Nicholson OpenLAWA 2019A PartialStudentSolutionManualsss
Nicholson OpenLAWA 2019A PartialStudentSolutionManualsss
LINEAR ALGEBRA
with Applications
Open Edition
PARTIAL STUDENT
SOLUTION MANUAL
VERSION 2019 – REVISION A
by W. Keith Nicholson
Creative Commons License (CC BY-NC-SA)
a d v a n c i n g l e a r n i n g
All digital forms of access to our high-quality We have been developing superior online for-
open texts are entirely FREE! All content is mative assessment for more than 15 years. Our
reviewed for excellence and is wholly adapt- questions are continuously adapted with the
able; custom editions are produced by Lyryx content and reviewed for quality and sound
for those adopting Lyryx assessment. Access pedagogy. To enhance learning, students re-
to the original source files is also open to any- ceive immediate personalized feedback. Stu-
one! dent grade reports and performance statistics
are also provided.
SUPPORT INSTRUCTOR
SUPPLEMENTS
Access to our in-house support team is avail- Additional instructor resources are also freely
able 7 days/week to provide prompt resolution accessible. Product dependent, these supple-
to both student and instructor inquiries. In ad- ments include: full sets of adaptable slides and
dition, we work one-on-one with instructors to lecture notes, solutions manuals, and multiple
provide a comprehensive system, customized choice question banks with an exam building
for their course. This can include adapting the tool.
text, managing multiple sections, and more!
BE A CHAMPION OF OER!
Contribute suggestions for improvements, new content, or errata:
A new topic
A new example
An interesting new question
A new or better proof to an existing theorem
Any other suggestions to improve the material
Contact Lyryx at [email protected] with your ideas.
CONTRIBUTIONS
Author
W. Keith Nicholson, University of Calgary
LICENSE
Creative Commons License (CC BY-NC-SA): This text, including the art and illustrations, are available
under the Creative Commons license (CC BY-NC-SA), allowing anyone to reuse, revise, remix and
redistribute the text.
To view a copy of this license, visit http://creativecommons.org/licenses/by-nc-sa/4.0/
a d v a n c i n g l e a r n i n g
2 Matrix Algebra 13
2.1 Matrix Addition, Scalar Multiplication, and Transposition . . . . . . . . . . . . . . . . . . 13
2.2 Matrix-Vector Multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3 Matrix Multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.4 Matrix Inverses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.5 Elementary Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.6 Matrix Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.7 LU-factorization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.8 An Application to Input-Output Economic Models . . . . . . . . . . . . . . . . . . . . . 36
2.9 An Application to Markov Chains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Supplementary Exercises: Chapter 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4 Vector Geometry 59
4.1 Vectors and Lines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
iii
iv CONTENTS
6 Vector Spaces 89
6.1 Examples and Basic Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.2 Subspaces and Spanning Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
6.3 Linear Independence and Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
6.4 Finite Dimensional Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
6.5 An Application to Polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
6.6 An Application to Differential Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
Supplementary Exercises: Chapter 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
8 Orthogonality 117
8.1 Orthogonal Complements and Projections . . . . . . . . . . . . . . . . . . . . . . . . . . 117
8.2 Orthogonal Diagonalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
8.3 Positive Definite Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
8.4 QR-Factorization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
8.5 Computing Eigenvalues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
CONTENTS v
B Proofs 175
4. Given the equation 4x − 2y + 0z = 1, take y = s and z = t and solve for x: x = 14 (2s + 3). This is the
general solution.
5. a. If a = 0, no solution if b 6= 0, infinitely many if b = 0.
b. If a 6= 0 unique solution x = b/a for all b.
h i
7. b. The augmented matrix is 10 21 01 .
1 1 0 1
d. The augmented matrix is 0 1 1 0 .
−1 0 1 2
1
2 Systems of Linear Equations
h i h i
3 −2 5 3 −2 5
11. b. −12 8 16 → 0 0 36 . The last equation is 0x + 0y = 36, which has no solution.
17. As in the Hint, multiplying by (x2 + 2)(2x − 1) gives x2 − x + 3 = (ax + b)(2x − 1) + c(x2 + 2).
Equating coefficients of powers of x gives equations 2a + c = 1, −a + 2b = −1, −b + 2c = 3.
Solving this linear system we find a = − 19 , b = − 59 , c = 11
9.
19. If John gets $x per hour and Joe gets $y per hour, the two situations give 2x+3y = 24.6 and 3x+2y =
23.9. Solving gives x = $4.50 and y = $5.20.
3. b. The matrix is already in reduced row-echelon form. The nonleading variables are parameters;
x2 = r, x4 = s and x6 = t.
The first equation is x1 − 2x2 + 2x4 + x6 = 1, whence x1 = 1 + 2r − 2s − t.
The second equation is x3 + 5x4 − 3x6 = −1, whence x3 = −1 − 5s + 3t.
The third equation is x5 + 6x6 = 1, whence x5 = 1 − 6t.
d. First carry the matrix to reduced row-echelon form.
" 1 −1 2 4 6 2 # " 1 0 4 5 5 1
# " 1 0 4 0 5 −4
#
0 1 2 1 −1 −1 0 1 2 1 −1 −1 0 1 2 0 −1 −2
0 0 0 1 0 1 → 0 0 0 1 0 1 → 0 0 0 1 0 1
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
1.2. Gaussian Elimination 3
Hence x = − 17 , y = − 37 .
d. hNote that the variables
i h in the second
i h equation are
i in the wrong order.
3 −1 2 3 −1 2 1 − 31 2
−6 2 −4 → 0 0 0 → 0 0 0 . 3
6. b. Label the rows of the augmented matrix as R1 , R2 and R3 , and begin the gaussian algorithm on
the augmented matrix keeping track of the row operations:
1 2 −3 −3 R1 1 2 −5 5 R2
1 3 −5 5 R2 → 0 1 −2 8 R2 − R1
1 −2 5 −35 R3 0 −4 8 −32 R3 − R1
At this point observe that R3 − R1 = −4(R2 − R1 ), that is R3 = 5R1 − 4R2 . This means that
equation 3 is 5 times equation 1 minus 4 times equation 2, as is readily verified. (The solution
is x1 = t − 11, x2 = 2t + 8 and x3 = t.)
" 1 −1 1 −1 0 # " 1 −1 1 −1 0 # " 1 −1 1 −1 0 #
−1 1 1 1 0 0 0 2 0 0 0 1 −1 1 0
7. b. 1 1 −1 1 0 → 0 2 −2 2 0 → 0 0 1 0 0
1 1 1 1 0 0 2 0 2 0 0 1 0 1 0
1 0 0 0 0 1 0 0 0 0
" # " #
0 1 −1 1 0 0 1 0 1 0
→ 0 0 1 0 0 → 0 0 1 0 0 . Hence x4 = t; x1 = 0, x2 = −t, x3 = 0.
0 0 1 0 0 0 0 0 0 0
4 Systems of Linear Equations
1 1 2 −1 4 1 1 2 −1 4 1 0 7 −7 8
" # " # " #
0 3 −1 4 2 0 3 −1 4 2 0 0 14 −14 14
d. 1 2 −3 5 0 → 0 1 −5 6 −4 → 0 1 −5 6 −4
1 1 −5 6 −3 0 0 −7 7 −7 0 0 −7 7 −7
1 0 7 −7 8 1 0 0 0 1 1 0 0 0 1
" # " # " #
0 1 −5 6 −4 0 1 −5 6 −4 0 1 0 1 1
→ 0 0 14 −14 14 → 0 0 1 −1 1 → 0 0 1 −1 1 .
0 0 −7 7 −7 0 0 0 0 0 0 0 0 0 0
Hence x4 = t; x1 = 1, x2 = 1 − t, x3 = 1 + t.
h i h i
8. b. a1 2b −15 → 10 2 −bab 5−1 +a .
−2−5b
h i
1 b −1 1 0
Case 1 If ab 6= 2, it continues → 0 1 5+a → 2−ab
5+a
.
2−ab 0 1 2−ab
Case3 If a 6= 1 and a 6=
0, there is a uniquesolution:
1 − a1
1 a −1 1 1 a −1 1 1 0 0
0 a−1 0 0 → 0 1 0 0
1
→ 0 1 0 0 .
0 0 a −1 0 0 1 −a 0 0 1 − 1a
Hence x = 1 − a1 , y = 0, z = − 1a .
h i h i
2 1 −1 3 1 12 − 21 3
10. b. 0 0 0 0 → 0 0 0 0 ; rank is 1.
2
f. False. The system 2x−y = 0, −4x+2y = 0 is consistent, but the system 2x−y = 1, −4x+2y =
1 is not consistent.
h. True. A has 3 rows so there can be at most 3 leading 1’s. Hence the rank of A is at most 3.
1 a b+c 1 a b+c
14. b. We begin the row reduction 1 b c + a → 0 b − a a − b . Now one of b − a and c −
1 c a+b 0 c−a a−c
is nonzero(by hypothesis) so that row provides the second leading 1 (its row becomes
a
0 1 −1 ). Hence further row operations give
1 a b+c 1 0 b+c+a
→ 0 1 −1 → 0 1 −1
0 0 0 0 0 0
1 1 1 −2 1 1 1 −2 1 1 1 −2
5 −3 1 −34 → 0 −8 −4 −24 → 0 1 1
2 3
3 3 −1 18 0 0 −4 24 0 0 1 −6
" 1
# " #
1 0 2 −5 1 0 0 −2
→ 0 1 1
2 3 → 0 1 0 6 .
0 0 1 −6 0 0 1 −6
Hence a = −2, b = 6, c = −6, so the equation is x2 + y2 − 2x + 6y − 6 = 0.
18. Let a, b and c denote the fractions of the student population in Clubs A, B and C respectively. The
4 2
new students in Club A arrived as follows: 10 of those in Club A stayed; 10 of those in Club B go to
2
A, and 10 of those in C go to A. Hence
4 2 2
a= 10 a + 10 b + 10 c
5 1 6
c= 10 a + 10 b + 10 c
Hence
−6a + 2b + 2c = 0
a − 3b + 2c = 0
5a + b − 4c = 0
" #
1 0 − 85 0
−6 2 2 0 1 −3 2 0 1 −3 2 0
1 −3 2 0 → 0 −16 14 0 → 0 1 − 87 0 → 0 1 − 87 0 .
5 1 −4 0 0 16 −14 0 0 0 0 0 0 0 0 0
1 2 1 0 1 2 1 0 1 0 −9 0
2. b. 1 3 6 0 → 0 1 5 0 → 0 1 5 0 .
2 3 a 0 0 −1 a−2 0 0 0 a+3 0
Hence there is a nontrivial solution when a = −3: x = 9t, y = −5t, z = t.
a 1 1 0 1 1 −1 0 1 1 −1 0
d. 1 1 −1 0 → a 1 1 0 → 0 1 − a 1 + a 0 .
1 1 a 0 1 1 a 0 0 0 a+1 0
Hence if a 6= 1 and a 6= −1, there is a unique, trivial solution. The other cases are as follows:
1.3. Homogeneous Equations 7
1 1 −1 0 1 1 0 0
a=1: 0 0 2 0 → 0 0 1 0 ; x = −t, y = t, z = 0.
0 0 2 0 0 0 0 0
1 1 −1 0 1 0 −1 0
a = −1 : 0 2 0 0 → 0 1 0 0 ; x = t, y = 0, z = t.
0 0 0 0 0 0 0 0
so the general solution is a = −t, b = 3 + t and c = t. Taking t = −1 gives the linear combina-
tion v = a + 2y − z.
4. b. We must determine if x, y and z exist such that y = xa1 + ya2 + za3 . Equating entries here gives
equations −x + 3y + z = −1, 3x + y + z = 9, 2y + z = 2 and x + z = 6. Carrying the coefficient
matrix to reduced form gives
" −1 3 1 −1 # " 1 0 0 2 #
3 1 1 9 0 1 0 −1
0 2 1 2 → 0 0 1 4
1 0 1 6 0 0 0 0
T
Hence the general solution x = x1 x2 x3 x4 x5 is
−t 0 −1
2s + 3t 2 3
x= s = s 1 +t 0
t 0 1
0 0 0
f2 = 60 − f4 − f7
f3 = −75 + f4 + f6
f5 = 40 − f6 + f7
Hence I1 = − 15 , I2 = 3
5 and I3 = 4
5.
