM OF HOMOGENEOUS SYSTEMS OF EQUATIONS solutions.

PROOF: Having more columns than rows, rref augS must have free variables, and therefore the system has infinitely many solutions. EXAMPLE 1.10 Determine the solution set of: 2x ++03y –54z w =   – 3x ++y 4zw += 0  x ++07y –114z w =  1.2 Consistent and Inconsistent Systems of Equations 21

SOLUTION: Theorem 1.3 tells us that the system has nontrivial solu- tions. Let’s find them: x y z w 2x ++03y –54z w =  23–50 4  aug S S: – 3x ++y 4zw += 0 –14103  x ++07y –114z w =  17–110 4

x y z w free variables 16 2 16 2 211 1 0 –------0 x = ------z – ------w 17 11 11 11 11 11 rref 4 17 4 17 0 1 –------0 y = ------z – ------w 11 11 11 11 00 0 0 0 Figure 1.5 Assigning arbitrary values to the two free variables z and w we arrive at the solution set of the system: x y z w

} 16 2 4 17} ------a – ------b ------a – ------bab ab  R = 16a – 2b 4a – 17b11a 11b ab  R 11 11 11 11 If S is a homogeneous system of equations, then the last column of rref augS will always consist entirely of zeros (Exercise 20). Con- sequently, when solving a homogeneous system of equations, one might as well start with coefS rather than with augS (one less col- umn to carry along in the rref-process, that’s all). In particular, the solu- tion set of the homogeneous system of the last example can easily be read from rref coefS (just mentally add a column of zeros to the right of the matrix):

x y z w x y z w 2 1 0 –-----16------0 2x ++03y –54z w =  23–5 4 11 11 coef S rref – 3x ++y 4zw += 0 4 17 S :  –1413 0 1 –------0  11 11 x ++07y –114z w =  17–11 4 0 0 0 0 0

CHECK YOUR UNDERSTANDING 1.7 Determine the solution set of: 2x +++3y 4z 5w = 0  3xy++4zw += 0  Answer: x +++7y 4z 11w = 0 4r –2r–23r r rR 22 Chapter 1 Matrices and Systems of Linear Equations

While underdetermined (“wide”) homogeneous systems of equations are guaranteed to always have non-trivial solu- tions, this is not the case with overdetermined (“tall”) sys- tems of equations [see Exercises 27-28], or with square systems of equations [see Exercises 29-30].

We end this section with a rather obvious result, but one that will play an important role in future developments; so much so, that we label it accordingly: THEOREM 1.4 A homogeneous system S of m linear equa- LINEAR INDEPEN- tions in n unknowns has only the trivial solu- DENCE THEOREM tion if and only if rref coefS has n leading ones.

PROOF: Since there are n unknowns, to say that rref coefS has n leading ones, is to say that it has no free variables. 1.2 Consistent and Inconsistent Systems of Equations 23

EXERCISES

Exercises 1-6. Determine if the system S with given rref augS is consistent. If it is, find its solution set.

01 2 0101 3 100 –2 1. 00 0 2. 0010 –2 3. 010 –2 00 0 0000 0 001 –2

10 3 5 1000130 100005–0 2 5. 4. 01– 4 2 0010021 6. 010002–3 1 00 0 1 0001201 001000 1 2 000010 0 0

Exercises 7-12. Determine if the system of equations is consistent. If it is, find its solution set.

2x ++3yz= 4 x –44y –1z =  4x – 2y +4z =  7.  8.   xy++2z = 5 2xy+8– 2z =  9. –42x ++y 2z = 10  5xy–4+2z = 

x –44y –1z =  2x ++3yz–2w = 4 – x +++wyz= 3    10. 2x +88y – 12z =  xy–2+3zw– =  6x ++44z –32y w =   11.  12.  –12x ++y 2z = 1 2yz+1– 2w =  5y –63x – w –1z = –    6x –63y +15z =  –720x –10w –8z +18y = – 

Exercises 13-14. Does the system of equations have a solution for all a, and b? If not, find some specific values of a and b for which a solution does not exist, and some specific values of a and b, not both zero, for which a solution does exist.

2x ++3yz= a  x –44y – z = a  13.  14.  xy++2z = b  2xy+ – 2z = b 

Exercises 15-16. Does the system have a solution for all a, b, and c? If not, find some specific val- ues of a, b, and c for which a solution does not exist, and some specific values of a, b, and c, not all zero, for which a solution does exist.

x –44y – z = a  4x – 2y + z = a    15. 2x + 8y – 12z = b  16. –42x ++y 2z = b    –12x ++y 2z = c  5xy–4+ z = c  24 Chapter 1 Matrices and Systems of Linear Equations

Exercises 17-19. Use the Spanning Theorem to determine if the system of equations has a solu- tion for all values of a, b, c, and d.

2xy– = a  4x – 2y + z = a    z – 3w = b  –42x ++y 2z = b  4x – 2y + z = a  17.  18.  19.  2x + 2z = c  5xy–4+ z = c  –42x ++y 2z = b    y + 2z = d  2xyz++ = d 

20. Let S is a homogeneous system of equations. Prove that the last column of rref augS contains only zeros. 21. Prove that if mn , then the system of equations:

a11x1 ++a12x2 a1nxn = b1   a x ++a x a x = b 21 1 22 2 2n n 2  . ... . ... . ... am1x1 ++am2x2 amnxn = bm 

cannot have a solution for all values of b1b2  bm .

22. (a) Show that if xx==0 y y0 and xx==1 y y1 are solutions of the system: ax+ by = c  , then, xx==0 + kx1 – x0  y y0 + ky1 – y0 is also a solution for any dx+ ey = f  given k   . Suggestion: Substitute the above expressions for x and y into the given system. (b) Generalize the above argument to show that if a system of equations has more than one solution, then it must have infinitely many solutions Exercises 23-26. Determine the solution set of the given underdetermined (“wide”) homogeneous system of equations.

2x +03yz– =  2x +03yz– =  23.  24.  4x ++6y 2z = 0 4x +03y – 2z = 

2x ++3yz–4w = 0 2x ++03y –42z w =    25. –53x –2y +0z – 3w =  26. –53x –2y +0z – 3w =    –3x –2y ++z 7w = 0 –3x –2y ++z 7w = 0 1.2 Consistent and Inconsistent Systems of Equations 25

Exercises 27-28. Determine if the given overdetermined (“tall”) homogeneous system of equa- tions has a unique solution. 2x +++3y 4z 6w = 0 2x +03y – 4z =    x +++3y 5z 2w = 0 3x ++2yz= 0  27.  2xy++6z +7w = 0 x +04y – 9z =  28.   5x +++3y 2zw= 0 – 4x –6y –0z =   2x +++4y 6z 2w = 0 3xy++4zw += 0 Exercises 29-30. Determine if the given square homogeneous system of equations has a unique solution.

29. 5x ++03y –54z w =  30. 2x +++5yz4w = 0   – x –2y +0z – 9w =  –23x –4y ++z 6w = 0   3x –23y +0zw– =  4xy++–62z w = 0   11xy+0–92z – w =  9x –23y –0z +0w =  Exercises 31-33. For what values of a will the given homogeneous system of equations have a unique solution?

xay+0=  xyz++= 0 xyz++= 0 31.    ax+0 y =  32. x +02yz– =  33. xayz+0– =    – x ++yaz= 0 – x ++yaz= 0

Exercises 34-36. For what values of a and b will the given homogeneous system of equations have a unique solution?

xay+0=  xay+0=  xyz++= 0 34.  35.   2xby+0=  bx+0 y =  36. xayz+0– =   – x ++ybz= 0

ax+0 by =  37. For what values of a, b, c, and d will the homogeneous system of equations  cx+0 dy =  have a unique solution:

38. Show that if x0 y0 is a solution of a given two by two homogeneous system of equations,

then kx0 ky0 is also a solution for any k   .

39. Show that if x0 y0 and x1 y1 are solutions of a given two by two homogeneous system

of equations, then x0 + x1 y0 + y1 is also a solution. 26 Chapter 1 Matrices and Systems of Linear Equations

a11xa+ 12y = b1  40. Let M be the solution set of S:  and let T be the solution set of the corre- a21xa+ 22y = b2 

a11xa+012y =  sponding homogeneous system H:  . Show that: a21xa+022y = 

(a) If x0 y0  M and x1 y1  M , then x0 – x1 y0 – y1  T .

(b) If x0 y0  M and x1 y1  T , then x0 + x1 y0 + y1  M

a11xa+ 12y = b1  41. Let M be the solution set of S:  and let T be the solution set of the corre- a21xa+ 22y = b2 

a11xa+012y =  sponding homogeneous system H:  . Show that for any x0 y0  M , a21xa+022y = 

Mx= 0 + x1 y0 + y1 x1 y1  T .

PROVE OR GIVE A COUNTEREXAMPLE

42. The system of equations associated with the augmented matrix abc 0 is consistent, def a independent of the values of the entries a through f.

43. The system of equations associated with the augmented matrix abc 1 is consistent, def 0 independent of the values of the entries a through f.

44. The system of equations associated with the augmented matrix 16 3 is consistent if 0 d 0 and only if d = 0 . 45. If a homogeneous system of equations has a nontrivial solution, then it has infinitely many solutions.

a11xa+012y =  46. If the homogeneous system  has only the trivial solution, then the system a21xa+022y =  a11xa+ 12y = b1   has a unique solution for all b1 b2 . a21xa+ 22y = b2 

47. Any system S of m linear equations in n unknowns with nm has nontrivial solutions.

48. A system of n linear equations in m unknowns S is consistent if and only if rref coefS has m leading ones. Chapter Summary 27

CHAPTER SUMMARY

N-TUPLE An (ordered) n-tuple is an expression of the form c1c2  cn , where each ci is a real number, for 1 in .

SOLUTION SET OF A An n-tuple c1c2  cn is a solution of the system of m SYSTEM OF EQUATIONS equations in n unknowns:

a11x1 ++a12x2 a1nxn = b1   a21x1 ++a22x2 a2nxn = b2  S: . ... . ... . ... am1x1 ++am2x2 amnxn = bm 

if each of the m equations is satisfied when ci is substituted for

xi , for 1 in . The solution set of S is the set of all solutions of S.

CONSISTENT AND A system of equations is said to be consistent if it has non- INCONSISTENT SYS- empty solution set. A system of equations that has no solution is said to be inconsistent. TEMS OF EQUATIONS

EQUIVALENT SYSTEMS Two systems of equations are said to be equivalent if they have OF EQUATIONS equal solution sets.

OVERDETERMINED, A system of m equations in n unknowns is said to be: UNDERDETERMINED, Overdetermined if nm (more equations than unknowns). AND SQUARE SYSTEMS Underdetermined if nm (fewer equations than unknowns). OF EQUATIONS Square if nm= .

ELEMENTARY The following three operations on a system of linear equations EQUATION OPERATIONS are said to be elementary equation operations: Interchange the order of any two equations in the system. Multiply both sides of an equation by a nonzero number. Add a multiple of one equation to another equation. Elementary row operations Performing any number of elementary equation operations on a do not alter the solution sets system of linear equations will result in an equivalent system of of systems of equations. equations (same solution set). 28 Chapter 1 Matrices and Systems of Linear Equations

MATRICES Matrices are arrays of numbers arranged in rows and columns, such as the matrix A below:

a11 a12 a13 a14

A ===a21 a22 a23 a24 also: A34 aij  or A aij

a31 a32 a33 a34 Since A has 3 rows and 4 columns, it is said to be a three-by- four matrix. When the number of rows of a matrix equals the number of columns, the matrix is said to be a .

ELEMENTARY ROW The following three operations on any given matrix are said to OPERATIONS be elementary row operations: Interchange the order of any two rows in the matrix. Multiply each element in a row of the matrix by a nonzero num- ber. Add a multiple of one row of the matrix to another row of the matrix.

EQUIVALENT MATRICES Two matrices are equivalent if one can be derived from the other by means of a sequence of elementary row operations.

AUGMENTED MATRIX The augmented matrix of a system of equations S is that matrix aug(S) composed of the of the equations in, along with the constants to the right of the equations in S. For example:

2xyz+2– = –  21– 1 –2  aug(S) S: x ++3y 2z = 9 132 9  – x –2y +1z =  –11 –2 1

Equivalent systems of equa- Two systems of equations, S1 and S2 , are equivalent if and tions corresponding to equiv- only if their corresponding augmented matrices, aug S and alent augmented matrices. 1 aug S2 , are equivalent.

ROW-REDUCED-ECHELON All rows consisting entirely of zeros are at the bottom of the FORM OF A MATRIX matrix. All of the other rows contain a leading-1 (with zeros in all entries above or below it), and those leading-ones “move” to the right, as you “move” down the matrix.

Gauss-Jordan Elimination The Gauss-Jordan Elimination Method of page 10 can be Method. used to obtain the row-reduced-echelon form rrefA of a given matrix A. Chapter Summary 29

COEFFICIENT MATRIX The coefficient matrix of a system of equations S is that matrix coef (S) composed of the coefficients of S. For example:

2xyz+2– = –  21– 1  coef (S) S: x ++3y 2z = 9 132  – x –2y +1z =  –11 –2 forget about the constants on the right of the equal signs

Spanning Theorem The system of equations:

a11x1 ++a12x2 a1nxn = b1   a x ++a x a x = b S: 21 1 22 2 2n n 2  . ... . ... . ...  am1x1 ++am2x2 amnxn = bm 

has a solution for all values of b1b2  bm if and only if rref coefS does not contain a row consisting entirely of zeros.

HOMOGENEOUS SYSTEM OF A system of equations of the form: EQUATIONS a x + a x +  + a x =0  11 1 12 2 1n n  a x + a x +  + a x =0  21 1 22 . 2 2n. n .  S: ......   am1x1 + am2x2 +  + amnxn =0  with zeros to the right of the equal sign, is said to be homoge- neous.

TRIVIAL SOLUTION x1 ==0 x2 0 xn =0 is a solution of the above homoge- neous system of equation. It is said to be the trivial solution of the system.

Fundamental Theorem of Any homogeneous system S of m linear equations in n Homogeneous Systems unknowns with nm has nontrivial solutions.

You can use rref coefS to Let S be a homogeneous system of equations. The only differ- solve a homogeneous system ence between rref augS and rref coefS is that the of equations S former contains an additional column of zeros. Being aware of this, you might as well focus on rref coefS to derive the solution set of S. Linear Independence A homogeneous system S of m linear equations in n unknowns Theorem has only the trivial solution if and only if rref coefS has n leading ones. 30 Chapter 1 Matrices and Systems of Linear Equations 2.1 Vectors in the Plane and Beyond 31

2 CHAPTER 2 VECTOR SPACES We begin this chapter with a geometrical consideration of vectors as directed line segments in the plane and in three dimensional space, and then extend the vector concept to higher dimensional Euclidean spaces. Abstraction is the nature of mathematics, and we let the “essence” of spaces guide us, in Section 2, to the definition of an abstract vector space. In Section 3 we begin to uncover some of the beautiful (and very practical) theory of abstract vector spaces, an exca- vation that will keep us well-occupied for the remainder of the text. Subsets of vector spaces which are themselves vector spaces are con- sidered in Section 4. In Section 5, we return to the two and three dimen- sional Euclidean spaces of the first section and derive a vector representation for lines and planes in those spaces.

§1. VECTORS IN THE PLANE AND BEYOND

We begin by considering vectors in the plane, such as those in Figure 2.1, which are depicted as directed line segments (“arrows”). In that geometrical setting, the arrow is pointing in the direction of the vector, with the of the arrow representing its . B B v = AB D . v = AB . A A v = CD C. (a) (b) Figure 2.1 Vectors will be denoted by boldface lowercase letters. The vector v = AB in Figure 2.1 is said to have initial point A, and terminal point B. One defines two vectors to be equal if they have the same mag- nitude and direction. If you pick up the vector v in Figure 2.1(a) and move it in a fashion as we did in Figure 2.1(b) to the vector with initial point C and terminal point D, then you will still have the same vector: v ==AB CD

In particular, the vector v in Figure 2.2 with initial point Ax= 0 y0 and terminal point Bx= 1 y1 can be moved in a parallel fashion so that its initial point coincides with the . When so placed, the vector is said to be in standard position. 32 Chapter 2 Vector Spaces

y Bx= 1 y1 . v Ax= 0 y0

v x1 – x0 y1 – y0 tor in standard position vec . x Figure 2.2

EXAMPLE 2.1 Sketch the vector with initial point –23 and terminal point 41 – . Position that vector in standard position, and identify its terminal point.

SOLUTION: The figure below tells the whole story: y 6 –23 . Pick up the top vector and move it 2 units down and 3 4 units to the right to the right v so that its initial point x –23 . In the process, the original terminal point v 41 – 41 – is also moved 2 4 units to the right at 3 units – 1 – 3 down, coming to rest at terminal point 64 – . 6 64 – 42– – Figure 2.3 A standard position vector in the plane is completely determined by the coordinates of its terminal point. This observation enables us to identify the vector in Figure 2.3 as an ordered pair of numbers or 2- Note that the two-tuple in tuple; namely: the expression v = 64 – v = 64 – appears in bold-face, so as to distinguish it from the In a similar fashion we may refer to the vectors v and w in Figure 2.4(a) form 64 – which rep- resents a point in the plane. as v = 23 and w = –32 – . z y 23 4 v 134 u x -3 -2 -1 1 2 3 y w 3 1 –32 – x (a) (b) Figure 2.4 2.1 Vectors in the Plane and Beyond 33

Likewise, the vector u in the 3 dimensional space of Figure 2.4(b) can be described by the bold-faced 3-tuple u = 134 The beauty of all of this is that while we cannot geometrically repre- sent a vector in 4 dimensional space, we can certainly consider 4- tuples, and beyond. With this in mind, let us agree to denote the set of n all (ordered) n-tuples a1a2  an by the symbol  : n  = a1a2  an ai  

The real numbers ai in the n-tuple a1a2  an are said to be the components of the n-tuple, and we define two n-tuples to be equal if their corresponding components are identical:

DEFINITION 2.1 a1a2  an = b1b2  bn if n-TUPLE EQUALITY ai = bi , for 1 in .

SCALAR PRODUCT AND SUMS OF VECTORS Vectors evolved from the need to adequately represent which are characterized by both magnitude and direction. In a way, these quantities themselves tell us how we should go about defining cer- tain algebraic operations on vectors. Suppose, for example, that the vec- tor v = 32 of Figure 2.5(a) represents a . Doubling the magnitude of that force without changing its direction would result in the vector force labeled 2v in that figure, as that vector is in the same direction as v = 32 , with length twice that of v.

2v = 64 v = 32

v = 32 1 –---v = –1 –--2- 3 3 (a) (b) Figure 2.5 Length of v = 32 : Similarly, if a force that is one-third that of v = 32 is applied in the 32 +1322 = opposite direction of v, then the vector representing that new force is 1 the vector –---v in Figure 2.5(b); for that vector is in the opposite direc- 3 Length of –1 –--2- : 3 tion of v = 32 , with length one-third that of v.

2 2 4 –1 2 +1–--- = + --- This stretching or shrinking of a vector, in one direction or the other,  3 9 is an important operation which we now formalize and extend to the set ==-----13------13- 9 3 of n-tuple-vectors for any positive integer n: 34 Chapter 2 Vector Spaces

DEFINITION 2.2 n To any vector vv= 1v2  vn in  , and any r   , we let: PRODUCT rvrv= 1rv2  rvn The vector rv is said to be a scalar multiple of v. For example: 315 = 315 –510– 4 = –5020 213– 45 = 23 2– 4 25 2

VECTOR If two people pull on an object positioned at the origin with v and w, then the observed combined effect is the same as that of one individual pulling with force z, where z is the vector coinciding with the diagonal in the parallelogram formed by the vectors v and w [Fig- ure 2.6(a)]. y y v1 + w1 v2 + w2

w2 w v1 v2 z z v v v2 w

x x v1 w1 (a) (b) Figure 2.6 The above vector z is said to be the sum of the vectors v and w, and is denoted by vw+ . Figure 2.6(b) reveals that:

zvw==+ v1 v2 + w1 w2 =v1 + w1 v2 + w2 Generalizing, we have: While identical in shape, the “+” in vw+ differs in DEFINITION 2.3 The sum of the vectors vv= 1v2  vn spirit from that in vi + wi : ECTOR UM n the latter represents the V S and ww= 1w2  wn in  , is denoted familiar sum of two num- bers, as in 37+ , while the by vw+ and is given by: former represents the vw+ = v + w  v + w  v + w newly defined sum of two 1 1 2 2 n n n-tuples, as in: 32 – + 711 2.1 Vectors in the Plane and Beyond 35

EXAMPLE 2.2 For v ==–2315  w 150– 2, and r = 2 , determine the vector rvw+ .

SOLUTION: rvw+2= –2315 + 150– 2 Definition 2.2: = –46210 + 150– 2 Definition 2.3: = – 4 + 16 + 52 + 0102 – = –31128

CHECK YOUR UNDERSTANDING 2.1 For v ====32– 2 w –310  r 2 s –3 , determine the Answer: 15 1– 4 vector rv + sw

EUCLIDEAN VECTOR SPACES

The set of n-tuples  n , together with the above defined operations of vector addition: v1v2  vn + w1w2  wn = v1 + w1 v2 + w2 vn + wn and : rv1v2  vn = rv1rv2  rvn is called the Euclidean n-space. Every contains a most distinguished vector: DEFINITION 2.4 The zero vector in  n , denoted by 0, is that ZERO VECTOR vector with each component the number 0: 000=  0 For example 0=  000 is the zero vector in 3 , and 0=  0000 is the zero vector in 4 . No direction is associated with the zero vector. A zero force, for example, is no force at all, and its “direction” would be a moot point.

Every real number r has an –r , namely that number which when added to r yields 0. Similarly:

DEFINITION 2.5 n For given vv= 1v2  vn  the ADDITIVE INVERSE additive inverse of v is denoted by –v and is OF A VECTOR given by: –v = –v1–v2  –vn

n Note that for every vv= 1v2  vn  : vv+ – ==v1v2  vn + –v1–v2  –vn 0 36 Chapter 2 Vector Spaces

We are now in a position to list the properties of Euclidean spaces which will morph into the definition of an abstract vector space in the next section: THEOREM 2.1 Let u, v, and w be vectors in the Euclidean n- space n , and let r and s be scalars (real num- bers). Then: (i)uv+ = vu+ ()

(ii)uv+ + w = uvw+ + (Vector ) Addition: (iii) v + 0 = v (Zero Property)

{ (iv)vv+ – = 0 (Inverse Property)

(v)ruv+ = ru + rv (Vector ) Scalar and Addition: { (vi)rs+ u = ru + su (Scalar Distributive Property) (vii)rsu = rs u (Scalar Associative Property) Scalar: { (viii)1uu= (Identity Property)

PROOF: We establish (ii) in 2 , (v) in 3 , and (vii) in 3 , and invite you to verify the rest on your own.

To emphasize the important role played by definitions, the symbol  instead of = will tem- porarily be used to indicate a step in the proof which follows directly from a definition. In addi- tion, the abbreviation “PofR” will be used to denote that a step follows directly from a Property of the Real numbers, all of which will be assumed to hold; for example, the additive associative property of the real numbers: ab+ + c = abc+ + .

(ii): uv+ + w = uvw+ + (in 2 ): This associative property If uu== u  v v  v  and w =w  w , then: eliminates the need for 1 2 1 2 1 2 including parenthesis when summing more than two uv+ + w  u1 u2 + v1 v2 + w1 w2 vectors. In particular, Definition 2.3: uvw++  u1 + v1 u2 + v2 + w1 w2 is perfectly well defined. Definition 2.3:  u1 + v1 + w1 u2 + v2 + w2

POFR: = u1 + v 1 + w1  u2 + v2 + w2  Definition 2.3:  u1 u2 + v1 v2 + w1 w2  uvw+ + 2.1 Vectors in the Plane and Beyond 37

(v): ruv+ = ru + rv (in 3 ):

If uu==1u2 u3 and vv1v2 v3 , then:

ruv+  ru1u2 u3 + v1v2 v3 Definition 2.3:  ru1 + v1 u2 + v2 u3 + v3 Definition 2.2:  ru1 + v1  ru2 + v2  ru3 + v3 PofR: = ru1 + rv1ru2 + rv2 ru3 + rv3 Defintion 2.3:  ru1 ru2 ru3 + rv1 rv2 rv3 Definition 2.2:  ru1u2 u3 + rv1v2 v3  ru + rv (vii): rsu = rs u (in Rn ):

rsu  rsu1u2  un

Definition 2.2:  rsu1su2  sun

Definition 2.2:  rsu1  rsu2  rsun PofR: = rs u1rs u2  rs un Defintion 2.2:  rs u1u2  un  rs u

In this, and any other abstract math course: DEFINITIONS RULE! Just look at the above proof. It contains but one “logical step,” the step labeled PofR; all other steps hinge on DEFINITIONS.

CHECK YOUR UNDERSTANDING 2.2

Answer: See page B-3. Establish Theorem 2.1(iv), vv+ – = 0 , in 3 and in n . 38 Chapter 2 Vector Spaces

EXERCISES

Exercises 1-6. Sketch the vector with given initial point A and terminal point B. Sketch the same vector in standard position in the plane, identifying its terminal point.

1.A ==–12  B 01 2.A ==33  B 01 – 3. A ==11  B –22

4.A ==10  B 01 – 5.A ==–12 –  B 11 – 6. A ==22  B 12 –

Exercises 7-10. Express, as a 3-tuple, the vector with given initial point A and terminal point B.

7.A ==123  B 321 8. A ==–450  B 251 – 

9.A ==01– 9 B –902 10. A ==–35– 3 B 353 – 

Exercises 11-14. Perform the indicated vector operations.

11. 532 – ++01 –24 – 12. 25 ++13 ––23

13.– 231– 5 + –1 – 200 14. – ––1234 + 312– 2

Exercises 15-18. Find the vector v such that:

15.v + 24 – = –42 16. v +204135 = –

17.42 +35–v =  + –12 – 18. 43– 1 +2132–v = –

19. For u = 13 , v = 24 , and w = 62 – , find scalars r and s such that: (a) ru + sv = w (b) –ru + sw = v (c) rv + –sw = u

20. Find scalars r, s, and t, such that: r130 ++s216 t146 = 756

21. Find scalars r, s, and t, such that: –r130 ++s216 –t146 = 756 22. Show that there do not exist scalars r, s, and t, such that r235 ++s325 t123 = 124

23. Find the vector ab  2 of length 5 that has the same direction as the vector with initial point 13 and terminal point 31 .

24. Find the vector ab  2 of length 5 that is in the opposite direction of the vector with ini- tial point 13 and terminal point 31 .

25. Prove Theorem 2.1(ii) for: (a) 3 (b) n . 2.1 Vectors in the Plane, and Beyond 39

26. Prove Theorem 2.1(v) for: (a) 2 (b) n .

27. Prove Theorem 2.1(i) for: (a) 2 (b) 3 (c) n

28. Prove Theorem 2.1(iii) for: (a) 2 (b) 3 (c) n

29. Prove Theorem 2.1(vi) for: (a) 2 (b) 3 (c) n

30. Prove Theorem 2.1(viii) for: (a) 2 (b) 3 (c) n

31. Prove that if v, w, and z, are vectors in 3 such that vw+ = vz+ , then wz= .

PROVE OR GIVE A COUNTEREXAMPLE

32. For v  n , if rv = sv then rs= .

33. For v  0  n , if rv = sv then rs= .

n 34. For v1 v2   and r  0 , if rv1 = rv2 then v1 = v2 .

35. For v  n , rv = 0 if and only if r = 0 or v = 0 . 40 Chapter 2 Vector Spaces

2

§2. ABSTRACT VECTOR SPACES

One of the main objectives of abstract mathematics is to isolate and analyze a particular structure of the real number system, so as to better focus on its “essence.” The essence of the vector structure in n tabu- lated in Theorem 2.1, page 36, leads us to the definition of an abstract vector space: DEFINITION 2.6 A (real) vector space is a nonempty set V along with two operations, called vector addi- The elements of a vector VECTOR SPACE space V are called vec- tion and scalar multiplication. The operation tors, and will be denoted of addition assigns to any two element u and by bold-faced letters (like v). Scalars will continue v in V, another element uv+ in V. The oper- to be denoted by non- ation of scalar multiplication assigns to any bold-faced letters (like r). real number r (also called a scalar), and any element v in V, another element rv in V. These operations must satisfy the following eight axioms for all uvw  V , and all rs   :

(i)uv+ = vu+ (Commutative Axiom) (ii)uv+ + w = uvw+ + (Vector Associative Axiom) (iii) There is a vector in V, denoted Addition: by 0 such that v + 0 = v for every vector v in V. (Zero Axiom) (iv) For every vector v in V, there is a vector in V, denoted by –v such that vv+ – = 0 . (Additive Inverse Axiom) (v)ruv+ = ru + rv (Vector Distributive Axiom) Scalar and Addition: { (vi)rs+ v = rv + sv (Scalar Distributive Axiom)

Scalar: (vii)rsv = rs v (Scalar Associative Axiom) { (viii)1vv= (Identity Axiom) While eight axioms are specifically listed in the above definition, two A set is said to be closed, more are lurking within the above so-called closure statements: with respect to an operation, if elements of that set sub- V is closed under addition: For every v and w in V, vw+  V jected to that operation remain in the set. For exam- V is closed under scalar ple, the set of positive inte- For every v  V and r   , rv  V gers is closed under addition multiplication: (the sum of two positive integers is again a positive integer), but that set is not closed under . 2.2 Abstract Vector Spaces 41

It is important to note that while the two addition signs in rs+ v = rv + sv are identical in appearance, they do not represent a common operator: The “+” in rs+ v denotes the sum of two real numbers, as in 25+ v = 7v , while the “+” in rv + sv denotes the sum of two vectors, as in 2v + 5v . By the same token, the two “products” in the expression rs v also denote We also point out that, by distinct operators: convention, no meaning is The rs in rs v denotes the product of two real numbers, attributed to an expression of the form vr , wherein a resulting in another number, as in 25 = 10 , while the scalar vector v appears to the left product 10v represents a vector. of a scalar r. We already have infinitely many vector spaces at our disposal, namely the Euclidean n-spaces. We now turn our attention to several others.

MATRIX SPACES

EXAMPLE 2.3 The set of two-by-two matrices: ab M22 = abcd  R cd with addition and scalar multiplication given by:

ab + a b = aa+ bb+  cd c d cc+  dd+ 

r ab = ra rb cd rc rd is a vector space.

SOLUTION: We content ourselves with verifying Axiom (iii) (the zero axiom), and Axiom (iv) (the additive inverse axiom).

00 ab Axiom (iii):Let 0 = . Then, for any v = we have: 00 cd

We are again using  to PofR indicate that equality fol- lows from a definition, v + 0 ab + 00 a + 0 b + 0 = ab  v and “PofR” for “Property cd 00 c + 0 d + 0 cd of the Real numbers.”

Axiom (iv): For any given v = ab , we show there exists a vector –v , cd

namely –v = –a –b , such that vv+ – = 0 : –c –d 42 Chapter 2 Vector Spaces

PofR

vv+ – ab + –a –b aa– bb– = 00  0 cd –c –d cc– dd– 00

Generalizing Example 2.3 to accommodate matrices of all dimen- sions, we have:

THEOREM 2.2 Let Mmn denote the set of all mn matrices. MATRIX SPACE For A = aij and B = bij in Mmn , let:

AB+ ==aij + bij aij + bij (The ijth entry of the sum matrix is the sum of the ijth entry in matrix A with the ijth entry in the matrix B.)

For r   and A = aij  Mmn , let:

rA ==raij raij (The ijth entry in the matrix rA is r times the ijth entry in the matrix A.)

The set Mmn with the above operations is a vector space.

PROOF: We content ourselves with verifying Axioms (i) and (vi).

Axiom (i): For every A = aij and B = bij : PofR

AB+ aij + bij aij + bij = bij + aij bij + aij BA+

Axiom (vi): For scalars r and s, and A = aij : PofR

rs+ A rs+ aij rs+ aij = raij + saij

 raij + saij

raij + saij rA + sA

CHECK YOUR UNDERSTANDING 2.3 Verify the associative axiom rsv = rs v for the vector space Answer: See page B-3. Mmn of Theorem 2.2. 2.2 Abstract Vector Spaces 43

POLYNOMIAL SPACES

n n – 1  A function of the form px= anx +++an – 1x a0 , with an  0 is said to be a polynomial function of degree n. For any given In particular: integer n  0 , Pn will represent the set of polynomials of degree less P0x ==a0 a0    than or equal to n We note that the polynomial: n n – 1 anx +++an – 1x  + a1xa0 can be written the other way around: 2 n a0 ++a1xa2x ++ anx and can also be expressed in Sigma-notation form: n The Greek letter  i 2 n (Sigma) is used to denote  aix = a0 ++a1xa2x ++ anx a sum. i = 0

THEOREM 2.3 The set of polynomials Pn of degree less than or equal to n, with operations: POLYNOMIAL SPACES n n n a xi + b xi = a + b xi  i  i  i i i = 0 i = 0 i = 0

n n  raxi = ra xi  i  i i = 0 i = 0 is a vector space.

PROOF: We establish the two distributive axioms, and relegate the remaining six to the exercises. Axiom (v) ruv+ = ru + rv :

n n n n raxi + b xi ra+ b xi ra+ b xi  i  i  i i  i i i = 0 i = 0 i = 0 i = 0 n n n = ra + rb xi  ra xi + rb xi  i i  i  i PofR i = 0 i = 0 i = 0 n n  raxi + rbxi  i  i i = 0 i = 0 44 Chapter 2 Vector Spaces

Axiom (vi) rs+ v = rv + sv : n n PofR n i i i i rs+  aix   rs+ aix =  raix + saix i = 0 i = 0 i = 0 n n n n ra xi + sa xi raxi + saxi  i  i  i  i i = 0 i = 0 i = 0 i = 0

CHECK YOUR UNDERSTANDING 2.4 Referring to Theorem 2.3, verify the commutative axiom: n n n n i i i i  aix +  bix =  bix +  aix Answer: See page B-3. i = 0 i = 0 i = 0 i = 0

FUNCTION SPACES You are probably accustomed of thinking that a function as some sort of dynamic creature that “takes numbers into numbers.” At this point, however, you want to think of a function as being an object, in the same All “objects” in mathemat- way that you see the number 5 as an object. Indeed, the set of all such ics are sets, and functions are no exceptions. The function functions f : X   from a set X (the domain of the function) to the 2 f given by fx = x , for set  of real numbers, can be turned into a vector space: example, is that subset of the plane, typically called the graph of f: THEOREM 2.4 Let FX denote the set of all real-valued 2 f = xx xR functions defined on a non-empty set X: Pictorially: FX= f f : X  R For f and g in FX , and r  R , let fg+ , and rf be given by: fg+ x = fx+ gx (*) and rf x = rf x With respect to these operations, FX is a vector space.

A function f : X   is defined to be equal to a function g : X   , if fx= gx for every xX . 2.2 Abstract Vector Spaces 45

PROOF: We verifying Axioms (i), (iii), (iv) and (v): The fact that FX is Axiom (i) (Commutative Axiom). For fg  FX and xX : closed under addition and scalar multiplication is PofR self-evident. fg+ x  fx+ gx= gx+ fx gf+ x Since fg+ x = gf+ x for every xX , fg+ = gf+ .

Axiom (iii) (Zero Axiom). Let Z : X   be the function given by As you can see, we Zx = 0 for every xX . For any fF and xX : elected to use the letter Z, rather than the symbol 0, PofR for our zero vector. It’s just that an expression fZ+ x fx+ Zx fx+ 0 = fx fx like 0x would strongly suggest that a multiplica- Since fZ+ x = fx for every xX , fZ+ = f . tion by zero is being per- formed, which is not the case. Axiom (iv) (Additive Inverse Axiom). For given f  FX let –f be the function given by –f x = –fx . Then, for any xX : PofR ff+ – x  fx+ –fx = 0  Zx

Since ff+ – x = Zx for every xX , ff+ – = Z .

Axiom (v) [Distributive Axiom (vector)]: For any fg  FX , xX and r  R : PofR rfg+ x rfg+ x rfx+ gx = rf x + rg x  rf x + rg x Since rfg+ x = rf x + rg x for every xX , rfg+ = rf + rg .

CHECK YOUR UNDERSTANDING 2.5

Verify the distributive axiom rs+ v = rv + sv for the func- Answer: See page B-3. tion space of Theorem 2.4.

ADDITIONAL EXAMPLES

As you will see in the next two examples, addition and scalar multi- plication in a vector space can be somewhat “counter-intuitive.” In addition, both the zero vector 0 and the additive inverse vector –v , may appear somewhat strange in a vector space 46 Chapter 2 Vector Spaces

EXAMPLE 2.4 Show that the set + = aa 0 of positive real numbers with addition and scalar multipli- cation given by: ab+ ==ab and raar is a vector space.

SOLUTION: + is certainly closed under both of the above opera- tions. Moreover: PofR Axiom (i): ab+  ab = ba  ba+

Axiom (ii): PofR ab+ + c  ab c = abc abc+ +

Axiom (iii):This axiom professes the existence of a vector 0 such that a + 0 = a for every a  + . It certainly cannot be the number 0, for that number is not even in the set + . Looking for a clue, we ask ourself: 5 +5what = ? which is to say: 5w hat = 5?  since rs+ = rs Yes, 1 is the zero vector in this space: a + 1  a  1 = a

Axiom (iv): Looking for a clue, we ask ourselves “What is the additive inverse of 7:” 7 + what = 1? (remember that 1 is the zero vector) which is to say: 7w hat = 1?

Yes, in this space, –a = --1- : a 1 aa+ –  a--- = 1  (the zero vector) a

Axiom (v): rab+ = ra + rb : PofR rab+  ab r = arbr ar + b r ra + rb You are invited to establish the remaining three axioms of Definition 2.6 in the exercises, thereby establishing the fact that + , with given operations, is a vector space. 2.2 Abstract Vector Spaces 47

EXAMPLE 2.5 Show that the set Vxy=  xy   under addition: xy + x y = xx+  – 1 yy++ 1 and scalar multiplication rxy = rx–1 r +  ry+ r – 1 is a vector space:

SOLUTION: V is certainly closed under both of the above operations. We content ourselves by establishing the zero and inverse axioms, and leave it for you to verify the remaining six axioms in the exercises. Zero Axiom: Does there exist a vector 0 such that v + 0 = v for every v  V ? Don’t be to quick to say “no,” basing you answer on the observation that xy + 00 ==x + 01–  y ++01 x – 1 y + 1  xy But you have no right to assume that if a zero vector exists, then it must be the one you would like it to be! Putting partiality aside, let’s see if we can find a vector 0 = ab such that xy + 0 = xy , for every xy  V : xy + 0 = xy xy + a b = xy  xa+ – 1 ==x  a 1 xa+ – 1 yb++1 = xy   yb++1 ==y  b –1 That’s right, in this vector space, 0 = 11 – , for: xy + 1 –1 ==x + 1 – 1 y ++–1 1 xy Additive Inverse Axiom: Now that we have a zero vector, 0 = 11 – , we can ask if, for any given v = xy , there exists a vector –v = ab such that xy +11ab =  – . There does: xy +11a b =  –  xa+1– 1 == a – x + 2 xa+ – 1 yb++1 = 11 –  yb++1 ==– 1 b – y – 2 So, the additive inverse of xy turns out to be –2x +  –2y – , for: xy + – x + 2 – y – 2 ==x – x + 2 – 1 y – y – 2 + 1 11 –

the zero vector

CHECK YOUR UNDERSTANDING 2.6 Verify that Axioms (iii) and (iv) of Definition 2.6 are satisfied for the set V=  xyz xyz   with imposed addition: Answer: Zero vector: 12– 3 xyz + x y z = xx+  – 1 yy++ 2 zz+  – 3 Inverse of xyz : and scalar multiplication: – x + 2– y + 4 – z + 6 rxyz = rx– r + 1 ry – 2r – 2 rz – 3z + 3 48 Chapter 2 Vector Spaces

EXAMPLE 2.6 Let V denote the upper-half plane: y Vxy=  xy   y  0 V with standard addition and scalar multiplication: xy + x y = xx+  yy+  x rxy = rx ry Is V a vector space?

SOLUTION: No, since V is not closed under scalar multiplication. But we can’t just say this, for our claim has to be established. We have to demonstrate that the scalar product involving some specific element of V and some specific real number ends up being outside of V, and so we shall: 52  V but – 7 5 2 = –35 – 14  V

EXAMPLE 2.7 Let X be the set of two-tuples with operations: xy + x y = xx+  yy+  rxy = rx y Is X a vector space?

SOLUTION: X is certainly closed under both operations. We need not challenge the first four axioms of Definition 2.6, as they involve only the addition operator which coincides with that of Euclidean two- space. The scalar operator, however, is a bit different from that of 2 , and we must therefore determine whether or not Axioms (v) through (viii) are satisfied. What you may want to do is to quickly check to see if they hold for some specific vectors and scalars: Let’s Challenge Axiom (vi), rs+ v = ru + sv , with r = 7 , s = 3 , and u = 42 – : 73+ 42 – ==10 4 – 2 40 – 2 and 7 4 – 2 +28234 – 2 == – +40412 – 2  – Oops! We need go no further, for the above shows that Axiom (vi) does not hold, and can therefore conclude that under the given operations, X is not a vector space.

CHECK YOUR UNDERSTANDING 2.7

THE TRIVIAL VECTOR SPACE: Define addition and scalar multiplication on the set V = 0 and verify Answer: See page B-4. that V is a vector space with respect to those operations. 2.2 Abstract Vector Spaces 49

EXERCISES

Exercises 1-13. Verify that the set S, with given operations, fails to be a vector space. 1.S =  , xy+ = xy– , and rx = rx.

2.Sxy=  xy   , xy + x y = xx+  0 , and rxy = rx ry .

3.Sxy=  xy   , xy + x y = xx yy , and rxy = rx ry .

4.Sxy=  xy   , xy +00x y =  , and rxy = rx ry .

5.Sxy=  xy   , xy +2x y = x + 2x yy+  , and rxy = rx ry .

6.Sxy=  xy   , xy + x y = xx+  yy+  , and rxy = 00 .

7.Sxy=  xy   , xy + x y = xy+  yx+  , and rxy = rx ry .

8.Sxy= 0 xy   , xy0 + xy 0 = xx+  yy+  0 , and rxy0 = rx ry 0 .

ab a b aa+0 ab ra rb 9.SM= 22 , + = , and r = . cd c d 0 dd+  cd rc rd

ab a b ac+  bd+  ab ra rb 10.SM= 22 , + = , and r = . cd c d ca+  db+  cd rc rd

11.Sax= 2 ++bx c , ax2 ++bx c + ax2 ++bxc = aa+  x2 + bb+  x and rax2 ++bx c = ra x2 ++rb xrc.

12.S = 01 ; 00+0= , 11+0= , 01+10==+1 ; and r00= , r11= .

13.S = a   a  0 , ab+ = lnab+ ln , and ra = lnar .

Exercises 14-16. Verify that the set V, with given operations, is a vector space. 14.V = 1 , 11+1= , and r11= .

15.Vxy= 0 xy   , xy0 + xy 0 = xx+  yy+  0 , and rxy0 = rx ry 0 .

16.Vxy=  xy   , xy + x y = xx++ 2 yy+  , rxy = rx + 2r – 2 ry . 50 Chapter 2 Vector Spaces

17. Complete the proof of Theorem 2.2. 18. Complete the proof of Theorem 2.3. 19. Complete the proof of Theorem 2.4. 20. Establish the remaining three axioms for the space of Example 2.4. 21. Establish the remaining six axioms for the space of Example 2.5.

 i 22. A polynomial is an expression of the form px=  aix for which there exists an m such i = 0

that ai = 0 for im . Show that, with respect to the following operations, the set P of all polynomials is a vector space:      i i i i i  aix +  bix = ai + bi x and r  aix =  rai x i = 0 i = 0 i = 0 i = 0 i = 0

PROVE OR GIVE A COUNTEREXAMPLE

23. Let V be a vector space, and let v  V . If rv = 0 , then r = 0 .

24. Let V be a vector space, and let v  V . If rv = 0 and v  0 , then r = 0 .

25. Let V be a vector space, and let v  V . If rv = 0 and r  0 , then v = 0 .

26. Let V be a vector space, and let v  V . If rv = sv , then rs= .

27. Let V be a vector space, and let v  V . If rv = sv and v  0 , then rs= . 2.3 Properties of Vector Spaces 51

2

§3. PROPERTIES OF VECTOR SPACES

For the sake of convenience, we again list the vector space axioms: (i)uv+ = vu+ (Commutative Axiom) (ii)uv+ + w = uvw+ + (Vector Associative Axiom) (iii) There is a vector in V, denoted Addition: by 0 such that v + 0 = v for every vector v in V. (Zero Axiom) (iv) For every vector v in V, there is a vector in V, denoted by –v such that vv+ – = 0 . (Additive Inverse Axiom) (v)ruv+ = ru + rv (Vector Distributive Axiom) Scalar and Addition: { (vi)rs+ v = rv + sv (Scalar Distributive Axiom)

Scalar: (vii)rsv = rs v (Scalar Associative Axiom) { (viii)1vv= (Identity Axiom)

For aesthetic reasons, a set of axioms should be independent, in that no part of an axiom is a consequence of the rest. One should not, for example, replace Axiom (iii) with: There is a vector in V, denoted by 0 such that v + 0 = v and 0 + v = v for every vector v in V. Reason: Axiom (i) already implies that the 0 of Axiom (iii) can be on either side of the v: THEOREM 2.5 Let V be a vector space. (a) For every vector v in V: v +00 ==+ v v (b) For every vector v in V, there exists a vector –v such that: vv+ – ==– v +0v

PROOF: (a): Since v + 0 = v [Axiom (iii)], and vw+ = wv+ [Axiom (1)]: v ==v + 0 0 + v (b): Since vv+ – = 0 [Axiom (v)], and v + 0 = v [Axiom (iii)], and vw+ = wv+ [Axiom (1)]: 0 ===vv+ – –v + v – v + v 52 Chapter 2 Vector Spaces

Our next theorem tells us that there is but one 0 vector in any vector Axiom (iii) asserts the space, and that every vector v has a unique additive inverse –v . While existence of a zero vec- tor, but makes no claim you might have taken these two facts for granted, neither is given to as to its uniqueness, and you free of charge: Axiom (iv) only asserts that every vector has an Let V be any vector space, then: additive inverse (could THEOREM 2.6 it have several?). (a) There is but one vector 0 which satisfies the property that v + 0 = v for every v in V. (b) For any given vector v in V, there is but one vector –v in V such that vv+ – = 0 .

PROOF:

Strategy for (a): Assume that 0 and 0 are any two zeros, and then go on to show 00= .

Let 0 and 0 be such that, for every vector v: v + 0 ==v and v + 0 v (*) v + 0 ==v and (**) v + 0 v

Substituting 0 for v in (*), we have (i): 00+ = 0 .

Substituting 0 for v in (**), we also have (ii): 00+ = 0 . Then: (i) (ii)

000===+ 00+ 0 00= commutativity

Strategy for (b): Assume that a vector v has two additive inverses, –v and v , and then go on to show that –v = v .

Let –v and v be such that:

vv+ – = 0 and v +  v (*) vv+ – ==0 and (**) v +  v 0 Adding v to both sides of (*) we have: v + vv+ – = v + 0 Axioms (ii) and (iii): vv+ + –v = v Axiom (i) and (**): 0 + –v = v Theorem 2.5(a): –v = v –v = v The above proof illustrates the important fact that a mathematical the- ory is based on a set of rules or axioms, on which sit logically derived 2.3 Properties of Vector Spaces 53

results, or theorems. Once established, a1.5 theorem can be used to prove other theorems. At some point, the axioms and theorems kind of blend into each other—they are just facts, with some of them being dic- tated (the axioms), while others are established (the theorems).

CHECK YOUR UNDERSTANDING 2.8 Show that in any vector space: If: vz+ = wz+ Answer: See page B-4. Then: vw=

Two different zeros come into play in the following theorem: The real number 0 that is involved in the scalar prod- uct at the left of the equality, and the vector 0 that appears to the right of the equality.

THEOREM 2.7 For any vector v in a vector space V: 0v = 0

PROOF: At times, as is the case here, a proof almost writes itself, once an appropriate initial step is taken (in this case, to write 0 as 00+ ): PofR Axiom (ii) (Distributive Axiom) 0v ==00+ v 0v + 0v The end is now in sight: just add the additive inverse of the vector 0v to both sides of the equation: 0v = 0v + 0v – 0v + 0v = – 0v + 0v + 0v Axiom (iv) and (ii): 0 = – 0v + 0v + 0v Axiom (iv): 00= + 0v Theorem 2.5: 0 = 0v In words, the above theorem tells us that:

Multiplying any vector by the scalar 0 results in the vector 0.

CHECK YOUR UNDERSTANDING 2.9 Answer: See page B-4. Prove that in every vector space V, r00= for every r   . The above CYU together with Theorem 2.7 tells us that if either r = 0 or v = 0 , then the scalar product rv = 0 . The converse also hods: 54 Chapter 2 Vector Spaces

THEOREM 2.8 In any vector space V: If rv = 0 then r = 0 or v = 0

PROOF: If r = 0 , then surely the statement r = 0 or v = 0 holds, and we are done. If r  0 , then: rv = 0 1 1 ---rv = ---0 r r 1 Axiom (vii) and CYU 2.8: ---  r v = 0 r 1v = 0 Axiom (viii): v = 0

CHECK YOUR UNDERSTANDING 2.10 Establish the following Cancellation Properties: (a) If r  0 and rv = rw , then vw= . Answer: See page B-4. (b) If v  0 and rv = sv , then rs= . Here is what the next theorem is saying: Multiplying any vector by the scalar –1 results in the additive inverse of that vector.

THEOREM 2.9 For any vector v in a vector space V: –1v = –v PROOF: Strategy: Show that if you add –1v to v you end up with the vector 0.

Axiom (viii) – 1v +1v ==–1v +1v –1+ v ==0v 0

Axiom (vi) PofR Theorem 2.7

CHECK YOUR UNDERSTANDING 2.11 Establish the following results, for v in a vector space V, and r   .

Answer: See page B-5. (a) ––v = v (b) –r v = –rv (c) r–v = –rv 2.3 Properties of Vector Spaces 55

SUBTRACTION

We all want to replace the expression vw+ – with vw– . Let’s do it, but officially: DEFINITION 2.7 For vectors v and w in a vector space V, we “SUBTRACTION” define v minus w, denoted by vw– , to be the vector given by: vw– = vw+ –

A definition is the introduction of a new word in the language of mathematics. As such, one must understand all of the words used in its description. This is so in Definition 2.7, where the “new word “vw– ” on the left of the equal sign is described by previ- ously defined words “vw+ – ” on the right of the equal sign. Here are a few results featuring the operation of subtraction. They are very reminiscent of the familiar subtraction operation of real numbers. This should come as no surprise since the real number system is itself a vector space.

THEOREM 2.10 For any vectors v, w, and z in a vector space V, and scalars r, and s in  : (a) vv– = 0 (b) vw+ – z = vwz+ – (c) vw+ – w = v

PROOF: (a) vv– ==vv+ – 0 Definition 2.7 (b) vw+ – z ===vw+ + –z vw+ + – z vwz+ –

Definition 2.7 (c) vw+ – w ===vww+ – v + 0 v

(b) (a)

CHECK YOUR UNDERSTANDING 2.12 (a) Show that for any two vectors v and w: –vw+ = – v – w (b) Use the Principle of Mathematical Induction (see Appendix A) to show that for any n vectors v1v2  vn : Answer: See page B-5.   –v1 ++v2 +vn = – v1 – v2 – – vn 56 Chapter 2 Vector Spaces

We complete this section with a list of results; some of which we proved, some of which appeared in Check Your Understanding boxes, and others which you are invited to establish in the exercises.

THEOREM 2.11 For every v , w , and z in a vector space V, and every rst  : (i) There exists a unique vector 0  V v +00 ==+ v v .

(ii) There exists a unique vector –v  V such that vv+ – ==– v +0v . Cancellation (iii) If vz+ = wz+ , then vw= Properties { (iv) If r  0 and rv = rw , then vw= . (v) If v  0 and rv = sv , then rs= . (vi) 0v = 0 (vii) r00= (viii)rv = 0 if and only if r = 0 or v = 0

(ix)rsv + tw = rs v + rt w (x) –1v = –v

(xi)––v = v (xii) r–v ==–r v –rv (xiii) rvw– = rv – rw (xiv) rs– v = rv – sv (xv) –vw+ = – v – w (xvi) vw– = – w + v

(xvii) vwz– + = vw– – z (xviii) vwz– – = vw– + z 2.3 Properties of Vector Spaces 57

EXERCISES

Exercises 1-8. Prove: 1. Theorem 2.11 (iv): If r  0 and rv = rw , then vw= .

2. Theorem 2.11 (v): If v  0 and rv = sv , then rs= .

3. Theorem 2.11 (ix): rsv + tw = rs v + rt w .

4. Theorem 2.11 (xiii): rvw– = rv – rw .

5. Theorem 2.11 (xiv): rs– v = rv – sv .

6. Theorem 2.11 (xvi): vw– = – w + v .

7. Theorem 2.11 (xvii): vwz– + = vw– – z .

8. Theorem 2.11 (xviii): vwz– – = vw– + z .

9. Show that for any vector v in a vector space V, and any r   : rv = ––rv . 10. Show that for any vector v in a vector space V and any integer n  1 : nv = n – 1 vv+ . 11. Let v, w, and z be any vectors in a vector space V, and let abc   , with a  0 . Show that c b if av + bw = cz , then v = ---z – ---w . a a 12. Let v and w be vectors in a vector space V, with v  0 . Show that if rvw+ = svw+ , then rs= . 13. Let v and w be vectors in a vector space V. Show that if r  1 and rvw+ = v + rw , then vw= . 14. Show that for any v and w in a vector space V, and for any ab   : ab+ vw+ = av ++bv aw +bw 15. Let v and w be non-zero vectors in a vector space V. Show that if rv + sw = 0 , with not both r and s equal to 0, then there exist unique numbers a and b such that v = aw and w = bv . Hint: Show that the condition that not both r and s equal 0 implies that neither is 0. 58 Chapter 2 Vector Spaces

PROVE OR GIVE A COUNTEREXAMPLE

16. All vector spaces contain infinitely many vectors. 17. Any vector space that contains more than one vector must contain an infinite number of vec- tors. 18. For any vector v in a vector space V and any r   : rv = r – 1 vv+ 19. Let V and W be vector spaces. Let VW = vw v  V w  W with operations given by: v1 w1 + v2 w2 ==v1 + v2 w1 + w2 and rvw rv rw Then VW is a vector space.

20. Let V and W be vector spaces. Let VW = vw v  V w  W with operations given by:

v1 w1 + v2 w2 ==v1 – v2 w1 – w2 and rvw rv rw Then VW is a vector space. 2.4 Subspaces 59

2

§4. SUBSPACES

DEFINITION 2.8 A subspace of a vector space V is a non- SUBSPACE empty subset S of V which, together with the imposed operations of addition and scalar multiplication of V, is itself a vector space.

If S is to be a subspace of V, then it is itself a vector space and must therefore be closed under addition and scalar multiplication:

If s1 and s2 are in S, then s1 + s2  S . If s  S and r   , then rs  S . In addition, the eight axioms listed in Definition 2.6, page 40, must also hold for S. Actually, the eight axioms come “free of charge,” once clo- sure is established; for: THEOREM 2.12 A subset S of a vector space V is a subspace of V if and only if: 1. S is nonempty. 2. S is closed under addition. 3. S is closed under scalar multiplication.

PROOF: If S is a subspace of the vector space V, then it is itself a vec- tor space and must therefore satisfy the three stated conditions. We now show that if those three conditions hold, then S is a subspace of V. Of the eight axioms of Definition 2.6, we need not worry about Axi- oms (i), (ii), (v), (vi), (vii), and (viii): (i)vu+ = uv+ (ii) uv+ + w = uvw+ + (v)ruv+ = ru + rv (vi) rs+ u = ru + su (vii)rsu = rs u (viii)1uu= Why not? Because, since they hold for all u, v and w in the given vec- tor space V, then they will surely hold for all u, v and w in the subset S of V. Why do we have to worry about the zero axiom? Because though we know that there is a vector 0 in V such that v + 0 = v for every v  V (and therefore for every s  S ), we have no assurance that 0 sits in S. To see that it does, take any vector v in S (we are given that S is nonempty), and then scalar multiply that vector by 0: 60 Chapter 2 Vector Spaces

Since v  S and S is closed under scalar mutliplication 0v = 0  S Theorem 2.7, page 53 We now complete the proof by showing that if v  S , then its addi- tive inverse –v is also in S: Since v  S and S is closed under scalar mutliplication –1vv= –  S Theorem 2.9, page 54

When challenging the “S is nonempty” condition of Theorem 2.12, one typically looks to see if the zero vector is contained in S. For: If 0  S , then S is certainly nonempty. If 0  S , then S is not a subspace, period.

EXAMPLE 2.8 Verify that S = abc cab= + is a subspace of 3 .

SOLUTION: We show that S satisfies the three conditions of Theorem 2.12. 1. Since the sum of the first two components of 0=  000 is equal to its third component, 0  S (and therefore S is not empty). 2. To show that S is closed under addition, we take two arbitrary ele- The “ticket” to be in S is ments of S: that the third component is equal to the sum of its s1 ==abab+  s2 cdcd+ first two components. and consider their sum:

s1 + s2 ==abab+ + cdcd+ acbd+  +  abcd+++

Since s1 + s2 has the Since the third component of s1 + s2 , abcd+++ , equals the “ticket,” it is in S. sum of its first two components, s1 + s2  S . 3. We now show S is closed under scalar multiplication. For aba + b  S and r   : rabab+ ==ra rb r a+ b ra rb ra+ rb  S

third component is sum of first two 2.4 Subspaces 61

CHECK YOUR UNDERSTANDING 2.13 Show that: a 2a S = a   –0a Answer: See page B-5. is a subspace of the vector space M22 of Example 2.3, page 41.

The following theorem merges two of the properties of Theorem 2.12 into one: THEOREM 2.13 A nonempty subset S of a vector space V is a subspace of V if and only if:

For every s1 s2  S and r   , rs1 + s2  S .

PROOF: For S a nonempty subspace of V, let s1 s2  S and r   . Since S is closed under scalar multiplication: rs1  S . Since S is closed under addition: rs1 + s2  S .

Conversely, assume that for every s1 s2  S and r   :

rs1 + s2  S (*) We show that S is closed under addition and scalar multiplication:

For any given s1 s2  S , simply choose r = 1 , and apply (*):

1s1 + s2 = s1 + s2  S To show that S is also closed under scalar multiplication, we first observe that (*) implies that 0  S : Since S is nonempty, we can choose an s  S . Letting

s1 ==s2 s , and r = –1 in (*) brings us to: – 1s + s  S 0  S Now consider any s  S and r   . Appealing to (*)

with s1 = s and s2 = 0  S , we find that: rs + 0 = rs  S

EXAMPLE 2.9 Let u and v be any two vectors in a vector space V. Show that the set: Sa= u + bv ab   is a subspace of V.

SOLUTION: Since 0 = 0u + 0v  S , 0  S : S is not empty. 62 Chapter 2 Vector Spaces

For au + bv  S and cu + dv  S , and for r   :

s1 s2  S and rR rau+ bv + cu+ dv = rau + cu + rbv + dv = ra+ c u + rb+ d v  S

It is of the form Au+ Bv rs1 + s2  S (has the “ticket”)

CHECK YOUR UNDERSTANDING 2.14 Answer: See Page B-5. Show that Sxyz=  xyz++= 0 is a subspace of 3 .

EXAMPLE 2.10 Let F denote the function space FX of Theorem 2.4, page 44, with domain X =  . Show that: S = f  F f9 = 0 is a sub- The “ticket” needed for a function f to get into S is space of F . that it maps 9 to 0. SOLUTION: As you recall, the zero in F turned out to be the function Z given by: Zx= 0 for every number x. In particular, since Z9 = 0 , Z  S . Hence, S   . If fg  S and r   , then: fg  S and rR since f and g are in S

rfg+ 9 ===rf9 + g9 r00+0 rfg+  S rfg+  S (has the “ticket”)

CHECK YOUR UNDERSTANDING 2.15 Let F denote the function space of Theorem 2.4, page 44 (with domain X =  ). Determine whether or not S = f  F f0 = 9 Answer: Not a subspace. is a subspace of FR .

EXAMPLE 2.11 Show that for any ab   : Sxy=  ax+ by = c is a subspace of 2 if and only if c = 0 .

SOLUTION: We have to establish two results: The “if -condition:” If c = 0 , then S is a subspace of 2 . The “only if-condition:” If c  0 , then S is not a subspace of 2 . 2.4 Subspaces 63

The “only if-condition” is easily dispensed with: If c  0 , then 000=   S , for: a  0 +0b  0 =  c . Since S does not contain 0, S is not a subspace of 2 . To establish the “if-condition,” we assume the c = 0 , and first observe that S is not empty: Since a  0 +0b  0 ==c , 000=   S .

We complete the proof by showing that if x1 y1  x2 y2  S and r   , then:

rx1 y1 + x2 y2 = rx1 + x2 ry1 + y2  S which is to say, that:

arx1 + x2 +0bry1 + y2 = Here goes:

arx1 + x2 + bry1 + y2 = arx1 ++ax2 bry1 +by2

= rax1 + by1 + ax2 + by2

Since x1 y1  x2 y2  S: ==r00+ 0

CHECK YOUR UNDERSTANDING 2.16 In Example 1.10, page 20, we showed that: S = 16a – 2b 4a – 17b11a 11b ab   is the solution set of the homogeneous system of equations: 2x ++03y –54z w =   – 3x ++y 4zw += 0  x ++07y –114z w =  Answer: See page B-5. Show that S is a subspace of 4 .

INTERSECTION AND UNION OF SUBSPACES

Let S and T be subspaces of a vector space V. Is their intersection ST = vv S and v  T [Figure 2.7(a)] necessarily a subspace of V?Is their union ST = vv S or v  T [Figure 2.7(b)] neces- sarily a subspace of V? The answer to the first question is “Yes” (Theo- rem 2.14 below), while the answer to the second question is “No” (Example 2.12 below). 64 Chapter 2 Vector Spaces

V V ST ST

ST = vv S and v  T ST = vv S or v  T S intersect T S union T (a) (b) Figure 2.7

In the exercises you are THEOREM 2.14 If S and T are subspaces of a space V, then so asked to show that the is their intersection: intersection of any num- ber of subspaces of V is ST = vv S and v  T again a subspace of V. PROOF: ST is not empty: Since 0  S and 0  T (why?), 0  ST . uv  ST and rR Let uv  ST and rR . Being in the intersection of S and T, u and v are both in S and in T. Since S and T are subspaces, ruv+  S ruv+  ST and ruv+  T . Being in both S and T, ruv+  ST .

EXAMPLE 2.12 Show that the union of two subspaces S and T of a vector space V: ST = vv S or v  T need not be a subspace of V. SOLUTION: While one can easily show that the set ST is non- empty, and that it is closed under scalar multiplication, one cannot show that it is closed under addition, for it need not be! And how do we show that this is the case? By exhibiting a specific vector space V, along with two specific subspaces S and T, such that their union fails T to be closed under addition. Let’s do it: Let V = 2 , S = x 0 x   , and T = 0 y y   . We 1 leave it for you to verify that both S and T are subspaces of 2 . To see S 1 that ST is not closed under addition, simply note that while 10  ST and 01  ST , 10 + 01 = 11  ST (for 11 is neither in S nor in T).

CHECK YOUR UNDERSTANDING 2.17

PROVE OR GIVE A COUNTEREXAMPLE: If S and T are subsets of a vector space V, and if ST is a subspace Answer: See page B-6. of V, then either S is a subspace of V or T is a subspace of V. 2.4 Subspaces 65

EXERCISES

Exercises 1-6. Determine if the given subset S of the vector space 2 is a subspace of 2 . Jus- tify your answer.

1.S = xy y = 2x 2. S = xy y = 2x + 1

3.S = xy xy+0= 4. S = xy xy

5. S = xy yx= 2 6. S = xy xy = 0

Exercises 7-12. Determine if the given subset S of the vector space 3 is a subspace of 3 . Jus- tify your answer.

7.S = xyz z = 2xy– 8. S = xyz z = 2xy+

9.S = xyz z =12xy– + 10. S = xyz z = 2xy++1

11. S = xyz xyz++= 0 12. S = xyz y = 0

Exercises 13-18. Determine if the given subset S of the matrix space M22 of Example 2.3, page

41 is a subspace of M22 . Justify your answer.

ab ab 13.S = d = 0 14. S = dab= + cd cd

ab ab 15. S = ad==0 c =2b 16. S = ab+2= c – 3d cd cd

ab ab 17.S = abcd+++= 1 18. S = abcd+++= 0 cd cd

Exercises 19-22. Determine if the given subset S of the polynomial space P2x of Theorem 2.3, page 43, is a subspace of P2x . Justify your answer.

19.Sax= 2 ++bx c a = 2 20. Sax= 2 ++bx c b = 0

21.Sax= 2 ++bx c a ++0 b c = 22. Sax= 2 ++bx c b = 2a 66 Chapter 2 Vector Spaces

Exercises 23-26. Determine if the given subset S of the polynomial space P3x of Theorem 2.3, page 43, is a subspace of P3x . Justify your answer.

23.Sax= 2 ++bx c b = 0 24. Sax= 2 ++bx c b = –2a + c 25.Sax= 2 ++bx c a + b = 0 26. Sax= 2 ++bx c b = 2ac = a + 1

Exercises 27-38. Determine if the given subset S of the function space F of Theorem 2.4 (with X =  ), page 44, is a subspace of F . Justify your answer. 27. Sff= 0 = 0 28. Sff= 1 = 0 29. Sff= 1 = 1 30. Sff= 2x = 2fx 31. The subset of even functions: Sf=  fx– = fx 32. The subset of odd functions: Sf=  fx– = –fx 33. The subset of increasing functions: Sf=  a  bfa  fb 34. The subset of decreasing functions: Sf=  a  bfa  fb 35. The subset of bounded functions: Sf=  fx M for every x  , for some M  0 36. ( dependent) Sf=  f is continuous 37. (Calculus dependent) Sf=  f is differentiable 38. (Calculus dependent) Sf=  f is integrable 39. Let V be a vector space. Show that: (a)0 is a subspace of V. (b) V is a subspace of V.

40. (PMI) Establish the following generalization of Theorem 2.14.

(a) If S1S2  Sn are subspaces of a vector space V, then so is their intersection: n  Si = S1 S2  Sn i = 1

(b) If S1S2  Sn  are subspaces of a vector space V, then so is their intersection:   Si = S1 S2  Sn  i = 1

(c) Let A be a nonempty set. If S is a subspace of a vector space V for every   A , then the set  S is also a subspace of V.   A 2.4 Subspaces 67

41. (PMI) Let v1v2  vn be vectors in a vector space V. Show that Sa= 1v1 +++a2v2  anvn ai   1 in is a subspace of V.

42. Let S and T be subspaces of a vector space V. Show that ST+ = st+ s  S and t  T is a subspace of V.

43. Let S and T be subspaces of a vector space V, with ST = 0 . Show that every vector in the subspace ST+ of the previous exercise can be uniquely expressed as a sum of a vector in S with a vector in T.

Suggestion: Show that if st+ = s1 + t1 , then ss= 1 and tt= 1 .

PROVE OR GIVE A COUNTEREXAMPLE

44. If S and T are both subsets of a vector space V, and if neither S nor W is a subspace of V, then ST cannot be a subspace of V. 45. If S and T are both subsets of a vector space V, and if neither S nor W is a subspace of V, then ST cannot be a subspace of V. 46. If S and T are subspaces of a vector space V, then ST+ = ST (see Exercise 43). 47. If S and T are subspaces of a vector space V, then STST  + (see Exercise 43). 48. If S, T, and W are subspaces of a vector space V, then ST + TW is also a subspace of V (see Exercise 43). 49. If S, T, and W are subspaces of a vector space V, then STW + = ST + W (see Exercise 42). 50. If S and T are subspaces of a vector space V with ST = 0 , then ST is a subspace of V. 51. If S is a subspace of a vector space V, and if T is a subspace of S, then T is a subspace of V. 52. If a vector space has two distinct subspaces, then it has infinitely many distinct subspaces. 68 Chapter 2 Vector Spaces

2

§5. LINES AND PLANES

This chapter began with a consideration of vectors in the plane and in three dimensional space, both from a geometrical point of view (directed line segments, or “arrows”), and from an analytical perspec- tive (2-tuples and 3-tuples). The main focus of this section is to deter- mine and classify all of the subspaces of those Euclidean spaces. Additional insight for the material of this section will surface in the fol- lowing chapter on Dimension Theory. This section, in turn, is a nice lead-in to the following chapter, for Euclidean spaces have a dimension component built right into their terminology. It should come as no sur- prise, for example, to find that the Euclidean spaces 2 and 3 will turn out to have dimensions 2 and 3, respectively.

SUBSPACES OF 2

A subset of a vector space V that is neither V or 0 is said to be a proper subset of V. The following theorem serves to characterize the proper subspaces of 2 :

THEOREM 2.15 S is a proper subspace of 2 if and only if St= ab t   for ab  00

PROOF: Let St= ab t   for ab  00 . Appealing to Theorem 2.13, page 61, we first observe that S is not empty [it clearly contains ab ]. Moreover, for t1ab  t2ab  S and r   :

rt1ab + t2ab = rt1 + t2 ab  S .ab At this point we know that S is a non-empty subspace of 2 . Can it be all of 2 ? No, for the set ta tb t   is the line in the plane passing through the origin and the point ab (see margin). 2 tab ab   Conversely, assume that T is a proper subspace of  . Since it is not empty, it contains a nonzero vector ab . Since T is a subspace, every scalar multiple of ab must be in T, which is to say: St= ab t    T . We now show that, in fact, ST= , by demonstrating that if T were to contain any vector cd  S then T For any vectorvR 2 : must be all of 2 (see margin)”: Can we find scalars A, B, such that: Aa+ Bc = x (a,b) .v Aab + Bcd = xy  ? . Ab+ Bd = y

Yes, providing the last row of rref ac does not consist entirely (c,d) bd . of zeros (see the Spanning Theorem, page 18) — and it doesn’t (Exercise 61). 2.5 Lines and Planes 69

While no line in the plane that does not pass through the origin can represent a subspace of 2 (why not?), every line in the plane is paral- lel to one that does pass through the origin—a of a subspace of 2 : THEOREM 2.16 Let L be the line passing through two distinct The vector v is said to be a points Px=  y and Qx=  y in the direction vector for the line, 1 1 2 2 and the vector u is said to be plane [Figure 2.9(a)]. Then, in terms of vectors: a translation vector. Lu= + rv r   (*)

where v ==PQ x2 – x1 y2 – y1 , and

ux= 0 y0 , for x0 y0 any chosen point on L [see Figure 2.8(b)]. [(*) is said to be a vector form repre- sentation for the line L.]