1 0 0 0 −1 −1 0 1 0 0 0 0 −2 1
0 1 0 0 −2 0 −2 0 1 0 0 0 −2 0
0 0 1 0 1 0 2 0 0 1 0 0 1 1
→
0 0 0 1 −2 −1 −2
→
0 0 0 1 0 −3 0
0 0 0 0 1 −1 1 0 0 0 0 1 −1 1
0 0 0 0 3 1 5 0 0 0 0 0 4 2
1 0 0 0 0 0 2
0 1 0 0 0 0 1
1
0 0 1 0 0 0
→ 2
. Hence I1 = 2, I2 = 1, I3 = 12 , I4 = 23 , I5 = 32 , I6 = 21 .
3
0 0 0 1 0 0 2
3
0 0 0 0 1 0
2
1
0 0 0 0 0 1 2
2. Suppose xNH3 + yCuO → zN2 + wCu + vH2 O where x, y, z, w and v are positive integers. Equating
the number of each type of atom on each side gives
N : x = 2z Cu : y = w
H : 3x = 2v O:y=v
Taking v = t these give y = t, w = t, x = 23 t and z = 12 x = 13 t. The smallest value of t such that there
are all integers is t = 3, so x = 2, y = 3, z = 1 and v = 3. Hence the balanced reaction is
1. b. No. If the corresponding planes are parallel and distinct, there is no solution. Otherwise they
either coincide or have a whole common line of solutions.
6 6 16
" 1 4 −1 1 2 # " 1 4 −1 1 2
# 1 0 10 10 10
0 1 −4 1 1
2. b. 31 −62 13 20 51 → 00 −10 −10
4 −1 −1
4 −1 −1 → 0 0 010 100 100 .
1 14 −5 2 3 0 10 −4 1 1
0 0 0 0 0
1
Hence x3 = s, x4 = t are parameters, and the equations give x1 = 10 (16 − 6s − 6t) and x2 =
1
10 (1 + 4s − t).
1 1 3 a 1 1 3 a 1 1 3 a
3. b. a 1 5 4 → 0 1 − a 5 − 3a 4 − a2 → 0 1 − a 5 − 3a 4 − a22 .
1 a 4 a 0 a−1 1 0 0 0 3(2 − a) 4 − a
1 1 3 1 1 1 3 0
If a = 1 the matrix is 0 0 2 3 → 0 0 1 1 , so there is no solution.
0 0 3 3 0 0 0 1
1.6. An Application to Chemical Reactions 11
1 1 3 2 1 0 2 2
If a = 2 the matrix is 0 −1 −1 0 → 0 1 1 0 , so x = 2 − 2t, y = −t, z = t.
0 0 0 0 0 0 0 0
If a 6= 1 and a 6= 2 there is a unique solution.
" 1 1 3 2 −a+4 −5a+8
# " #
1 1 3 a a 1 0 a−1 a−1
1 0 0 3(a−1)
2 a2 −4
0 1−a 5 − 3a 4 − a2
2
→ 0 1 3a−5 a−1
a −4
a−1 → 0 1 3a−5
a−1 a−1 → 0 1 0 −a−2
3(a−1)
.
0 0 3(2 − a) 4−a a+2 a+2 a+2
0 0 1 3 0 0 1 3 0 0 0 3
8−5a −a−2 a+2
Hence x = 3(a−1)
, y= 3(a−1)
, z= 3 .
4. If R1 and R2 denote the two rows, then the following indicate how they can be interchanged using
row operations of the other two types:
R1 R1 + R2 R1 + R2 R2 R2
→ → → →
R2 R2 −R1 −R1 R1
Note that only one row operation of Type II was used — a multiplication by −1.
3 − a + 2c = 0 a − 2c = 3
3b − c − 6 = 1 that is 3b − c = 9
3a − 2 + 2b = 5 3a + 2b = 7
This
system oflinear
equations for a, b and
c has uniquesolution:
1 0 −2 3 1 0 −2 3 1 0 −2 3 1 0 −2 3 1 0 0 1
0 3 −1 7 → 0 3 −1 7 → 0 1 −7 9 → 0 1 −7 9 → 0 1 0 2 .
3 2 0 7 0 2 6 −2 0 2 6 −2 0 0 20 −20 0 0 1 −1
Hence a = 1, b = 2, c = −1.
1 1 1 5 1 1 1 5 1 1 1 5 1 0 0 2
8. 2 −1 −1 1 → 0 −3 −3 −9 → 0 1 1 3 → 0 1 1 3 .
−3 2 2 0 0 5 5 15 0 0 0 0 0 0 0 0
d. [ 3 −1 2 ] − 2 [ 9 3 4 ] + [ 3 11 −6 ] = [ 3 −1 2 ]−[ 18 6 8 ]+[ 3 11 −6 ]
= [ 3 − 18 + 3 −1 − 6 + 11 2 − 8 − 6 ] = [ −12 4 −12 ]
2 T
0 −1 0 1 −2
f. 1 0 −4 = −1 0 4
−2 4 0 2 −4 0
h iT h i h i h i h i h i h i
2 1 1 −1 2 −1 1 −1 6 −3 2 −2 4 −1
h. 3 −1 0 −2 2 3 =3 1 0 −2 2 3 = 3 0 − 4 6 = −1 −6
h i h i
3 −1 15 −5
3. b. 5C − 5 2 0 = 10 0
13
14 Matrix Algebra
4X + 3Y = A
6. b. Given , subtract the first from the second to get X +Y = B − A. Now subtract 3
5X + 4Y = B
times this equation from the first equation: X = A − 3(B − A) = 4A − 3B. Then X +Y = B − A
gives Y = (B − A) − X = (B − A) − (4A − 3B) = 4B − 5A.
Note that this also follows from the Gaussian Algorithm (with matrix constants):
h i h i h i
4 3 A 5 4 B 1 1 B−A
5 4 B → 4 3 A → 4 3 A
h i h i
1 1 B−A 1 0 4A − 3B
→ 0 −1 5A − 4B → 0 1 4B − 5A
7. b. Given 2X − 5Y = 1 2 let Y = T where T is an arbitrary 1 × 2 matrix. Then 2X = 5T +
1 2 so X = 25 T + 12 1 2 , Y = T . If T = s t , this gives X = 52 s + 12 25 t + 1 ,
Y = s t , where s and t are arbitrary.
p + q + r = a
q + s = b
r + s = c
p = d
11. b. A + A′ = 0
−A + (A + A′ ) = −A + 0 (add − A to both sides)
(−A + A) + A′ = −A + 0 (associative law
0 + A′ = −A + 0 (definition of − A)
A′ = −A (property of 0)
···
a1 0 0
b1 0 ··· 0
0 a2 ··· 0 0 b2 ··· 0
13. b. If A = .. .. .. and B = .. .. .. ,
. . . . . .
0 0 ··· an 0 0 ··· bn
a1 − b1 0 ··· 0
0 a2 − b2 ··· 0
then A − B = .. .. .. so A − B is also diagonal.
. . .
0 0 ··· an − bn
h i
s t
14. b. st 1 is symmetric if and only if t = st; that is t(s − 1) = 0; that is s = 1 or t = 0.
2.2. Matrix-Vector Multiplication 15
3. b. By Definition 2.4:
h i x1
h i h i h i h i
1 2 3 1 2 3 x1 + 2x2 + 3x3
Ax = 0 −4 5 x2 = x1 0 + x2 −4 + x3 5 = −4x2 + 5x3
x3
h i x1
h i h i
1 2 3 1 · x1 + 2 · x2 + 3 · x3 x1 + 2x2 + 3x3
By Theorem 2.2.5: Ax = 0 −4 5
x2 = 0 · x1 + (−4) · x2 + 5 · x3 = −4x2 + 5x3
x3
16 Matrix Algebra
6. To say that x0 and x1 are solutions to the homogeneous system Ax = 0 of linear equations means
simply that Ax0 = 0 and Ax1 = 0. If sx0 + tx1 is any linear combination of x0 and x1 , we compute:
using Theorem 2.2.2. This shows that sx0 + tx1 is also a solution to Ax = 0.
h i h i 2
1 1 −1
f. False. If A = 2 and x =
2 then Ax = 14 , and this is not a linear combination
0
0
1
h i h i h i
of 12 and 12 because it is not a scalar multiple of 12 .
h i 1 h i
1 −1
h. False. If A = −1 1 −1 , there is a solution 2 for b = 00 . But there is no solution
1
1
h i h i x h i
for b = 10 . Indeed, if −11 −11 −11 y = 10 then x − y + z = 1 and −x + y − z = 0.
z
This is impossible.
h i h i
11. b. If xy is reflected in the line y = x the result is xy ; see the diagram for Example 2.4.12. In
h i h i h ih i h i
other words, T xy = yx = 01 10 x
y . So T has matrix 0 1
1 0 .
h i h i
d. If xy is rotated clockwise through π2 the result is −xy ; see Example 2.2.14. Hence
h i h i h ih i h i
x y 0 1 x 0 1
T y = −x = −1 0 y so T has matrix −1 0 .
x
13. b. The reflection of y in the yz-plane keeps y and z the same and negates x. Hence
z
x −x −1 0 0 x −1 0 0
T y = y = 0 1 0 y , so the matrix is 0 1 0 .
z z 0 0 1 z 0 0 1
16. Write A = a1 a2 · · · an where ai is column i of A for each i. If b = x1 a1 + x2 a2 + · · · + xn an
T
where the xi are scalars, then Ax = b by Theorem 2.2.1 where x = x1 x2 · · · xn ; that is x is
a solution to the system Ax = b.
18. b. We are given that x1 and x2 are solutions to Ax = 0; that is Ax1 = 0 and Ax2 = 0. If t is any
scalar then, by Theorem 2.2.2, A(tx1 ) = t(Ax1 ) = t0 = 0. That is, tx1 is a solution to Ax = 0.
T
22. Let A = a1 a2 · · · an where ai is column i of A for each i, and write x = x1 x2 · · · xn
T
and y = y1 y2 · · · yn . Then
T
x+y = x1 + y1 x2 + y2 · · · xn + yn
Hence we have
2a + a1 = 7 2b + b1 = 2
−a + 2a1 = −1 −b + 2b1 = 4
h ih i h i h i 1 0
1 −1 −9 −16 −14 −17 −2 −1 −2
5. b. A(BC) = 0 1 5 1 = 5 1 = 3 1 0
2 1 = (AB)C
5 8
h i h i h i h i h i
a b 0 0 0 0 b 0 0 0
6. b. If A = c d then A 1 0 = 1 0 A becomes d 0 = a b whence b = 0 and
h i
a 0
a = d. Hence A has the form A = c a , as required.
h ih i h i h i
P2 0 P2 0 P4 0 P6 0
A4 = 0 Q2 0 Q2 = 0 Q4 , A6 = A4 A2 = 0 Q6 , . . . ; in general we claim that
h i
P2k 0
A2k = 0 Q2k for k = 1, 2, . . . (∗)
h i
1 −m
Pm = 0 1 for m = 1, 2, . . . (∗∗)
Finally
1 −(2k + 1) 2 −1
" #
h ih i h i
P2k 0 P X P2k+1 P2k X 0 1 0 0
A2k+1 = A2k · A = 0 I 0 Q = 0 Q = 0 0 −1 1
0 0 0 1
h ih i h i h i
I X I −X I 2 + X0 −IX + XI I 0
13. b. 0 I 0 I = 0I + I0 −0X + I 2 = 0 I = I2k
h i
−X T
d. [ I XT ][ −X I ]T = [ I XT ] I = −IX T + X T I = Ok
h i2 h ih i h i
f. 0I X
0 = 0
I
X
0
0
I
X
0 = X
0
0
X
h i3 h ih i h i
0 X 0 X X 0 0 X2
I 0 = I 0 0 X = X 0
h i4 h ih i h 2 i
0 0 X2
I
X
0 = I
X
0
0
X 0 = X0 X02
h i2m h m i
Continue. We claim that 0I X
0 = X0 0
Xm for m ≥ 1. It is true if m = 1 and, if it holds
for some m, we have
h i2(m+1) h i2m h i2 h ih i h i
0 X Xm X m+1
I 0 = 0I X
0
0
I
X
0 = 0
0
Xm
X
0
0
X = 0
0
X m+1
18. b. We are given that AC = CA, so (kA)C = k(AC) = k(CA) = C(kA), using Theorem 2.3.3. Hence
kA commutes with C.
20. Since A and B are symmetric, we have AT = A and BT = B. Then Theorem 2.3.3 gives (AB)T =
BT AT = BA. Hence (AB)T = AB if and only if BA = AB.
a x y
22. b. Let A = x b z . Then the entries on the main diagonal of A2 are a2 + x2 + y2 , x2 + b2 + z2 ,
y z c
y2 + z2 + c2 . These are all zero if and only if a = x = y = b = z = c = 0; that is if and only if
A = 0.