PROOF: Figure 2.8(b) may serve as a geometrical “proof” of the theo- rem; for if you take the vector v which is in the same direction as the line L, and stretch it every which way, then you will get a line that is parallel to L passing through the origin. Adding the vector u to every rv “moves” that line up to L, in a parallel fashion. Those not satisfied with y y u + rv L this geometrical proof L are invited to consider . x0 y0 Exercise 62. P Q r . to c u . e v n o ti a sl n v a r ctor rv x t ion ve direct x (a) (b) Figure 2.8 EXAMPLE 2.13 Find a vector form representation for the line L passing through the points 15 and 23 Note that the set: L = 15 + r12 – r   SOLUTION: We take v ==2135–  – 12 – to be our direction = 1 + r 52– r r   vector, and u = 15 as our translation vector, leading us to the vec- This brings us to the so- tor form: L ==u + rv r   15 + r12 – r   . called parametric repre- sentation of L: x ==1 +52ry – r CHECK YOUR UNDERSTANDING 2.18

(a) Referring to Example 2.13, find the vector form representation for the line L, when v is the vector from 23 to 15 , and u = 23 . (b) Your vector form representation in (a) will look different from Answer: that of Example 2.12: L = 15 + r12 – r   . Appear- (a) 2 +32r – r ances aside, show that your set of two-tuples in (a) and ours of (b) 1 +52r – r r   Example 2.13 are one and the same. 70 Chapter 2 Vector Spaces

SUBSPACES OF 3

We now move up a notch and focus our attention on the Euclidean space 3 . The first order of business is to arrive at vector form repre- sentations for lines and planes in 3 . As it was in 2 , a line in three- space is determined by any two points, P and Q, in 3 . In the exer- cises, you are invited to establish the following result (compare with Theorem 2.16): THEOREM 2.17 Let L be the line passing through two distinct points Px= y z and Qx= y z in One cannot envision a line in 1 1 1 2 2 2 n for n  3 . We can, three-space. Then, in vector form: however define, in vector L = u + rv r   form, the line passing through: where v is the direction vector: Pa= 1a2  an v ==PQ x2 – x1 y2 – y1 z2 – z1 and Qb= 1b2  bn in n to be the set: and ux= 0y0 z0 is a translation vector, L = u + rv r   with x0y0 z0 any chosen point on L. where: v = b – a  b – a z . L 1 1 n n rv and: Q x y z . 0 0 0 u = a a  a 1 2 n P . . u translation vector v = PQ direction vector y

x

EXAMPLE 2.14 Find a vector form representation for the line L passing through the points 20– 3 and 142 .

SOLUTION: We take the vector from 20– 3 to 142 The line can also be v ==12402–  –  – – 3 –145 as our direction vector. expressed in parametric form (see margin note Selecting the translation vector u = 20– 3 , we have: of Example 2.13): x ===2 –4ry rz 35+ r L ==u + rv r   20– 3 + r–145 r  

CHECK YOUR UNDERSTANDING 2.19

Consider the line L = 135 + r21– 1 r   . Find two points on L, both different from 135 , and proceed as in Example 2.14 to obtain another representation for the set L. Verify that “your Answer: See page B-6 set” is equal to the set 135 + r21– 1 r   . 2.5 Lines and Planes 71

In Theorem 2.16, we noted that the line L in 2 which passes through the origin and the point ab (distinct from the origin) can be expressed in vector form: Lr= v r   , where vab=  . A sim- ilar result, which you are invited to establish in the exercises, holds for a plane in 3 :

THEOREM 2.18 Let P be a plane in 3 passing through the ori-

gin [Figure 2.9(a)]. Let x1y1 z1 and x2y2 z2 be any two points on P which do not both lie on a common line passing through the origin. Then, in vector form: Pr= u + sv rs  

where ux= 1y1 z1 and vx= 2y2 z2 .

z

.B P . A x y z . u 0 0 0 . .C v . w u . y y 0 v x x (a) (b) Figure 2.9 In this general setting, we have (compare with Theorem 2.17): THEOREM 2.19 Let P be the plane passing through three non- colinear (not lying on a common line) points

Ax= 1y1 z1 , Bx= 2y2 z2 and

Cx= 3y3 z3 [Figure 2.9(b)]. Then, in vec- u and v are said to be direction vectors, and w tor form: is said to be a translation P = w ++ru sv rRs    vector. where u ==AB x2 – x1y2 – y1 z2 – z1

v ==AC x3 – x1y3 – y1 z3 – z1 ,

and wx= 0y0 z0 for x0y0 z0 any cho- sen point on P. 72 Chapter 2 Vector Spaces

EXAMPLE 2.15 Find a vector form representation for the Plane P passing through the points 32–2 , 25– 3 and 41– 3 .

SOLUTION: We elect 32–2 to play the role of both w and of

x1y1 z1 in Theorem 2.19; with: u ==25– 3 – 32– 2 –175– and P consists of all points v ==41– 3 – 32– 2 13– 5 xyz such that: Then: x = 3 – r + s P = w ++ru sv rs   y = –72 ++r 3s z = 25–5r – s = 32–2++r–751– s13– 5 rs   The above is said to be a parametric representa- = 3 – r + s –72 ++r 3s 25–5r – s rs   tion of the plane (with parameters r and s). CHECK YOUR UNDERSTANDING 2.20

(a) Repeat our solution to Example 2.15, but this time letting 41– 3 play the role of w, instead of 32–2 ; all of the rest remaining as before. (b) Your set representation for P in (a) will look different from that of Example 2.15. Appearances aside, show that your set in (a) Answer: See page B-6 and ours of Example 2.15 are one and the same.

We previously showed that the only proper subspaces of 2 are the lines passing through the origin (as sets). The following theorem, a proof of which is relegated to the exercises, settles the subspace issue in 3 :

THEOREM 2.20 S is a proper subspace of 3 if and only if S is the set of points on a line that passes through the origin, or the set of points on a plane that passes through the origin. 2.5 Lines and Planes 73

EXERCISES

Exercises 1-4. Determine a vector form representation for the line in 2 passing through the ori- gin and the given point. 1.15 2.–24 3.51 – 4. –22 –

Exercises 5-8. Determine a vector form representation for the line in 2 passing through the two given points. 5.13  24 – 6.35  55 7.35  37 8. 24  –52 –

Exercises 9-12. Determine two different vector form representations for the line in 2 passing through the two given points, and then proceed to show that the set of points associated with the two vector forms are one and the same. 9.13  24 – 10.35  55 11.35  37 12. 24  –52 –

Exercises 13-20. Determine a vector form representation for the line in 2 that passes through the point 37 and is parallel1 to the line of: 13. Exercise 1 14. Exercise 2 15. Exercise 3 16. Exercise 4 17. Exercise 5 18. Exercise 6 19. Exercise 7 20. Exercise 8

Exercises 21-28. Determine a vector form representation for the line in 2 that passes through the point 37 and is perpendicular2 to the line of: 21. Exercise 1 22. Exercise 2 23. Exercise 3 24. Exercise 4 25. Exercise 5 26. Exercise 6 27. Exercise 7 28. Exercise 8

Exercises 29-32. Determine a vector form representation for the line in 3 passing through the origin and the given point. 29.245 30.100 31.–240 32. –441 –  –

Exercises 33-36. Determine a vector form representation for the line in 3 passing through the two given points. 33.245  311 34. 012  102

35.34– 1 210 36. 511 –  –  222

1. Two lines are parallel they have equal slopes, or if both lines are vertical. 2. Two lines are if and only if the slope of one is the negative reciprocal of the slope of the other, or if one line is horizontal and the other is vertical. 74 Chapter 2 Vector Spaces

Exercises 37-40. Determine two different vector form representations for the line in 3 passing through the two given points, and then proceed to show that the set of points associated with the two vector forms are one and the same. 37.245  311 38. 012  102

39.34– 1 210 40. 511 –  –  222

Exercises 41-48. Determine a vector form representation for the line in 3 that passes through the point 12– 1 and is parallel to the line of: 41. Exercise 29 42. Exercise 30 43. Exercise 31 44. Exercise 32 45. Exercise 33 46. Exercise 34 47. Exercise 35 48. Exercise 36 Exercises 49-52. Determine a vector form representation for the plane passing through the origin and the two given points. 49.132  214 50. 314  200

51.200  020 52. –321 –  –  241 Exercises 53-56. Determine a vector form representation for the plane passing through the three given points. 53.341 215 11– 1 54. 210 –  211 –  123

55.24– 3515 41– 1 56. –112 –  231 –  121 Exercises 57-60. Determine two different vector form representations for the plane passing through the two given points, and then proceed to show that the set of points associated with the two vector forms are one and the same. 57.341 215 11– 1 58. 210 –  211 –  123

59.24– 3515 41– 1 60. –112 –  231 –  121 61. Complete the proof of Theorem 2.15. Incidentally:

You can also show directly that if v1 = ab and v2 = cd are such that v2  rv1 for 2 any r   , then for any v   there exist r1 r2   such that v = r1v1 + r2v2 . 62. Prove Theorem 2.16. 63. Prove Theorem 2.17. 64. Prove Theorem 2.18. 65. Prove Theorem 2.19. 66. Prove Theorem 2.20. Chapter Summary 75

CHAPTER SUMMARY

EUCLIDEAN VECTOR The set of n-tuples, with addition and scalar multiplication given by: SPACE n v1v2  vn + w1w2  wn = v1 + w1 v2 + w2 vn + wn

rv1v2  vn = rv1rv2  rvn ABSTRACT VECTOR A nonempty set V, closed under addition and scalar multiplication, satisfying the following eight properties: SPACE (i)uv+ = vu+ (ii)uv+ + w = uvw+ + (iii) There is a vector in V, denoted by 0 such that v + 0 = v for every vector v in V. (iv) For every vector v in V, there is a vector –v in V, such that vv+ – = 0 . (v) ruv+ = ru + rv (vi) rs+ u = ru + su (vii) rsu = rs u (viii) 1uu= SUBTRACTION vw– = vw+ – Uniqueness of 0 and –v There is but one vector 0 which satisfies the property that v + 0 = v for every v in V. For any given vector v in V, there is but one vector –v in V such that vv+ – = 0 . Cancellation Properties If vz+ = wz+ , then vw= If r  0 and rv = rw , then vw= . If v  0 and rv = sv , then rs= . Zero Properties 0v = 0 r00= rv = 0 if and only if r = 0 or v = 0 Inverse Properties –1v = –v ––v = v r–v ==–r v –rv 76 Chapter 2 Vector Spaces

SUBSPACE A nonempty subset S of V which is itself a vector space under the vector addition and scalar multiplication operations of the space V. Closure says it all A nonempty subset S of a vector space V is a subspace of V if and only if it is closed under addition and under scalar multipli- cation. A one liner A nonempty subset S of a vector space V is a subspace of V if

and only if for every s1 s2  S and r   , rs1 + s2  S . Intersection of subspaces The intersection of any collection of subspaces in a vector space V is again a subspace of V. PROPER SUBSPACES A subspace of a vector space V distinct from the trivial sub- space 0 and V itself is said to be a proper subspace of V.

Vector form of lines Let L be the line in 2 passing through the origin, and let ab be any point on L distinct from the origin, then, in vector form: Lr= v r   , where v = ab

Let L be the line in the plane passing through Px= 1 y1 and Qx= 2 y2 . Then, in vector form, L = u + rv r  

where v = PQ , and ux= 0 y0 , with x0 y0 any chosen point on L. Let L be the line in 3 passing through two distinct points Px= 1y1 z1 and Qx= 2y2 z2 . Then L = u + rv rR where v is the direction vector

v ==PQ x2 – x1 y2 – y1 z2 – z1 and u = x0y0 z0 is a translation vector, with x0y0 z0 any chosen point on L.

Vector form of planes Let P be a plane in 3 passing through the origin. Let

x1y1 z1 and x2y2 z2 be any two points on P which do not both lie on a common line passing through the origin. Then, in

vector form: Pr= u + sv rs   where ux= 1y1 z1

and vx= 2y2 z2 . Let P be the plane passing through three non-colinear points

P1 = x1y1 z1 , P2 = x2y2 z2 and P3 = x3y3 z3 .

Then P = w ++ru sv rRs    where u = P1P2 ,

v = P1P3 , and w = x0y0 z0 is any chosen point on P 3.1 Spanning Set 77

3 CHAPTER 3 BASES AND DIMENSION It is easy to see that every vector xyz in 3 can uniquely be expressed in terms of the three vectors 100 , 010 , and 001 . For example: 52– 9 = 5100 ++2010 –9 001 In this chapter, we consider an arbitrary vector space V to see if we can

find a set of vectors v1v2  vn in V, called a basis for V, such that every vector v in V can be uniquely expressed in the form:

v = c1v1 +++c2v2  cnvn for some scalars c1c2  cn . As you will see, while many such sets of vectors v1v2  vn may exist, the number of vectors in those sets will always be the same. For example, if you find a basis for a vec- tor space V that contains 5 vectors, v1v2 v3 v4 v5 , and someone else finds another basis, that other basis must also consist of 5 vectors. That being the case, we will then be in a position to say that the vector space V has dimension 5.

§1. SPANNING SETS

Using vector addition and scalar multiplication in 2 , one can build the vector v = 914 from the vectors v1 = 34 and v2 = 12 : 914 = 234 + 312 and we say that v = 914 is a of v1 = 34 and v2 = 12 . In general:

DEFINITION 3.1 A vector v in a vector space V is said to be a linear combination of vectors v v  v in LINEAR 1 2 n COMBINATION V, if there exist scalars c1c2  cn such that: v = c1v1 +++c2v2  cnvn

EXAMPLE 3.1 Determine whether or not the vector 0224 in 3 is a linear combination of the vectors 138 and 254 . 78 Chapter 3 Bases and Dimension

SOLUTION: We are to see if we can find scalars a and b such that: 0224 = a138 + b254

or: a + 2b 3a + 5b 8a + 4b = 0224 Equating coefficients, we come to the following system of three equations in two unknowns: a b a b a +02b =  120 104  augS S: 3a +25b =  352 rref 01– 2  8a +244b =  8424 000

Solution: a ==4 b –2 Conclusion: 0224 is a linear combination of the vectors 138 and 254 : 0224 = 4138 + –2 254

Some Added Insight on Example 3.1

The set of all linear combinations of z 138 138 and 254 is the plane in 3 containing those vectors. As such, were we to pick an arbitrary point in 3 , say 254 243 , then there would be little chance that it would lie in that plane, and would therefore not be a linear combina- tion of the two given vectors: x a b a b a +22b =  Note that, except for the last  augS 12 2 10 0 3a +45b = rref column, this augmented S:  35 4 01 0 matrix is the same as that of  Example 3.1. 8a +34b =  84 3 00 1

a +00b =   0ab+0=   0a +10b =  No solution!

CHECK YOUR UNDERSTANDING 3.1

Determine if the given vector is a linear combination of the vectors 138 and 254 .

Answer: (a) No. (b) Yes. (a) –238 –  (b) –248 –  3.1 Spanning Set 79

THEOREM 3.1 The set of linear combinations of the set of vec- tors v1v2  vn in a vector space V is a sub- space of V.

PROOF: Let Sc= 1v1 +++c2v2  cnvn . We first observe that since 0 = 0v1 +++0v2  0vn  S , S is not empty. Moreover, for u = a1v1 +++a2v2  anvn and w = b1v1 +++b2v2  bnvn in uw  S and r   S, and r   :

ruw+ = ra1v1 +++a2v2  anvn + b1v1 +++b2v2  bnvn

= ra1 + b1 v1 +++ra2 + b2 v2  ran + bn vn  S ruw+  S

a linear combination of the vis

DEFINITION 3.2 The set of linear combinations of a set of vec- tors v v  v in a vector space V is SPANNING 1 2 n called the subspace of V spanned by v1v2  vn and will be denoted by Spanv1v2  vn .

In the event that V = Spanv1v2  vn , v1v2  vn is said to span V.

EXAMPLE 3.2 Determine if the vectors 2x2 + 3x – 1 , x – 5 , 2 and x – 1 span the space P2x of polynomi- als of degree less than or equal to two.

2 SOLUTION: We consider an arbitrary vector ax ++bx c in P2x to see whether or not we can find scalars rsand t such that: r2x2 + 3x – 1 ++sx– 5 tx2 – 1 = ax2 ++bx c Expanding and combining the left side of the above equation brings us to: 2rt+ x2 ++3rs+ xr–5– s – t = ax2 ++bx c

Equating coefficients we come to the following system of three equa- tions, in the unknowns r, s, and t (the a, b, and c are not variables, they are the fixed coefficients of the polynomial ax2 ++bx c ): System S was solved directly in Example 1.7, page 16. In that example, we labeled the variables x, y, and z, instead of r, s, and t. 80 Chapter 3 Bases and Dimension

2rt+ = a   S: 3rs+ = b   –5r – s – t = c  Can we find values for r, s, and t such that the above three equations hold? The Spanning Theorem (page 18), tells us that the answer is “yes” if and only if rref coefS does not contain a row consisting entirely of zeros. Let’s see:

Since the rref-matrix does not contain a row of zeros, system S has a solution for all values of a, b, and c, and we conclude that the vectors 2 2 2x + 3x – 1 , x – 5 , and x – 1 span the space P2x . While the above does not tell you how to build ax2 ++bx c from 2x2 + 3x – 1 , x – 5 , and x2 – 1 , it does tell you that it can be done, for

each and every polynomial in P2x . If you want to see how to build any particular polynomial, say the polynomial 4x2 + 10x – 6 , then that’s not a problem, for the task reduces to finding scalars r, s, and t, such that: 4x2 + 10x – 6 = r2x2 + 3x – 1 ++sx– 5 tx2 – 1 Or: 4x2 +210x – 6 = rt+ x2 ++3rs+ xr–5– s – t Equating coefficients, we have: r s t r s t 2rt+4=  2014 1003  augS rref S: 3rs+10=  31010 0101  –5r – s –6t = –  –51 –1–6– 001– 2 Conclusion: 4x2 +3210x – 6 = x2 + 3x – 1 ++1x – 5 –2 x2 – 1 3.1 Spanning Set 81

CHECK YOUR UNDERSTANDING 3.2 (a) Use the Spanning Theorem (page 18) to show that the vectors 12 10 01 04  span M22 . 34 10 01 20

(b) Express –51 as a linear combination of the above four vectors. Answer. See pare B-7. 113

EXAMPLE 3.3 Do the vectors 2101 , 12– 20 , 231– 1 and 124– 4 span 4 ?

SOLUTION: Let abcd be an arbitrary vector in 4 . Are there scalars x, y, z, and w such that: abcd = x2101 +++y12– 20 z231– 1 w124– 4 which is to say: does the following system of equations have a solu- tion for all values of a, b, c, and d? 2xy++2zw + = a   x –32y ++z 2w = b  S:  0x +++2yz4w = c   x + 0yz–4– w = d  The Spanning Theorem tells us that it does not, as rref coefS con- tains a row consisting entirely of zeros:

While the above argument establishes the fact that the vectors 2101 , 12– 20 , 231– 1 and 124– 4 do not span 4 , it does not define for us the subspace of 4 spanned by those four vectors; bringing us to:

EXAMPLE 3.4 Determine the subspace of 4 spanned by 2101 , 12– 20, 231– 1 and 124– 4. 82 Chapter 3 Bases and Dimension

SOLUTION: We are to find the set of all vectors abcd for which there exist scalars x, y, z, and w such that: abcd = x2101 +++y12– 20 z231– 1 w124– 4 which again boils down to a consideration of a system of equations: 2xy++2zw + = a   x –32y ++z 2w = b  S:  0x +++2yz4w = c   x + 0yz–4– w = d  for to say that abcd is in the space spanned by the four given vec- tors is to say that system S has a solution for those numbers a, b, c, and d. That system appeared earlier in Example 1.8, page 17 where it was noted that the given system of equation has a solution if and only if: – 10a ++7b 12c +13d = 0 Bringing us to a representation for the space spanned by the four given vectors: Span 2101 12– 20 231– 1 124– 4 = abcd –710a ++b 12c +13d = 0

CHECK YOUR UNDERSTANDING 3.3 Determine the space spanned by the vectors 215 , 12– 2 , Answer: 3 3 abc b = 5c – 12a 051 . If it is not all of  , exhibit a vector in  that is not in See Page B-8. Span 215 12– 2 051 . The following example differs from the previous two in that it is not confined to a specific concrete vector space, like 4 .

EXAMPLE 3.5 Let the set of vectors v1v2 v3 and w1 w2 be such that wi  Spanv1v2 v3 for 1 i 2 . Show that Spanw1 w2  Spanv1v2 v3 .

SOLUTION: In any non-routine problem, it is important that you chart out, in one way or another, what is given and that which is to be established:

We are given that the vectors w1 and w2 are in the space spanned by the three vectors v1v2 and v3 , and are to show that for any given scalars a and b, the vector aw1 + bw2 can be written as a linear combination of the vectors v1v2 and v3 ; which is to say that we can find scalars c1c2 c3 such that:

aw1 + bw2 = c1v1 ++c2v2 c3v3 3.1 Spanning Set 83

From the given information, we know that there exist scalars

a1a2 a3 and b1b2 b3 such that:

w1 ==a1v1 ++a2v2 a3v3 and w2 b1v1 ++b2v2 b3v3 Consequently:

aw1 + bw2 = aa1v1 ++a2v2 a3v3 + bb1v1 ++b2v2 b3v3

= aa1v1 +++++aa2v2 aa3v3 bb1v1 bb2v2 bb3v3

= aa1 + bb1 v1 ++aa2 + bb2 v2 aa3 + bb3 v3

= c1v1 ++c2v2 c3v3

The following theorem generalizes the situation of Example 3.5.

THEOREM 3.2 Let the set of vectors v1v2  vn and w1w2  wm be such that

wi  Spanv1v2  vn for 1 im.Then:

Spanw1w2  wm  Spanv1v2  vn .

PROOF: If w  Spanw1w2  wm then, for some scalars ci :

w = c1w1 +++c2w2  cmwm We also know that for each 1 im , there exist scalars ai1ai2  ain , such that:

wi = ai1v1 +++ai2v2  ainvn Consequently: w = c1w1 ++ cmwm

= c1a11v1 ++ a1nvn ++cmam1v1 ++ amnvn

= c1a11 ++ cmam1 v1 ++ c1a1n ++ cmamn vn

Since w can be expressed as a linear combination of v1v2  vn , w  Spanw1w2  wm . Consequently:

Spanw1w2  wm  Spanv1v2  vn

CHECK YOUR UNDERSTANDING 3.4

Prove that for any three vectors v1v2 v3 in a vector space V:

Spanv1v2 v3 = Spanv1v1 + v2 v1 ++v2 v3 Answer: See Page B-8. 84 Chapter 3 Bases and Dimension

EXERCISES

Exercises 1-4. Determine whether or not the given vector in 3 is a linear combination of the vectors –121 and 203 . 1.234 2.14– 2 3.–560 4. –110 11

Exercises 5-8. Determine whether or not the given vector in M22 is a linear combination of the

vectors 1201 12. 03 23 30

5.59 6.49 7.11 6 8. 611 73 73 67 76

Exercises 9-12. Determine whether or not the given vector in P3 is a linear combination of the vectors x3 +21x2 –3 and x + 1 .

9. 2x3 ++x 1 10.3x2 –2x – 11.2x3 + 4x2 –6x – 12. 2x3 ++6x2 x – 6

Exercises 13-16. Show that the given vector in the function space Fx of Theorem 2.4, page 44, is in the space spanned by the vectors sinxx cos sin2x cos2x tan2x cot2x .

13. cos2x 14.sin--- – x 15.sin--- – x 16. sin--- – x 2 7 7

Exercises 17-21. Determine if the given vectors span 4 . If not, find a specific vector in 4 which is not contained in that subspace. 17. 1000 0200 0030 0004 18. 1000 1100 1110 1111 19. 1111 1001 0110 6444 20. 2131 1213 3112 1123 21. –1234 –  –  – –3112 1213 7032–  –

Exercises 22-25. Determine if the given vectors span P3 . If not, find a specific vector in P3 which is not contained in that subspace.

22.1 x2x2 3x3 23. 1 +1x2 + xx – + x2 1 ++xx2 +x3 24.x 1 + x 1 ++xx2 1 ++xx2 +x3 25. 1xx2 + 1 x3 – x2 26. For what values of c do the vectors 213 435 00c span 3 ?

2 2 27. For what values of c do the vectors c 2x 2x  2x ++2xc span P2 ? 3.1 Spanning Set 85

28. For what values of a and b do the vectors ab and –ba span 2 ?

29. Show that for any given set of vectors v1v2  vn , vi  Spanv1v2  vn for every 1 in.

30. Let the set of vectors v1v2  vn and w1w2  wm be such that wi  Spanv1v2  vn for 1 im and vi  Spanw1w2  wm for 1 in . Prove that Spanw1w2  wm = Spanv1v2  vn .

31. Show that if v1v2 v3 span a vector space V, then for any vector v4 the vectors v1v2 v3 v4 also span V. 32. Show that a nonempty subset S of as vector space V is a subspace of V if and only if Spanv1v2  vn  S for every v1v2  vn  S . 33. Let P denote the vector space of all polynomials of Exercise 22, page 50. Show that no finite set of vectors in P spans P . 34. Let S be a subset of a vector space V. Prove that SpanS is the intersections of all subspaces of V which contain the set S.

PROVE OR GIVE A COUNTEREXAMPLE

35. If the vectors u and v span V, then so do the vectors u and uv+ . 36. If the vectors u and v span V, then so do the vectors u and uv– . 37. If the vectors u and v are contained in the space spanned by the vectors w and z, then Spanuv = Spanwz .

38. If Spanv1v2 v3 = V , and if vi  Spanw1 w2 for 1 i 3 , then Spanw1 w2 = V .

39. If S1 and S2 are finite sets of vectors in a vector space V, then: SpanS1  S2 = SpanS1  SpanS2 .

40. If S1 and S2 are finite sets of vectors in a vector space V, then: SpanS1  S2  SpanS1  SpanS2

41. If S1 and S2 are finite sets of vectors in a vector space V, then: SpanS1  S2 = SpanS1  SpanS2 .

42. If S1 and S2 are subspaces of a vector space V, then:

SpanS1  S2 = SpanS1  SpanS2 . 86 Chapter 3 Bases and Dimension

3

§2. LINEAR INDEPENDENCE

The subspace of 3 spanned by the vectors 100 , 010 , 250 is built from those vectors. But there is a kind of inefficiency with the three building blocks 100 010 250 , in that whatever can be built from those three vectors can be built with just two of them; as one of them, say the vector 250 , can itself be con- structed from the other two: 250 = 2100 + 5010 The following concept, as you will soon see, addresses the above “inefficiency issue:” Note that if each ci = 0 , then surely DEFINITION 3.3 A set of vectors v1v2  vn are linearly c v +++c v  c v 1 1 2 2 n n Linearly independent if: will equal zero. Independent c v +c v + +c v == 0 each c 0 To say that v1v2  vn is 1 1 2 2 n n i linearly independent, is to A collection of vectors that are not linearly say that no other linear combination of the vectors independent is said to be linearly dependent. equals 0. EXAMPLE 3.6 Is x2 x2 + xx 2 ++x 1 a linearly indepen- dent set in the vector space P2x ? SOLUTION: To resolve the issue we consider the equation: ax2 ++bx2 + x cx2 ++x 1 = 0x2 ++0x 0 Equating coefficients, brings us to the following system of equations: abc++= 0  bc+0=   c = 0  Working from the bottom up, we see that c ===0 b 0 a 0 is the only solution for the above system of equations, and therefore con- clude that the given set of vectors is linearly independent. EXAMPLE 3.7 212 120 10– 2 13–0 Is   340 –011 –213 26– 1

a linearly independent set in M23 ?

SOLUTION: If:

a 212 ++b 120 c 10– 2 +d 13–0= 000 340 –011 –213 26– 1 000 3.2 Linear Independence 87

Then: Most graphing calculators do not have the capability of 2abcd+++ a +22b – 3d a – 2c = 000 “rref-ing” a “tall matrix.” 3ab–3–2c +4d a ++2c 6d bcd+ – 000 But you can always add enough zero columns to Leading us to the homogeneous system: arrive at a square matrix: a b c d a b c d 2abcd+++= 0  2111 1 001 a +02b – 3d =   120– 3 0 1 02– 2a –02c =  20–0 2 S:  coef(S) rref 001 1 3ab–3–2c +0d =  31–3–2 0000  4a ++2c 6d = 0 4026 0000 bcd+0– =  011– 1 0000 Since rref[coef(S)] has a free variable, the system has nontrivial solu- tions, and we therefore conclude that the given set of vectors is lin- early dependent.

CHECK YOUR UNDERSTANDING 3.5

2 2 Answer: Yes. Is x 2x + x x – 3 a linearly independent set in P2 ?

EXAMPLE 3.8 Let v1v2 v3 be a linearly independent set of vectors in a vector space V. Show that v1v2 + v1 v3 – v2 is also linearly independent.

SOLUTION: We start with: av1 ++bv2 + v1 cv3 – v2 = 0 and go on to show that abc===0 :

av1 ++bv2 + v1 cv3 – v2 = 0

regroup: ab+ v1 ++bc– v2 cv3 ==ab+0

Since v1v2 v3 is linearly independent: ab+0=   bc–0=   a ===0 b 0 c 0  c = 0 

CHECK YOUR UNDERSTANDING 3.6

Let v1v2 v3 v4 be a linearly independent set of vectors in a vec-

tor space V. Show that v1v1 + v2 v1 ++v2 v3 v1 +++v2 v3 v4 Answer: See page B-8. is also a linearly independent set. 88 Chapter 3 Bases and Dimension

Here is a useful consequence of the Linear Independence Theorem of page 22:

THEOREM 3.3 n A set of vectors v1v2  vm in  is lin- LINEAR INDEPENDENCE early independent if and only if the row-reduced- THEOREM FOR n echelon form of the nm matrix with ith col-

umn the (vertical) n-tuple vi has m leading ones.

PROOF: For 1 im , let vi = a1ia2i  ani . To challenge lin- ear independence of those m vectors, we consider the vector equation: c1v1 ++ cmvm = 0:

c1a11a21  an1 +++0c2a12a22  an2 cma1ma2m  anm =  0   c1a11 + c2a12 ++cma1m  c1an1 + c2an2 ++cmanm = 0 0 Equating coefficients brings us to the following homogeneous sys- tem of n equations in m unknowns: c1 c2 cm a c + a c +  + a c =0  a a  a 11 1 12 2 1m m  11 12 1n H: a21c1 + a22c2 +  + a2ncn =0  coefS a a  a S:  . 21 .22 .2n        . . .  . . . an1c1 + an2c2 +  + anmcm =0  an1 an2  anm Applying the Linearly Independent Theorem of page 22, we conclude that the above homogeneous system of equations has only the trivial

solution c1 ==c2  =cm =0 if and only if rref coefS has m leading ones. EXAMPLE 3.9 Determine if: 13– 21 2132 1314 4727 is a linearly independent set of vectors in 4 . SOLUTION: 1214 1 00 1 3137 rref 0 1 0 1 –3122 001 1 1247 000 0 Since the above matrix does not have 4 leading ones, the given set of vectors is not linearly independent.

CHECK YOUR UNDERSTANDING 3.7

Use Theorem 3.3 to show that there cannot exist a set of five linearly Answer: See page B-9. independent vectors in 4 . 3.2 Linear Independence 89

THEOREM 3.4 Any set of vectors containing the zero vector 0 is linearly dependent.

PROOF: Let 0 v1v2  vm be a subset of a vector space V. Since: 10 ++++0v2 0v3  0vm = 0  0 0 v1v2  vm is not linearly independent. The above theorem tells us that 0 is a linearly dependent set in any vector space V. As for the rest:

THEOREM 3.5 A finite set of vectors, distinct from 0 , is linearly independent if and only if no vec- tor in the set can be expressed as a linear combination of the rest.

PROOF: To establish the fact that linear independence implies that no CONTRAPOSITIVE PROOF vector in the set can be expressed as a linear combination of the rest, Let P and Q be two proposi- we show that if some vector in the set can be expressed as a linear tions. You can prove that: combination of the rest, then the set is not linearly independent (see PQ margin). Here goes: by showing that: Assume that one of the vectors in v1v2  vn can be Not-Q  Not-P expressed as a linear combination of the rest. Since one can (After all if Not-Q implies Not-P, always reorder the given vectors we may assume, without loss of then you certainly cannot have P without having Q: think about it) generality, that v1 is a linear combination of the rest:  v1 = a2v2 +++a3v3 anvn Leading us to:  1v1 – a2v2 – a3v3 – –0anvn = Since the coefficient of v1 is not zero, we conclude that v1v2  vn is linearly dependent. To establish the converse we again turn to a contrapositive proof: If v v  v is not linearly independent, then we can find sca- v1v2  vn NOT linearly 1 2 n independent lars c1c2  cn , not all zero, such that:  c1v1 +++c2v2 cnvn = 0 Assuming, without loss of generality, that c1  0 we find that we can express v1 as a linear combination of the rest: c2 c3 cn some v can be expressed as a  i v1 = – -----v2 – -----v3 – – -----vn linear combination of the rest c1 c1 c1 The next theorem tells us that whatever can be built from a collection of linearly independent vectors, can only be built in one way:

In the exercises you are THEOREM 3.6 Let v1v2  vn be a linearly independent set. invited to establish the If a v ++ a v = b v ++ b v , then: converse of this theo- 1 1 n n 1 1 n n rem. ai = bi , for 1 in . 90 Chapter 3 Bases and Dimension

PROOF:   n n a1v1 +++a2v2 anvn = b1v1 +++b2v2 bnvn

a v = b v a v +++a v  a v – b v +++b v  b v = 0  i i  i i 1 1 2 2 n n 1 1 2 2 n n i = 1 i = 1  a1 – b1 v1 +++a2 – b2 v2 an – bn vn = 0

by linear Ind.: a1 – b1 ==0 a2 – b2 = 0 an – bn 0

ai = bi a1 = b1 a2 = b2 an = bn

CHECK YOUR UNDERSTANDING 3.8

Show that the vectors 213 502 11311 are linearly dependent in 3 , and that 8412 is in the space spanned by those vectors. Express 8412 as a linear combination of those Answer: See page B-9. vectors in two distinct ways. A linearly independent set of vectors S in a vector space V may not be able to accommodate additional vectors without losing its indepen- dence. This is not the case if V  SpanS :

THEOREM 3.7 If v1v2  vn is a linearly independent set EXPANSION of vectors, and if v  Spanv v  v , then THEOREM 1 2 n v1v2  vn v is also linearly independent.

PROOF: Let S = v1v2  vn be linearly independent, and let v  SpanS . We show that v1v2  vn v is linearly indepen- dent by showing that no vector in v1v2  vn v can be expressed as a linear combination of the rest. To begin with, we observe that v cannot be expressed as a linear com- bination of the rest, as v  SpanS . Suppose then that some other vector in v1v2  vn v , say for definiteness the vector v1 , can be expressed as a linear combination of the rest:

v1 = c2v2 ++++c3v3  cnvn cv

Since S = v1v2  vn is linearly independent, c  0 (why?). But then: 1 c c c v = ---v – ----2-v – ----3-v –  – ----n-v c 1 c 2 c 3 c n again contradicting the given condition that v  SpanS .

CHECK YOUR UNDERSTANDING 3.9

Find a linearly independent set of four vectors in P3 which includes Answer: See page B-9. the two vectors x3 + x and – 7 . 3.2 Linear Independence 91

EXERCISES

Exercises 1-6. Determine if the given set of vectors is a linearly independent set in 3 . 1.215  4110 2. 000  –64– 5

5 3. 215 4110 4110 –  4. 32 –  --- –64– –5 120 2 5.134 25– 1 010 220 6. 101 201 102 123

Exercises 7-12. Determine if the given set of vectors is a linearly independent set in M22 .

12 21 12 11 21 11– 7. 8.  34 34 43 34 23 –65 –

12 21 12 21 12 21 12 12 9. 10.  34 34 43 25 34 34 43 25

12 21 12 12 11 11 22 01 –22 00 11.  12.   34 34 43 25 11 11 21 20 22 01

Exercises 13-17. Determine whether the given set of vectors is a linearly independent set in P3 .

13. x + 1 x2 + x + 1 x3

14. x + 1 x2 + 13 x – 517

15. 3x3 3x2 ++3x 33 x3 +++3x2 6x 63 x + 3

16. 2x3 3x2 ++3x 33 x3 +++3x2 6x 63 x + 3

17. 2x3 3x2 ++3x 33 x3 +++3x2 6x 63 x + 5

Exercises 18-27. Determine if the given set of vectors in the function space FX of Theorem 2.4, page 44, is linearly independent.

18.5 sinx  F 19. sin2x cos2x  F

20.5sin 2x cos2x  F 21. sin2x cos2x cos2x  F

22. x2 sinx  F 23. sin2x cos2x 2  F 24.ex e–x  F 25. ex e2x  F

26. 1ex + e–x ex – e–x  F 27.lnx lnx2  F+ , where+ denotes the set of positive numbers. 92 Chapter 3 Bases and Dimension

28. For what real numbers a is 11a 1a 1 a11 a linearly dependent set in 3 ?

29. For what real numbers a is 2xa++1 x + 2 a linearly dependent set in P1x ?

x 1 x2 1 x2 x 30. For what real numbers a is ax2 – --- – ---– ----- + ax – --- – ----- – --- + a a linearly dependent 2 2 2 2 2 2 set in P2x ? 31. Find a value of a for which cosxa+  sinx is a linearly dependent set in the function space F ? 32. Find a value of a for which sinxa+  sinxx cos is a linearly dependent set in the func- tion space F ? 33. Let v be any nonzero vector in a vector space V. Prove that v is a linearly independent set.

34. Prove that every nonempty subset of a linearly independent set v1v2  vn is again lin- early independent.

35. Prove that if v1v2  vn is a linearly dependent set in a vector space V, then so is the set v1v2  vn  w1w2  wm for any set of vectors w1w2  wm in V. 36. Establish the converse of Theorem 3.6.

37. Let S = v1v2  vm be a set of vectors in a space V. Show that if there exists any vector v  SpanS which can be uniquely expressed as a linear combination of the vectors in S then S is linearly independent. 38. Show that 10  01 is a linearly independent set in the vector space of Example 2.5, page 47.

39. Let S = u1u2  un and T = v1v2  vm be linearly independent sets of vectors in a vector space V with SpanS  SpanT = 0 . Prove that ST is also a linearly independent set. 3.2 Linear Independence 93

PROVE OR GIVE A COUNTEREXAMPLE

40. If uv is a linearly dependent set, then u = rv for some scalar r.

41. If uvw is a linearly dependent set, then u = rv + sw for some scalars r and s.

42. If uvw is a linearly independent set of vectors in a vector space V, then uu+ vuvw++ is also linearly independent.

43. If uvw is a linearly independent set of vectors in a vector space V, then uvvwwu–  –  – is also linearly independent.

44. For any three nonzero distinct vectors uvw in a vector space V, uvvwwu–  –  – is linearly dependent.

45. If uvw is a linearly independent set of vectors in a vector space V, and if a   then auvw is also linearly independent.

46. If uvw is a linearly independent set of vectors in a vector space V, and if a is any non- zero number, then auu+ av uv++aw is also linearly independent.

47. If S = u1u2  un and T = v1v2  vm are linearly independent sets of vectors in a vector space V, then ST is also a linearly independent set.

48. If uv and vw are linearly independent sets of vectors in a vector space V, then uvw is also a linearly independent set. 94 Chapter 3 Bases and Dimension

3

§3. BASES

So far we have considered sets of vectors S = v1v2  vn in a space V that satisfy one of two properties: (1) S spans V: Every vector in V can be built from those in S. (2) S is linearly independent: No vector in S can be built from the rest. In a way, (1) and (2) are tugging in numerically opposite directions: From the spanning point of view: The more vectors in S the better. From the linear independence point of view: The fewer vectors in S the better. Sometimes, the set of vectors S is not too big, nor too small — it’s just right:

DEFINITION 3.4 A set of vectors  = v1v2  vn in a BASIS vector space V is said to be a basis for V if: (1)  spans V and: (2)  is linearly independent.

In the exercises you are asked to verify the fact that: 2 S2 ==e1 e2 10  01 is a basis for  ,

S3 ==e1e2 e3 100 010 001 is a basis for 3 , and that, in general: S = e e  e n 1 2 n , where each entry in the n-tuple ei is 0 with the exception of the ith entry which equals 1, is a n STANDARD BASES IN  basis for n , called the of n .

EXAMPLE 3.10 Show that 130 204 012 is a basis for 3 .

SOLUTION: Appealing to Definition 3.4, we challenge the given set of vectors on two fronts: spanning, and linear independence. Spanning. For any given abc  3 , we are to determine if there exist scalars x, y, and z such that: x130 ++y204 z012 = abc 3.3 Bases 95

Expanding the left side, and equating coefficients, we come to the following 33 system of equations:  x + 2y = a 1 2 0 1 00  coefS rref S: 3xz+ = b  3 0 1 0 1 0  4y + 2z = c  0 4 2 001 the three given vectors

If you take the time to Figure 3.1 solve the system directly, Applying the Spanning Theorem (page 18) we conclude that you will find that: 130 204 012 spans 3 . 2a + 2bc– x = ------8 Linear independence. A consequence of: y = ------6a – 2b + c 1 2 0 1 00 16 rref 3 0 1 0 1 0 z = –2------6a ++b 3c- 8 0 4 2 001

the three given vectors and the Linear Independence Theorem for n (page 88). Since 130 204 012 is a linearly independent set which spans 3 , it is a basis for 3 .

EXAMPLE 3.11 Determine if the following set of matrices con- stitute a basis for the vector space M22 : 21 11 –03 04   30 22 –56 – 15

SOLUTION: The problem boils down to a consideration of the coeffi- cient matrix of a system of equations. What system? Well, if you take the spanning approach, then you will be looking at the vector equation:

21 11 –03 04 ab x ++y z +w = (*) 30 22 –56 – 15 cd

to see if it can be solved for any given matrix ab . cd On the other hand, if you take the linear independent approach, then you will consider the vector equation:

21 11 –03 04 00 x ++y z +w = (**) 30 22 –56 – 15 00 to see if it has only the trivial solution. 96 Chapter 3 Bases and Dimension

In either case, by equating entries on both sides of the vector equa- tions, you arrive at systems of equations: Form (*): From (**): 2xy++–03z w = a 2xy++–03z w = 0 xy++0z +4w = b xy++0z +4w = 0

3x ++2y – 6z w = c 3x ++02y – 6z w = 0x ++2y –55z w = d 0x ++02y –55z w = Yes, the two systems differ to the right of the equal signs, but both share a common coefficient matrix, which “twice” reveals the fact

that the four given vectors do not constitute a basis for M22 . Once by the Spanning Theorem of page 18, and then again by the Linear Independence Theorem of page 22:

only 3 leading ones and 4 variables 21–0 3 1 00– 1 so the four vectors are not linearly independent 1104 rref 0 1 05 32–1 6 001 1 02–5 5 0000 contains a zero row so the four vectors do not span

CHECK YOUR UNDERSTANDING 3.10 Determine if the following set of matrices constitute a basis for the

matrix space M22 : 21 11 –03 04   Answer: It is a basis. 30 22 –55 – 15

In Example 3.10 we found that the vector space 3 has a basis con- sisting of three vectors, and every fiber in your body is probably sug- gesting that each and every basis for 3 must also consist of 3 vectors. Those fibers are correct. Indeed we will prove that if a vector space V has a basis consisting of n vectors, then every basis for V must again contain n vectors. Our proof will make use of the following fun- damental result:

In words: There cannot THEOREM 3.8 If w w  w spans V, and if be more lineally indepen- 1 2 m SPAN dent vectors than the v1v2  vn is a linearly independent sub- number of vectors in any VERSES spanning set. INDEPENDENT set of V, then nm .

PROOF: Assume that nm . (We will show that this assumption contradicts the given con- dition that v1v2  vn is a linearly independent set). 3.3 Bases 97

We begin by expressing each vi as a linear combination of the vec- tors in the spanning set w1w2  wm :

v1 = c11w1 +++c12w2  c1nwm

v2 = c21w1 +++c22w2  c2nwm . . . . (*) ...... vn = cn1w1 +++cn2w2  cnmwm Now Consider the following homogeneous system of n linear equa- tions in m unknowns, with nm : c x +++c x  c x = 0 Note that the coefficients of the 11 1 21 2 n1 m ith equation coincide with those c x +++c x  c x = 0 12 1 22 2 n2 m (**) th    of the i column of (*). 

c1nx1 +++c2nx2  cnmxm = 0 The Fundamental Theorem of Homogeneous Systems of Equations (page 20), tells us that (**) has a nontrivial solution: r1r2  rn (not all of the ris are zero). We now show that r1v1 +++r2v2  rnvn equals 0, contradicting the assumption that v1v2  vn is a linearly independent set (for some ri  0 ): From (*)

r1v1 +++r2v2  rnvn = r1c11w1 +++c12w2  c1nwm + r c w +++c w  c w . 2 21 1 22 2 2n m . . + rncn1w1 +++cn2w2  cmnwm

Regrouping: = c11r1 +++c21r2  cn1rn w1 + c r +++c r  c r w . 12 1 22 2 n2 n 2 Since r r  r is a . 1 2 n . solution of (**): + c1nr1 +++c2nr2  cmnrn wm c11r1 +++c21r2  cn1rn = 0 . see margin: = 0w1 + 0w2 ++ 0wn = 0 c1nr1 +++c2nr2  cmnrn = 0 We are now in a position to show that all bases of a vector space must contain the same number of elements:

THEOREM 3.9 If w1w2  wn and v1v2  vm are bases for a vector space V, then nm= . 98 Chapter 3 Bases and Dimension

PROOF: Since w1w2  wn is a basis, it spans V, and since v1v2  vm is a basis, it is a linearly independent set. Applying Theorem 3.8, we have nm .

One more time: Since v1v2  vm is a basis, it spans V, and since w1w2  wn is a basis, it is linearly independent. Applying The- orem 3.8, we also have mn . Since nm and mn , nm= .

DEFINITION 3.5 A vector space with basis v1v2  vn is DIMENSION said to be finite dimensional of dimension n. The symbol dimV will be used to denote the dimension of the vector space V.

The trivial vector space V = 0 of CYU 2.6, page 48, has no basis In the exercises you are (see Theorem 3.4, page 89). Nonetheless, it is said to be of dimension asked to show that the zero. We also point out that a vector space that is not finite dimensional polynomial space of Exer- is said to be infinite dimensional. cise 22, page 50, is an infinite dimensional space. CHECK YOUR UNDERSTANDING 3.11

Prove that S = v1v2 vn is a basis for a vector space V if and only if every vector in V can uniquely be expressed as a linear com- Answer: See page B-10. bination of the vectors in S. Let V be a space of dimension n. The following theorem says that any So, if the number of vectors spanning set of n vectors in V must also be linearly independent, and equals the dimension of the that any linearly independent set of n vectors must also span V: space, then to show that those vectors is a basis you THEOREM 3.10 Let V be a vector space of dimension n  0 , and do not have to establish both let S be a set of n vectors in V. Then: linear independence and spanning, for either implies S spans V if and only if S is linearly independent. the other. PROOF: We first show that the assumption that S = v1v2  vn spans V and is not linearly independent leads to a contradiction: If S is not linearly independent, then some vector in S is a linear combination of the remaining elements in S (Theorem 3.5, page 89). Assume, without loss of generality, that it is the vector vn : vn = a1v1 +++a2v2  an – 1vn – 1 Let vV . Since S spans V:

v = c1v1 + c2v2 ++ cnvn

= c1v1 +++c2v2  cn – 1vn – 1 +cna1v1 +++a2v2  an – 1vn – 1

= c1 + cna1 v1 +++c2 + cna2 v2  cn – 1 + cnan – 1 vn – 1 3.3 Bases 99

The above argument shows that the set v1v2  vn – 1 spans V — a contradiction, since a space of dimension n, which neces- sarily contains a basis of n elements, and therefore a linearly inde- pendent set of n elements, cannot contain a spanning set of n – 1 vectors (Theorem 3.8).

We now show that the assumption that v1v2  vn is a linearly independent set which does not span V will also lead to a contradiction:

Let vn + 1  Spanv1v2  vn . The Expansion Theorem of

page 90 tells us that v1v2  vn vn + 1 is still a linearly inde- pendent set. This leads to a contradiction since a space of dimen- sion n, which necessarily contains a spanning set of n elements, cannot contain a linearly independent set of n + 1 vectors (Theo- rem 3.8). The following result is essentially a restatement of Theorem 3.10. It underlines the fact that you can show that a set of n vectors in an n- dimensional vector space is a basis by EITHER showing that they span the space, OR by showing that it is a linearly independent set—you DON’T have to do both:

THEOREM 3.11 Let v1v2  vn be a set of n vectors in a vector space V of dimension n. The following are equivalent:

(i)v1v2  vn is a basis.

(ii)v1v2  vn spans V.

(iii)v1v2  vn is linearly independent. The cycle: (i) PROOF: We can easily show that i ii iii i : i  ii : By Definition 3.4. (iii) (ii) ii  iii : By Theorem 3.10. iii  i : By Theorem 3.10 and Definition 3.4. insures that the validity of any of the three prop- CHECK YOUR UNDERSTANDING 3.12 ositions implies that of Prove that the vector space of Example 2.5, page 47: the other two. Vxy=  xy   , under the operations: xy + x y = xx+  – 1 yy++ 1 rxy = rx–1 r +  ry+ r – 1 Answer: See page B-10. has dimension 2. 100 Chapter 3 Bases and Dimension

STRETCHING OR SHRINKING TO A BASIS A basis has to be both a spanning and a linearly independent set of vectors. Help is on the way for any set of vectors that falls short on either of those two counts: THEOREM 3.12 Let V be a nonzero space of dimension n. Expansion Theorem: (a) Any linearly independent set of vectors in V can be extended to a basis for V. Reduction Theorem: (b) Any spanning set of vectors in V can be reduced to a basis for V.

PROOF: (a) Let v1v2  vm be a linearly independent set of vec- Procedure: Keep adding tors in V. If it spans V, then it is already a basis for V and we are done. vectors, while maintaining linear independence, till If not, then take any vector vm + 1  Spanv1v2  vm and add it you end up with n linearly to the set: v1v2  vm vm + 1 . This brings us to a larger set of independent vectors. vectors which, by the Expansion Theorem of page 90, is again lin- early independent. Continue this process until the set contains n lin-

early independent vectors: v1v2  vm  vn , and then apply Theorem 3.11 to conclude that it is a basis for V.

(b) Let v1v2  vm be a spanning set of vectors in V. Since V Procedure: Keep throwing contains a linearly independent set of n elements (any basis will do), vectors away, while main- taining spanning, till you mn (Theorem 3.8). If mn= then we are done (Theorem 3.11). If end up with n spanning vectors. mn , then the spanning set v1v2  vm cannot be linearly independent; for if it were, then it would be a basis, and all bases have

n elements. Find a vector in v1v2  vm that is a linear combina- tion of the rest and remove it from that set to arrive at a smaller span- ning set. Continue this “tossing out” process until you arrive at a spanning set consisting of n elements—a basis for V (Theorem 3.11).

EXAMPLE 3.12 (a) Expand the linearly independent set: L = 1234 –4101 3312 to a basis for 4 . (b) Reduce the spanning set: Sx= + 12 x2 – 32 x – 3 x2 + 4 x2 –3x – to a basis for P2 . SOLUTION: (a) We are given that L is linearly independent, and need to comple- ment it with an additional 4-tuple, while maintaining linear indepen- dence. From earlier discussions, you know that if you randomly pick a 4-tuple, then the probability that it will be in Span(L) is nil. Let’s go with 2 – 517 :  = 1234 –4101 3312 2 – 517 3.3 Bases 101

We, of course, have to demonstrate that the “gods were not against us,” and do so via the Linearly Independence Theorem for n of page 88:

14–32 1 000 rref 0 1 00 213– 5 3011 001 0 4127 0001 The above shows that  is a linearly independent set. Since it con- sists of 4 vectors, and since 4 is of dimension 4 [Exercise 1(c)],  is a basis for 4 . (b) It is easy to see that 1xx2 is a basis for the vector space P2x , and that consequently P2 is of dimension 3. Since Sx= + 12 x2 – 32 x – 3 x2 + 4 x2 –3x – contains five vectors, we have to throw two of them away in such a manner so as to

end up with three vectors that still span P2 ; or, equivalently, with three linearly independent vectors. We leave it for you to verify that x + 12 x2 – 32 x – 3 is a linearly independent set. As such, it

must be a basis for the three dimensional space P2 .

CHECK YOUR UNDERSTANDING 3.13

21 22 (a) Expand the linearly independent set L =  to a 12 11 basis for M22 (b) Reduce the set: S = 312 –  –93– 6 12– 2 –54– 6 624 –  Answer: See page B-11. to a basis for SpanS . Dose S span 3 ?

The next example reveals a systematic approach that can be used to reduce a set of vectors S in n to a basis for SpanS : EXAMPLE 3.13 Find a subset of: S = 12– 1302 540 66– 1 which is a basis for SpanS .

SOLUTION: To determine which of the four vectors can be discarded, we challenge their linear independence, and turn to the equation: a12– 1 +++b302 c540 d66– 1 = 000 (*) Equating coefficients leads us to the following homogeneous system of equations: 102 Chapter 3 Bases and Dimension

a b c d a b c d  a +++3b 5c 6d = 0 1356 1 0 2 3 coefS rref S: 2a +++0b 4c 6d = 0 2046 0 1 1 1  –2a ++b 0cd– = 0 –2011 – 0 0 0 0

equivalent systems Note that these are the four given vectors, but written in column form a ++2c 3d = 0  bcd+0+ =  Figure 3.2 Let’s agree to call a vectors in S = 12– 1302 540 66– 1 that occupies the same column-location in coef(S) as that of a lead- ing-one-column in the rref-matrix, a leading-one vector [12– 1 and 302 are leading-one vectors]. We now proceed to show that those leading-one vectors constitute a basis for SpanS . Figure 3.2 tells us that system S will be satisfied for any a, b, c, and a ++2c 3d = 0 d for which:  (**) bcd+0+ =  a +02 =  Note that c and d are the free Setting c = 1 and d = 0 in (**) leads us to  , with variables in rref[coef (s)] b +01 =  solution a = –2 and b = –1 . Substituting these values in (*) we have: –212– 1 – 1302 ++1540 066– 1 = 000

or: 540 = 212– 1 + 1302

Conclusion: the non-leading-one vector (5, 4, 0) can be expressed as a linear combination of the two leading-one vectors. . Setting d = 1 and c = 0 in (**) we find that a = –3 and b = –1 , bringing us to: –1312– 1 –0302 ++540 166– 1 = 000

or: 66– 1 = 312– 1 + 1302 Conclusion: the non-leading-one vector (6, 6, -1) can be expressed as a linear combination of the two leading-one vectors. Since 540 and 66– 1 are linear combinations of 12– 1 302 : Span12– 1302 540 66– 1 = Span12– 1 302 3.3 Bases 103

Covering up the two non-leading-one columns in the development of Figure 3.2: a b c d a b c d a +++3b 5c 6d = 0 1356 1 0 2 3 coefS rref S: 2a +++0b 4c 6d = 0 2046 0 1 1 1  –2a ++b 0cd– = 0 –2011 – 0 0 0 0

a ++2c 3d = 0  bcd+0+ =  we see that the only solution of: a12– 1 + b302 = 000 is the trivial solution, and that the vectors 12– 1 302 are therefore linearly independent.

SUMMARIZING To find a basis for the space spanned by the vectors v1 = 12– 1, v2 = 302 , v3 = 540 , and v4 = 66– 1, we constructed the 34 matrix 1356 A = 2046 with columns the given four-tuples. We –2011 – then showed that the leading-one vectors, namely v1 = 12– 1 and v2 = 302 , constituted a basis for the space spanned by the four given vectors.

We state, without proof, a generalization of the above observation:

THEOREM 3.13 n Let v1v2  vm be vectors in  . If A is the nm matrix whose ith column is the n-tuple vi , then the set consisting of those vectors vi , where the ith column of rrefA contains a leading one, is a basis for the space spanned by v1v2  vm.

CHECK YOUR UNDERSTANDING 3.14

Use the above theorem to address the problem of CYU 3.13(b). Answer: See page B-11. 104 Chapter 3 Bases and Dimension

‘ EXERCISES

1. (a) Prove that S ==e  e 10  01 is a basis for 2 . Express –3 --5- as a linear 2 1 2 2 combination of the vectors in S2 . 3 (b) Prove that S3 ==e1e2 e3 100 010 001 is a basis for  . Express

320 as a linear combination of the vectors in S3 .

n (c) Prove that Sn = e1e2  en is a basis for  . 2. (a) Prove that  = 205 0110 120 is a basis for 3 , and express 70– 5 as a linear combinations of the vectors in  . (b) Show that S = 205 0110 120 111 is not a basis for 3 , and find two different representations for the vector 70– 5 as a linear combination of the vec- tors in S.

2 2 2 3. (a) Prove that  = 2x + 3x – x x – 5 is a basis for P2 , and express x + 3x – 1 as a lin- ear combinations of the vectors in  . 2 (b) Show that S = 2x 3x 5 x – 4 is not a basis for P2 , and find two different representa- tions for the vector x2 + 3x – 1 as a linear combination of the vectors in S.

Exercises 4-7. Determine if the given set of vectors is a basis for 3 . 4.215 4110 123 5. 000 –12– 5 215 –  –

6. 215 4110 4310 7. 2 e e 2 e2 

8. (a) Prove that the matrix space M22 has dimension 4.

(b) Prove that the matrix space Mmn has dimension mn .

Exercises 9-12. Determine if the given set of vectors is a basis for M22 .

12 23 34 41 –21 12– 12 12 9. 10.  34 41 12 23 34 34 –43 34–

 12 12 10 02 2 --1- 13– 11 --1- –1 21– 11.  12. 2 3 30 04 34 34 13 1 31– ------4 01 --- 0 2 6 3.3 Bases 105

Exercises 13-15. Determine if the given set of vectors is a basis for M32 .

01 10 03 30 00 10 11 10 01 00 00 11 13.3003 01 10 01 00 14. 1100 00 11 11 00 01 10 01 10 03 00 11 00 00 00 00 11

12 22– 32 –21 –21 – 15. 31– –12 – 12– 31 31 –32 – 34– –13 – –32 23–

3 2 16. (a) Prove that the polynomial space P3 = a3x +++a2x a1xa0 ai  R 0 i 3 is of dimension 4. (b) Prove that the polynomial space: n n – 1 Pn = anx +++an – 1x  + a1xa0 ai  R 0 in is of dimension n + 1 .

Exercises 17-20. Determine if the given set of vectors is a basis for P3 .

17. x3 +21 x3 –23x2 +  x + 14 x3 –9x2 ++x 7

18. x3 +21 x3 –23x2 +  x + 14 x3 –9x2 ++x 8

19. 2x3 ++3x2 x – 1 –9x3 –2x2 ++x 2 x3 –2x2 ++x 23 –2x3 ++x2 –1x

20. 2x3 ++3x2 x – 1 – x3 –2x2 ++x 2 x3 –2x2 ++x 23 –2x3 ++x2 –1x

Exercises 21-24. Extend the given linearly independent set of vectors to a basis for 4 . 21.1341 –1201 1122 22. 121 – 21212  1234

23.213– 1 1302 24. 2211

Exercises 25-28. Extend the given linearly independent set of vectors to a basis for P4 .

25.x4 + 1x3 +32x2 x3 +9x – 5  26. 3x –32x2 –32 x3 –32 x4 – 2

27.x4 ++x3 x2x3 ++x2 x x2 ++x 1 28. 2x4 ++++x3 x2 x 1 x4 – 5 106 Chapter 3 Bases and Dimension

Exercises 29-30. Does the given set of vectors span M22 ? If so, reduce the set to obtain a basis

for M22 .

13 20 33 01 21 12 29.  –21 11– 01 12 01 34

–01 50 10 10 00 00 01 30.  20 –010 01 02 00 01– 10

Exercises 31-32. Does the given set of vectors span P3 ? If so, reduce the set to obtain a basis

for P3 .

31. 2x 3x2 6x2 + 4xx3 +31 x2 + x x3 +++3x2 x 1

32. 2x3 +4x – 1 x3 –642x2 –  x2 + 25 x3 + 6x2 – 2 x3 + x – 1 x3 – 2x2 –2x3 – x2

Exercises 33-37. Use Theorem 3.13 to find a subset of the given set of vectors S in n which is a basis for SpanS . If necessary expand that basis for SpanS to a basis for the corresponding Euclidean space.

33.S = 214 –132 51– 6 428

34. S = 113 –132 158 32– 1

35. S = 113 –132 158 32– 1

36. S = 2130 0442 –2312 1158 3288

37. S = 13132 24142 11010 11202 22111 12345 

Exercises 38-39. Find a subset of the given set of vectors S which is a basis for SpanS . If nec- essary, expand that basis for SpanS to a basis for P4 .

38. S = 2x4 ++x3 3x +1 – x4 – x3 + x2 x3 –22x3 ++x2 3x –22x4 –2x3 + x2 – 1

39. S = 5 – x3 – xx 4 ++++x3 x2 x 12 x4 – 2x2 2x4 + 2x2

Exercises 40-41. Find a basis for the space spanned by the given set of vectors in the function space vectors F of Theorem 2.4, page 44.

40.5 sinxx cos  sin2x cos2x sin5 + x 41. sinxx cos  sin2x cos2x cos2x sin2x

42. Show that abc a ++2bc= 0 is a subspace of 3 , and then find a basis for that sub- space. 3.3 Bases 107

43. Show that abcd ab+ = cd– is a subspace of 4 , and then find a basis for that sub- space.

2 44. Show that ax ++bx c abc= – is a subspace of P2 , and then find a basis for that sub- space.

3 2 45. Show that ax +++bx cx d abc== abd++ is a subspace of P3 , and then find a basis for that subspace.

46. Find all values of c for which 110 c210 0c 1 is a basis for 3 .

c 1 12c 01 c 0 47. Find all values of c for which  is a basis for M22 . 00 10 –0c 0 c

48. Find a basis for the vector space of Example 2.5, page 47.

49. Suppose v1v2 v3 is a basis for a vector space V. For what values of a and b is av1 bv2 ab– v3 a basis for V? 50. Let S is a subspace of V with dimS ==dimV n . Prove that SV= .

51. Suppose that v1v2  vs is a linearly independent set of vectors in a space V of dimen- sion n, and that w1w2 wt spans V. Prove that snt . 52. A set of vectors S in a finite dimensional vector space V is said to be a maximal linearly independent set if it is not a proper subset of any linearly independent set. Prove that a set of vectors is a basis for V if and only if it is a maximal linearly independent set. 53. A set of vectors S in a finite dimensional vector space V is said to be a minimal spanning set if no proper subset of S spans V. Prove that a set of vectors is a basis for V if and only if it is a minimal spanning set. 54. Let H and K be finite dimensional subspace of a vector space V with HK = 0 , and let HK+ = hk+ h  H and k  K . Prove that dimHK+ = dimH + dimK . (Note: you were asked to show that HK+ is a subspace of V in Exercise 42, page 67.) 55. Let H and K be finite dimensional subspace of a vector space V, and let HK+ = hk+ h  H and k  K . Prove that:

dimHK+ = dimH + dimK – dimHK Suggestion: Start with a basis for HK and extend it to bases for H and K.

56. Prove that the polynomial space P of Exercise 22, page 50, is not finite dimensional by show- ing that it does not have a finite base. 57. (Calculus dependent) Show that SpP=  p0 = 0 is a subspace of the polynomial space P of Exercise 22, page 50. Find a basis for S. 58. Prove that a vector space V is infinite dimensional (not finite dimensional) if and only if for any positive integer n, there exists a set of n linearly independent vectors in V. 108 Chapter 3 Bases and Dimension

PROVE OR GIVE A COUNTEREXAMPLE

59. If v1v2 v3 is a basis for a vector space V, and if c1 , c2 , and c3 are nonzero scalars, then

c1v1c2v2 c3v3 is also a basis for V.

60. If v1v2  vn – 1 is a linearly independent set of vectors in a space V of dimension n, and

if vn  v1v2  vn – 1 , then v1v2  vn – 1 vn is a basis for V.

61. If v1v2  vn – 1 is a linearly independent set of vectors in a space V of dimension n, and

if vn  Spanv1v2  vn – 1 , then v1v2  vn – 1 vn is a basis for V.

62. If v1v2  vn vn + 1 is a spanning set of vectors in a space V of dimension n, then

v1v2  vn is a basis for V.

63. If v1v2  vn vn + 1 is a spanning set of vectors in a space V of dimension n, and if

vn + 1  Spanv1v2  vn , then v1v2  vn is a basis for V.

64. If v1v2 v3 is a basis for a vector space V, then v1v1 + v2 v1 ++v2 v3 is also a basis for V.

65. It is possible to have a basis for the polynomial space P2x which consists entirely of poly- nomials of degree 2.

66. Let v1v2  vn + 1 be a spanning set for a space V of dimension n satisfying the property

that v1 ++v2  +vn + 1 = 0 . If you delete any vector from the set v1v2  vn + 1 , then the resulting set of n vectors will be a basis for V. Note: In set notation, an element cannot be repeated. In particular, no two of the vs in

v1v2  vn + 1 are the same. 67. If V is a space of dimension n, then V contains a subspace of dimension m for every integer 0 mn. Chapter Summary 109

CHAPTER SUMMARY

LINEAR A vector v in a vector space V is said to be a linear combination COMBINATION of vectors v1v2  vn in V, if there exists scalars c1c2  cn such that:

v = c1v1 +++c2v2  cnvn

SPANNING The set of linear combinations of v1v2  vn is a subspace of V. It is said to be the subspace of V spanned by those vectors,

and is denoted by Spanv1v2  vn :

Spanv1v2  vn = c1v1 +++c2v2  cnvn

If V = Spanv1v2  vn , then v1v2  vn is said to span the vector space V. If every vector in a set S is 1 Let the set of vectors v v  v and w w  w be contained in the space 1 2 n 1 2 m such that wi  Spanv1v2  vn for 1 im . Then: spanned by another set S2 , then SpanS1 is a subset Spanw1w2  wm  Spanv1v2  vn of SpanS2 .

If v1v2  vn  Spanw1w2  wm and w1w2  wm  Spanv1v2  vn

then: Spanv1v2  vn = Spanw1w2  wm

LINEARLY INDEPENDENT The vectors v1v2  vn are linearly independent if: SET  a1v1 ++anvn ==0  ai 01 in

If the vectors v1v2  vn are not linearly independent then they are said to be linearly dependent. Unique representation. The vectors v1v2  vn are linearly independent if and only if

a1v1 ++ anvn = b1v1 ++ bnvn

implies that ai = bi , for 1 in . No vector can be built from A collection of two or more vectors is linearly independent if the rest. and only if none of those vectors can be expressed as a linear combination of the rest. Expanding a linearly inde- Let S = v1v2  vn be a linearly independent set. If pendent set. v  SpanS , then v1v2  vn v is again a linearly indepen- dent set. 110 Chapter 3 Bases and Dimension

Linear Independence Theo- A homogeneous system of m linear equations in n unknowns S has rem. only the trivial solution if and only if rref coefS has n leading ones. n n Linear independence in R . A set of vectors v1v2  vm in  is linearly independent if and only if the row-reduced-echelon form of the nm matrix th whose i columns is the (vertical) n-tuple vi has m leading ones.

BASIS A set of vectors  = v1v2  vn in a vector space V is said to be a basis for V if: (1)  spans V and: (2)  is linearly independent.

A spanning set can’t have If w1w2  wn spans a space V and if v1v2  vm is a fewer elements than the linearly independent subset of V, then nm . number of elements in any linearly independent set.

All bases for a vector space If w1w2  wn and v1v2  vm are bases for a vector contain the same number of space V, then nm= . vectors.

You can show that a set of Let v1v2  vn be a set of n vectors in a vector space V of n vectors in an n-dimen- dimension n. The following are equivalent: sional vector space is a basis by EITHER showing (i)v1v2  vn is a basis. that they span the space, OR by showing that it is a lin- (ii)v1v2  vn spans V. early independent set—you (iii)v v  v is linearly independent. DON’T have to do both: 1 2 n

Expansion Theorem Any linearly independent set of vectors in V can be extended to a basis for V. Reduction Theorem Any spanning set of vectors in V can be reduced to a basis for V.

Reducing a set of vectors S n Let v1v2  vm be vectors in  . If A is the nm matrix n in  to a basis for th Span(S) whose i column is the n-tuple vi , then the set consisting of th those vectors vi , where the i column of rrefA contains a leading one, is a basis for the space spanned by v1v2  vm . 4.1 Linear Transformations 111

4115 CHAPTER 4 LINEARITY In this chapter we turn our attention to functions from one vector space to another which, in a sense, preserve the algebraic structure of those spaces. Such functions, called linear transformations, are introduced in Section 1. As you will see, each linear transformation T: VW gives rise to two important subspace; one, called the kernel of T, resides in the vector space V, while the other, called the image of T, lives in W. Those two subspaces are investigated in Section 2. Linear transformations which are also one-to-one and onto are called isomorphisms, and they are considered in Section 3.

§1. LINEAR TRANSFORMATIONS Up to now we have focused our attention exclusively on the internal nature of a given vector space. The time has come to consider functions from one vector space and another: A linear transformation is also called a linear func- DEFINITION 4.1 A function T: VW from a vector space V to tion, or a . A LINEAR a vector space W is said to be a linear trans- linear map T: VV from a formation if for every vv   V and r   : vector space to itself is said TRANSFORMATION to be a linear operator. (1): Tvv+  = Tv + Tv and (2): Trv = rTv Let’s focus a bit on the statement: Tvv+  = Tv + Tv for T: VW . There are two plus signs in the equation, but the one on the left is happening in the space V, while the one on the right takes place in the space W. What the statement is saying is that you can per- form the sum in the vector space V and then carry the result over to the vector space W via the transformation T, or you can first carry those two vectors v and v over to W (via T) and then perform their sum in the space W. In the same fashion, Trv = rTv is saying that you can perform scalar multiplication in V and then carry the result over to W, or you can first carry the vector v to W and then perform the scalar multiplication in that vector space.

EXAMPLE 4.1 Show that the function T: 3  2 given by Tabc = 2ab+  –c is linear. 112 Chapter 4 Linearity

SOLUTION: Let abc  a b c  3 and r   . T preserves sums: A smoother approach: T abc + a b c Tabc + a b c = Taa+  bb+  cc+  = Taa+  bb+  cc+  = 2aa+  + bb+   –cc+  Tabc = 2ab+  –c : = 2aa+  + bb+   –cc+  = 2a +++2a bb – c – c = 2a +++2a bb – c – c

= 2ab+  –c + 2a + b –c same = Tabc + Ta b c and: Tabc +2Ta b c = ab+  –c + 2a + b –c = 2ab++2a +b – c – c T preserves scalar products: Trabc ==Tra rb rc 2ra+ rb –rc ==r2ab+  –c rTabc

EXAMPLE 4.2 Determine if the function f: 3  3 given by fabc = 2cb2 –a is linear.

SOLUTION: If you are undecided on whether or not f is linear, you may want to challenge its linearity with a of specific voters ? f 213 + 401 = f213 + f401 ? f614 = f213 + f401 ? fabc = 2cb2 –a : 81– 6 = 61– 2 + 20– 4 81– 6 = 81– 6 Yes! The above establishes nothing. It certainly does not show that the function f is not linear, nor does it establish its linearity since we have but demonstrated that it “works” for the two chosen vectors 213 and 401 . Let’s try another pair: ? f 32– 5 + 49– 3 = f32– 5 + f49– 3 ? f772 = f32– 5 + f49– 3 ? 4497– = 10 4– 3 + –681– 4 ? 4497– = 4857– No! Since f 32– 5 + 49– 3  f32– 5 + f49– 3, the func- tion is not linear. You can also show that the above function is not linear by dem- onstrating, for example, that f2123  2f123 . To show that a function is not linear you need only come up with a specific counterexample which “shoots down” EITHER (1) or (2) of Definition 4.1. 4.1 Linear Transformations 113

CHECK YOUR UNDERSTANDING 4.1

Is the function f: 2  3 given by fab = ab+  2ba – b Answer: Ye. linear? Justify your answer.