24. If AB = 0 where A 6= 0, suppose BC = I for some matrix C. Left multiply this equation by A to get
A = AI = A(BC) = (AB)C = 0C = 0, a contradiction. So no such matrix C exists.
" 1 0 1 0 # " 2 1 1 1 #
1 0 0 1 3 0 2 2
26. We have A = 0 0 0 1 , and hence A3 = 2 0 1 1 . Hence there are 3 paths of length 3
1 1 0 0 3 1 2 1
from v1 to v4 because the (4, 1)-entry of A3 is 3. Similarly, the fact that the (3, 2)-entry of A3 is 0
means that there are no paths of length 3 from v2 to v3 .
h i
27. b. False. If A = 10 00 = J then AJ = A, but J 6= I.
d. True. Since A is symmetric, we have AT = A. Hence Theorem 2.1.2 gives (I +A)T = I T +AT =
I + A. In other words, I + A is symmetric.
h i
f. False. If A = 00 10 then A 6= 0 but A2 = 0.
h. True. We are assuming that A commutes with A + B, that is A(A + B) = (A + B)A. Multiplying
out each side, this becomes A2 + AB = A2 + BA. Subtracting A2 from each side gives AB = BA;
that is A commutes
h with i B. h i
j. False. Let A = 21 42 and B = −21 −24 . Then AB = 0 is the zero matrix so both columns
are zero. However
h B has ino zero column.
h i
−2
l. False. Let A = −21 4 and B = 2
1
4
2 as above. Again AB = 0 has both rows zero, but A
has no row of zeros.
28. b. If A = [ai j ] the sum of the entries in row i is ∑nj=1 ai j = 1. Similarly for B = [bi j ]. If AB =
C = [ci j ] then ci j is the dot product of row i of A with column j of B, that is ci j = ∑nk=1 aik bk j .
Hence the sum of the entries in row i of C is
!
n n n n n n
∑ ci j = ∑ ∑ aik bk j = ∑ aik ∑ bk j = ∑ aik · 1 = 1
j=1 j=1 k=1 k=1 j=1 k=1
2.3. Matrix Multiplication 21
Easier Proof: Let X be the n × 1 column matrix with every entry equal to 1. Then the entries
of AX are the row sums of A, so these all equal 1 if and only if AX = X . But if also BX = X
then (AB)X = A(BX ) = AX = X , as required.
30. b. If A = [ai j ] then the trace of A is the sum of the entries on the main diagonal, that is tr A =
a11 + a22 + · · · + ann . Now the matrix kA is obtained by multiplying every entry of A by k, that
is kA = [kai j ]. Hence
e. If A = [ai j ] the transpose AT is obtained by replacing each entry ai j by the entry a ji directly
across the main diagonal. Hence, write AT = [a′i j ] where a′i j = a ji for all i and j. Let bi denote
the (i, i)-entry of AAT . Then bi is the dot product of row i of A and column i of AT , that is
bi = ∑nk=1 aik a′ki = ∑nk=1 aik aik = ∑nk=1 a2ik . Hence we obtain
!
n n n n n
tr (AAT ) = ∑ bi = ∑ ∑ a2ik =∑ ∑ a2ik
i=1 i=1 k=1 i=1 k=1
Similarly,
36. See the article in the mathematics journal Communications in Algebra, Volume 25, Number 7
(1997), pages 1767 to 1782.
2. In each case we need row operations that carry A to I; these same operations carry I to A−1 . In short
A I → I A−1 . This is called the matrix inversion algorithm.
h.
We begin by subtracting
row 2 from twice row1:
3 1 −1 1 0 0 1 0 −2 2 −1 0 1 0 −2 2 −1 0
5 2 0 0 1 0 → 5 2 0 0 1 0 → 0 2 10 −10 6 0
1 1 −1 0 0 1 1 1 −1 0 0 1 0 1 1 −2 1 1
2
" − 24
# " #
1 0 −2 2 −1 0 1 0 −2 2 −1 0 1 0 0 4 0
→ 0 1 5 −5 3 0 → 0 1 5 −5 3 0 → 0 1 0 − 45 2
4
5
4 .
0 0 −4 3 −2 1 0 0 1 − 43 2
− 14
4 0 0 1 − 43 2
4 − 14
2 0 −2
1
Hence A−1 = 4
−5 2 5 .
−3 2 −1
−1 4 5 2 1 0 0 0 1 −4 −5 −2 −1 0 0 0
" # " #
0 0 0 −1 0 1 0 0 0 0 0 1 0 −1 0 0
j. 1 −2 −2 0 0 0 1 0 → 0 2 3 2 1 0 1 0
0 −1 −1 0 0 0 0 1 0 1 1 0 0 0 0 −1
1 −4 −5 −2 −1 0 0 0 1 0 −1 −2 −1 0 0 −4
" # " #
0 1 1 0 0 0 0 −1 0 1 1 0 0 0 0 −1
→ 0 2 3 2 1 0 1 0 → 0 0 1 2 1 0 1 2
0 0 0 1 0 −1 0 0 0 0 0 1 0 −1 0 0
2.4. Matrix Inverses 23
1 0 0 0 0 0 1 −2 1 0 0 0 0 0 1 −2
" # " #
0 1 0 −2 −1 0 −1 −3 0 1 0 0 −1 −2 −1 −3
→ 0 0 1 2 1 0 1 2 → 0 0 1 0 1 2 1 2 .
0 0 0 1 0 −1 0 0 0 0 0 1 0 −1 0 0
0 0 1 −2
" #
−1 −2 −1 −3
Hence A−1 = 1 2 1 2 .
0 −1 0 0
1 2 0 0 0 1 0 0 0 0 1 0 0 0 0 1 −2 6 −30 210
0 1 3 0 0 0 1 0 0 0 0 1 0 0 0 0 1 −3 15 −105
l. 0 0 1 5 0 0 0 1 0 0 → 0 0 1 0 0 0 0 1 −5 35 .
0 0 0 1 1 0 0 0 1 0 0 0 0 1 0 0 0 0 1 −7
0 0 0 0 1 0 0 0 0 1 0 0 0 0 1 0 0 0 0 1
1 −2 6 −30 210
0 1 −3 15 −105
Hence A−1 = 0 0 1 −5 35 .
0 0 0 1 −7
0 0 0 0 1
h i h i h i
3. b. The equations are Ax = b where A = 21 −3−4 , x = x
y , b = 0
1 . We have (by the algo-
4 −3
rithm or Example 2.4.4) A−1 = 51 . Left multiply Ax = b by A−1 to get
1 −2
h ih i h i
x = A−1 Ax = A−1 b = 15 41 −3 −2
0
1 = 1 −3
5 −2
Hence x = − 35 and y = − 52 .
1 4 2 x 1
d. Here A = 2 3 3 , x = y , b = −1 .
4 1 4 0
z
9 −14 6
−1 1
By the algorithm, A = 5 4 −4 1 .
−10 15 −5
" 23 #
9 −14 6 1 5
1
5
4 −4 1 −1 = 8
5
−10 15 −5 0 −5
23
Hence x = 5, y = 85 , and z = − 25
5 = −5.
1 −1 2
4. b. We want B such that AB = P where P = 0 1 1 . Since A−1 exists left multiply this
1 0 0
equation by A−1 to get B = A−1 (AB) = A−1 P. [This B will satisfy our requirements because
AB = A(A−1 P) = IP = P]. Explicitly
1 −1 3 1 −1 2 4 −2 1
−1
B=A P= 2 0 5 0 1 1 = 7 −2 4
−1 1 0 1 0 0 −1 2 −1
h i−1 h i
1 −1 1 3 1
5. b. By Example 2.4.4, we have (2A)T = 2 3 = 5 −2 1 . Since (2A)T = 2AT , we get
h i h i
2AT = 15 −23 11 so AT = 101 3
−2
1
1 . Finally
h iT h i
T T 1 3 1 1 3 −2
A = (A ) = 10 −2 1 = 10 1 1
24 Matrix Algebra
h i
2 1
d. We have (I − 2AT )−1 = 1 1 so (because (U −1 )−1 = U for any invertible matrix U )
h i−1 h i
2 1 1 −1
(I − 2AT ) = 1 1 = −1 2
h i h i h i h i
1 −1 1 0 1 −1 0 1
Thus 2AT =I− −1 2 = 0 1 − −1 2 = 1 −1 .
h i
1 0 1
This gives AT = 2 1 −1 , so
h iT h i
1 0 1 1 0 1
A = (AT )T = 2 1 −1 = 2 1 −1
h i −1 h i
1 0 1 0
f. Given 2 1 A = 2 2 , take inverses to get
h i h i−1 h i
1 0 1 0 1 2 0
2 1 A= 2 2 = 2 −2 1
h i−1 h i
1 0 1 0
Now 2 1 = −2 1 , so left multiply by this to obtain
h i h i h i
1 0 1 2 0 1 2 0
A= −2 1 2 −2 1 = 2 −6 1
h i
1 1
h. Given (A−1 − 2I)T = −2 1 0 , take transposes to get
h iT h iT h i
1 1 1 1 1 1
A−1 − 2I = −2 1 0 = −2 1 0 = −2 1 0
h i h i h i h i h i
1 1 2 0 1 1 0 −2 0 −1
Hence A−1 = 2I − 2 1 0 = 0 2 −2 1 0 = −2 2 =2 −1 1 . Finally
h i−1 h i−1 h i h i
−1 −1
A = (A−1 )−1 = 2 0
−1 1 = 1
2
0
−1 1 = 1
2
1
−1
1
1
1
0 = −1
2
1
1
1
0
−1
0 1 −1 2 −1 3
1
6. b. Have A = (A−1 )−1 = 1 2 1 = 2
0 1 −1 by the algorithm.
1 0 1 −2 1 −1
h i h
h i ih h i h i i
x′
8. b. The equations are A x
y where A = 34 45 and B = −54
= 7
1 and x
y =B y′
4
−3 .
h i h i h i
x′
Thus B = A−1 (by Example 2.4.4) so the substitution gives 71 = A xy = AB y′ =
h ′ i h ′ i h i h i h ih i h i
x x ′ ′ x 7 −5 4 7 −31
I y′ = y′ . Thus x = 7, y = 1 so y = B 1 = 4 −3 1 = 25 .
h i h i h i
1 0 1 0 2 0
9. b. False. A = 0 1 and B = 0 −1 are both invertible, but A + B = 0 0 is not.
18. b. Suppose column j of A consists of zeros. Then Ay = 0 where y is the column with 1 in the
position j and zeros elsewhere. If A−1 exists, left multiply by A−1 to get A−1 Ay = A−1 0, that
is Iy = 0; a contradiction. So A−1 does not exist.
d. If each column of A sums to 0, then xA = 0 where x is the row of 1s. If A−1 exists, right
multiply by A−1 to get xAA−1 = 0A−1 , that is xI = 0, x = 0, a contradiction. So A−1 does not
exist.
2 1 −1
19. b. (ii) Write A = 1 1 0 . Observe that row 1 minus row 2 minus row 3 is zero. If x =
1 0 −1
1 −1 −1 , this means xA = 0. If A−1 exists, right multiply by A−1 to get xAA−1 =
24. b. The condition can be written as A(A3 + 2A2 − I) = 4I, whence A[ 14 (A3 + 2A2 − I)] = I. By
Corollary 2.4.1 of Theorem 2.4.5, A is invertible and A−1 = 41 (A3 + 2A2 − I). Alternatively,
this follows directly by verifying that also [ 14 (A3 + 2A2 − I)]A = I.
25. b. If Bx = 0 then (AB)x = 0 so x = 0 because AB is invertible. Hence B is invertible by Theorem
2.4.5. But then A = (AB)B−1 is invertible by Theorem 2.4.4 because both AB and B−1 are
invertible.
26 Matrix Algebra
h 2 −1
i
b. As in Example 2.4.11, −B−1YA−1 = −(−1)−1 1 3
26. −5 3 =[ −13 8 ], so
h ih ih i h i
1 −1 5 2 −2 1 −14 8
d. As in Example 2.4.11, −A−1 X B−1 = − −1 2 −1 0 −1 1 = 16 −9 , so
31. b. If A = B then A−1 B = A−1 A = I. Conversely, if A−1 B = I left multiply by A to get AA−1 B = AI,
IB = A, B = A.
32. a. Since A commutes with C, we have AC = CA. Left-multiply by A−1 to get C = A−1CA. Then
right-multiply by A−1 to get CA−1 = A−1C. Thus A−1 commutes with C too.