Given that a linear transformation preserves vector space structures, it may come as no surprise to find that it maps zeros to zeros, and inverses to inverses: THEOREM 4.1 If T: VW is linear, then: In order to distinguish where the different zero in the space W zeros preside, we are (a) T0V = 0W using 0V and 0W to indicate the zero is in zero in the space V the vector space V and W, respectively. inverse of the vector Tv in the space W (b) T–v = –Tv inverse of v in the space V

PROOF: (a) Note that three different zeros are featured below: by linearity

T0V ===T0  0V 0T0V 0W Theorem 2.7, page 53

by linearity

(b) T–v ===T–1 v –1 Tv –Tv Theorem 2.8, page 54

CHECK YOUR UNDERSTANDING 4.2

2 Use Theorem 4.1(a) to show that the function f:   P2x given Answer: See page B-12. by fab = ax2 ++bx 1 is not linear. The following Theorem compresses the two conditions of Definition 4.1 into one: You can perform the vector operations in V THEOREM 4.2 T: VW is linear if and only if: and then apply T to that Trvv+  = rTv + Tv result: Trvw+ , or you can first apply T for all vv   V and r   and then perform the vector operations in W: PROOF: If T is linear, then for all vv   V and r   : rTv + Tw . Either way, you will end up at Trvv+  ==Trv + Tv rTv + Tv the same vector in W. (1) of Definition 4.1 (1) of Definition 4.1 114 Chapter 4 Linearity

Conversely, if Trvv+  = rTv + Tv for all vv   V and r   , then: (1): Tvv+  ==T1vv+  1Tv + Tv =Tv + Tv Axiom (viii), page 40

Theorem 4.1(a) (2): Trv ==Trv + 0 rTv + T0 ==rTv + 0 rTv Axiom (iii), page 40

You are invited to establish the following generalization of the above result in the exercises:

THEOREM 4.3 Let T: VW be linear. For any vectors

v1v2  vn in V, and any scalars

a1a2  an :

Ta1v1 ++ anvn = a1Tv1 ++anTvn

CHECK YOUR UNDERSTANDING 4.3 Answer: See page B-12. Use Theorem 4.2 to establish the linearity of the function of CYU 4.1. The following theorem tells us that linear transformations map sub- spaces to subspaces: THEOREM 4.4 Let T: VW be linear. If S is a subspace of V, then TS= TssS is a subspace of W.

See Theorem 2.13, page 61 PROOF: For Ts1  Ts2  TS , and r   :

rT s1 + Ts2 = Trs1 + s2  TS since S is a subspace, since T is linear rs1 + s2  S

CHECK YOUR UNDERSTANDING 4.4

PROVE OR GIVE A COUNTEREXAMPLE: Let T: VW be linear, and SV . If TS is a subspace of W, then S must be a subspace of V.

A LINEAR MAP IS COMPLETELY DETERMINED BY ITS ACTION ON A BASIS

Suppose you have a run-of-the-mill function f: 2  3 , and you know that f10 = 123 and that f01 = 23– 5 . What can you say about f35 ? Nothing. But if T: 2  3 is linear, with T10 = 123 and T01 = 23– 5 , then: 4.1 Linear Transformations 115

T35 = T 310 + 501 by linearity: = 3T10 + 5T01 ==3123 + 523– 5 13 21– 16

Yes: A linear transformation T: VW is COMPLETELY DETERMINED by its action on a basis of V In particular, if you have two linear transformations that act identi- cally on a basis, then those two transformations must, in fact, be one and the same: THEOREM 4.5 Let V be a finite dimensional space with basis v1v2  vn . If T: VW and L: VW

are linear maps such that Tvi = Lvi for 1 in, then Tv = Lv for every v  V .

PROOF: Let v  V . Since v1v2  vn is a basis for V, there exist

scalars a1a2  an such that:

v = a1v1 + a2v2 ++ avn We then have: linearity of T

Tv ==Ta1v1 + a2v2 ++ anvn a1Tv1 +++a2Tv2 anTvn Since Tv= Lv: i i = a1Lv1 +++a2Lv2 anLvn linearity of L: ==La1v1 + a2v2 ++ anvn Lv

CHECK YOUR UNDERSTANDING 4.5

PROVE OR GIVE A COUNTEREXAMPLE: Theorem 4.5 still holds if

Answer: See page B-12. v1v2  vn is a spanning set for V (not necessarily a basis).

The next theorem gives a method for constructing all linear maps from one vector space to another:

THEOREM 4.6 Let v1v2  vn be a basis for a vector LINEAR space V, and let w1w2  wn be n arbitrary CONSTRUCTION vectors (not necessarily distinct) in a vector space W. There is a unique linear transforma-

tion T: VW which maps vi to wi for 1 in; and it is given by:

Ta1v1 ++ anvn = a1w1 ++ anwn 116 Chapter 4 Linearity

PROOF: From Theorem 4.4, we know that there can be at most one

linear transformation T: VW such that Tvi = wi for 1 in . We complete the proof by establishing the fact that the above func- tion T is indeed linear:

For v = a1v1 ++ anvn and v = b1v1 ++ bnvn in V, and r   :, n n  Trv+ v = Trav + b v Trv+ v  i i  i i i = 1 i = 1 n n  definition of ==Tra + b v ra + b w  i i i  i i i i = 1 i = 1 n n

==raiwi + biwi rT v + Tv T rT v + Tv   i = 1 i = 1

EXAMPLE 4.3 Let T: 2  3 be the linear transformation which maps 11 to 324 and 41 to 10– 5. Determine: (a) T96 (b) Tab

SOLUTION: (a) It is easy to see that 11  41 is a basis for 2 . We express 96 as a linear combination of the vectors in that basis: 9 = r + 4s 96 = r11 + s41 6 = rs+ _ : 3== 3ss 1 6 ==r +51  r Applying the linear transformation to 96 = 511 + 141 we have: T96 ==T 511 + 141 5T11 + 1T41 = 5324 + 10– 5 = 15 10 20 + 10– 5 = 16 10 15 (b) Repeating the above process for ab , we have: ab = r11 + s41 ar= + 4s brs= + ab– _: ab–== 3ss ------3 ab– 4ba– br==+ ------ r ------3 3 4.1 Linear Transformations 117

Consequently: 4ba– ab– Tab ==rT11 + sT41 ------324 + ------10– 5 3 3 – 2a + 11b – 2a + 8b – 9a + 21b details omitted: = ------------3 3 3

CHECK YOUR UNDERSTANDING 4.6

3 Answer: Let T: R  P2x be the linear transformation which maps (a) 4x2 + 4x – 10 100 and 020 to 2x2 + x , and maps 111 to x – 5 . 2ab+ – 3c x2 (b) Determine: c + a + --b- – --- x – 5c 2 2 (a) T342 (b) Tabc .

COMPOSITION OF LINEAR FUNCTIONS

As you may recall from earlier courses, for given functions f: XY and g: YZ , the composition of f followed by g is that function, denoted by gf , that is given by gf x = gfx (first apply f, and then apply g): f g X Y Z x z . .y .

gf The following theorem tells us that the composition of linear functions is again a :

THEOREM 4.7 If T: VW and L: WZ are linear, then the composition LT: VZ is also linear. PROOF: For vv   V and r   we show that the function LT sat- isfies the condition of Theorem 4.1: linearity of T L T rv+ v  LT rv+ v ==LTrvv+  LrTv + Tv linearity of L: = rL Tv + LTv

rLT v + LT v Definition of composition: = rLT v + LT v 118 Chapter 4 Linearity

EXAMPLE 4.4 3 Let T:   M22 and L: M22  P2 be given by:

Tabc = ab+0 cb– and: ab L= bx2 ++ac– xc cd (a) Show that T and L are linear. (b) Show directly that the composite function 3 LT:   P2 is again linear.

SOLUTION: Linearity of T: Trabc + a'b' c' = Tra+ a' rb+ b' rc+ c'

= ra+ a' +0rb+ b' rc+ c' –rb+ b'

= r ab+0+ a' +0b' cb– c' –b' = rTabc + Ta'b' c' Linearity of L:

ab a' b' ra+ a' rb+ b' Lr+ = T cd c' d' rc+ c' rd+ d' = rb+ b' x2 ++ra+ a' – rc+ c xrcc+ ' = rbx2 ++ac– xc+ b'x2 ++a' – c' xc'

ab a' b' = rT+ T cd c' d'

3 The composite function LT:   P2 : ab+0 LT abc ==LTabc L cb– = 0x2 ++ab+ – c xc = abc+ – xc+ 4.1 Linear Transformations 119

3 Linearity of LT:   P2 :

LT rabc + a'b' c' = LT ra+ a' rb+ b' rc+ c' = r aa++'  r bb +'  – rc+ c' xrcc+ + ' = ra+ b– cxc+ + a' + b' – c' xc+ ' = rLT abc + LT a'b' c'

CHECK YOUR UNDERSTANDING 4.7

(a) Show that 10  11 and 010 110 011 are bases for 2 and 3 , respectively. (b) Let T: 2  3 be the linear transformation which maps 10 to 020 and 11 to 101 . Let L: 3  2 be the linear transformation which maps 010 to 01 and both 110 and Answer: (a) See page B-13. 011 to 10 . Determine the map L T: 2  2 . (b) 2b 2a – 4b  120 Chapter 4 Linearity

EXERCISES

Exercises 1-18. Determine whether or not the given function f is a linear transformation.

1.f:  , where fx = –5x . 2.f:  , where fx = 0 . 3.f:  , where fx = 1 .

4.f: 2   , where fxy = 2x – 3y .

5.f: 2   , where fxy = xy .

6.f: 2  2 , where fxy = xyxy+  – .

7.f: 2  3 , where fxy = xxyy  .

x y 8.f: 2  3 , where fxy = ---y --- . 2 2

3 2 9.f:   P2x , where fabc = ax ++bx c .

2 10.f: P2   , where fax++bx c = abc++ .

2 11.f: P2   , where fax++bx c = abc .

2 2 12.f: P2  P2 , where fax++bx c = ax+ 1 + bx+ 1 + c .

2 3 2 13.f: P2  P3 , where fax++bx c = ax ++bx cx .

3 0 ab 14.f:   M23 , where fabc = . cba

ab 15.f: M22   , where f = ad– bc . cd

ab ab–0 16.f: M22  M22 , where f = . cd 0 cd

17.f:  , where fx = sinx .

18.f:   V , where fx = 2x , and V is the vector space of Example 2.5, page 47. 4.1 Linear Transformations 121

19. Let the linear map T: 2  3 be such that: T11 ==120  T02 102 Find: (a) T53 (b) Tab

20. Let the linear map T: 3  2 be such that: T100 ==32  T010 23  and T001 =11 Find: (a) T53– 2 (b) Tabc

2 21. Let the linear map T:   M22 be such that:

T10 ==10 T11 21 21 32 Find: (a) T53 (b) Tab

2 22. Let the linear map T:   M22 be such that:

T11 ==10 T31 21 21 32 Find: (a) T53 (b) Tab

23. Show that there cannot exist a linear transformation T: R2  R2 such that: T12 ==53  T53 12  and T11 =22

2 24. Show that there cannot exist a linear transformation T: R  P2 such that: T12 ==x2 + 2 T53 5x +3 and T11 =x2 ++x 1

25. Show that the identity function I: V  V , given by Iv = v for every v in V, is linear.

26. Show that the zero function Z: VW , given by Zv = 0 for every v in V, is linear. (Refer- ring to the equation Zv = 0 , where does 0 live?)

27. In and calculus, functions of the form yaxb= + are typically called linear functions. Give necessary and sufficient conditions for a function of that form to be a linear operator on the vector space  .

28. Show that for any a   the function Ta: F   , where Taf = fa is linear. (See Theorem 2.4, page 44.)

29. (Calculus Dependent) Let D be the subspace of the function space F consisting of all differentiable functions. Let T: D  F be given by Tf= f  , where f  denotes the of f. Show that T is linear. 122 Chapter 4 Linearity

30. (Calculus Dependent) Show that the function T: P2  P4 , given by d Tpx = 5x3 px + 5 is linear. dx

31. (Calculus Dependent) Show that if the linear function T: P2  P1 is such that Tx2 ==2xTx  1 , and T1 = 0 , then T is the derivative function.

32. (Calculus Dependent) Let Iab denote the vector space of all real-valued functions that are integrable over the interval ab . Let T: Iab  be given by Tf= bfxdx . a Show that T is linear.

33. Let T: VW be linear and let S be a subspace of V. Show that Tv v  S is a subspace of W. 34. (PMI) Use the Principle of Mathematical Induction to prove Theorem 4.3. 35. Let LVW = T: VW T is linear , with addition and scalar multiplication given by:

T1 + T2 v ==T1v + T2v and rT v rTv Show that LVW is a vector space.

36. (a) Show that if a function f:  satisfies the property that frx = rfx for every r   and x   , then f is a linear function: which is ti say, that it must also satisfy the

property that fx1 + x2 = fx1 + fx2 for every x1x2  . (b) Give an example of a function f: 2  2 satisfying the property that frab = rfab for every r   and every ab  2 but which does not satisfy the property that f: ab + cd = fab + fcd . (Note: It is by no means a trivial task to establish the existence of a non-linear function f: VW which

satisfies the property that fv1 + v2 = fv1 + fv2 for every v1 v2  V .)

37. Let f: VW satisfy the condition that fv1 + v2 = fv1 + fv2 for every v1 v2  V . Show that: (a)f3v = 3fv for every v  V . (b) f --2-v = --2- fv for every v  V . 3 3

38. Let f: VW satisfy the condition that fv1 + v2 = fv1 + fv2 for every v1 v2  V . Show that: (a) (PMI) fnv = nfv for every v  V , and for every integer n. a a (b)f---v = --- fv for every v  V , and for every rational number --a- . b b b 4.1 Linear Transformations 123

Exercises 39-44. Determine the indicated composition of the given linear functions, and then directly verify its linearity. 39.LT: 2  2 , where T: 2   is given by Tab = ab+ and L:  2 by La = 2aa – .

40.LT: 2   , where T: 2  2 is given by Tab = –a ab+ and L: 2   by Lab = a + 2b

2 2 41.LT:   P2 , where T:    is given by Tab = 3ab+ and L:   P2 by La = ax2 + 2a .

3 2 42.LT: P2  M22 , where T: P2   is given by Tax++bx c = abc and

3 2ab L:   M22 by Labc = . 0 –c

43.KLT:  3 , where T:  2 is given by Ta = aa – , L: 2  2 by Lab = 2aa + b, and K: 2  3 by Kab = –a 2ba + b .

3 3 2 44.KLT:   P2 , where T:    is given by Tabc = abac+  – , 2 2 L:    by Lab = ab– , and K:   P2 by Ka = ax ++2ax 3a .  45. (PMI) Let Li: Vi  Vi + 1 be linear, for 1 in . Show that Ln L2L1: V1  Vn + 1 is linear.

PROVE OR GIVE A COUNTEREXAMPLE

46. For any a   the function Ta: VV given by Tav = av is linear.

47. For any v  V the function T : VV given by T v = vv+ is linear. 0 v0 v0 0

48. Let T: VW be linear. If Tv1 Tv2 Tvn is a linearly independent subset of W

then v1v2  vn is a linearly independent subset of V.

49. Let T: VW be linear. If v1v2  vn is a linearly independent subset of V then

Tv1 Tv2 Tvn is a linearly independent subset of W.

50. If for given functions f: VW and g: WZ the composite function gf: VZ is lin- ear, then both f and g must be linear.

51. If for given functions f: VW and g: WZ the composite function gf: VZ is lin- ear, then f must be linear.

52. If, for given functions f: VW and g: WZ , the composite function gf: VZ is lin- ear, then g must be linear. 124 Chapter 4 Linearity

4

§2. KERNEL AND IMAGE For any given transformation T: VW , we define the kernel of T to be the set of vectors in V which map to the zero vector in W [see Figure 4.1(a)], and we define the set of all vectors in W which are “hit” by some Tv to be the image of T [see Figure 4.1(b)].

V W V W T . T 0

Kernel of T Image of T (a) (b) Figure 4.1 More formally:

DEFINITION 4.2 Let T: VW be linear. The kernel (or null KERNEL space) of T is denoted by KerT and is given by: KerT = v  VTv = 0 The image of T is denoted by ImT and is IMAGE given by: ImT = w  WTv = w for some v  V

Both the kernel and image of a linear transformation turn out to be subspaces of their respective vector space:

THEOREM 4.8 Let T: VW be linear. Then: (a)KerT is a subspace of V. (b)ImT is a subspace of W.

PROOF: We employ Theorem 2.13, page 61 to establish both parts of the theorem. v1 v2  KerT and rR (a) Since T0 = 0 , KerT is nonempty. Let v1 v2  KerT and r   . Then: linearity

Trv1 + v2 ===rTv1 + Tv2 r  0 + 0 0 since v  v  kerT rv1 +Kerv2  T 1 2

Since T maps rv1 + v2 to 0, rv1 +Kerv2  T . 4.2 Kernel and Image 125

(b) Since T0 = 0 , ImT is nonempty. w  w  ImT and rR 1 2 Let w1 w2  ImT and rR . Choose vectors v1 v2 such that Tv1 = w1 and Tv2 = w2 (how do we know that such vec- tors exist?). Then:

Trv1 + v2 ==rTv1 + Tv2 rw1 + w2 Since we found a vector in V which maps to rw1 + w2 , rw +Imw  T 1 2 rw1 +Imw2  T .

DEFINITION 4.3 Let T: VW be linear. NULLITY The dimension of KerT is called the nul- lity of T, and is denoted by nullityT . The dimension of ImT is called the RANK of T, and is denoted by rankT . The following theorem will be useful in determining the rank of a lin- ear transformation.

In particular, if THEOREM 4.9 Let T: VW be linear. If  = v1v2  vn is a basis for V, then Spanv1v2  vn = V Then: Tv1 Tv2 Tvn will span ImT . SpanTv1 Tv2 Tvn = ImT

PROOF: We show that any w  ImT can be written as a linear com-

binations of the vectors Tv1 Tv2 Tvn : Let w  ImT , and let v  V be such that Tv = w . Since w  ImT S = v1v2  vn spans V, we can express v as a linear combination of the vectors in S:.

v = a1v1 +++a2v2  anvn n By linearity: w = a Tv Tv ==w a Tv +++a Tv a Tv  i i 1 1 2 2 n n i = 1

EXAMPLE 4.5 3 (a) Show that the functionT:   P2 given by: Tabc = ab+ x2 ++cx c is linear. (b) Determine ImT and rankT . (c) Determine KerT and nullityT . 126 Chapter 4 Linearity

SOLUTION: (a) For abc  a b c  3 and r   : Trabc + a b c = Tra+ arb+ b rc+ c = ra+ a + rb+ b x2 ++rc+ c xrcc+  regrouping: = ra+ bx2 ++cx c + a + b x2 ++cxc = rTabc + Ta b c (b) By Theorem 4.9, we know that the vectors: T100 = x2 T010 ==x2 T001 x + 1 span the image of T. Since x2 and x + 1 are linearly independent, x2 x + 1 is a basis for ImT . Consequently, rankT = 2 .

(c) By definition: KerT = abc Tabc = 0 = abc ab+ x2 ++cx c = 0x2 ++0x 0 This leads us to the following system of equations: ab+0=    c ==0 and ab– c = 0  Thus: KerT = aa –  0 aR . Since 110 –  is a basis for KerT , nullityT = 1 .

CHECK YOUR UNDERSTANDING 4.8

(a) Show that the function T: P2  M22 given by: Tax2 ++bx c = ab is linear. Answer: (a) See page B-13. ca (b) rankT = 3 , nullityT = 0 (b) Determine rankT and nullityT .

THEOREM 4.10 Let V be a vector space of dimension n, and DIMENSION let T: VW be linear. Then: THEOREM rankT + nullityT = n

PROOF: Start with a basis v1v2 vk for KerT and extend it (if

necessary) to a basis v1v2  vk vk + 1  vkt+ for V [see Theo- rem 3.12(a), page 100], where: kt+ = n —the dimension of V.

We will show that the t vectors Tvk + 1 Tvkt+ constitute a basis for ImT . This will complete the proof, for we will then have:

rankT + nullityT ==k + t n 4.2 Kernel and Image 127

As you know, to establish the fact that Tvk + 1 Tvkt+ is a basis for ImT we have to verify that:

(1) The vectors Tvk + 1 Tvkt+ span the space ImT .

Why can’t you simply And that: show just one of the two? (2) The vectors Tvk + 1 Tvkt+ are linearly independent. Let’s do it: (1) Let w  ImT . Choose v  V such that Tv = w .

Since v1v2  vk vk + 1  vkt+ is a basis for V, we can express v as a linear combination of those kt+ vectors:   v = a1v1 ++++a2v2 akvk ak + 1vk + 1 ++akt+ vkt+ By linearity:   w ==Tv a1Tv1 ++++a2Tv2 akTvk ak + 1Tvk + 1 ++akt+ Tvkt+

Since v1v2  vk  KerT

==0 +++ak + 1Tvk + 1 akt+ Tvkt+ ak + 1Tvk + 1 ++akt+ Tvkt+

a linear combination of the vectors Tvk + 1 Tvkt+

(2): Assume that: b Tv ++b Tv = 0 1 k + 1 t kt+ b1Tvk + 1 + b2Tvk + 2 ++btTvkt+ = 0 By linearity:

Tb1vk + 1 + b2vk + 2 ++btvkt+ = 0 And therefore:

b1vk + 1 +Kerb2vk + 2 ++btvkt+  T

Since v1v2 vk is a basis for KerT , we can then find scalars

c1 ck such that:

b1vk + 1 + + btvkt+ = c1v1 ++ckvk Or:

–c1 v1 ++–ck vk +b1vk + 1 + + btvkt+ = 0

Since v1v2  vk vk + 1  vkt+ is a basis, it is a linearly inde- pendent set of vectors. Consequently those c’s and b’s must all be b ==0 b 0 b =0 1 2 t zero; in particular: b1 ==0 b2 0 bt =0 . 128 Chapter 4 Linearity

EXAMPLE 4.6 Determine bases for the kernel and image of the linear function T: P3  M22 given by:

Tax3 +++bx2 cx d = ab– c 2ca+ d

SOLUTION: We go for the kernel, as it is generally easier to find than the image apace: To say that: Tax3 +++bx2 cx d = 0

Is to say that: ab– c = 00 2ca+ d 00 Equating coefficients leads us to a homogeneous system of equations: System S is certainly easy each of these four enough to solve directly. equations is equal to 0. Still: a b cd a b cd ab–0=  ad= –   11–00 1 001  c = 0  coef[S] 0010 0 1 01 bd= –  S:  rref  2c = 0  0020 001 0 c = 0  ad+0=    1001 0000 d is free  From the above, we see that: KerT ==– dx3 – dx2 + dd R dx– 3 –1x2 + dR Conclusion: – x3 –1x2 + is a basis for KerT . Knowing that nullityT = 1 , we turn to Theorem 4.10 and con- clude that

rankT ===dimP3 –41nullityT –3 See Exercise 16, page 105 At this point, the easiest way to find a basis for ImT is to apply T to

3 vectors in P3 making sure that you end up with 3 linearly indepen- dent vectors in ImT —a basis for ImT . Which 3 vectors should we start with? Basically, you can take any 3 randomly chosen vec- tors, and the chances are that they will do fine (think about it); we will go with the 3 vectors x3 , x2 , and x + 1 : Recall that: 1x3 +++0x2 0x 0 0x3 +++1x2 0x 0 3 2 Tax3 +++bx2 cx d = 0x +++0x 1x 1

ab– c 3 10 2 –01 11 2ca+ d Tx == Tx  Tx + 1 = 01 00 21 4.2 Kernel and Image 129

We leave it to you to verify that above three vectors, which are cer- tainly in Im(T) (why?), are linearly independent, and therefore consti- tute a basis for the 3-dimensional space Im(T).

CHECK YOUR UNDERSTANDING 4.9

Determine bases for the kernel and image of the linear function Answer: See page B-14. T: 3  4 given by Tabc = 2ab + ccb .

ONE-TO-ONE AND ONTO FUNCTIONS As you may recall: ...... DEFINITION 4.4 A function f from a set A to a set B is said . . . . ONE-TO-ONE to be one-to-one if: A B A B fa==fa  a a one-to-one not one-to-one A function f from a set A to a set B is said to ...... ONTO ...... be onto if for every bB there exist aA such that fa= b . A B A B onto not onto When dealing with a linear transformation, we have:

The first part of this theorem THEOREM 4.11 (a) A linear transformation T: VW is one- is telling is that if a linear map is “one-to-one at 0,” to-one if and only if KerT = 0 . then it is one-to-one every- where. Certainly not true for (b) A linear transformation T: VW is onto other functions: if and only if ImT = W .

PROOF: (a) Let T be one-to-one, and suppose that Tv = 0 . Since T0 = 0 [Theorem 4.1(a), page 113], and since there can be but one v  V that is mapped to 0  W (T is one-to-one), v = 0 . Conversely, assume that Tv = 0  v = 0 . Then:

Tv1 = Tv2

Tv1 – Tv2 = 0

T is linear: Tv1 – v2 = 0

Tv = 0  v = 0: v1 – v2 = 0

v1 = v2 (b) Follows directly from the definition of onto (Definition 4.4) and the definition of ImT (Definition 4.2). 130 Chapter 4 Linearity

EXAMPLE 4.7 Show that the linear function T: 3  3 given by Tabc = ac+  3bc – b is one-to-one.

SOLUTION: We show that Tv ==0  v 0 : Tabc = 0 Tabc ==ac+  3bc – b 000 equating coefficients: 3b ==0  b 0 cb–0=== c b  c 0 abc = 0 ac+0== a –c  a =0 Applying Theorem 4.11, we conclude that T is one-to-one.

Consider the adjacent graph of the func- tion fx= x3 – x . As you can see, there 2. are some y’s which are only “hit by one x” 1 (like y = 2 ), and there are some y’s . . . which are “hit by more than one x” (like -1 1 y = 0 ) — the function is kind of one-to- -1 one in some places, and not one-to-one in -2 other places. Linear transformations are not so fickle; if a linear transformation is “one-to-one anywhere” (not just at 0) then it is one-to-one everywhere:

CHECK YOUR UNDERSTANDING 4.10

Let T: VW be linear. Show that if there exists v  V such that Answer: See page B-14. Tv ==Tv  v v then T is one-to-one. 4.2 Kernel and Image 131

EXERCISES

Exercises 1-18. Determine if the given function is linear. If it is, find a basis for its kernel and its image space.

1.f:  , where fx = –5x 2.f:  2 , where fx = x 2x

3.f:  2 , where fx = xx – 4.f:  2 , where fa = aa 2

5.f: 2  2 , where fab = ab a 6.f: 2  2 , where fab = –2ba

2 3 3 7.f:   P1 , where fab = ax+ b 8.f:    , where f abc = 0bc

ab 2 9.f: M22  P2 , where f = ab– ++bc– xca– x cd

2 ab 10.f:   M22 , where fab = ab+ ab–

4 a 2b 11.f:   M22 , where fabcd = cd c+ d

3 12.f: P2   , where fpx = p1  p2  p3

2 13.f: P2   , where fpx = p0  p1

14.f: P3  P4 , where fpx = xp x

15.f: P3  P3 , where fpx = px+ p1

p1 p2 16.f: P3  M22 , where fpx = p3 p4

3 ab– bc– 17.f:   M22 , where fabc = ab+ bc+

4 2ac+ d 18.f:   M22 , where fabcd = cd– b 132 Chapter 4 Linearity

Exercises 19-27. (Calculus Dependent) Show that the given function is linear. Find a basis for its kernel and its image space.

19.f: P3  P3 , where fpx = px

20. f: P3  P3 , where fpx = px

21. f: P3  P3 , where fpx = px + px

22. f: P2  P , where fpx = px

23.f: P2  P3 , where fpx = 2px– 3px

1 24.f: P1   where fpx = pxdx 0

1 25.f: P2   where fpx = pxdx 0

x 26.f: P1  P2 where fpx = ptdt 0

x 27.f: P2  P3 where fpx= ptdt 0

28. Let T: 3  3 be given by Tabc = ab– ac+ c . (a) Which, if any of the following 3-tuples are in the kernel of T? (i) –330 (ii) 033 –  – (iii) 300 (b) Which, if any of the following 3-tuples are in the image T? (i) –330 (ii) 033 –  – (iii) 300

3 3 29. Let T:   P3 be given by Tabc = ax ++ax b+ c . (a) Which, if any of the following 3-tuples are in the kernel of T? (i) 100 (ii) 033 –  (iii) 001 (b) Which, if any of the following 3-tuples are in the image of T? (i) x3 + x (ii) 5 (iii) x3 + 5 30. Determine a basis for the kernel and image of the linear transformation T: 3  3 which maps 100 to 111 , 010 to 325 –  , and 001 to –23– 4.

3 31. Determine a basis for the kernel and image of the linear transformation T:   M22

which maps 100 to 21 , 010 to 22 , and 001 to 01 . 01 13 10 4.2 Kernel and Image 133

32. Determine a basis for kernel and image of the linear transformation T: M22  P4 which

maps 10 to 2x4 + 1 , 11 to x4 – x , 11 to 5, and 11 to x2 . 00 00 10 11

33. Find, if one exists, a linear transformation T: 2  3 such that: (a) 913 is a basis for ImT . (b)900  010 is a basis for ImT . (c)913 319 931 is a basis for ImT . 34. Find, if one exists, a linear transformation T: 3  3 such that: (a) 913 is a basis for KerT . (b)900  010 is a basis for KerT . (c)913 319 931 is a basis for KerT . 35. Let dimV = n , and let T: VV be the linear operator Tv = rv , for r   . Enumer- ate the possible values of nullityT and rankT . 36. Let T: VW be linear with dimV = 2 and dimW = 3 . Enumerate the possible val- ues of nullityT and rankT . 37. Let T: VW be linear with dimV = 3 and dimW = 2 . Enumerate the possible val- ues of nullityT and rankT . 5 38. Let T: P3   be a one-to-one linear map. Determine the rank and nullity of T. 5 39. Let T:   P3 be an onto linear map. Determine the rank and nullity of T. 40. Let T: VV be a linear operator, with dimV = n . Prove that ImT = V if and only if T is one-to-one.

41. Let T: VW be linear, with dimV = dimW . Prove that v1v1  vn is a basis for

V if and only if Tv1 Tv2 Tvn is a basis for W. 42. Give an example of a linear transformation T: VW such that ImT = KerT . 43. Let T: VW be a linear transformation, with dimV = n . Prove that T is one-to-one if and only if rankT = n . 44. Let T: VW be linear, with dimV = dimW . Prove that T is one-to-one if and only if T is onto. 45. Let T: VW and L: WZ be linear. (a) Prove that KerT  KerLT . (b) Give an example for which KerT = KerLT . (c) Give an example for which KerT  KerLT . 134 Chapter 4 Linearity

PROVE OR GIVE A COUNTEREXAMPLE

46. If spanTv1 Tv2 Tvn = ImT then spanv1v2  vn = V .

47. There exists a one-to-one linear map T: M23  P4 .

48. There exists a one-to-one linear map T: P4  M23 .

49. There exists an onto linear map T: M23  P4 .

50. There exists an onto linear map T: P4  M23 . 51. If T: VW is linear and dimV  dimW , then T cannot be onto. 52. If T: VW is linear and dimV  dimW , then T cannot be one-to-one. 53. If T: VW is linear and dimV  dimW , then T cannot be onto. 54. If T: VW is linear and dimV  dimW , then T cannot be one-to-one. 55. There exists a linear transformation T: 2  4 such that rankT = nullityT . 56. There exists a linear transformation T: 3  4 such that rankT = nullityT . 57. There exists a linear transformation T: 2  4 such that rankT  nullityT . 58. There exists a linear transformation T: 2  4 such that rankT  nullityT . 59. If T: VW is linear and if W is finite dimensional, then V is finite dimensional. 60. If T: VW is linear and if V is finite dimensional, then W is finite dimensional. 61. If T: VW is linear and if V is finite dimensional, then ImT is finite dimensional. 62. If T: VW is linear and if ImT is finite dimensional, then V is finite dimensional.If T: VW is linear and if KerT is finite dimensional, then either V or W is finite dimen- sional. 63. Let T: VW and L: WZ be linear. If dimV = 3dim W = 2 , and nullityT = 1 , then rankLT  1 . 64. Let T: VW and L: WZ be linear. If dimV = 3dim W = 3 , rankT = 1 , and nullityL = 2 , then rankLT = 1 . 65. Let T: VW and L: WZ be linear. If dimV = 3dim W = 3 , rankT = 1 , and nullityL = 2 , then rankLT  1 . 66. Let T: VW and L: WZ be linear, with dimW = n and dimZ = m . If T is one- to-one and L is onto, then dimV = nm– . 4.3 Isomorphisms 135

4

§3. ISOMORPHISMS

We can all agree that there is little difference between the vector n space  with its horizontal n-tuples, and the space Mn  1 of “vertical n-tuples”. In this section, we show that there is, in fact, little difference between the vector space n and any n dimensional vector space what- soever.

BIJECTIONS AND INVERSE FUNCTIONS One-to-one and onto functions were previously defined on page 129. Of particular interest are functions that satisfy both properties:

DEFINITION 4.5 A function f: A  B that is both one-to-one BIJECTION and onto is said to be a bijection. . . Roughly speaking: ...... A bijection f: A  B serves to pair of each A B elements of A with those of B (see margin). a bijection THEOREM 4.12 If f: A  B and g: BC are bijections, then the composite function gf: AC is also a bijection. PROOF: gf is one-to-one: gf a = gf b gf a = gf b Definition of composition: gfa = gfb Since g is one-to-one: fa= fb ab= Since f is one-to-one: ab=

f g gf is onto: Let cC be given (see margin). Since g is onto there . c exists bB such that gb= c . Since f is onto, there a. b . exists aA such that fa= b . We then have: A B C g f  gf a ===gfa gb c this a “does the trick”

Figuratively speaking, if you reverse the arrows of the bijection f: 1234  6510 of Figure 4.2(a), you end up with the bijection f –1: 6510  1234 of Figure 4.2(b), which is called the inverse of f. 136 Chapter 4 Linearity

–1 Only bijections f f 1. .6 1 6 f: X  Y . . 2 2 . 5 . 5 have inverses 3. . 1 3 . 1 –1 . . . f : Y  X 0 4. . 4. .0 (a) (b) Figure 4.2 The relationship between the functions f and f –1 depicted in the mar- gin reveals the fact that each function “undoes” the work of the other. f For example: 1. 6 . –1 –1 –1 2. f  f 2 ===f f2 f 5 2 .5 3 1 and .4 0. . f –1 . f f –1 5 ===ff–15 f2 5  In general: DEFINITION 4.6 The inverse of a bijection f: XY is the –1 INVERSE function f : YX such that: –1 FUNCTIONS f  f x = x for every x in X and –1 ff y = y for every y in Y.

CHECK YOUR UNDERSTANDING 4.11

–1 Answer: See page B-14. Prove that if f: XY is a bijection, then so is f : YX .

BACK TO LINEAR ALGEBRA

THEOREM 4.13 If the bijection T: VW is linear, then its inverse, T –1: WV is also linear. PROOF: Let ww   W and r   be given. Let vv   V be such that Tv = w and Tv = w . Then: linearity of T T –1rww+  ==T –1rTv + Tv T –1Trvv+  ===T–1T rvv+  rvv+  rT–1w + T–1w Definition of composition Definition 4.6 since Tv ==w and Tv w 4.3 Isomorphisms 137

EXAMPLE 4.8 Show that the linear map T: 3  3 given by: Tabc = ac+  3bc – b is a bijection. Find its inverse and show, directly, that T –1 is linear.

SOLUTION: T IS ONE-TO-ONE. We start with Tabc = Tab c and go on to show that abc = ab c (see Definition 4.4, page 129): Tabc = Tab c  ac+  3bc – b = a + c 3b c – b

ac+ = a + c  aa=   3b = 3b  bb=  then  then cb– = c – b  cc=  T IS ONTO. We start with ab c and find an abc such that T Tabc = ab c (see Definition 4.4, page 129):

abc Tabc ==ac+  3bc – b ab c . .abc Equating coefficients brings us to the system of equations: b a = a – c – ---- ac+ = a  3  b 3bb=    b = ---- then  3 then cb– = c b  c = c + ---- 3 b b b Let’s make sure that abc = a – c – ---- ---- c + ---- works: 3 3 3 b b b Tabc = Ta – c – ---- ---- c + ---- 3 3 3 b b b b b ==a – c – ---- +3c + ---- ---- c + ---- – ---- ab c 3 3 3 3 3

FINDING THE INVERSE OF T. Since b b b Ta – c – -------- c + ---- = ab c : 3 3 3

–1 b b b T ab c = a – c – ---- ---- c + ---- 3 3 3 Dropping the primes, we show, directly, that T–1: 3  3 given by b b b T–1abc = ac– – ------ c + --- IS LINEAR: 3 3 3 138 Chapter 4 Linearity

–1 –1 T –1rabc + ab c T rabc + ab c = T ra+ a rb+ b rc+ c rb b rb b rb b = ra+ a – rc – c – ----- – ---- ----- + ---- rc++ c ----- +---- 3 3 3 3 3 3 b b b b b b = = rac– – ------ c + --- + a – c – ---- ---- c + ---- 3 3 3 3 3 3

–1 –1 –1 rT abc + T –1a b c = rT abc + T a b c

CHECK YOUR UNDERSTANDING 4.12

2 Show that the linear map T:   P1 given by: Tab = ab+ xa– Answer: See page B-14. is a bijection. Find its inverse, and show directly that it is linear. Roughly speaking, two vector spaces will be considered to be the “same” if the vectors of one space can be pared off with those of the other, while preserving the vector space structures of those spaces. More formally:

DEFINITION 4.7 A linear map T: VW which is one-to-one ISOMORPHISM and onto is said to be an isomorphism from the vector space V to the vector space W.

3 EXAMPLE 4.9 Show that the function T:   P2 given by: Tabc = – ax2 ++bc+ x 3c is an isomorphism.

SOLUTION: We start off by establishing linearity, for we can then take advantage of previously established theory to show that the function is one-to-one and onto: Linearity: Trabc + ab c = Tra + a rb + b rc + c' = – ra + a x2 ++rb + b + rc + c x 3rc + c = ra– x2 ++bc+ x 3c + –ax2 ++b + c x 3c = rTabc + Tab c One-to-one: We show that KerT = 0 [see Theorem 4.11(a), page 129]: Tabc ==0  – ax2 ++bc+ x 3c 0x2 ++0x 0 Equation coefficients: –0a === bc+ 03 c 0 or: abc===0 4.3 Isomorphisms 139

Onto: We show that ImT = P2 [see Theorem 4.11(b)]: The Dimension Theorem of page 126, tells us that: rankT +3nullityT = Knowing that the nullity is 0, we conclude that rankT = 3 . Since P2 is of dimension 3: ImT = P2 (Exercise 50, page 107).

THEOREM 4.14 (a) Every vector space is isomorphic to itself. This theorem asserts that “isomorphic” is an equiv- (b) If V is isomorphic to W, then W is isomor- alence relation on any set phic to V. of vector spaces. See Exer- cises 37-39. (c) If V is isomorphic to W, and W is isomor- phic to Z, then V is isomorphic to Z.

PROOF: (a) The identity map IV : VV is easily seen to be an iso- morphism. (b) To say that V is isomorphic to W is to say that there exists an iso- morphism T: VW . Since T –1:WV is also a linear bijection (CYU 4.12 and Theorems 4.13), W is isomorphic to V. (c) Let T: VW and L: WZ be isomorphisms. Since LT: VZ is both a bijection (Theorem 4.12) and linear (Theorem 4.7, page 117), V is isomorphic to Z. Theorem 4.14(b) enables us to formulate the following definition: DEFINITION 4.8 Two vector spaces V and W are isomorphic, ISOMORPHIC written VW , if there exists an isomor- SPACES phism from one of the vector spaces to the other.

CHECK YOUR UNDERSTANDING 4.13

4 Prove that   M22 . (You have to exhibit an isomorphism from Answer: See page B-15. one of the spaces to the other, whichever you prefer). The following lovely result says that all n dimensional vector spaces are isomorphic to the Euclidean n-space. THEOREM 4.15 If V is a vector space of dimension n, then: V  n

PROOF: Let v1v2  vn be a basis for V, and let e1e2  en be the standard basis for n (see page 94). Consider the function n T: V   , given by: Ta1v1 ++ anvn = a1e1 ++ anen . Theorem 4.6, page 115 assures us that T is linear. We complete the proof by showing that T is both one-to-one and onto (and therefore an isomorphism). 140 Chapter 4 Linearity

Tv = Tv T is one-to-one: Let Tv = Tv , for: v ==a1v1 ++ anvn and v b1v1 ++ bnvn Then: Ta1v1 ++ anvn = Tb1v1 ++ bnvn

a1Tv1 ++anTvn = b1Tv1 ++bnTvn

a1e1 ++ anen = b1e1 ++ bne1n Since e1e2  en is a linearly independent set of vectors we have, by Theorem 3.6, page 89, that ai = bi , for vv=  1 in; in other words: vv=  . n T is onto: For X = a1e1 ++ anen   : For X  Rn Ta1v1 ++ anvn = a1Tv1 ++anTvn

==a1e1 ++ anen X Tv = X CHECK YOUR UNDERSTANDING 4.14 Let V and W be finite-dimensional vector spaces. Show that VW if Answer: See page B-15. and only if dimV = dimW .

A ROSE BY ANY OTHER NAME

Let T: VW be an isomorphism. Being a bijection it links every element in V with a unique element in W (every element in V has its own W-counterpart, and vice versa). Moreover, if you know how to T–1 v1 w1 function algebraically in V, then you can also figure out how to function T–1 v w algebraically in W. Suppose, for example, that you forgot how to add or 2 2 scalar multiply in the space W, but remember how to add and scalar rw1 + w2 rv + v –1 1 2 multiply in V. To figure out rw1 + w2 in W you can take the “T bridge” back to V and find the vectors v1 and v2 such that Tv1 = w1 T and Tv2 = w2 . Do the calculations rv1 + v2 in V and then take the “T bridge” back to W to find the value of the vector rw1 + w2 , for it coincides with the vector Trv1 + v2 :

Trv1 + v2 ==rTv1 + Tv2 rw1 + w2 Indeed, the intimacy between isomorphic vector spaces is so great that isomorphic spaces are said to be equal up to an isomorphism. Basi- cally, if a vector space W is isomorphic to V, then the two spaces can only differ from each other in “appearance.” For example, M22 looks different than 4 , but you can easily link its elements with those of V:

ab link abcd cd And that linkage preserving the algebraic structure:

r ab ++a b link rabcd a b c d cd c d 4.3 Isomorphisms 141

We note that all vector space properties are preserved under an iso- morphisms. In particular, as you are asked to verify in the exercises, a linear map T: VW maps: Linearly independent sets in V to linearly independent sets in W. Spanning sets in V to spanning sets in W. CHECK YOUR UNDERSTANDING 4.15

Let L: VW be an isomorphism. Prove that if v1v2  vn is a Answer: See page B-16. basis for V, then Lv1 Lv2 Lvn is a basis for W. At times, one can take advantage of established properties of Euclid- ean spaces to address issues in other vector spaces: EXAMPLE 4.10 Find a subset of the set:  12 –23 11– 2 36 –146 S =     32– 13 313– 96– –46 which is a basis for SpanS .

SOLUTION: Lets move the elements of S over to 4 via the isomor- ab phism T= abcd : cd

For f: XY , and SX : 123– 2–3 213 11–313 2 – TS=  fS= fssS 369– 6 –1466–4 Applying Theorem 3.13, page 103, to the vectors in TS we see that the first, second, and fifth vector in TS constitute a basis for SpanTS:

13–1136– 1 0230 22–614 2 0 1 –003 rref 31 3 9– 6 00 0 01 –3132 –6–4 00 0 00

Utilizing the result of the CYU 4.16, we conclude that the corre- sponding first, second, and fifth matrix in S constitute a basis for SpanS .

CHECK YOUR UNDERSTANDING 4.16 Proceed as in Example 4.10 to find a subset of the set

S = 2x3 –53x2 + x – 1x3 –8x2 + x – 3 x2 + 11x – 5 –2x3 ++x2 3x – 2

Answer: See page B-16. in P3 which is a basis for Span (S). 142 Chapter 4 Linearity

Let f be a bijection from a vector space V (with addition denoted by v1 + v2 and scalar multiplication by rv ) to a set X. Just as the bijection f can be thought of as simply “renaming” each v  V with its counter- part in X, so then can f be used to “carry” the vector space structure of V onto the set X; specifically: THEOREM 4.16 Let f be a bijection from a vector space V to a set X. With addition and scalar multiplication on X defined as follows: (*) (**) –1 –1 –1 x1  x2 = ffx1 + f x2 and rx = frfx Go back to V and and sum in V. Similarly Then carry the sum back to X. the set X evolves into a vector space. Moreover f itself turns out to be an isomorphism from the vector space V to the vector space X. PROOF: The set X is clearly closed with respect to both (*) and (**) operations. We will content ourselves by verifying the zero and inverse axioms of Definition 2.6, page 40: Zero Axiom. Let 0 be the zero vector in V. We show that f0 is the f0 turns out to zero vector in X. be the zero in X. For xX , let v be the vector in V such that fv = x . Then: f0  x ==f0  fv ff–1f0 + f –1fv ===f0 + v fv x

fv– turns out to be Inverse Axiom: For xX , let v be the vector in V such that the inverse of fv . fv = x . We show that f–v is the inverse of xf= v : . xf –v ==fv  f–v ff–1fv+ f –1fv–

==fv+ – v f0 the zero in X

In the exercises you are invited to verify that the remaining axioms of Definition 2.6 are also satisfied, thereby establishing the fact that X with the above specified addition and scalar multiplication is indeed a vector space. We now show that the given bijection f from the vector space V to the above vector space X is an isomorphism. Actually, since f was given to be a bijection, we need only establish the linearity of f. Let’s do it:

For v1 v2  V , let x1 ==fv1 and x2 fv2 . Then, for any r   : frv1 + v2 ==rx 1  x2 rfv1 + fv2 by (*) and (**) 4.3 Isomorphisms 143

EXAMPLE 4.11 Let 2 denote the Euclidean two-space, and let X be the set Xxy=  xy  R . (a) Show that the function f: 2  X given by fxy = x – 1 xy+ is a bijection, and find its inverse f –1: X  2 . (b) Determine the vector space structure on X induced by the function f, as is described in Theorem 4.16. (c) Identify the zero in the above vector space X, and the inverse of the vector xy  X .

SOLUTION: (a) f is one-to-one:

fx1 y1 = fx2 y2

 x1 – 1 = x2 – 1 x1 = x2 x1 – 1 x1 + y1 = x2 – 1 x2 + y2   x1 + y1 = x2 + y2 y1 = y2

f is onto: Let xy  X . Then: fx+ 1 yx–1– ==x + 11–  x ++1 yx–1– xy From the above we can easily see that: f –1xy = x + 1 yx–1– (b) Theorem 4.16 assures us that the set X achieves “vector space- hood” when it is augmented with the following operations:

x1 y1  x2 y2 = fx1 + 1 y1 –1x1 – + x2 + 1 y2 –1x2 –

= fx1 ++x2 2 y1 + y2 –2x1 – x2 – fxy = x – 1 xy+ : = x1 ++x2 1 y1 + y2 rxy  = frx+ 1 – x + y – 1 ==frx+ r – rx + ry– r rx+ r – 1 ry

(c) Theorem 4.16 assures us that fxy = x – 1 xy+ is a linear map (in fact an isomorphism). That being the case, f00 = –10 must be the zero vector in X. Less there be any doubt: –10  xy = ff–1–10 + f –1xy –1 f xy = x + 1 yx–1– : = f 00 + x + 1 – x + y – 1 ==fx+ 1 – x + y – 1 xy fxy = x – 1 xy+ 144 Chapter 4 Linearity

As for the inverse of xy  X : –xy ==ff– –1xy fx–+ 1 yx–1– –1 f xy = x + 1 yx–1– : ==fx–1–  – y ++x 1 –2x –  –y

fxy = x – 1 xy+

Let’s challenge the above formula with the vector 32  X . The formulas tells us that –332 ==–2–2 – –25 – . If that is correct, then 32  –25 – has to be the zero vector –01 , and it is: 32  –25 – = ff–132 + f –1–25 –

===f42 – + –24 f00 –01

CHECK YOUR UNDERSTANDING 4.17

Let 3 denote the Euclidean three-space, and let X be the set X=  xyz xy  R . (a) Show that the function f: 3  X given by fxyz = 2xx + z – z is a bijection, and find its inverse. (b) Determine the vector space structure on X induced by the func- tion f, as is described in Theorem 4.16. (c) Identify the zero in the above vector space X, and the inverse of Answer: See page B-17. the vector xyz  X .

It can be shown that there exists a bijection from n to R for any positive integer n. Consequently:

THEOREM 4.17 Every Euclidean vector space n (and there- fore every finite dimensional vector space) sits (isomorphically) in the set R of real numbers. PROOF: A direct consequence of Theorem 4.16. 4.3 Isomorphisms 145

EXERCISES

Exercises 1-6. Determine if the given linear function f is a bijection. If so, find its inverse f –1 and show directly that it is also linear. 1.f:  , where fx = –5x .

2.f:  2 , where fx = xx – .

3.f: 2  2 , where fab = –2ba .

2 4.f:   P1 , where fab = ax+ b .

3 2 5.f:   P2 , where fabc = bx + cx– a .

2 ab 6.f:   M22 , where fab = . ab+ ab–

Exercises 7-17. Determine if the given function is an isomorphism. 7.f:  , where fx = –5x . 8.f:  , where fx = x + 1 .

2 9.f:   P1 , where fab = ab+ x .

2 10.f:   P1 , where fab = ax+ b .

3 2 11.f:   P2 ,where fabc = cx + bx– a.

12.f: P2  P2 , where fpx = px+ 1 .

13.f: P2  P3 , where fpx = xp x .

3 14.f: P2   , where fpx = p1  p2  p3 .

15.f: P3  P3 , where fpx = px+ p1 .

p1 p2 16.f: P3  M22 , where fpx = . p3 p4

4 2ac 17.f:   M22 , where fabcd = . db 146 Chapter 4 Linearity

Exercises 18-22. (Calculus dependent) Determine if the given function is an isomorphism.

18. f: P2  P2 , where fpx = px .

19. f: P2  P2 , where fpx = px + px .

20. f: P2  P2 , where fpx = px .

21.f: P2  P2 , where fpx = 2px– 3 .

1 22.f: P2  P2 , where fpx = px+ pxdx . 0

Exercises 23-24. As is done in Example 4.11, show that the given function f is a bijection and find its inverse. Determine the vector space structure on the given set X induced by f and identify the zero and the inverse of the vector xy  X in the resulting space X.

23. f: 2  X = xy xy  R given by fxy = xy++3 x – 4 .

24. f: 3  X = xyz xyz   R given by fxyz = x +21y –32 z + 4 .

25. Show that if the functions f: XY and g: YZ have inverses, then the function –1 –1 –1 gf: XZ also has an inverse and that gf = f g .

26. For r   , let fr: VV be given by frv = rv . For what values of r is fr an isomor- phism? 27. For v a vector in the space V let f : VV be given by f v = vv+ . 0 v0 v0 0 (a) Show that f is a bijection. v0 (b) Give necessary and sufficient conditions for f to be an isomorphism. v0 28. Find a specific isomorphism from Vax= 3 ++ab+ x cabc  to 3 .

29. Show that the vector space + of Example 2.4, page 46, is isomorphic to the vector space of real numbers,  .

30. Find an isomorphism between the vector space of Example 2.5, page 47 and 2 .

31. Suppose that a linear transformation T: VW is one-to-one, and that v1v2  vn is a linearly independent subset of V. Show that Tv1 Tv2 Tvn is a linearly indepen- dent subset of W. (In particular, the above holds if T is an isomorphism.)

32. Suppose that a linear transformation T: VW is onto, and that v1v2  vn is a span- ning set for V. Show that Tv1 Tv2 Tvn is a spanning set for W. (In particular, the above holds if T is an isomorphism.) 4.3 Isomorphisms 147

33. Prove that a linear transformation T: VW is an isomorphism if and only if for any given basis v1v2  vn for V, Tv1 Tv2 Tvn is a basis for W. 34. Let V be a vector space of dimension n, and let LV  be the vector space of linear trans- formations from V to  (see Exercise 35, page 122). Prove that LV  is also of dimen- sion n and is therefore isomorphic to V. (The space LV  is called the of V.)

Suggestion: For v1v2  vn a basis for V, show that T1T2  Tn is a basis for

 1ifij= LV  , where Ti is the linear transformation given by: Tivj =  .  0ifij 35. Let V be a vector space of dimension n, and let W be a vector space of dimension m. Let LVW be the vector space of linear transformations from V to W (see Exercise 35, page 122). Prove

that LV   Mmn .

Suggestion: For v1v2  vn a basis for V, and w1w2  wn a basis for W, show that Tij is a basis for LVW , where Tij is the linear transformation given by:

 wj if ij= Tijvj =  .  0ifij Exercises 37-39. () A relation on a set X is a set Sab=  ab  X . If ab  S , then we will say that a is related to b, and write ab . An equivalence relation on a set X is a relation which satisfies the following three properties: (i) Reflexive property: aa . (ii) Symmetric property: If ab , then ba . (iii) Transitive property: If ab and bc , then ac . 36. A partition of a set X is a collection of mutually disjoint (nonempty) subsets of X whose union equals X. (In words: a partition breaks the set X into disjoint pieces.) (a) Let  be an equivalence relation of a set X, and for any given xX define the of x to be the set of all elements of X that are related to x: x = x  Xx x . Prove that x xX is a partition of X. (b) Let S be a partition of a set X. Prove that there exists an equivalence relation  on X such that the Sx= xX , where x = x  Xx x . a c 37. Show that the relation defined by ---  --- if and only if ad= bc is an equivalence relation b d on the set Q of rational numbers (“fractions”). 38. Show that the relation VW if the vector space V is isomorphic to the vector space W is an equivalence relation on any set of vector spaces. 148 Chapter 4 Linearity

PROVE OR GIVE A COUNTEREXAMPLE

39. (a) If f: XY is an onto function, then so is the function gf: XZ onto for any function g: YZ .

(b) If g: YZ is an onto function, then so is the function gf: XZ onto for any func- tion f: XY .

40. (a) Let f: XY and g: YZ . If gf: XZ is onto, then f must also be onto. (b) Let f: XY and g: YZ . If gf: XZ is onto then g must also be onto.

41. (a) If f: XY is a one-to-one function, then so is the function gf: XZ one-to-one for any function g: YZ .

(b) If g: YZ is a one-to-one function, then so is the function gf: XZ one-to-one for any function f: XY . 42. If T: VW and L: VW are isomorphisms, then TL= .

43. If T: VW is an isomorphism, and if a  0 , then Ta: VW given by Tav = aTv is also an isomorphism.

44. Let T: VW and L: WZ be linear. If gf: VZ is an isomorphism, then T and L must both be isomorphisms. 45. If T: VW and L: VW are isomorphisms, then so is the function TL+ : V  W given by TL+ v = Tv + Lv an isomorphism. 149 Chapter 4 Linearity

CHAPTER SUMMARY

LINEAR A function T: VW from a vector space V to a vector space W is TRANSFORMATION said to be a linear transformation if for all vw  V and r   : Tvw+ ==Tv + Tw and Trv rTv

The two conditions for T: VW is linear if and only if: linearity can be incor- Trvw+ = rTv + Tw porated into one state- ment. for all vw  V and r   . The above result can be Let T: VW be linear. For any vectors v1v2  vn in V, and any extended to encompass scalars a1a2  an : n-vectors and scalars. Ta1v1 ++ anvn = a1Tv1 ++anTvn Linear transformations If T: VW is linear, then: map zeros to zeros and T0 = 0 and T–v = –Tv inverses to inverses. A linear transformation Let V be a finite dimensional space with basis v1v2  vn . If is completely deter- T: VW and L: VW are linear maps such that Tv = Lv mined by its action on a i i basis. for 1 in , then Tv = Lv for every v  V . A method for construct- Let v1v2  vn be a basis for a vector space V, and let ing all linear transfor- w w  w be n arbitrary vectors (not necessarily distinct) in a mations from a finite 1 2 n dimensional vector vector space W. There is then a unique linear transformation space to any other vec- L: VW which maps vi to wi for 1 in ; and it is given by: tor space. La1v1 ++ anvn = a1w1 ++ anwn The composition of lin- If T: VW and L: WZ are linear, then the composition ear maps is linear. LT: VZ is also linear. KERNEL Let T: VW be linear. The kernel (or null space) of T is denoted by KerT and is defined by: KerT = v  VTv = 0 IMAGE The image of T is denoted by ImT and is defined by: ImT = w  WTv = w for some v  V

Both the kernel and Let T: VW be linear. Then: image of a linear transformation are sub- KerT is a subspace of V spaces. ImT is a subspace of W Chapter Summary 150

NULLITY Let T: VW be linear. The dimension of KerT is called the nul- lity of T, and is denoted by nullityT .

RANK The dimension of ImT is called the rank of T, and is denoted by rankT .

The Dimension Let V be a vector space of dimension n, and let T: VW be linear. Theorem. Then: rankT + nullityT = n

ONE-TO-ONE A function f from a set A to a set B is said to be one-to-one if: fa==fa  a a

ONTO A function f from a set A to a set B is said to be onto if for every bB there exist aA such that fa= b .

A linear transformation T: VW is one-to-one if and only if KerT = 0 . A linear transformation T: VW is onto if and only if ImT = W .

BIJECTION A function f: A  B that is both one-to-one and onto is said to be a bijection.

The composite of bijec- If f: A  B and g: BC are bijections, then the composite function tions is again a bijec- gf: AC is also a bijection. tion.

INVERSE FUNCTION The inverse of a bijection f: XY is the function f –1: YX such that f –1 f x = x for every x in X, and ff –1 y = y for every y in Y.

The inverse of a linear If the bijection T: VW is linear, then its inverse, T –1: WV is bijection is again lin- also linear. ear.

ISOMORPHISM A bijection T: VW that is also a linear transformation is said to be an isomorphism from the vector space V to the vector space W. If there exists an isomorphism from V to W we then say that V and W are isomorphic, and write: VW .

Every vector space is isomorphic to itself. If V is VV isomorphic to W, then W is isomorphic to V. If V is VW  WV isomorphic to W, and W is isomorphic to Z, then V is isomorphic to Z. VW and WZ  VZ All n-dimensional vector spaces are isomorphic to If dimV = n , then V  Rn . Euclidean n-space. 5.1 151

5 CHAPTER 5 MATRICES AND LINEAR MAPS It turns out that linear transformations can, in a sense, be represented by matrices. Such representations are developed and scrutinized in Sec- tions three and four — a development that rests on the matrix theory presented in the first two sections of the chapter.

§1. MATRIX MULTIPLICATION

The matrix space Mmn of Theorem 2.2, page 42, comes equipped with a scalar multiplication. We now turn our attention to another form of multiplication, one that involves a pair of matrices (instead of a sca- lar and a matrix). Left to one’s own devices, one would probably define the product of matrices in the following fashion: Take two matrices of equal dimension, and simply mul- As with: 23 53 = 10 9 tiply corresponding entries 54 06 024 to obtain their product.

The above might be “natural,” but as it turns out, not very useful. Here, as you will see, is a most useful definition: In general, we will use: A = a a DEFINITION 5.1 If Amr = aij and Brn = bij , then: mn ij or ij mn MATRIX to denote an m by n matrix Amr Brn ==Cmn cij with entries a . ij MULTIPLICATION where:  cij = ai1b1j ++++ai2b2j ai3b3j airbrj Using Sigma notation, we have: r

cij = aibj

b  1 j ai1air   = 1 b rj th IN WORDS: To get cij of CAB= , run across the i row of A and r down the j th column of B, multiplying and adding along the way (see c = a b ij  i j margin).  = 1 Note: The above is meaningful only if the number of columns of the matrix on the left equals the number of rows of the matrix on the right. 152 Chapter 5 Matrices and Linear Maps

EXAMPLE 5.1 Find the product CAB= , if: 1235 201 A == and B 2043 524 1532

SOLUTION: Since the number of columns of A equals the number of rows of B, the product AB is defined:

A2  3 B3  4 = C24

same

Below, we illustrate the process leading to the value of c11 (run

across the first row of A and down the first column of B), and for c23 (run across the second row of A and down the third column of B):

1 235 123 5 201 3 201 3 2 043 == 204 3 524 524 35 1 532 153 2 21 ++02 11 53 ++24 43 At this point, we are confident that you can find the remaining entries:

1235 201 3 9912 2043 = 524 13 30 35 39 1532 In any event: GRAPHING CALCULATOR GLIMPSE 5.1

CHECK YOUR UNDERSTANDING 5.1

(a) Perform the product AB for: 33 37 35 Answer: (a) 30 26 64 54 36 A ==42 and B 35 (b) Number of columns 90 in A does not equal the number of rows in B. (b) Explain why the product BA is not defined. 5.1 Matrix Multiplication 153

In some ways, matrices behave differently than numbers. For one thing, even when both products AB and BA are defined, as is the case when the matrices A and B are square matrices of the same dimension, those products need not be equal: 12 56 19 22 56 12 23 34 Matrix multiplication is == while not commutative. 34 78 43 50 78 34 31 46

Some familiar multiplication properties do however carry over to matrices: THEOREM 5.1 Assuming that the matrix dimensions are such that the given operations are defined, we have: (i) AB+ C = AB+ AC Distributive Properties { (ii) AB+ CACBC= + (iii) Associative ABC= AB C Properties { (iv) rAB==rA BArB

PROOF: We establish (i): AB+ C = AB+ AC, and invite you to verify the rest in the exercises. Let: The properties of this A = a B = b C =c theorem are not particu- mr ij rn ij rn ij larly difficult to estab- lish. The trick is to Dmn ==AB+ C dij Emn ==AB+ AC eij carefully keep track of the entries of the matri- We need to show that d = e . Let’s do it: ces along the way, ij ij r distribute r r

dij == aitbtj + ctj  aitbtj +  aitctj =eij t = 1 t = 1 t = 1 down the down the across the th the across run across the i row of A th and down the j column of BC+ j j th i th i th th column of column of colum of colum of A A C B

POWERS OF SQUARE MATRICES Why we are restricting this 2 discussion to square matrices? If A is a square matrix, what should A represent? That’s right: Because: A2 = AA. In a more general setting, we have: 2 21 35 = ? DEFINITION 5.2 For A a square matrix: 42 POWERS A2 = AA A3 ==AAA and An AAn – 1 154 Chapter 5 Matrices and Linear Maps

While the definition of An mimics that of an for aR , not all of the familiar properties of exponents hold for matrices, even when the matrix expressions are well-defined. The property ab n = anbn , for example, does not carry over to matrices: 2 2 12 22 42 42 42 20 8 == = 02 10 20 20 20 84

2 2 12 22 12 12 22 22 while: =  02 10 02 02 10 10

==16 64 18 16 04 22 88

In  = M , ab= ba . 2 2 2 11 12 22 12 22 Why not for M22 ? And we see that:  02 10 02 10

The following familiar properties are, however, a directly consequence of Definition 5.2: THEOREM 5.2 For A a square matrix, and positive integers n and m: AnAm ==Anm+ and An m Anm

CHECK YOUR UNDERSTANDING 5.2

PROVE OR GIVE A COUNTEREXAMPLE:

For any two matrices AB  M22 : Answer: See page B-18. AB+ 2 =2A2 ++AB B2

COLUMN AND ROW SPACES

DEFINITION 5.3 The columns space of a matrix AM mn , m COLUMN AND is the subspace of  spanned by the col- umns of A. ROW SPACE The row space of a matrix AM mn , is the subspace of n spanned by the rows of A. The following result may be a bit surprising in that the rows and col-

umns of the matrix AM mn need not even live in the same Euclid- ean spaces: The rows of A are in n while its columns are in m : 5.1 Matrix Multiplication 155

THEOREM 5.3 The dimension of the column space of any matrix AM mn equals the dimension of its row space. PROOF: We know, from Theorem 3.13, page 103, that the dimension

of the column space of AM mn equals the number of leading ones in rrefA . To see that the dimension of the row space of A is also equal to the number of leading ones in rrefA we reason as follows: Let B be a matrix obtained by performing any of the three ele- mentary row operations on A. Since each row in B is a linear combination of the rows in A, and vice version, the space spanned by the rows of A equals that spanned by the rows of B . Since rrefA is derived from A through a sequence of ele- mentary row operations, the row space of A equals that of rrefA , which is easily seen to be the non-zero rows of rrefA ; each of which contains a leading one.

DEFINITION 5.4 For AM mn , the rank of A, denoted by RANK OF A MATRIX rankA is the common dimension of the row and column space of AM mn .

EXAMPLE 5.2 Find a basis for both the row and the column 39–15–6 space of A = –32421 – . 26–2–6–4

39–15–6 13–02–2 rref SOLUTION: A = –32421 – 00110. 26–2–6–4 00000 A basis for the row space of A: 13– 0.– 2 2  00110 . A basis for the column space of A: 31– 2  12– 2 (see Theorem 3.13, page 103) While Theorem 3.13 assures us that he columns of A associated with the leading one columns of rrefA constitute a basis for the column space of A, the rows of A associated with the leading one rows of rrefA need not be a basis for the row space of A. A case in point: 156 Chapter 5 Matrices and Linear Maps

CHECK YOUR UNDERSTANDING 5.3

Find a basis for the column and the row space of: 25–4 3 A = –104 –68– . Answer: 240 –   5101 –  012– 4

SYSTEM OF EQUATIONS REVISITED As you know, a system of equations can take the form of an augmented matrix multiplication matrix 24– 4 x 6 matrix [see page 3, and Figure (a) and (b) below]. That system can also 26 4 y = 0 be represented as the product of its coefficient matrix with a (column) 11 2 z –2 variable matrix [see Figure 5.1(c) and margin]. 2x + 4y – 4z 6 2x +64y – 4z =  24–6 4 24– 4 x 6 2x ++6y 4z = 0  2x ++6y 4z = 0 26 4 0 26 4 y = 0 xy++2z –2  xy++2z = – 2 11 2– 2 11 2 z –2 2x +64y – 4z =   System of Equations Augmented Matrix Matrix Product Form 2x ++6y 4z = 0  (a) (b) (c) xy++2z = – 2 Figure 5.1

The next result tells us that the product of AM mn with a column

variable matrix Xn  1 is a linear combination of the columns of A. Spe- cifically: C C C 1 2 n THEOREM 5.4 a11 a12 a1n x1 Let Amn = aij . For each 1 j n let a a a x 21 22 2n 2 = C = a (the jth “column matrix” of A). j ij m  1 Then, for any Xx= i : am1 am2 amn xn n  1 n a a a 11 12 1n AX= x C (see margin). a a a  i i x 21 ++x 22 x 2n 1 2 n i = 1 n am1 am2 amn PROOF: For BAXb== , and Dx==C d : C1 C2 Cn i m  1  i i i m  1 i = 1 m m b ==a x = x a d i  ij j  j ij i j = 1 j = 1 5.1 Matrix Multiplication 157

Consider the following system of equations AX= B : A X B x 2134 b1 2x +++1y 3z 4w b1 y 183– 5 = b  1x ++8y 3z – 5w = b z 2 2 12–16 b 1x –12y ++z 6w b w 3 3 Invoking Theorem 5.2, we find that:

2 1 3 4 b1

x 1 +++y 8 z 3 w –5 = b2

1 –2 1 6 b3 It follows that the given system AX= B is consistent if and only if

b1

b2 is in the column space of A.

b3 In general: THEOREM 5.5 A system of linear equations AXB= is consis- tent if and only if B is in the column space of A.

CHECK YOUR UNDERSTANDING 5.4 Prove that the solution set of any homogeneous system of m Answer: See page B-18. equations in n unknowns is a subspace of n .

FROM MATRICES TO LINEAR TRANSFORMATIONS While matrix spaces only come equipped with vector addition and sca- Note that: lar multiplication, matrix multiplication can serve to define a linear map Amn Xnz  Mmz between such spaces — providing their dimensions “match up:”

THEOREM 5.6 For AM mn and any positive integer z, the

map TA:Mnz  Mmz given by TAB = AB is linear.

PROOF: For B1 B2  Mnz and r   Theorem 5.1(iv):

TArB1 + B2 ==ArB1 + B2 ArB1 + AB2 Theorem 5.1(iv): ==rAB1 + AB2 rTAB1 + TAB2 158 Chapter 5 Matrices and Linear Maps

For notational convenience we will let the symbol n denote the space Mn  1 (“vertical n-tuples”). We then have:

Note that X is a is a THEOREM 5.7 n m n  1 For AM mn the map TA:    given “vertical n-tuple,” and that by T X = AX is linear. TAX is a vertical m-tuple. A PROOF: Simply set z = 1 in Theorem 5

DEFINITION 5.5 For A  Mmn , the null space of A, denoted NULL SPACE by nullA , is the set: nullA = X  n AX = 0

THEOREM 5.8 n The null space of AM mn is a subspace of  . PROOF: Follows from Theorem 4.8(a), page 124, and the fact that:

nullA = KerTA Note: The dimension of the null space of A is called the nullity of A Fair terminology, in that nullA = KerTA .

EXAMPLE 5.3 Find a basis for the null space of the matrix: 2134 A = 18310– 12–16 SOLUTION: By definition, the null space of A is the solution set of the homogeneous system of equations: 2134x 0 y 18310– = 0 z 12–16w 0 x y z w x y z w 1 0 --7------14- 2134 5 5 rref 1 8 From: 18310– 0 1 --- –--- 5 5 12–16 0000 5.1 Matrix Multiplication 159

vertical 3-tuple 7 14 1 8 nullA = –---a – ------b –---a + ---bab ab  R We see that:  5 5 5 5 = –147a – ba – + 8b5a 5b ab  R

Letting a = 1 and b = 0 we obtain the vector –71–50 . Letting a = 0 and b = 1 we arrive at the vector –14 8 0 5 . The above two vectors are clearly independent, and also span nullA : –7a – 14ba – + 8b5a 5b  = a–71–50+ b–14805 It follows that –71–50 –14805 is a basis for nullA .

We arrived at the two vectors in the basis for nullA in the above example by setting each of the two free variables to 1 and the other free variable to 0. Generalizing:

THEOREM 5.9 The nullity of AM mn equals the number of free variables in rrefA .

CHECK YOUR UNDERSTANDING 5.5

Find a basis for:  21 3 0 null14–7 2–  Answer: 11– 11 30 1– 2

You get to draw the final curtain of this section: CHECK YOUR UNDERSTANDING 5.6

n m Let AM mn and let TA:    be the linear map given by TAX = AX. Show that: Answer: Seepage B-18. nullityA = nullityTA and rankA = rankTA 160 Chapter 5 Matrices and Linear Maps

EXERCISES

Exercises 1-5. Perform the given matrix operations.

12 12 23 230 1.01– 2. 01– –41 –401 13 13

010 100 100 010 3. 200 002 and 002 200 003 030 030 003

12 12 12 13031 211 13 031 13 211 4. + 13 and 13 + 13 22123 420 22 123 22 420 01 01 01 Exercises 5-8. Find a basis for the null space of A. Determine rankA along with a basis for the column and row space of A.

321 5 201 1101 11–55 5. A = 10 6 1 10 6. A = 31–0 7. A = –011 –4 8. A = 10 3 2 42–0 1 633 53–7 2 01–2–1– 142 52–2 1 14 7 2

9. (a) Show that each column of 25 371 (as a vertical two-tuple) is a linear combination 13 204 of the columns of 25 . 13 (b) Show that each row of 25 371 (as a horizontal three-tuple) is a linear combination 13 204 of the rows of 371 . 204

th 10. Let AM mn and let Xi  Mn  1 be the column matrix whose i entry is 1 and all other th entries are 0. Show that AXi is the i column of A, for 1 in .

r 0 2 2 11. (a) (Dilation and Contraction) Let A = for r  0 . Show that TA:    maps 0 r every point in the plane to a point r times a far from the origin.

10 2 2 (b) (Reflection about the x-axis) Let A = . Show that TA:    reflects every 01– point in the plane about the x-axis. 5.1 Matrix Multiplication 161

2 2 (c) Find the matrix A for which TA:    reflects the point xy to the point twice the length of xy and reflected about the y-axis.

TAv = x y (d) ( about the Origin) Let A = cos –sin . Show that . r sin cos v = xy 2 2 r . T :    rotates the vector xy by  in a counterclock-  y A  wise direction. x Suggestion: Start with x = rcos+ and y = rsin+ . 12. Show that for any given linear transformation T: n  m there exists a unique matrix AM mn such that TX= AX .