33. b. The condition (AB)2 = A2 B2 means ABAB = AABB. Left multiplication by A−1 gives BAB =
ABB, and then right multiplication by B−1 yields BA = AB.
34. Assume that AB is invertible; we apply Part 2 of Theorem 2.4.5 to show that B is invertible. If
Bx = 0 then left multiplication by A gives ABx = 0. Now left multiplication by (AB)−1 yields
x = (AB)−1 0 = 0. Hence B is invertible by Theorem 2.4.5. But then we have A = (AB)B−1 so A is
invertible by Theorem 2.4.4 (B−1 and AB are both invertible).
−1
35. b. By the hint, Bx = 0 where x = 3 so B is not invertible by Theorem 2.4.5.
−1
2.5. Elementary Matrices 27
36. Assume that A can be left cancelled. If Ax = 0 then Ax = A0 so x = 0 by left cancellation. Thus
A is invertible by Theorem 2.4.5. Conversely, if A is invertible, suppose that AB = AC. Then left
multiplication by A−1 yields A−1 AB = A−1 AC, IB = IC, B = C.
U T = InT − 2(X X T )T = In − 2X T T X T = In − 2X X T = U
U 2 = (In − 2X X T )(I − 2X X T )
= In − 2X X T − 2X X T + 4X X T X X T
= In − 4X X T + 4X Im X T
= In
I = (I − 2P)2 = I − 4P + 4P2
Hence 4P = 4P2 ; so P = P2 .
41. b. If A and B are any invertible matrices (of the same size), we compute:
A−1 (A + B)B−1 = A−1 AB−1 + A−1 BB−1 = B−1 + A−1 = A−1 + B−1
Hence A−1 + B−1 is invertible by Theorem 2.4.4 because each of A−1 , A + B, and B−1 is
invertible. Furthermore
(A−1 + B−1 )−1 = [A−1 (A + B)B−1 ]−1 = (B−1 )−1 (A + B)−1 (A−1 )−1 = B(A + B)−1A
h i
−1 0
2. b. A → B is accomplished by negating row 1, so E = 0 0 .
h i
1 −1
d. A → B is accomplished by subtracting row 2 from row 1, so E = 0 1 .
h i
f. A → B is accomplished by interchanging rows 1 and 2, so E = 01 1
0 .
h i h i h i h i h i
0 1 k 0 1 0 1 k 1 0
3. b. The possibilities for E are 1 0 , 0 1 , 0 k , 0 1 and k 1 . In each case EA
has a row different from C.
h i h i 2 2 2 2
1 0 7 1 12 −2
UA = R = 0 1 −3 where U = 2 −5 1 . This matrix U is the product of the elementary
matrices used at each stage:
h i
1 2 1
5 12 −1 =A
↓
h i h i
1 2 1 1 0
0 2 −6 = E1 A where E1 = −5 1
↓
h i h i
1 2 1 1 0
0 1 −3 = E2 E1 A where E2 = 0 1
2
↓
h i h i
1 0 7 1 −2
0 1 −3 = E3 E2 E1 A where E3 = 0 1
8. b. h i
2 3
1 2 = A
h i ↓ h i
1 2 0 1
2 3 = E1 A where E1 = 1 0
h i ↓ h i
1 2 1 0
0 −1 = E2 E1 A where E2 = −2 1
h i ↓ h i
1 2 1 0
0 1 = E3 E2 E1 A where E3 = 0 −1
h i ↓ h i
1 0 1 −2
0 1 = E4 E3 E2 E1 A where E4 = 0 1
Thus E4 E3 E2 E1 A = I so
A = (E4 E3 E2 E1 )−1
= E1−1 E2−2 E3−1 E4−1
h ih ih ih i
= 01 10 1 0
2 1
1
0
0
−1
1
0
2
1
10. By Theorem 2.5.3, UA = R for some invertible matrix U . Hence A = U −1 R where U −1 is invertible.
h i h i h i h i
12. b. A I = 32 21 10 01 → 12 11 10 −11 → 10 −11 −21 −13 → 10 01 −12 −32
h i
−1 2 −1
so U = 2 −3 . Hence, UA = R = I2 in this case so U = A . Thus, r = rank A = 2 and,
taking V = I2 , UAV = UA = I2 .
1 1 0 −1 1 0 0 1 1 0 −1 1 0 0 1 0 1 3 −2 1 0
d. A I = 3 2 1 1 0 1 0 → 0 −1 1 4 −3 1 0 → 0 1 −1 −4 3 −1 0 .
0 −1 1 4 −1 0 1
1 0 1 3 0 0 1
0 0 0 0 2 −1 1
−2 1 0 1 0 1 3
Hence, UA = R where U = 3 −1 0 and R = 0 1 −1 −4 . Note that rank A = 2.
2 −1 1 0 0 0 0
Next,
1 0 0 1 0 0 0 1 0 0 1 0 0 0 1 0 0 0
" # " # " #
T 0 1 0 0 1 0 0 0 1 0 0 1 0 0 0 1 0 0
so V T =
R I = 1 −1 0 0 0 1 0 → 0 0 0 −1 1 1 0 −1 1 1 0 .
3 −4 0 0 0 0 1 0 0 0 −3 4 0 1 −3 4 0 1
1 0 0
" #
1 0 0 0
0 1 0
Hence, (UAV )T = (RV )T = V T RT = 0 0 0 , so UAV = 0 1 0 0 .
0 0 0 0
0 0 0
I U −1 A = U −1U U −1 A = U −1 U A
(2.1)
30 Matrix Algebra
row-echelon form.
r r r
17. b. A ∼ A because A = IA. If A ∼ B, let A = U B, U invertible. Then B = U −1A so B ∼ A.
r r
Finally if A ∼ B and B ∼ C, let A = U B and B = VC where U and V are invertible. Hence
r
A = U (VC) = (UV )C so A ∼ C.
h i
19. b. The matrices row-equivalent to A = 00 00 01 are the matrices UA where U is invertible. If
h i h i
a b 0 0 b
U = c d then UA = 0 0 d where b and d are not both zero (as U is invertible). Every
h i
such matrix arises — use U = −ba ba — it is invertible as a2 + b2 6= 0 (Example 2.3.5).
1 −1 5
" # " # " #
1 1 −1
2. b. Let a = 1 ,b= 2 and x = 2 . We know T (a) and T (b); to find T (x) we express
1 −4 4
x as a linear combination of a and b, and use the assumption that T is linear. If we write
x = ra + sb, equate entries, and solve the linear equations, we find that r = 2 and s = −3.
Hence x = 2a − 3b so, since T is linear,
5 2 4
T (x) = 2T (a) − 3T (b) = 2 1 −3 0 = 2
−3 1 −9
h i h i
3. b. In R2 , we have e1 = 10 and e2 = 01 . We are given that T (x) = −x for each x in R2 . In
particular, T (e1 ) = −e and T (e2 ) = −e2 . Since T is linear, Theorem 2.6.2 gives
h i
A = T (e1 ) T (e2 ) = −e1 −e2 = −10 −10
2.6. Matrix Transformations 31
h i h i h i h ih i h i
x x −x −1 0 x x
Of course, T y =− y = −y = 0 −1 y in R2 , so in this case
for all y
h i
we can easily see directly that T has matrix −10 −10 . However, sometimes Theorem 2.6.2 is
necessary.
h i h i
d. Let e1 = 0 and e2 = 01 . If these vectors are rotated counterclockwise through π4 , some
1
√ √
2
− 22
simple trigonometry shows that T (e1 ) = √2 and T (e2 ) =
2 √
2
. Since T is linear, the
2 2
h √ √ i
matrix A of T is A = T (e1 ) T (e2 ) = 12 √22 −√22 .
1 0 0
4. b. Let e1 = 0 , e2 = 1 and e3 = 0 denote the standard basis of R3 . Since T : R3 → R3
0 0 1
is reflection in the uz-plane, we have:
T (e1 ) = −e1 because e1 is perpendicular to the uz-plane; while
T (e2 ) = e2 and T (e3 ) = e3 because e2 and e3 are in the uz-plane.
−1 0 0
So A = T (e1 ) T (e2 ) T (e3 ) = −e1 e2 e3 = 0 1 0 .
0 0 1
5. b. Since y1 and y2 are both in the image of T , we have y1 = T (x1 ) for some x1 in Rn , and
y2 = T (x2 ) for some x2 in Rn . Since T is linear, we have
This shows that ay1 + by2 = T (ax1 + bx2 ) is also in the image of T .
Note that T 1 also fails for this transformation T , as you can verify.
h i h i h ih i h i
x x+y 1 1 x
8. √1
b. We are given T y = 2 −x + y = 2 −1 1 √1
y for all xy , so T is the matrix
h i √1 √1
1 1 1
transformation induced by the matrix A = √ −1 1 = − √12 √12 . By Theorem 2.6.4
2 2 2
xz-plane through the angle θ from the x axis to the z axis. By Theorem 2.6.4 the effect of T on
the xz-plane is given by
h i h ih i h i
x cos θ − sin θ x x cos θ − z sin θ
z → sin θ cos θ z = x sin θ + z cos θ
x x cos θ − z sin θ
Hence T 0 = 0 , and so
z x sin θ + z cos θ
x 0 x 0 x cos θ − z sin θ
T y =T y +T 0 = y + 0
z 0 z 0 x sin θ + z cos θ
x cos θ − z sin θ cos θ 0 − sin θ x
= y = 0 1 0 y
x sin θ + z cos θ sin θ 0 cos θ z
cos θ 0 − sin θ
Hence the matrix of T is 0 1 0 .
sin θ 0 cos θ
13. b. Since R has matrix A, we have R(x) = Ax for all x in Rn . By the definition of T we have
T (x) = aR(x) = a(Ax) = (aA)x
for all x in Rn . This shows that the matrix of T is aA.
14. b. We use Axiom T 2: T (−x) = T [(−1)x] = (−1)T (x) = −T (x).
17. b. The matrix of T is B, so T (x) = Bx for all x in Rn . Let B2 = I. Then
T 2 (x) = T [T (x)] = B[Bx] = B2 x = Ix = x = 1R2 (x) for all x in Rn .
Hence T 2 = 1Rn since they have the same effect on every column x.
Conversely, if T 2 = 1Rn then
B2 x = B(Bx) = T (T (x)) = T 2 (x) = 1R2 (x) = x = Ix for all x in Rn .
This implies that B2 = I by Theorem 2.2.6.
2.6. Matrix Transformations 33
h i h i h i h i
1 0 0 1 0 −1 0 −1
18. The matrices of Q0 , Q1 , Q−1 and R π are 0 −1 , 1 0 , −1 0 and 1 0 , respectively.
2
We use Theorem 2.6.3 repeatedly: If S has matrix A and T has matrix B then S ◦ T has matrix AB.
h ih i h i
0 −1
b. The matrix of Q1 ◦ Q0 is 01 10 1
0 −1
0
= 1 0 , which is the matrix of R π .
2
h ih i h i
1 0 0 −1 0 −1
d. The matrix of Q0 ◦ R π is 0 −1 1 0 = −1 0 which is the matrix of Q−1 .
2
19. b. We have Pm [Qm (x)] = Pm (x) for all x in R2 because Qm (x) lies on the line y = mx. This means
Pm ◦ Qm = Pm .
T T
20. To see that T is linear, write x = x1 x2 · · · xn and y = y1 y2 · · · yn . Then:
T
T (x + y) = T x1 + y1 x2 + y2 · · · xn + yn
= (x1 + y1 ) + (x2 + y2 ) + · · · + (xn + yn )
= (x1 + x2 + · · · + xn ) + (y1 + y2 + · · · + yn )
= T (x) + T (y)
T
T (ax) = T ( ax1 ax2 · · · axn )
= ax1 + ax2 + · · · + axn
= a(x1 + x2 + · · · + xn )
= a T (x)
Hence T is linear, so its matrix is A = T (e1 ) T (e2 ) · · · T (en ) = 1 1 · · · 1 by Theorem
2.6.2.
Note that this can be seen directly because
x1 x1
x2 x2
T .
..
= x1 + · · · + xn = 1 1 ··· 1 .
..
xn xn
so we see immediately that T is the matrix transformation induced by 1 1 · · · 1 . Note that
this also shows that T is linear, and so avoids the tedious verification above.
22. b. Suppose that T : Rn → R is linear. Let e1 , e2 , . . . , en be the standard basis of Rn , and write
T (e j ) = w j for each j = 1, 2,. . . , n. Note that each w j is in
R.As T is linear, Theorem
2.6.2
asserts that T has matrix A = T (e1 ) T (e2 ) · · · T (en ) = w1 w2 · · · wn .
x1
x2
Hence, given x = . in Rn , we have
..
xn
x1
x2
T (x) = Ax = w1 w2 · · · wn .
..
= w1 x1 + w2 x2 + · · · + wn xn = w · x = Tw (x)
xn
34 Matrix Algebra
T
for all x in Rn where w = w1 w2 · · · wn . This means that T = Tw . This can also be
T (x) = T (x1 e1 + x2 e2 + · · · + xn en )
= x1 T (e1 ) + x2 T (e2 ) + · · · + xn T (en )
= x1 w1 + x2 w2 + · · · + xn wn
= w·x
= Tw (x)
2.7 LU-factorization
2 4 2 1 2 1 1 2 1 1 2 1
1. b. 1 −1 3 → 1 −1 3 → 0 −3 2 → 0 1 − 32 = U.
−1 7 −7 −1 7 −7 0 9 −6 0 0 0
2 0 0
Hence A = LU where U is above and L = 1 −3 0 .
−1 9 1
−1 −3 1 0 −1 1 3 −1 0 1 1 3 −1 0 1
" # " # " #
1 4 1 1 1 0 1 2 1 0 0 1 2 1 0
d. 1 2 −3 −1 1 → 0 −1 −2 −1 0 → 0 0 0 0 0 = U.
0 −2 −4 −2 0 0 −2 −4 −2 0 0 0 0 0 0
−1 0 0 0
" #
1 1 0 0
Hence A = LU where U is as above and L = 1 −1 1 0 .
0 −2 0 1
2 2 −2 4 2 1 1 −1 2 1 1 1 −1 2 1
" # " # " #
1 −1 0 2 1 0 −2 1 0 0 0 1 − 21 0 0
f. 3 1 −2 6 3 → 0 −2 1 0 0 → 0 0 0 0 0 = U.
1 3 −2 2 1 0 2 −1 0 0 0 0 0 0 0
2 0 0 0
" #
1 −2 0 0
Hence A = LU where U is above and L = 3 −2 1 0 .
1 2 0 1
The elementary
matrices corresponding
(in order) to the interchanges
are
1 0 0 0 1 0 0 0 1
P1 = 0 0 1 and P2 = 1 0 0 , so take P = P2P1 = 1 0 0 .
0 1 0 0 0 1 0 1 0
We apply the LU -algorithm to PA:
−1 2 1 1 −2 −1 1 −2 −1 1 −2 −1
PA = 0 −1 2 → 0 −1 2 → 0 1 −2 → 0 1 −2 =U
0 0 4 0 0 4 0 0 4 0 0 1
−1 0 0
Hence PA = LU where U is as above and L = 0 −1 0 .
0 0 4
d. "
The reduction to#row-echelon
" 1 2 −3form #requires
" 1 2two row interchanges:
−1 −2 −3
3 0 0 0
# " 1 2 −3 0
# " 1 2 −3 0
#
2 4 −6 5 0 0 0 5 0 0 0 5 0 1 −2 −3 0 1 −2 −3
1 1 −1 3 → 0 −1 2 3 → 0 1 −2 −3 → 0 0 0 5 → 0 0 −2 4
2 5 −10 1 0 1 −4 1 0 0 −2 4 0 0 −2 4 0 0 0 5
The elementary matrices corresponding (in #order) to the interchanges are #
1 0 0 0 1 0 0 0 1 0 0 0
" # " "
0 0 1 0 0 1 0 0 0 0 1 0
P1 = 0 1 0 0 and P2 = 0 0 0 1 so P = P2P1 = 0 0 0 1 .
0 0 0 1 0 0 1 0 0 1 0 0
We apply
" the LU -algorithm
# to" PA:
−1 −2 3 0 1 2 −3 0 1 2 −3 0
# " #
1 1 −1 3 0 −1 2 3 0 1 −2 −3
PA = 2 5 −10 1 → 0 1 −4 1 → 0 0 −2 4
2 4 −6 5 0 0 0 5 0 0 0 5
1 2 −3 0 1 2 −3 0
" # " #
0 1 −2 −3 0 1 −2 −3
→ 0 0 1 −2 → 0 0 1 −2 =U
0 0 0 5 0 0 0 1
−1 0 0 0
" #
1 −1 0 0
Hence PA = LU where U is as above and L̇ = 2 1 −2 0 .
2 0 0 5
" x1
#
2 0 0 1 1 0 −1 y1
x2
3. b. Write L = 1 3 0 ,U= 0 1 0 1 , x= x3 , y= y2 . The system Ly = b
−1 2 1 0 0 0 0 y3
x4
2y1 = −2
is y1 + 3y2 = −1 and we solve this by forward substitution: y1 = −1, y2 =
−y1 + 2y2 + y3 = 1
x1 + x2 − x4 = −1
1
3 (−1−y 1 ) = 0, y3 = 1+y1 −2y 2 = 0. The system U x = y is x2 + x4 = 0 and
0 = 0
we solve this by back substitution: x4 = t, x3 = 5, x2 = −x4 = −t, x1 = −1 + x4 − x2 = −1 + 2t.
" 2 # " 8 − 2t #
8 6−t
d. Analogous to (b). The solution is: y = −1 ,x= −1 − t , t arbitrary.
0 t
7. We proceed
h h on niwhere A and B are n × n. It is clear if n = 1. In
iby induction h general, writei
A = X A1 and B = Y B1 where A1 and B1 are lower triangular. Then AB = Xb +abA1Y A10B1
a 0 b 0
Thus the solution to (I − E)p is p1 = p2 = p3 = t. Thus all three industries produce the same output.
h i h i h i
1−a −b 1 − a −b bt
4. I − E = −1 + a b → 0 0 so the possible equilibrium price structures are p = (1 − a)t ,
h i
t arbitrary. This is nonzero for some t unless b = 0 and a = 1, and in that case p = 11 is a solution.
h i
If the entries of A are positive then p = 1 −b a has positive entries.
h i h i
.4 .8 −1 5 8 8
7. b. One such example is E = .7 .2 , because (I − E) = − 4 7 6 .
h i h i
−a −b
8. If E = ac db then I − E = 1−c 1 − d . We have det (I − E) = (1 − a)(1 − d) − bc = 1 −
(a + d) +h (ad − bc) = −1
i 1 − tr E + det E. If det (I − E) 6= 0 then Example 2.3.5 gives (I − E) =
1 1−d b −1 ≥ 0 if
det (I−E) c 1 − a . The entries 1 − d, b, c, and 1 − a are all between 0 and 1 so (I − E)
det (I − E) > 0, that is if tr E < 1 + det E.
2.9. An Application to Markov Chains 37
3
9. b. If p = 2 then p > Ep so Theorem 2.8.2 applies.
1
3
d. If p = 2 then p > Ep so Theorem 2.8.2 applies.
2
1. b. Not regular. Every power of P has the (1, 2)- and (3, 2)-entries zero.
h 1 i h i h i
2. b. I − P = −21 −11 → 10 −20 so (I − P)s = 0 has solutions s = 2tt . The entries of s sum
2
2 1
h i
1 1
to 1 if t = 3 , so s = 31 is the steady state vector. Given s0 = 0 , we get s1 = Ps0 = 21 ,
3 2
3 5
s2 = Ps1 41 , s3 = Ps2 = 83 . So it is in state 2 after three transitions with probability 38 .
4 8
.6 −.1 −.5 1 −2 1 1 0 −1
d. I − P = −.2 .4 −.2 → 0 11 −11 → 0 1 −1 so (I − P)s = 0 has solution s =
−.4 −.3 .7 0 −11 11 0 0 0
" 1 #
t 3 1
t . The entries sum to 1 if t = 1
3 so the steady state vector is s = 1
3 . Given s 0 = 0 ,
t 1 0
3
.4 .38 .350
s1 = Ps0 = .2 , s2 = Ps1 = .28 , s3 = Ps2 = .312 . Hence it is in state 2 after three
.4 .34 .338
transitions with probability .312.
1 0 − 85
.9 −.3 −.3 1 −3 2
f. I − P = −.3 .9 −.6 → 0 24 −21 → 0 1 − 8 , so (I − P)s = 0 has solution s =
7
1 1
6. Let States 1 and 2 be “late” and “on time” respectively. Then the transition matrix is P = 3
2
2
1 .
3 2
Here column 1 describes what happens if he was late one day: the two entries sum to 1 and the top
entry is twice the bottom entry by the information we are given.
Column
2 is determined similarly.
3
h 3 i
Now if Monday is the initial state, we are given that s0 = 1 . Hence s1 = Ps0 = 85 and
4
4 8
7
7 9
s2 = Ps1 = 169 . Hence the probabilities that he is late and on time Wednesdays are 16 and 16
16
respectively.
8. Let the states be the five compartments. Since each tunnel entry is equally likely,
1 1 1
0 2 5 0 2
1 1
3 0 0 4 0
P= 1
3 0 2
5
1
4
1
2
1 1 2
0 2 5 4 0
1 1
3 0 5 0 0
7
Hence the probability that he is in compartment 1 after three moves is 75 .
b. The steady state vector s satisfies (I − P)s = 0. As
1 1 1
1 −2 −5 0 −2 1 0 0 0 − 32
− 13 1 0 − 41 0 0 1 0 0 −1
(I − P) = − 13 0 3
5 − 41 − 12 → 0 0 1 0 − 52
− 21 − 15 1
0 2 0 0 0 0 1 −2
− 13 0 − 15 0 1 0 0 0 0 0
3
2
1
so the steady state is s = 16
5 . Hence, in the long run, he spends most of his time in
4
2
5
compartment 3 (in fact 16 of his time).
h i h i h i h i h i
1− p q 1 q 1 (1 − p)q + qp 1 q 1 q
12. a. p 1−q · p+q p = p+q pq + (1 − q)p = p+q p . Since the entries of p+q p add
to 1, it is the steady state vector.
b. If m = 1
h i h i h i
q + p − p2 − pq q − q + pq + q2
1
p+q
q
p
q
p + 1−p−q
p+q
p
−p
−q
q = 1
p+q p − p_p2 + pq p + q − pq − q2
h i
1 (p + q)(1 − p) (p + q)q
= p+q (p + q)p (p + q)(1 − q)
=P
2.9. An Application to Markov Chains 39
h i h i
q q p −q
In general, write X = p p and Y = −p q . Then PX = X and PY = (1 − p − q)Y .
m
Hence if Pm = 1
p+q X + (1−p−q)
p+q Y for some m ≥ 1, then
1 (1−p−q)m
Pm+1 = PPm = p+q PX + p+q PY
m
= 1
p+q X + (1−p−q)
p+q (1 − p − q)Y
m+1
= 1
p+q X + (1−p−q)
p+q Y
2. b. We have 0 = p(U ) = U 3 −5U 2 +11U −4I so that U (U 2 −5U +11I) = 4I = (U 2 −5U +11I)U .
Hence U −1 = 14 (U 2 − 5U + 11I).
6. d. Using (c), I pq AIrs = ∑ni=1 ∑nj=1 ai j I pq Ii j Irs . Now (b) shows that I pq Ii j Irs = 0 unless i = q and
j = r, when it equals I ps . Hence the double sum for I pq AIrs has only one nonzero term — the
one for which i = q, j = r. Hence I pq AIrs = aqr I ps .
because AIrs = Irs A. Hence aqr = 0 if q 6= r by Exercise 6(b). If r = q then aqq I ps = I ps A is the
same for each value of q. Hence a11 = a22 = · · · = ann , so A is a scalar matrix.