13. Let AB  Mmn . Prove that if AC= BC for every CM n  1 , then AB= .

14. Determine all AM 22 such that AB= BA for every BM 22 . 15. Prove Theorem 5.1(ii). 16. Prove Theorem 5.1(iii). 17. Prove Theorem 5.1(iv).

18. A square matrix Aa= ij nn for which aij = 0 if ij is said to be a diagonal matrix. n Show that if Aa= ij nn is a diagonal matrix and if Xx= i   is a column matrix, 200 9 18 n then AX=  aiixi  . For example: 050 6 = 30 . 007 –1 –7 T 19. The of a matrix Aa= ij  Mmn is the matrix A = aij  Mnm , where aij = aji . In other words, the transpose of A is that matrix obtained by interchanging the rows and columns of A. (a) Show that for any matrix A, AT T = A . (b) Show that for any matrix A and any scalar r, rA T = rAT . T T T (c) Show that for any AB  Mmn , AB+ = A + B . T T T (d) Show that for any AB  Mmn , AB– = A – B . T T T (e) Show that for AM nn , AA = A A . T T T (f) Show that for AM mn and BM nr , AB = B A .

20. A square matrix A is symmetric if the transpose of A equals A: AT = A (see Exercise 19).

(a) Show that the set of symmetric matrices in the space Mnn is a subspace of Mnn .

(b) Let AB  Mmn be symmetric matrices. Show that AB+ is symmetric. T (c) Let AM nn . Show that AA+ is symmetric.

T (d) Let AM nn . Show that AA is symmetric.

(e) Let AB  Mnn be symmetric. Show that AB is symmetric if and only if AB= BA 162 Chapter 5 Matrices and Linear Maps

21. A square matrix A is said to be skew-symmetric if AT = –A (see Exercise 19).

(a) Show that if the square matrix Aa= ij is skew symmetric, then each diagonal entry aii must be zero. (b) Show that the square matrix Aa= ij is skew symmetric if and only if aij = –aji for all i, j. T (c) Let AM nn . Show that AA– is skew symmetric. (d) Show that every square matrix A can be uniquely written as ASK= + , where S is a symmetric matrix and K is a skew symmetric matrix.

2 22. A matrix AM nn is said to be idempotent if A = A . (a) Find an idempotent 22 matrix A containing no zero entry. (b) Let A and B be idempotent square matrices of the same dimension. Prove that if AB= BA , then AB is idempotent.

k 23. A matrix AM nn is said to be nilpotent if A = 0 for some integer k. (a) Find a nilpotent 22 matrix A distinct from the . (b) Let A and B be two nilpotent matrices of the same dimension. Prove that if AB= BA , then AB is again nilpotent.

24. The sum of the diagonal entries in the matrix Aa= ij  Mnn is called the trace of A and is n

denoted by traceA : traceA ==a11 ++a22 ann  aii . i = 1 (a) Show that for any square matrix A: traceA = traceAT (see Exercise 19). (b) Show that for any two square matrices A and B of the same dimension: TraceAB+ = TraceA + TraceB (c) Show that for any two square matrices A and B of the same dimension: TraceAB = TraceBA Exercises 25-30. (PMI) Determine An for the given matrix A. Use the Principle of Mathematical Induction to substantiate your claim.

25.11 26.11 27. 10 01 11 02

28. a 0 29.10 30. 1 a 0 b 21 01

31. (PMI) Show that if AB= BA , then for any positive integer n, AB n = AnBn .

n 32. (PMI) Let AM mm and BM ms . Show that if AX= X , then A XX= for every positive integer n.

33. (PMI) Show that if the entries in each column of AM nn sum to 1, then the entries in each column of Am also sum to 1, for any positive integer m. 5.1 Matrix Multiplication 163

n 34. (PMI) Show that if AM nn is a diagonal matrix, then so is A . (See Exercise 18.)

n 35. (PMI) Show that if AM nn is an idempotent matrix, then A = A for all integers n  1 . (See Exercise 22.)

n T T n 36. (PMI) Show that for any AM nn , and for any positive integer n, A = A . (See Exercise 19.)

37. (PMI) Let Ai  Mmn , for 1 in . Show that:  T T T  T A1 +++A2 An = A1 +++A2 An . (See Exercise 19.)

38. (PMI) Let Ai  Mnn for 1 in . show that:  T T T  T A1  A2 An = An  An – 1 A1 . (See Exercise 19.)

PROVE OR GIVE A COUNTEREXAMPLE

39. For AM mn and BM nr , if AB = 0 then either A = 0 or B = 0 . 40. Let A and B be two-by-two matrices with A  0 . If AB= AC , then BC= . 41. If A and B are square matrices of the same dimension, and if AB is idempotent, then AB= BA . (See Exercise 22.)

2 2 42. For all AB  M22 , A – B = AB+ AB– .

2 43. For any given matrix AM 22 , all entries in the matrix A are nonnegative.

44. For AM mn and BM nr , if A has a column consisting entirely of 0’s, then so does AB.

45. For AM mn and BM nr , if A has a row consisting entirely of 0’s, then so does AB.

46. For AM mn and BM nr , if A has two identical columns, then so does AB.

47. For AM mn and BM nr , if A has two identical rows, then so does AB.

48. For AM nn , BM nn AB = 0 is a subspace of Mnn .

49. For AM nn , BM nn AB= BA is a subspace of Mnn .

50. For AM nn , BM nn TraceAB  0 is a subspace of Mnn . (See Exercise 24.)

51. If A is a nilpotent matrix, then so is A2 . (See Exercise 23.) 52. A is idempotnet if and only if AT is idempotent. (See Exercise 22 and 19.) 164 Chapter 5 Matrices and Linear Maps

5

§2. INVERTIBLE MATRICES

Since all identity matrices The identity matrix of dimension n, denoted by In , is the nn are square we can get away by specifying just one of matrix that has 1’s along its main diagonal (top-left corner to bottom- its dimensions, as with: right corner), and 0’s elsewhere. In the event that the dimension n of

1 0 0 the identity matrix is understood, we may simply write I rather than In . I = 3 0 1 0 Just as 1  a = a and a  1 = a , for any number a, it is not difficult 0 0 1 to show that for any AM : I AA== and AI A . In particu- instead of I33 . mn m n

lar, for any square matrix AM nn : InAAI==n A .

INVERTIBLE MATRICES

DEFINITION 5.6 A square matrix AM nn is said to be INVERTIBLE invertible if there exists a matrix BM nn MATRIX such that: AB== BA I We will soon show, that a The matrix B is then said to be the inverse of matrix A can have but one –1 inverse. A, and we write BA= . If no such B exists, then A is said to be non-invertible, or singular.

EXAMPLE 5.4 Determine if the matrix A = –25 is 94– invertible. If it is, find its inverse.

ab SOLUTION: Let us see if there exists a matrix B = such that cd AB= I2 :

–25 ab = 10 94– cd 01

– 5a + 2c –25b + d = 1 0 9a –94c b – 4d 01 Equating corresponding entries leads us to the following pair of sys- tems of two equations in two unknowns:

– 5a +12c =  –25b +0d =    9a –04c =  9b –14d =  5.2 Invertible Matrices 165

If you take the time to solve the above systems, you will find that the one on the left has solution a = –2 , c = –--9- ; and the one on the 2

--5- right has solution b = –1 , and d = –2 ; which is to say: –12 – B = –--9- –--5- 2 2 Actually, you need not verify We leave it for you to verify that both AB and AB equal I and that there- that both AB and BA equal I, for if one does, then so must –12 – the other (Theorem 5.19). fore A–1 = (see margin). –--9- –--5- 2 2

EXAMPLE 5.5 Determine if the matrix A = 23 is –64 – invertible. If it is, find its inverse. SOLUTION: As we did in the previous example, we again try to find a ab matrix B = such that AB= I2 : cd

23 ab = 10 –64 – cd 01

2a +23c b + 3d = 10 –64a –4c –6b – d 01 Equating corresponding entries of the two matrices leads us to the following two systems of equations: 2a +13c =  2b +03d =    –64a –0c =  –64b –1d = 

The system of equation on Multiplying the equation 2a +13c = in the system on the left by 2, the right also has no solution. and adding it to the bottom equation, leads to an absurdity: 4a +26c = –64a –0c = add: 02=

We have just observed that there does not exist a matrix ab such that cd

23 ab = 10, which tells us that 23 is a singular matrix. –64 – cd 01 –64 – 166 Chapter 5 Matrices and Linear Maps

GRAPHING CALCULATOR GLIMPSE 5.2 Example 5.4 Example 5.5

CHECK YOUR UNDERSTANDING 5.7 Answer: Invertible with 32– inverse –2515  . Determine if is invertible. If it is, find its inverse. –3545  41– While a square matrix need not have an inverse (Example 5.5), if it does, then it is unique: THEOREM 5.10 An A has a unique inverse.

PROOF: For B and C inverses of A, we have: BBIBAC=====BA CICC

since C is an since B is an inverse of A inverse of A Here are three additional results pertaining to invertible matrices:

THEOREM 5.11 (i) If A is invertible, then so is A–1 , and: A–1 –1 = A (ii) If A is invertible and r  0 , then rA is also invertible, and: 1 rA –1 = ---A–1 r This is sometimes refered (iii) If A and B are invertible matrices of the to as the same dimension, then AB is also invertible, shoe-sock theorem Can you guess why? and: AB –1 = B–1A–1

PROOF: (i) We must suppress the temptation to appeal to the familiar exponent rule a–1 –1 ==a–1 –1 a , for we are not dealing with matrices and not numbers. What we do, instead, is to turn to Defini- tion 5.6 which tells us that: 5.2 Invertible Matrices 167

If A–1 = A–1 for some matrix in the box, then the matrix A–1 is invertible, and the matrix in the box is its inverse: A–1 –1 = Now put A in the box:

A–1 AA== A–1 I

We have shown that A–1 –1 = A .

(ii) Exercise 25. (iii) Returning to our “box game,” we see that: From the given condi- AB B–1A–1 ====ABB–1 A–1 AIA–1 AA–1 I tions, we know that B–1 and A–1 exist. What we and: B–1A–1 AB ====B–1A–1A BB–1IB B–1BI do here is to show that the –1 –1 product B A is, in –1 –1 –1 fact, the inverse of AB. The above shows that the inverse of AB: AB = B A . In words: The inverse of a product of invertible matrices is the product of their inverses, in the reverse order.

CHECK YOUR UNDERSTANDING 5.8 Use the Principle of Mathematical Induction to show that if

A1A2  An are invertible, then so is their product, and: Answer: See page B-19.  –1 –1 –1 –1 A1A2 An = An An – 1 A1 As you know, for any nonzero number a: 1 (i) a0 = 1 and (ii) a–n = ----- (for a  0 ) an While we can adopt the notation in (i) for invertible matrices, we can- 1 not do the same for (ii), since the expression --- is simply not defined A for a matrix A. We can however rewrite (ii) in the form a–n = a–1 n , and use that form to define A–n : DEFINITION 5.7 For A, an invertible matrix, and n a positive integer, we define: 0 –n A and A (i) A0 = I (ii) A–n = A–1 n 168 Chapter 5 Matrices and Linear Maps

ELEMENTARY MATRICES

Elementary row operations DEFINITION 5.8 An elementary matrix is a matrix that is were introduced on page 3. ELEMENTARY obtained by performing an elementary row MATRIX operation on an identity matrix.

Here are some examples of elementary matrices: 100 0 01 0 050 0 1 6 E1 = 10 0 E2 = E3 = 001 0 0 1 00 1 000 1 Interchange Add 6 times the I the first and Multiply the second second row of 2 second row of I row of by 5 to the first row of I 3 I4 2

CHECK YOUR UNDERSTANDING 5.9

Show that the above three elementary matrices, E1 , E2 and E3 are Answer: See page B-19 invertible, and that the inverse of each is itself an elementary matrix. The situation in the above Check Your Understanding box is not a fluke. You are invited to establish the following theorem in the exercises: THEOREM 5.12 Every elementary matrix is invertible, and its inverse is also an elementary matrix. Indeed: (a) If E is the elementary matrix obtained by interchanging rows i and j of the identity matrix I, then E–1 is the elementary matrix obtained by again interchanging rows i and j of I. (b) If E is obtained by multiplying row i of I by c  0 , then E–1 is obtained by multi- 1 plying row i of I by --- . c (c) If E is obtained by adding c times row i to row j of I, then E–1 is obtained by adding –c times row i to row j. Consider the row operation that is performed in the first and second line of Figure 5.2. As you can see in line 3, the matrix resulting from performing the elementary row operation on A coincides with the prod- uct matrix EA : 5.2 Invertible Matrices 169

multiply row 3 by -2 and add it to the first row

2103– 2R + R  R –4 –5 –8 1 3 1 1 A = 1326 1326

3341 3341 SAME

10 0– 2R + R  R 10 –2 3 1 1 I3 ==01 0 01 0 E 00 1 00 1

10 –22103 –4 –5 –8 1 EA ==01 01326 1326 00 13341 3341 Figure 5.2 The following theorem, a proof of which is relegated to the exercises, tells us that the situation depicted in Figure 5.2 holds in general: THEOREM 5.13 The matrix obtained by performing an elemen- tary row operation on a matrix AM mn equals that matrix EA , where E is the matrix obtained by performing the same elementary

row operation on the identity matrix Im .

001 CHECK YOUR UNDERSTANDING 5.10 Answer: (a) E = 010 (a) Switch the first and third rows of the matrix A of Figure 5.2 to 100 arrive at a matrix B, and then find the elementary matrix E such that EA= B . 001 (b) Multiply the second row of the matrix A of Figure 5.2 by 2 to (b) E = 0 2 0 arrive at a matrix B, and then find the elementary matrix E such 100 that EA= B . We remind you that two matrices are equivalent if one can be derived from the other via a sequence of elementary row operations (Definition 1.1, page 3). The following theorem asserts that if a matrix A is invert- ible, then every matrix equivalent to A is also invertible. THEOREM 5.14 For A and B are equivalent square matrices: A is invertible if and only if B is invertible.

PROOF: Since A and B are equivalent, there exist elementary matrices

E1E2  Es such that:  BE= s E2E1A 170 Chapter 5 Matrices and Linear Maps

Theorem 5.11 tells us that each Ei is invertible. Consequently, if A is invertible then so is B, for it is a product of invertible matrices. By sym- metry we also have that if B is invertible then so is A. You can quickly determine whether or not a matrix is invertible by looking at its row-reduced-echelon form:

THEOREM 5.15 AM nn is invertible if and only if:

rrefA = In

PROOF: If rrefA = In , then A and I are equivalent. Since I is invertible (with inverse I), A is also invertible (Theorem 5.15).

We establish the converse by showing that if rrefA  In , then A is not invertible:

Let rrefA = CI n . Since Cc= ij has less than n leading ones (otherwise it would be In ), its last row must consist entirely of zeros. This being the case, for any given nn matrix Dd= ij , the product matrix CD=  eij cannot be the identity matrix, as its lower-right-corner entry is not 1: n n

enn === cndn  0  dn 01  = 1  = 1 Since there does not exist a matrix D such that CD= I , rrefA = C is not invertible. It follows, from Theorem 5.155, that A is not invertible. The following theorem provides a systematic method for finding the inverse of an invertible matrix: THEOREM 5.16 If a sequence of elementary row operations reduces the invertible matrix A to I, then applying the same sequence of elementary row operations on I will yield A–1 .

PROOF: Let E1E2  Es be the elementary matrices correspond- ing to elementary row operations which take A to I:  Es E2E1AI= Since elementary matrices are invertible, their product is also invert- ible and we have (CYU 5.9):

–1 –1 –1 –1 –1 AE===sE2E1 IEsE2E1 E1 E2 Es Thus: 5.2 Invertible Matrices 171

–1 –1 –1 –1 –1 –1 –1 –1 –1 –1 –1 A ==E1 E2 Es Es E2 E1

==EsE2E1 EsE2E1I

We have shown that EsE2E1I is the inverse of the matrix A; to put it another way: The exact same sequence of row operations that trans- form A to I can also be used to transform I to A–1 . We now illustrate how the above theorem can effectively be used to find the inverse of an invertible matrix, by hand:

EXAMPLE 5.6 12 3 Determine if A = 25 3 is invertible. If it 10 8 is, find its inverse.

SOLUTION: We know that A is invertible if and only if rrefA = I . Rather than obtaining the row-reduced-echelon form of A, we will find the row-reduced echelon form of the matrix AI which is that matrix obtained by “adjoining” the 33 identity matrix to A: 12 3100 AI = 25 3010 10 8001 for that can feed two birds with one seed: First bird: A is invertible if and only if rrefAI = I (By Theorem 5.15) Second bird: If A is invertible, then rrefAI = IA–1 (By Theorem 5.17) Using the Gauss-Jordan Elimination Method (page 10), we have:

12 3100 12 3 100 10 9 5–0 2 – 2R1 + R2  R2 – 2R + R  R A I = 25 3010 01 –23 –10 2 1 1 – 1R + R  R 2R + R  R 01 –23–10 1 3 3 2 3 3 10 8001 02– 51–01 00 –51–21

–1R  R 10 9 5–0 2 10 0–169 40 3 3 – 9R + R  R 01 –23–10 3 1 1 = I A–1 3R + R  R 01 0 13–3 5– 3 2 2 00 1 5–1 2– 00 1 5–1 2–

Oh well: 172 Chapter 5 Matrices and Linear Maps

Answer: A is invertible CHECK YOUR UNDERSTANDING 5.11 with inverse: –557 – 6 101 2 –233 – 2 021 4 10–10 7 –8 Determine if A = is invertible. If it is, find its inverse. 111 0 –111 – 1 031 1 Theorem 5.16 says that a matrix A is invertible if and only rrefA = I . Here are some other ways of saying the same thing:

THEOREM 5.17 Let AM nn . The following are equivalent: (i) A is invertible. (ii)AX= B has a unique solution for

every BM n  1 . (iii)AX = 0 has only the trivial solution. (iv)rrefA = I . (v) A is a product of elementary matrices.

PROOF: We show i ii iii iv v i : i  ii : Let A be invertible. For any given equation AX= B , we have: AX= B A–1AX = A–1B A–1A XA= –1B IX= A–1B XA= –1B unique solution

ii  iii : If AX= B has a unique solution for any B, then AX = 0 has a unique solution. But AX = 0 has the trivial solution X = 0 . It follows that AX = 0 has only the trivial solution.

iii  iv : Assume that AX = 0 has only the trivial solution. Since A and rrefA are equivalent matrices, rrefA X = 0 has only the trivial solutions. This tells us that rrefA does not have a free variable, and must therefore have n leading ones. As such: rrefA = I . 5.2 Invertible Matrices 173

iv  v : If rrefA = I then:  Es E2E1AI= for some elementary matrices E1E2  Es . Since elementary matrices are invertible (Theorem 5.12), their product is also invertible and we have: –1 –1 –1 –1 –1 AE===sE2E1 IEsE2E1 E1 E2 Es

a product of elementary matrices (see CYU 5.8)

v  i : If A is a product of elementary matrices, then it is a product of invertible matrices, and is therefore invertible [CYU 5.9]. We defined a matrix A to be invertible if there exists a matrix B such that AB== BA I . As it turns out, B need only “work” on the left (or right side) of A:

THEOREM 5.18 Let AM nn . If there exists BM nn such that BA= I , then A is invertible and A–1 = B .

PROOF: Assume that BA= I , and consider the equation AX = 0 . Multiplying both side of that equation on the left by B we have: AX = 0 BAX= B0 IX = 0 X = 0 Since the equation AX = 0 has only the trivial solution, the matrix A is invertible (Theorem 5.18). Upon multiplying both sides of BA= I on the right by A–1 , we have: BA= I BAA–1 = IA–1 BI= A–1 BA= –1 In a similar fashion: Answer: See page B-20. CHECK YOUR UNDERSTANDING 5.12

Let AM nn . Show that if there exists BM nn such that AB= I , then A is invertible and A–1 = B . 174 Chapter 5 Matrices and Linear Maps

EXERCISES

Exercises 1-6. Find the inverse of the given invertible matrix A, and then check directly that AA–1 ==A–1AI.

1.20 2.12 3. 22 05 32 13

10 0 01 1 21 3 4.12 0 5.12 0 6. 12 3 12 3 01 2 20 4 Exercises 7-12. Determine if the given matrix A is invertible and if so, find its inverse.

7.23 8.32 9. –41 –64 – 31 520–

12 3 140 13 0 10. 23 2 11.22–3 12. 22– 3 33 3 13212 12 1 Exercises 12-15. For what values of a and b is the given matrix invertible?

13.a 3 14.a 2 15. ab 4 b b 1 1 ab

Exercises 16-18. (a) Find elementary matrices E1 and E2 such that E2E1IA= for the given matrix A. (b) Find two elementary row operations which will take A to I.

1 00 0 0 1 1 0 0 16.A = 03– 0 17.A = 01 0 18. A = 00 1 0 0 1 5 0 0 7 1 0 Exercises 19-24. Assume that A, B, and X are square matrices of the same dimension and that A and B are invertible. (a) Solve the given matrix equation for the matrix X.

(b) Challenge your answer in (a) using A = 12 and B = 21 . 01 31 19.2XA= AB 20. 2XA= BA 21. AXB–1 = BA 22.BXA= B2 23.BXB=  BAB 2 24. BXA=  AB BA –1

25. Prove Theorem 5.11(ii). 5.2 Invertible Matrices 175

26. Prove Theorem 5.12. 27. Prove Theorem 5.13. 28. Prove that if A is invertible, then A–2A–3 = A–5 .

29. Let AM mn . Prove that there exists an invertible matrix CM mm such that CA = rrefA .

30. Let A1A2  As be a linearly independent set of vectors in Mnn , and let AM nn

be invertible. Show that AA1AA2  AAs is linearly independent.

31. Let A1A2  An2 be a basis for Mnn , and let AM nn be invertible. Show that

AA1AA2  AAn2 is also a basis. 32. Show that a (square) matrix that has a row consisting entirely of zeros cannot be invertible. 33. Show that a (square) matrix that has a column consisting entirely of zeros cannot be invertible. 34. Show that if a row of a (square) matrix is a multiple of one of its other rows, then it is not invertible. 35. State necessary and sufficient conditions for a diagonal matrix to be invertible. (See Exercise 18, page 161.)

n 36. Prove that AM nn is invertible if and only if the rows of A constitute a basis for  .

n 37. Prove that AM nn is invertible if and only if the columns of A constitute a basis for  .

38. Prove that the transpose AT of an invertible matrix A is invertible, and that AT –1 = A–1 T . (See Exercise 19, page 161.) 39. Prove that if a symmetric matrix is invertible, then its inverse is also symmetric. (See Exer- cise 20, page 161.)

40. Prove that if AM nn is an idempotent invertible matrix, then AI= n . (See Exercise 22, page 162.) 41. Prove that every nilpotent matrix is singular. (See Exercise 23, page 162.)

42. (a) Prove that A = ab is invertible if and only if ad–0 bc  . cd

(b) Assuming that A = ab is invertible, find A–1 (in terms of a, b, c, and d). cd

2 43. Let AM nn be such that A – 2A +0I = . Show that A is invertible.

2 44. Let AM nn be such that A ++sA tI = 0 , with t  0 . Show that A is invertible. 176 Chapter 5 Matrices and Linear Maps

3 45. Let AM nn be such that A –23A +0I = . Show that A is invertible. 46. (PMI) Show that if A is invertible, then so is An for every positive integer n. 47. (PMI) Let A and B be invertible matrices of the same dimension with AB= BA . Sow that: (a)B–1 nA–1 = A–1B–1 n for every positive integer n. (b)AB –n = A–nB–n for every positive integer n.

PROVE OR GIVE A COUNTEREXAMPLE

48. If A is invertible, then so is –A , and –A –1 = –A–1 .

49. If A1A2  As is a linearly independent set in the vector space Mnn , and if AM nn is not the zero vector, then AA1AA2  AAs is linearly independent.

50. Let A be an nn invertible matrix, and BM nm . If AB = 0 , then B = 0 .

51. Let AM nn be invertible, and BM mn . If BA = 0 , then B = 0 52. If A and B are nn invertible matrices, then AB+ is also invertible, and AB+ –1 = A–1 + B–1 .

2 53. If AM nn and A = 0 , then A is not invertible. 54. If a square matrix A is singular, then A2 = 0 .

55. If A and B are nn matrices, and if AB is invertible, then both A and B are invertible.

56. If A and B are nn matrices, and if AB is singular, then both A and B are singular. 5.3 Matrix Representation of Linear Maps 177

5

§3. MATRIX REPRESENTATION OF LINEAR MAPS

The main objective of this section is to associate to a general linear transformation T: VW , where dimV = n and dimW = m , a matrix AM mn which can be used to find the value of Tv , for every v  V . The importance of all of this is that, in a way, the linear transfor- mation T is “linked to a finite object:” an mn matrix A. The notion of an ordered basis for a vector space plays a role in the current development. The only difference between a basis and an ordered basis is that, in the latter, the listing of the vectors is of conse- quence. For example, while 12  35 and 35  12 are one-and-the-same bases, they are not the same ordered bases (different ordering of the elements). Choosing an ordered basis for a vector space V of dimension n enables us to associate vertical n-tuple to each vector of V:

DEFINITION 5.9 Let  = v1v2  vn be an ordered COORDINATE basis for a vector space V, and let VECTOR v = c1v1 +++c2v2  cnvn We define the of v rela- tive  to be the column vector:

c1 c2 v  = 

cn

EXAMPLE 5.7 In Example 3.10, page 94, we showed that  = 130 204 012 is a basis for 3 . Find the coordinate vector of 1118 with respect to  .

SOLUTION: As is often the case, the problem boils down to that of solving a system of linear equations: If: 1118 = a130 ++b204 c012 , then: a b c a b c 100 2 a = 2 a ++2b 0c = 1 120 1 2  augS rref 1 1 S: 010–--- b = –---  1118 = –--1- 3a ++0bc= 11 30111 2   2 2 0a ++4b 2c = 8 042 8  001 5 c = 5 5 178 Chapter 5 Matrices and Linear Maps

CHECK YOUR UNDERSTANDING 5.13

–14 Find the coordinate vector of –112  3 with respect to: Answer: –38 74  = 130 204 012

Throughout this sec- tion the term “basis” THEOREM 5.19 If  = v1v2  vn is a basis for a vector will be understood to n mean “ordered basis.” space V, then the function L: V   given

We remind you that we are by Lv = v  (a vertical n-tuple) is linear. n using  to denote Mn  1 [Indeed, it is an isomorphism (Exercise 32).] n n

PROOF: For v =  aivi and v =  bivi in V, and r   , we have: i = 1 i = 1 n n n L rvv+  L rv+ v ==L rav + b v L ra + b v    i i  i i  i i i i = 1 i = 1 i = 1

ra1 + b1 ra1 b1 a1 b1 ==  +  =r  + 

ran + bn ran bn an bn

n n = rL a v + L b v = rL v + L v = rLv + Lv  i i  i i   i = 1 i = 1

Consider a linear transformation T: VW , with dimV = n and

dimW = m . If we fix a basis  = v1v2  vn for V and a basis

 (gamma) is the Greek  = w1w2  wm for W, then for every v  V we can determine letter c. the coefficient matrix v  of v with respect to  , as well as the coeffi-

cient matrix Tv  of Tv with respect to  . The following defini-

tion provides an important relation between v  and Tv  : 5.3 Matrix Representation of Linear Maps 179

DEFINITION 5.10 Let T: VW be a linear map from a vector Note the order of the two MATRIX REPRESEN- space V of dimension n to a vector space W subscripts in T . It kind  TATION OF A LIN- of dimension m. Let  = v1v2  vn of “looks backward,” EAR MAP since T “goes from  to and  = w1w2  wm be bases for V  .” As you will soon see, and W, respectively. We define the matrix however, the chosen order is the more suitable representation of T with respect to  and  for the task at hand, that th of representing a linear to be the matrix T   Mmn whose i map in matrix form. column is Tvi  .

The above definition looks intimidating, but appearances can be mis- leading. Consider the following example.

EXAMPLE 5.8 3 Let T:   P2 be the linear map given by: Tabc = 2ax2 ++ab+ xc

Determine the matrix representation T  of T with respect to the bases:  = 130 204 012 and  = 2x2 + x 3x2 – 1 x

SOLUTION: Definition 5.10 tells us that the first column of T  is the coefficient matrix of T130 with respect to  [entries are the values of a, b, c stemming from (1) below], the second column is T204  [values of a, b, c stemming from (2)], and the third column is T012  [values stemming from (3)]: (1): T130 ==2x2 ++4x 0 a2x2 + x ++b3x2 – 1 cx (2): T204 ==4x2 ++2x 4 a2x2 + x ++b3x2 – 1 cx (3): T012 ==0x2 ++1x 2 a2x2 + x ++b3x2 – 1 cx Equating coefficients in each of the above we come to the following three systems of equations:

0x2 ++1x 0 3x2 + 0x – 1 2 2x ++1x 0 a b c a b c 2302 2304 2300 (1): 1014 (2): 1012 (3): 1011 01–00 01–04 01–02 180 Chapter 5 Matrices and Linear Maps

We want the solutions of the above systems of three equations in the three unknowns a, b, c. Noting that the coefficient matrix of all three systems are one and the same, we can “compress” all three systems into a single matrix form: system (3) solutions of system (3) system (2) solutions of system (2) system (1) solutions of system (1) a b c 2302 4 0 100183 rref 18 3 1014 2 1 01004– –2  T  = 04– –2 T  01–00 4 2 00136– –2 36– –2 Figure 5.2

CHECK YOUR UNDERSTANDING 5.14

3 Let T: M22   be the linear map given by:

T ==ab bac + d cd Determine T  for: Answer: 33 72 11– 26 12–14 3 4   =   10–238 1 8  22 11 51 59 0 1 13 4 54 and  = 130 204 012 . To know the coordinates of a vector v in a finite dimensional vector space V with respect to a fixed ordered basis for that space is the same as knowing the vector v itself. This being the case, the following theo- rem tells us that the action of a linear transformation T: VW can, in effect, be realized by means of a matrix multiplication. THEOREM 5.20 Let T: VW be a linear map from a vector Note that the dimensions space V of dimension n to a vector space W of match up: dimension m. Let  = v1v2  vn and Tv  = T v      = w1w2  wm be bases for V and W, M M N n mn m    respectively. Then, for every v  V : 1 1 Tv  = T  v  m m PROOF: Noting that T  v : V   and Tv  : V   are Since  both linear, we need only show that T  vi  = Tvi  for each vi = 0v1 ++1vi ++ 0vn, 0 1 in (Theorem 4.6, page 115):

0 th v = th T v is the i column of T , namely: Tv (see i 1 i position  i   i  margin and Exercise 10, page 160). By Definition 5.10, 0 th Tvi  is also the i column of the matrix T  . 5.3 Matrix Representation of Linear Maps 181

EXAMPLE 5.9 Let T: 2  3 be the linear map given by: Tab = ab+ b 2a

Determine T  of T with respect to the bases:  = 12  30 and  = 100 110 111 and then show directly that:

T57  = T 57  SOLUTION: Proceeding as in Figure 5.2 of Example 5.8 we find

T  . Since T12 ==322 and T30 306 :

T  a b c 11133 1001 3 (*) 01120 rref 0100– 6 00126 0012 6 INCIDENTALLY, noting that the coefficient matrix The three vectors in  Solving the two systems of equations: of system (*) is identical 322 a100 ++b110 c111 =  to that of (**) we could 306 save a bit of time by  doing this Finding T57  . Since T57 = 12710 :

T57  11112 100 5 rref (**) 011 7 010– 3 00110 00110 T  Finding 57 : T57  

57 

7 10--- 135 rref2 2 207 01--1- 2 The two vectors in  We leave it for you to verifying that 7 5 13 --- 2 T57  = T 57 ; that is: –3 = 06– . --1- 10 26 2 182 Chapter 5 Matrices and Linear Maps

CHECK YOUR UNDERSTANDING 5.15 Referring to the situation described in CYU 5.15, verify directly that:

13 13 T= T Answer: See page B-20.  20  20  The next result is particularly nice — it tells us that a matrix of a com- posite of linear transformations is a product of the matrices of those transformations:

THEOREM 5.21 Let T: VW and L: WZ be linear V T W L Z   THE COMPOSITION maps, and let  , and  be bases for the LT THEOREM finite dimensional spaces V, W, and Z, respec- tively. Then: LT  = L T  LT  = L  T  PROOF: For any v  V we have: Theorem 5.20

LT v  = LT v 

==LTv  L Tv  =L T v  The result now follows from Exercise 34 which asserts that if A is a matrix such that LT v  = Av for every v  V , then the matrices LT  and A must be one and the same.

EXAMPLE 5.10 2 Let T:   M22 and L: M22  P2 be the linear maps given by:

Tab = aa+ b 02b and ab L= ax2 ++bx c+ d cd Show directly that:

LT  = L  T  For bases:  = 12  04

11 02 00 00  =  00 10 11 01  = x2x2 + x x2 ++x 1 5.3 Matrix Representation of Linear Maps 183

SOLUTION: Determining T  :

T04 = 04 08

T12 = 13 04

11 02 00 00 00 10 11 01

1 00 0 1 0 1 00 0 1 0 10 120034 010012 12  T  = 011000 0010–2 1– –21 – 001148 0001510 510

Determining L  : L L     L 01 00   L 11 00   10 02 00 11 = = = 0 0 =1 x x 0 2 1 1 1 2 1 x ++ x x x x 2 ++ 2 2 2 2 0 ++ 0 ++ ++ ++ ++ x x 2 0 1 1 x x x x x 1 2 1 0 0 0 1

1 11 1 0 0 0 1 000–00 2 02–00 rref 0111200 01011–1 2–  L  = 11–1 2– 0010121 00101 2 1 01 2 1

Determining LT  :

LT04 ===LT04 L04 0x2 ++4x 8 1 1 1 08 x x x 2 2 2 ++ ++ ++ 13 2

0 1 LT12 ===LT12 L x ++3x 4 1

x x 

x 04 0 0 1

1 11 1 0 100–4 2– –42 – 01134 rref 010–4 1–  LT  = –41 – 00148 001 4 8 48 And it does all fit nicely together, as you can easily check:

LT  L  T 

1 0 –42 – 02–00 1 2 –41 – = 11–1 2– –1 –2 48 01 2 1 5 10 184 Chapter 5 Matrices and Linear Maps

CHECK YOUR UNDERSTANDING 5.16

3 2 Let T:   P2 and L: P2   be the linear maps given by: Tabc = bx2 ++ax c and Lax2 ++bx c = aa ++ b c Show directly that:

LT  = L  T  for bases:  = 111 110 100  = x2x 2 Answer: See page B-21.  = 01  11

A proof of the following result is relegated to the exercises: We recall that I denotes n THEOREM 5.22 Let  be a basis for a vector space V of the identity matrix of dimension n. Then: dimension n, and that IV denotes the identity map I = I from V to V. V  n As might be expected: THEOREM 5.23 Let T: VW be an isomorphism. Let  and  be bases for V and W, respectively. Then: –1 –1 T  = T 

PROOF: We have:

V T W T – 1 V   

–1 T TI= V –1 T T  = IV  –1 Theorems 5.21 and 22: T  T  = I –1 –1 T  = T 

CHECK YOUR UNDERSTANDING 5.17

Show that if the matrix representation of T: VW with respect to any chosen basis  for V and  for W is invertible, then the linear Answer: See page B-21. map T is itself invertible (an isomorphism). 5.3 Matrix Representation of Linear Maps 185

EXERCISES

Exercises 1-3. Let T: 2  2 be the linear operator given by Tab = ab+2 b . Find 23  and T23  for the given (ordered) basis  . 1. = 10  01 2. = 01  10 3.  = 12  –2 1

Exercises 4-5. Let T: 3  3 be the linear operator given by Tabc = ab+  bac–  . Find 231  and T231  for the given basis  . 4. = 100 010 001 5.  = –101010 –211

2 2 Exercises 6-7. Let T: R  P2 be the linear map given by T ab = ax – bx + ab– . Find 12  and T12  for the given bases  and  . 6.  ==21  10  2x2 –x 2

7.  ==31 –  12  2x2 + 12 xx 2 ++x 1

3 2 ad– Exercises 8-9. Let T: P3  M22 be the linear map given by Tax+++bx cx d = . bc 2 2 Find x ++x 1  and Tx++x 1  for the given bases  and  .

11 01 00 00 8.  ==x32x2 x + 1 x – 1   11 11 11 01

11 01 00 00 9.  ==x3 + x2 x2 + xx + 11   11 11 11 01

3 2 Exercises 10-11. Let T:    be the linear map given by Tabc = ab+2 c . Find T  with respect to the given bases  and  , and show directly that T121  = T 121  . 10.  ==110 101 111  22  01 11.  ==011 101 110  01  22

2 2 Exercises 12-13. Let T:   P2 be the linear map given by T ab = ax – bx + ab– . Find T  with respect to the given bases  and  , and show directly that T12  = T 12  . 12.  ==22  01  xx2 x + 1 13.  ==12  21  xx2 + x x2 ++x 1 186 Chapter 5 Matrices and Linear Maps

Exercises 14-15. Let T: P1  P2 be the linear map given by Tpx = xp x + p0 . Show directly that T2x + 1  = T 2x + 1  for the given bases  and  . 14.  ==42 x  42 x 4x2 15.  ==x + 12 x + 3  xx2 + x x2 ++x 1

3 3 3 Exercises 16-17. Let I3:    be the identity map on  . Find I3  with respect to the given bases  and  , and show directly that I3121  = I3 121  16.  ==110 101 111  110 101 111 17.  ==110 101 111  101 111 110

ab 12 ab Exercises 18-19. Let T: M22  M22 be the linear map given by T = . Find cd 01 cd

T for the given bases  and  , and show directly that T 12 = T 12 .    21–  21– 

11 01 00 00 01 00 11 00 18.  ==    11 11 11 01 11 01 11 11

01 00 11 00 10 00 01 01 19.  ==    11 01 11 11 01 11 00 01

Exercises 20-22. Let T: 2  2 be the linear operator given by Tab = ab+2 b . Find T  with respect to the given basis  , and show directly that T13  = T 13  . 20. = 10  01 21. = 01  10 22.  = 12  –2 1

23. (Calculus Dependent) Let T: P2  P3 be the linear map given by Tpx = xp x and let D: P3  P2 be the differentiation linear function: Dpx = px . Determine the given 2 2 3 matrices for the basis  = 1xx of P2 , and the basis  = 1xx x of P3 .

(a) T  (b) D  (c) DT 

(d) TD  (e) TDT  (f) DTD 

24. (Calculus Dependent) Let V be the subspace of F spanned by the three vectors 1, sinx , and cosx . Let D: VV be the differentiation operator. Determine D  for  = 1 sinxx cos , and show directly that D52+ sinx  = D 52+ sinx  .

25. (Calculus Dependent) Let D: P3  P3 be the differentiation operator. Determine D  for 3 2 3 2 3 2  = 4x 3x 2x 1 , and show directly that D5x + 3x  = D 5x + 3x  . 5.3 Matrix Representation of Linear Maps 187

2 2 11 26. Find the linear function T:    , if T  = for  = 12  20 . 02

2 2 11 27. Find the linear function T:    if T  = for  = 12  20 and 02  = 11  12 . 10 2 28. Find the linear function T:   P2 if T  = 01 for  = 12  20 and 11  = x2 2x + 12 . 1111 0111 29. Find the linear function T: M22  M22 if T  = for 0011 0001

10 11 11 11 10 01 00 00  =  and  =  . 00 00 10 11 00 00 10 01

2 3 ab– 30. Let T:   M22 and L: M22   be the linear maps given by: Tab = –0a

ab and L= bac + d . Show directly that LT  = L  T  for bases: cd

11 02 00 00  = 12  01   =    = 111 110 100 00 10 11 01

2 3 3 31. Let T:    and L:   P2 be the linear maps given by: Tab = –a0 ab+

2 and Labc = bx – cx + a . Show directly that LT  = L  T  for bases:  = 12  01   = 111 110 100  = x2x +31 32. Prove that the linear function of Theorem 5.21 is an isomorphism. 33. Prove Theorem 5.22. 34. Let T: VW be a linear map from a vector space V of dimension n to a vector space W of dimension m. Let  and  be bases for V and W, respectively. Show that if AM mn is such that Av  = Tv  for every v  V , then AT=  .

2 Exercises 35-36. Prove that the function T :   P1 given by T ab = ax + 2b is an isomor- –1 2 –1 phism. Find the linear map T : P1   . Determine T  and T  for the given bases  –1 –1 and  , and show directly that T  = T  . 35.  ==22  01  xx + 1 36.  ==12  21  x 2 188 Chapter 5 Matrices and Linear Maps

4 aa+ b Exercises 37-38. Prove that the function T :   M22 given by T abcd = is an dc– d –1 4 –1 isomorphism. Find the linear map T : M22  R . Determine T  and T  for the –1 –1 given bases  and  , and show directly that T  = T  .

11 01 00 00 37.  = 1100 0110 0011 1000  =  11 11 11 01

01 00 00 11 38.  = 1111 1110 1100 1000  =  11 11 01 11

T 39. Let L: M32  M23 be given by LA= A (See Exercise 19, page 161). Let:

 11 11 11 11 11 10 100 110 111 111 111 111  = 1111 11 10 00 00 ,  =  . 000 000 000 100 110 111 11 10 00 00 00 00

(a) Determine L  . –1 (b) Show that L is an isomorphism, and use Theorem 5.25 to find L  .

40. Let T: VV be a linear operator. A nontrivial subspace S of V is said to be invariant under T if TS= Tv v  S  S . Assume that dimV = n and dimS = m . Show AB that there exists a basis  for V such that T  = , where 0 is the zero nm–  m 0 C matrix.

41. Let T: VW be a linear function and let  and  be bases for the finite dimensional vec- tor spaces V and W, respectively. Let AT=  . Show that:

(a)v  KerT if and only if v   nullA .

(b)w  ImT if and only if w  is in the column space of A.

42. Let V and W be vector spaces of dimensions n and m, respectively. Prove that the vector space LVW of Exercise 35, page 122, is isomorphic to Mmn . Suggestion: Let  and  be bases for V and W, respectively. Show that the function : LVW  Mmn given by T = T  is an isomorphism.

43. (PMI) Let V1V2  Vn be vector spaces and let i be a basis for Vi , 1 in . Let Ti: Vi  Vi + 1 be a linear map, 1 in– 1 . Use the Principle of Mathematical Induction to show that T T T = T T T T . n – 1 n – 2 1 n1 n – 1 nn – 1 n – 2 n – 1n – 2 2 32 1 21 5.3 Matrix Representation of Linear Maps 189

PROVE OR GIVE A COUNTEREXAMPLE

44. Let T: VW be an isomorphism, and let  = v1v2  vn be a basis for V. Then, for

every v  V , v  = Tv  , where  = Tv1 Tv2 Tvn .

45. Let T: VW be linear, and let  and  be bases for V and W, respectively. Let

rT: VW be defined by rT v = rTv . Then: rT  = rT .

46. Let  = v1v2  vn be a basis for V, and let  = 2v12v2  2vn . If T: VV is

a linear operator on V, then T  = 2T  .

47. If Z: VW is the zero transformation from the n-dimensional vector space V to the m-

dimensional vector space W, then Z  is the mn zero matrix for every pair of bases  and  for V and W, respectively.

48. Let IV : VV be the identity map on a space V of dimension n, and let  and  be

(ordered) basis for V. Then IV  = In if and only if = .

49. Let T: 2  2 be given by Tab = a + 3b 2a + 2b . There exists a basis  such that

T  is a diagonal matrix (See Exercise 18, page 161).

50. Let T: 2  2 be given by Tab = a + 3b –2a + 2b . There exists a basis  such

that T  is a diagonal matrix (See Exercise 18, page 161).

n n n n n 51. For T:    and T:    and any basis  for  : TL+  = T  + L  . 190 Chapter 5 Matrices and Linear Maps

5

§4.

In the previous section we observed that by choosing a basis  in an n-dimensional vector space V one can associate to each vector its coor- dinate vector relative to  (Definition 5.9, page 177). The following result tells us how the coordinate vector of v  V changes when switching from one basis to another:

THEOREM 5.24 For  and  bases for the finite dimensional vector space V, and for v  V we have:

v  = IV v 

where IV denotes the identity map from V to V.

PROOF: Consider the identity map IV: VV along with accompa- nying chosen bases  and  : V V v . IV .v  

Applying Theorem 5.20, page 180 we have our result:

v  ==IVv  IV IVv  =IV  v 

CHANGE OF BASE The above matrix IV  is called the MATRIX change of base matrix from  to .

EXAMPLE 5.11 Find the change-of-base matrix I for the P2  basis  = x2x 1 ,  = 1 + xx + x2 x of P2 , and then verify directly that v = I v for v = 2x2 + 3 .  P2  

SOLUTION: Definition 5.10, page 179, tells us that the first column of I is the coefficient matrix of I x2 with respect to  , the P2  P2 second column is I x , and the third column is I x ; P2  P2  bringing us to the following three vector equations:] I x2 ==x2 a1 + x ++bx+ x2 cx P2 I x ==xa1 + x ++bx+ x2 cx P2 I 1 ==1 a1 + x ++bx+ x2 cx P2 5.4 Change of Basis 191

2 Noting that we also have to find 2x + 3  we might as well throw in the fourth vector equation: 2x2 + 3 = a1 + x ++bx+ x2 cx Equating coefficients in each of the above four vector equations we come to the following four systems of equations in the three unknowns a, b, and c: xx 1 +1 +0 x

2 2 2 2 x I x ==x 1x ++0x 0 P2 = = = 2 0 I x ==x 0x ++1x 0

x P x

x 2 2 2 2 ++ ++ ++ I 1 ==10x2 ++0x 1

1 P 1

1 2 x x x 2x2 +23 = x2 ++0x 3 0 0 1 a b c 0 10 1 0 0 2 1 00 0 0 1 3 rref 1110100 0101002

1000013 001–11 1} –5

IP 2 2  2x + 3  At this point we have two of the three matrices in the equation 2x2 + 3 = I 2x2 + 3  P2   2 2 2 As for 2x + 3  , it is simply 0 , since  = x x 1 . 3 We leave it for you to verify that:

3 001 2 2 = 100 0 5 –111 – 3

CHECK YOUR UNDERSTANDING 5.18 Let V = 2,  = 12  21 , and  = 03  21 – . Determine the change-of-base matrix I and verify directly 2 

Answer: See page B-21. that 23 = I 2 23 .    

EXAMPLE 5.12 Find the coordinates of the point P = 13 with respect to the coordinate axes obtained by rotating the standard axes by 45 in a counterclockwise direction. 192 Chapter 5 Matrices and Linear Maps

SOLUTION: We are to find the coordinate vector of the point 13 with respect to vectors of length 1 (unit vectors) in the direction of the x - and y-axis depicted in Figure 5.3. y P. ==13 x y y x 2 1 45 1 1 1 1 1 1 cos45 sin 45 = ------ ------–45cos45 sin  = –------ ------.  . 2 2 2 2 1 45 x 1

Figure 5.3

We begin by finding the change-of-base matrix I 2 , for   1 1 1 1 ------ ------ –------ ------ = 10  01 and  =  : 2 2 2 2 I 10 = 10  2 I 201 = 01 

1 1 2 2 ------–10------10------2 2 rref 2 2 1 1 2 ------01 01–------2------2 2 2 2

Applying Theorem 5.20, page 180, we have:

------2------2- 2 2 1 22 13  ===I  13  2 3 –------2------2 2 2 Conclusion: In the x y - of Figure 5.3:

P = 22 2 Check: 1 1 1 1 22------ ------+212–------ ------==–21 + 23 2 2 2 2

CHECK YOUR UNDERSTANDING 5.19 Find the coordinates x y of the point P = 13 with respect to Answer: 13– 31+ the coordinate axes obtained by rotating the standard axes by 60 in a clockwise direction. 5.4 Change of Basis 193

The adjacent identity map is pointing in two directions. The left-to-right direction gives rise to the change-of- V V base matrix I , while the right-to- v I V  . V .v   left directions brings us to IV  . Are IV  and IV  related? Yes:

I In other words: V  THEOREM 5.25 Let  and  be two bases for a finite I and V  are invert- dimensional vector space V. Then: ible, with each being the –1 IV  = IV  inverse of the other. PROOF: A direct consequence of Theorem 5.23, page 184, with –1 IV: V  V playing the role of T: V  W (note that IV = IV ).

We now turn our attention to the matrix representations of a linear operator T: VV .

A generalization of this THEOREM 5.26 Let T: VV be a linear operator on a finite result appears in Exer- CHANGE OF BASIS dimensional space V. Let  and  be two cise 24. bases for V. Then: T  = IV T IV 

PROOF: Consider the following figure: V T V  

IV IV In reading the composition T of functions, you kind of V V have to read from right to   left: the right-most func- T = IV TIV tion being performed first. Top path

Dotted path: IV at left of figure, then T, then IV at right of figure Figure 5.4

Since the identity map IV does not move anything, we have:

T = IVTIV

and therefore: T  = IVTIV  Applying Theorem 5.21, page 182, to the above equation, we have:

top path in Figure 5.4 dotted path in Figure 5.4 } T  = IV  T  IV  this function first then this function and finally this function 194 Chapter 5 Matrices and Linear Maps

EXAMPLE 5.13 Verify, directly, that Theorem 5.26 holds for the linear operator T: 3  3 given by: Tabc = 2ab + c 0 and bases:  = 111 110 100  = 101 010 001

SOLUTION: We determine the four matrices:

T  T  I3  I3 

For T 

T101 = 210  T010 = 010 T001 = 010

1 0 0 20 0 100 200 rref 010111 010111 101000 001–00 2

For T 

T111 = 220  T110 = 210 T100 = 200

1 1 1 22 2 100 000 rref 1 1 0 2 10 010 210 1 0 0 00 0 001 012

For I3 

I111 = 111 I 110 = 110    I100 = 100 1 0 0 11 1 100 111 rref 010 1 10 010 110 1 0 1 10 0 001 01–1– 5.4 Change of Basis 195

For I3 

I101 = 101 I 010 = 010    I001 = 001 1 1 1 10 0 100 101 rref 110 0 10 010–11 1 – 1 0 0 10 1 001 11–0 As advertised:

111 00 0 101 200 I T I  ===110 21 0 –111 – 111 T  01–1– 01 2 11–0 –002

CHECK YOUR UNDERSTANDING 5.20 Verify, directly, that Theorem 5.26 holds for the linear operator 2 2 T: P2  P2 given by Tax++bx c = bx –2ax + c , and bases: Answer: See page B-22.  = x2x2 + x x2 ++x 1   = x2 + 1x2 –1x

In Exercise 21 you are asked DEFINITION 5.11 AB  M are similar if there exists an to show that “similar” is an nn equivalence relation on SIMILAR MATRICES invertible matrix PM nn such that –1 Mnn . (See Exercises 37-39, BP= AP. page 147 for the definition of an equivalence relation). Theorem 5.26 tells us that if T: V  V is a linear operator on a vector

space of dimension n, and if  and  are basis for V, thenT  and

T  are similar matrices. The following result kind of goes in the oppo- site direction:

THEOREM 5.27 Let  = v1v2  vn be a basis for V, and let T be a linear operator on V. If A is similar to T  , then there exists a basis  for V, such that AT=  .

PROOF: Since A is similar to T  , there exist a matrix Pp= ij such –1 that AP= T P . In Exercise 22 you are asked to verify that

 = v1v2vn , where vi = p1iv1 +++p2iv2  pnivn , is a basis for V. Applying Theorem 5.26 we have:

T  = IV T IV  . 196 Chapter 5 Matrices and Linear Maps

ith I By its very construction: I = P (see margin). Moreover: The column of V  , V  namely: –1 I v  = v  I = P (Theorem 5.23, page 184). Hence: V i  i  V  th –1 equals the i column of P, T  ===IV T IV P T P A since:   vi = p1iv1 +++p2iv2  pnivi EXAMPLE 5.14 Let T: 2  2 be the linear map given by Tab = 3a +66b a .

(a) Find T  for  = 12  21 . –812 (b) Show that is similar to T  . –1518 (c) Find a basis  for 2 such that –812 T  = –1518

T12 = 15 6 T21 = 12 12  T  1 2 15 12 10–4 1 SOLUTION: (a): rref 21612 0184

(b) We determine P = ab such that cd –1 –812 = ab –1 4ab –1518 cd 8 4 cd

ab –812 = –1 4ab cd –1518 8 4 cd The above leads us to a homogeneous system of four equations in four unknowns: –1812a – b = –4a + c  –1811a –4b –0c +0d =    8a + 15b = –4b + d  8a ++16b 0c – 4d = 0    –1812c –8d = a + 4c  –08a +0b –1816c – d =    8c +815d = b + 4d  0a –88b ++c 11d = 0

a b c d a b c d 102--9- –1811 –4–0 4 81604– 01– 1–-----11- rref 8 (*) –0168 –18– 0000 08–811 0000 5.4 Change of Basis 197

Matrix (*) has two free variables, telling us that there are infinitely –1 many ab for which –812 = ab –1 4ab . Letting cd –1518 cd 8 4 cd d = 0 and c = 1 , we arrive at a solution, namely: a ====–12 b  c 1 d 0 . And so we have: –1 –812 = –12 –1 4–12 –1518 10 8 4 10

(c) Following the procedure spelled out in the proof of Theorem 5.27,

with P = –12 , and  = 12  21 , we determine a basis 10

–812  = v1 v2 such that T  = : –1518

v1 ==– 212 + 21 03 –

v2 ==112 + 021 12 –812 Let’s verify that T  = for  = 03 –  12 : –1518 T03 – = –18 0 T12 = 15 6  T 

01 –1518 rref 10 –812 –23 06 01 –1518

CHECK YOUR UNDERSTANDING 5.21 Referring to Example 5.14, determine a basis,  distinct from

–812  = 03 –  12 , for which T  = . Answer: See page B-22. –1518 198 Chapter 5 Matrices and Linear Maps

EXERCISES

Exercises 1-7. Verify directly that v  = IV v  holds for the given vector space V, the vector v  V , and the bases  and  : 1.V = 2 , v = 25 ,  = 10  01 , and  = 12  –2 1 . 2.V = 2 , v = 31 – ,  = 10  01 , and  = 12  –2 1 . 3.V = 3 , v = 23– 1,  = 110 101 111 , and  = 100 010 001 . 4.V = 3 , v = 30– 2,  = 102 001 112 , and  = 101 011 110 , 2 2 2 2 5.VP= 2 , v = 2x ++x 1 ,  = x xx+ 1 , and  = xx+ x x ++x 1 . 2 2 2 6.VP= 2 , v = x + 1 ,  = 22 x 2x , and  = 1x + 1 x + x .

20 11 11 11 10 7.VM= 22 , v = ,  =  , and 11 11 10 00 00

00 00 01 11  =  . 01 11 11 11

8. Find the coordinates of the point Pxy=  in the xy-plane with respect to the coordinate axes obtained by rotating the standard axes  in a counterclockwise direction. (See Example 5.12.)

Exercises 9-13. Verify directly that T  = IV T IV  holds for the given vector space V, the linear operator T, and the bases  and  : 9.V = 2 , T: VV given by Tab = –b a ,  = 10  01 , and  = 12  –2 1 .

10.V = 3 , T: VV given by Tabc = –bac ,  = 010 101 111 , and  = 120 02– 1 101 .

2 2 2 11.VP= 2 , T: VV given by Tax++bx c = cx + b ,  = x x +11 , and  = 2x 1 + x2 .

ab –b c 12.VM= 22 , T: VV given by T= , cd da–

11 11 11 10 00 00 01 11  =  , and  =  . 11 10 00 00 01 11 11 11 5.4 Change of Basis and Similar Matrices 199

13. (Calculus Dependent) VP= 3 , T: VV given by Tpx = px ,  = 1xx2 x3 , and  = 1 x 2x2 3x3 .

14. Let T: 2  2 be the linear operator given by Tab = 2ab – . Find a basis  for 2

–1 10 such that T  = P T P , where  = 10  01 and P = . 11 15. Let T: 3  3 be a linear operator. Find the basis  for 3 such that –021 – –1 T  = P T P , where:  = 100 110 111 and P = 011 . 101

16. Let T: P2  P2 be a linear operator. Find the basis  for P2 such that 110 –1 2 T  = P T P , where  = 1x + 1 x + x and P = 011 . 001

17. Show that 20 and 20 are similar. –12 – 11–

18. Show that 21 and 20 are not similar. –12 – 11–

19. Find all matrices that are similar to the identity matrix In . 20. Let T: 2  2 be the linear map given by Tab = ab+  b .

(a) Find T  for  = 12  21 . 20 8 (b) Show that is similar to T  . –1738 –

2 20 8 (c) Find a basis  for  such that T  = . –1738 –

21. Show that “similar” is an equivalence relation on Mnn . (See Exercises 37-39, page 147 for the definition of an equivalence relation).

22. Show that  = v1v2vn in the proof of Theorem 5.27 is a basis for V.

n n 23. Let AB  Mnn be similar. Show that there exists a linear operator T:    and bases n  and  for  such that AT=  and BT=  . 24. (A generalization of Theorem 5.26) Let T: VW be a linear map from the finite dimen- sional vector space V to the finite dimensional vector space W. Let  and  be bases for V,

and let  and  be bases for W. Prove that: T  = IW T IV  . 200 Chapter 5 Matrices and Linear Maps

Exercises 25-29. Referring to Exercise 24, show directly that T  = IW T IV  holds for the given linear transformation T: VW , the bases  and  for V, and the bases  and  for W.

25.V ==2W 3 , Tab = –baa+ b,  = 10  01 ,  = 12  01 ,  = 110 –201 111 , and  = 111 001 110

26.V ==3W 2 , Tabc = abc+  – ,  = 110 –201 111 ,  = 111 001 110 ,  = 10  01 , and  = 12  01 .

3 2 27.V ==  W P2 , Tabc = bx + cx– a,  = 110 –201 111 ,  = 111 001 110 ,  = 21 + x 2 – x2 , and  = 1xx2 .

3 2 2 28.VP==2W , Tax++bx c = ab+ 0 –c ,  = 1xx ,  = 21 + x 2 – x2 ,  = 21 + x 2 – x2 ,  = 101 001 110 and  = 110 –201 111 .

2 2 29.VP= 2 W = P1 , Tax++bx c = bx+  a+ c ,  = x xx+ 1 ,  = 21 + x 2 – x2 ,  = xx – 1 and  = x 2x + 1 .

30. Let T: VW be linear. Let  be bases for the n-dimensional space V, and let  be bases for the m-dimensional space W. Prove that there exists an invertible matrix QM mm and an invertible matrix PM nn such that T  = QTP . Suggestion: Consider Exercise 24. 31. Let T: VW and L: WZ be linear maps. Let  be bases for V,  be bases for W,

and  be bases for Z. Show that LT  = IZ L T IV  .

PROVE OR GIVE A COUNTEREXAMPLE

32. Let T: VV be a linear operator, and let  = v1v2  vn and  be a bases for V. If

T  = 2T  , then  = 2v12v2  2vn . 33. If A and B are similar matrices, then A2 and B2 are also similar.

34. If A and B are similar invertible matrices, then A–1 and B–1 are also similar. 35. If A and B are similar matrices, then at least one of them must be invertible. 36. If A and B are similar matrices, then so are their transpose. (See Exercise 19, page 161.) 37. If A and B are similar matrices, and if A is symmetric, then so is B. (See Exercises 20, page 161.) 38. If A and B are similar matrices, and if A is idempotent, then so is B. (See Exercises 22, page 162.) 39. If A and B are similar matrices, then TraceA = TraceB . (See Exercises 24, page 162.) Chapter Summary 201

CHAPTER SUMMARY

MULTIPLYING MATRICES You can perform the product Cmn = Amr Brn of two matrices, if the number of columns of A equals the number of rows of B, and you get cij of the product matrix C by running across the ith row of A and down the jth column of B:  cij = ai1b1j ++++ai2b2j ai3b3j airbrj Properties Assuming that the matrix dimensions are such that the given operations are defined, we have: (i) AB+ C = AB+ AC (ii) AB+ CACBC= + (iii) ABC= AB C (iv) rAB==rA BArB A connection between matrix For AM mn and any positive integer z, the map multiplication and linear T :M  M given by T X = AX is linear. transformations. A nz mz A In particular:

n m For AM mn the map TA:    given by

TAX = AX is linear, where X is a vertical n-tuple and TAX is a vertical m-tuple.

NULL SPACE OF A MATRIX, n For AM mn , the null space of A is the subspace of R con- AND NULLITY sisting of the solutions of the homogeneous of equations AX = 0 . It is denoted by nullA .

INVERTIBLE MATRIX A square matrix A is said to be invertible if there exists a matrix B (necessarily of the same dimension as A), such that: AB== BA I The matrix B is then said to be the inverse of A, and we write BA= –1 . If no such B exists, then A is said to be non-invert- ible, or singular. Need only “work” on one Let A be a square matrix. If B is a square matrix such that side either AB= I or BA= I, then A is invertible and A–1 = B . Uniqueness An invertible matrix has a unique inverse. 202 Chapter 5 Matrices and Linear Maps

Properties If A is invertible and r  0 , then rA is also invertible, and: A–1 –1 = A 1 rA –1 = ---A–1 r

 –1 –1 –1 –1 A1A2 An = An An – 1 A1

ELEMENTARY MATRIX A matrix that is obtained by performing an elementary row operation of an identity matrix

Invertibility Every elementary matrix is invertible. Inverses by means of The matrix obtained by performing an elementary row opera- multiplication tion on a matrix A  Mmn equals that matrix obtained by multiplying A on the left by the elementary matrix obtained by performing the same elementary row operation on the identity

matrix Im . Inverses by row-reduction If a sequence of elementary row operations reduces the invert- ible matrix A to I, then applying the same sequence of elemen- tary row operations on I will yield A–1 .

EQUIVALENCES OF Let AM nn . The following are equivalent: INVERTIBILITY (i) A is invertible.

(ii)AXB= has a unique solution for every B  Mn  1 . (iii)AX = 0 has only the trivial solution. (iv)rrefA = I . (v) A is a product of elementary matrices.

COORDINATE c where v = c v +++c v  c v VECTOR 1 1 1 2 2 n n v = and  

cn  = v1v2  vn is a basis for V

MATRIX Let T: VW be linear,  = v1v2  vn be a basis for V, and REPRESENTATION OF and  = w w  w a basis for W. A LINEAR MAP 1 2 m The matrix representation of T with respect to  and  is that th matrix T   Mmn whose i column is Tvi  Chapter Summary 203

The matrix representa- Let T: VW be linear. If  = v1v2  vn is a basis for V, and tion of a linear map T  = w w  w is a basis for W, then: describes the “action” 1 2 m of T. Tv  = T v 

The matrix of a com- Let T: VW and L: WZ be linear maps, and let  and  be position function is the bases for the finite dimensional spaces V, W, and Z, respectively. product of matrices of Then: those functions. LT  = L T 

Relating coordinate Let  and  be bases for V. Then: vectors with respect to different bases. v  = I v 

The matrix of the Let T: VW be an invertible transformation and let  and  be inverse of a transforma- bases for V and W, respectively. Then: tion is the inverse of the T –1 = T –1 matrix of that transfor-   mation. Relating matrix repre- Let T: VV be a linear operator, and let  and  be two bases for sentations of a linear V. Then: operator with respect to different bases. T  = IV T I 

SIMILAR The matrices AB  Mnn are similar if there exists an invertible MATRICES –1 matrix PM nn such that BP= AP . Similar matrices repre- Let  = v1v2  vn be a basis for V, and let T be a linear opera- sent linear maps with tor on V. If A is similar to T , then there exists a basis  for V, respect to different  basis. such that AT=  . 204 Chapter 5 Matrices and Linear Maps 6.1 Determinants 205

6 CHAPTER 6 DETERMINANTS AND EIGENVECTORS As you know, a linear operator T: VV has a matrix representation

T  , which depends on the chosen basis  for V. A main goal of this

chapter is to determine if there exists a basis  for which T  turns out to be a diagonal matrix. Determinants, as you will see, play an essential role in that endeavor. §1. DETERMINANTS

We define a function that assigns to each square matrix a (real) num- jthcolumn ber: 1

st out row DEFINITION 6.1 For a 22 matrix:

out det ab = ad– bc cd For AM , with n  2 , let A denote the nn nn 1j n – 1  n – 1 matrix obtained by deleting the first row and jth column of the matrix A resulting in a (see margin). Then: asquare matrix n of dimension 1 + j n – 1 detA =  –1 a1j detA1j j = 1

IN WORDS: Multiply each entry a1j in the first row of A by the determinant of the (smaller) matrix obtained by discarding the first row and the jth column of A, and then sum those n products with alternating signs (starting with a + sign). It’s not as bad as you may think:L

EXAMPLE 6.1 29– 3 Evaluate: det 32–4 57– 6

SOLUTION:

2 93– 2 9 –3 29–3 3 –2 4 3 –2 4 32–4 5 76– 5 7 –6 57–6

29– 3 –2 4 34 32– det 32–4= 2det – 9det – 3det 76– 56– 57 57– 6 ==22– –6 – 47 –3375293– 6 – 45 –217 – – 206 Chapter 6 Determinants and Eigenvectors

GRAPHING CALCULATOR GLIMPSE 6.1

Definition 6.1 defines the determinant of a matrix by an expansion process involving the first row of the given matrix. The next theorem, known as the Laplace Expansion Theorem, enables one to expand along any row or column of the matrix. A proof of this important result is offered at the end of the section.

THEOREM 6.1 For given AM nn , Aij will denote the n – 1  n – 1 submatrix of A obtained th th Note that the sign of the by deleting the i row and j column of A. ij+ We then have: –1 has an alternat- n ing checkerboard pattern EXPANDING ALONG i + j _ _ _ _ detA = –1 a detA + + + + THE th ROW ij ij _ _ _ _ i  + + + + + _ + _ + _ + _ j = 1 _ _ _ _ + + + + + _ + _ + _ + _ and: _ _ _ _ + + + + EXPANDING ALONG n _ _ _ _ + + + + th _ _ _ _ THE j COLUMN i + j + + + + detA =  –1 aij detAij i = 1

Note: detAij is called the of aij , and ij+ th Cij = –1 aij detAij is called the ij cofactor of A

EXAMPLE 6.2 Evaluate: 29– 3 det 32–4 57– 6 by expanding about its second row.

SOLUTION:

_ 2 93– 2 9 –3 29–3 as in Example 6.1 same +_ _+ + 3 –42 3 –2 4 32– 4 + _ + 5 76– 5 7 –6 57–6

29– 3 93– 23– 2 9 det 32– 4 = –det3 –2det –4det 76– 56– 5 7 57– 6 ==–22639– 6 – 73– –– – 53– –217 427 –  59 6.1 Determinants 207

CHECK YOUR UNDERSTANDING 6.1

29– 3 Evaluate det 32–4 expanding along: 57– 6 Answer: See page B-23. (a) The third row. (b) The second column.

While it was not so bad to calculate the 33 determinant of Example 6.1, the task gets increasingly more tedious as the dimension of the matrix increases. If we were only interested in calculating determinants An upper triangular matrix of matrices with numerical entries, then we could avoid the whole mess is a square matrix with zero entries below its main diag- entirely and simply use a calculator. But this will not always be the case. onal. For example: In any event, one can easily calculate the determinant of any upper triangular matrix (see margin): 2501– 2 017–0 2 The determinant of an upper diagonal matrix 00034 THEOREM 6.2 00091 equals the product of the entries along its diagonal. 00005 PROOF: By induction on the dimension, n, of Mnn . A lower triangular matrix is a square matrix with zero ab entries above its main diago- I. Claim holds for n = 2 : det ==ad– b  0 ad . nal. For example: 0 d 2000 II. Assume claim holds for nk= . 5100 –3402 III. We establish validity at nk= + 1 : 4025 Let Aa= ij  Mk + 1  k + 1 . Since all entries in the first

column below its first entry a11 is zero, expanding about the

first column of A we have detA = a11detA11 (where A11 is the k by k upper triangular matrix obtained by removing the first row and first column from the matrix A. As such, by II:  detA11 = a22a33 ak + 1 k + 1 . Consequently:  detA ==a11detA11 a11a22a33 ak + 1 k + 1

COROLLARY For any n: detIn = 1 .

CHECK YOUR UNDERSTANDING 6.2

Prove that the determinant of a lower diagonal matrix equals the prod- Answer: See page B-23. uct of the entries along its diagonal. 208 Chapter 6 Determinants and Eigenvectors

ROW OPERATIONS AND DETERMINANTS Since it is easy to find the determinant of an upper triangular matrix, ad since any square matrix can be transformed into an upper triangular matrix by means of elementary row operations, it would be nice to have relations between the determinant of a matrix and that obtained by per- forming an elementary row operations on that matrix. Niceness is at hand:

THEOREM 6.3 (a) If two rows of AM nn are inter- changed, then the determinant of the resulting matrix is –detA . (b) If one row of A is multiplied by a con- stant c, then the determinant of the result- ing matrix is cdetA . (c) If a multiple of one row of A is added to another row of A, then the determinant of the resulting matrix is detA .

PROOF: (a) By on the dimension of the matrix A. For n = 2 :

det ab ==ad– bc and det cd cb– da cd ab negative of each other Assume the claim holds for matrices of dimension k  2 (the hypothesis).

Let Aa= ij be a matrix of dimension k + 1 , and let Bb= ij denote the matrix obtained by interchanging rows p and q of A. Let i be the index of a row other than p and q. Expanding about row i we have: k + 1 k + 1 ij+ ij+ detA =det –1 aij Aij and det B =det –1 bij Bij j = 1 j = 1 Since rows p and q were switched to go from A to B, row i of B still equals that of A, and therefore: bij = aij . Since Bij is the matrix Aij with two of its rows interchanged, and since those matrices are of dimension k, we have: detBij = –detAij (the hypothesis). Consequently: k + 1 k + 1 detB = –1 ij+ b detB = –1 ij+ a –detA  ij ij  ij ij j = 1 j = 1

k + 1 ij+ ==–  –1 aijdetAij –detA j = 1 6.1 Determinants 209

(b) Let B denote the matrix obtained by multiplying row i of matrix A by c. Expanding both matrices about the ith row, we have: n n ij+ ij+ detA =det –1 aij Aij and detB =det –1 bij Bij j = 1 j = 1

Matrix A and B differ only Since bij = caij , and since Bij = Aij (margin), we have: in the ith row, and that row n n has been removed from ij+ ij+ both A and B to arrive at detB =det –1 bij Bij =  –1 caij detAij the matrices A and B . ij ij j = 1 j = 1

n pull out that common factor, c: ij+ ==c  –1 aijdetAij cdetA j = 1

(c) Let B be the matrix obtained by multiplying row r of A by c and adding it to row i. If ir= , then the result follows from (b). Assume ir . Expanding B about its ith row, we have: n ij+ detB =det –1 aij + carj Aij j = 1 n n ij+ ij+ =  –1 aijdetAij + c  –1 arjdetAij j = 1 j = 1 n ===detA + c –1 ij+ a detA detA + c detA  rj ij 0 j = 1

n ij+  –1 arjdetAij is the determinant of a matrix with th th j = 1 two equal rows: the i and r row The result now follows from CYU 6.3 below

CHECK YOUR UNDERSTANDING 6.3

Answer: See page B-23. Show that if two rows of a matrix A are identical, then detA = 0 . 210 Chapter 6 Determinants and Eigenvectors

The following example illustrates how Theorem 6.3 can effectively be used to calculate determinants.

EXAMPLE 6.3 1268 Evaluate: det 1002 2135 –1111

SOLUTION: Theorem 6.3(a) Theorem 6.3(c) (applied three times)

1268 1002 100 2 det 1002 ==–det 1268 –det 026 6 2135 2135 013 1 –1111 –1111 011 3

Theorem 6.3(b) (see margin) a11  a1n  100 2 100 2 100 2 det ca  ca i1 in 013 3 013 3 013 3  ==–det2 –det2 =2det =8 013 1 000 –2 00– 2 0 an1  ann 011 3 00– 2 0 000– 2 a  a 11 1n Theorem 6.3(c)  Theorem 6.3(a) Theorem 6.2 (applied two times) = c  det ai1  ain 

an1  ann CHECK YOUR UNDERSTANDING 6.4

Evaluate: 2 101 det 012 2 101 4 Answer: 15 411 3

To establish any of the following three elementary matrix results, you

need but substitute the identity matrix In for A in Theorem 6.3:

Note that det In = 1 THEOREM 6.4 (a) If EM nn is obtained by interchang- ing two rows of In , then detE = –1 .

(b) If EM is obtained by multiplying The restriction c  0 is nn imposed in (b) since we a row of In by c  0 , then detE = c . are concerned with ele- mentary row operations (see page 3). (c) If EM nn is obtained by adding a multiple of one row of In to another row, then detE = 1 . 6.1 Determinants 211

Soon, we will be in a position to show that the determinant of a prod- uct of any two nn matrices is equal to the product of their determi- nants. For now:

THEOREM 6.5 For any BM nn and any nn elementary matrix E: detEB = detE detB

PROOF: Let EM nn be an elementary matrix obtained by inter- changing two rows of In . By Theorem 5.13, page 169: (*) EB is the matrix obtained by interchanging the same two rows in the matrix B. Consequently: detEB ==–detdetB E detB (*) and Theorem 6.4(a) detE = –1 [Theorem 6.4(a)] As for the rest of the proof: CHECK YOUR UNDERSTANDING 6.5

(a) Establish Theorem 6.5, for the elementary matrix EM nn : (i) Obtained by multiplying a row of In by c  0 . (ii) Obtained by adding a multiple of one row of In to another row of In .

(b) Let BM nn and E1E2  Es  Mnn be elementary matri- ces. Use the Principle of Mathematical Induction to show that:   detEs E2E1B = detEs E2E1 detB  Answer: See page B-24. = detEs detE2 detE1 detB

We now come to one of the most important results of this section: You can add this result to the list of equivalences for THEOREM 6.6 A matrix AM is invertible if and only invertibility appearing in nn Theorem 5.17, page 172: if detA  0 . (vi) detA  0 PROOF: Let E1E2  Es be a sequence of elementary matrices  such that Es E2E1A = rrefA (Theorem 5.13, page 169). Appeal- ing to CYU 6.5(b), we have:  detEs detE2 detE1 detA = det rrefA

By Theorem 6.4, detEs detE2 detE1  0 . Consequently: Ifrref AI= , then: det rrefA  0 if and only if detA  0 det rrefA = 10 If rrefA  I , then its last margin: if and only if rrefA = I row consists entirely of zeros, and Theorem 5.17(iv), page 172: if and only if A is invertible det rrefA = 0 212 Chapter 6 Determinants and Eigenvectors

We are now in a position to establish another powerful result, Austin Cauchy, a prolific attributed to Cauchy: French mathematician (1789-1857). THEOREM 6.7 For AB  Mnn : detAB = detA detB

PROOF: Case 1: A is invertible. By Theorem 5.17, page 172, A can be expressed as a product of elementary matrices:

AE= sE2E1

Then: detAB = detEsE2E1B CYU 6.5(b): ==detEsE2E1 detB detA detB

Case 2: A is not invertible. AB is not invertible; for: AB invertible  CAB C = I  ABC= IA invertible -- a contradiction. It follows, from Theorem 6.6, that both detAB and detA are zero, and we again have equality: detAB ==0 0 detB =detA detB

CHECK YOUR UNDERSTANDING 6.6 Prove that if A is invertible, then: 1 detA–1 = ------Answer: See page B-24. detA

For the brave at heart: PROOF OF THE LAPLACE EXPANSION THEOREM

The column-expansion part We use on n to show that the determinant of AM can be eval- of the theorem is relegated nn to the exercises. uated by expanding about any row of A: The claim is easily seen to hold when n = 2 :

a11 a12 det ==a11a22 – a12a21 –a21a12 + a11a22 a21 a22 Assume that the claim holds for nk= (the hypothesis).

Let AM k + 1 k + 1 . We show that for any 1  t  k + 1 : k + 1 k + 1 This will show that the 1 + j ts+ expansion about any row –1 a1jdetA1j = –1 atsdetAts equals that of expanding   j = 1 s = 1 } about the first row. } (*) (**) expanding about first row expanding about row t 6.1 Determinants 213

Working with (*), we employ the hypothesis and evaluate the deter- minant of each kk matrix A1j along its t – 1 row, which is row t of A (see Figure 6.1). In so doing, we need to keep in mind that just as st th each A1j has a row and a column removed from A (1 row and j column), so then will each submatrix in the expansion of detAij have two rows and two columns of A removed. column numbers for A } 1 2 j – 1 j j + 1 n

row numbers for a11 a12  a1j – 1 a1j a1j + 1  a1n

1 row numbers for 2 a21 a22  a2j – 1 a2j a2j + 1  a2n 1      

t as1 as2  asj– 1 asj asj+ 1  asn t – 1       A A { a a  a a a  a } n1 n2 nj– 1 nj nj+ 1 nn n n – 1 ij 1 2 j – 1 j n – 1 }

column numbers for Aij Figure 6.1

Let A1tjs denote the submatrix of A with rows 1 and t, and columns j and s of A removed. Using the hypothesis we can obtain the determi- th nant of A1j by expanding about its t – 1 row (which is the t row of A), breaking that sum into two pieces the “before-j” piece, and the “after-j” piece we have: j – 1 k + 1 t – 1 + s t – 1 + s – 1 detA1j =1 –1 ats detA1tjs +  – ats detA1tjs s = 1 sj= + 1 Bringing us to: k + 1 1 + j  –1 a1jdetA1j j = 1

k + 1  j – 1 k + 1  1 + j  t – 1 + s t – 1 + s – 1  = –1 a1j –1 ats detA1tjs + –1 ats detA1tjs       j = 1 s = 1 sj= + 1 

jts++ jts++– 1 =1 –1 a1jatsdetA1tjs +  – a1jatsdetA1tjs (A) sj sj 1 + J + t – 1 + s 1 ++jt– 1 +s – 1 214 Chapter 6 Determinants and Eigenvectors

Turning to (**), we again appeal to the hypothesis, and expand about the first row to calculate each detAts : k + 1 ts+  –1 atsdetAts s = 1 k + 1 s – 1 k + 1 ts+ 1 + j 1 + j – 1 = –1 ats–1 a detA1tjs + –1 a detA1tjs   1j  1j s = 1 j = 1 js= + 1

= –1 tsj+++1 a a detA + –1 tsj++ a a detA  ts 1j 1tjs  ts 1j 1tjs (B) js js To complete the proof, we observe that the left summation in (A) is equal to the right summation in (B), and that the right summation in (A) equals the left in (B):

–1 jts++a a detA = –1 tsj++ a a detA  1j ts 1tjs  ts 1j 1tjs sj js –1 jts++– 1a a detA = –1 tsj+++1 a a detA  1j ts 1tjs  ts 1j 1tjs sj js

note that – 1 jts++– 1 = –1 tsj+++1 6.1 Determinants 215

EXERCISES

Exercises 1-8. Compute the determinant of the given matrix.

210 153 224 124 1.–211 2.–011 3.5711 4. –199 101 429 369 461

2101 0301 1304 6339 5.–2121 6.–2421 7.–2461 8. 33–36 1010 4040 0042 30312 0202 1235 1035 9036

ab c Exercises 9-14. Given that det def = 9 , find: ghi

gh i a b c a +2d b +2e c + 2f 9.det ab c 10.det 2d 2e 2f 11. det de f def –3g – 3h – 3i gh i

2 ab c ab c adg+– beh+– cfi+ – 12. det def 13. det ghi 14. det de f ghi def gh i

Exercises 15-18. Find all values of k for which the given matrix is invertible.

15. k 1 k –01 kk2 0 k 00 1 16. 2 k k 0 k –1 17. 0 k k 18. 0 k 1 0 241 k2 k 0 10k 0 011 k Exercises 19-22. Verify:

x –01 111 2 19. det 0 x –1 = ax ++bx c 20. det abc = ab– bc– ca– cba a2 b2 c2

111 1 +11a 21. det abc = ab– bc– ca– abc++ 22. det 11+1b = abc+++ bc ac bc a3 b3 c3 111+ c

23. While one can certainly find matrices AB  Mnn such that AB BA , prove that one can not find matrices AB  Mnn such that detAB  detBA . 216 Chapter 6 Determinants and Eigenvectors

111 24. Show that the matrix abc is invertible if and only if the numbers a, b, and c, are all distinct. a2 b2 c2 25. Prove that if a matrix A contains a row (or column) consisting entirely of zeros, then detA = 0 .

26. If Dd= ij  Mnn is a diagonal matrix and if Xi  Mn  1 is the column matrix whose th i entry is 1 and all other entries are 0, then DXi = diiXi . T T 27. Let AM nn . Prove that detA = detA , where A denotes the transpose of A (see Exercise 19, page 161). n 28. Prove that if AM nn is skew-symmetric, then detA = –1 detA (see Exercise 21, page 162). What conclusion can you draw from this result?

29. For AM nn , let B be obtained from A by interchanging pairs of rows of A m times. Express detB as a function of m and detA . 30. Let A be similar to B (see Definition 5.11, page 195). Prove that: (a) detA = detB (b) detAcI– = detBcI– for every c   . Suggestion: Consider Theorem 5.1, page 153.

x y 1 x y 1 31. Show that det 1 1 = 0 is an equation of the line passing through the points x1 y1 x2 y2 1 2 and x2 y2 in  .

x y z 1 x1 y1 z1 1 32. Show that det = 0 is an equation of the plane passing through the points x2 y2 z2 1 x3 y3 z3 1 3 x1y1 z1 , x2y2 z2 , and x3y3 z3 in  .

33. Show that the of the with vertices at x1 y1 , x2 y2 , and x3 y3 is given by

x1 y1 1 1 ---det x y 1 , where the sign ( ) is chosen to yield a positive number. 2 2 2 x3 y3 1 34. (Cramer’s Rule) If AX= B is a system of n equations in n unknowns, with A invertible, then the system has a unique solution x1x2  xn [Theorem 5.17(ii), page 172]. Cramer’s rule asserts that: detA detA detA x = ------1  x = ------2  x = ------n 1 detA 2 detA n detA th where Ai is the matrix obtained by replacing the i column of A with B. Use Cramer’s rule to solve the system of: (a) Example 1.3, page 9. (b) CYU 1.3, page 10. 6.1 Determinants 217

35. Prove the “column-expansion-part” of Theorem 6.3 (Laplace Expansion Theorem).

Exercises 36-39. Use the Principle of Mathematical Induction to show that:   36. For any m and Ai  Mnn , detA1  A2 Am = detA1 detA2 detAm .

n 37. For any AM nn and cR : detcA = c detA .

m m 38. Prove that for AM nn and any positive integer m: detA = detA .

cI X 39. If AM nn is of the form A = , where I is the rr identity matrix, 0 is 0 Y the nr–  r zero matrix, and X and Y are rnr – and nr–  nr– matrices, respectively, then: detA = crdetY .

X Y 40. If AM nn is of the form A = , where X and Z are square matrices and 0 Z

0 is a zero matrix, then: detA = det X det Z .

PROVE OR GIVE A COUNTEREXAMPLE

41. For AM nn , if detA = 1 , then AI= .

42. For AM nn , if detA = 0 , then A is the zero matrix.

–1 43. For AB  Mnn , if detAB = 1 , then both A and B are invertible and AB= .

44. For any AB  Mnn , detAB+ = detA + detB .

45. For any AM nn , det–A = –detA .

46. If AM nn is nilpotent, then detA = 0 (see Exercise 23, page 162).

X Y 47. If AM 44 and XYZW  M22 , and if A = , then: Z W detA = detX detW – detY detZ . 218 Chapter 6 Determinants and Eigenvectors

6

The German word eigen §2. EIGENSPACES translates to: characteristic. At one time, eigenvalues were called latent values, We begin by defining eigenvalues and eigenvectors for matrices: and it is for this reason that  (lamba), the Greek letter DEFINITION 6.2 An eigenvalue of a matrix AM nn is a for “l” is used. EIGENVALUES AND scalar  (which may be zero) for which , EIGENVECTORS there exists a nonzero vector X  n such We remind you that we use (FOR MATRICES) that: n to denote M , and n  1 AX = X that v  n is the vector Any such vector X is said to be an eigenvec- v  n in “column form.” tor corresponding to the eigenvalue  .

EXAMPLE 6.4 Show that 21 and –1 3 are eigenvectors of

the matrix A = 32 . 32–

32 2 8 2 SOLUTION: Since ==4 , 21 is an eigenvec- 32– 1 4 1 tor of A corresponding to the eigenvalue 4. By the same token, –13 is an eigenvector corresponding to the eigenvalue –3 :

32 –1 ==3 –3 –1 32– 3 –9 3

Since the set of eigenvectors corresponding to an eigenvalue  of a matrix AM nn does not contain the zero vector, it cannot be a sub- space of n . If you throw in the zero vector, however, you do end up with a subspace: eigenvectors, along with the zero vector n-by-n identity matrix Recall that null(A) denotes the solution set of the XAX= X ==XAX– X = 0 XAX– InX = 0 homogeneous system of equations AX = 0. ==A – In X = 0 nullA – In a subspace of n (Theorem 5.4, page 159)

Bringing us to: DEFINITION 6.3 The eigenspace associated with an eigen- value  of a matrix AM nn , denoted by E , is given by:

E = nullA – In 6.2 Eigenspaces 219

EXAMPLE 6.5 Find a basis for the eigenspace E4 of the

matrix A = 32 of Example 6.4. 32–

32 10 –21 SOLUTION: E4 ==null– 4 null 32– 01 36– –21 null is the solu- 36– –21 rref 1 –2 tion set of the homoge- From we see that E4 = 2rr rR neous system: 36– 00 –2x +0y =   with basis 21 . 3x –06y =  CHECK YOUR UNDERSTANDING 6.7

Find a basis for the eigenspace E–3 of the matrix A = 32 of Answer: 13 – 32– Example 6.4.

CHARACTERISTIC POLYNOMIALS At this point, there is a gap in our eigenvector development; namely: How does one go about finding the eigenvalues of a matrix? The answer hinges on the following objects:

DEFINITION 6.4 For AM nn , the n-degree polynomial CHARACTERISTIC detA – I is said to be the characteristic POLYNOMIAL n (FOR MATRICES) polynomial of A, and detA – In = 0 is said to be the characteristic equation of A. We are now in the position to state the main theorem of this section:

THEOREM 6.8 The eigenvalues of AM nn are the solutions of the characteristic equation detA – In = 0 .

PROOF: To say that  is an eigenvalue of A is to say that there exists a nonzero vector X  n such that: AX = X AX – X = 0

AX – In X = 0

A – In X = 0 220 Chapter 6 Determinants and Eigenvectors

But, to say that A – In X = 0 has a nontrivial solution is to say that detA – In = 0 (Theorem 6.6, page 211, and Theorem 5.17(iii), page 172).

EXAMPLE 6.6 Find the eigenvalues and corresponding eigens- paces of the matrix: 1 0 1 A = 2 2 1 1 0 1

SOLUTION: The eigenvalues are the solutions of the equation: 1 – 0 1 detA – I3 ==det 2 2– 1 0 1 0 1–  Expanding about the third row, we have: A better choice is to expand 01 1 –1 1 –  0 about the second column. If 1  det – 0det + 1 –  det = 0 you do, pay particular atten- 2 –  1 21 22–  tion to the checkerboard sign pattern of page 206. –12 –  + –  1 –  2 –  = 0 2 –  –11 + –  2 = 0 2 –  –121 ++–  2 = 0 2 –  – 2 –  = 0 –02 –  2 =  ==0  2 We now determine the eigenspaces for the two eigenvalues, 0 and 2. Finding E0 :  1 0 1 1 00 1 0 1 E0 ==null2 2 1– 0 0 10 null2 2 1  1 0 1 0 01 1 0 1 x y z x y z 1 0 1 1 0 1 rref 1 2 2 1 0 1 –--- 2 1 0 1 0 0 0 Setting the free variable z equal to r, we have: r E0 ==–r --- r r   –2rr2r rR  2 with basis: –212 . 6.2 Eigenspaces 221

Finding E2 :  1 0 1 1 00 –1 0 1 E2 ==null2 2 1– 2 0 10 null2 0 1  1 0 1 0 01 1 0 – 1

x y z x y z –1 0 1 1 00 rref 2 0 1 001 1 0 – 1 000 Setting the free variable y equal to r, we have: E2 = 0r 0 r   with basis 010 .

EXAMPLE 6.7 Find the eigenvalues and corresponding eigenspaces of the matrix: 02–0 2 A = 110– 1 –121 –1 –121 –1

SOLUTION: To find the eigenvalues of A we need to solve the charac- teristic equation: A TI-92 teaser: –22 –0 11–0 – 1 detA – I4 ==det 0 –121 – –1 –121 –1– 

In this endeavor, we first use Theorem 6.3, page 208, to express

detA – I4 as the determinant of a matrix in upper triangular form, and then take advantage of Theorem 6.2, page 207, to finish the job: 222 Chapter 6 Determinants and Eigenvectors

switch the first two rows R1 + R2  R2 ;R1 + R3  R3 ;R1 + R4  R4

–22 –0 11–0 – 1 11–01 – 2 det 11–0 – 1==–detdet –22 –0– 0 2+2– – – –121 – –1 –121 – –1 02– – 2– 0 –121 –1–  –121 –1–  02–2 – – switch rows 2 and 4 (introduces another minus sign) – 1R + R  R 2 3 3 – 1 –  R2 + R4  R4

11–01 – 11–01 – 11–0 – 1 02–2 – – 02–2 – – 02–2 – – ===det det det 02– – 2– 0 0 0 –  00–  0 2+2– 2 – – 0 2+2– 2 – – 0022

2R3 + R4  R4

11–0 – 1 02–2 – – ===det 1 2 –  – 2 + 2 2 – 2 + 2 00–  2 000 + 2 Theorem 6.2, page 207

2 Setting detA – I4 =  – 2 + 2 to 0, we see that there are three distinct eigenvalues:  = 0  = 2 , and  = –2 . As for their corresponding eigenspaces:

02–0 2 1000 02–0 2  E0 ==null110– 1– 0 0100 null110– 1  –121 –1 0010 –121 –1 –121 –1 0001 –121 –1 x y z w x y z w 02–0 2 1 01– 1 rref 110– 1 0 1 –01 –121 –1 0000 –121 –1 0000 Setting the free variables z and w equal to r and s respectively, we find that E0 = – r + srrs rs . By letting r ==1 s 0 , and then r ==0 s 1 we arrive at the basis –1110  1001 . 6.2 Eigenspaces 223

02–0 2 1000 –222 –0  E 2 ==null110– 1– 2 0100 null11–01–   –121 –1 0010 –141 –1 –121 –1 0001 –121 –1–

x y z w x y z w –222 –0 1 –001 rref 11–01– 001 0 –141 –1 0001 –121 –1– 0000 Setting the free variable y equal to r, we have E2 = rr00 rR with basis 1100 .

02–0 2 1000 22–0 2  E –2 ==null110– 1+ 2 0100 null130– 1   –121 –1 0010 –1011 –121 –1 0001 –121 –3

x y z w x y z w 22–0 2 1 00– 1 130– 1 rref 0 1 00 –1011 001 –1 –121 –3 0000

Setting the free variable w equal to r, we find that E–2 = r0 rr r   with basis 1011 .

CHECK YOUR UNDERSTANDING 6.8 Find the eigenvalues and corresponding eigenspaces of the matrix: 16 3 2 A = –384 – Answer: See page B-24. –62 –11 224 Chapter 6 Determinants and Eigenvectors

TURNING TO LINEAR OPERATORS Shifting our attention from matrices to linear operators we have: DEFINITION 6.5 Let T: VV be a linear operator. An eigen- Compare with Definition 6.3, EIGENVALUES AND value of T is a scalar  (which may be page 218. EIGENVECTORS zero) for which there exists a nonzero vector (FOR LINEAR OPERATORS) v  V such that: Tv = v Any such v is then said to be an eigenvector corresponding to the eigenvalue  .

EXAMPLE 6.8 Show that 21 and –1 3 are eigenvectors of the linear operator T: 2  2 given by: Note that the linear map T stretches the eigenvector Tab = 3a +32b a – 2b 21 by its eigenvalue 4, SOLUTION: and –31 by –3 : (a) T21 ===6262+  – 84 421 by d” che ret f 4 “st r o 421 and T–13 ===– 3 + 63 – – 6 39 – –3–13 . acto –13 a f We see that 21 is an eigenvector corresponding to the eigenvalue 4, 21 and that –13 is an eigenvector corresponding to the eigenvalue –3 . Since the set of eigenvectors corresponding to an eigenvalue  of a

“s linear operator T: VV does not contain the zero vector, it cannot be tre f tc ac h to ed r ” a subspace of V. As it was with matrices, however, if you throw in the of by -3 a zero vector, then you do end up with a subspace, for: –3–13 eigenvectors, along with the zero vector vTv= v = v Tv – v = 0

==v T – IV v = 0 KerT – IV

a subspace of V (Theorem 4.8, page 126) Bringing us to: Compare with Definition 6.4, DEFINITION 6.6 The eigenspace associated with an eigen- page 219. value  of a linear operator T: VV , denoted by E , is given by:

E = KerT – IV

EXAMPLE 6.9 Find a basis for the eigenspace E4 of the lin- Compare with Example 6.5. ear operator: Tab = 3a +32b a – 2b of Example 6.8. 6.2 Eigenspaces 225

SOLUTION: The kernel of the linear operator:

T – 4I 2 ab ==Tab – 4ab 3a + 2b – 4a 3a – 2b – 4b  = – a +32b a – 6b is, by definition, the set: ab – a +32b a – 6b = 00 Equating coefficients, we have: a b a b –2a +0b =  –21 rref 1 –2  3a –06b =  36– 00 Setting the free variable b equal to r, we have E4 = 2rr r   with basis 21 .

CHECK YOUR UNDERSTANDING 6.9 Find a basis for the eigenspace E–3 of the linear operator Answer: –31 Tab = 3a +32b a – 2b of Example 6.8. How does one go about finding the eigenvalues of a linear operator? Like this: THEOREM 6.9 The eigenvalues of a linear operator T: VV on a vector space of dimension n are the eigen- values of the matrix T   Mnn , where  is any basis for V.

PROOF: We show that v is an eigenvector for the linear operator T corresponding to the eigenvalue  , if and only if v  is an eigenvec-

tor for the matrix T  corresponding to the eigenvalue  : Tv = v with v  0

Note that Tv  = v 

v  ==0  v 0 Theorem 5.22, page 180: T v  = v  with v   0 (Why?)

The above theorem leads us to the following definition: DEFINITION 6.7 Let T: VV be a linear operator on a vector Theorem 5.26, page 193, and Exercise 30(b), page CHARACTERISTIC space V of dimension n. The characteristic 216, tell us that POLYNOMIAL polynomial of T is the n-degree polynomial detT – I (FOR LINEAR OPERATORS)  n detT  – In where  is any basis for V, = detT  – In and detT  – In = 0 is said to be the for any bases  and  characteristic equation of T. 226 Chapter 6 Determinants and Eigenvectors

Embedding the above terminology in the statement of Theorem 6.9, we come to: THEOREM 6.10 Let V be a vector space of dimension n. The Compare with Theorem 6.8 eigenvalues of the linear operator T: VV are the solutions of the characteristic equation detT  – In = 0 , where  is any basis for V.

EXAMPLE 6.10 Find the eigenvalues and corresponding eigenspaces of the linear operator T: P2  P2 given by: Tax2 ++bx c = ac+ x2 ++2a ++2bcxac+ 2 PROOF: With respect to the basis  = x x 1 in P2 , we have:

Tx2 = 1x2 ++2x 1 1 01 Tx= 0x2 ++2x 0 T  = 2 21 2 T1 = x ++x 1 1 01

Theorem 6.10 tells us that the eigenvalues are the solutions of the equation: 1 – 0 1 see Example 6.6 2 detT  – I3 ===det 2 2– 1 –02 –  1 0 1–  We now determine the eigenspaces associated with the two eigenval- ues,  = 0 and  = 2 . Finding E0 : E0 =kerkerT – 0I ==T ax2 ++bx c Tax2 ++bx c = 0 P2

ac+ x2 ++02a ++2bcxac+ = x2 ++0x 0 Equating coefficients brings us to the following homogeneous system of equations: a b c a b c 1 0 1 ac+0=  101  1 rref 0 1 –--- 2a ++2bc= 0  221 2  ac+0=  101 0 0 0 6.2 Eigenspaces 227

As shown in Example 6.6, –2rr2r r   is the solution set of the above system of equations. Thus: E0 = – 2rx2 ++rx 2rr  with basis – 2x2 ++x 2 Finding E2 : E2 = kerT – 2I = ax2 ++bx c T – 2I ax2 ++bx c = 0 P2 P2

ac+ x2 ++2a ++2bcxac+ – 2ax2 ++bx c = 0 ==– a + c x2 ++2ac+ xac– 0x2 ++0x 0 Equating coefficients: a b c a b c – a +0c =  –1 0 1 100  rref 2ac+0=   2 0 1 001  ac–0=  1 0 – 1 000 As shown in Example 6.6, 0r 0 r   is the solution set of the above system of equations. Thus: E2 = rx r   with basis x

CHECK YOUR UNDERSTANDING 6.10 Find the eigenspaces of the linear operator of Example 6.10 using Answer: See page B-25.  = x2 ++x 1 x + 11 instead of  = x2x 1 . 228 Chapter 6 Determinants and Eigenvectors

EXERCISES

Exercises 1-14. Determine the eigenvalues and corresponding eigenspaces of the given matrix.

1.11 2.14 3. 71– 234 62 21– 62 4. 230 005

32– 1 1111 1000 2 1 000 5. 26– 2 6.1111 7. 0200 01– 000 –21 –3 1111 31–0 1 8. 0 0 401 1111 –0022 0 0 530 0 0 008 A factorization for the characteristic polynomial in the next six exercises can be obtained with the help of the following result: A zero of a polynomial px is a Since –1 is a zero px= x3 –67x – , number which when substituted for x – –1 = x + 1 must be a factor, and we the variable x yields zero. For have: example, –1 is a zero of the poly- x2 –6x – 3 nomial px= x3 –23x – , since x + 1 x –67x – 3 2 p–1 = 0 . One can show that: c x + x –7x2 –6x – is a zero of a polynomial if and only – x2 – x if xc– is a factor of the polyno- –66x – mial. –66x – The adjacent example illustrates So: x3 –67x – = x + 1 x2 –6x – how the above result can be used to factor certain polynomials. = x + 1 x + 2 x – 3

20 1 401 12– 2 9.A = 03 4 10.A = 232 11. A = –522 – 00 1 –021 –636 –

32– 2 –1101 42–2 2 12. A = –13 –3 13.A = 000– 1 14. A = 131– 1 120 0100 0020 011– 1 11–5 3 6.2 Eigenspaces 229

Exercises 15-31. Determine the eigenvalues and corresponding eigenspaces of the given linear operator. 15.T:  given by Tx= –5x . 16.T: 2  2 where T10 = 20 and T01 = 11 . 17.T: 2  2 given by Tab = 8a – 6b 12a – 19b . 18.T: 3  3 given by Tabc = 0 ac+  3bc– . 19.T: 3  3 given by Tabc = a –99b + ca –35b + c 2a –46b + c . 20.T: 4  4 , where Tabcd = adbc . 21.T: 4  4 , where Tabcd = 2aa – b 3a + 2cdab–  – + c – 2d .

22.T: P1  P1 , where Tax+ b = ab+ xb– .

23.T: P1  P1 , where T1 = x and Tx= 1 . 2 24.T: P2  P2 given by Pax++bx c = –2bx – c . 2 2 25.T: P2  P2 given by Pax++bx c = – cx + bx– a . 2 2 26.T: P2  P2 , if Tx= 3x –42x + , Tx= 7x – 8 , and T1 = 1 . 3 2 3 2 27.T: P3  P3 , if Tax+++bx cx d = ad+ x +++– 2a – c + d x 2c – 2d xb– d .

ab –0c 28.T: M22  M22 given by T= . cd ad–

10 00 01 –29 – 00 00 29.T: M22  M22 , where T = , T = , T = , and 01 11– 00 00 10 –12

T 00 = 00. 01 –27

30.I: VV , where I is the identity map: Iv = v . 31.Z: VV , where Z is the zero map: Zv = 0 .

32. (Calculus Dependent) Let V be the vector space of differentiable functions, and let D: VV be the derivative operator. Show that v = e2x is an eigenvector for D. 33. Prove that a square matrix A is invertible if and only if 0 is not an eigenvalue of A. 34. Let A be an invertible matrix with eigenvalue   0 and corresponding eigenvector v. Prove 1 that --- is an eigenvalue of A–1 with corresponding eigenvector v. 

35. Let 1 and 2 be distinct eigenvalues of AM nn . Prove that E1  E2 = 0 230 Chapter 6 Determinants and Eigenvectors

36. (a) Show that similar matrices have equal characteristic polynomials (see Definition 5.11, page 195). (b) Let T: VV be a linear operator on a finite dimensional space V. Show that if  and  are bases for V then: T  and T  have equal characteristic polynomials.

–1 37.Let AP  Mnn , with P invertible. Prove that if v is an eigenvector of A, then P v is an eigenvector of P–1AP .

38. Let A = ab . Prove that a, and c are eigenvalues of A. 0 c

39. For A = ab , find necessary and sufficient conditions for A to have: cd (a) Two eigenvectors. (b) One eigenvector. (c) No eigenvector.

a11 a12 a13 40. Let A = 0 a22 a23 . Prove that a11 a22 , and a33 are eigenvalues of A. 00a33

41. (a) Let T: VV be a linear operator with eigenvalue  . Prove that: E = vv is an eigenvector of T  0

(b) Let AM nn with eigenvalue  . Prove that: E = vv is an eigenvector of A  0

42. For AM nn , show that nullA – I = kerTA – I . 43. Prove that 0 is an eigenvalue for a linear operator T: VV if and only if kerT  0 . 44. Show that if v is an eigenvector for the linear operator T: VV be a linear operator, then so is rv for any r  0 . 45. Let T: VW be an isomorphism. Show that v is an eigenvector in V if and only if Tv is an eigenvector in W. 46. Let v be an eigenvector for the linear operators T: VV and L: VV . Show that v is also an eigenvector for the linear operator LT: VV . Find a relation between the eigenvalues corresponding to v for T, L, and LT .

47. Show that if 1 and 2 are distinct eigenvalues of a linear operator T: VV , then E1  E2 = 0 .

48. Let v1 and v2 be eigenvectors corresponding to distinct eigenvalues 1 and 2 of a linear operator T: VV . Show that v1 v2 is a linearly independent set. 6.2 Eigenspaces 231

49. Let T: VV be a linear operator on a vector space V of dimension n, and let L: V  n be an isomorphisms. Prove that  is an eigenvalue of T if and only if  is an eigenvalue of –1 n the matrix AL= TL S , where S is the standard basis of  , and that E = L–1v v  nullA – I . 50. Let T: VW be an isomorphism. Show that if v is an eigenvector of the linear operator L: VV , then Tv is an eigenvector of the linear operator TLT–1 : WW . 51. Let  be a basis for a space V of dimension n, and L: VV a linear operator. Prove that if v  V is an eigenvector of T with eigenvalue  , then v  is an eigenvector of T : Rn  Rn with eigenvalue  . L 

52. Show that if  is an eigenvalue of AM n  n then  is also an eigenvalue of the transpose AT . (See Exercise 19, page 162)

53. Show that if AM nn is nilpotent, then 0 is the only eigenvalue of A. (See Exercise 23, page 163.)

54. Show that the characteristic polynomial of AM 22 can be expressed in the form 2 –detTraceA + A , where Trace(A) denotes the trace of A (see Exercise 24, page 163).

55. Let AM 22 . Prove that the characteristic polynomial of A is of the form p = 2 ++b detA , and that A2 ++bA detA I = 0 . (This is the Cayley-Hamilton Theorem for square matrices of dimension 2.)

56. (PMI) Let AM nn . Use the Principle of Mathematical Induction to show that the coeffi- cient of the leading term of the characteristic polynomial of A is 1 .

57. (PMI) Let AM nn . Show that the constant term of the characteristic polynomial of A is detA .

58. (PMI) Let 12 k be the distinct eigenvalues of A for AM mm . Prove that n n n n 12 k are the distinct eigenvalues of A . 59. (PMI) Let A be a square matrix with eigenvalue  and corresponding eigenvector v. Show that for any positive integer n, n is an eigenvalue of An with corresponding eigenvalue v.

60. (PMI) Let A be a square matrix with eigenvalue  and corresponding eigenvector v. Show that for any integer n, n is an eigenvalue of An with corresponding eigenvalue v. 232 Chapter 6 Determinants and Eigenvectors

61. (PMI) Let  be an eigenvalue for a linear operator T: VV . Use the Principle of Mathe- matical Induction to show that n is an eigenvalue for T n: VV , where Tn is defined 1 k + 1 k inductively as follows: T = T , and T = TT .

PROVE OR GIVE A COUNTEREXAMPLE

62. If  is an eigenvalue for T: VV then it is also an eigenvalue for kT : VV , where kT v = kTv .

63. If  is an eigenvalue for the two operators T: VV and L: VV , then it is also an eigenvalue for the operator TL+ : VV , where TL+ v = Tv + Lv .

64. For AB  M22 , if A and B are eigenvalues of A and B, respectively, then A + B is

an eigenvalue of TAB+ .

65. For AB  M22 , if A and B are eigenvalues for A and B, respectively, then AB is an eigenvalue for AB.

66. If  is an eigenvalue of the linear operator T: VV , then 2 is an eigenvalue of TT : VV .

67. If T and L are eigenvalues for the linear operators T: VV and L: VV , respectively,

then AB is an eigenvalue for TL : VV .

68. If T and L are eigenvalues for the linear operators T: VV and L: VV , respectively,

then A + B is an eigenvalue for TL+ : VV .

69. If v is an eigenvector for T: VV and L: VV , then v is also an eigenvector for TL+ : VV .

70. If T: VV is a linear operator with eigenvector v, then v + rv is also an eigenvector of T for every r   .

2 71. For AM 22 , A = 0 if and only if 0 is the only eigenvalue of TA .

72. Let T be a linear operator on a vector space V of dimension n. Let  be an eigenvalue for T

and let v1v2  vm be a basis for E . Then, for any r   ,  + r is an eigenvalue for

TrI+ n , and v1v2  vm is a basis for E + r . 6.3 Diagonalization 233

6

§3. DIAGONALIZATION

We begin with: DEFINITION 6.8 A square matrix Aa= ij nn for which a 0  0 11 DIAGONAL MATRIX aij = 0 if ij is said to be a diagonal 0 a22  0    matrix (see margin). 00 a DIAGONALIZABLE nn A linear operator T: VV on a finite OPERATOR dimensional vector space V is said to be diagonalizable if there exists a basis  for

which T  is a diagonal matrix. There is an intimate connection between diagonalizable matrices and eigenvectors, namely: THEOREM 6.11 Let T: VV be a linear operator on a finite dimensional vector space. Then: T is diagonalizable if and only if there exists a basis for V consisting of eigenvectors of T.

PROOF: Assume that T is diagonalizable. Let  = v1v2  vn be such that T  is a diagonal matrix: T  = aij with aij = 0 th for ij . Since the i column of T  consists of the coefficients of the vector Tvi with respect to the basis  , we have: Tvi ==0v1 ++ 0vi – 1 +aiivi +0vi + 1 ++ 0vn aiivi

From the above we see that vi is an eigenvector for T corresponding to the eigenvalue aii .

Conversely, let  = v1v2  vn be a basis for V consisting of eigenvectors, and let 12 n be the eigenvalues corresponding to v1v2  vn . From:  Tvi ==ivi 0v1 +++++0v2 ivi  0vn we have (see Definition 5.10, page 179):

1 0  0 The i ‘s can be zero and need not be distinct (sev- 0 2  0 T = eral of the eigenvectors     in  may share a com- 00  n – 1 0 mon eigenvalue). 00n 234 Chapter 6 Determinants and Eigenvectors

CHECK YOUR UNDERSTANDING 6.11

Let T: 3  3 be the linear map given by: Tabc = 3ab– – c 2a – 2c 2ab– – c Show that  = 120 122 211 is a basis for 3 con- sisting of eigenvectors of T. Determine T  and show that its diag- Answer: See page B-25. onal elements are eigenvalues of T.

In our quest for bases consisting of eigenvectors, we note that:

THEOREM 6.12 If 12 m are distinct eigenvalues of a

linear operator T: VV , and if vi is any

eigenvector corresponding to i , for

1 im, then v1v2  vm is a linearly independent set.

PROOF: By induction on m:

If m = 1 , then v1 consists of a single nonzero vector and is there- fore linearly independent (Exercise 33, page 92). Assume the assertion holds for mk= (the induction hypothesis).

Let v1v2  vk + 1 be a set of eigenvector corresponding to distinct

eigenvalues 12 k + 1 . We are to show that v1v2  vk + 1 is a linearly independent set. With this in mind, we consider the linear combination:

a1v1 +++a2v2  akvk + ak + 1vk + 1 = 0 (*) Applying T to both sides, we have:

Ta1v1 +++a2v2  akvk + ak + 1vk + 1 = T0 Since vi is an eigenvector a Tv +++a Tv a Tv + a Tv = 0 corresponding to i : 1 1 2 2 k k k + 1 k + 1 a  v +++a  v  a  v + a  v = 0 (**) Tvi = ivi 1 1 1 2 2 2 k k k k + 1 k + 1 k + 1

Multiply both sides of (*) by k + 1 :

a1k + 1v1 +++a2k + 1v2  akk + 1vk + ak + 1k + 1vk + 1 = 0 (***) Subtract (***) from (**):

a11 – k + 1 v1 +++a22 – k + 1 v2  akk – k + 1 vk = 0

By the induction hypothesis, the k eigenvectors v1v2  vk corre- sponding to the distinct eigenvalues 12 k are linearly inde- pendent. Consequently:

a11 – k + 1 ====a22 – k + 1  akk – k + 1 0 6.3 Diagonalization 235

Since the eigenvalues 12 k + 1 are distinct, none of the above i – k + 1 is equal to 0. Hence:

a1 ====a2  ak 0 Returning to (*), we then have:

ak + 1vk + 1 = 0

Being an eigenvector, vk + 1  0 , and therefore ak + 1 = 0 (Theorem 2.8, page 54). We have just observed that if you take eigenvectors corresponding to different eigenvalues you will end up with a linearly independent set of vectors. More can be said:

THEOREM 6.13 Let 12 m be distinct eigenvalues of a linear operator T: VV , and let S = v v  v be any linearly inde- i i1 i2 iri pendent subset of Ei . Then:

SS= 1 S2  Sm is a linearly independent set.

PROOF: Consider the vector equation:

The r1 vectors in S1 The rm vectors in Sm linear combination = 0 a v ++ a v ++a v ++ a v = 0 (*) 11 11 1r1 1r1 m1 m1 mrm mrm (we will show that every coefficient must be zero) For 1 im , let v = a v ++ a v . Assume, without loss of i i1 i1 iri iri generality, that vi  0 for 1 it and that the rest are zero vectors. As for any of the zero vectors, v ==a v ++ a v 0 , its i i1 i1 iri iri coefficients must be zero, as v v  v is given to be a i1 i2 iri linearly independent set. As for the nonzero vectors, we begin by rewriting (*) in the form:  v1 ++v2 +vt = 0 (**)

Since the nonzero vectors v1v1  vt are eigenvectors associated

with distinct eigenvalues 12 t , they are linearly independent

(Theorem 6.12). It follows, from (**), that each vi must be 0: v ==a u ++ a u 0 i i1 i1 iri iri Using, again, the fact that each S = u u  u is a linearly each coefficient is 0 i i1 i1 iri independent set, we conclude that each scalar aij must be zero CHECK YOUR UNDERSTANDING 6.12

Let T: VV be a linear operator on a space of dimension n. Prove Answer: See page B-26. that if T has n distinct eigenvalues, then T is diagonalizable. 236 Chapter 6 Determinants and Eigenvectors

RETURNING TO MATRICES

To say that T: VV is diagonalizable is to say that V contains a basis consisting of eigenvectors of T (Theorem 6.11). Let’s modify this characterization to accommodate matrices:

DEFINITION 6.9 A matrix AM nn is diagonalizable if DIAGONALIZABLE n MATRIX there exists a basis for  consisting of eigenvectors of A. Here is a link between diagonalizable matrices and diagonalizable linear operators:

THEOREM 6.14 AM nn is diagonalizable if and only if the n n linear map TA:    given by TAXAX= is diagonalizable. (X is a vertical n-tuple: a column matrix.)

n n PROOF: The linear map TA:    is diagonalizable if and only if: n there exists a basis X1X2  Xn of  , and scalars

12 n , such that TAXi = iXi if and only if:

AXi = iXi (Definition of TA ) if and only if A is diagonalizable (Definition 6.9). Definition 6.9 is okay, but just where do diagonal matrices come into play? Here:

This theorem asserts that THEOREM 6.15 Let AM nn be diagonalizable. Let any diagonalizable matrix n is similar to a diagonal  = X1X2  Xn be any basis for  matrix. The converse also consisting of eigenvectors of A, with associated holds (Exercise 37). And eigenvalues    . If PM is the so we have: 1 2 n nn th AM nn is diagonal- matrix whose i column is Xi , then: izable if and only if it is DP= –1AP where Dd=  is the diag- similar to a diagonal ij matrix. onal matrix, with diagonal entry dii = i .

PROOF: Let Sn = e1e2  en denote the standard basis for n (see page 94). Employing Theorem 5.26 (page 193), and The- n n orem 5.23 (page 184) to the linear map TA:    given by TAX = AX, we have: 6.3 Diagonalization 237

T = P–1T P (*) A  A SnSn th n T n where PI=  (see margin). Since IX= X , the i col-  A  Sn i i   umn of P is simply the vector Xi (recall that Sn is the standard IIbasis). Now: th n TA n The i column of T is:   A SnSn Sn Sn T e ==Ae the ith column of A (See page 193) A i Sn i Sn Hence: T = A . A SnSn Since  is a basis of eigenvectors:

TAXi == iXi 0X1 +++++0X2 iXi  0Xn

It follows that TA  is the diagonal matrix Dd= ij with dii = i . Putting all of this together we have (see *): DP==–1AP, or: APDP–1

EXAMPLE 6.11 Show that the matrix: 02–0 2 A = 110– 1 –121 –1 –121 –1 is diagonalizable, and find a matrix P such that P–1AP is a diagonal matrix.

SOLUTION: In Example 6.7, page 221, we found that 0, 2, and –2 are the eigenvalues of A. We also observed that –1110 1001 is a basis for E0 , 1100 is a basis for E2 , and that 1011 is basis for E–2 . It is easy to see that the four eigen- vectors –10101101 1100 1011 are lin- early independent, and therefore constitute a basis for 4 . Taking P to be the matrix with columns the above four eigenvectors: –1111 P = 1010 1001 0101 we have: 238 Chapter 6 Determinants and Eigenvectors

02–0 2 000 0 P–1 110– 1P = 000 0 –121 –1 002 0 –121 –1 000– 2 Q CHECK YOUR UNDERSTANDING 6.13 Q–1AQ Determine if the given matrix is diagonalizable. If it is, use Theorem 6.15 to find a matrix P such that P–1AP is a diagonal matrix. –011 32– 1 (a) A = –301 (b) A = 26– 2 Answer: See page B-26. –1314 – –21 –3

The next result plays an important role in many eigenvector applica- tions:

THEOREM 6.16 If AM mm is diagonalizable with APDP= –1 , then An = PDnP–1 .

PROOF: (By induction on n) For n = 1 we have:

P–1AP= D1  APDP= –1

Assume that Ak = PDkP–1 (the induction hypothesis). Then:

P–1AP= D  A = PDP–1

Ak + 1 ==AAk APDkP–1 =PDP–1 PDkP–1

induction hypothesis ==PDDk P–1 PDk + 1P–1 Answer: 0 0 0 0 CHECK YOUR UNDERSTANDING 6.14 0 0 0 0 0 01024 0 10 0 0 0 1024 Calculate A for the diagonalizable matrix A of Example 6.11.

ALGEBRAIC AND GEOMETRIC MULTIPLICITY OF EIGENVALUES

An eigenvalue 0 of a matrix AM nn (or of a linear operator T on a vector space of dimension n) has algebraic multiplicity k if k  0 –  is a factor of A’s (or T ’s) characteristic polynomial, and k + 1  0 –  is not. We also define the geometric multiplicity of 0 to be the dimension of E0 (the eigenspace corresponding to 0 ). 6.3 Diagonalization 239

EXAMPLE 6.12 Find the algebraic and geometric multiplicity of the eigenvalues of the matrix: 1 0 1 A = 2 2 1 1 0 1

SOLUTION: In Example 6.6, page 220, we showed that –2 –  2 is the characteristic polynomial of A. It follows that the eigenvalue 0 has algebraic multiplicity 1, and that the eigenvalue 2 has algebraic multiplicity 2. Since both of the eigenspaces E0 and E2 were seen to have dimension 1, the geometric multiplicity of both eigen- values is 1. The above example illustrates the fact that the geometric multiplicity of an eigenvalue can be less than its algebraic multiplicity; it cannot go the other way around:

THEOREM 6.17 If 0 is an eigenvalue of AM nn with algebraic multiplicity ma and geometrical multiplicity mg , then mg  ma .

PROOF: (By contradiction) Assume that ma  mg and let v v  v be a basis for E . Expand v v  v to 1 2 mg 0 1 2 mg a basis  = v v  v v  v for n . Since, for 1 2 mg mg + 1 n Recall that T : Rn  Rn is A 1 img : the linear operator given by: TAvi ==Avi 0vi TAv = Av = 0v1 ++ 0vi – 1 +0vi +0vi + 1 ++ 0vn

the matrix T is of the following form: Recall that the ith column A  m of T consists of the g A   coefficients of the vector  0  0  0  T v A i with respect to the 0 0  0     mg X basis  = v v  v .  1 2 n  TA  = 00 0   0 0  0 0 0  0    Y

0 0  0 nn In the proof of Theorem 6.15 we observed that T = A (where A SnSn n Sn is the standard basis in  ). It follows, from Exercise 36(a), page 230, that the characteristic polynomial of A equals that of TA  , namely, detTA  – I : 240 Chapter 6 Determinants and Eigenvectors

mg cI X  det = crdetY  –0  0  0  Exercise 40, page 217 0 Y 0  –   0  (see margin) 0 m X

 g     m det 00 0 –   =  –  gdetY  0 0 0  0 0 0  0    Y 0 0  0 nn

This leads to a contradiction, for the factor 0 –  cannot appear with exponent greater than ma in the characteristic polynomial of A (remember that ma is the algebraic multiplicity of 0 ). In certain cases, the algebraic and geometric multiplicities of a linear operator can be used to determine if the operator is diagonalizable: THEOREM 6.18 Assume that the characteristic polynomial of a linear operator T: VV (or of a matrix), can be factored into a product of linear fac- tors (with real coefficients). Then T is diago- nalizable if and only if the algebraic multiplicity of each eigenvalue of T is equal to its geometric multiplicity.

PROOF: Assume that T is diagonalizable. By Theorem 6.11, there exists a basis  = v1v2  vn for V consisting of eigenvectors of T. Let 12 k be the set of T’s (several of the vi ’s may cor- respond to the same eigenvalue). If necessary, reorder  so that  = S1 S2  Sk , where Sj consists of the eigenvectors in  corresponding to  j . Since the vectors in Sj are linearly independent, and since their com- bined sum equals the dimension of V, it follows, from Theorem 6.13, that the number of vectors in Sj must equal gj = dimE j , the geo-  metric dimension of j . Hence: ng= 1 +++g2 gk . We are given that the characteristic polynomial of T can be factored into a product of linear factors:

a1 a2 ak (*) detT  – I = – 1 – 2 – k It follows that:   ng==1 +++g2 gk  a1 +++a2 ak n

Theorem 6.17 degree of the characteristic polynomial (*) 6.3 Diagonalization 241

Consequently:   a1 +++a2 ak = g1 +++g2 gk Or: a1 – g1 +++0a2 – g2 ak – gk =

Knowing that ai  gi (Theorem 6.17), we conclude that ai = gi for each 1 ik.

Conversely, assume that the multiplicity of each eigenvalue  j , 1 jk, is equal to its degree: aj ==dimEj gj . Let Sj be a basis for Ej . By Theorem 6.13, the set S1 S2  Sk , which  contains g1 +++g2 gk vectors, is linearly independent. It is in fact a basis, since it contains n vectors (Theorem 3.11, page 99): degree of the characteristic polynomial   g1 +++g2 gk ==a1 +++a2 ak n Possessing a basis of eigenvectors, T is diagonalizable (Theorem 6.11).

EXAMPLE 6.13 Appeal to the previous theorem to show that the linear operator T: 4  4 given by: Tabcd = 2b – 2ca + b– d– a ++b – 2c d – a ++b – 2c d

is diagonalizable. Find a basis  for which T  is a diag- onal matrix.

SOLUTION: For S the standard basis of 4 , we have: T1000 = 01– 1 – 1 02–0 2 110– 1 T SS = –121 –1 –121 –1 The above matrix was encountered in Example 6.7, page 221, where we found its characteristic polynomial to be 2 – 2 + 2 . We also showed that: The eigenspace corresponding to the eigenvalue 0, of multiplicity 2, has dimension 2 (with basis –1110  1001 ). The eigenspace corresponding to the eigenvalue 2, of multiplicity 1, has dimension 1 (with basis 1100 .) The eigenspace corresponding to the eigenvalue –2 , of multiplicity 1, has dimension 1 (with basis 1011 ). Since the algebraic multiplicity of each eigenvalue equals its geometric multiplicity, the linear operator is diagonalizable. Moreover, since  = –1110 1001 1100 1011 242 Chapter 6 Determinants and Eigenvectors

is a basis for V of eigenvectors, we know that T  will be a diagonal matrix with diagonal entries equal to the eigenvalues corresponding to the eigenvalues of  ; namely: 000 0 000 0 T  = 002 0 000–2

CHECK YOUR UNDERSTANDING 6.15 Verify that the algebraic multiplicity of each eigenvalue of the diago- nalizable matrix of CYU 6.13(b) equals that of its geometric multi- Answer: See page B-27. plicity. 6.3 Diagonalization 243

EXERCISES

Exercises 1-19. Determine if the given linear operator T: VV is diagonalizable. If it is, find a basis  for V such that T  is a diagonal matrix. (Note: To factor the characteristic polynomial of the given operator, you may need to use the division process discussed above Exercise 9 on page 228.)

1.T: 2  2 given by Tab = 2aa + 2b . 2.T: 2  2 given by Tab = 7ab–  6a + 2b . 3.T: 2  2 given by Tab = – 4a +8b – a + 2b . 4.T: 2  2 where T10 = 41 – and T01 = 12 . 5.T: 2  2 where T11 = 12 and T01 = 20 . 6.T: 3  3 given by Tabc = –a 2c 2b + 3c . 7.T: 3  3 given by Tabc = 13a – 4b 8b – 2c 5c . 8.T: 3  3 where T100 = 011 –  – , T010 = 145 , and T001 = 112 –  – . 9.T: 3  3 where T111 = 101 , T011 = 110 , and T002 = 110 . 10.T: 3  3 given by Tabc = 2a +34b – 4c – b + 5c – 6b + 8c . 11.T: 4  4 given by Tabcd = badc . 12.T: 4  4 given by Tabcd = –a0 a + 2c –2b . 13.T: 4  4 where T1000 = 1000 , T0100 = 32– 00 , T0010 = 2 – 140 , and T0001 = 605– 3 . 14.T: 4  4 where T1100 = 1001 , T0011 = 0010 , T1001 = 0111 , and T0001 = 1111 . 15.T: 5  5 given by Tabcde = 2aa – b 4c + 5d 3dc + 8e .

16.T: P2  P2 given by Tpx = px+ 1 .

2 2 17.T: P2  P2 given by Tax++bx c = 3a – 2b + c x ++2bc– xb .

2 2 2 18.T: P2  P2 where Tx= x , Tx= 2x + 1 and T1 = x – 1 .

ab ab– 19.T: M22  M22 given by T= . cd 3c 0 244 Chapter 6 Eigenvectors and Diagonalization

Exercises 20-34. Determine if the given matrix A is diagonalizable. If it is, find a matrix P such that P–1AP is a diagonal matrix.

20.A = 11 21.A = 16– 22. A = 24– 11 –21 33–

135 500 2 34 23.A = 02–6 24.A = –501 25. A = 2 30 004 232 0 05

203 000 12– 1 26.A = –031 27.A = 123 28. A = 24– 2 102 –231 – –21 –1

1234 5000 02–0 2 29.A = 02–3 1 30.A = 2200 31. A = 110– 1 0051 –9101 –121 –1 0001 3357 –121 –1

253– 6 312–4 2 21000 32. A = –4022 01442 01– 000 –104 –6–12 33.A = 00211 34. A = 00401 173– 5 00030 00530 00003 00008

2 35. Let AM nn be such that A = I . Show that: (a) If  is an eigenvalue of A, then  = 1 or  = 0 . (b) A is diagonalizable.

36. Let AM nn be diagonalizable. Prove that the rank of A is equal to the number of nonzero eigenvalues of A.

37. Prove that if AM nn is similar to a diagonal matrix, then A is diagonalizable.

T 38. Let AM nn . Prove that A and its transpose A have the same eigenvalues, and that they occur with equal algebraic multiplicity (see Exercise 19, page 161).

39. Let AM nn . Prove that if  is an eigenvalue of A with geometric multiplicity d, then  is an eigenvalue of its transpose AT with geometric multiplicity d (see Exercise 19, page 161). 40. Let L: VW be an isomorphism on a finite dimensional vector space. Prove that: (a) The linear operator T: VV and LTL–1: WW have equal characteristic polyno- mials. (b) The eigenspace corresponding to an eigenvalue  of LTL–1 is isomorphic to the eigenspace corresponding to that eigenvalue of T. (c) T is diagonalizable if and only if LTL–1 is diagonalizable. 6.3 Diagonalization 245

PROVE OR GIVE A COUNTEREXAMPLE

41. Let T: VV be a linear operator on a space of dimension n. If 12 m are distinct eigenvalues of T, and if there exists a basis  for V such that T  is a diagonal matrix, then mn= .

42. Let 12 k be the distinct eigenvalues of a linear operator T: VV on a vector space V of dimension n. The operator T is diagonalizable if and only if kn= .

43. If AB  Mnn are both diagonalizable, then so is AB .

44. If AB  Mnn are such that AB is diagonalizable, then both A and B are diagonalizable. 246 Chapter 6 Determinants and Eigenvectors

6

§4. APPLICATIONS

Applications of eigenvectors surface in numerous fields. In this sec- tion we focus on recurrence relations, and on differential equations.

FIBONACCI NUMBERS AND BEYOND

The Fibonacci sequence is that sequence whose first two terms are 1, Leonardo Fibonacci (Ital- and with each term after the second being obtained by summing its ian; circa 1170 - 1250), is two immediate predecessors: considered by many to be 11+ the best mathematician of 813+ the Middle Ages. The 1123581321 34 55  sequence bearing his name 12+ 34+ 55 evolved from the follow- ing question he posed and What is the 100th Fibonacci number? Given enough time, we could resolve in 1220: keep generating the numbers of the sequence, eventually arriving at its Assume that pairs of rab- th bits do no produce off- 100 element. There is a better way: spring during their first th Letting sk denote the k Fibonacci number we have s1 ==s2 1 , month of life, but will produce a new pair of and, for k  3 : offspring each month sk = sk – 1 + sk – 2 thereafter. Assuming that We observe that s is the top entry of the matrix product: no rabbit dies, how many k pairs of rabbits will there s s + s s be after k months? 11 k – 1 ==k – 1 k – 2 k (*) 10 sk – 2 sk – 1 sk – 1

11 sk Letting F = and Sk = we can express (*) in the form: 10 sk – 1

Sk = FSk – 1 In particular 1 1 2 1 k – 2 1 S3 ==FS2 F  S4 ==FS3 FF =F  Sk =F 1 1 1 1 th Note that the k Fibonacci number sk in (*) is the top entry in the

sk k – 2 1 matrix Sk ==F , which is simply the sum of the s 1 ab 1 ab+ k – 1 = k – 2 cd 1 cd+ entries in the first row of F (see margin). But this is of little benefit

unless we can readily find the powers of the matrix F = 11 . We can: 10 6.4 Applications 247

Since the characteristic polynomial of F is 2 –1 – (margin), the 11 10 15– 15+ det–  = 0 matrix F has eigenvalues 1 = ------and 2 = ------. Let’s find 10 01 2 2 eigenvectors associated with those eigenvectors: 1 –1  det = 0 15+ 15– 1 – ------a ------ab+0=  15+ 2  11 a = ------a  ab+ = 2  2 –1 – = 0 2  10 b b a 15+ 15+  114 + ------b a–------b = 0   = ------2 2 2 homogeneous system of equations  a b a b solution 15– ------1 rref 15+ 15+ 2 1 –------ a = ------b 2 2 1 –------15+ 00 2

15+ Setting the free variable b to 1 we find that a = ------. It follows that 2 a ------15+ = 2 is an eigenvector associated with the eigenvalue 1 . b 1

------15– ------15– -1+ In the same manner one can show that is an eigenvector 11 15------– - 2 2 2 = 10 1 1 ------15– - 2 associated with the eigenvalue 2 — a fact that is verified in the margin. Theorem 6.15, page 236, tell us that: ------35– - = 2 a diagonal matrix 15– ------–1 15+ 2 15+ 15– 15+ 15– ------0 ------11 ------2 D ==2 2 2 2 15– 15– 15– = ------11 10 11 0 ------2 2 2 1 P–1FP Leading us to:

15+ –1 15+ 15– ------0 15+ 15– –1 ------2 ------F ==PDP 2 2 2 2 15– 11 0 ------11 2 Applying Theorem 6.16, page 238, we have:

k 15+ –1 15+ 15– ------0 15+ 15– k k –1 ------2 ------F ==PD P 2 2 2 2 15– k 11 0 ------11 2 248 Chapter 6 Determinants and Eigenvectors

You are invited to show that the first row of the above matrix product can be expressed in the following form:

15+ k + 1 15– k + 1 15+ k 15– k ------–------------– ------1   Fk = ------2 2 2 2 5 ********************** ****************** Recalling that the kth Fibonacci number is the sum of the entries in the first row of Fk – 2 , we have: k – 1 k – 1 k – 2 k – 2 k – 1 k – 2 15+ 15+ 1 15+ 15– 15+ 15– ------– ------sk = ------– ------+ ------– ------2 2 5 2 2 2 2 15+ k – 2 15+ = ------------– 1 2 2 1 15+ k – 1 15– k – 2 15– k – 1 15– k – 1 = ------------– ------– ------– ------15+ k – 2 15+ 2   = ------------5 2 2 2 2 2 2  k k 15+ k – 2 = ------1 15+ 15– 2 = ------– ------(see margin) 5 2 2 Looking at the above “5 -expression” from a strictly algebraic point of view, one would not expect to find that each sk is an integer. Being a Fibonacci number, that must be the case. In particular, the 100th Fibo- nacci number is: 100 100 1 15+ 15– s100 ==------– ------354,224,848,179,261,915,075 5 2 2

Using the TI-92: 15+ The number  = ------has an interesting history dating back to the 2 time of Pythagoras (c. 500 B.C.). It is called the golden ratio ( is the first letter in the Greek spelling of Phydias, a sculptor who used the golden ratio in his work). Basically, and for whatever aesthetic rea- son, it is generally maintained that the most “visually appealing” partition of a line seg- L l ment into two pieces is that for which the ratio of the length of the longer piece L to the length of the sorter piece l equals the ratio of the entire to that of the longer piece, leading us to: L --L- = Ll------+ - l L L2 – lL –0l2 = l ll+ 5 L 15+ L ==------ ------2 l 2 6.4 Applications 249

th The formula sk = sk – 1 + sk – 2 for the k element of the Fibonacci RECURSIVE RELATION sequence describes each element of the sequence in terms of previous elements. It is an example of a recurrence relation. You are invited to consider addition recurrence relations in the exercises, and in the fol- lowing Check Your Understanding box.

CHECK YOUR UNDERSTANDING 6.16 Answer: th 5 1 Find a formula for the k term of the sequence s1s2 s3  , if s = ---  2k – 1 + ---–1 k – 1 k 3 3 s1 = 2 , s2 = 3 , and sk = sk – 1 + 2sk – 2 for k  3 .

SYSTEMS OF DIFFERENTIAL EQUATIONS (CALCULUS DEPENDENT)

We begin by extending the concept of a matrix to allow for function entries; as with: 2x Ax==3xe and Bx 0 lnx 4 sinx x –25 x The arithmetic of such matrices mimics that of numerical matrices. For example:

2x 2x 3xe+ 0 lnx = 3xe+ lnx 4 sinx x –25 x x – 1 sinx + 2x and:

2x 2x 2x 3xe 0 lnx = e x – 5 3xxln+ 2xe 4 sinx x –25 x sinxx– 5 4lnx + 2xxsin We also define the derivative of a function-matrix to be that matrix obtained by differentiating each of its entry. For example:

2x 2x 2x If Ax= 3xe then Ax ==3x e x 2e 4 sinx 4 sinx 0 cosx In the exercises, you are invited to show that the following familiar derivative properties: fx+ gx = f x + gx fxgx = fxgx + gxf x cf x = cfx extend to matrices: 250 Chapter 6 Determinants and Eigenvectors

THEOREM 6.19 Let the entries of the matrices Ax and Bx be differentiable function, and let C be a matrix with scalar entries (real numbers). Then: (i) Ax+ Bx = A x + Bx (ii) AxBx = AxBx + BxA x (iii) CA x = CAx (assuming, of course, that the matrix dimen- sions are such that the operations are defined) Differential equations of the form: f x = af x  or y = ay play an important role in the development of this subsection. As you may recall:

THEOREM 6.20 The solution set of f x = af x , consists of those functions of the form fx= ceax for some constant c.

PROOF: If fx= ceax , then: f x ==ceax aceax =af x At this point, we know that every function of the form yce= ax is a solution of the differential equation f x = af x . Moreover, if fx is any solution of f x = af x , then the derivative of the function fx gx= ------ is zero: eax since f x = af x fx gxf x – fxgx ------= ------gx 2 eaxf x – fxaeax f x – af x eax 0 gx gx ====------0 e2ax e2ax eax fx If the derivative of a func- It follows that gx==------c for some constant c, or that eax tion is zero, then the func- fx= ceax . tion must be constant.

We now turn our attention to systems of linear differential equation of the form: f1x = a11f1x +++a12 f2x  a1n fnx   f2x = a21f1x +++a22f2x  a2n fnx      fnx = an1f1x +++an2f2x  ann fnx  where the coefficients aij are real numbers. As it is with systems of lin- ear equations, the above system can be expressed in the form: Fx = AF x were Fx= fix  Mn  1 and Aa= ij  Mnn . 6.4 Applications 251

In the event that A is a diagonal matrix, the system Fx = AF x is easily solved: THEOREM 6.21 f x  0  0 f x c e1x f1x = 1f1x  1 1 1 f1x 1  f x 0   0 f x f x = f x 2 2 2 f x 2x     2 2 2  =  2 c e  If then: = 2   fnx 0  0 n fnx f x fnx = nfnx n nx  cne where c1c2 cn  . PROOF: Simply apply Theorem 6.20 to each of the n differential

equations: fix = ifix . We now consider systems of differential equations of the form: Fx = AF x (*) where A is a diagonalizable matrix. In accordance with Theorem 6.15, page 236, we know that for any chosen basis  = v1v2  vn of eigenvectors of A: APDP= –1 th where the i column of PM nn is the eigenvector vi  a1ia2i  ani with eigenvalue i , and Dd= ij is the diago- nal matrix with dii = i . Substituting in (*), we have: Fx = PDP–1 Fx –1 Multiply both sides by P : P–1Fx = DP–1Fx Theorem 6.19(iii): P–1Fx = DP–1Fx Letting Gx= P–1Fx , brings us to: Gx = DGx Applying Theorem 6.21, we have:

1x 1x 1x c1e c1e c1e  x  x  x 2 –1 2 2 Gx===c2e  P Fx c2e  Fx P c2e

nx nx nx cne cne cne At this point we have: v1 v2 vn

a a  a 1x 11 12 1n c1e

a a  a 2x Fx= 21 22 2n c2e   a a a nx n1 n2 nn cne Appealing to Theorem 5.3, page 154, we conclude that: 1x 2x  nx Fx= c1e v1 +++c2e v2 cne vn 252 Chapter 6 Determinants and Eigenvectors

Summarizing, we have:

THEOREM 6.22 Let AM nn be diagonalizable, and let n v1v2  vn be a basis for  consisting entirely of eigenvectors of A with corre- sponding eigenvalues i . Then, the general solution of: Fx = AF x is of the form: 1x 2x  nx c1e v1 +++c2e v2 cne vn

for c1c2 cn  .

In alternate notation form: EXAMPLE 6.14 Find the general solution for: f  x = – 3 f x + f x 1 1 2  y1 = –3y1 + y2   f2 x = 6 f1x + 2f2x   y2 = 6y1 + 2y2 

SOLUTION: Our first order of business is to find (if possible) a basis

2 –13 v1 v2 of  consisting of eigenvectors of A = . 62

– 3 –1 From:det = – 3 –  2 –  – 6 62–  ==2 +  – 12  + 4 – 3

we see that the matrix A = –13 is diagonalizable, with eigenvalues 62 –4 and 3. Here are their corresponding eigenspaces: –13 –04 11 E–4 ==null– null 62 04– 66 x y x y homogeneous 11 rref 1 1 system of equations: 66 00 Setting the free variable b equal to r, we have: E–4 = –rr rR And: –13 30 –16 E3 ==null– null 62 03 61–

a b a b 1 –16 1 –--- rref 6 61– 00 Bringing us to: E3 = r 6r rR 6.4 Applications 253

Choosing the eigenvector v = –11 for the eigenvalue –4 and Any other two eigenvec- 1 tors corresponding to the the eigenvector v2 = 16 for the eigenvalue 3, we obtain a basis of two eigenvalues will do 2 consisting of eigenvectors for A. Applying Theorem 6.22, we just as well. conclude that the general solution of the given system of differential equations is give by: –4x 3x y1 –4x –1 3x 1 –c1e + c2e ==c1e + c2e –4x 3x y2 1 6 c1e + 6c2e Which is to say: –4x 3x –4x 3x y1 ==–c1e + c2e and y2 c1e + 6c2e

y1 = –3y1 + y2  Let’s check our result in the given system  : y2 = 6y1 + 2y2 

–4x 3x –4x 3x y1 ==–c1e + c2e 4c1e + 3c2e and: –4x 3x –4x 3x –4x 3x – 3y1 + y2 ==– 3–c1e + c2e + c1e + 6c2e 4c1e + 3c2e

–4x 3x –4x 3x Similarly: y2 ==c1e + 6c2e –4c1e + 18c2e –4x 3x –4x 3x ==6–c1e + c2e + 2c1e + 6c2e 6y1 + 2y2

CHECK YOUR UNDERSTANDING 6.17 Find the general solution for:

y1 = 2y2 – 2y3   y2 = y1 + y2 – y4   y3 = – y1 ++y2 – 2y3 y4   y4 = – y1 ++y2 – 2y3 y4  Answer: See page B-28. Suggestion: Consider Example 6.11, page 237. Let us return momentarily to the system of equations of Example 6.14: –4x 3x y1 = –3y1 + y2  y1 = –c1e + c2e  with general solution: (*) y  = 6y + 2y –4x 3x 2 1 2  y2 = c1e + 6c2e To arrive a particular or specific solution for the system, we need some additional information. Suppose, for example, that we are given the ini-

y 0 tial condition Y0 ==1 –2 . Substituting in (*), we then y20 3 have: 254 Chapter 6 Determinants and Eigenvectors

–4  0  15 3  0 –2 = –c + c c1 = ------–2 = – c1e + c2e  1 2  7    1 3 = c e–4  0 + 6c e3  0  3 = c1 + 6c2  c = --- 1 2  2 7 1 –4x 15 3x y = – ---e + ------e 1 7 7 Solution: y = --1-e–4x + -----90-e3x 2 7 7

CHECK YOUR UNDERSTANDING 6.18 Answer: 2x 2x y1 = 1 – e  y2 = 2 – e Find the specific solution of the system in CYU 6.17, if y = 2 y = 3 3 4 y10 = 0y20 = 1 y30 = 2 and y40 = 3

EXAMPLE 6.15 In the forest of Illtrode lived a small peaceful community of 50 elves, when they were sud- denly invaded by 25 trolls. The wizard Callan- dale quickly determined that: d -----Tt= --1-Tt– --1-Et dt 2 9

d -----Et= –Tt+ --1-Et dt 2 where Tt and Et represents the troll and elves populations t years after the troll inva- sion. Analyze the nature of the two popula- tions as time progresses. SOLUTION: To find the general solution of the system: 1 1  --- –--- T t = 2 9 Tt Et –1 --1- Et 2 we first find the eigenvalues of the above 22 matrix:

1 1  = --5- --- –  –--- 1 1 6 det 2 9 ==--1- –  2 –0---  --1- –  = ---  1 2 9 2 3 1 –1 --- –   = --- 2 6

1 1 5 1 1 5 --- –------0 –--- –--- Then: E --- ==null2 9 – 6 null3 9 6 5 1 –1 --1- 0 --- –1 –--- 2 6 3 x y x y

homogeneous 1 1 1 –--- –--- rref 1 --- system of equations: 3 9 3 –1 –--1- 0 0 3 6.4 Applications 255

Setting the free variable y to 3, we obtain the eigenvector –31 for the eigenvalue --5- . In a similar fashion, you can show that 13 is an 6 1 eigenvector for --- . This leads us to the general solution: 6

--5-t --1-t 5 1 ---t ---t –c e6 + c e6 Tt 6 –1 6 1 1 2 ==c1e + c2e Et 3 3 --5-t --1-t 6 6 3c1e + 3c2e Turning to the initial conditions, we have: T0 25 25 = – c1 + c2  25 125 === c1 –------and c2 ------E0 50 50= 3c1 + 3c2  6 6 Leading us to the specific solution:

5 1 25 ---t 125 ---t Tt= ------e6 + ------e6 6 6 25 --5-t 125 --1-t Et= –------e6 + ------e6 2 2 A consideration of the graphs of the two functions reveals that while the troll population will continue to flourish in the region, the poor elves vanish around two-and-a-half years following the invasion:

trolls

elves

CHECK YOUR UNDERSTANDING 6.19

Turtles and frogs are competing for food in a pond, which currently contains 120 turtles and 200 frogs. Assume that the turtles’ growth rate and the frogs’ growth rate are given by 5Tt Ft 5Ft T t ==------– ------and F t ------– Tt 2 4 2 respectively; where Tt and Ft denote the projected turtle and frog population t years from now. Find, to one decimal place, the number of years it will take for the turtle population to equal that of Answer: 1.3 years the frog population. 256 Chapter 6 Determinants and Eigenvectors

EXERCISES

th Exercises 1-8. Find a formula for the k term of the sequence s1s2 s3  , if:

1. s1 = 2 , s2 = 2 , and sk = sk – 1 + sk – 2 for k  3 .

2. s1 = a , s2 = a , and sk = sk – 1 + sk – 2 for k  3 .

3. s1 = 1 , s2 = 2 , and sk = sk – 1 + sk – 2 for k  3 .

4. s1 = 1 , s2 = 6 , and sk = 6sk – 1 – 9sk – 2 for k  3 .

5. s1 = 1 , s2 = 4 , and sk = 3sk – 1 – 2sk – 2 for k  3 .