3. Determinants and Diagonalization
a+1 a 1 1
d. Subtract row 2 from row 1: a a−1 = a a−1 = (a − 1) − a = −1
f. Subtract 2 times row 2 from row 1, then expand along row 2:
2 0 −3 0 −4 −13
−4 −13
1 2 5 = 1 2 5 =− 3 0 = −39
0 3 0 0 3 0
0 a 0
b d
h. Expand along row 1: b c d = −a 0 0 = −a(0) = 0
0 e 0
l. Subtract multiples of row 1 from rows 2, 3 and 4, then expand along column 1:
1 0 3 1 1 0 3 1
2 0 −2
2 2 6 0 0 2 0 −2
−1 0 −3 1 = 0 0 0 2 = 0 0 2 =0
1 0 −4
4 1 12 0 0 1 0 −4
n. Subtract multiples of row 4 from rows 1 and 2, then expand along column 1:
4 −1 3 −1 0 −9 7 −5
−9 7 −5
3 1 0 2 0 −5 3 −1
0 1 2 2 = 0 1 2 2 =− −5 3 −1
1 2 2
1 2 −1 1 1 2 −1 1
Again, subtract multiples of row 3 from rows 1 and 2, then expand along column 1:
4 −1 3 −1
0 25 13
3 1 0 2 25 13 −1 −5
0 1 2 2 =− 0 13 9 = 13 9 =− 13 9 = −(−9 + 65) = −56
1 2 2
1 2 −1 1
−1 3 1 −1 3 1 −1 3 1 −1 3 1
5. b. 2 5 3 = 0 11 5 =− 0 1 2 =− 0 1 2 = −(17) = −17
1 −2 1 0 1 2 0 11 5 0 0 −17
41
42 Determinants and Diagonalization
2 3 1 1 1 1 2 5 1 1 2 5 1 1 2 5
0 2 −1 3 0 2 −1 3 0 2 −1 3 0 1 −3 −9
d. 0 5 1 1 =− 0 5 1 1 =− 0 5 1 1 = 0 5 1 1 =
1 1 2 5 2 3 1 1 0 1 −3 −9 0 2 −1 3
1 1 2 5 1 1 2 5
0 1 −3 −9 0 1 −3 −9
0 0 16 46 = 0 0 1 −17 = 106
0 0 5 21 0 0 0 106
a b c a b c
6. b. Subtract row 1 from row 2: a+b 2b c+b = b b b = 0 by Theorem 3.1.2(4).
2 2 2 2 2 2
7. b. Take −2 and 3 out of rows 1 and 2, then subtract row 3 from row 2, then take 2 out of row 2:
−2a −2b −2c a b c a b c a b c
2p + x 2q + y 2r + z = −6 2p + x 2q + y 2r + z = −6 2p 2q 2r = −12 p q r = 12
3x 3y 3z x y z x y z x y z
Now subtract row 1 from rows 2 and 3, and then add row 2 plus twice row 3 to row 1, to get
a+ p+x b+q+y c+r +z 3x 3y 3z
=3 p−a q−b r−c =3 p−a q−b r−c
x− p y−q z−r x− p y−q z−r
Next take 3 out of row 1, and then add row 3 to row 2, to get
x y z x y z
=9 p−a q−b r−c =9 −a −b −c
−p −q −r −p −q −r
h i
1 1
9. b. False. The matrix A = has zero determinant, but no two rows are equal.
2 2
h i h i
2 0 1 0
d. False. The reduced row-echelon form of A = 0 1 is R = 0 1 , but det A = 2 while
det R = 1.
h i
1 1
f. False. A = 0 1 , det A = 1 = det AT .
h i h i
1 1 1 0
h. False. If A = 0 1 and B = 1 1 then det A = det B = 1. In fact, it is a theorem that
det A = det AT holds for every square matrix A.
14. b. Follow the Hint, take out the common factor in row 1, subtract multiples of column 1 from
columns 2 and 3, and expand along row 1:
x−1 −3 1 x−2 x−2 x−2 1 1 1
det 2 −1 x−1 = 2 −1 x−1 = (x − 2) 2 −1 x−1
−3 x+2 −2 −3 x+2 −2 −3 x+2 −2
1 0 0
−3 x−3
= (x − 2) 2 −3 x−3 = (x − 2) x+5 1
−3 x+5 1
2 −1
15. b. If we expand along column 2, the coefficient of z is − 1 3 = −(6 + 1) = −7. So c = −7.
16. b. Compute det A by adding multiples of row 1 to rows 2 and 3, and then expanding along column
1:
1 x x 1 x x
x2 − 2 x2 + x
det A = −x −2 x = 0 x2 − 2 x2 + x = x2 − x x2 − 3
−x −x −3 0 x2 − x x2 − 3
where the determinant at the second step is expanded along column 1. Similarly, T (ax) = aT (x) for
any scalar a.
44 Determinants and Diagonalization
24. Suppose A is n × n. B can be found from A by interchanging the following pairs of columns: 1 and
n, 2 and n − 1, . . . . There are two cases according as n is even or odd:
Case 1. n = 2k. Then we interchange columns 1 and n, 2 and n − 1, . . . , k and k + 1, k interchanges
in all. Thus det B = (−1)k det A in this case.
Case 2. n = 2k + 1. Now we interchange columns 1 and n, 2 and n − 1, . . . , k and k + 2, leaving
column k fixed. Again k interchanges are used so det B = (−1)k det A.
Thus in both cases: det B = (−1)k det A where A is n × n and n = 2k or n = 2k + 1.
Remark: Observe that, in each case, k and 12 n(n − 1) are both even or both odd, so (−1)k =
1 1
(−1) 2 n(n−1) . Hence, if A is n × n, we have det B = (−1) 2 n(n−1) det A.
d. In computing the cofactor matrix, we use the fact that det 31 M = 91 det M for any 2 × 2 matrix
3. b. det (B2C−1 AB−1CT ) = det B2 det C−1 det A det B−1 det CT
= ( det B)2 det1 C det A det1 B det C
= det B det A
= −2
1 1
4. b. det (A−1 B−1 AB) = det A−1 det B−1 det A det B = det A det B det A det B = 1.
Note that the following proof is wrong:
The reason is that A−1 B−1 AB may not equal A−1 AB−1 B because B−1 A need not equal AB−1 .
6. b. Since C is 3 × 3, the same is true for C−1 , so det (2C−1) = 23 · det C−1 = det8 C . Now we
compute det C by taking 2 and 3 out of columns 2 and 3, subtracting column 3 from column
2:
2p −a + u 3u p −a + u u p −a u
det C = 2q −b + v 3v =6 q −b + v v =6 q −b v
2r −c + w 3w r −c + w w r −c w
Now take −1 from column 2, interchange columns 1 and 2, and apply Theorem 3.2.3:
p a u a p u a b c
det C = −6 q b v =6 b q v =6 p q r = 6 · 3 = 18
r c w c r w u v w
8 8
Finally det 2C−1 = det C = 18 = 49 .
7. b. Begin by subtracting row 2 from row 3, and then expand along column 2:
2b 0 4d 2b 0 4d
2b 4d b 2d b d
1 2 −2 = 1 2 −2 =2 a 2c =4 a 2c =8 a c
a+1 2 2(c − 1) a 0 2c
9 4 3 9
−1 −1 −5 5 2 −1 −21 21
8. b. x = 3 4
= −11 = 11 , y= 3 4
= −11 = 11
2 −1 2 −1
46 Determinants and Diagonalization
1 1
T
b. A−1 =
9. det A adj A = det A Ci j where Ci j is the cofactor matrix. Hence the (2, 3)-entry of
1 2 −1 1 2 −1
1 1 −1
A−1 is det A C32 . Now C32 = − 3 1 = −4. Since det A = 3 1 1 = 0 −5 4 =
0 4 7 0 4 7
−5 4 −4 4
4 7 = −51, the (2, 3) entry of A−1 is −51 = 51 .
10. b. If A2 = I then det A2 = det I = 1, that is ( det A)2 = 1. Hence det A = 1 or det A = −1.
d. If PA = P, P invertible, then det PA = det P, that is det P det A = det P. Since det P 6= 0 (as
P is invertible), this gives det A = 1.
f. If A = −AT , A is n × n, then AT is also n × n so, using Theorem 3.1.3 and Theorem 3.2.3,
If n is even this is det A = det A and so gives no information about det A. But if n is odd it
reads det A = − det A, so det A = 0 in this case.
15. Write d = det A, and let C denote the cofactor matrix of A. Here
AT = A−1 = 1
d adj A = d1 CT
2 −1 −1 −1 −1 2
−
−c c c c c −c
c 0 −c
c −c 0 −c 0 c
Ci j = −
−c c c c
−
c −c
= 0 c2 c2
c c c
c −c 0 −c 0 c
−
2 −1 −1 −1 −1 2
T c 0 c 1 0 1
1 1 1 1
Hence A−1 = det A adj A = c2
Ci j = c2
0 c2 c = c
0 c 1 for any c 6= 0.
−c c2 c −1 c 1
3.2. Determinants and Matrix Inverses 47
4 c 3
d. Write A = c 2 c . Then det A = 2 (Exercise 2) and the cofactor matrix is
5 c 4
2 c c c c 2
−
c 4 5 4 5 c
8 − c2 c2 − 10
c
c 3 4 3 4 c
[Ci j ] = −
c 4 5 4
−
5 c
= −c 1 c
c2 − 6 −c 8 − c2
c 3 4 3 4 c
−
2 c c c c 2
T 1 8 − c2 −c c2 − 6
1 1
Hence A−1 = det A adj A = Ci j = 2 2 c 2
1 −c .
c − 10 c 8 − c2
1 c −1
f. Write A = c 1 1 . Then det A = −(c3 + 1) (Exercise 2) so det A = 0 means c 6= −1 (c
0 1 c
is real). The cofactor matrix is
1 1 c 1 c 1
−
1 c 0 c 0 1
−c2
c−1 c
c −1 1 −1 1 c
Ci j = −
1 c 0 c
−
0 1
= −(c2 + 1) c −1
c+1 −(1 + c) 1 − c2
c −1 1 −1 1 c
−
1 1 c 1 c 1
c − 1 −(c2 + 1) 1 − c c2 + 1 −c − 1
c+1
1 −1 −1 1
Hence A−1 = det A adj A = [C ]T
c3 +1 i j
= c3 +1
−c2 c −(c + 1) = c3 +1
c2 −c c+1 ,
c −1 1 − c2 −c 1 c2 − 1
where c 6= −1.
20. b. True. Write d = det A, so that d · A−1 = adj A. Since adj A = A−1 by hypothesis, this gives
dA−1 = A−1 , that is (d − 1)A−1 = 0. It follows that d = 1 because A−1 6= 0 (see Example
2.1.7).
d. True. Since AB = AC we get A(B − C) = 0. As A is invertible, this means B = C. More
precisely, left multiply by A−1 to get A−1 A(B − C) = A−1 0 = 0; that is I(B − C) = 0; that is
B −C = 0, so B = C.
1 1 1
f. False. If A = 1 1 1 then adj A = 0. However A 6= 0.
1 1 1
h i h i
1 1 0 −1
h. False. If A = 0 0 then adj A = 0 1 , and this has no row of zeros.
h i
−1 1
j. False. If A = 1 −1 then det (I + A) = −1 but 1 + det A = 1.
h i h i
l. False. If A = 10 11 then det A = 1, but adj A = 10 −11 6= A.
r0 = p(0) = 5
r0 + r1 + r2 = p(1) = 3
r0 + 2r1 + 4r2 = p(2) = 5
r0 = p(0) = 1
r0 + r1 + r2 + r3 = p(1) = 1
r0 − r1 + r2 − r3 = p(−1) = 2
r0 − 2r1 + 4r2 − 8r3 = p(−2) = −3
−5
The solution is r0 = 1, r1 = 3 , r2 = 12 , r3 = 76 , so p(x) = 1 − 53 x + 12 x2 + 76 x3 .
r0 = p(0) = 1
r0 + r1 + r2 + r3 = p(1) = 1.49
r0 + 2r1 + 4r2 + 8r3 = p(2) = −0.42
r0 + 3r1 + 9r2 + 27r3 = p(3) = −11.33
The solution is r0 = 1, r1 = −0.51, r2 = 2.1, r3 = −1.1, so
to two decimals.
26. b. Let A be an upper triangular, invertible, n × n matrix. We use induction
h i on n. If n h= 1 it isi
a X
clear (every 1 × 1 matrix is upper triangular). If n > 1 write A = 0 B and A−1 = Zb YC
in block form. Then h i h i
1 0 −1 ab + XZ aY + XC
0 I = AA = BZ BC
34. b. Write d = det A so det A−1 = d1 . Now the adjugate for A−1 gives
Take inverses to get ( adj A−1 )−1 A = dI. But dI = ( adj A)A by the adjugate formula for A.
Hence
( adj A−1 )−1 A = ( adj A)A
−1
Since A is invertible, we get adj A−1
= adj A, and the result follows by taking inverses
again.
3.3. Diagonalization and Eigenvalues 49
Thus AB adj (AB) = AB · adj B · adj A, and the result follows because AB is invertible.
h i
Hence b1 = so, as λ1 is dominant, xk ∼
7
3 λ k x = 7 2k 2 .
= 1 1 1 3
b 1
1 1 1
d. Here λ1 = 3, λ2 = −2 and λ3 = 1; x1 = 0 , x2 = 1 and x3 = −2 , and P =
−3
1 3
1 1 1 3 6 3 9
0 1 −2 . Now P−1 = 1 2 −2 −2 , so P−1 v0 = 1 2 and hence b1 = 23 . Hence
6 0 6
1 −3 1 −4 −1
3 1
1
vk ∼
= 32 3k 0 .