2 6. s1 = 1 , s2 = 2 , and sk = ask – 1 – bsk – 2 for k  3 and a – 4b  0 .

7. s1 = 1 , s2 = 2 , s3 = 3 , and sk = 2sk – 1 + sk – 2 – 2sk – 3 for k  4 .

21– 2 3 Hint: Note that S4 = 10 0 2 . 01 0 1

8. s1 = 1 , s2 = 2 , s3 = 3 , and sk = –2sk – 1 ++sk – 2 2sk – 3 for k  4 .

th 2 k + 1 9. (PMI) Let sk denote the k Fibonacci number. Prove that sksk – 2 –1sk – 1 = – , for k  3 . Suggestion: Use the Principle of Mathematical Induction to show that for A = 11 and 10 s s k  3 , Ak = k k – 1 . sk – 1 sk – 2

10. Let s0 and s1 be the first two elements of a sequence and let sk = ask – 1 + bsk – 2 be a recur- rence relation which defines the remaining elements of the sequence. Prove that if the quadratic 2 n n equation  – a –0b = has two distinct solutions, 1 and 2 , then sk = c11 + c22 for some c1 c2  .

Suggestion: Replace the matrix 11 in the development of the Fibonacci sequence with 10 the matrix ab . 10 6.4 Applications 257

11. Let the entries of the matrices Ax and Bx be differentiable function, and let C be a matrix with scalar entries (real numbers). Given that the dimensions of the matrices are such that the operations can be performed, prove that: (i) Ax+ Bx = A x + Bx (ii) AxBx = AxBx + BxA x (iii) CA x = CAx

Exercises 12-17. Find the general solution of the given system of differential equations, then check your answer by substitution.

f1 x = 2f1x  y1 = y1 – y2  12. 13.  f2x = 3f1x – f2x  y2 = 2y1 + 4y2 

y1 = 3y1 + 2y2  f x = fx+ 2gx– hx  14.   y2 = 6y1 – y2  15. gx = 2fx+ 4gx– 2hx  hx = –fx – 2gx+ hx 

x = 4x ++3y 3z   x –261 x 16. y = –3x ++y 8z  17. y = –351 y  z = –86x ++y 6z  z 025 z

Exercises 18-21. Solve the given initial-value problem.

y  y y 0 18. 1 ==31– 1  1 1 y2 62– y2 y20 –1

f x f x f 0 19. 1 ==21– 1  1 2 f2x –21 f2x f20 0

x 02–1x x0 1 20. y ==003 y  y0 0 z 100 z z0 2

f x –261 fx f0 1 21. gx ==–351 gy g0 –1 hx 025 hz h0 0 258 Chapter 6 Determinants and Eigenvectors

22. Given enough space and nourishment, the rate of growth of plants A and B are given by At = --3-At and Bt = --3-Bt , respectively, where t denotes the number of months 2 2 after planting. One year, 50 of A and 30 of B were planted, and in such a fashion that the rates of growth of each of the two plants were compromised by the presence of the other; in accordance with: At = --3-At– Bt and Bt ==--3-Bt --1-At . Analyze the nature of 2 2 4 the two plant populations as time progresses. 23. Assume that initially, tank A contains 20 gallons of a liq- uid solution that is 10% alcohol, and that tank B contains 30 gallons of a solution that is 20% alcohol. At time t = 0 , the mixture in A is pumped to B at a rate of 1 gal- lons/minute, while that of B is pumped to A at a rate of 1.5 gallons/minute. Find the percentage of alcohol con- centration in each tank t minutes later. A B 6.5 Markov Chains 259

6

§5. MARKOV CHAINS Certain systems can occupy a number of distinct states. The transmis- Stochos: Greek for “guess.” sion in a car, for example, may be in the neutral state, or reverse (state), Stochastices: Greek for “one or first gear (state), etc. When chance plays a role in determining the who predicts the future.” current state of a system, then the system is said to be a stochastic pro- Andrei Markov: Russian cess, and if the probabilities of moving from one state to another Mathematician (1856-1922). remain constant, then the stochastic process is said to be a Markov process, or Markov chain. Here is an example of a two-state Markov process: State Y: Person x was involved in an automobile accident within the previous 12 month period. State N: Person x was not involved in an automobile accident within the previous 12 month period. Let’s move things along a bit by citing the following study: Probability of x being involved .23 if x is in Y in an accident within the next = 12 month period } { .19 if x is in N The above information is reflected in Figure 6.2(a) (called a transi- tion diagram), wherein each arrow is annotated with the probability of taking that path. The same information is also conveyed in the transi- Transition matrices are also tion matrix, Aa=  , of Figure 6.2(b), where t represents the called probability matrices. ij ij probability of moving from the jth state to the ith state in the next move. The 0.19 in the upper right-hand corner of the matrix, for exam- ple, gives the probability of moving from state N to state Y in the next Since the entries in the transi- move, while the entry 0.81 is the probability of remaining in state N. tion matrix are probabilities, current state they must lie between 0 and Y N 1 (inclusive). Moreover, since .23 .77 .81 next state the entries down either col- .23 .19 Y umn account for all possible YN A : outcomes (staying in Y, or leaving Y, for example), their .19 .77 .81 N sum must equal 1. In particu- lar, since there is a 0.23 prob- Transition Diagram Transition Matrix ability that a person in state Y (a) (b) returns to state Y, there has to Figure 6.2 be a 0.77 probability that the Assume that, initially, 25% of the population was involved in an auto- person will leave that state and, consequently, move to mobile accident within the previous 12 month period (and 75% was state N. not). This given condition brings us to the so called initial state matrix .25 of the system: S0 = . .75 260 Chapter 6 Determinants and Eigenvectors

Utilizing matrix multiplication we can arrive at the next state, S1 :

T S0 S1 .23 .19 .25 (.23)(.25) + (.19)(.75) .20 Y (*) == .77 .81 .75 (.77)(.25) + (.81)(.75) .80 N The above tells us that there is a 0.20 probability that a person will be involved in an accident in the first 12 month period.

To get to the next state matrix, we replace S0 with S1 in (*): T S1 S2 .23 .19 .20 ==(.23)(.20) + (.19)(.80) .198 Y .77 .81 .80 (.77)(.20) + (.81)(.80) .802 N The above tells us that there is a 0.198 probability that a person will be involved in an accident in the next (second) 12 month period.

T S2 S3 .23 .19 .198 .19792 Y Similarly: = .77 .81 .802 .80208 N

Working backwards, we find that we can also arrive at S3 by multi- 3 plying the initial state matrix S0 by T :

2 2 3 S3 ==TS2 TTS1 =T S1 =T TS0 =T S0 Generalizing, we have: THEOREM 6.23 If T is the transition matrix of a Markov pro- cess with initial-state matrix S0 , then the nth state matrix in the chain is given by: n Sn = T S0

CHECK YOUR UNDERSTANDING 6.20 Of the 1560 freshmen at Bright University, 858 live in the dorms. There is a 0.8 probability that a freshman, sophomore, or junior cur- rently living in the dorms will do so in the following year, and a 0.1 probability that a currently commuting student will live on campus Answer: 757, 686, and 636 next year. Assuming (big assumption) that all 1560 freshmen will of the current freshmen will live in the dorm in their soph- graduate, determine (to the nearest integer) the number of the current omore, junior, and senior freshman that will be living in the dorms in their sophomore, junior, year, respectively. and senior years. 6.5 Markov Chains 261

POWERS OF THE TRANSITION MATRIX Let us formally define the concept of a transition matrix:

DEFINITION 6.10 A transition matrix TM nn is a TRANSITION MATRIX matrix that satisfies the following two properties: (1) T contains no negative entry. (2) The sum of the entries in each column of T equals 1.

Consider the three-state Markov process with transition matrix: I II III .3 .1 .6 I T: .2 .1 .4 II .5 .8 0 III Assume that at the start of the process we are in state II:

0 I S0 = 1 II 0 III I II III .3 .1 .6 0 .1 I Observe that: S1 ==TS0 .2 .1 .4 1 =.1 II .5 .8 0 0 .8 III same 2 .3 .1 .6 0 .41 .52 .22 0 .52 I 2 And that: S2 ==T S0 .2 .1 .4 1 =.28 .35 .16 1 =.35 II .5 .8 0 0 .31 .13 .62 0 .13 III same In general: THEOREM 6.24 Let T denote the transition matrix of a Mar- kov chain. If the process starts in state j, then the element in the ith row of the jth col- umn of T m represents the probability of end- ing up at state i after m steps.

EXAMPLE 6.16 Analyze the nature of the second column of T 2 , T 4 , and T 8 for the transition matrix: I II III .2 .6 .4 I T = .3 .1 .3 II .5 .3 .3 III Given that the system is initially in state II. SOLUTION: 262 Chapter 6 Determinants and Eigenvectors

.2 .6 .4 .2 .6 .4 .42 .30 .38 I 2 T ==.3 .1 .3 .3 .1 .3 .24 .28 .24 II .5 .3 .3 .5 .3 .3 .34 .42 .38 III The second column of T 2 tells us that if you start in state II, then there is a 0.30, 0.28, and 0.42 probability that you will end up at states I, II, and III, respectively, after two steps. .42 .30 .38 .42 .30 .38 .3776 .3696 .3760 I 4 2 2 T ==T  T .24 .28 .24 .24 .28 .24 =.2496 .2512 .2496 II .34 .42 .38 .34 .42 .38 .3728 .3792 .3744 III The second column of P 4 tells us that if you start in state II, then there is a 0.3696, 0.2512, and 0.3792 probability that you will end up at states A, B, and C, respectively, after four steps.

.3776 .3696 .3760 .3776 .3696 .3760 .3750 .3750 .3750 8 4 4 T ==T  T .2496 .2512 .2496 .2496 .2512 .2496 =.2500 .2500 .2500 .3728 .3792 .3744 .3728 .3792 .3744 .3750 .3750 .3750 Whoa! The three columns of T 8 are identical (to four decimal places). Moreover, if you take higher powers if T, you will find that you will again end up at the above T 8 matrix. This suggests that eventually there is, to four decimal places, a 0.375, 0.250, and 0.375 probability, respectively, that you will end up at states I, II, and III, independently of whether you start at state I, or II, or III! Indeed, no matter what initial state you start with, say the state .2 S0 = .5 , it looks like you will still end up at the same situation: .3 .2 .3750 .3750 .3750 .2 .3750 .3750 .3750 8 A T .5 ==.2500 .2500 .2500 .5 .2500 .2500 .2500 B .3 .3750 .3750 .3750 .3 .3750 .3750 .3750 C It appears that for this Markov chain, there is a probability of 0.375 that you will eventually end up in state A, a probability of 0.250 that you will end up in state B, and a probability of 0.375 that you will end up in state C, independently of the initial state of the process! Even more can be said; but first, a definition:

DEFINITION 6.11 n SF   is a fixed state for a transition

matrix TM nn if TSF = SF .

EXAMPLE 6.17 .2 .6 .4 Show that the transition matrix T = .3 .1 .3 .5 .3 .3 of Example 6.16 has a fixed state. 6.5 Markov Chains 263

x SOLUTION: We are to show that there exists a state SF = y such z that TSF = SF : .2 .6 .4 x x .2x ++.6y .4z x .3 .1 .3 y = y  .3x ++.1y .3z = y .5 .3 .3 z z .5x ++.3y .3z z Equating entries brings us to a system of three equations in three unknowns:  –.6.8x ++y .4z = x 0 .2x +.6y +.4z = x   .3x +.1y +.3z = y   .3x –.3.9y +0z =    .5x +.3y +.3z = z  .5x +0.3y – .7z =  It can be shown, however, that for any transition matrix T, the system x x of equations stemming from T y = y will always have more than z z one solution (Exercise 22). By adding the equation xyz++= 1 (the sum of the entries in any state matrix of the system must equal 1) to the system, we do end up with a unique solution: –.6.8x ++y .4z = 0  –.6.40.8 1003 8 .3x –.3.9y +0z =  .3–.30 .9 rref 0101 4   .5x +0.3y – .7z =  .5 .3–0 .7 0013 8  xyz++= 1 1111 000 0 We see that the matrix T has a unique fixed state; namely: .3750 .3750 .3750 38 .3750 8 T = .2500 .2500 .2500 SF ==14 .2500 , which, to four decimal places, coincides with .3750 .3750 .3750 38 .3750 the columns of the matrix T 8 in Example 6.14 (margin). The above rather surprising result, as you will soon see in Theorem 6.26, actually holds for the following important class of Markov chains: DEFINITION 6.12 A Markov chain with transition matrix T is For example: REGULAR MARKOV said to be regular if T k consists solely of 0.3.2 CHAIN positive entries for some integer k. The T = .8 .4 .5 transition matrix of a regular Markov chain .2 .3 .3 is said to be a regular transition matrix. is regular, since: Note that it is possible to eventually go from any state to any .28 .18 .21 2 other state in a regular Markov chain (see Theorem 6.24). T = .42 .55 .51 .30 .27 .28 264 Chapter 6 Determinants and Eigenvectors

In other words, AT is that matrix obtained by inter- DEFINITION 6.13 The transpose of a matrix changing the rows and col- TRANSPOSE Aa= ij  Mmn is the matrix umns of A. For example: T A = aij  Mnm , where aij = aji . 12 103 T If A =  A = 04 245 35 The following results will be called upon within the proof of Theo- rem 6.25 below:

T T T A-1: If AM mn and BM nr , then AB = B A . [Exercise 19(f), page 161.]

A-2: If  is an eigenvalue of AM n  n then  is also an eigenvalue of AT . (Exercise 52, page 231.)

THEOREM 6.25  = 1 is an eigenvalue of every transitional

matrix TM nn .

PROOF: Let w  n be the n-tuple with every entry equal to 1. Being

a transition matrix, the columns of Tt= ij sum to 1. Hence:

t11 t12  t1n

t21 t22  t2n  wT ===111    111 w

tn1 tn2  tnn

t11 +++t21  tn1 = 1 Taking the transpose of wT = w , we have T TwT = wT (A-1), and T T Note that wT is an eigen- this tells us that w is an eigenvector of T corresponding to the vector of the transpose of eigenvalue  = 1 . Applying A-2 we conclude that  = 1 is also an T, and not necessarily of T. eigenvalue of T.

Every regular transition matrix TM has In a sense, independently of THEOREM 6.26 nn FUNDAMENTAL THEOREM r its initial state: 1 OF REGULAR MARKOV The fixed state of a reg- r 2 ular transition matrix is CHAINS a unique fixed state vector SF = , and: also the final state of the r matrix n

r1 r1  r1  lim Ts = r2 r2 r2 s  

rn rn  rn (each column of the matrix Ts approaches SF as s increases). 6.5 Markov Chains 265

PROOF: Assume that Tt= ij consists of positive entries. (You are invited, in Exercise 27, to establish the result under the assumption that T k consists solely of positive entries for some integer k  1 .)

s s Consider the matrix T = cij , which must also consist of positive s s s s s That “(s)” in Mi is not entries. Let M = c and m = c denote the largest and i ijM i ijm an exponent; it is there to th s indicate that we are con- smallest entry in the i row of T . We will show that s s s th sidering the matrix T limMi – mi = 0 . This will tell us that all entries in the i row s   of lim Ts are equal, which is the same as saying that the columns of s   lim Ts are all equal. s   s + 1 s + 1 s s From T ===cij T Tcij tij we have: n n s + 1 s c ==c t cs t + cs t ij  i j ijm jmj  i j

 = 1   jm n = ms t + cs t i jmj  i j

  jm n  ms t + Ms t i jmj i  j   j The entries in the jth column of the m s s transition matrix T sum to 1: = m t + M 1 – t i jmj i jmj

s + 1 th s + 1 We have shown that for every entry cij in the i row of T : cs + 1  ms t + Ms 1 – t ij i jmj i jmj In particular, for the largest entry in that row we have: s + 1 M  ms t + Ms 1 – t i i jmj i jmj A similar argument (Exercise 26) can be used to show that for the s + 1 th s + 1 smallest entry mi in the i row of T we have: ms + 1  Ms t + ms 1 – t i i jMj i jMj Consequently: s + 1 s + 1 Mi – mi s  ms t + M 1 – t – Ms t + ms 1 – t i jmj i jmj i jMj i jMj = Ms – ms 1 – t – t i i jmj jMj 266 Chapter 6 Determinants and Eigenvectors

Let t be the smallest entry in T. Since T consists of positive entries, and since the entries in every column of T sums to 1, we have 1 0  t  --- , and, in particular that 1 – t – t  12– t . Hence: n jmj jMj s + 1 s + 1 s s Mi – mi  Mi – mi 12– t Leading us to:

s s s – 1 s – 1 Mi –12mi  – t Mi – mi

2 s – 2 s – 2  12– t Mi – mi 

s – 1  12– t Mi – mi Since 012 – t  1 , 12– t s – 1  0 as s   , and this tells us that the elements in the ith row of the matrix Ts must get arbitrarily close to each other as s tends to infinity. In turn, the columns of Ts

r1 must all tend to a common vector r2 . We complete the proof by

rn

r1 r showing that SF = 2 is the unique fixed state of T:

rn

x1 Employing Theorem 6.25, we start with an eigenvector X = x2 of T,

xn of eigenvalue 1. Since TXX= , TsXX= for all k. Hence:

x x r1 r1  r1 1 1 x x lim TsX ==r2 r2  r2 2 2 s  

rn rn  rn xn xn From Theorem 5.4, page 156:

r r  r x r1 r1 r1 r 1 1 1 1 n 1 r r  r x r r r r 2 2 2 2 ==x 2 +++x 2  x 2 x 2 1 2 1  i i = 1 r r  r x n n n n rn rn rn rn 6.5 Markov Chains 267

Since TX= X :

x1 r n 1 x X ==2 x r2  i (**) i = 1 xn rn n

Since X  0 (it is an eigenvector), cx=0 i  , and dividing both sides of (**) by c brings us to: i = 1 x 1 r1 1 x S ==--- 2 r2 F c

xn rn We then have: x x x r1 1 1 1 r1 r 1 x 1 x 1 x r T 2 ====T--- 2 ---T 2 --- 2 2 c c c  rn xn xn xn rn The above argument also establishes the uniqueness of the fixed state vector, for if X is to be a (fixed) state vector, then c must equal one. EXAMPLE 6.18 An automobile insurance company classifies its customers as Preferred, Satisfactory, or Risk. Each year, 10% of those in the Preferred category are downgraded to Satisfactory, while 12% of those in the Satisfactory cate- gory move to Preferred. Twenty percent of Satisfactory drop to Risk, while 15% of Risk goes to Satisfactory. No customer is moved more than one slot in either direction in a sin- gle year. Find the fixed state of the system. SOLUTION:

.85 R .15 Pr S R Pr .9 .2 .9 .12 0 .1 Pr S .68 T = .1 .68 .15 S .12 0 .2 .85 R 268 Chapter 6 Determinants and Eigenvectors

The easiest way to go, is to take a “large” power of the transition How large is large enough? If matrix, and let any of its rows represent an approximation for the the rows look different, then fixed state matrix of the regular Markov process: take a higher power.

This establishes the fact that we are in a regular Markov situation (how)?

We conclude that roughly 34% of the company’s clients will (eventu- ally) fall in its Preferred category; 28% in its Satisfactory category; and 38% in its Risk category. But that is but an approximation, for:

.9 .12 0 .34 .3396 .1 .68 .15 .28 = .2814 0 .2 .85 .38 .3790 You can, however, find the exact steady state by the method of Exam- ple 6.17: .9x ++.12y 0z = x  .9 .12 0 x x  .1 .68 .15 y = y  .1x ++.68y .15z = y   0 .2 .85 z z 0x ++.2y .85z = z 

1 00-----18- 53 –.1x ++.12y 0z = 0  15 0 1 0 ------.1x –.32 +0 .15z =  53  rref 0x +0.2y – .15 =  001 -----20-  53 xyz++= 1 0000 see solution of Example 6.15.

18 15 20 We found ------------to be the (exact) steady state of the given Mar- 53 53 53 kov chain; telling us that the longer the process, the closer it will be that -----18-% of the customers, for example, will be in the preferred cate- 53 gory. 6.5 Markov Chains 269

CHECK YOUR UNDERSTANDING 6.21 The transition matrix T below represents the probabilities that an individual that voted the Democratic, Republican, or Green party ticket in the last election will vote D, R, or G in the next election. D R G D .73 .32 .09 T: .21 .61 .04 R Answer: Approximately 41%, .06 .07 .87 G 26%, 33% of the population, will vote democratic, republi- Determine the eventual percentage of the population in each of can, green, respectively. the three category. 270 Chapter 6 Determinants and Eigenvectors

EXERCISES

Exercises 1-6. Indicate whether or not the given matrix represents a transition matrix for a Markov Process. If not, state why not. If so, indicate whether or not the given transition matrix is regular.

.2 .1 01 1. 2. 3. 0 .1 .8 .9 10 –1 .9

.2 .4 .1 1 .4 .1 0 .3 .4 4..7 0 .3 5.0 0 .3 6. .5 .3 .6 .1 .6 .6 0 .6 .6 .5 .1 0 Exercises 7-8. Determine the transition matrix associated with the given transition diagram. 8. 7. .3 .4 .6 .1 .1 .2 A .5 B A .9 B .5 .4 .7 C .3

Exercises 9-11. Determine a transition diagram associated with the given transition matrix.

A B A B C A B C D A A A 9. .3 .4 1 0 1 0 1 .2 .6 10. .7 .6 B 000B 11. .5 0 .5 .4 B 010C 0 0 0 0 C .5 0 .3 0 D

12. Determine the probability of ending up at states A and B after two steps of the Markov chain associated with the transition matrix in Exercise 9, given that you are initially in state: (a) A (b) B 13. Determine the probability of ending up at states A, B and C after two steps of the Markov chain associated with the transition matrix in Exercise 10, given that you are initially in state: (a) A (b) B (c) C 14. Determine the probability of ending up at states A, B, C and D after two steps of the Markov chain associated with the transition matrix in Exercise 11, given that you are initially in state: (a) A (b) B (c) C (d) D 6.5 Markov Chains 271

Exercises 15-20. (a) Proceed as in Example 6.15 to find the stationary state matrix of the given regular transition matrix. (b) Use Theorem 6.26 and a graphing utility to check your answer in (a).

15..7 .2 16..1 .6 17. .5 .3 .3 .8 .9 .4 .5 .7

.8 .5 0 .5 .3 0 .6 .2 .1 18..2 .1 .6 19..1 .7 .6 20. .3 .5 .5 .0 .4 .4 .4 0 .4 .1 .3 .4

001 21. Show that the matrix A = 100 is not a regular matrix, by: 010 (a) Demonstrating that for each k, Ak will contain a row that does not consist solely of posi- tive entries. (b) Showing that A does not have a fixed state vector.

x x 22. Show that for any transition matrix T, the system of equations stemming from T y = y has z z infinitely many solutions. Suggestion: Use the fact that the sum of the elements in each column of T sum to 1.

23. Let TM nn be a regular transition matrix. Prove that x – 1 is a factor of the character- istic polynomial of T.

24. Show that if the entries in each column of AM nn sum to k, then k is an eigenvalue of A. 25. Referring to the proof of Theorem 6.26, show that: mk + 1  Mk b + mk 1 – b i i jMj i jMj 26. Establish Theorem 6.26 for an arbitrary transitional matrix T. Suggestion: Let r be such that AT= r consists of positive entries, and consider the matrix TA.

27. Prove that if  is any eigenvalue of a regular transition matrix, then   1 .

28. Show that if  is an eigenvalue of a regular transition matrix, then   –1 .

29. (Rapid Transit) A study has shown that in a certain city, if a daily (including Saturday and Sunday) commuter uses rapid transit on a given day, then he will do so again on his next com- mute with probability 0.85, and that a commuter who does not use rapid transit will do so with probability 0.3. Assume that on Monday 57% of the commuters use rapid transit. Deter- mine, to two decimal places, the probability that a commuter will use rapid transit on: (a) Tuesday (b) Wednesday (c) Sunday 272 Chapter 6 Determinants and Eigenvectors

30. (Dental Plans) A company offers its employees 3 different dental plans: A, B, and C. Last year, 550 employees were in plan A, 340 in plan B, and 260 were in plan C. This year, there are 500 employees in plan A, 360 in plan B, and 290 in plan C. Assuming that the number of employees in the company remains at 1150, and that the current trend continues, determine the number of employees in each of the three plans: (a) A year from now. (b) Two years from now. (c) In 4 years (Suggestion: use the square of the 2-year matrix). (d) In 8 years (Suggestion: use the square of the 4-year matrix). (e) In 12 years (Suggestion: use the product of 4-year and the 8-year matrix). 31. (Campus Life) The following transition matrix gives the probabilities that a student living in the Dorms, at Home, or Off-campus (but not at home), will be living in the Dorms, at Home, or Off-campus (but not at home) next year (assume that all freshmen will graduate from the college in four years). D H O .62 .23 .25 D .11 .64 .09 H .27 .13 .66 O Currently, 55%, 24%, and 21% of the freshman class are living in the Dorms, at Home, and Off-campus (but not at home), respectively. Determine (to two decimal places) the probability that a current freshman will, three years from now, be living in the: (a) Dorms (b) Home (c) Off-campus. 32. (Higher Learning) The transition matrix below represents the probabilities that a female child will receive a Doctorate, a Masters, or a Bachelors (terminal degree), or No degree; given that her mother received a D, M, B (terminal degree), or No degree. mother D M B N

.31 .24 .11 .06 D daughter .25 .26 .09 .05 M .37 .42 .52 .49 B .07 .08 .28 .40 N

Given the initial state matrix (in column form), determine the S0 = .05 .09 .39 .47 probability that: (a) A granddaughter will receive a Bachelors degree. (b) A great granddaughter will earn a Doctorate. (c) A fifth generation daughter will receive no degree. 6.5 Markov Chains 273

33. (HMO Plans) A company offers its employees 5 different HMO health plans: A, B, C, D, and E. An employee can switch plans in January of each year, resulting in the following tran- sition matrix: this year A B C D E .54 .13 .08 .10 .06 A next year .11 .61 .17 .12 .18 B .17 .10 .56 .17 .15 C .06 .05 .08 .49 .10 D .12 .11 .11 .12 .51 E

Given the initial state matrix S0 = .11 .20 .31 .14 .24 (in column form), deter- mine, to three decimal places, the probability that: (a) An employee will chose plan B in the next enrollment period. (b) An employee will chose plan B two enrollment periods from now. (c) An employee will chose plan B three enrollment periods from now. (d) Determine to 5 decimal places, the fixed state of the system.

(e) Repeat (a) through (d) with initial state matrix S0 = .24 .31 .0 .26 .19 34. (Mouse in Maze) On Monday, a mouse is placed in a maze consisting of paths A and B. At the end of path A is a cheese treat, and at the end of path B there is bread. Experience has shown that if the mouse takes path A, then there is a 0.9 probability that it will take path A again, on the following day. If it takes path B, then there is a 0.6 probability that it will take that path again, the next day. The mouse takes path B on Monday. Determine the probability that the mouse will take path A on:

(a) Tuesday (b) Wednesday (c) Sunday (d) Answer parts (a), (b), and (c), under the assumption that the mouse takes path A on Monday. (e) Show that the transition matrix is regular, and then proceed as in Example 6.14 to determine the exact stationary state of that matrix. (f) Indicate the long-term state of the system (the probability that the mouse will take path A and the probability that the mouse will take path B, at the nth step of the process, for n “large”). 35. (Cities, Suburbs, and Country) Within the period of a year, 2% of a population currently residing in cities will move to the suburbs, while 2% of them will move to the country. 4% of those living in the suburbs will move to the cities, while 3% of them will move to the country. One percent of the country folks will move to the cities, while 2% of them will go to the sub- urbs. Currently, 65% of the population are in cities, and 20% are in the suburbs. Determine, to two decimal places, the percentage of city dwellers:

(a) Next year. (b) Two years from now. (c) Four years from now. 274 Chapter 6 Determinants and Eigenvectors

(d) Answer parts (a), (b), and (c), under the assumption that 50% of the population are in cities, and 35% are in the suburbs. (e) Determine to 5 decimal places, the fixed state of the system. 36. (Crop Rotation) A farmer rotates a field between crops of beans, potatoes and carrots. If she grows beans this year, then next year she will grow potatoes or carrots, each with 0.5 proba- bility. If she grows carrots, then she will grow beans with probability 0.2, potatoes with prob- ability 0.5 (and carrots with probability 0.3). If she grows potatoes, then she will grow beans with probability 0.5, and potatoes with probability 0.25. If she grows beans this year, what is the probability that she will grow beans again:

(a) Next year? (b) Two years from now? (c) Three years from now? (d) Answer parts (a), (b), and (c) under the assumption that she grows potatoes this year. (e) Determine to 5 decimal places, the fixed state of the system.

37. (Wolf Pack) A wolf pack hunts on one of four regions: A, B, C, and D:

A

D B

C

If the pack hunts in any given region one day, then it is as likely to hunt there again the next day as it is for it to hunt in either of its neighboring regions. On Monday, it hunted in region A. (a) Determine, to two decimal places, the probability that the pack will hunt in Region B on Tuesday. (b) Determine, to two decimal places, the probability that the pack will hunt in Region B on Sunday. (c) Determine the fixed state of the system. Chapter Summary 275

CHAPTER SUMMARY

DETERMINANTS The determinant of an nn matrix A, denoted det (A), is defined inductively as follows:

For a 11 matrix Aa= 11 , detA = a11 .

For a nn matrix A, with n  1 , let Aij denote the n – 1  n – 1 matrix obtained by deleting the ith row and jth column of the matrix A; Then: n 1 + j detA =  –1 a1j detA1j j = 1

Laplace’s Theorem For AM nn and any 1 in : n n + _ + _ + _ + ______+ _+ _+ +_ detA = –1 ij+ a detA and det A = –1 ij+ a detA + + + +  ij ij  ij ij _ + _ + _ + _ + _ _ _ _ + + + + j = 1 i = 1 _ + _ + _ + _ + th + _ + _ + _ + _ Expanding along the i row Expanding along the jth row _ + _ + _ + _ + Determinants of diago- The determinant of a diagonal matrix or of an upper triangular matrix nal and upper triangu- is the product of the entries in its diagonal. lar matrices Determinants and (a) If two rows of AM nn are interchanged, then the determinant row operations of the resulting matrix is –detA . (b) If one row of A is multiplied by a constant c, then the determi- nant of the resulting matrix is cdetA . (c) If a multiple of one row of A is added to another row of A, then the determinant of the resulting matrix is detA Invertibility A matrix AM nn is invertible if and only if detA  0 . Product Theorem For AB  Mnn : detAB = detA detB 276 Chapter 6 Determinants and Eigenvectors

EIGENVALUE AND An eigenvalue of a linear operator T: VV is a scalar   R for EIGENVECTOR which there exists a nonzero vector v  V such that: Tv = v Any such v is then said to be an eigenvector corresponding to the eigenvalue  .

An eigenvalue of a matrix AM nn is a scalar   R for which there exists a nonzero vector XR n such that: AX = X Any such X is then said to be an eigenvector corresponding to the eigenvalue  .

IGENSPACE E The eigenspace of an eigenvalue  of a matrix AM nn is denoted by E and is given by: E = nullA – I The eigenspace of an eigenvalue  of a linear operator T: VV is denoted by E and is given by: E = kerT – I

CHARACTERISTIC For AM nn , the n-degree polynomial detA – I is said to be the POLYNOMIAL AND characteristic polynomial of A, and detA – I = 0 is said to be CHARACTERISTIC the characteristic equation of A. EQUATION For T: VV a linear operator on a vector space V of dimension n,

the n-degree polynomial detT  – I , where  is a basis for V, is said to be the characteristic polynomial of T, and

detT  – I = 0 is said to be the characteristic equation of T. Finding Eigenvalues The eigenvalues of AM nn are the solutions of the characteristic equation detA – I = 0 . The eigenvalues of a linear operator T: VV on a vector space of

dimension n are the eigenvalues of the matrix T   Mnn , where  is any basis for V.

IAGONAL ATRIX D M A diagonal matrix is a square matrix Aa= ij with aij = 0 for ij . Chapter Summary 277

IAGONALIZABLE D A matrix AM nn is diagonalizable if A is similar to a diagonal MATRICES AND matrix. LINEAR OPERATORS A linear operator T: VV on a finite dimensional vector space V is

said to be diagonalizable if there exists a basis  for which T  is a diagonal matrix. Diagonalization Let T: VV be a linear operator on a finite dimensional vector Theorem space. The following are equivalent: (i) T is diagonalizable.

(ii)T  is a diagonalizable matrix, for any basis  of V. (iii)There exists a basis for V consisting of eigenvectors of T. Eigenvectors corre- If 12 m are distinct eigenvalues of a linear operator sponding to different T: VV , and if v is an eigenvector corresponding to  , for eigenvalues are lin- i i early independent. 1 im, then v1v2  vm is a linearly independent set. The union of linearly Let 12 m be distinct eigenvalues of a linear operator independent subsets of T: VV , and let S = v v  v be a linearly independent different eigenspaces is i i1 i1 iri again a linearly inde- subset of Ei . Then: pendent set. SS= 1 S2  Sm is a linearly independent set.

LGEBRAIC AND A An eigenvalue 0 of a matrix AM nn (or of a linear operator T on GEOMETRIC a vector space of dimension n) has algebraic multiplicity k if MULTIPLICITY OF k – 0 is a factor of A’s (or T ’s) characteristic polynomial, and EIGENVALUES k + 1 – 0 is not. We also define the geometric multiplicity of 0

to be the dimension of E0 (the eigenspace corresponding to 0 ) The geometric multi- If 0 is an eigenvalue of AM nn with algebraic multiplicity ma plicity cannot exceed and geometrical multiplicity m , then m  m . the algebraic multi- g g a plicity. Another Diagonaliza- Assume that the characteristic polynomial of a linear operator tion Theorem. T: VV (or of a matrix), can be factored into a product of linear factors. Then T is diagonalizable if and only if the algebraic multiplic- ity of each eigenvalue of T is equal to its geometric multiplicity. 278 Chapter 6 Determinants and Eigenvectors

Diagonalizing a Matrix Let AM nn be diagonalizable. Let the columns of PM nn be any basis for Rn consisting of eigenvectors of A. Then DP= –1AP

is a diagonal matrix. Moreover, the diagonal entry dii  D is the th eigenvalue i corresponding to the i column of P.

Power Theorem for –1 Let AM mm be diagonalizable with P AP= D . Then for any n: diagonalizable matri- ces An = PDnP–1 .

FIBONACCI The Fibonacci sequence is that sequence whose first two terms are 1, SEQUENCE and with each term after the second being obtained by summing its two immediate predecessors. 1 15+ k 15– k The kth Fibonacci number is given by ------------– ------. 5 2 2

SYSTEMS OF DIFFER- The solution set of f x = af x , consists of those functions of the ENTIAL EQUATIONS form fx= ceax for some constant c.

Let AM nn be diagonalizable, and let v1v2  vn be a basis for n consisting entirely of eigenvectors of A with corresponding eigenvalues i . Then, the general solution of: Fx = AF x is of the form: 1x 2x  nx c1e v1 +++c2e v2 cne vn

for c1c2 cn  .

MARKOV CHAINS Markov chain: When the probabilities of moving from one state of a system to another remain constant.

TRANSITION MATRIX A transition matrix TM nn is a matrix that satisfies the follow- ing two properties: (1) T contains no negative entry. (2) The sum of the entries in each column of T equals 1.

n FIXED STATE SF   is a fixed state for a transition matrix TM nn if TSF = SF . If T is the transition matrix of a Markov process with initial-state th matrix S0 , then the n state matrix in the chain is given by: n Sn = T S0

UNDAMENTAL HEO F T - Every regular transition matrix TM nn has a unique fixed state vec- REM OF REGULAR tor and each column of the matrix Ts approaches S as s increases. MARKOV CHAINS F 7.1 Dot Product 279

7 CHAPTER 7 INNER PRODUCT SPACES Basically, an is a vector space augmented with an additional structure, one that will enable us to generalize the familiar concepts of and in the plane to general vector spaces.

§1. DOT PRODUCT

We begin by introducing a function which assigns a real number to each pair of vectors in n :

DEFINITION 7.1 The dot product of uu= 1u2  un DOT PRODUCT and vv= 1v2  vn , denoted by uv , is the real number: uv = u1v1 +++u2v2  unvn For example: 24– 3 1  507– 1 ==25 + 40 + – 3 71+  – 1 –12

The following four properties will lead us to the definition of an inner product space in the next section, much in the same way that the eight properties of Theorem 2.1, page 36, lead us to the definition of an abstract vector pace on page 40.

THEOREM 7.1 Let uvw  n , and r   . Then: positive-definite property: (i)vv  0 , and vv = 0 only if v = 0 commutative property: (ii) uv = vu homogeneous property: (iii) ruv ==ruv u  rv distributive property: (iv) uv+  w = uw + vw

PROOF: We establish (iii) and invite you to verify the remaining three properties in the exercises:

ruv = ru1u2  un  v1v2  vn scalar multiplication: = ru1ru2  run  v1v2  vn definition 7.1: = ru1 v1 +++ru2 v2  run vn associative property: = ru1v1 +++ru2v2 runvn distributive property: = ru1v1 +++u2v2  unvn definition 7.1: = ruv 280 Chapter 7 Inner Product Spaces

CHECK YOUR UNDERSTANDING 7.1

Let uv  n , and r   . Prove that: Answer: See page B-30. ruv = u  rv

DEFINITION 7.2 The of a vector vv= 1v2  vn , n NORM IN  denoted by v , is given by: vvv= 

2 2 2 For vv= 1 v2  , vvv== v1 + v2 represents the length (magnitude) of v [Figure 7.1(a)], and the same can be said about 3 v for vv= 1v2 v3  [Figure 7.1(b)].

2 2 2 2 2 v = v1 + v2 v = v1 ++v2 v3 vv=  v 1 2 vv= 1v2 v3

v2 v3

v1 v1 v2

(a) (b) Figure 7.1 In general, for v  n : v is defined to be the length of v. Moreover:

uv– is defined to be the distance between uv n .

2 In particular, for uu= 1 u2 and vv= 1 v2 in  :

vv= 1 v2

2 2 uv– = u1 – v1 + u2 – v2 v2 – u2 v1 – u1 uu= 1 u2

7.1 Dot Product 281

CHECK YOUR UNDERSTANDING 7.2

Prove that for uv  n and c   : (a) cv = c v (b) uv– 2 = u 2 – 2uv + v 2 2 2 2 Answer: See page B-30. [Reminiscent of: ab– = a – 2ab + b ]

ANGLE BETWEEN VECTORS Applying the law of cosines [Figure 7.2(a)] to the vectors uv  2 in Figure 7.2(b), we see that: uv– 2 = u 2 + v 2 – 2 uvcos

c uv– a u  b  v c2 = a2 + b2 – 2abcos Law of Cosines uv– 2 = u 2 + v 2 – 2 uvcos (a) (b) Figure 7.2 From CYU 7.2, we also have: uv– 2 = u 2 – 2uv + v 2 Thus: u 2 + v 2 – 2 uvcos = u 2 – 2uv + v 2 –22 uvcos = – uv For any –1 x 1 , uv cos = ------ –1 cos x is defined to be uv that 0  –1 uv see margin:  = cos ------whose cosine is x. uv Leading us to: In Exercise 44 you are asked to verify that DEFINITION 7.3 The angle  between two nonzero vectors n ------uv  1 ANGLE BETWEEN uv   is given by: uv VECTORS –1 uv Assuring us that:  = cos ------uv –1 uv cos ------exists. uv 282 Chapter 7 Inner Product Spaces

EXAMPLE 7.1 Determine the angle between the vectors u = 120– 2 and v = –1312 .

SOLUTION: –1 uv –1 120– 2  –1312  ==cos ------cos ------uv 144++ 1914+++

–1 1 = cos ------ 85 315

CHECK YOUR UNDERSTANDING 7.3 Answer: –1 5 Determine the angle between the vectors u = 120 and cos ------ 83 55 v = –131.

ORTHOGONAL VECTORS IN Rn

The angle  between the vectors uv  2 depicted in the adjacent figure has a measure of u We remind you that, for any  –1 90 (--- radians ), and we say that those vectors are –1 x 1 , cos x is that angle 2  = 90 0  such that cos = x . perpendicular or orthogonal. Appealing to Defini- tion 7.3 we see that: v –1 uv So, if cos ------= 90 , uv –1 uv cos ------= 90 or uv = 0 uv then------ = 0 , or:uv = 0 . uv uv (see margin) Bringing us to:

DEFINITION 7.4 Two vectors u and v in n are orthogonal ORTHOGONAL VECTORS if uv = 0 .

Note: The zero vector in n is orthogonal to every vector in n .

CHECK YOUR UNDERSTANDING 7.4

Let v  n . Show that the set v of vectors perpendicular to v: v = u  n uv = 0 Answer: See page B-31. is a subspace of n .

It is often useful to decompose a vector v  n into a sum of two vec- tors: one parallel to a given nonzero vector u, and the other perpendicular to u. The parallel-vector must be of the form cu for some scalar c (see Figure 7.3). 7.1 Dot Product 283

v – cu

v

cu = projuv v – cu u

Figure 7.3 ORTHOGONAL The vector cu in Figure 7.3 is said to be the orthogonal projection of v onto u and is denoted by projuv . To determine the value of c, we note that for v – cu to be orthogonal to u, we must have:

Theorem 7.1(iv): v – cu  u = 0 Theorem 7.1(iii): vu – cu  u = 0 vu – cuu = 0 vu vu c ==------ ------ uu u 2 Summarizing, we have: vv= – projuv + projuv THEOREM 7.2 Let v  n and let u be any nonzero vector in n . VECTOR Then: DECOMPOSITION vv= – proj v + proj v v – projuv u u where: u uv projuv = ------u and v – projuv  projuv = 0 uv uu proj v = ------u u uu The vector projuv is said to be the vector component of v along u, and the vector v – projuv is said to be the vector component of v orthogonal to u.

EXAMPLE 7.2 Express the vector 21– 3 as a sum of a vector parallel to 140 and a vector orthog- onal to 140 .

SOLUTION: For v = 21– 3 and u = 140 we have: uv 140  21– 3 6 proj v ==------u ------140 =------140 u uu 140  140 17 6 24 28 7 and v – proj v ==21– 3 – ------------0 ------–------–3 u 17 17 17 17 28 7 6 24 Check: ------–------–3 + ------------0 ==21– 3 v 17 17 17 17 28 7 6 24 28 6 7 24 and ------–------–3  ------------0 ==------------+ –------------0 17 17 17 17 17 17 17 17 284 Chapter 7 Inner Product Spaces

Answer: CHECK YOUR UNDERSTANDING 7.5 6 9 24 3–------– ------21 21 21 Express the vector 301– 1 as the sum of a vector parallel to 6 12 3 0241 and a vector orthogonal to 0241 . + 0------21 21 21

EXAMPLE 7.3 Find the distance from the point P = 313 to the line L in 3 which passes through the points 102 and 316 .

SOLUTION: We first find a direction vector for the given line (see Theorem 2.17, page 70): P u ==316 – 102 214 . . want this distance Choosing the point A = 102 on L v – proj v we determine the vector v from A to P: u L

v ==313 – 102 211 v Applying Theorem 7.2, we have: uv proj v = ------u u uu u 214  211 = ------214 proj v 214  214 . u A 9 6 3 12 ==------214 ---------21 7 7 7 Thus: 6 3 12 v – proj v = 211 – ---------u 7 7 7 8 4 5 1 1 105 ==------ –------84– 5 =--- 64++ 16 25 =------7 7 7 7 7 7 CYU 7.2(a)

CHECK YOUR UNDERSTANDING 7.6 (a) Find the distance from the point P = 25 to the line L in 2 passing through the points 12 – and 24 .

1 (b) Find the distance from the point P = 1013 to the line L in Answer: (a) ------(b) 4 4 37  passing through the points 1201 and 1221 . 7.1 Dot Product 285

PLANES REVISITED We now offer an alternative representation for a plane in 3 than that given in Theorem 2.19, page 71. Just as a line in 2 is determined by a point on the line and its slope, so then is a plane in 3 determined by a point on the plane and a nonzero vector orthogonal to the plane (a vector to the plane). To be more specific, suppose we want the equation of the plane with normal P n . vector n = abc that contains the point A0 = x0y0 z0 . For any . point Pxyz=  on the plane we have: A0 n  A0P = 0 or: normal form abc  xx– 0 yy– 0 zz– 0 = 0 scalar form or: ax– x0 ++by– y0 cz– z0 = 0

or: ax++ by cz =d where d = ax0 ++by0 cz0 general form

Note that a normal to the plane can easily be spotted from any of the above forms. For example, n = 25– 4 is a normal to the plane: 25– 4  x – 1 y – 3 z + 2 = 0 2x – 1 +05y – 3 – 4z + 2 = 2x +255y – 4z =

EXAMPLE 7.4 Find a normal, scalar, and general form equa- tion of the plane passing through the point 13– 2 with normal vector n = 41– 5 . SOLUTION: normal: 41– 5  x – 1y – 3 z + 2 = 0 scalar: 4x – 1 –51y – 3 +0z + 2 = general: 4xy–5+9z = –

CHECK YOUR UNDERSTANDING 7.7

Find an equation of the plane passing through the point 13– 2 with normal parallel to the line passing through the points Answer: See page B-31. 110  021 .

EXAMPLE 7.5 Express the plane 3xy+6– 2z = in the vec- tor form of Theorem 2.19, page 71. 286 Chapter 7 Inner Product Spaces

SOLUTION: In order to write the plane in vector form we need two direction vectors and a translation vector. We chose 200 as our translation vector (corresponding to the vector wx= 0y0 z0 in Figure 2.9, page 71). Any two linearly independent vectors orthogonal to n = 31– 2 31– 2  021 = 0 can serve as direction vectors, say 021 and 203 [corre- and 31– 2  203 = 0 sponding to u = AB and v = AC in Figure 2.9]. This leads us to the following vector form equation of the plane: 200 ++r021 s203 rs  R

CHECK YOUR UNDERSTANDING 7.8 With reference to Example 7.5, show, directly that: Answer: See page B-31. xyz 3xy+6– 2z = = 200 ++r021 s203 rs  R

EXAMPLE 7.6 Find the distance from the point P = 432 to the plane 2x – 3y +5z = .

SOLUTION: We begin by choosing n Any point xyz satis- the point A = 005 on the plane fying the equation (note that 20 –530 +5= ). We P proj v 2x – 3y +5z = would position the normal vector n v . do just as well. n = 231 –  so that its initial A point is at A, and then determine the we want this distance . vector v from A to P: B v ==432 – 005 43– 3 Applying Theorem 7.2, we have: vn proj v = ------n n nn 43– 3  231 –  = ------231 –  231 –   231 –  –4 4 6 2 ==------231 –  –------ –--- 14 7 7 7 Hence: 1 1 56 214 proj v ==--- –462– --- 16++ 36 4 ==------n 7 7 7 7

CYU 7.2(a)

Answer: See page B-32. CHECK YOUR UNDERSTANDING 7.9 Note that if you apply this formula in Example 7.6 Prove that the distance D from a point x0y0 z0 to the plane you obtain: ax++ by cz = d is given by the formula: 24– 33+ 12– 5 D = ------   22 ++32 12 ax ++by cz – d D = ------0 ------0 ------0 ==------4 ------214 a2 ++b2 c2 14 7 7.1 Dot Product 287

CROSS PRODUCT Here is a handy result:

THEOREM 7.3 If v1 = a1a2 a3  v2 = b1b2 b3 are lin- early independent vectors in 3 , then: va= 2b3 – a3b2– a1b3 + a3b1 a1b2 – a2b1 is perpendicular to both v1 and v2

PROOF: Rolling up our sleeves, we simply show that vv 1 = 0 , and leave it for you to verify that vv 2 is also zero: vv 1 = a2b3 – a3b2– a1b3 + a3b1 a1b2 – a2b1  a1a2 a3

= a2b3 – a3b2 a1 ++– a1b3 + a3b1 a2 a1b2 – a2b1 a3

==a1a2b3 – a1a3b2 –0a1a2b3 ++a2a3b1 a1a3b2 – a2a3b1

Handy or not, how is one to remember that complicated three-tuple v = a2b3 – a3b2– a1b3 + a3b1 a1b2 – a2b1 of Theorem 7.3? By formally evaluating the following determinant about its first row:

e1 e2 e3 a2 a3 a1 a3 a1 a2 det a1 a2 a3 = det e1 – det e2 + det e3 b2 b3 b1 b3 b1 b2 b1 b2 b3

= a2b3 – a3b2 e1 – a1b3 – a3b1 e2 + a1b2 – a2b1 e3 and then replacing e1e2 and e3 with the standard basis vectors 100 010 and 001 , respectively, to arrive at the three- tuple of Theorem 7.3:

va= 2b3 – a3b2– a1b3 + a3b1 a1b2 – a2b1

The above vector is called the of v1 = a1a2 a3 and v2 = b1b2 b3 , and is denote by: v1  v2 . Moreover, to conform with standard notation, we shall replace the standard basis vectors e1e2 and e3 with the symbols i, j, and k, respectively; bringing us to:

DEFINITION 7.5 The cross product of v1 = a1a2 a3 CROSS PRODUCT and v2 = b1b2 b3 is denoted by v1  v2 , and is expressed in the form: ijk

v1  v2 = det a1 a2 a3

b1 b2 b3 288 Chapter 7 Inner Product Spaces

For example: ij k 234  31– 2 = det 23 4 31– 2

= det 34i – det 24j + det 23k 12– 32– 31 ==–46 – i –29–124 – j + – k –10 16– 7

EXAMPLE 7.7 Find the general form equation of the plane that contains the points A = 12– 1 , B = 231 , C = 312 –  .

SOLUTION: Noting that the vectors AB ==231 – 12– 1 112 and AC ==312 –  – 12– 1 23– 3 are parallel to the plane, we employ Theorem 7.3 to find a normal to the plane:

ijk n ===det 112 9ij+ – 5k 91– 5 23–3

Choosing the point A = 12– 1 on the plane, we proceed as in Example 7.4 to arrive at the general form equation of the plane: 91– 5  x – 1y – 2 z + 1 = 0 9x – 1 +0y – 2 – 5z + 1 = 9xy+0–165z – =

CHECK YOUR UNDERSTANDING 7.10

(a) Find the general form equation of the plane that contains the points A = 32–2, B = 25– 3, C = 412  – 3 . (b) Verify that your answer in (a) coincides with that of Example Answer: See page B-33. 2.15, page 72. 7.1 Dot Product 289

EXERCISES

Exercises 1-2. Evaluate uv for the given n-tuples. 1.u ==53  v 61 2. u ==0357  v 2104 Exercises 3-5. Determine the norm v for the given vector.

3.v = 32 4.v = 10– 5 5. v = 31– 21–

6. Find all values of c such that c231 = 9 . 7. Find all values of a such that the vector a 3 is orthogonal to the vector 2a –5 . 8. Find all values of a such that the vector 3a 2a is orthogonal to the vector a2 a . 9. Find all values of a and b such that the vector a3 b is orthogonal to the vector b3 a . Exercises 10-11. Determine the angle between the vectors u and v. 10.u ==53  v 61 11. u ==0357  v 2104 Exercises 12-13. Express the given vector v as a sum of a vector parallel to the given vector u and a vector orthogonal to u. 12.v ==15  u 32 13. v ==23– 1 u –204 Exercises 14-15. Find a normal form, the general form, and a vector form representation (Theo- rem 2.20, page 72) of the plane passing through the point A0 with given normal vector n.

14.A0 ==213  n 12– 1 15. A0 ==12– 1 n 213 Exercises 16-18. Find both a normal form equation and a vector form representation (Theorem 2.20, page 72) for the given plane. 16.2x – 3y +2z = 17.4xz+1= 18. x –23y –2z = –

19. Find the distance from the point P = 14 and the line L in 2 passing through the points 12 and 21 . 20. Find the distance from the point P = 14– 1 and the line L in 3 passing through the points 121 and 210 . 21. Find the distance from the point P = 14– 11 and the line L in 4 passing through the points 121– 2 and 2102 . 22. Find the distance from the point P = 21– 2 to the plane x –24y +3z = . 23. Find the distance from the point P = 14– 2 to the plane 3xy++4z = 2 . 24. Determine the angle of intersection of the planes x –23y +1z = and 2xyz+2– = . Suggestions: Consider the normals to those planes. 290 Chapter 7 Inner Product Spaces

25. Find the set of vectors in 3 orthogonal to: (a) the vector 132 . (b) the vectors 132 and 22– 1 . (c) the vectors 132 , 20– 1 , and 25– 3 . Exercises 26-27. Find the general form equation of the plane that contains the given points. 26. 21– 210– 1 –301 27. 001 200 030

28. Find the angle between a main diagonal and an adjacent edge of a cube of volume 8 in.3 . 29. Prove Theorem 7.1(i). 30. Prove Theorem 7.1(ii). 31. Prove Theorem 7.1(iv). 32. Establish the following properties for uvw  n and r   : (a) 0  v ==v  0 0 (b) uv  w= uv + uw (c) u  rv = ruv (d) uv–  w = uw – vw (e) uvw – = uv – uw (f) –v  w ==vw – –vw 33. Show that two nonzero vectors v and u are normal to a given plane if and only if each is a scalar multiple of the other. 34. (Normal form equation of a line in R2 ) Express the line ax+ by = c in the form nv = np , where vxy=  , p is a point on the line, and n  0 is a normal to the line

n T 35. Let AM nn , and let uv   . Show that Auv = u  A v . (See Exercise 19, page 161).

n n 36.u   . Show that the function pu:    given by puv = uv is linear. What is the

kernel of pu ? 37. Let u  n . Show that if uv = 0 for every v  n , then u = 0 . 38. ( in n ) Let uv  n . Show that uv+ 2 = u 2 + v 2 if and only if uv = 0 . 39. (Parallelogram Law in n ) Let uv  n . Show that: uv+ 2 +2uv– 2 = u 2 + 2 v 2 40. Let uv  Rn . Prove that uv= if and only if uv+ and uv– are orthogonal.

n 41. Prove that if v1v2  vn  is such that vi  vj = 0 if 1  ijn  , then n v1v2  vn is a basis for  .

n 42. Let uv1 v2  vm   . Prove that if u is orthogonal to each vi , 1 im , then u is

orthogonal to every v  v1v2  vm . 7.1 Dot Product 291

43. (Cauchy-Schwarz Inequality in n ) Show that if uv  n , then uv  uv . Suggestion: (If u = 0 , then equality holds). For u  0 , use the fact that 0  ruv+  ruv+ = uu x2 ++2uv x vv to conclude that the discriminant of the quadratic polynomial uu x2 ++2uv x vv cannot be positive. 44. Use the above Cauchy-Schwuarz Inequality to show that for any nonzero vectors uv  n : ------uv  1 . uv 45. Establish the following properties for uvw  3 and rs : (a) ru  sv = rs uv (b) –u  v ==uv – –uv (c) uvw + = uv + uw (d) uvw  = uv  w 46. (Metric Space Structure of n ) Define the distance between two vectors uv  n to be duv = uv– . Prove that for all uvw  n : (a)duv  0 . (b)duv = 0 if and only if uv= . (c) duv = dvu (d)duw  duv + dvw Suggestion: Use the Cauchy-Schwuarz Inequality of Exercise 41. 47. (PMI) Use the principle of mathematical induction to show that for any n uv1 v2  vm   and any a1a2  am   : .

u  a1v1 +++a2v2  amvm = a1uv 1 +++a2uv 2  amuv m

PROVE OR GIVE A COUNTEREXAMPLE

48. Let uvw n . If u  0 and if uv = uw , then vw= . 49. Let uv  n . If vw = uw for every w  n , then uv= .

n 50. Let uv   . If vw= 1 + w2 and vz= 1 + z2 with w1 and z1 multiples of v, and if w2 and

z2 are orthogonal to u, then w1 = z1 and w2 = z2 . 51. Let uvz n , with u  0 . If u is orthogonal to both v and z, then v = cz for some c   . 52. The function N: n   given by Nv = v is linear. 53.uv = vu for all uv  3 . 292 Chapter 7 Inner Product Spaces

7

§2. INNER PRODUCT

As you know, a vector space V comes equipped with but two opera- tions: addition, and scalar multiplication. We now enrich that algebraic structure by adding another binary function on V — one inspired by the dot-product properties of Theorem 7.1, page 279:

While the scalar product DEFINITION 7.6 An inner product on a vector space V is a rv assigns a vector to a INNER PRODUCT function which assigns a real number uv scalar r and a vector v, to any two vectors uv  V , such that: the inner product u v assigns a real number to positive-definite axiom: (i)vv  0 , and vv = 0 only if v = 0 a pair of vectors. commutative axiom: (ii) uv = vu

homogeneous axiom: (iii) ruv = r uv

distributive axiom: (iv) uv+  w = uw + vw

INNER PRODUCT SPACE A vector space together with an inner product is said to be an inner product space. The Euclidean vector space n with dot product uv = uv is called the Euclidean inner product space of dimension n. There are other inner products that can be imposed on n , among them:

EXAMPLE 7.8 For any positive real numbers c1c2  cn : Why are we requiring WEIGHTED u1u2  un  v1v2  vn the c’s to be positive? EUCLIDEAN INNER PRODUCT SPACE = c1u1v1 +++c2u2v2  cnunvn is an inner product on n .

SOLUTION: We show that the distributive axiom (iv) holds and invite you to establish the remaining axioms in the exercises:

For uu===1u2  un  v v1v2  vn  w w1w2  wn :

uv+  w = u1 + v1u1 + v1  u1 + v1  w1w2 wn

= c1u1 + v1 w1 +++c1u1 + v1 w1  c1u1 + v1 w1

= c1u1w1 +++c2u2w2  cnunwn + c1v1w1 +++c2v2w2  cnvnwn = uw + vw 7.2 Inner Product 293

CHECK YOUR UNDERSTANDING 7.11

Verify that 2 2 a2x ++a1xa0 b2x ++b1xb0 = a2b2 ++a1b1 a0b0

is an inner product on P2 . Answer: See page B-33. Generalize the above to obtain an inner product on P3

The following theorem extends the Euclidean dot-product results of Exercise 32, page 290 to general inner product spaces: THEOREM 7.4 For every u, v, and w in an inner product space V: In the exercises you are (a) 0 v ==v 0 0 asked to establish the fol- lowing generalization and (b) uv + w = uv + uw combination of (b) and (c). (c) u rv = r uv For uv1 v2 vn  V

and c1 c2 cn  : (d) uv–  w = uw – vw n n (e) uv – w = uv – uw u  civi =  ci uv i i = 1 i = 1 (f) –v w ==vw – – vw

PROOF: We verify (d), and leave it for you to establish the rest.

Definition 2.7, page 55: uv–  w = uv+ –  w Axiom (iv): = uw + –vw Theorem 2.11 (x), page 56: = uw + –1vw Axiom (iii): = uv – uw

CHECK YOUR UNDERSTANDING 7.12

Answer: See page B-33. Prove: ru rv = r2 uv

DISTANCE IN AN INNER PRODUCT SPACE In the previous section we defined the norm (or magnitude) of a vec- tor v inn in terms of the dot product: vvv=  . The dot product was also used to describe the distance between two vectors u and v in n :uv– = uv–  uv– . Replacing “dot-product” with “inner product” enables us to extend the notion of magnitude and dis- tance to any inner product space: 294 Chapter 7 Inner Product Spaces

DEFINITION 7.7 The norm (or magnitude) of a vector v in an inner product space V, denoted by v , is NORM AND given by: DISTANCE vvv=  The distance between two vectors u and v in V is given by uv– .

CHECK YOUR UNDERSTANDING 7.13

Show that for any vectors u and v in an inner product space V and any r   : Answer: See page B-33. (a) rv = r v (b) uv+ 2 = u 2 ++2 uv v 2

EXAMPLE 7.9 Find the distance between the two vectors 2 p1x = 2x –1x + and p2x = 3x + 4 in the inner product space P2 of CYU 7.11

SOLUTION: Utilizing the inner product: 2 2 a2x ++a1xa0 b2x ++b1xb0 = a2b2 ++a1b1 a0b0 we have: 2 2 p1x –2p2x ==x –1x + –23x + 4 x –34x –

2 2 p1 – p2 ==p1 – p2 p1 – p2 2x –34x –2 x –34x – ==22 ++–4 2 –3 2 29

CHECK YOUR UNDERSTANDING 7.14

With reference to the weighted inner product space: u1u2 u3  v1v2 v3 = 5u1v1 ++2u2v2 4u3v3 on 3 (see Example 7.8), determine: (a) The magnitude of the vector 35– 8 . (b) The distance between the vector 35– 8 and the vector Answer: (a) 710 (b) 645 102 . 7.2 Inner Product 295

THE CAUCHY-SCHWARZ INEQUALITY

The following theorem will enable us to extend the concept of an angle between two vectors in n to vectors in an inner product space: THEOREM 7.5 For any two vectors u and v in an inner product CAUCHY- space: SCHWARZ uv  uv

PROOF: If either u = 0 or v = 0 , then uv = 0 and we are done. For u  0 and v  0 , we first show that uv  – uv : 1 1 1 1 Definition 7.5(i): ------u + ------v ------u + ------v  0 The proof sketched out in u v u v Exercise 43, page 291, can 1 1 1 1 1 1 also be used to establish 7.5(iv): ------u + ------v ------u + ------u + ------v ------v  0 this result. 7 u v u u v v .5 ( iv ) : 1 1 1 1 1 1 1 1 ------u ------u +++------v ------u ------u ------v ------v ------v  0 u u v u u v v v 1 2 1 CYU 7.10: ------uu ++0------uv ------vv  u 2 uv v 2 2 1 ++0------uv 1  uv 2 ------uv  –2 uv uv  – uv In the following Check Your Understanding box you are asked to show that uv  uv . Putting the two inequalities together, we come up with – uvuv uv , which is to say: uv  uv

CHECK YOUR UNDERSTANDING 7.15

Verify: uv  uv 1 1 1 1 Answer: Seepage B-34. Suggestion: Begin with ------u–------v ------u–------v  0 . u v u v 296 Chapter 7 Inner Product Spaces

Here are norm properties that are reminiscent of prop- erties in  : THEOREM 7.6 Let V be an inner product space. For all uv  V and r   : (a)v  0 , and v = 0 if and only if v = 0 (b) rv = r v

TRIANGLE INEQUALITY (c) uv+  uv+

PROOF: (a) A consequence of Definition 7.5(i) and vvv=  .

(b) rv ===rv rv r2 vv r2 vv =r v

CYU 7.11

CYU 7.12(b) (c) uv+ 2 = u 2 ++2 uv v 2 Cauchy-Schwarz inequality:  u 2 ++2 uv v2 = uv+ 2 Taking the square root of both sides of uv+ 2  uv+ 2 yields the desired result. We now extend the angle concept of Definition 7.3, page 281, to inner product spaces:

The Cauchy-Schwarz ine- DEFINITION 7.8 The angle  between two nonzero vectors u quality plays a hidden role ANGLE BETWEEN and v in an inner product space is given by: in this definition. (Where?) VECTORS –1 uv  = cos ------uv

EXAMPLE 7.10 Find the angle between the two vectors 2 p1x = 2x –1x + and p2x = 3x + 4 in the inner product space P2 of CYU 7.10 SOLUTION: While it is admirable that we were able to extend the geometrical notion of the angle between vectors in the plane to vectors in an arbi- trary inner product space, the real benefit of that generalization surfaces in the next section, where the concept of orthogonality takes center stage. 7.2 Inner Product 297

–1 2x2 –1x +3 x + 4  = cos ------2x2 –1x +3x + 4 –1 20 + –1  314+  = cos ------ 2x2 –1x +2 x2 –1x + 3x +34 x + 4 –1 1 = cos ------22 ++33–1 –1 11  + 44

–1 1 = cos ------ 84 615

CHECK YOUR UNDERSTANDING 7.16

With reference to the weighted inner product space:

u1u2 u3  v1v2 v3 = 5u1v1 ++2u2v2 4u3v3 Answer: 3 –1 –49 on  (see Example 7.7), determine the angle between the vectors cos ------ 119.1 351 29 35– 8 and 102 . 298 Chapter 7 Inner Product Spaces

EXERCISES

Exercises 1-8. With reference to the weighted inner product space:

u1u2 u3  v1v2 v3 = 5u1v1 ++2u2v2 4u3v3 of Example 7.7, determine: 1. The magnitude of the vector 12– 3 . 2. The magnitude of the vector 320 . 3. The distance between the vectors 12– 3 and 102 . 4. The distance between the vectors 320 and 12– 3 . 5. The angle between the vectors 12– 3 and 102 . 6. The angle between the vectors 320 and 12– 3 . 7. Verify that the Cauchy-Schwarz inequality holds for the vectors 12– 3 and 102 . 8. Verify that the Cauchy-Schwarz inequality holds for the vectors 320 and 12– 3 . Exercises 9-16. Referring to the inner product space: 2 2 a2x ++a1xa0 b2x ++b1xb0 = a2b2 ++a1b1 a0b0

on P2 of CYU 7.11, determine: 9. The magnitude of the vector 2x2 –3x + . 10. The magnitude of the vector – x2 + x – 5 . 11. The distance between the vectors 2x2 –3x + and – x2 + x – 5 . 12. The distance between the vectors 3x2 + 1 and 2x – 5 . 13. The angle between the vectors 2x2 –3x + and – x2 + x – 5 . 14. The angle between the vectors 3x2 + 1 and 2x – 5 . 15. Verify that the Cauchy-Schwarz inequality holds for the vectors 2x2 –3x + and – x2 + x – 5 . 16. Verify that the Cauchy-Schwarz inequality holds for the vectors 3x2 + 1 and 2x – 5 .

a11 a12 b11 b12 17. For A =  B = in the vector space M22 , define: a21 a22 b21 b22

AB = a11b11 +++a12b12 a21b21 a22b22

Show that the above operator is an inner product on M22 . 7.2 Inner Product 299

Exercises 18-22. Referring to the inner product space on Exercise 17, determine:

18. The magnitude of the vector 13 . 02

19. The magnitude of the vector 21– . 10

20. The distance between the vectors 13 and 21– . 02 10

21. The angle between the vectors 13 and 21– . 02 10

22. Verify that the Cauchy-Schwarz inequality holds for the vectors 13 and 21– . 02 10

n n n i i 23. Verify that  aix   bix =  aibi is an inner product on the polynomial space Pn . i = 0 i = 0 i = 0

24. (Calculus Dependent) (a) Show that Cab = f: ab  Rf is continuous is a sub- set of the function vector space Fab of Theorem 2.4, page 44. b (b) Show that fg = fxgxdx is an inner product on Cab (called the standard a inner product on Cab ). Exercises 24-35. Calculus Dependent) Referring to the inner product space on Exercise 24, determine: 25. The magnitude of the vector 2x2 –3x + in the inner product space C01 . 26. The distance between the vectors 2x2 –3x + and – x2 + x – 5 in the inner product space C01 . 27. The angle between the vectors 2x2 –3x + and – x2 + x – 5 in the inner product space C01 . 28. The magnitude of the vector ex in the inner product space C01 . 29. The distance between the vectors ex and x in the inner product space C01 . 30. The angle between the vectors ex and x in the inner product space C01 . 31. The magnitude of the vector sinx in the inner product space C–  . 32. The distance between the vectors sinx and cosx in the inner product space C–  . 33. The angle between the vectors sinx and cosx in the inner product space C–  . 300 Chapter 7 Inner Product Spaces

34. Verify that the Cauchy-Schwarz inequality holds for the vectors ex and x in the inner product space C01 . 35. Verify that the Cauchy-Schwarz inequality holds for the vectors sinx and cosx in the inner product space C–  . 36. Prove that ordinary multiplication in the set of real numbers R is an inner product on the vec- tor space  . 37. Prove Theorem 7.3(a). 38. Prove Theorem 7.3(b). 39. Prove Theorem 7.3(c). 40. Prove Theorem 7.3(e). 41. Prove Theorem 7.3(f).

42. Let uv  V , V an inner product space. Show that uv+ 2 –4uv– 2 = uv 43. Let uv  V , V an inner product space. Show that uv+  uv– = u 2 – v 2 . 44. Let uv  V , V an inner product space. Show that uv+ 2 = u 2 + v 2 if and only if uv = 0 . 45. Let u  V , V an inner product space. Show that v  V uv = 0 is a subspace of V. 46. (PMI) Let V be an inner product space.Use the principle of mathematical induction to show

that for any uv1 v2  vn  V and any a1a2  am   :

u a1v1 +++a2v2  amvn = a1 uv 1 +++a2 uv 2 am uv n and:

a1v1 +++a2v2  amvn u = a1 uv 1 +++a2 uv 2 am uv n

PROVE OR GIVE A COUNTEREXAMPLE

47. Let uvw  V , V an inner product space. If u  0 and if uv = uw , then vw= .

48. Let uv  V . If vw = uw for every w  V , then uv= .

49. There exists an inner product on 3 for which 111 = 1 .

50. There exists an inner product on 3 for which 111 = 222 .

51. There exists an inner product on 3 for which 111  222 . 7.3 Orthogonality 301

7

§3. ORTHOGONALITY Having extended the concept of angles between vectors in n to vec- tors in inner product spaces, we can now extend the definition of orthogonality to those spaces: DEFINITION 7.9 Two vectors u and v in an inner product ORTHOGONAL VECTORS space V are orthogonal if uv = 0 . A set S of vectors in an inner product space ORTHOGONAL SET V is an orthogonal set if uv = 0 for every uv  S , with uv .

EXAMPLE 7.11 Verify that: S = 2x2 + 2x – 1 –2x2 ++x 22 x2 –2x + is an orthogonal set in the inner product space

P2 of CYU 7.11, page 293, wherein 2 2 a2x ++a1xa0 b2x ++b1xb0

= a2b2 ++a1b1 a0b0

SOLUTION: All pairs of vectors from S are orthogonal: 2x2 + 2x – 1 –2x2 ++x 2 ==21– +22 + – 12 0 (check) 2x2 +22x – 1 x2 –2x + ==22 +21– +–1 2 0 (check) –2x2 ++2x 2 x2 –2x + ==–1 221+– +22 0 (check)

You can check directly that the set S in the above example is a lin-

early independent in P2 . In general:

THEOREM 7.7 If v1v2  vn is an orthogonal set of non- zero vectors in an inner product space V , then v1v2  vn is a linearly independent set in V. PROOF: Let: c1v1 ++c2v2 cnvn = 0 c1v1 ++c2v2 cnvn = 0 For each vi , 1 in , we have:   Therem 7.4(a), page 277: vi c1v1 ++civi ++cnvn ==vi 0 0   c1 vi v1 ++ci vi vi ++cn vi vn = 0

vi vj = 0 if ij : ci vi vi = 0 Definition 7.5 (i), page 276: c = 0 each ci = 0 i 302 Chapter 7 Inner Product Spaces

CHECK YOUR UNDERSTANDING 7.17

Let v1v2  vm be an orthogonal set of vectors in an inner prod- uct space V, and let u  V be such that uv i = 0 for 1 im . Answer: See page B-34. Prove that u is orthogonal to every vector in Spanv1v2  vm .

DEFINITION 7.10 A in an inner product space is UNIT VECTOR a vector v of magnitude 1.

NORMALIZATION To normalize a nonzero vector v in an inner product space simply mul- 1 tiply it by ------: v 1 1 ------v ==------v 1 v v CYU 7.13(a), page 294

DEFINITION 7.11 An orthonormal set of vectors in an inner ORTHONORMAL SET product space is a set of orthogonal unit vectors.

The standard basis S = e1e2  en of page 94 can easily be shown to be an orthonormal set in the Euclidean inner product space n  . Moreover, for any v = c1 ci  cn :   vve=  1 e1 ++ve i ei ++ve n en since: ve i ==c1 ci  cn  0 1  0 ci ith entry This nicety extends to any orthonormal basis in any inner product space:

THEOREM 7.8 If  = v1v2  vn is an orthonormal basis in an inner product space V, then, for any v  V :

vvv=  1 v1 +++vv 2 v2  vv n vn n

PROOF: Let v =  cjvj . We show ci = vv i for 1 in : j = 1 n n 2 vv i ===== cjvj vi  cj vj vi ci vi vi ci vi ci j = 1 j = 1

1 if ji= Exercise 46, page 300. vj vi =  vi vi = 1 0 if ji 7.3 Orthogonality 303

CHECK YOUR UNDERSTANDING 7.18

Let  = v1v2 vn be an orthonormal basis for an inner prod- uct space V. Show that for any a1v1 +++a2v2  anvn and n

w = b1v1 +++b2v2  bnvn , vw =  aibi . Answer: See page B-34. i = 1 The following theorem spells out a procedure that can be used to con- struct an orthogonal basis in any given finite dimensional inner product space. Basically, the construction process is such that each newly added vector in an evolving basis is orthogonal to all of its predecessors.

THEOREM 7.9 Let  = v1v2  vn be a basis for an inner product space V, GRAHM-SCHMIDT and let V be the following subspaces of V: PROCESS i

u1 = v1 V1 = Spanv1

u1 v2 u2 = v2 – ------u1 V2 = Spanv1 v2 u1 u1

u1 v3 u2 v3 u3 = v3 – ------u1 – ------u2 V3 = Spanv1v2 v3 . u1 u1 u2 u2 . . . u1 vn u2 vn un – 1 vn un = vn – ------u1 – ------u2 –  – ------un – 1 Vn = Spanv1v2  vn u1 u1 u2 u2 un – 1 un – 1

Then, u1 ui is an orthogonal basis for Vi . In particular,

u1u2  un is an orthogonal basis for V. PROOF: By Induction on the dimension of V:

Since u1 = v1 and since u1 is an orthogonal set, the claim is seen to hold for Vi . Assume that u1 uk is an orthogonal basis for Vk , for kn .

We show that u1 uk + 1 is an orthogonal basis for Vk + 1 , where: u  v u  v 1 k + 1  k k + 1 uk + 1 = vk + 1 – ------u1 – – ------uk (*) u1 u1 uk uk

We are assuming that u1 uk is an orthogonal set. Conse- quently, to establish orthogonality of u1 uk + 1 , we need but show that ui uk + 1 = 0 for 1 ik : 304 Chapter 7 Inner Product Spaces

u1 vk + 1 uk vk + 1 ui uk + 1 = ui vk + 1 – ------u1 –  – ------uk u1 u1 uk uk u  v u  v u  v 1 k + 1  i k + 1 i k + 1 = ui vk + 1 – ------ui u1 – – ------ui ui –  – ------ui uk u1 u1 ui ui ui ui u  v ------i ------k + 1 = ui vk + 1 – ui ui since ui uj = 0 if ij ui ui

ui ui ===ui vk + 1 – ------ui vk + 1 ui vk + 1 –01 ui vk + 1 ui ui

Being an orthogonal set, u1 uk + 1 is linearly independent. To show that Span u1 uk + 1 = Spanv1v2  vk + 1 we need but show that Spanv1v2  vk + 1  Span u1 uk + 1 (why?). Let’s do it: k + 1 k + 1 k a v  Spanv  v  i i 1 k + 1 a v = a v + a v i = 1  i i  i i k + 1 k + 1 i = 1 i = 1 k Induction Hypothesis: =  biui + ak + 1vk + 1 i = 1 k k uk vk + 1 from (*): = b u + a u + ------u  i i k + 1k + 1  k uk uk i = 1 i = 1 k + 1 k + 1  aivi  Spanu1 uk + 1 = ciui i = 1  i = 1

u1 v2 u2 v3 where: c1 = b1 + ak + 1------c2 = b2 + ak + 1------ ck + 1 = ak + 1 u1 u1 u2 u2

Note: To obtain an orthonormal basis for an inner product space, simply normalize the orthogonal basis generated by the Gram-Schmidt process.

EXAMPLE 7.12 Extend 120 to an orthogonal basis for the Euclidean inner product space 3 . 7.3 Orthogonality 305

SOLUTION: ONE APPROACH: Extend 120 to a basis for 3 : 120 100 001 :

v1v2 v3 and then apply the Grand-Schmidt process to the above basis:

u1 = 120 u  v Multiplying any ui in the 1 2 120  100 u2 ==v2 – ------u1 100 – ------120 Gram-Schmidt process u1 u1 120  120 by a nonzero constant 1 4 2 will not alter that vectors ==100 – ---120 ---–--- 0 42– 0 21– 0 “orthogonality-feature,” 5 5 5 but will simplify subse- see margin quent calculations. u1 v3 u2 v3 u3 = v3 – ------u1 – ------u2 u1 u1 u2 u2 120 001 21– 0 001 = 001 – ------ 120 – ------ 21– 0 120  120 21– 0  21– 0 0 0 ==001 – ---120 – ---21– 10 001 5 5 The above process led us to the following orthogonal basis:

u1u2 u3 = 120 21– 0 001

ANOTHER APPROACH: Since we are dealing with a vector space of dimension 3, we can sim- ply roll up our sleeves and construct an orthogonal bases by “brute This brute force approach is not always practical. force.” First, add a nonzero vector, abc , to 120 with: Software, such as Maple and MATLAB, include the abc  120 = 0 Gram-Schmidt process as a ++2b 0  c = 0 a built-in procedure. Yes, the Gram-Schmidt process This can be done in many ways. One way: abc = 001 , works off of a basis for the brings us to the orthogonal set 120  001 . We still need inner product space, but another nonzero vector abc — one that is orthogonal to both that is not a problem: if you randomly choose n 120 and 001 : vectors in an n dimen- sional space, even if abc  120 ==0 and abc  001 0 n = 100 , there is little a +2b +0  c = 0 and 0  a ++00  b c = chance that those vectors a +2b = 0 and c = 0 end up being linearly dependent! How about abc = 630 ? Sure. Leading us to the orthogonal basis 120 001 630 306 Chapter 7 Inner Product Spaces

CHECK YOUR UNDERSTANDING 7.19 Apply the Gram-Schmidt Process to construct an orthonormal basis for the subspace S of the Euclidean inner product space of 4 spanned by the vectors 2110 , 1010 , 3120 , Answer: See page B-35. 0101 .

ORTHOGONAL COMPLEMENT The set of vectors orthogonal to any subspace W of 3 , denoted by the symbol W , is itself a subspace of 3 . More specifically:

W W 0 3 Line W passing through the origin. Plane passing through the origin with normal W. Plane W passing through the origin. Line passing through the origin orthogonal to W. 3 0 In a more general setting, for W a subspace of an inner product space ORTHOGONAL V we define the orthogonal complement of W to be: COMPLEMENT W = u  V uw = 0 for every w  W We then have: THEOREM 7.10 If W a subspace on an inner product space V, then: (i)W is a subspace of V. (ii) WW  = 0 (iii) Every vector in V can be uniquely expressed as a sum of a vector in W and a vector in W .

(iv) If  is a basis for W and   is a basis W W  for W , then     is a basis for V. W W  PROOF: (i) For u1 u2  W , r   , and w  W , we have:

ru1 + u2 w ===r u1 w + u2 w r  00+ 0 The result now follows from Theorem 2.13, page 61. (ii) Let v  WW  . Being in both W and W , vv = 0 . It fol- lows, from Axiom (i) of Definition 7.6 (page 292) that v = 0 . In this part of the theo- (iii) Let w1w2  wm be an orthonormal basis for W. For v  V , rem we assume that W is let: finite dimensional (the v = vw w +++vw w  vw w  W result does, however, W 1 1 2 2 m m hold in general).  We show that the vector vW = vv– W is in W by showing that

vW wi = 0 , for 1 im (see CYU 7.17): 7.3 Orthogonality 307

vW wi ==vv– W  wi vw i – vW  wi

= vw i – vw 1 w1 +++vw 2 w2  vw m wm wi

===vw i – vw i wi wi vw i –0vw i since w  w = 0 for ij i j since wi wi = 1 At this point, we have shown that v can be expressed as a sum of a  vector in W and a vector in W : vv= W + vW . Uniqueness of the decomposition follows from part (ii) of this theorem, and Exercise 43, page 67. As for (iv): CHECK YOUR UNDERSTANDING 7.20 Answer: See page B-35. Establish Theorem 7.10(iv) Lets highlight an important observation lurking within the proof of Theorem 7.10(iii):

THEOREM 7.11 Let w1w2  wm be an orthonormal basis for a subspace W of an inner product space V. For any v  V there exist unique  vectors v  W and v   W such that v W W v W vv= + vW , where Compare with Theorem W v = vw w +++vw w  vw w 7.2, page 283. W 1 1 2 2 m m

W v  W and vW = vv– W .

Note: vW is said to be the orthogonal projection v of v onto W, and we write: vW = projW .

EXAMPLE 7.13 Let W = Span = v1v2  vn , in the Euclidean inner product space 4 . (a) Find a basis for W . (b) Express 1234 as the sum of a vector in W and a vector in W . SOLUTION: (a) Since 1002 and 1110 are linearly inde-   pendent, they constitute a basis for W. To say that abcd  W is We know that W will to say that: turn out to be of dimen- sion 2. How? abcd  1002 = 0 and abcd  1110 = 0 a +02d = abc+ +0= d = –--a- ba= – – c 2 308 Chapter 7 Inner Product Spaces

Choosing a and c to be our free variables, we have: a W = aa– – c c –--- a c    2 First setting a = 0 and c = 1 , and then setting c = 0 and a = 2 leads us to the basis: 0 – 110  22– 01– for W . (b) One approach: Simply express 1234 as a linear combination of the basis 1002 1110 0 – 110 22– 01– :

1234 = a1002 ++b1110 c0 – 110 +d22– 01–

ab++2d = 1  bc–2–2d =  3 3 3   a ====--- b --- c --- d –1 bc+3=  2 2 2  2ad–4=  Bringing us to: in W in W 3 3 3 1234 = ---1002 + ---1110 + ---0– 110 – 22– 01– 2 2 2

= 3--3- --3- 3 + –2--1- --3- 1 2 2 2 2

Another Approach: First apply the Gram-Schmidt method on 1002  1110 to obtain an orthonormal basis for W:

u1 = 1002 1110  1002 u = 1110 – ------1002 2 1002  1002 1 4 2 Since u2 is orthogonal ==1110 – ---1002 ---11–--- Or: 455– 2 to u , so is 5u 5 5 5 1 2 see margin Orthonormal basis for W: 1 1 w1 w2 = ------1002  ------455– 2 5 70 Applying Theorem 7.11 we know that

1234 = 1234 W + 1234 W where:

1234 W = 1234  w1 w1 + 1234  w2 w2 1 2 1 2 4 5 5 –2 4 5 5 –2 = 1234  ------00------------00------+ 1234  ------------------------5 5 5 5 70 70 70 70 70 70 70 70 9 1 2 21 4 5 5 –2 3 3 ==------------00------+ ------------------3------3 55 5 7070 70 70 70 2 2 and: 3 3 1 3 1234  ===1234 – 1234 1234 – 3------3 –2------1 W W 2 2 2 2 7.3 Orthogonality 309

Note that both approaches lead to the same decomposition, as must be the case: 3 3 1 3 1234 = 3------3 + –2------1 2 2 2 2 in W in W

CHECK YOUR UNDERSTANDING 7.21 Find the orthogonal projection of the vector v = 201 onto the subspace W = Span 101  120 of the Euclidean inner Answer: See page B-35. product space 3 . The shortest distance between a vector v in an inner product space V and Consider Example 7.3, any vector w in a subspace W of V turns out to be the distance between page 284. v and vW : THEOREM 7.12 Let W be a subspace of the inner product space V, and let v  V . Then: v – proj v  vw– for every w  W W

vW

ROOF v P : Let vW = projW . For any w  W : vw– 2 = vw–  vw–

= vv– W + vW – w  vv– W + vW – w Theorem 7.4(b), page 293: = vv– W vv– W +++vv– W vW – w vW – w vv– W vW – w vW – w

2 2 = vv– W +++vv– W  vW – w vW – w vv– W vW – w  Since vW – w  W and vv– W  W , the two middle terms in the above expression are 0, bringing us to: 2 2 2 vw– = vv– W + vW – w

To complete the proof we need but note that vW –0w  .

CHECK YOUR UNDERSTANDING 7.22

Find the shortest distance between the vector v =33x2 + x and the 2 3 subspace W = Span x + 1 x + 1 in the inner product space P3 of CYU 7.11, page 293: 3 2 3 2 3 a0 ++a1xa2x +a3x  b0 ++b1xb2x +b3x =  aibi . Answer: 3 i = 0 310 Chapter 7 Inner Product Spaces

EXERCISES

Exercises 1-7. Determine if the given set of vectors is an orthogonal set in the given inner product space. If so, modify the set to arrive at an orthonormal set. 1.111 –12– 1 –101 in the Euclidean inner product space 3 . 2.111 –12– 1 –101 in the weighted inner product space of Example 7.8, page 292, with u1u2 u3  v1v2 v3 = 5u1v1 ++2u2v2 4u3v3 . 3.111 –12– 1 105 in the weighted inner product space of Example 7.8, page 292, with u1u2 u3  v1v2 v3 = 5u1v1 ++u2v2 u3v3 . 4.x2 ++2x 13 x2 + x – 511 x2 –58x + in the polynomial inner product space of CYU 7.11, page 293.  12 21– 00 5.1 1 in the inner product space of Exercise 17, page 298. 34 00 --- –--- 3 4

4 6. (Calculus Dependent) x2 –1---x + in the inner product space C01 of Exercise 24, 3 page 299. 4 7. (Calculus Dependent) x2 –1---x + in the inner product space C12 of Exercise 24, 3 page 299. 8. Use Theorem 7.8 to express 352 in the Euclidean inner product space 3 as a linear 1 1 combination of the vectors in the orthonormal basis ------210 001 ------120–  . 5 5

9. Use Theorem 7.8 to express 2x2 + 3x – 1 in the polynomial inner product space of CYU 7.11, page 293, as a linear combination of the vectors in the orthonormal basis x2 x 1 x2 2x 1 x2 1 ------++------------– ------+ ------– ------. 3 3 3 6 6 6 2 2 10. Find all values of a for which 132  1a 1 is an orthogonal set in the Euclidean inner product space 3 . 11. Find all values of a and b for which 13b  1a 1 is an orthogonal set in the Euclid- ean inner product space 3 . 12. Find all values of a and b for which 11a –12b –1 –101 is an orthogonal set in the weighted inner product space of Example 7.8, page 292, with u1u2 u3  v1v2 v3 = 2u1v1 ++u2v2 u3v3 . 7.3 Orthogonality 311

13. Find all values of a and b for which 111 –12– 1 –101 is an orthogonal set in the weighted inner product space of Example 7.8, page 292, with u1u2 u3  v1v2 v3 = au1v1 ++bu2v2 u3v3. 14. Find all values of a, b, and c for which 111 –12– 1 105 is a n orthogonal set in the weighted inner product space of Example 7.8, page 292, with u1u2 u3  v1v2 v3 = au1v1 ++bu2v2 cu3v3. 15. (Calculus Dependent) Find all values of a and b for which ax2 + 1 – x + b is an orthog- onal set in the inner product space C01 of Exercise 24, page 299. 16. (Calculus Dependent) Find all values of a, and b for which x2 –1x + is an orthogonal set in the inner product space Cab of Exercise 24, page 299. Exercises 17-26. Find an orthonormal basis for the given inner product space. 17.Span201  –120 in the Euclidean inner product space 3 . 18.Span111  –12– 1 in the Euclidean inner product space 3 . 19.Span1111  –12– 12 in the Euclidean inner product space 4 20.Span1010 0202 2000 in the Euclidean inner product space 4 . 21.Span201  –120 in the weighted inner product space of Example 7.8, page 292, with u1u2 u3  v1v2 v3 = 2u1v1 ++3u2v2 u3v3 . 22.x2 + 1 x – 5 in the polynomial inner product space of CYU 7.11, page 293. 2x ++03yz– w =  23. The solution set of  in the Euclidean inner product space 4 . 4x +03y –22z – w = 

x ++03y – 2z w =  24. The solution set of  in the Euclidean inner product space 4 . 4x +03y – 2z =  25. (Calculus Dependent) Spanx2 2x + 1 in the inner product space C01 of Exercise 24, page 299. 26. (Calculus Dependent) Spanx3x + 1 x2 – 1 in the inner product space C–11 of Exercise 24, page 299.

27. Find an orthonormal basis for a 2a 0 aR in the Euclidean inner product space 3 .

28. Find an orthonormal basis for aba– b 2b ab   in the Euclidean inner product space 4 . 29. Find an orthonormal basis for a2aca– c ac   in the Euclidean inner product space 4 . 30. Find an orthonormal basis for abc abc++= 0 in the weighted inner product space

of Example 7.8, page 292, with u1u2 u3  v1v2 v3 = –5u1v1–2u2v2 + u3v3 . 31. Find an orthonormal basis for ax3 +++ab+ x2 cx 2a abc  R in the polynomial inner product space of CYU 7.11, page 293. 312 Chapter 7 Inner Product Spaces

Exercise 32-36. (a) Find a basis for the orthogonal complement of the given Euclidean inner product subspace W . (b) Express the given vector v as a sum of a vector in W and a vector in W . (c) Determine the distance from v to W. 32.W = Span10  11 , v = 13 . 33.W = Span102 , v = 13– 2. 34.W = Span102  13– 2 , v = 111 . 35.W = Span1020  0101 , v = 4 – 133 . 36.W = Span1023 2101 1101 , v = 13– 22 . 37. Find a basis for the orthogonal complement of the subspace W = Span201  –120 in the weighted inner product space of Example 7.8, page 292, with u1u2 u3  v1v2 v3 = u1v1 ++2u2v2 u3v3 , and express the vector v = 12– 1 as a sum of a vector in W and a vector in W .

38. Find a basis for the orthogonal complement of the subspace W = Spanx2 + 1 x – 5 in the polynomial inner product space of CYU 7.11, page 293, and express the vector v = 2x2 – x as a sum of a vector in W and a vector in W .

39. Prove that the standard basis e1e2  en of page 94 is an orthonormal basis in the Euclidean inner product space n . 40. Prove that xnxn – 1  x 1 is an orthonormal basis in the polynomial inner product space of CYU 7.11, page 277. 41. Let V be an inner product space. Prove that V = 0 and that 0  = V .

 42. Let W = Spanw1w2  wm in an inner product space V. Prove that v  W if and only if vw i = 0 for all wi , 1 im .

43. Let v1v2  vk vk + 1  vm be an orthogonal set in an inner product space V. Show that if w  Spanv1v2  vk and zv k + 1 vm , then wz = 0 .

44. Let W be a subspace in an inner product space V. Prove that W  = W . 45. Let w be a vector in an inner product space V of dimension n. Prove that w = v  V vw = 0 is a subspace of V of dimension n – 1 . 46. Let S be a subset of an inner product space V. Prove that S = v  V vw = 0 for all w  S is a subspace of V. 47. Let S be a subspace of an inner product space V of dimension n. Prove that dimS + dimS = n

48. Let  = v1v2  vn be an orthonormal basis in an inner product space V. Show that for any uv  V uv = u   v  . (See Definition 5.9, page 178.) 7.3 Orthogonality 313

Exercises 51-59 (Orthogonal Matrices) AM nn is an orthogonal matrix if the columns of A is an orthonormal set in the Euclidean inner product space n . (Orthogonal matrices would better have been named “orthonormal matrices,” no?) 49. Sow that the following are equivalent:

(i) AM nn is orthogonal. (ii) ATAI= . (See Exercise 19, page 162). (iii)AXX= for every X  n (iv)AX  AY = XY for every XY n . 50. Prove that every orthogonal matrix is invertible, and that its inverse is also orthogonal. 51. Prove that a product of orthogonal matrices (or the same dimension) is again orthogonal. 52. Prove that if A is orthogonal, then detA = 1 . 53. Prove that if A is orthogonal then the rows of A also constitute an orthonormal set. 54. Prove that if A is orthogonal, and if B is equivalent to A, then B is also orthogonal.

55. Prove that every 22 orthogonal matrix is of the form ab– or ab where ba ba– a2 +1b2 = .

56. Show that every 22 orthogonal matrix is of the form cos –sin or cos sin . sin cos sin –cos 57. Show that every 22 orthogonal matrix corresponds to either a rotation or a reflection about a line through the origin in 2 .

58. (a) Prove that the null space of AM mn is the orthogonal complement of the row space of A. T (b) Prove that the null space of A  Mmn is the orthogonal complement of the column space of A. (See Exercise 19, page 162.)

(c) Verify directly that the null space of A = 1320 is the orthogonal complement of the –0121 row space of A. (d) Verify directly that the null space of AT = 1 320 is the orthogonal complement of –1 012 the column space of A. (See Exercise 19, page 162.) 314 Chapter 7 Inner Product Spaces

59. (Bessel’s Equality) Let v1v2  vn be an orthonormal basis for an inner product space n 2 2 V. Prove that for any w  V :  wv i = w . i = 1

PROVE OR GIVE A COUNTEREXAMPLE

60. If v1v2  vm is an orthogonal set in an inner product space V, then

a1v1a2v2  amvm is an orthogonal set for all scalars a1a2  am .

61. If v1v2  vm is an orthonormal set in an inner product space V, then

a1v1a2v2  amvm is an orthonormal set for all scalars a1a2  am .

62. Let W be a subspace of an inner product space V. If wv = 0 with w  W , then v  W .

63. Let v1v2  vk  vn be a basis for an inner product space V such that each vj for kj  n is orthogonal to every vm for 1 mk . If W = Spanv1v2  vk , then  W = Spanvk + 1 vn .

64. Let v1v2  vk  vn be an orthogonal basis for an inner product space V. If  W = Spanv1v2  vk hen W = Spanvk + 1 vn . 7.4 The Spectral Theorem 315

7

§4. THE SPECTRAL THEOREM

We begin by recalling Definition 6.13 of page 264: In other words, the ith row The transpose of a matrix A = aij  Mmn is the th of A is the i column of T matrix A = bij  Mnm , where bij = aji AT . For example:

12 103 T If A =  A = 04 DEFINITION 7.12 T 245 A  Mnn is symmetric if A = A. 35 SYMMETRIC MATRIX As it turns out:

THEOREM 7.13 If A  Mnn is a (real) symmetric matrix, then its eigenvalues are real numbers. An outline of a proof for the above theorem, which involves a bit of terminology, is relegated to the exercises.

2 1 n n We remind you that we are using  to denote Mn  1 . For XY 3  4 = 235  140 5 0 we now define XY to be the dot product of the corresponding vertical = 21 + 34 + 50 n-tuples (see margin). It is easy to show that n ,with XY defined to = 14 be XY , is an inner product space (see Definition 7.6, page 292). Note, that the above dot product can also be effected by means of matrix mul- 2 1 235 1 tiplication:  = 3 4 4 XY = XTY (see margin) 5 0 0 = 21 + 34 + 50 THEOREM 7.14 Any two eigenvectors in the inner product ==14 14 space n corresponding to distinct eigen- values of a symmetric matrix A  Mnn are orthogonal.

PROOF: Let 1 2 be distinct eigenvalues of a symmetric matrix A  Mnn , with corresponding eigenvectors XY , so that: AX = 1X and AY = 2Y We show that XY = 0 : T 1XY ==1X  Y AX Y Exercise 19(f), page 162 = XTAT Y Definition 7.12: ==XTA YXTAY T T ==X 2Y 2X Y =  XY 2 Then: 1XY = 2XY  1 – 2 XY = 0

Since 1  2 , XY = 0 ; which is to say: X and Y are orthogonal. 316 Chapter 7 Inner Product Spaces

EXAMPLE 7.14 Verify the result of Theorem 7.14 for the sym- metric matrix: 11–0 A = –211 – 01–1

SOLUTION: To determine the eigenvalues of A, we turn to Theorem 6.8, page 219, and calculate the determinant of A – I :

1 –1 –0 detA – I ==det –21 –1 – –– 1 – 3 01–1–  details omitted

We see that A has three distinct eignevalues: 1 = 0 , 2 = 1 and 3 = 3 , with corresponding eigenspaces:

11–0 Note: E0 ==nullA – 0I null –211 – =aaa aR 11–0 10– 1 01–1 rref –211 – = 01– 1 margin 01–1 00 0 01–0 01–0 101 E1 ==nullA – 1I null –111 – =–0bb bR rref –111 – = 010 01–0 01–0 000 –12 –0 –12 –0 10– 1 E3 ==nullA – 3I null –11 –1– =c–2c c cR rref = –11 –1– 01 2 01–2– 01–2– 00 0 You can easily verify that for 1  ij  3 , every vector in Ei is orthogonal to every vector Ej . For example: aaa  c–2c c ==ac – 2ac +0ac

CHECK YOUR UNDERSTANDING 7.23

211 Verify the result of Theorem 7.13 for the matrix A = 121 . Answer: See page B-37. 112

THEOREM 7.15 A  Mnn is symmetric if and only if AX  Y = XAY 

for all vectors XY  Rn . 7.4 The Spectral Theorem 317

PROOF: If A is symmetric, then: AX  Y ==AX TYXTAT YX =TATY e e 2 3 : T Exercise 19(f), page 161 ==X AY XAY   Conversely, assume that Aa=  is such that 135 0 0 3 0 ij 267 1  0 = 6  0 (*)  AX  Y = XAY  4 5 2 0 1 5 1 th Turning to the n-tuple ek , with 1 as its k entry and 0 elsewhere, we = 5 show that A is symmetric, by showing that aij = aji : a32 aij = Aej  ei (see margin) By (*): = ej  Aei Theorem 7.1(ii), page 279: ==Aei  ej aji

CHECK YOUR UNDERSTANDING 7.24

111 (a) Let A = 121 . Show directly that AX  Y = XAY  for 113 every XY  3 .

(b) Write down an arbitrary non-symmetric matrix AM 33 and Answer: See page B-37. exhibit XY  3 for which AX  Y  XAY  .

SYMMETRIC OPERATORS As you know, there is an intimate relation between matrices and lin- ear maps; bringing us to: DEFINITION 7.13 Let V be an inner product space. A linear SYMMETRIC OPERATOR operator T: VV is symmetric if Compare with Theorem 7.15. Tv  w = v Tw for all vectors vw  V .

Here is the linear operator version of Theorem 7.14: THEOREM 7.16 Let T: VV be a symmetric linear opera- tor on an inner product space V. If v1 and v2 are eigenvectors associated with distinct eigenvalues 1 and 2 , then v1 and v2 are orthogonal. 318 Chapter 7 Inner Product Spaces

PROOF: Tv1  v2 –0Tv1  v2 =

Definition 7.13: Tv1  v2 –0v1 Tv2 =

Definition 6.5, page 224: 1v1 v2 –0v1 2v2 =

Theorem 7.4, page 293: 1 v1 v2 –02 v1 v2 =

1 – 2 v1 v2 = 0

Since 1  2 , v1 v2 = 0 .

The matrix representation T  of a symmetric linear operator need not be symmetric for every basis  (see Exercise 18). However: THEOREM 7.17 If T: VV is a symmetric linear operator

on an inner product space V, then T  is a symmetric matrix for any orthonormal basis  of V.

PROOF: Employing Theorem 7.15, we show that for any two (column) n n-tuples vv= 1v2  vn , ww= 1w2  wn in 

T v  w = v  T w :

T v  w = T v   w 

Theorem 5.22, page 180: = Tv   w  CYU 7.18, page 303: = Tv  w By symmetry: ==v Tw v   Tw  =v  T w

CHECK YOUR UNDERSTANDING 7.25

For 3 the Euclidean inner product space, let T: 3  3 be the linear transformation given by: Tabc = ab– – a + 2bc– – b + c (a) Show that T is symmetric.

(b) Verify that T  is symmetric for the orthonormal basis Answer: See page B-38.  = 010 001 100 of 3 . Theorem 7.9, page 303, assures us that every finite dimensional inner product space contains an orthonormal basis  . It follows that every symmetric linear operator T: VV has a symmetric matrix represen- tation T  . Indeed,  can be chosen so that T  is a diagonal matrix: 7.4 The Spectral Theorem 319

Note that V contains an THEOREM 7.18 A linear operator T: VV on an inner orthonormal basis if and THE SPECTRAL product space V is symmetric if and only if only if it contains a nor- THEOREM mal basis. V contains an orthonormal basis of eigen- vectors of T.

PROOF: Assume that  = v1v2 vn is an orthonormal basis of eigenvectors of T with corresponding eigenvalues 12 n . We show that T is symmetric: n n Let vw  V , with v =  aivi and w =  bivi . Then: i = 1 i = 1 n n

Tv  w =  aiTvi   bivi i = 1 i = 1 n n

=  aivi  bivi i = 1 i = 1 n n n

== aibi  aivi  bivi =v Tw i = 1 i = 1 i = 1 1 if ij= Since vi vj =  [along with Theorem 7.4, page 293]  0 if ij To establish the converse, we apply the Principle of Mathematical Induction on the dimension n of V: v I. Let dimV = 1 . For any v  0 , ----- is an orthonormal v v v basis for V. Since T -----  V , and since ----- is a basis for V,   v v v v T ----- = ----- for some  . v v II. Assume that the claim holds for dimV = k . III. We establish validity for dimV = k + 1 : Applying the Grahm-Schmidt process, we can obtain an ortho- normal basis  for V. By Theorem 7.17, T  is symmetric. Let  be a (real) eigenvalue of T  (Theorem 7.13). Let n v   be an eigenvector associated with  , and let vn be such

that vn  = v . Then: 320 Chapter 7 Inner Product Spaces

T v = v  T vn  = vn  Theorem 5.21, page 181: Tvn  = vn 

Tvn = vn

Let W = Spanvn . Theorem 7.10, page 306 tells us that:  W = v  V vv n = 0 is a subspace of V of dimension k – 1 . Nothing that for any v  W :

Tv  vk ====v Tvk v vk vv k 0 we conclude that Tv  W for every v  W .    Let TW:W  W denote the restriction of T to W . The lin-

earity of T assures us that TW is linear. Moreover, since Tv  w = v Tw holds for every vw  V , it must cer- tainly hold for every vw  W . Invoking the induction hypothesis (II) to the symmetric linear operator   TW:W  W , we let v1v2 vk – 1 denote an orthonor- mal basis of eigenvectors of TW . Since each vector in that basis is orthogonal to the eigenvector vk , the set

vk v1v2 vk – 1 ------is seen to be an orthonormal basis for V vk consisting of eigenvectors.

CHECK YOUR UNDERSTANDING 7.26 Find an orthonormal basis of eigenvectors for the symmetric linear Answer: See page B-38. operator Tabc = ab– –2a + bc– – b + c of CYU 7.25.

MATRIX VERSION OF THE SPECTRAL THEOREM Here is the link between symmetric matrices and symmetric linear operators:

THEOREM 7.19 AM nn is symmetric if and only if n n TA:    given by TAX = AX is a symmetric linear operator. 7.4 The Spectral Theorem 321

n PROOF: Assume that AM nn is symmetric. For any XY   : Recall that for XY  n : TAX  Y ==TAX  Y AXY XY = XY Theorem 7.15: (see page 307) ==X  AY X  TAY =X TAY

Conversely, if TA is symmetric, then:

AXY ==TAX  Y TAX  Y

Definition 7.13: ===X TAY X  TAY X  AY

DEFINITION 7.14 AM nn is an orthogonal matrix if the ORTHOGONAL AND columns of A constitute an orthogonal set in ORTHONORMAL the Euclidean inner product space n . MATRICES A is an orthonormal matrix if its columns constitute an orthonormal set in n .

THEOREM 7.20 Every orthogonal matrix is invertible.

PROOF: If AM nn is orthogonal, then the columns of A constitute a linearly independent set of vectors in n (Theorem 7.7, page 301), and are therefore a basis for n (Theorem 3.11, page 99). The result now follows from Exercise 37, page 175. Yes, every orthogonal matrix is invertible; but more can be said for orthonormal matrices:

THEOREM 7.21 AM nn is orthonormal if and only if A–1 = AT

T T PROOF: Let Aa= ij , A = aij , and AA = cij . Then: n n

cij == aiaj  aiaj  = 1  = 1

the dot product of the ith column of A with the jth column of A It follows that A–1 = AT , if and only if AAT = I , if and only if n 1 if ij=  aiaj =   0 if ij  = 1 if and only if the columns of A constitute an orthonormal set in n .

CHECK YOUR UNDERSTANDING 7.27

Prove that the product of any two orthonormal matrices in Mnn is Answer: See page B-39. again orthonormal. 322 Chapter 7 Inner Product Spaces

Note: In the literature the DEFINITION 7.15 AM is if there exists an orthonor- term orthogonally diago- nn nalizable is typically ORTHONORMALLY mal matrix P and a diagonal matrix D such used to refer to what we DIAGONALIZABLE that: are calling . P–1AP= D As it turns out:

THEOREM 7.22 AM nn is orthonormally diagonalizable THE SPECTRAL if and only if it is symmetric. THEOREM n n PROOF: If AM nn is symmetric, then TA:    given by TAX = AX is a symmetric operator (Theorem 7.19). Employing Theorem 7.18, we chose an orthonormal basis  = X1X2  Xn of eigenvectors of TA with corresponding eigenvalues 12 n . n For S = e1e2  en the standard basis of  we have: See Theorem 5.26, page 193 TA  = I STA SSI S (*) We now show that: (1), (2), (3) and (*) tell us (1) T is a diagonal matrix with the  s along its diagonal. that: A  i P–1AP is a diagonal matrix, (2) TA SS = A with PI=  an orthonor- S (3)The columns of I are the X s — an orthonormal set. mal matrix. S i In particular: (1):A consequence of Definition 5.10, page 179, and the fact that A is ! TAXi ==AXi iXi . Moreover: P–1AP= PTAP (2):A consequence of Definition 5.10 and the fact that TAei S is th th with Xi the i column of P. the i column of A. (3):A consequence of Definition 5.10 and the fact that

IXi S ==Xi S Xi . Conversely, assume that A is orthogonally diagonalizable. Let P be an orthogonal matrix and D a diagonal matrix with: P–1AP= D APDP= –1 Theorem 7.21: APDP= T By Then: AT = PDPT T Exercise 24(f), page 1643: = PT TDTPT Exercise 24(a), page 164: ===PDTP–1 PDP–1 A Since every diagonal matrix is symmetric Since AT = A , A is symmetric. 7.4 The Spectral Theorem 323

CHECK YOUR UNDERSTANDING 7.28 Find an orthonormal diagonalization for the symmetric matrix: 2 1 1 121 Answer: See page B-40. 112 324 Chapter 7 Inner Product Spaces

EXERCISES

Exercises 1-4. Verify that the given matrix is orthonormal.

1 1 1 001 3 2 6 –------3------–------2 3. 100 7 7 7 1. 2 2. 2 2 1 1 010 4. 2 6 3 --1------3- –------–------2 2 2 2 7 7 7 6 3 --- –--2- --- 7 7 7

Exercises 5-8. Verify that the given matrix AM nn is symmetric. Show directly that Av  w = v  Aw for every vw  n .

5.51 6. 73– 010 21–0 15 –43 7.101 8. –311 012 012 Exercises 1-4. Find an orthonormal diagonalization for the symmetric matrix of: 9. Exercise 5. 10. Exercise 6. 11. Exercise 67 12. Exercise 8.

Exercises 9-12. Verify that the given linear operator T: n  n on the Euclidean (dot product) inner product space Rn is symmetric. Determine T where S denotes the standard basis in Sn Sn n n . 13.Tab = 2ab+  a + 2b 14.Tab = – a +33b a + 5b

15.Tabc = a ++2b 3c 2ab+  3a + 2c

16.Tabc = a ++22bc a + 2ba + 3c

17. Verify that Tab = 3a 2ab+ is a symmetric operator on the weighted inner product 1 space R2 with ab cd = 4ac+ bd . Verify that = ---0 01 is an orthonor-      2

mal basis in this inner product space, and determine T  . 18. (a) Verify that Tax2 ++bx c = a ++2bcx2 ++2ab x 3a + 2c is a symmetric oper- 2 2 ator on the standard inner product space P2 : ax ++bx c  ax ++bx c = aa+ bb + cc . (b) Use the Grahm-Schmidt process of page 303 on the basis  = x2 ++x 1 x + 1 x2 + 1 1 1 1 1 1 1 1 to arrive at the orthonormal basis  = ------x2 ++------x ------– x2 ++------– ---x + --- . Verify that 3 3 3 x 2 4 4 T is not symmetric, and that T is symmetric.   7.3 The Spectral Theorem 325

19. Let 2 denote the standard Euclidean dot product inner product space. Find a symmetric lin-

2 2 32 ear operator T:    and a basis  for which T  = . 21 20. Let 2 denote the weighted inner product space R2 with ab  cd = 3ac + 2bd . Find

2 2 32 a symmetric linear operator T:    and a basis  for which T  = . 21

21. Let P1 denote the standard inner product space P1 : ax+ b  ax+ b = aa+ bb + cc . 32 Find a symmetric linear operator T: P1  P1 and a basis  for which T  = 21

T T 22. Show that for any A  Mnn both AA+ and AA are symmetric.

23. Show that if AB  Mnn are orthonormally diagonalizable, then so is: (a) cA for every c   . (b) AB+ (c) A2

n 24. (PMI) Show that if A  Mmm is orthonormally diagonalizable, then so is A for any posi- tive integer n.

25. (PMI) Show that if Ai  Mmm is orthonormally diagonalizable for 1 in , then so is  A1 +++A2 An .

26. Show that if A  Mmm is an invertible orthonormally diagonalizable matrix, then so is A–1 . 27. Prove that if A is a real symmetric matrix, then the eigenvalues of A are real. Suggestion: For Av = v , show that Av = v , where here  denotes the (complex) conju- gate or  and v is the n-tuple obtained by taking the conjugate of each entry in the n-tuple v. Proceed to show that vTv = vTv .

PROVE OR GIVE A COUNTEREXAMPLE

T 28. If A  Mnn is a symmetric matrices, then so is A .

–1 29. If A  Mnn is a symmetric matrices, then so is A .

30. If AB  Mnn are symmetric matrices, then so is AB+ .

31. If AB  Mnn are symmetric matrices, then so is AB . 326 Chapter 7 Inner Product Spaces

32. If AB  Mnn are orthonormally diagonalizable, then so is AB .

T 33. If A  Mnn is orthonormally diagonalizable, then so is A .

–1 34. If A  Mnn is orthonormally diagonalizable, then so is A . 35. Let V be an inner product space. If T: VV is a symmetric operator, then so is cT for every c   . 36. Let V be an inner product space. If T: VV and L: VV are symmetric operators, then so is TL+ . 37. Let V be an inner product space. If T: VV and L: VV are symmetric operators, then so is LT . Chapter Summary 327

CHAPTER SUMMARY

OT RODUCT D P The dot product of u = u1u2  un and v = v1v2  vn , denoted by uv , is the real number:

uv = u1v1 +++u2v2  unvn

PROPERTIES Let uvw  n , and r   . Then: positive-definite property: (i)vv  0 , and vv = 0 only if v = 0

commutative property: (ii) uv = vu

homogeneous property: (iii) ruv = ruv

distributive property: (iv) uv+  w = uw + vw

n NORM IN  The norm of a vector v = v1v2  vn , denoted by v , is given by: vvv=  Denotes length of vector.

ANGLE BETWEEN The angle  between two nonzero vectors uv  n is given by: VECTORS –1 uv  = cos ------uv

ORTHOGONAL VECTORS Two vectors u and v in n are orthogonal if uv = 0 .

VECTOR Let v  n and let u be any nonzero vector in n . Then: DECOMPOSITION vv= – projuv + projuv where: vu proj v = ------u and v – proj v  proj v = 0 u uu u u

INNER PRODUCT SPACE An inner product on a vector space V is an operator which assigns to any two vectors, u and v in V, a real number uv , satisfying the following four axioms: positive-definite axiom: (i)vv  0 , and vv = 0 only if v = 0 commutative axiom: (ii) uv = vu

homogeneous axiom: (iii) ruv = r uv

distributive axiom: (iv) uv+  w = uw + vw 328 Chapter 7 Inner Product Spaces

PROPERTIES For every u, v, and w in an inner product space V: (a) 0 v ==v 0 0 (b) uv + w = uv + uw (c) u rv = r uv (d) uv–  w = uw – vw (e) uv – w = uv – uw (f) –v w ==vw – – vw

NORM AND The norm (or magnitude) of a vector v in an inner product space V, DISTANCE denoted by v , is given by: vvv=  The distance between two vectors u and v in V is given by uv– .

CAUCHY-SCHWARZ For any two vectors u and v in an inner product space: INEQUALITY uv  uv

PROPERTIES Let V be an inner product space. For all uv  V and r   : (a)v  0 , and v = 0 if and only if v = 0 (b) rv = r v (c) uv+  uv+

ANGLE BETWEEN The angle  between two nonzero vectors u and v in an inner prod- VECTORS uct space is given by: –1 uv  = cos ------uv

ORTHOGONAL VECTORS Two vectors u and v in an inner product space V are orthogonal if uv = 0 . ORTHOGONAL SET A set S of vectors in an inner product space V is an orthogonal set if uv = 0 for every uv  S , with uv .

THEOREM If v1v2  vn is an orthogonal set of non-zero vectors in an inner

product space V , then v1v2  vn is a linearly independent set in V.

UNIT VECTOR A unit vector in an inner product space is a vector v of magnitude 1.

ORTHONORMAL SET An orthonormal set of vectors in an inner product space is an orthog- onal set of unit vectors. Chapter Summary 329

THEOREM If  = v1v2  vn is an orthonormal basis in an inner product space V, then, for any v  V :

vvv=  1 v1 +++vv 2 v2  vv n vn

GRAHM-SCHMIDT An algorithm (page 303) for generating an orthogonal base in any finite PROCESS dimensional inner product space.

ORTHOGONAL The orthogonal complement of a subspace W of an inner product COMPLEMENT space: W = u  V uw = 0 for every w  W

PROPERTIES If W a subspace on an inner product space V, then: (i)W is a subspace of V. (ii) WW  = 0 (iii) Every vector in V can be uniquely expressed as a sum of a vec- tor in W and a vector in W .  (iv) If  is a basis for W and   is a basis for W , then     W W W W is a basis for V.

VECTOR Let w1w2  wm be an orthonormal basis for a subspace W of an DECOMPOSITION inner product space V, and let v  V . Then, there exists a unique vec- tor w  W and u  W such that: v vw= + u u where:

w = vw 1 w1 +++vw 2 w2  vw m wm w W and: u = vw– .

SYMMETRIC MATRIX T AM nn is symmetric if A = A .

THEOREMS If AM nn is a (real) symmetric matrix, then its eigenvalues are real.

Any two eigenvectors in the inner product space n corresponding to

distinct eigenvalues of a symmetric matrix AM nn are orthogo- nal.

AM nn is symmetric if and only if Av  w = v  Aw for all vectors vw  Rn (in column form). 330 Chapter 7 Inner Product Spaces

SYMMETRIC Let V be an inner product space. A linear operator T: VV is sym- OPERATOR metric if Tv  w = v Tw for all vectors vw  V .

THEOREMS Let T: VV be a symmetric linear operator on an inner product space V. If v1 and v2 are eigenvectors associated with distinct eigen- values 1 and 2 , then v1 and v2 are orthogonal. If T: VV is a symmetric linear operator on an inner product space

V, then T  is a symmetric matrix for any orthonormal basis

 = o1o2 on of V.

SPECTRAL THEOREM (Linear Operator) A linear operator T: VV on an inner product space V is symmetric if and only if V contains an orthonormal basis of eigenvectors of T.

(Matrix) AM nn is symmetric if and only if there exists a diago- nal matrix D and a matrix P with columns an orthonormal set in n such that P–1AP= D . PRINCIPLE OF MATHEMATICAL INDUCTION A-1

APPENDIX A PRINCIPLE OF MATHEMATICAL INDUCTION We introduce a most powerful mathematical tool, the Principle of Mathematical Induction (PMI). Here is how it works: (PMI) Let Pn denote a proposition that is either true or false, depend- ing on the value of the integer n. If: I.P1 is True. And if, from the assumption that: II. Pk is True one can show that: III.Pk+ 1 is also True. then the proposition Pn is valid for all integers n  1 Step II of the induction procedure may strike you as being a bit strange. After all, if one can assume that the proposition is valid at nk= , why not just assume that it is valid at nk= + 1 and be done with it? Well, you can assume whatever you want in Step II, but if the proposition is not valid for all n you simply are not going to be able to demonstrate, in Step III, that the proposition holds at the next value of n. Its sort of like the domino theory. Just imagine that the propositions P1  P2  P3 Pk Pk+ 1  are lined up, as if they were an infinite set of dominoes:

P(1 P(2 P P P(6) P(7) P(8) P(9) P(10) ) ) (3) (4) P(5) ......

If you knock over the first domino (Step I), and if when a domino falls (Step II) it knocks down the next one (Step III), then all of the domi- noes will surely fall. But if the falling kth domino fails to knock over the next one, then all the dominoes will not fall.

The Principle of Mathemati- To illustrate how the process works, we ask you to consider the sum cal Induction might have been of the first n odd integers, for n = 1 through n = 5 : better labeled a Principle of Mathematical Deduction; for: n Sum of the first n odd integers Sum n Sum Inductive reasoning is a pro- 11 1 1 cess used to formulate a 2 1 + 3 4 2 4 hypotheses or conjecture, 1 + 3 + 5 9 3 9 while deductive reasoning is 4 16 a process used to rigorously 1 + 3 + 5 + 7 4 16 establish whether or not the 1 + 3 + 5 + 7 + 9 25 5 25 conjecture is valid. 6 ? Figure 1.1 A-2 Principle of Mathematical Induction

Looking at the pattern of the table on the right in Figure 1.1, you can probably anticipate that the sum of the first 6 odd integers will turn out to be 62 = 36 , which is indeed the case. In general, the pattern cer- tainly suggests that the sum of the first n odd integers is n2 ; a fact that we now establish using the Principle of Mathematical Induction.

Let Pn be the proposition that the sum of the first n odd integers equals n2 . I. Since the sum of the first 1 odd integers is 12 , P1 is true. The sum of the first 3 odd  2 integers is: II. Assume Pk is true; that is: 135+++ +2k – 1 = k see margin 135++ 2  3 – 1 III. We show that Pk+ 1 is true, thereby completing the proof: The sum of the first 4 odd integers is: the sum of the first k + 1 odd integers

1357+++ 2  4 – 1  2 2 Suggesting that the sum of 135+++ +2k – 1 + 2k + 1 ==k + 2k + 1 k + 1 the first k odd integers is: induction hypothesis: Step II 13++ +2k – 1 (see Exercise 1).

EXAMPLE 1.1 Use the Principle of Mathematical Induction to establish the following formula for the sum of the first n integers: nn+ 1 123+++ +n = ------(*) 2

SOLUTION: Let Pn be the proposition: nn+ 1 123+++ +n = ------ 2 11+ 1 I.P1 is true: 1 =------ Check! 2 kk+ 1 II. Assume Pk is true: 123+++ +k = ------2 III. We are to show that Pk+ 1 is true; which is to say, that (*) holds when nk= + 1 : k + 1 k + 1 + 1 k + 1 k + 2 123+++ ++kk+ 1 ==------------------- 2 2 Let’s do it: 123+++ ++kk+ 1 = 123+++ +k + k + 1 kk+ 1 induction hypothesis: = ------+ k + 1 2 ==k------k + 1------+ 2k------+ 1------k + 1 ------k + 2 2 2

PRINCIPLE OF MATHEMATICAL INDUCTION A-3

The “domino effect” of the Principle of Mathematical Induction need not start by knocking down the first domino P1 . Consider the fol- lowing example where domino P0 is the first to fall. EXAMPLE 1.2 Use the Principle of Mathematical Induction to establish the inequality n  2n for all n  0 .

SOLUTION: Let Pn be the proposition n  2n . I.P0 is true: 02 0 , since 20 = 1 . II. Assume Pk is true: k  2k III: We need to show that III. We show Pk+ 1 is true: n n  2 holds for k +221  2k + 12 k + 2k ==k 2k + 1 nk= + 1 ; which is to II k say, that: k + 12 k + 1 : 12

Recall that:. EXAMPLE 1.3 Use the Principle of Mathematical Induction to 2 n!12=  n show that n!  n for all integers n  4 .

SOLUTION: Let Pn be the proposition n!  n2 :

I.P4 is true: 4!1234== 24 42 . II. Assume Pk is true: k!  k2 (for k  4 ) III. We show Pk+ 1 is true; namely, that k + 1 !  k + 1 2 : k + 1 ! = k!k + 1  k2k + 1 II Now what? Well, if we can show that k2k + 1  k + 1 2 , then we will be done. Let’s do it: Since k  4 (all we need here is that k  2 ): k2  k + 1 Multiplying both sides by the positive number k + 1 : k2k + 1  k + 1 2 . A-4 Principle of Mathematical Induction

Our next application of the Principle of Mathematical Induction involves the following Tower of Hanoi puzzle: Start with a number of washers of differing sizes on spindle A, as is depicted below:

A B C The objective of the game is to transfer the arrangement cur- rently on spindle A to one of the other two spindles. The rules are that you may only move one washer at a time, without ever placing a larger disk on top of a smaller one. EXAMPLE 1.4 Show that the tower of Hanoi game is winna- ble for any number n of washers.

SOLUTION: If spindle A contains one washer, then simply move that washer to spindle B to win the game (Step I). Assume that the game can be won if spindle A contains k washers (Step II —the induction hypothesis). We now show that the game can be won if spindle A contains k + 1 washers (Step III): Just imagine that the largest bottom washer is part of k + 1 wahsers the base of spindle A. With this sleight of hand, we are looking at a situation consisting of k washers on a modified spindle A (see margin). By the induction . . k washers hypothesis, we can move those k washers onto spindle } B. We now take the only washer remaining on spindle { new

base A (the largest of the original k + 1 washers), and move it to spindle C, and then think of it as being part of the base of that spindle. Applying the induction hypotheses one more time, we move the k washers from spindle B onto the modified spindle C, thereby winning the game.

CYU SOLUTIONS B-1

APPENDIX B CHECK YOUR UNDERSTANDING SOLUTIONS CHAPTER 1 MATRICES AND SYSTEMS OF LINEAR EQUATIONS

9 10713 100--- 1 10 7 13 5 10 7 13 ------R  R 3R + R  R 4 15 1 3 01–4 3– 3 2 2 010 --- –7R3 + R1  R1 4 CYU 1.1 01–4 3– 5 010--- 8 5 0015 24 00 1 --- 8 5 001 --- 8 5 001--- 5

CYU 1.2 (a) Yes (b) No [fails (ii)] (c) Yes (d) Yes

x y z x y z xyz++= 6 11 1 6 1001 x = 1 CYU 1.3  rref 3x +42yz– =   32–4 1 0102  y = 2  3xy++2z = 11 31211 0013 z = 3

CYU 1.4 (a) Inconsistent: The last row 0002 corresponds to the equation 0x ++0y 0z = 2 which clearly has no solution. free variable x y z (b) 10– 2 1 x +10y – 2z =  x = 12+ z  01 5 4    : 12+ r 45– rr r   0xy++5z = 4 y = 45– z 00 0 0  

free variables

1 2 01 1 x +++2y 0zw= 1 x = 12– y – w  0 014 –2   : 0x +++0yz4w = – 2 z = –42 – w  (c) 0000 0 12– r – sr –42 – ssrs   R B-2 CYU SOLUTIONS

CYU 1.5

7 4 7 4 100a + ------b – ---c xa= + ------b – ---c (a) 4x – 2y + z = a  18 9 18 9  42–1a –42x ++y 2z = b  rref 11 5  –422 b 010 a + ------– ------: ya= + -----11- – -----5-  18 9c 18 9c 5xy–4+ z = c  51–4c 1 2 1 2 001– a – ---b + ---c za= – – ---b + ---c 3 3 3 3 Solution for all a, b, and c

10– 5a + ------b – 2a- 4 x –44y – z = a  14–4– a  after two cycles b – 2a (b) 2x + 8y – 12z = b   28– 12b 01–--1------ 4 16 –12x ++y 2z = c  –1221 c 00 0 c ++3ab The system is consistent if and only if c ++3ab= 0

3x + 7yz– = a  37– 1 100  rref coefS CYU 1.6 (a) S: 13x –24y + z = b  coef(S) 13–2 4 010  2x –24y + z = c  24–2 001

does not contain a row of zeros: system has a solution for all values of a, b, and c

x – 3y + w = a   13–01 100 10 3xy–2+ z – 3w = b coef(S) 31– 23– rref coefS 010 3 (b) S:  xz+ – 5w = c  101 –5 001–15  2xy–3+ z – 2w = d  21– 1 2 000 0 contain a row of zeros: system does not have a solution for all values of a, b, c, and d CYU SOLUTIONS B-3

CYU 1.7 x y z w x y z w 1 0 0 –2 2x +++3y 4z 5w = 0  234 5 coef S rref 0 1 0 1 S : 3xy++4zw += 0  314 1 3  0 0 1 --- x +++7y 4z 11w = 0 17411 2 Setting the free varialble w to r we arrive at the system: x –02r =  yr+0=  3  Solution set:  2rr – –---r r rR 3 2 z +0---r =   2  = 4r –2r–23r r rR

CHAPTER 2 VECTOR SPACES

CYU 2.1 rv +2sw ===32– 2 + –3 –310 64– 4 + 93– 0 15 1– 4

CYU 2.2 (iv): vv+ – = 0 (in 2 ): PofR

If vv= 1 v2 , then: vv+ – v1 v2 + –v1 –v2 v1 – v1 v2 – v2 = 00  0 Definition 2.5 Definition 2.3 Definition 2.4 n (in  ): If vv= 1v2  vn , then:

vv+ – v1v2  vn + –v1–v2  –vn v1 – v1v2 – v2  vn – vn = 000  0

CYU 2.3 For A = aij  Mmn and rs   :

rsA rsaij rsaij rsaij = rs aij rs aij rs A

n n n n n n i i i i i i CYU 2.4  aix +  bix   ai + bi x =  bi + ai x   bix +  aix i = 0 i = 0 i = 0 i = 0 i = 0 i = 0

CYU 2.5 rs+ f x  rs+ fx = rfx+ sfx  rf x + sf x B-4 CYU SOLUTIONS

CYU 2.6 The Zero Axiom: We need to find 0 = abc such that for any xyz  V :

xyz + abc = xyz , which is to say: xa+ – 1 = x a = 1 xa+ – 1 yb++2 zc+ – 3 = xyz  yb++2 = y  b = –2 zc+ – 3 = z c = 3 Let’s verify directly that 12– 3 + xyz does indeed equal xyz for every xyz  V : 12– 3 + xyz ==1 +2x – 1– ++y 2 3+ z – 3 xyz . The Inverse Axiom: For given xyz  V we are to find abc such that xyz + abc = 0 , which is to say: xa+1– 1 = ax= –2+ xa+ – 1 yb++2 zc+ – 3 = 12– 3  yb++2 = – 2 by= –4+ zc+3– 3 = cz= –6+ It is easy to verify, directly that: xyz++ + – x + 2– y + 4 – z + 6 = 12– 3 .

CYU 2.7 Since V = 0 must be closed under addition and scalar multiplication we have no choice but to define: 00+ = 0 and r00= for every r   . It is easy to see that all eight axioms of Defi- nition 2.6 hold. We establish (v): If uv  V , then they must both be 0 . Consequently: ruv+ r00+  r0 0 and ru + rv  r0 + r0 000+  . Hence: ruv+ = ru + rv . As for (iii) and (iv), simply note that 0 is certainly the zero vector in V and that it is also its own inverse.

CYU 2.8 vz+ = wz+ CYU 2.9 Start with: r0 = r00+ vz+ + –z = wz+ + –z r0 = r0 + r0 vz+ + – z= wz+ – z r0 + –r0 = r0 + r0 + –r0 v + 0 = w + 0 0 = r0 + r0 + –r0 vw= 0 = r00+ Conclusion:0 = r0 CYU 2.10 (a) rv = rw (b) rv = sv 1 1 rv + –sv = sv + –sv ---rv = ---rw r r rv + –sv = 0 1 1 rs– v = 0 ---r v = ---r w r r Theorem 2.8: rs–0= 1v = 1w rs= Axiom (viii): vw= CYU SOLUTIONS B-5

CYU 2.11 (a) –1–v ====– –1 v –1 –1 v 1vv (b) –r v ===–1 r v –1 rv –rv (c) r–v ====r–1 v r–1 v –1 rv –rv

CYU 2.12 (a) –1vw+ ==– vw+ –1 v + –1 w ==– v + –w – v – w

(b) I. Valid at n = 2 : –v1 + v2 = – v1 – v2 [by part (a)].   II. Assume –v1 ++v2 +vk = – v1 – v2 – – vk . (The induction hypothesis)   III. We show that –v1 ++v2 ++vk vk + 1 = – v1 – v2 – – vk – vk + 1 :   –v1 ++v2 ++vk vk + 1 = –v1 ++v2 +vk + vk + 1  By I: = – v1 ++v2 +vk – vk + 1   By II: ==– v1 – v2 – – vn – vk + 1 – v1 – v2 – – vk – vk + 1

CYU 2.13 Since 00  S , S   . S is closed under addition: 00

For any a 2a  b 2b  S: a 2a + b 2b = ab+ 2ab+  S –0a –0b –0a –0b –0ab+ S is closed under scalar multiplication:

For any a 2a  S and r  : r a 2a = ra 2ra  S –0a –0a –0ra

CYU 2.14 Since 000  S , S   . For any xyz  abc  S and r   : rxyz + abc = rx+ a ry+ b rz+ c is back in S, since:

rx+ a ++ry+ b rz+ c ===rxyz++ + abc r00+ 0

CYU 2.15 Since the zero of that vector space is the zero function Z:  which maps every number to zero, and since S = f  F f9 = 0 , S does not contain the zero vec- tor and is therefor not a subspace of F .

CYU 2.16 Since 000  S , S   . For any 16a – 2b 4a – 17b11a 11b  16x – 2y 4x – 17y11x 11y  S and r   : r16a – 2b 4a – 17b11a 11b + 16x – 2y 4x – 17y11x 11y = 16ra+ x – 2rb+ y  4ra+ x – 17rb+ y  11ra+ x  11ra+ x ===16A – 2B 4A – 17B 11A 11B , where Arax+ and B ra+ x B-6 CYU SOLUTIONS

CYU 2.17 False. The subsets S = 01 and T = 02 are not subspaces of  since neither is close under addition (nor under scalar multiplication), yet ST = 0 is a subspace of  .

CYU 2.18 (a) For v ==2135–  – 12 – and u = 23 : u + rv r   ==23 + r12 – 2 +32r – r

(b) Since 1 +52r – r = 2 +32r – 1  – r – 1 : 1 +52r – r r   = 2 +32r – r r  

CYU 2.19 Choosing r = –1 and r = 1 in L = 135 + r21– 1 r   we obtain the two points 135 –121– 1 = –26 , 135 +21– 1 =  344 on L. Preceding as in Example 2.14 we arrive at a direction vector v ==344 – –126 42– 2 . Select- ing u = 344 as our translation vector, we have: L ==344 + r42– 2 r   34+42r+42r – r r   , which we now show to be equal to the set L ==135 + r21– 1 r   12+3r+5r – r r   : r – 1 34+42r+42r – r ==12+3r+5r – r , were r ------2

CYU 2.20 Choosing 41– 3 to play the role of w, instead of 32–2 we obtain:

P = w ++ru sv rs   = 41– 5 ++r–175– s13– 5 rs   = 4 – r +5s 17++r 3s – – 5r – 5s rs  

which we show to be equal to the set P = w ++ru sv rs   : = 32– 2 ++r–175– s13– 5 rs   = 3 – r + s – 2 ++7r 3s 25– r – 5s rs  

1 : 4 – r +3s = – r + s Equating the first two components of P and P we obtain: . 2 : 1++ 7r 3s = –7 2++r 3s Multiplying equation (1) by –3 and adding it to equation (2) we find that rr= . Substituting in (1) we then find that ss= – 1 ; bringing us to: 4 – r +5s 17++r 3s – – 5r – 5s = 3 – r + s – 2 ++7r 3s 25– r – 5s CYU SOLUTIONS B-7

CHAPTER 3 BASES AND DIMENSIONS

a b a b a +22b = –  CYU 3.1 (a) No: 12 – 2 augS rref 10 0 S: 3a +35b = –  35– 3  01 0 8a +84b = No solution! 848 00 1

a b a b (b) Yes: a +22b = –  12 – 2 augS rref 10 2 S: 3a +45b = –  35– 4  01– 2 8a +84b =  848 000

–2 –4 – 8 = 2138 – 2254

CYU 3.2 (a) We are to show that for any given matrix ab there exist scalars xyzw for which cd

(*) x 12 +++y 10 z 01 w 04 = ab: 34 10 01 20 cd

xy+2xz++4w = ab 3xy++2w 4xz+ cd

xy++0z +0w = a   1100 1000 2x +++0yz4w = b coef(S)2014 [rref(coef(S)] 0100 S:  3xy++0z +2w = c  3102 0010  4x +++0yz0w = d  4010 0001 Since rref coefS does not contain a row consisting entirely of zeros, the system S [which stems from (*)] has a solution for any given abcd .

xy++0z +0w = – 1 (b)  1100– 1 1000 2 2x +++0yz4w = 5 aug(S) 2014 5 rref 0100–3 S:  3xy++0z +2w = 1 3102 1 0010 5  4x +++0yz0w = 13 401013 0001–1

from the above we see that: 2 12–3 10 + 5 01–1 04 = –51 34 10 01 20 113 B-8 CYU SOLUTIONS

CYU 3.3 We are to find the set of vectors abc for which there exist scalars x, y, z, such that: abc = x215 ++y12– 2 z051 :

12–5 b  2xy++0z = a 210a 12–5b 12–5 b a – 2b  01– 2 ------x –52y + z = b   12–5b 210a 05– 10a – 2b  5  5x ++2yz= c  521c 521c 012– 24c – 5b 00 0c – 5b – 12------a – 2b 5 Note that in the above we didn’t bother to reduce the coefficient matrix to its row-reduced-echelon forms with 0’s above and below leading ones. Rather, we obtained its row-echelon form with 0’s only below the leading ones [see Exercises18-22, page 12]. This still enables us to determine Span215 12– 2 051 , for it consists of all vectors abc for which 12 a – 2b c – 5b – ------= 0 , which is equivalent to: 12ab+0– 5c = or b = 5c – 12a . 5 Conclusion: Span215 12– 2 051 = abc b = 5c – 12a . Any vector abc for which b  5c – 12a , say (1,2,3), will not be in the spanning set.

CYU 3.4 For given vV we are to find scalars A,B,C such that vAv= 1 ++Bv1 + v2 Cv1 ++v2 v3 . Since v1v2 v3 span V, there exist scalars a,b,c such that vav= 1 ++bv2 cv3 . Equating these two expressions for v we have: Av1 ++Bv1 + v2 Cv1 ++v2 v3 = av1 ++bv2 cv3

ABC++ v1 ++BC+ v3 Cv3 = av1 ++bv2 cv3

ABC++ = a  Cc=  Bringing us to: BC+ = b  with solution: BbC==– bc–  Cc=  AaB==– – C abc– – – c =ab–

CYU 3.5 ax2 ++b2x2 + x cx– 3 = 0 a ++2b 0c = 0 12 0 100 2  rref a + 2b x + bc+ x – 3c = 0 0abc++= 0  01 1 010  0a +00b – 3c =  00– 3 001

No free variable Linearly independent

CYU 3.6 av1 ++bv1 + v2 cv1 ++v2 v3 +dv1 +++v2 v3 v4 = 0 (we are to show abcd====0  abcd+++ v1 +++bcd++ v2 cd+ v3 dv4 = 0

Since  v 1  v 2 v 3 v 4 is abcd+++= 0 linearly independent:  bcd++= 0   a ====bcd0 cd+0=   d = 0  CYU SOLUTIONS B-9

CYU 3.7 Aa1a2 a3 a4 ++++Bb1b2 b3 b4 Cc1c2 c3 c4 Dd1d2 d3 d4 Ee1e2 e3 e4 = 0

a1 b1 c1 d1 e1

a2 b2 c2 d2 e2 rref must have a free variable (5 variables and 4 equations ) a3 b3 c3 d3 e3

a4 b4 c4 d4 e4

2511 rref 103 CYU 3.8 Linear dependent since: 10 3 011 . 3211 000

2 5 11 8 1034 8412  Span213 502 11311 since: 10 3 4 rref 0110 . 3 2 11 12 0000

From the above rref-matrix we see that a213 ++b502 c11311 = 8412 for any a +43c = a,b, and c for which . Letting c = 0 we get a = 4 and b = 0 , giving us the linear bc+0= combination 8412 = 4213 ++0502 011311 . Letting c = 1 we get a = 1 and b = –1 , giving us another linear combination 8412 = 213 –502 + 11311 .

CYU 3.9 x3 + x and –7 are clearly linearly independent. Since x2 cannot be “built” from those vectors, x3 +7x –  x2 is linearly independent. Can x3 be “built” by those three vectors? No; so x3 +7x–  x2 x3 is linearly independent. Just to make sure: a b c d 1001 1000 ax3 + x +++b–7 cx2 dx3 = 0 0010 rref 0100 3 2 ad+ x +++cx ax b–7 = 0 1000 0010 07–00 0001

3 (Incidentally, if you throw in any two randomly chosen vectors from P3 into x +7x – chances are really good that you will end up with a linearly independent set. Try it.) B-10 CYU SOLUTIONS

CYU 3.10 For spanning: x 21 ++y 11 z –03 +w 04 = ab 30 22 –55 – 15 cd

21 11 –03 04 00 For linear independence: x ++y z +w = 30 22 –55 – 15 00

21–0 3 1000 Spans: Does not contain a row consisting entirely of zeros. coefficient matrix 1104 rref 0100 32–1 5 0010 Linearly Independent: each row 02–5 5 0001 has a leading one.

CYU 3.11 If S = v1v2 vn is a basis, then it spans V insuring us that every vector in V can be expressed as a linear combination of the vectors in S. Being a basis, S is also linearly independent, and Theorem 3.6, page 89 insures us that the representation is unique. Conversely, if every vector in V can uniquely be expressed as a linear combination of the vectors in S then S certainly spans V. To show that it is also linearly independent consider c1v1 +++c2v2  cnvn = 0. Since 0v1 +++0v2  0vn = 0 , and since we have unique repre- sentation: ci = 0 for 1 in.

CYU 3.12 We show that S = 10  01 is a basis for the space V of Example 2.5. S spans V: For ab  V can we find rs   such that ab = r10 + s01 ? Yes: ab = r10 + s01 Since rxy = rx–1 r –  ry+ r – 1 : = rr–1+  r – 1 + –1s +  ss+ – 1 = 1 r – 1 + –1s +2 s – 1 Since xy + x y = xx+  – 1 yy++ 1 : = –1s +  r –21 ++s –11 = –1s +2 sr+ – 1 equating coefficients as= –1+    s ==–a + 1 and r b + 2a – 1 b = 2sr+ – 1  Check: r10 + s01 = b + 2a – 1 10 + –1a + 01 = b –12a – –1b + 2a – 1 +  b + 2a – 1 – 1 + –1–1a + +  –1a + – 1 ==1 b + 2a – 2 +1a –12a + + a – 1 b + 2a – 2 + a –12a + – 1 =ab S is linearly independent: Recalling that 11 – is the zero vector in S we start with the equation a10 +11b01 =  – and go on to show that ab==0 : a10 +11b01 =  – 1 a – 1 +11–1b +  bb+ – 1 =  – 1 a – 1 +11–1b +2 b – 1 =  – 1 –1b =   solution: ab==0 1 –11b + –  a –21 ++b –11 = 11 – a –21 +1b = –  1 – b a –21 + b = 11 – CYU SOLUTIONS B-11

CYU 3.13 (a) Knowing that M22 has dimension 4, we simply have to add two vectors to 21 22 L =  without rupturing linear independence. By now, you may be convinced that 12 11 if you add any two randomly chosen vectors, say 15 and 01 , chances are good that you –11 –62 will end up with a set of four independent vectors, and therefore a basis for M22 . Let’s make sure that we do: a b c d 22 1 0 1000 21 22 15 01 00 rref a ++b c +d = 12 5 1 0100 12 11 –11 –62 00 11–2 1– 0010 21 1 6 0001

(b) Here is a brute force approach to obtain a basis for Span(S). We start with the first vector in S = 312 –  –93– 6 12– 2 –54– 6 624 –  : 31– 2 . Since the second vector is easily seen to be a multiple of the first, we discard it and turn our attention to the third vector 12– 2 . Since that vector is not a multiple of 31– 2 , we throw it into that set to obtain the two independent vectors 31– 2  12– 2 . Can the vector –54– 6 be built from those two independent vectors? Yes:

31– 5 10– 2 –241 rref 01 1  –12312 –  + 12– 2 = –54– 6 22–6– 00 0 Since we are looking for a maximal independent set, we discard –54– 6 and turn our attention to the last remaining vector 624 –  . Is it independent of the vectors 31– 2  12– 2 ? Clearly not, since 624 –  = 2312 –  . Conclusion: 31–2 12– 2 is a basis for Span(S). Since 3 is of dimension 3, S does not span 3 .

CYU 3.14: first vector

39–15–6 13–02–2 –32421 – rref 00110 Basis for SpanS = 31–2 12– 2 26–2–6–4 00000 third vector B-12 CYU SOLUTIONS

CHAPTER 4 LINEARITY CYU 4.1 The function f: 2  3 given by fab = ab+  2ba – b is linear. f preserves sums: fab + a b = faa++ b b = aa+  +2bb+  bb+  aa+  – bb+  ==ab+2 ba – b + a+2b b a – b fab + fa b

f preserves scalar products: frab ==fra rb ra+ rb 2rb ra– rb ==rab+2 ba – b rfab

CYU 4.2 Since f00 ==0x2 ++0x 1 1 0 , f is not linear.

CYU 4.3 frab + a b = fra++ a rb b = ra+ a +2rb+ b rb+ b ra+ a – rb+ b = ra+ rb +2a + b  rb + 2b ra– rb + a – b = ra+2 rbrb ra– rb + a +2bb a – b ==rab+2 ba – b + a +2bb a – b rfab + fa b

CUU 4.4 A counterexample: The trivial function T:  given by Tx= 0 for every x   is linear. The set S = + is not a subspace of  , but fS= 0 is a subspace of  .

CUU 4.5 True: The proof of Theorem 4.5 makes no mention of linear independence.

CUU 4.6 (a) We first express 342 as a linear combination of the basis 100 020 111 :

ac+3=  342 = a100 ++b020 c111  = ac+2bc+ c 2bc+4=   c ===2 b 1 a 1  c = 2 

Then: T342 = T100 ++020 2111 By linearity: = T100 ++T020 2T111 ==2x2 + x ++2x2 + x 2x – 5 4x2 + 4x – 10 (b) Expressing abc as a linear combination of the given basis we have: AC+ = a  abc = A100 ++B020 C111  bc– C ==cB ------A =ac– = AC+2BC+ C 2BC+ = b      2 Cc=  CYU SOLUTIONS B-13

bc– Then: Tabc = Tac– 100 ++------020 c111 2 bc– = ac– T100 ++------T020 cT111 2 bc– = ac– 2x2 + x ++------2x2 + x cx– 5 2 b c = 2ab+ – 3c x2 + a + --- – --- x – 5c 2 2

CYU 4.7 (a) 10  11 is easily seen to be linearly independent, and therefore a basis of 2 . 010 110 011 is easily seen to be linearly independent and therefore a basis of 3 . (b) From the given information, we have: LT 10 = LT10 ==L020 L2010 =2L010 ==201 02 To determine LT 11 ==LT11 L101 , we need to express 101 as a linear combination of the basis 010 110 011 . Let’s do it:

b = 1  101 = a010 ++b110 c011  abc++= 0  c ===1 b 1 a –2 = babc++ c  c = 1 

Then: LT 11 ==LT11 L101 = L– 2010 ++110 011 = – 2L010 ++L110 L011 = –201 ++10 10 = 22 – To determine LT ab , we first express ab as a linear combination of 10  11 : ab ====A10 + B11 ABB+   B b and Aab– ; so: ab = ab– 10 + b11 Putting this together we find that: LT ab = LT ab– 10 + b11 = ab– LT 10 + bLT 10 ==ab– 02 + b22 – 2b 2a – 4b

CYU 4.8 (a):

Trax2 ++bx c + a'x2 ++b'x c' = Tra+ a' x2 ++rb + b' xrc+ c'

==ra + a' rb + b' r ab + a' b' rc + c' ra + a' ca c' a' = rT ax2 ++bx c + Ta'x2 ++b'x c' B-14 CYU SOLUTIONS

(b) The vectors Tx2 = 10 , Tx= 01 , T1 = 00 span ImT (Theorem 4.9), and 01 00 10 they are easily seen to be linearly independent. Consequently: rankT = 3 .