1
20. b. If µ is an eigenvalue of A−1 then A−1 x = µ x for some column x 6= 0. Note that µ 6= 0 because
A−1 is invertible and x 6= 0. Left multiplication by A gives x = µ Ax, whence Ax = µ1 x. Thus,
1 1 1
µ is an eigenvalue of A; call it λ = µ . Hence, µ = λ as required. Conversely, if λ is any
eigenvalue of A then λ 6= 0 by (a) and we claim that λ1 is an eigenvalue of A−1 . We have
Ax = λ x for some column x 6= 0. Multiply on the left by A−1 to get x = λ A−1 x; whence
A−1 x = λ1 x. Thus λ1 is indeed an eigenvalue of A−1 .
27. a. If A is diagonalizable and has only one eigenvalue λ , then the diagonalization algorithm asserts
that P−1 AP = λ I. But then A = P(λ I)P−1 = λ I, as required.
b. Here the characteristic polynomial is cA (x) = (x − 1)2 , so the only eigenvalue is λ = 1. Hence
A is not diagonalizable by (a).
h 1 1 i h 1 1 i
31. b. The matrix in Example 3.3.1 is 22 04 . In this case A = 34 04 so
h i
x − 14 − 41
cA (x) = det −3 x = x2 − 14 x − 34 = (x − 1)(x + 34 )
stabilizes if α = 15 . In fact it is easy to see that the population becomes extinct (λ1 < 1) if and only
if α < 15 , and the population diverges (λ1 > 1) if and only if α > 15 .
(−2)k (−2)k
h i
1 k
xk = 34 − 13 (−2)k = 13 [4 − (−2)k ] =
3 1−4 −2 ≈− 3 for large k.
h i h i h i
xk 1 1
Now xk+1 = 45 2k 2 + 15 (−3)k −3 , and so, looking at the top entries we get
xk = 12 (−1)k + 12 1k = 12 [(−1)k + 1]
Note that the sequence xk here is 0, 1, 0, 1, 0, 1, . . . which does not converge to any fixed
value for large k.
3. b. If a bus is parked at one end of the row, the remaining spaces can be filled in xk ways to fill it in;
if a truck is at the end, there are xk+2 ways; and if a car is at the end, there are xk+3 ways. Since
one (and only one) of these three possibilities must occur, we have xk+4 = xk + xk+2 + xk+3
must hold for all k ≥ 1. Since x1 = 1, x2 = 2 (cc or t), x3 = 3 (ccc, ct or tc) and x4 = 6 (cccc,
cct, ctc, tcc, tt, b), we get successively, x5 = 10, x6 = 18, x7 = 31, x8 = 55, x9 = 96, x10 = 169.
5. Let xk denote the number of ways to form words of k letters. A word of k + 2 letters must end in
either a or b. The number of words that end in b is xk+1 — just add a b to a (k + 1)-letter word.
But the number ending in a is xk since the second-last letter must be a b (no adjacent a’s) so we
simply add ba to any k-letter word. This gives the recurrence xk+2 = xk+1 + xk which is the same as
in Example 3.4.2, but with different initial conditions: x0 = 1 (since the “empty” word is the only
one formed with no letters) and x1 = 2. The eigenvalues, eigenvectors, and diagonalization remain
the same, and so h i h i
vk = b1 λ1k λ11 + b2 λ2k λ12
√ √
where λ1 = 12 (1 + 5) and λ2 = 12 (1 − 5). Comparing top entries gives
xk = b1 λ1k + b2 λ2k
h i
By Theorem 2.4.1, the constants b1 and b2 come from bb12 = P0−1 v0 . However, we vary the method
and use the initial conditions to determine the values of b1 and b2 directly. More precisely, x0 = 1
means√ 1 = b1 + b2 while
√ x1 = 2 means 2 = b1 λ1 + b2 λ2 . These equations have unique solution
5−3 5−3
b1 = √ and b2 = √ . It follows that
2 5 2 5
√ 1+√5 k √ 1−√5 k
1
xk = √ (3 + 5) 2 + (−3 + 5) 2 for each k ≥ 0
2 5
54 Determinants and Diagonalization
7. In a stack of k + 2 chips, if the last chip is gold then (to avoid having two gold chips together)
the second last chip must be either red or blue. This can happen in 2xk ways. But there are xk+1
ways that the last chip is red (or blue) sohthere are
i 2xk+1 ways these possibilities can occur. Hence
0 1
√ √
xk+2 = 2xk + 2xk+1 . The matrix is A = 2 2 with eigenvalues λ1 = 1 + 3 and λ2 = 1 − 3
h i h i
1
and corresponding eigenvectors x1 = λ1 and x2 = λ12 . Given the initial conditions x0 = 1 and
x1 = 3, we get
h i h ih i h √ i h √ i
b1 −1 1 λ2 −1 1 1√ −2 −√ 3 1 2 + √3
b2 = P 0 v 0 = √
3 −λ1 1 3 = −2 3 2− 3
= √
2 3 −2 + 3
1
h √ √ √ √ i
xk = b1 λ1k + b2 λ2k = √
2 3
(2 + 3)(1 + 3)k + (−2 + 3)(1 − 3)k
yk +yk+1
9. Let yk be the yield for year k. Then the yield for year k + 2 is yk+2 = =i12 yk + 12 yk+1
h . The
i 2 h
eigenvalues are λ1 = 1 and λ2 = − 12 , with corresponding eigenvectors x1 = 1 and x2 = −21 .
1
Given that k = 0 for the year 1990, we have the initial conditions y0 = 10 and y1 = 12. Thus
h i h ih i h i
b1 −1 1 1 2 10 1 34
b2 = P0 v 0 = 3 −1 1 12 = 3 2
Since h i k h i
34 1 −2
vk = 3 (1)
k
1 + 23 − 12 1
then k k
34 2
yk = k
3 (1) + 3 (−2) − 12 = 34
3 − 43 − 21
34
For large k, yk ≈ so the long term yield is 11 13 million tons of wheat.
3
0 1 0
11. b. We have A = 0 0 1 so cA (x) = x3 − (a + bx + cx2 ). If λ is any eigenvalue of A, and we
a b c
1
write x = λ2 , we have
λ
0 1 0 1 λ λ
Ax = 0 0 1 λ = λ2 = λ2 = λx
a b c λ2 a + bλ + cλ 2 λ3
qk+2 = rk+2 − pk+2 = (ark+1 + brk + c(k)) − (apk+1 + bpk + c(k)) = aqk+1 + bqk
Hence, f1 (x) = c1 e4x + 5c2 e−2x , f2 (x) = c1 e4x − c2 e−2x . The boundary condition is f1 (0) = 1,
f2 (0) = −1; that is h i h i h i
1 1 5
−1 = f(0) = c1 1 + c2 −1
= (x + 1)(x − 2)(x − 4)
56 Determinants and Diagonalization
8
−3 −1 −2 1
5 −6 1 0 7 −8
λ1 = −1: −2 −3 2 → 2
3 −2 → 0 1 − 10 7 ; x1 = 10
−3 −1 −2 0
0 0
0 0 0
7
0 −1 −2 1 0 −1 1 0 −1 1
λ2 = 2: −2 0 2 → 0 1 2 → 0 1 2 ; x2 = −2
−3 −1 1 0 −1 −2
0 0 0
1
2 −1 −2 2 −1 −2 1 0 −1 1
λ3 = 4: −2 2 2 → 0 2 0 → 0 1 0 ; x3 = 0
−3 −1 −3 −1
3 3
0 0
0 1
−1 0 0 −8 1 1
Thus P−1 AP = 0 2 0 where P = 10 −2 0 . The general solution is
0 0 4 7 1 1
−8 1 1
−x 2x 4x −x 2x
f = c1 x1 e + c2 x2 e + c3 x3 e = c1 10 e + c2 −2 e + c3 0 e4x
7 1 1
That is
−8c1 + c2 + c3 = 1
10c1 − 2c2 = 1
2c1 + c2 + c3 = 1
3. b. Have m′ (t) = km(t), so m(t) = cekt by Theorem 3.5.1. Then the requirement that m(0) = 10
gives c = 10. Also we ask that m(3) = 8, whence 10e3k = 8, e3k = 45 . Hence (ek )3 = 45 , so
(ek ) = ( 45 )1/3 . Thus m(t) = 10( 45 )t/3 . Now, we want the half-life t0 satisfying m(t0) = 12 m(0),
3 ln(1/2)
that is 10( 45 )t0 /3 = 5 so t0 = ln(4/5) = 9.32 hours.
6. b. Assume that f1′ = a1 f1 + f2 and f2′ = a2 f1 . Differentiating gives f1′′ = a1 f1′ + f2′ = a1 f1′ + a2 f1 .
This shows that f1 satisfies (∗).
3.6. Proof of the Cofactor Expansion Theorem 57
√
1 p
1. b. −1 = 12 + (−1)2 + 22 = 6
2
√
−1 p
d. 0 = (−1)2 + 02 + 22 = 5
2
√ √
1
f. −3 1 = |−3| 12 + 12 + 22 = 3 6
2
−2 −2
2. b. A vector u in the direction of must have the form u = t
−1 for a scalar t > 0. Since
−1
2 2
q
u is a unit vector, we want kuk = 1; that is 1 = |t| (−2)2 + (−1)2 + 22 = 3t, which gives
−2
1 1
t = 3 . Hence u = 3 −1 .
2
2 2
4. b. Write u = and v =
−1 0 . The distance between u and v is the length of their difference:
2 1
√
0 p
ku − vk = −1 = 02 + (−1)2 + 12 = 2.
1
4 3 1 p
d. As in (b), the distance is 0 − 2 = −2 = 12 + (−2)2 + (−2)2 = 3.
−2 0 −2
−→ −
→
6. b. In the diagram, let E and F be the midpoints of sides BC and AC respectively. Then FC = 12 AC
−→ −
→
and CE = 12 CB. Hence
−→ −→ −→ 1 − → −
→ −
→ − → −
→
FE = FC + CE = 2 AC + 12 CB = 12 (AC + CB) = 12 AB
7. Two nonzero vectors are parallel if and only if one is a scalar multiple of the other.
59
60 Vector Geometry
√ √
−→ 1 2 −1 −→ p
9. b. PQ = −1 − 1 = −1 , so PQ = (−1)2 + (−1)2 + 52 = 27 = 3 3.
6 0 5
−→ −→
d. Here P = Q are equal points, so PQ = 0. Hence PQ = 0.
√
−→ 1 3 −2 −1 −→ p
f. PQ = 1 − −1 = 2 =2 1 . Hence PQ = |2| (−1)2 + 12 + (−1)2 = 2 3.
4 6 −2 −1
x 3 −→
10. b. Given Q(x, y, z) let q = y and p = 0 be the vectors of Q and P. Then PQ = q − p.
z −1
2
Let v = −1 .
3
−→ 5
(i) If PQ = v then q − p = v, so q = p + v = . Thus Q = Q(5, −1, 2). −1
2
−→ 1
(ii) If PQ = −v then q − p = −v, so q = p − v = 1 . Thus Q = Q(1, 1, −4).
−4
a 0 c a+c
12. b. We have au + bv + cw = + + a b 0 = a+b . Hence setting
2a 2b −c 2a + 2b − c
1
au + bv + cw = x = 3 gives equations
0
a + c = 1
a + b = 3
2a + 2b − c = 0
5
− 54 , − 12 .
Hence P = P 4,
−→ −→
17. b. Let p =
OP and q = OQ
denote
the vectors of the points
and
P Then q − p =
Q respectively.
−→ −1 1 −1 1 0
PQ = 4 and p = 3 , so q = (q − p) + p = 4 + 3 = 7 .
7 −4 7 −4 3
Hence Q = Q(0, 7, 3).
18. b. We have kuk2 = 20, so the given equation is 3u + 7v = 20(2x + v). Solving for x gives
6 26 −20
40x = 3u − 13v = 0 − 13 = −13
−12 −26 14
−20
1
Hence x = 40
−13 .