KerT = ax2 ++bx c Tax2 ++bx c = 0 By definition, 2 ab 00 ====ax ++bx c =  a bc0 ca 00 Consequently, KerT = 0 and therefore nullityT = 0

CYU 4.9 We first show that KerT = 0 :

2a = 0   Tabc ==02 ab + ccb 0000 bc+0=    a ===bc0 c = 0   b = 0 

At this point we know that nullityT = 0 and that the kernel has no basis. Since rankT +30 = , Im(T) has dimension 3. It is easy to see that the three vectors T100 ===2000  T010 0101  T001 0110 in the image of T are linearly independent. It follows that 2000 0101 0110 is a basis for Im(T).

CYU 4.10 Let v  V be such that Tv ==Tv  v v (*) . We establish that T is one -to-one by showing that KerT = 0 [see Theorem 4.11(a)]: Assume that Tv = 0 (we want to show that v = 0 ). Consider the vector v + v . By linearity we have: Tvv+ ===Tv+ Tv 0 + Tv Tv . From (*) and the fact that Tvv+ = Tv we have: vv+ = v v = 0 CYU 4.11 f –1: YX is one-to-one: –1 –1 –1 –1 –1 –1 f a ==f b  ffa ffb  ff a ==ff ba b –1 –1 f –1: YX is onto: For any xX , f fx ==f f xx .

CYU 4.12 T is one-to-one: Tab ==Ta' b'  ab+ xa– a' + b' xa– ' aa= '  Equating coefficients:   aa= ' and b = b' ab+ = a' + b'  CYU SOLUTIONS B-15

2 T is onto: For given ax+ b P1 , we need to find AB  such that: TAB ==AB+ xA– ax+ b

Equating coefficients: –A = b    A ===–b and B aA– ab+ AB+ = a 

Check: T–b ab+ ==– b ++abxb– – ax+ b.

Determining T –1 : Let T –1ax+ b = AB . Then: Applying T to both sides: ax+ b ==TAB  ax+ b AB+ xA– –A = b  Equating coefficients:   A ===–b and B aA– ab+ AB+ = a 

–1 2 From the above: T ax+ b = –b ab+ . Let’s show that the function L: P1   given by Lax+ b = –b ab+ is linear: Lrax+ b + a'xb+ ' = Lraa+ ' xrbb+ + ' = –rb+ b'  ra+ a' + rb+ b' ==r–ba + b + –b' a' + b' rL ax+ b + La'xb+ '

4 ab CYU 4.13 We show that T:   M22 given by Tabcd = is an isomorphism: cd T is one-to-one:

Tabcd ==Ta'b' c' d'  ab a' b'  a ====a' b b' c c' d d' cd c' d'

ab ab T is onto: For given  M22 , Tabcd = . cd cd

ab ab ra+ a rb+ b ab ab T is linear: Tr+ ==T rT + T . cd cd rc+ c rd+ d cd cd

CYU 4.14 Let dimV ==dimW n . By Theorem 4.15, V  n and W  n . By Theorem 4.14, VW .

Conversely, suppose that VW . Let T: VW be an isomorphism, and let V = v1v2  vn be a basis for V. We show STv= 1 Tv2 Tvn is a basis for W, thereby establishing that V and W are of the same dimension, n. B-16 CYU SOLUTIONS

S is linearly independent: since  is a linearly independent set n n n V a Tv = 0 Tav = 0 a v ==0 a 0 for 1 in  i i   i i   i i  i  i = 1 i = 1 i = 1 since T is linear Theorem 4.11(a), page 129 n S spans W: Let wW . Since T is onto, there exist va=  ivi such that Tv= w . Then: i = 1

n n wTv== Tav =a Tv  i i  i i i = 1 i = 1 since T is linear

CYU 4.15 By CYU 4.15, we know that the dimension of W equals that of V. We can therefore verify that Lv1 Lv2 Lvn is a basis for W by showing that the n vectors Lv1 Lv2 Lvn span W (see Theorem 3.11, page 99): n

Let w  W . Since L is onto, there exist va=  ivi such that Lv= w . Since L is n i = 1 linear: wLv== a .  i Lvi i = 1

CYU 4.16 Lets move the elements of: S = 2x3 –53x2 + x – 1x3 –8x2 + x – 3 x2 + 11x – 5 –2x3 ++x2 3x – 2

over to 4 via the isomorphism Tax3 +++bx2 cx d = abcd to arrive at the 4-tuples: TS= 23– 51– 1183 – – 0 1 11 – 5 –123– 2 Applying Theorem 3.13, page 103, we conclude that the first

two vectors of TS , 23– 51– and 1183 – – , consti- 210– 1 10–1 1– tute a basis for Span TS . It follows that –13 –12 rref 01 2 1 2x3 –53x2 + x – 1 and x3 –8x2 + x – 3 is a basis for Span(S). 58113 00 0 0 –31 –5–2– 00 0 0

CYU 4.17 (a) fxyz = 2x + 1 xyz+  – is one-to-one:

fx1y1 z1 = fx2y2 z2

2x1 +21 = x2 + 1 x1 = x2   2x1 + 1 x1 + y1 –z1 = 2x2 + 1 x2 + y2 –z2  x1 + y1 = x2 + y2  y1 = y2   –z1 = –z2 z1 = z2 CYU SOLUTIONS B-17 f is onto: For given xyz  X we need to find abc  3 such that: fabc = xyz

a = ------x – 1 2a + 1 = x  2   2a + 1ab+ –c = xyz  ab+ = y bya==– y – x------– 1   2 –c = z  cz= – x – 1 x – 1 From the above formula: f------y – ------–z = xyz , we conclude that: 2 2

–1 x – 1 x – 1 f xyz = ------y – ------–z 2 2

–1 x – 1 x – 1 (b) From Theorem 4.16 and the above formula f xyz = ------y – ------–z we have: 2 2 x – 1 x – 1 x – 1 x – 1 x y z  x y z = f ------1 -y – ------1 - –z + ------2 -y – ------2 - –z 1 1 1 2 2 2 2 1 2 1 2 2 2 2 x x x x = f ----1- + ----2-1– y + y – ----1- – ----2-1+ – z – z 2 2 1 2 2 2 1 2 x x x x x x Since fxyz = 2xx + y – z: = 2----1- + ----2-1– ----1- +++----2-1– y y – ----1- –1----2- +–– z – z 2 2 2 2 1 2 2 2 1 2

= x1 + x2 – 2y1 + y2 z1 + z2

x – 1 x – 1 rx– r rx– r and: r  xyz ==fr------y – ------–z f ------ry – ------–rz 2 2 2 2 rx– r rx– r rx– r = 2  ------------+ ry – ------––rz 2 Since fxyz = 2xx + y – z: 2 2 = rx– rry rz (c) The zero in the space X: f000 = 100 . The inverse of xyz in X:

–1 x – 1 x – 1 –1x + x – 1 ff– xyz ===f –------y – ------–z f------– y + ------z –2x +  –yz 2 2 2 2

–1 x – 1 x – 1 f xyz = ------y – ------–z fxyz = 2x + 1 xyz+  – 2 2 B-18 CYU SOLUTIONS

CHAPTER 5 MATRICES AND LINEAR MAPS

35 36 +3453  + 55 33 37 CYU 5.1 (a) 64 42 ==46 +4423  + 25 30 26 35 90 96 +9403  + 03 54 36 (b) Number of columns in A does not equal the number of rows in B.

2 CYU 5.2 False: 11 12 11 12 11 12 23 23 77 + ===+ + 11 00 11 00 11 00 11 11 34

11 11 11 12 12 12 22 24 12 58 While: ++2 = + + = 11 11 11 00 00 00 22 24 00 46

The columns associated with the leading 13 25–4 3 10–12------ones in rref(A); namely: 2 CYU 5.3 –104 –68– rref 240 –  and  5–1 10 01 2– 4 constitute a basis for the column space 012– 4 00 0 0 of A. 13 Since A and rref(A) share a row space, 10–12------ and  0 1 2– 4 constitutes a basis for the 2 row space of A. Note that the first two rows of A are not linearly independent. We are assured, however, that two of A’s rows will constitute a basis for its row space (either rows 1 and 3, or rows 2 and 3, will do the trick).

CYU 5.4 If X1 and X2 are solutions of the homogeneous system of m equations in n unknowns of equations, AX = 0 , and if r   , then:

ArX1 + X2 ===rAX1 + AX2 r00+ 0 It follows, from Theorem2.13, page 61, that the solutions set of AX = 0 is a subspace of n .

 21 3 0 100– 1 21 3 0 CYU 5.5 From 14–7 2– rref 010– 1 , we see that: null14–7 2– = cc– cc c    30 1– 2 001 1 30 1– 2 A basis:  11– 11

CYU 5.6 Let AM mn . By definition:

nullityA ==dimXAX= 0 dimXTAX = 0 =nullityTA Since nullityA equals the number of free variables in rrefA and since rankA equals the number of leading ones in rrefA : rankA = n – nullityA . By Theorem 4.10, page 126:

rankTA = n – nullityTA . Since nullityA = nullityTA : rankA = rankTA . CYU SOLUTIONS B-19

CYU 5.7 32– ab 10 3a –12c =  3b –02d =  1 2 4 3 =====  and   a –--- b --- c –--- d --- 41– cd 01 4ac–0=  4bd–1=  5 5 5 5

From the above, we see that the matrix 32– is invertible, with inverse –2515  . 41– –3545 

–1 –1 –1 CYU 5.8 I. Theorem 5.8 tells us that the claim is valid for n = 2 : A1A2 = A2 A1 .

 –1 –1 –1 –1 II. Assume validity at nk= : A1A2 Ak = Ak Ak – 1 A1 III. We establish validity at nk=1+ :

 –1 –1 A1A2 AkAk + 1 = A1A2Ak Ak + 1 –1 –1 By I: = Ak + 1 A1A2Ak –1 –1 –1 –1 –1 –1 By II: ==Ak + 1Ak Ak – 1 A1 Ak + 1Ak Ak – 1 A1

–1 100 0 –1 100 0 CYU 5.9 01 0 01 0 1 –1 050 0 0 --- 0 0 16 16– 10 0 ===10 0 5 001 0 01 01 00 1 00 1 001 0 000 1 000 1

E 2103 3341 100 001 001 2103 3341 CYU 5.10 (a) R  R R  R 1326 1 3 1326 010 1 3 010 and 010 1326 = 1326 3341 2103 001 100 100 3341 2103

E (b) 2103 334 1 100 001 001 2103 334 1 2R  R 2R  R 1326 2 2 26412 010 2 2 0 2 0  and 0 2 0 1326= 26412 3341 210 3 001 100 100 3341 210 3

A A–1 1 0 1 21 0 0 0 1000–586 7 – CYU 5.11 AI = 0 2 1 40 1 0 0 0100–232 3 – 1 1 1 00 0 1 0 0010107–10–8 0 3 1 10 0 0 1 0001–111 1 – B-20 CYU SOLUTIONS

CYU 5.12 Assume AB= I . Multiplying both sides of the equation BX = 0 by A we have: AB XA= 0  IX ==0  X 0  B is invertible (Theorem 5.17) Then: AB== I  AB B–1 B–1  ABB–1 ===B–1  A B–1  A–1 B .

120– 1 100– 1 4 –1 –14 CYU 5.13 301 1 rref 010– 3 8  1 = –38

042 2 001 7 4 2  74

CYU 5.14

33 72 11– 26 T ====334  T 272  T–116  T6214 22 11 51 59

12032–6 1 rref 10012–14 3 4  1234–14  30137 1 2 01010–238 1 8   T  = 1018–238  04242 6 14 00101 134 54 0 1 13 4 54

 T

rref  1203 100 3 4  34 CYU 5.15 13 13 T==312 and 3011 010 9 8  T 98 20 20 0422 001–4 5  –45 

37121 1000 6928 69 28 rref 32–6 1 3 0100– 1114  13 = –11 14 21552 0010 1 28 20  128 21190 0001– 1328 –13 28

69 28 1234–14  34 13 –11 14 13 T  ===1018–238  98 T 20 128 20  01134 54  –13 28 –45  From CYU 5.15 CYU SOLUTIONS B-21

 T 2 CYU 5.16 T : T111 = 1x ++1x 1   100110 100110 110 T110 = 1x2 ++1x 0   010111 010111 T  = 111  T100 = 0x2 ++1x 0  002100 0011 200 12 00

2  L L : Lx= 11   01100 10012 012 Lx= 01    L  =  11112 01100 100 L2 = 02 

2  LT  LT : LT 111 ===LT111 Lx++x 1 13   01110 10211 211 LT 110 ===LT110 Lx2 ++x 0 12    LT  =  11321 01110 110 LT 100 ==LT100 L0x2 ++x 0 =01 

CYU 5.17 Since T  is invertible, it must be a square matrix of dimension n. It follows that V

and W are both of dimension n. Let  = v1v2  vn ,  = w1w2  wn , and –1 T  = aij . Consider the linear transformation L: WV which maps wi to the vector –1 a1iv1 +++a2iv2  anivn . From its very definition, we see that L  ==aij T  (note,

for example that Lw1  is the first column of aij ). Applying Theorem 5.22, we see that: –1 LT  ===L T  T T  I . It follows that LT: VV is the identity map, and that therefore T is invertible with inverse L.   23  I

V   CYU 5.18 IV  and 23  : 0 2 122 rref 105 623 43 31– 213 011 2 11

122rref 104 3 56 23 43 43 23  :. IV 23  ===23 . 213 011 3  12 1 13 1

CYU 5.19 Rotating the standard basis  = 10  01 clockwise by 60 leads us to the basis 1 3 1 3  = --- –------ --- ------. Finding I and 13 :    2 2 2 2

I  13  1 1 ------101 101–13 3 2 – 2 2 rref –3232  013 011 3 2 31+

1 3 1 3 Check: 13– --- –------+1331+ --- ------=  2 2 2 2 B-22 CYU SOLUTIONS

2 2 2 CYU 5.20 For T  and I  . Noting that Tx+ 1 ==–2x +  Tx– x – x – x , and T1 = 2 ; and that  = x2x2 + x x2 ++x 1   = x2 + 1x2 –1x ; leads us to:

 T I T  I  1100–0111 1 100–2 1–0122 rref 01–01–1–0011 0101100–1 1– 101202001 001322–2 1–1–

Noting that Tx2 ==–xTx 2 + x x2 – x and Tx2 ++x 1 =x2 –2x + leads us to: T T  I  I  111011110 100122120 011–1 1–1–01–0 rref 010–1 1–3–1–1–1– 001002101 001002101

122 122 120 –21 –0 Then: I T I  ===01–1– –11 –3– –11 –1– –11 –1– T  –21 –1– 002 101 101

CYU 5.21 (One possible solution). Choosing to let c ==0 d 8 in system (*) at the bottom of page 197 we obtain the solution a ====–18 b 11 c 0 d 8 . At this point, we know that: –1 –812 = –1118 –1 4–1118 . Preceding as in part (c) of Example 5.13, we arrive –1518 08 8 4 08

at the basis  = v1 v2 , with:

v1 = –1812 + 021 = –18 –36 and v2 ==1112 + 821 27 30 . –812 Let’s verify that T  = for  = –18 –36  27 30 : –1518

T–18 – 36 = –270 –108 T27 30 = 261 162  –18 27– 270 261 rref 10–8 12 –36 30– 108 162 01–15 18 T  CYU SOLUTIONS B-23

CHAPTER 6 DETERMINANTS AND EIGENVECTORS

29– 3 93– 23– 29 CYU 6.1 (a-i) det 32–4=75det –det –det6 ==536– 6 –6478+ 9 ––27– 217 –42 34 32– 57– 6

29– 3 34 23– 23– (a-ii) det 32–4=–det9 – 2det –97det ==–18–20– –789212–15+ –+ 217 56– 56– 56– 57– 6

CYU 6.2 By induction on the dimension, n, of Mnn .

I. Claim holds for n = 2 : det a 0 ==ad – 0  c ad . cd II.Assume claim holds for nk= .

III.We establish validity at nk= + 1 : Let Aa= ij  Mk + 1  k + 1 . Since all entries in the

first row after its first entry a11 is zero, expanding about the first row of A we have

detA = a11detB , where B is the k by k lower triangular matrix obtained by removing the  first row and first column from the matrix A. As such, by II: detB = a22a33 ak + 1 k + 1 .  Consequently: detA ==a11detB a11a22a33 ak + 1 k + 1

CYU 6.3 Let the ith and jth row of A be identical, with ij . Multiplying row i by –1 and adding it to row j we obtain a matrix B whose jth row consists entirely of zeros. As such detB = 0 (just expand about the jth row of B). Applying Theorem 6.3(c), we conclude that detA = 0 .

CYU 6.4 Theorem 6.4(a) Theorem 6.4(c) (two times)

2 101 1014 1014 012 2 012 2 012 2 det ==–det –det 101 4 210 1 01– 2 –7 411 3 411 3 0 1 –3 –13

1014 1014 012 2 012 2 15 ==–det –4det ==– –------15 00– 4–9  00– 4 –9 4 15 00– 5 –15 000 –------4 CYU 6.2(b) B-24 CYU SOLUTIONS

CYU 6.5 (a-i) Let BM nn , and let E be the elementary matrix obtained by multiplying row i of

In by c  0 . By Theorem 5.12, EB is the matrix obtained by multiplying row i of B by c. Conse- quently: detEB ==cdetB detE detB Theorem 6.4(b) detE = c [Theorem 6.4(b)] th (a-ii) Let BM nn , and let E be the elementary matrix obtained by adding a multiple of the i th row of In to its j row. By Theorem 5.12, EB is the matrix obtained by adding a multiple of row i of B to its jth row. Consequently: detEB ==detB detE detB Theorem 6.4(b) detE = 1

(b) We use the Principle of Mathematical Induction to show that for any BM nn and elementary

matrices E1E2  Es  Mnn :    detEs E2E1B ==detEs E2E1 detB detEs detE2 detE1 detB I. Validity for s = 1 follows from Theorem 6.5.    II. Assume detEk E2E1B ==detEx E2E1 detB detEk detE2 detE1 detB   III. detEk + 1  Ek E2E1B = detEk + 1 Ek E2E1B  By I: = detEk + 1 detEk E2E1B By II:  = detEk + 1 detEk detE2 detE1 detB

1 CYU 6.6 IAA=det–1  I =1detAA–1  = detA detA–1  detA–1 = ------detA

32 10 62 62 11 3 CYU 6.7 E–3 ==null– –3 null . From rref we see 32– 01 31 31 00

r that E –3 = –--- r rR with basis 13– .    3 CYU 6.8 The eigenvalues are the solutions of the equation: 16 –3 2 detA – I3 ==det –4 3– –8 0 –2 –116 –  Which reduces to –0x – 15 2 = . It follows that 0 and 15 are the eigenvalues of A.

16 3 2 100 16 3 2 16 3 2 101 2  rref Then: E0 ==null–384 – – 0 010 null–384 – . From –384 – 01– 2 we  –62 –11 001 –62 –11 –62 –11 00 0 r see that E0 = –--- 2rr r   with basis –142 .  2 CYU SOLUTIONS B-25

 16 3 2 100 132 132 132 rref E15 ==null–384 – – 15 010 null–124 –8– . From –124 –8– 000 we  –62 –11 001 –62 –4– –62 –4– 000 see that E15 = – 3r – 2srs rs with basis –310  –201 .

CYU 6.9 The kernel of the linear operator:

T – –3 I 2 ab ==Tab + 3ab 3a ++2b 3a 3a – 2b + 3b =6a + 2b 3a + b  is, by definition, the set: ab 6a +32b ab+ = 00 . Equating coefficients, we have:

6a +02b =  62rref 11 3   . It follows that E–3 = –r 3r r   with basis 3ab+0=  56– 00 –31 .

CYU 6.10 With respect to the basis  = x2 ++x 1 x + 11 : Tx2 ++x 1 = 2x2 ++5x 2 Tx+ 1 = x2 ++3x 1  T1 = x2 ++x 1 T  100211 100 2 1 1 rref 110531 010 3 2 0 111211 001–2 3–0

T 2 – 1 1 2 From detT  – I3 ===det 3 2– 0 –02 –  we see that the eigenval- –3 –2 –  ues are 0 and 2: same as those found in the solution of Example 6.10. Since the determination of the corresponding eigenspaces only depends on T and the eigenvalues, the spaces E0 and E2 are identical to those determined in the solution of Example 6.10.

CYU 6.11 The three vectors 120 122 211 are easily seen to constitute a basis  for 3 . Since T120 = 120 , T122 = –122 , and T211 = 2211 ,  = 120 122 211 consists of eigenvector with corresponding eigenvalues 1 – 1 and 2 , respectively.

T120 = 1120 ++0122 0211 100  From:T 122 = 0120 ++–1 122 0211  we have: T  = 01–0 T211 = 0120 ++0122 2211  002 B-26 CYU SOLUTIONS

CYU 6.12 By Theorem 6.12, the n eigenvalues of T are linearly independent. Since V is of dimen- sion n, those n eigenvectors constitute a basis for V. It follows, from Theorem 6.11, that T is diag- onalizable.

CYU 6.13 (a) Characteristic polynomial:  – 1 –0 1 detA – I ===det–31 –0 – 3 +++2  2 – – 2 2 ++ 1  –1314 – –  From the above, we see that  = 2 is the only (real) eigenvalue of A (note that the discriminant of 2 ++ 1 is negative). Determining the dimension of E2  –21 –0 1 E2 ==nullA – 2I null–321 –0  –1314 –2–

–013 10– 1 3 Turning to the homogeneous system of equations: –101 rref 01– 1 3 we conclude –1334 – 00 0 that E2 = rr3r r   with basis 113 . It follows that there does not exist a basis for 3 consisting of eigenvectors of A, and that therefore A is not diagonalizable.

 3 –2 – 1 (b) det2 6 –2 – ==– 3 +++2  2 – – 2 2 – 8 .  –21 – 3 – 

  1 21– –215 – 12– 1 12– 1 E2 = null2 4 –2 and E8 = null2 –22 – . From 24– 2 rref 00 0   –21 – 1 –21 – –5 –21 –1 00 0

–215 – 101 rref and 22–2– 012 , we see that E2 = – 2r + srs rs   and –21 –5– 000 E8 = –r–2r r r   , with bases –210 101 and –1 –21 , respectively. It follows that –210101 –1 –21 is a basis for 3 consisting of eigenvectors of A, and that therefore A is diagonalizable. Theorem 6.15 tells us that the matrix P–1AP will turn

–112 – out to be a diagonal matrix with eigenvalues along its diagonal, where P = 10– 2 . 011 CYU SOLUTIONS B-27

–1 –112 – 32– 1–112 – 200 We leave it for you to verify that: 10– 2 26– 2 10– 2 = 020 . 011 –21 –3 011 008

CYU 6.14 From Theorem 6.19 and Example 6.11 we have: 10 10 02–0 2 000 0 0 00 0 0 00 0 0 00 0 110– 1 ==000 0 =0 00 0 –121 –1 002 0 0 0 210 0 0 01024 0 –121 –1 000– 2 0 00– 210 0 0 0 1024

32– 1 CYU 6.15 In the solution of CYU 6.13(b), we found the characteristic polynomial of 26– 2 –21 –3 to be – – 2 2 – 8 with E2 and E8 of dimensions 2 and 1, respectively.

th 12 CYU 6.16 Let sk denote the k element of the sequence (for k  3 ), F = , and 10

sk 3 3 2 3 k – 2 3 Sk = . From: S3 ==FS2 F  S4 ==FS3 FF =F  Sk =F , we see sk – 1 2 2 2 2 k – 2 3 that sk is the sum of the entries in the first row of F . We now set our sights on finding the 2 matrix Fk , and begin by finding a diagonalization for F:

det 1 –2 == – 2 + 1 ; E2 2rr r   ; E–1 =–r r r   1 – It follows that 21  –11 is a basis of eigenvectors for F. Employing Theorem 6.18 we have: –1 FPDP==–1 21– 20 21– =12 11 01– 11 10 steps omitted From Theorem 6.19: –1 1 1 k + 1 1 2 k + 1 21– 2k 0 21– ---  2k + 1 –---–1 ---  2k + 1 + ---–1 Fk ==PDkP–1 =3 3 3 3 11 k 11 01– ************** *************** steps omitted B-28 CYU SOLUTIONS

1 1 k – 1 1 2 k – 1 k – 2 3 ---  2k – 1 –---–1 ---  2k – 1 + ---–1 3 Thus: F = 3 3 3 3 2 2 ************** ***************

2 k – 1 4 5 1 2k – 1 – –1 k – 1 ++---  2 ---–1 k – 1 ---  2k – 1 + ---–1 k – 1 ==3 3 3 3 ****************************** *************** 5 1 Conclusion: s = ---  2k – 1 + ---–1 k – 1 . k 3 3

CYU 6.17 –11101001 1100 1011 is a basis of eigenvalues for 02–0 2 110– 1, with 0 the eigenvalue associated with –1110 and 1001 , 2 the eigenvalue –121 –1 –121 –1 associated with 1100 , and –2 the eigenvalue associated with 1011 (Example 6.11, page 237). Applying Theorem6.25 we conclude that: – c ++c c e2x +c e–2x –1 1 1 1 1 2 3 4 2x 0x 1 0x 0 2x 1 –2x 0 c1 + c3e c1e +++c2e c3e c4e = –2x 1 0 0 1 c1 + c4e 0 1 0 1 –2x c2 + c4e is the general solution of the given system of equations.

2x –2x y10 = 0 y1 = – c1 ++c2 c3e +c4e 2x y20 = 1 CYU 6.18 From CYU 6.16: y2 = c1 + c3e Since : –2x y30 = 2 y3 = c1 + c4e –2x y40 = 3 y4 = c2 + c4e

– c1 +++c2 c3 c4 = 0  –1 1110 1000 2 c1 +1c3 =  aug(S) 1 0101 0100 3 S:  rref c1 +2c4 =  1 0012 0010– 1  0 1013 0001 0 c2 +3c4 = 

2x 2x Solution: y1 = 1 – e y2 = 2 – e y3 = 2 y4 = 3 . CYU SOLUTIONS B-29

5 1 5 1 --- –------–  –---  5 2 1 5 1 3 T t = 2 4 Tt det 2 4 ===0  --- –  – --- 0  --- –  ---   = CYU 6.19 .   Ft 5 Ft 5 2 4 2 2 2 –1 --- –1 --- –  2 2

5 1 5 1 ---3– –------2– –--- E3 ==null 2 4 –r 2r r   and E2 ==null 2 4 r 2r r   –1 --5-3– –1 --5-2– 2 2 Choosing –12 and 12 as eigenvectors for the eigenvalues 3 and 2, respectively, we have:

–c e3t + c e2t Tt ==c e3t –1 + c e2t 1 1 2 . Turning to the initial conditions Tt = 120 : Ft 1 2 2 2 3t 2t Ft 200 2c1e + 2c2e

30 20 –c + c = 120 –c1e + c2e 120  1 2 === c1 –10 and c2 110 30 20 200 2c1 +2002c2 = 2c1e + 2c2e

10e3t + 110e2t Bringing us to:Tt = . Setting Tt= Ft we have: Ft 3t 2t –20e + 2202e 3t 2t 3t 2t 10e +20110e = – e + 2202e 30e3t = 110e2t ln30e3t = ln 110e2t ln30 + lne3t = ln 110 + lne2t ln30+ 3t = ln 110+ 2t  t =ln110 – ln 30  1.3years

CYU 6.20 Let D denote the state that a student is living in the dorm, and C denote the state that current state

D C next state .8 .1 D 858 the student is a commuter. Then:T = with S0 = . Then: .2 .9 C 702

------3783- .8 .1 858 5 757 D S1 == 803 .2 .9 702 ------4017- C 5

.8 .1 757 686 D .8 .1 686 636 D S2 =  and S2 =  .2 .9 803 874 C .2 .9 874 924 C Conclusion: 757, 686, and 636 of the current freshmen will live in the dorm in their sophomore, junior, and senior year, respectively. B-30 CYU SOLUTIONS

.73x ++.32y .09z = x  CYU 6.21 .73 .32 .09 x x  .21 .61 .04 y = y  .21x ++.61y .04z = y   .06 .07 .87 z z .06x ++.07y .87z = z 

479 100------–.27x ++.32y .09z = 0 1157  297 .21x –.39 +0 .04z =  rref 010------ 1157 .06x +0.07y – .13 =  381  001------xyz++= 1 1157 0000 see solution of Example 6.15.

479 297 381 Conclusion: Eventually ------------, or approximately 41%, 26%, 33% of the population, 1157 1157 1157 will vote democratic, republican, green, respectively.

CHAPTER 7 INNER PRODUCT SPACES

n n

CYU 7.1 ruv ====ru1u2  un  v1v2  vn  rui vi  uirvi u  rv i = 1 i = 1

CYU 7.2

(a) cv ==cv1v2  vn  cv1v2  vn cv1cv2  cvn  cv1cv2  cvn

n n 2 2 2 2 == c vi c  vi =c v i = 1 i = 1 (b) uv– 2 ==uv–  uv– uuv – – vuv – ==uu – uv – vu + vv u 2 – 2uv + v2

–1 uv –1 120  –131 –1 5 CYU 7.3  ==cos ------cos ------=cos ------ 83 uv 14+ 191++ 55 CYU SOLUTIONS B-31

CYU 7.4 For uu  v and r   : ruu+  v = ruv + uv ==r0 +00 . We see that ruu+  v . It follows, from Theorem 2.13 (page 61), that v is a subspace of n .

uv 0241  301– 1 3 6 12 3 CYU 7.5 proj v ==------u ------0241 =------0241 =0------u uu 0241  0241 21 21 21 21

6 12 3 6 9 24 and: v – proj v ==301– 1 – 0------3–------– ------. u 21 21 21 21 21 21 6 9 24 6 12 3 301– 1 = 3–------– ------+ 0------21 21 21 21 21 21

CYU 7.6 (a) A direction vector for the line L passing through 12 – and 24 : u ==24 – 12 – 16 . The vector from the point 12 – on L to P = 25 : v ==25 – 12 – 17 . Applying Theorem 7.2, we have: uv 16  17 43 proj v ==------u ------16 =------16 u uu 16  16 37

43 258 6 1 62 + 12 1 Hence: v – projuv ====17 – ------ ------–------ ------. 37 37 37 37 37 37

(b) u ==1221 – 1201 0020 . v ==1013 – 1201 02– 12 . uv 0020  02– 12 1 proj v ==------u ------0020 =---0020 =0010 u uu 0020  0020 2

Hence: v – projuv ====02–12–020010 –02 16 4 .

CYU 7.7 A normal to the desired plane will have the same direction is that of the line passing through the two given points; namely: n ==021 – 110 –111 . Normal form for the plane: –111  x – 1y – 3 z + 2 = 0 .

CYU 7.8 Let A=  xyz 3xy+6– 2z = , and B ==200 ++r021 s203 rs  R 22+2s rr + 3s rs  R. BA : xyz = 22+2s rr + 3s .  3xy+322– 2z ===+ s +662r – 2r + 3s ++s 2r –62r – s 6 AB : If xyz is such that 3xy+6– 2z = (*) , can we find rs  R such that x = 22+ s , y = 2r , and zr= + 3s ? Yes: x – 2 y In order for x = 22+ s , s = ------. In order for y = 2r , r = --- . We show that for 2 2 those particular values of r and s, zr= + 3s : z ==------3xy+ ------– 6------32+ 2------s + 2------r – 6- =------66++s ------2r – 6-2 =s + 2 2 2 2 (*) B-32 CYU SOLUTIONS

CYU 7.9 Let A=  xyz be on the plane ax++ by cz = d with normal n = abc .

Determine the vector v from A to x0y0 z0 : v ==xyz – x0y0 z0 xx– 0yy– 0 zz– 0 . Applying Theorem 7.2, we have: vn ax– ax + by– by + cz– cz proj v ==------n ------0------0 ------0- abc n nn a2 ++b2 c2 ax – ax + by – by + cz – cz = ------0------0 ------0- abc a2 ++b2 c2

ax0 ++by0 cz0 – d ax ++by cz – d Since ax++ by cz = d: ==------a2 ++b2 c2 ------0 ------0 ------0 2 2 2 a ++b c a2 ++b2 c2

CYU 7.10 (a)AB ==25– 3 – 32–2 –175– and AC ==41  – 3 – 32–2 13– 5 .

ijk Here is a normal to the plane: n ==det –751 – –1020i –10j – k =–20 – 10 –10 . 13– 5 1 Here is a “nicer” normal: n ==–------–20 – 10 –10 211 . 10 Choosing the point A = 32–2 on the plane, we arrive at the general form equation of the plane: 211  x – 3y + 2 z – 2 ==0 or: 2xyz++– 6 0. (b) In Example 2.15, page 72, we found the vector form representation for the above plane: P = 3 – r + s –72 ++r 3s 25–5r – s rs   . We are to show that: P ===3 – r + s –72 ++r 3s 25–5r – s rs   xyz 2xyz++= 6 Q . If xyz = 3 – r + s –72 ++r 3s 25–5r – s  P , then: 2zyz++= 23– r + s ++–72 ++r 3s 25–5r – s Thus: PQ . ==62–2r +s –72 +++r 3s 25–5r – s 6 If xyz is such that 2xyz++= 6 can we find real numbers r and s such that: 3 – r + s ==x –72 ++r 3s y and 2–5 5r – s ==z 62– x – y since 2xyz++= 6 which is to say: – r + s ==x –73 r + 3s y +2 and –5 5r – s =– 2x –4y + Turning to the system of equation: r s 1 0 –------3x ++y 19 – r + s = x – 3  –11 x – 3 10  augmented 7r + 3s = y + 2 rref 7xy+ – 19  matrix 73 y + 2 0 1 ------ 10 –5r –25s = – x –4y +  –55 –2– x –4y + 00 0 steps ommited 7xy+ – 19 We see that rx= –3+ and s = ------does the trick. Thus: QP . 10 CYU SOLUTIONS B-33

CYU 7.11 (i) For v = ax2 ++bx c , vv ==ax2 ++bx c ax2 ++bx c a2 ++b2 c2  0 and vv = 0 only if abc===0 . 2 2 (ii) For u ==a2x ++a1xa0 v b2x ++b1x b0 : 2 2 uv ==a2x ++a1xa0 b2x ++b1x b0 a2b2 ++a1b1 a0b0

= ba2 ++b1a1 b0a0 ==b x2 ++b x b  a x2 ++a x a vu 2 1 0 2 1 0 2 2 (iii) For u ==a2x ++a1xa0 v b2x ++b1xb0 , and r   : 2 2 ruv = ra2x ++a1xa0  b2x ++b1xb0 2 2 = ra2x ++ra1xra0 b2x ++b1xb0

===ra2b2 ++ra1b1 ra0b0 ra2b2 ++a1b1 a0b0 r uv 2 2 2 (iv) For u ==a2x ++a1xa0 v b2x ++b1xb0 and z = c2x ++c1x c0 : 2 2 2 uv+  z = a2x ++a1xa0 + b2x ++b1xb0  c2x ++c1xc0 = a + b x ++a + b xa+ b  c x2 ++c xc 2 2 2 1 1 0 0 2 1 0 = a2 + b2 c2 + a1 + b1 c1 + a0 + b0 c0

==a2c2 ++a1c1 a0c0 + b2c2 ++b1c1 b0c0 uz + vz

CYU 7.12 ru rv ==r u rv rruv =r2 uv

Definition 7.5(iii) Theorem 7.4 (c)

CYU 7.13 (a) rv ===rv rv r2 vv r2 vv =r vv CYU 7.12 Definition 7.5(iv) (b) uv+ 2 ==uv+  uv+ uu + v + vu + v =uu +++uv vu vv Definition 7.6 and 7.5(i): = u 2 +++uv uv v 2 = u 2 ++2 uv v 2

CYU 7.14 (a) 35– 8 ==35– 8  35– 8 532 ++552 58– 2 = 710.

(b) 35– 8 – 102 = 25– 10  25– 10 ==25– 10  25– 10 522 ++552 510– 2 = 645 B-34 CYU SOLUTIONS

CYU 7.15 1 1 1 1 1 1 1 1 1 1 ------u–------v ------u–------v  0  ------u – ------v ------u –0------u – ------v ------v  u v u v u v u u v v 1 1 1 1 1 1 1 1  ------u ------u – ------v ------u – ------u ------v +0------v ------v  u u v u u v v v 1 2 1  ------uu – ------uv +0------vv  u 2 uv v 2 2 2 1 – ------uv +01  2  ------uv  uv  uv uv uv

–1 35– 8 102 CYU 7.16  = cos ------35– 8 102

–1 531 ++250 48– 2 = cos ------533 ++255 48– –8 511 ++200 423 –1 –49 = cos ------ 119.1 351 29

m

CYU 7.17 For any v = civi  Spanv1v2  vm : i = 1 m m

uv ===u  civi  ci uv i 0 , since uv i = 0 for 1 im . i = 1 i = 1

CYU 7.18 Exercise 46, page 300

vw = a1v1 ++ anvn b1v1 ++ bnvn

= a1v1 ++ anvn b1v1 ++a1v1 +++a2v2  anvn bnvn

n n n n == aivi b1v1 ++  aivi bnvn  aib1 vi v1 ++  aibn vi vn i = 1 i = 1 i = 1 i = 1 n  1ifij= since vi vj =  : ==a b ++ a b a b  0ifij 1 1 n n  i i i = 1 CYU SOLUTIONS B-35

CYU 7.19 The first order of business is to determine a basis for the space S spanned by the vec- tors 2110 , 1010 , 3120 , 0101 . Applying Theorem 3.13, page 103 we see that first, second, and fourth of the above four vectors constitute a basis: v1 = 2110 , 2130 1010 1011 0110 v2 = 1010 , v3 = 0101 of S: . We now apply the Grahm- 1120 0001 0001 0000

Schmidt Process to that basis v1v2 v3 to generate an orthonormal basis u1u2 u3 for S:

u1 ==v1 2110 u  v 1 2 2110  1010 u2 ==v2 – ------u1 1010 – ------2110 u1 u1 2110  2110 3 1 1 ==101 – ---2110 0–------0 6 2 2

1 1 0–------0  0101  u1 v3 u2 v3 2110  0101 2 2 1 1 u3 ==v3 – ------u1 – ------u2 0101 – ------2110 – ------0 –------0 u  u u  u 2110  2110 1 1 1 1 2 2 1 1 2 2 0–------0  0–------0 2 2 2 2 1 –12 1 1 1 1 1 ==0101 – ---2110 – ------0–------0 –---------1 6 12 2 2 3 3 3

1 1 1 1 3 1 1 1 Conclusion: ---2110 ---0–------0 ---–---------1 is an orthonormal basis for  6 2 2 2 4 3 3 3 Span2110 1010 3120 0101 .

CYU 7.20 A consequence of Theorem 7.10(iii) and CYU 3.11, page 98.

CYU 7.21 We first use the Grahm-Schmidt Process to determine the orthonormal basis

w1 w2 of W stemming from the given basis v1 v2 = 101  120 : 1 u1 ==v1 101 : w1 = ------101 2 u  v 1 2 101  120 1 1 2 1 1 ------120 – ------101 ---2–---------2–--- u2 ==v2 – u1 =: w2 =  u1 u1 101  101 2 2 3 2 2

Turning to Theorem 7.11, we determine the orthogonal projection, vW , of the vector v = 201 onto W: B-36 CYU SOLUTIONS

vW = vw 1 w1 + vw 2 w2 1 1 2 1 1 2 1 1 = 201  ------101 ------101 + 201  ---------2 –---------2 –--- 2 2 3 2 2 3 2 2 3 1 1 1 7 5 ==---101 + ------2 –--- ---1 --- 2 22 2 4 4

CYU 7.22 NOTE: It is easy to see that the function 2 n fa0a1  an = a0 ++a1xa2x ++ anx from the standard (dot-product) inner product space n of Theorem 7.1 (page 279) to the polyno- mial inner product space Pn of Exercise 23 (page 299) is an isomorphism which ALSO preserves the inner product structure of the two spaces: n a0a1  an  b0b1  bn ==fa0a1  an  fb0b1  bn  aibi i = 0

As such, we could translate the given P3 -problem: Find the shortest distance between the vector v =33x2 + x and the 2 3 subspace W = Span x + 1 x + 1 in the inner product space P3 into the following 4 -form: Find the shortest distance between the vector v = 03 30  and the sub- space W = Span 0101  1001 in the inner product space 4 . You are invited to take the above 4 -approach. For our part, we will deal directly within the inner product space P3 : 2 Employing the Grahm-Schmidt Process we go from the basis v1 v2 , with v1 = x + 1 and 3 2 3 v2 = x + 1 , to an orthonormal bases w1 w2 for W = Span x + 1 x + 1 : u 2 1 1 2 u1 ==v1 x + 1 : w1 ==------x + 1 . u1 2

u1 v2 x2 + 1 x3 + 1 u ==v – ------u x3 + 1 – ------x2 + 1 2 2 1 2 2 u1 u1 x + 1 x + 1 u 1 1 1 2 2 1 1 ==x3 + 1 – ---x2 + 1 x3 – ---x2 + --- : w ==------x3 – ---x2 + --- 2 2 2 2 2 u2 3 2 2 Turning to Theorems 7.11 we determine the projection vW of v =33x + x onto W: vW = vw 1 w1 + vw 2 w2 1 1 2 1 1 2 1 1 = 3x2 + 3x ------x2 + 1 ------x2 + 1 + 3x2 + 3x ---x3 – ---x2 + --- ---x3 – ---x2 + --- 2 2 32 2 32 2 3 1 3 2 1 1 ==------------x2 + 1 – ------x3 – ---x2 + --- x3 ++x2 1 22 232 2 CYU SOLUTIONS B-37

Appealing to Theorem 7.12 we calculate the shortest distance between the vector v =33x2 + x and the subspace W = Span x2 + 1 x3 + 1 :

2 3 2 vv–3W = x + 3x – x ++x 1 ==–2x3 –31x2 ++ –2x3 –31x2 ++ –2x3 –31x2 ++

===–1 2 ++–2 2 3 2 +1 2 93

CYU 7.23 Determining the eigenvalues of A:

2 –  11 2 detA – I ==det 1 2 –  1 –4 – 4 – 1 Eignevalues:  ==  1 112–  details omitted Determining the corresponding eigenspaces: –112 –112 10– 1 E4 ==nullA – 4I null 12–1 =aaa a   since rref 12–1 =01– 1 11– 2 11– 2 00 0

111 111 111 E1 ===nullAI– null 111 – c – dcd cd since rref 111 =000 111 111 000 As is indicated in Theorem 7.14: aaa  – c – dcd ==ac– – d ++ac ad 0

CYU 7.24 (a) v w 111 1 1 121 v  w = v1 ++v2 v3v1 ++2v2 v3 v1 ++v2 3v3  w1w2 w3 2 2 113 v3 w3

= w1v1 ++v2 v3 ++w2v1 ++2v2 v3 w3v1 ++v2 3v3 v w 1 111 1 v  121 w = v1v2 v3  w1 ++w2 w3w1 ++2w2 w3 w1 ++w2 3w3 2 2 v3 113 w3

= v1w1 ++w2 w3 ++v2w1 ++2w2 w3 v3w1 ++w2 3w3 = w v ++v v ++w v ++2v v w v ++v 3v 1 1 2 3 2 1 2 3 3 1 2 3 B-38 CYU SOLUTIONS

111 0 1 (b) One possible answer:  000 1  0 ==100  100 1  000 0 0

 0 111 1 while: 1  000 0 ==010  100 0  0 000 0

CYU 7.25 (a) Tabc  ABC = ab– – a + 2bc– – b + c  ABC = Aa– b ++Ba–2+ bc– Cb– + c abc  TABC = abc  AB– – A + 2BC– – B + C = aA– B ++bA–2+ BC– cB– + C = Aa– b ++Ba–2+ bc– Cb– + c (b) For Tabc = ab– – a + 2bc– – b + c and  = 010 001 100 :

11–0 T  = –211 – 01–1

11–0 CYU 7.26 Consider the symmetric matrix AT== –211 – of CYU 7.25(b). 01–1 Employing Theorem 6.10 of page 226, we find a basis of eigenvectors for the linear operator Tabc = ab– –2a + bc– – b + c :

1 –  –01 detA – I ==det –1 2 –  –1 –0– 1 – 3 Eignevalues:  ===  1  3 01–1–  Here are the associated eigenspaces: CYU SOLUTIONS B-39

11–0 11–0 10– 1 E0 ==nullA null –211 – =aaa a  3 since rref –211 – =01– 1 01–1 01–1 00 0

01–0 01–0 101 E1 ==nullAI– null –111 – =–a0 a a  3 since rref –111 – =010 01–0 01–0 000

–12 –0 –12 –0 10– 1 E3 ==nullA – 3I null –11 –1– =a–2a a a  3 since rref –11 –1– =01 2 01–2– 01–2– 00 0

Letting a = 1 in each of the above eigenspaces we arrive at a normal basis for 3 consisting of eigenvectors of T: 111 –101 12– 1 (Theorem 6.15); which is easily turned 1 1 1 –1 1 1 –2 1 into an orthonormal basis: ------------------0------------.  3 3 3 2 2 6 6 6

CYU 7.27 If A–1 = A T and B–1 = B T , then: AB –1 ===B–1A–1 BTAT AB T Theorem 5.12(iii), page 167 Exercise 19(f), page 162

2 1 1 CYU 7.28 In CYU 7.23 we showed that the matrix 121 has eigenvalues  ==4  1 112 with E4 ==aaa a    E1 – a – bab ab , with 111 a basis for E4 , and –101  –101 a basis for E1 (set a = 0 and b = 1 , and then set b = 0 and a = 1 ). Applying the Grahm-Schmidt process (Theorem 7.9, page 303) to 1 1 –101  –101 , we arrive at the orthogonal basis –101  –1---–--- of E1 , and  2 2

1 1 1 1 1 1 1 1 to the orthonormal basis of eigenvectors ------–0------–------–------------–------.   3 3 3 2 2 6 6 6 Turning to the marginal comment on page 314 we conclude that:

2 1 1 400 13 – 12 – 16 T P 121P = 010, where P = 13 016 112 001 13 12 –16 B-40 CYU SOLUTIONS ANSWERS C-1

Appendix C Answers to Selected Exercises 1.1 Systems of Linear Equations, page 11. 5xy++4z = 6 1. 33–12 3.  5. 100 7. 100 –32x – y +4z =  55–1 9–  010 010 –43 –10 1  001 001 ---xy–0=  2 000

9. x = 1 y ==0 z 2 11. x1 = 1 x2 = 2 x3 = 2 x4 = 2 x5 = –1 1 1 1 3 13. x ==0 y –24 z = 15. x = --- y = --- z = –--- w = --- 2 4 2 4 19. No: first non-zero entry in last row is not 1. 23. x ==5 y –22 z = 25. x = 12 y = –5 z = 1 w = 0

1.2 Consistent and Inconsistent Systems of Equations, page 23. 1. r 2 r   3. –22–2– 5. –s – 3t r 12– t 12– sst rst 

12+ r 36+ r 7. 11–6 5r–3+ r r r   9. 79– 6 11. 2 – r ------------rr  5 5 13. Yes. 15. No. Solutions if and only if a, b, c, satisfy the equation 4ab–2+0c = .

3 17. No. 19. Yes. 23. –---rr 0 rR 25. –11r – s r + 6srs rs  R  2 27. No. 29. No. 31. a  1 33. None. 35. ab  1 37. ad– bc  0 2.1 Vectors in the Plane and Beyond, page 38.

1. 3. 5. 23 3_ B = –32 . 3_ 3 _ _ 2 –32 2 2 . A = 11 . B = 01 . 1_ . 1 1 2 3 – 2 _ –3 –2 –1 1 2 -3 -2 -1 1 2 3 –1 -1 AB. . _ . –2 -2 A = –12 ANSWERS C-2

7. 20– 2 9. –9111 –  11. 13 – 13 15. –6 6 17. 25 – 7 1 5 1 19. (a) r = –14 s = 10 (b) r = –--- s = ------(c) r = --- s = ------5 10 7 14

19 17 11 5 5 21. r ==–------ s ------ t = ------23. ------ –------6 6 6 2 2 2.2 Abstract Vectors Spaces. page 49. 1. No. 3. No. 5. Yes. 7. No. 9. Yes. 11. No. 13. No. 15. Yes.

2.3 Properties of Vectors Spaces, page 57. (All exercises call for either a proof or a counterexample)

2.4 Subspaces, page 65. 1. Yes. 3. Yes. 5. No. 7. Yes. 9. No. 11. No. 13. Yes. 15. Yes. 17. No. 19. No. 21. Yes. 23. Yes. 25. Yes. 27. Yes. 29. No. 31. Yes. 33. No. 35. Yes. 37. Yes.

2.5 Lines and Planes, page 73.

1. r15 r   3. r51 – r   5. 13 + r17 – r  

7. 35 + r02 r   13. 37 + r15 r   15. 37 + r51 – r  

17. 37 + r17 – r   19. 37 + r02 r   21. 37 + r51 – r  

23. 37 + r15 r   25. 37 + r71 r   27. 37 + r01 r  

29. r245 r   31. r–402 r   33. 245 + r134 –  –

35. 210 + r13  – 1 41. 12– 1 + r245 r  

43. 12– 1 + r–402 r   45. 12– 1 + r13–4– r  

47. 12– 1 + r13– 1 r   49.r132 + s211 rs  

51. r200 + s020 rs   53. 341 ++r13– 4 s232 rs  

55. 24– 3 ++r33–8 s23–2rs   ANSWERS C-3

3.1 Spanning Sets, page 84.

1. No. 3. Yes. 5. Yes. 7. No. 9. No. 11. No. 13. cos2x = cos2x – sin2x

15. sin--- – x = sin--- cosx – cos--- sinx 17. Span. 7 7 7 19. Do not span. Just about any randomly chosen four-tuple will not be a linear combination of the given vectors (check it out). 21. Do not span. Just about any randomly chosen four-tuple will not be a linear combination of the given vectors (check it out). 23. Do not span. Just about any randomly chosen four-tuple will not be a linear combination of the given vectors (check it out). 25. Span. 27. All c  0 . 3.2 Linear Independence, page 91.

1. Yes. 3. No. 5. No. 7. Yes. 9. No. 11. No. 13. Yes. 15. No. 17. Yes. 19. Yes. 21. No. 23. No. 25. Yes. 27. No. 29. a = 3 3.3 Bases, page 104.

5 5 1. (a) –3 --- = – 3e + ---e (b) 320 = 3e ++2e 0e 2 1 2 2 1 2 3

19 25 14 3. (a) x2 + 3x – 1 = ------2x2 + 3 – ------x2 – x + ------x – 5 13 13 13 5. No. 7. Yes. 9. Yes. 11. Yes. 13. Yes. 15. No. 17. No. 19. Yes.

13 20 01 21 29.  31. Do not span. –21 11– 12 01

33. A basis for SpanS : 214  –132 . A basis for 3 : 214 –132 111 .

35. A basis for SpanS = 3 : S = 113 –132 32– 1

37. A basis for SpanS = 5 : S = 13132 24142 11202 22111 12345

39. A basisfor SpanS : 5 – x3 – xx 4 ++++x3 x2 x 12 x4 – 2x2 . A basis for 3 : 5 – x3 – xx 4 ++++x3 x2 x 12 x4 – 2x2 1 + x

41. sinxx cos  sin2x cos2x sin2x 43. –1 1001010 –1001

45. x3 ++x2 2x x + 1 47. c  0 49. a  0 b  0 and ab ANSWERS C-4

4.1 Linear Transformations, page 120. 1. Yes 3. No 5. No 7. No 9. Yes 11. No 13. Yes 15. No ab+ 17. No 19. (a) 4102– (b) ------ 2aa – + b 2

83 ab+ b 21. (a) (b) 27. b = 0 39. LT ab = 2a + 2b – a – b 13 8 2ab+ ab+

41. LT ab = 3ab+ x + 6a + 2b 43. KLT a = –022aa 4.2 Kernel and Image, page 131. 1. Linear. Nullity: 0, Rank: 1. 3. Linear. Nullity: 0, Rank: 1. A basis for the image space: 11 – 5. Not linear. 7. Linear. Nullity: 0, Rank: 2.

00 11 9. Linear. Nullity: 2, Rank: 2. A basis for the kernel:  . 01 10

A basis for the image space: x – 1 –1x2 + . 11. Not linear. 13. Linear. Nullity: 1, Rank: 2. A basis for the kernel: x2 – x . A basis for the image space: 01  11 . 15. Linear. Nullity: 0, Rank: 4.

10 –11 01– 17. Linear. Nullity: 0, Rank: 3. A basis for the image space:  . 10 11 01

19. Nullity: 1, Rank: 3. A basis for the kernel: 1 . A basis for the image space: x2x 1 .

21. Nullity: 0, Rank: 4. A basis for the image space: x3 x2x 1

23. Nullity: 0, Rank: 3. A basis for the image space: x2x 1 .

x2 1 x 1 25. Nullity: 2, Rank: 1. A basis for the kernel: ----- – --- --- – --- . A basis for the image space: 1 . 3 3 2 2

27. Nullity: 0, Rank: 3. A basis for the image space: x3x2 x . 29. (a) 033 – 

21 22 01 (b) x3 + x and 5. 31. Nullity: 0, Rank: 3. A basis for the image space:  . 01 13 10

33. (a) Tab = 9aa3a (b) Tab = 9ab0 (c) None exists. ANSWERS C-5

35. nullityT = n , rankT = 0 ; nullityT = 0 , rankT = n . 37. nullityT = 1 , rankT = 2 ; nullityT = 2 , rankT = 1 ; nullityT = 3 , rankT = 0 . 39. nullityT = 1 , rankT = 4

4.3 Isomorphisms, page 145.

1 1 1. f –1x = –---x 3. f –1ab = b –---a 5. f –1ax2 ++bx c = –cab 5 2 7. Isomorphism. 9. Not an isomorphism. 11. Isomorphism. 13. Not an isomorphism. 15. Isomorphism. 17. Isomorphism. 19. Isomorphism. 21. Not an isomorphism. 23. f –1xy = y + 4 xy–7– . Zero: 34 – and –xy = –6x +  –8y – .

5.1 Matrix , page 161.

011 002 010 Column: 3 10 4  262 1. 14– 3. 200 and 006 5. Row: 3215  106110 –151 090 600

Column: 1–55 1 –1 032 01–2–1– n 7. 25. 11 = 1 n Row: 1101 –0141–  53– 27 01 01

n n 10 10 27. 10 = 29. 10 = 02 02n 02 02n

5.2 Invertible Matrices, page 175.

2 –1 –--8- --- 1 3 1 5 5 --- 0 --- –--- –124 2 4 2 1 2 1 1. 3. 5. 20– 1 7. Singular. 9. Singular. 11. ------–------1 1 1 2 5 10 0 --- –------–011 5 4 2 5 1 1 --- –--- 3 3

500 001 13. ab  12 15. b  0 and a  1 17. E1 = 010, E2 = 010 001 100

1 19. X = ---ABA–1 21. XA= –1BAB 23. XAB= 2A 2 ANSWERS C-6

5.3 Matrix Representation of Linear Maps, page 186.

--8------17- 2 5 5 5 1. 23  = , T23  = 3. 23  = , T23  = 3 6 –--1- –--4- 5 5

4 7 2 0 5. 231  = 6 , T231  = 7 7. 12  = , T12  = 12 1 –3 –6 –3

0 0 2 1 2 –2 9. x ++x 1  = , Tx++x 1  = 0 2 1 0

11– 2 –1 –33 – –3 11. T  = 1 1 , T121  ==T 121  3 13. 21 , T12  ==T 12  2 ------1 --- 2 2 2 –11 –1

01 –1 15.T  = 01– ,T2x + 1  ==T 2x + 1  1 13 1

010 –1 17. I  = 001 , I121  ==I 121  2 100 0

203 2 5 101 1 12 12 2 19. T  = , T==T  516 4 21–  21–  8 –132 –2– –8

000 0100 20 6 100 21. T  = , T13  ==T 13  23. (a) T  = (b) D  = 0020 11 4 010 0003 001

0000 000 100 0100 0100 100 (c) D T  = 020 (d) T D  = (e) TDT  = (f) DTD  = 0040 0020 020 003 0009 0003 003

0000 0 0130 3 2 3 2 3 25. D  = , D5x + 3x  ==D 5x + 3x  27. Tab = –43a + b 2a 0004 5 0000 0

00 –13 – 29. T is the identity map. 31. LT  ==L T  --2- --1- 3 3 ANSWERS C-7

--1- --1- –1 b –22 – –1 2 2 35. T ax+ b = a --- , T  = , T  = 2 42 –1 –--1- 2

1001 –011 –1– –1ab 1100 –1 111 1 37. T = ab – ad + cc , T  = , T  = cd –12 –11– 111 0 01–0 1 201 1

00 0 0 1 1 00 1 1 0 0 00–1 1–1–0 39. L  = 00 0 1 1 0 01 1 0 0 0 10 0 0 0 0

5.4 Change of Basis and Similar Matrices, page 199.

12 ------2 5 1. 25  ==I 25  3. 23– 1  ==I 23– 1  3 1 --- –1 5

0 –1 2 2 20 20 1 5. 2x ++x 1  ==I 2x ++x 1  1 7. ==I  11 11 –2 1   2

1 1 –1 --- –--- 01– 2 2 9. T  ==I T I  11. T  ==IV T IV 10   000 201

0100 0040 13. T  ==IV T IV 9 15. 011 110 021 19. I   000--- n 2 0000

0 –--1- 3 –011 – 1 25. T ==I T I 0 --- 27. T ==I T I 110  W   V  3  W   V  101 1 1 --- 3

–14 –2– 29. T  ==IW T IV   211 ANSWERS C-8

6.1 Determinants, page 215. 1. 6 3. 6 5. 0 7. 86 9. 9 11. 9 13. 81 15. k  0 , k  1 17. k  01 

6.2 Eigenspaces, page 228. 1.  = –14 ; E–1 = –r 2r r   , E4 = r 3r r  

3. = 45 ; E4 = r 3r r   , E5 = r 2r r  

5. = 28 ; E2 = – 2r + srs rs , E8 = r –2rr r  

7.  = 21  ; E2 = 03 rr0 r   , E1 = 2r03r 4r r   E–1 = 00r 0 r  

9.  = 123 ; E1 = 32 – rr r   , E2 = r00 r   , E3 = 0r 0 r  

11. = 3 ; E3 = rsrs–  rs , E–3 = rr3r r  

13. = 01 – ; E0 = r0 rr r   , E–1 = r00s rs

15. = 0 , E0 = 

17. = 516 – ; E5 = 2rr r   , E–16 = r 4r r  

19. = 42 – ; E4 = 3rr2r r   , E–2 = 3r – 3srs rs

21. = 21 –   3 ; E2 = 3rr34r 9r r   , E–1 = 0 –rr3r r   E3 = 00 2+ 3rr r   , E3 = 00 2– 3rr r  

23. = 1 ; E1 = rx+ r r   , E–1 = –rxr+ r  

25.  = 1 , E1 = –rx2 ++sx r rs , E–1 = rx2 + r r  

27.  = 012 , E0 = rx2 r   , E1 = –2rx3 + rx2 r   , E2 = –2rx2 + rx r  

57r 38r 29.  = 4 ; E4 = r   31.  = 0 ; E0 = V 84r –72r

39. (a) a2 ++d2 –42ad bc  0 (b) a2 ++0d2 –42ad bc = (c) a2 ++d2 –42ad bc  0 ANSWERS C-9

6.3 Diagonalization, page 243.

1. Not diagonalizable. 3. Diagonalizable. 5. Diagonalizable. 7. Diagonalizable. 9. Not diagonalizable. 11. Diagonalizable. 13. Diagonalizable. 15. Diagonalizable. 17. Not diagonalizable. 19. Diagonalizable. 21. Diagonalizable. 23. Diagonalizable. 25. Not diagonalizable. 27. Diagonalizable. 29. Not diagonalizable. 31. Diagonalizable. 33. Not diagonalizable.

6.4 Applications, page 256.

k k 2 15+ 15+ 1. sk = ------– ------5 2 2

k – 1 k – 1 k – 2 k – 2 1 15+ 15– 15+ 15– 3. sk = ------–2------+ ------– 2 ------5 2 2 2 2

k + 1 –2x –3x k – 1 2 1 1 k y1 –c1e – c2e 5. sk = 32 – 2 7. sk = ------++------–1 13. = 6 y –2x –3x 3 2 2 c1e + 2c2e

6x –7x x –x fx – 2c1 + c2 – c3e x c1e + c2e – 7c3e f1x e3x + ex 15. gx = c – 2c e6x 17. y = c e–7x – 2c ex – 3c e–x 19. = 1 3 1 2 3 3x x f2x – e + e hx 6x z –7x x –x c2 + c3e c1e ++c2e c3e

1 7 –--3-e–7x ++---ex ---e–x 5 8 2 8 –---t fx A-concentration of alcohol: 4.8– 2.8e 2 3 3 21. gx = –---e–7x – ex + ---e–x 23. 8 8 5 hx –---t 3 1 1 2 –---e–7x + ---ex – ---e–x B-concentration of alcohol: 3.2+ 2.8e 8 2 8

6.5 Markov Chains, page 270.

A B A .3 .7 1. Regular 3. No 5. Regular 7. .4 .6 9. A B .6 .9 .1 B .4 ANSWERS C-10

A .5 .2 (a) Probability 1 of ending up at A. 1 .5 (0 probability of ending up in B or C) BC(b) Probability 1 of ending up at A. 11. .5 .6 13. (0 probability of ending up in B or C) .4 .3 (c) Probability 1 of ending up at A D (0 probability of ending up in B or C)

-----3- 2 3 10 ------.40 5 8 1 .62 .64 .66 15. 17. 19. --- 29. (a) (b) (c) 31. .22 3 5 2 .39 .36 .34 ------.38 5 8 --1- 5

.17141 .27767 33. (a) 0.247 (b) 0.265 (c) 0.273 (d) .24065 (e-a) 0.281 (e-b) 0.274 (e-c) 0.274 .12204 .18823 (e-d) The initial state of the system has no bearing on the fixed or final state of the system. Inde- pendent of the initial state, eventually (to five decimal places): 17.141%, 17.767%, 24.065%, 12.204%, and 18.823% of the employees will be enrolled in plans A, B, C, D, and E, respectively.

.33333 35. (a) 0.63 (b) 0.62 (c) 0.59 (d-a) 0.50 (d-b) 0.49 (d-c) 0.48 (e) .22222 .44444

14 37. (a) 0.33 (b) 0.25 (c) 14 14 14 ANSWERS C-11

7.1 Dot Product, page 289.

15 9 1. 33 3. 13 5. 15 7. a =  ------9. ac=  b = ------for any c  0 2 2c

–1 31 4 8 6 3 11. cos ------ 42 13. ---0 –---  ---3 --- 83 21 5 5 5 5

15. A normal form: 213  x – 1y – 3 z + 1 = 0 ; a general form: 2xy++3z = 1 ; a vector form: 010 ++r120 –  s031 –  rs   .

17. A normal form: 401  x – 1yz+ 3 = 0 ; a vector form: 1 001 ++r10– 4 s010 rs   . 19. 2 21. ------5529 19 9 –3a – b 23. ------25 (a) ab------ab   (b) 7c–45c c c   (c) 000  21 2

27. 3x ++2y 6z = 6

7.2 Inner Product, page 298.

–1 –19 –1 –18 1. 7 3. 29 5. cos ------ 126 9. 14 11. 77 13. cos ------ 158 721 378

1 –1 –1 19. 6 21. cos ------ 96 25  2x2 –3x + 2dx  10.1 221 0

1 2x2 –3x + – x2 + x – 5 dx  1 27. cos–1------0 ------ 174 29. ex – x 2dx  1.2 1 1 0 2x2 –3x + 2dx – x2 + x – 5 2dx 0 0

 sinxxxcos d   31. sin2xxd  0.08 33. cos–1------–------= 0 –   sin2xxd cos2xxd – – ANSWERS C-12

7.3 Orthogonality, page 310

 00 1 1 1 –1 2 –1 –1 0 1 1 1 134 ------------------------------------12------21– ------1.  3. No 5. 1 1 3 3 3 6 6 6 2 2 2 30 34 5 00 5 --- –--- 3 4

4 x2 x 1 5 x2 2x 1 3 x2 1 7. No 9. ------------++------– ------------– ------+ ------+ ------------– ------33 3 3 66 6 6 22 2

–612r + 11. arb== –31 – r for r   . 13. a = 1 b = 2 15. br=  a = ------for r   . –43 + r

1 10 2 2 1 1 1 1 –7 8 –7 8 –------------ ------01 5 ------------ ------------17.  19.  105 105 105 5 5 5 5 5 226 226 226 226

2 1 1 18 4 1 0 2 0 36 –40 –18 6 ------0 ------ –------------------------ ------------21.  23.  5 5 341 341 341 5 5 5 5 3256 3256 3256 3256

25 12 6 1 2 1 1 2 22 2 2 5x2 ------x2 – ------x – ------------0 00------–------ ------------------25.  27.  29.  31 31 31 5 5 2 2 11 11 211211

x3 x2 2 x3 5x2 2 31. x------++------– ------+ ------33. (a) 010  20– 1 6 6 6 30 30 30

3 4 305 (b) 13– 2 = – ---102 + ---20– 1 + 3010 (c) ------5 5 25

9 35. (a) –2 010 01–01 (b) 4133 –  = 2141 + 2212 –  –  (c) ------5

1 1 37. 418 , 12– 1 = ------55 54 41 + ------–832–64– 23 23 7.4 The Spectral Theorem, page 324

123 21 21 13. 15. 210 17. 13 13 302 Index I-1

A E Abstract Vector Space, 40 Eigenvalues, 218, 224 Algebraic Multiplicity, 238 Eigenvectors, 218, 224 Angle Between Vectors, 281, 296 Eigenspaces, 218, 224 Augmented Matrices, 3 Elementary Operations, 2 Elementary Matrix, 168 B Equivalent Matrices, 3 Basis, 94 Equivalent Systems of Equations, 2 Ordered, 177 Consistent, 14 Bijection, 135 Inconsistent, 14 Euclidean Vector Space, 35 C Expansion Theorem, 90, 100 Cauchy-Schwarz Inequality, 295 Change of Basis Theorem, 193 F Characteristic Equation, 219, 225 Fibonacci Numbers, 246 Characteristic Polynomial, 219, 225 Formula for nth term, 248 Closure Axioms, 40 Function Spaces, 44 Coefficient Matrix,18 Fundamental Theorem of Homogeneous Cofactor, 206 Systems of Equations, 20 Column Space, 156 Composition, 117 G Composition Theorem, 117, 182 Gauss-Jordan Elimination Method, 10 Consistent Systems of Equations, 14 Geometric Multiplicity, 238 Coordinate Vector, 177 Golden Ratio, 248 Counterexample, 13 Grahm-Schmidt Process, 303 Cross Product, 287 H D Homogeneous Systems of Equations, 20 Determinant, 205 Cofactor, 206 I Minor, 206 Idempotent Matrix, 162 Diagonal Matrix, 161, 233 Image, 124 Diagonalizable Matrix, 233 Inconsistent Systems of Equations, 14 Diagonalizable Operator, 233 Inner Product, 292 Dimension, 98 Inner Product Space, 292 Dimension Theorem, 126 Distance between vectors, 294 Distance Between Vectors, 280, 294 Norm of a vector, 294 Dot Product, 279 Invertible Matrix, 164 Inverse Function, 136 Isomorphic Spaces, 139 Isomorphism, 138

I-2 INDEX

K Space, 42 Kernel, 124 Symmetric, 161 Trace, 162 L Transpose, 161 Upper Triangle, 207 Laplace Expansion Theorem, 206 Proof, 212 Leading One, 7 N Linear Combination, 77 n-Tuples, 33 Linear Extension Theorem, 115 Nilpotent Matrix, 162 Linear Independent, 86 Norm, 280, 294 Linear Independence Theorem, 22 Normal Vector, 269 Linear Transformation (map),111 Null Space, 155 Image, 124 Nullity, 124 Kernel, 124 M O Markov Chain, 259 One-To-One Function, 129 Regular, 263 Onto Function, 129 Fundamental Theorem, 264 Ordered Basis, 177 Transitional Diagram, 259 Orthogonal Complement, 290 Transitional Matrix, 259, 261 Orthogonal Matrix, 321 Initial State. 260 Orthogonal Vectors, 282, 302 Fixed State, 261 Orthogonal Projection, 307 Matrix Orthogonal Set of Vectors, 301 Augmented, 3 Orthonormal Matrix, 321 Coefficient, 18 Orthonormal Set of Vectors, 302 Cofactor, 207 Orthonormally Diagonalizable, 322 Column Space, 156 P Determinant, 205 Pivoting, 4 Diagonal, 233 Pivot Point, 5 Diagonalizable, 236 Plane Elementary, 168 General Form, 285 Equivalent, 3 Normal Form, 285 Idempotent, 162 Scalar Form, 285 Invertible, 164 Vector Form 71 Minor, 206 Polynomial Space, 43 Multiplication, 151 Properties, 153 R Nilpotent, 162 Rank, 124, 159 Null Space, 155 Recurrence relation, 249 Orthogonal, 321 Reduction Theorem, 100 Orthonormal, 321 Row-Echelon Form, 12 Powers, 159, 167 Row Operations, 2 Rank, 158 Row-Reduced-Echelon Form, 7 Representation of a Linear Map, 179 Row Space, 157 Row Space, 157 Similar, 195 Skew-Symmetric, 162 Index I-3

S U Scalar Product, 34 Unit Vector, 302 Similar Matrices, 195 Upper Triangle Matrix, 207 Skew-Symmetric Matrix, 163 Spanning, 79 V Spanning Theorem, 18 Vector, 31 Spectral Theorem, 319, 322 Abstract, 40 Stochastic Process, 259 Additive Inverse, 31 Subtraction, 55 Addition, 34 Subspaces, 59 Coordinate, 179 Of 2 , 68 Cross Product, 287 Decomposition, 283 3 Of  , 70 Normal, 285 Symmetric Matrix, 161, 315 Scalar product, 34 Symmetric Operator, 317 Sum, 25 Systems of Differential Equations, 249 Subtraction, 55 Systems of Linear Equations, 1 Standard Position, 32 Equivalent, 2 Unit, 302 Elementary Operations, 2 Zero, 35 Homogeneous, 20 Vector Space Abstract, 40 T Euclidean, 35 Theorems Function, 44 Cauchy-Schwarz Inequality, 279 Matrix, 42 Change of Basis, 193 Polynomial, 43 Composition Theorem, 182 Properties, 51, 56 Dimension Theorem, 126 Subspaces, 59 Expansion Theorem, 90, 100 Trivial, 48 Fundamental Theorem of Homogeneous systems of Equations, 20 W Grahm-Schmidt Process, 303 Laplace Expansion Theorem, 206 Weighted Inner Product Spaces, 292 Linear Extension Theorem, 115 Linear Independence Theorem, 22 Z For n , 88 Zero Vector, 35Weigh Reduction Theorem, 100 Spanning Theorem, 1 Spectral Theorem, 319, 322 Vector Decomposition Theorem, 267,306 Trace, 162 Translation Vector, 72 Transpose of a Matrix, 161, 264 , 296 Trivial Space, 48