14
−
→ −→
20. b. Let S denote the fourth point. We have RS = PQ, so
−
→ −→ − → −→ −→ 3 −4 −1
OS = OR + RS = OR + PQ = −1 + 4 = 3
0 2 2
is the vector of an arbitrary point on the line. Equating coefficients gives the parametric equa-
tions of the line
x = 3 + 2t
y = −1 − t
z = 4 + 5t
62 Vector Geometry
1 1
d. Now p0 = 1 because P1 (1, 1, 1) is on the line, and take d = because the line is to 1
1 1
1 1 x
be parallel to d. Hence the vector equation is p = p0 + td = 1 + t 1 . Taking p = y ,
1 1 z
x = 1+t
the scalar equations are y = 1 + t .
z = 1+t
x = 2−t
−1
f. The line with parametric equations y = 1 has direction vector d = 0 — the compo-
1
z=t
nents are the coefficients of t. Sinceour line is parallel to this one, d will do as direction vector.
2
We are given the vector p0 = −1 of a point on the line, so the vector equation is
1
2 −1
p = p0 + td = −1 +t 0
1 1
x = 2−t
y = −1
z = 1+t
x 4−t
23. b. P(2, 3, −3) lies on the line y = 3 since it corresponds to t = 2. Similarly Q(−1, 3, −9)
z 1 − 2t
corresponds to t = 5, so Q lies on the line too.
x = 1−t
y = 2 + 2t for some t because P lies on the first line.
z = −1 + 3t
x = 2s
y = 1 + s for some s because P lies on the second line.
z=3
If we eliminate x, y, and z we get three equations for s and t:
1 − t = 2s
2 + 2t = 1 + s
−1 + 3t = 3
x
d. If y is the vector of a point on both lines, then
z
x 4 1
y = −1 +t 0 for some t (first line)
z 5 1
x 2 0
y = −7 +s −2 for some s (second line).
z 12 3
x 4 1 2 0
Eliminating y gives −1 +t 0 = −7 +s −2 . Equating coefficients gives three
z 5 1 12 3
equations for s and t:
4+t = 2
−1 = −7 − 2s
5 + t = 12 + 3s
This has a (unique) solution t = −2, s = −3 so the lines do intersect. The point of intersection
has vector
4 1 4 1 2
−1 +t 0 = −1 −2 0 = −1
5 1 5 1 3
2 0 2 0 2
(equivalently −7 +s −2 = −7 −3 −2 = −1 ).
12 3 12 3 3
1 2 1
29. Let a = −1 and b = 0 be the vectors of A and B. Then d = b − a = 1 is a direction
2 1 −1
vector for the line through A and B, so the vector c of C is given by c = a + td for some t. Then
−
→ −
→
AC = kc − ak = ktdk = |t| kdk and BC = kc − bk = k(t − 1)dk = |t − 1| kdk
−
→ −→
Hence AC = 2 BC means |t| = 2 |t − 1|, so t 2 = 4(t −1)2, whence 0 = 3t 2 −8t +4 = (t −2)(3t −
" 5 #
3 3
2
2). Thus t = 2 or t = 3 . Since c = a + td, this means c = 1 or c = − 13 .
0 4
3
31. b. If there are 2n points, then Pk and Pn+k are opposite ends of a diameter of the circle for each
−
→ −→ −
→ − → −
→
k = 1, 2, . . . . Hence CPk = −CPn+k so these terms cancel in the sum CP1 + CP2 + · · · + CP2n .
Thus all terms cancel and the sum is 0.
−→ −→ −→ −→
33. We have 2EA = DA because E is the midpoint of side AD, and 2AF = FC because F is 31 the way
−→ − →
from A to C. Finally DA = CB because ABCD is a parallelogram. Thus
−→ −→ −−→ −→ −→ −→ −→ − → −→ −→
2EF = 2(EA + AF) = 2EA + 2AF = DA + FC = CB + FC = FB
−→ −→
Hence EF = 12 FB so F is in the line segment EB, 1
3 the way from E to B. Hence F is the trisection
point of both AC and EB.
64 Vector Geometry
1. b. u · v = u · u = 12 + 22 + (−1)2 = 6
d. u · v = 3 · 6 + (−1)(−7) + 5(−5) = 18 + 7 − 25 = 0
f. v = 0 so u · v = a · 0 + b · 0 + c · 0 = 0
−18−2+0
2. b. cos θ = u·v
kukkvk = √ √
10 40
= −20
20 = −1. Hence θ = π.
d. cos θ = u·v
= √ = 1 . Hence θ = π .
√6+6−3
kukkvk 6(3 6) 2 3
0−21−4 2π
f. cos θ = u·v
kukkvk = √ √ = − 12 . Hence θ = 3 .
25 100
2 1
3. b. Writing u = −1 and v = x , the requirement is
1 2
1
2 = cos π3 = u·v
kukkvk = √2−x+2
√
6 x2 +5
Hence 6(x2 + 5) = 4(4 − x)2 , whence x2 + 16x − 17 = 0. The roots are x = −17 and x = 1.
3x − y + 2z = 0
2x + z = 0
−1
The solutions are x = −t, y = t, z = 2t, so v = t 1 .
2
2x − y + 3z = 0
0=0
1 0
The solutions are x = s, y = 2s + 3t, z = t, so v = s 2 + t 3 .
0 1
2
−→ 2 3
6. b. PQ = −2 = 9 + 4 + 16 = 29
4
2
−→ 2 2
QR = 7 = 4 + 49 + 4 = 57
2
2
2
−
→ 5
PR = 5 = 25 + 25 + 36 = 86
6
−→ −→ 2 −→ 2
Hence PR = PQ + QR . Note that this implies that the triangle is right angled, that
PR is the hypotenuse, and hence that the angle at Q is a right angle. Of course, we can confirm
−→ −→
this latter fact by computing PQ · QR = 6 − 14 + 8 = 0.
4.2. Projections and Planes 65
−
→ 2 −
→ 1
8. b. We have AB = 1 and AC = 2 so the angle α at A is given by
1 −1
−
→− →
AB·AC 2+2−1 1
cos α = −
→ − → = √ √
6 6
= 2
AB AC
π −
→ −2 −
→ −1
Hence α = 3 or 60◦ . Next BA = −1 and BC = 1 so the angle β at B is given by
−1 −2
−
→− →
BA·BC 2−1+2 1
cos β = −
→ − → = √ √ = 2
BA BC 6 6
Hence β = π3 . Since
π π π
the angles in any triangle add to π , the angle γ at C is π − 3 − 3 = 3 .
−
→ −1 −
→ 1
However, CA = −2 and CB = −1 , this can also be seen directly from
1 2
−
→− →
CA·CB −1+2+2 1
cos γ = −
→ − → = √ √ = 2
CA CB 6 6
4 4
u·v 12−2+1 11
10. b. proj v u = v = 16+1+1 1 = 1
kvk2 1
18 1
−6 −6 3
d. proj v u = u·v
v = −18−8−2 4 = − 12 4 = −2
kvk2 36+16+4 2 2 −1
−2 −2 3 −2
11. b. Take u1 = proj v u = u·v
v = −6+1+0
= 1 −5 1= . Then u2 = u−u1 = 1 5
+ 21 1
kvk2 4+1+16 4
21 4 0 4
53
1 26 . As a check, verify that u2 · v = 0, that is u2 is orthogonal to v.
21 20
−6 6
u·v −18−8−1 27
d. Take u1 = proj v u = 2 v = 36+16+1 4 = 53 −4 . Then u2 is given by u2 = u − u1 =
kvk −1 1
3 6 −3
27 1
−2 − 53 −4 = 53 2 . As a check, verify that u2 · v = 0, that is u2 is orthogonal to v.
1 1 26
1 3 −→ 1 0
12. b. Write p0 = 0 , d= 1 , p=
and write u = P0 P = p − p0 = −1 . Write
−1
−1 4 3 4
−−→ −→ 0−1+16 3 3
u1 = P0 Q and compute it as u1 = proj d P0 P = 9+1+16 1 = 15 26
1 . Then the distance
4 4
√
−→ 1
−45
1
from P to the line is QP = ku − u1 k = 26 −41 = 26 5642. To compute Q let q be its
44
vector. Then
1 3 71
15 1
q = p0 + u1 = 0 + 26
1 = 26
15
−1 4 34
Hence Q = Q( 71 15 34
26 , 26 , 26 ).
i 3 −6 0
13. b. u × v = det j −1 2 = 0i − 0j + 0k = 0 =0
k 0 0 0
66 Vector Geometry
i 2 1 4
d. u × v = det j 0 4 = 4i − 15j + 8k = −15
k −1 7 8
−
→ − → −1 3 i −1 3 23
14. b. A normal is n = AB × AC = 1 × 8 = det j 1 8 = −32 .
−5 −17 k −5 −17 −11
Since the plane passes through B(0, 0, 1) the equation is
1 1
d. The given lines have direction vectors d1 = 1 and d2 = 2 , so
−2 −3
i 1 1 1
d = d1 × d2 = det j 1 2 = 1 is perpendicular to both lines.
k −2 −3 1
Hence d is a direction vector for the line we seek. As P(1, 1, −1) is on the line, the vector
equation is
x 1 1
y = 1 +t 1
z −1 1
−→ 1+t
f. Each point on the given line has the form Q(2 + t, 1 + t, t) for some t. So PQ = t . This
t −2
−→ 1
is perpendicular to the given line if PQ · d = 0 (where d = 1 is the direction vector of the
1
given line). This condition is (1 + t) + t + (t − 2) = 0, that is t = 31 . Hence the line we want
" 4 #
3 4
has direction vector 1
3 . For convenience we use d = 1 . As the line we want passes
− 35 −5
x 1 4
1 . [Note that Q 7 , 4 , 1 is
through P(1, 1, 2), the vector equation is y = 1 + t 3 3 3
z 2 −5
the point of intersection of the two lines.]
−→ 3
16. b. Choose a point P0 in the plane, say P0 (0, 6, 0), and write u = P0 P = −5 . Now write
−1
2
n= 1 for the normal to the plane. Compute
−1
2
u·n 2
u1 = proj n u = n = 1
knk2 6 −1
√
The distance from P to the plane is ku1 k = 13 6.
0
Since p0 = 6 and q are the vectors of P0 and Q, we get
0
0 3 2 7
1 1
q = p0 + (u − u1 ) = 6 + −5 − 3
1 = 3
2
0 −1 −1 −2
7 2 −2
Hence Q = Q 3, 3, 3 .
17. b. A normal to the plane is given by
−→ − → −2 −3 i −2 −3 −10
n = PQ × PR = 2 × −1 = det j 2 −1 = 6
−4 −3 k −4 −3 8
3x + y − 2z = 1
x+y+z = 5
Solve
− 23
h i h i h i
3 1 −2 1 1 1 1 5 1 0 −2
1 1 1 5 → 0 −2 −5 −14 → 0 1 5
7
2
3
21. b. The line has direction vector d = 0 which is a normal to all such planes. If P0 (x0 , y0 , z0 )
2
is any point, the plane 3(x − x0 ) = 0(y − y0 ) + 2(z − z0) = 0 is perpendicular to the line. This
can be written 3x + 2z = 3x0 + 2z0 , so 3x + 2z = d, d arbitrary.
a
d. If the normal is n = b 6= 0, the plane is a(x − 3) + b(y − 2) + c(z + 4) = 0, where a, b and
c
c are not all zero.
−→ −1 a
f. The vector u = PQ = is parallel to these planes so the normal n =
1 is orthogonal b
−1 c
a
to u. Thus 0 = u · n = −a + b − c. Hence c = b − a and n = b . The plane passes through
b−a
Q(1, 0, 0) so the equation is a(x−1)+b(y−0)+(b−a)(z−0) = 0, that is ax+by+(b−a)z =
a. Here a and b are not both zero (as n 6= 0). As a check, observe that this plane contains
P(2, −1, 1) and Q(1, 0, 0).
a
h. Such a plane contains P0 (3, 0, 2) and its normal n = b must be orthogonal to the direction
c
1 a
vector d = −2 of the line. Thus 0 = d · n = a − 2b − c, whence c = a − 2b and n = b
−1 a − 2b
(where a and b are not both zero as n 6= 0). Thus the equation is
where a and b are not both zero. As a check, observe that the plane contains every point
P(3 + t, −2t, 2 − t) on the line.
4.2. Projections and Planes 69
23. b. Choose P1 (3, 0, 2) on the first line. The distance in question is the distance
from P1 to
the
−−→ 4 3
second line. Choose P2 (−1, 2, 2) on the second line and let u = P2 P1 = −2 . If d = 1
0 0
is the direction vector for the line, compute
u·d 10
T T
u1 = proj d u = d = 3 1 0 = 3 1 0
kdk2 10
T √
then the required distance is ku − u1 k = 3 1 0 = 10.
1 3 −1
24. b. The cross product n = 1 × 1 =
of the two direction vectors serves as a normal to
3
1 0 −2
−−→ 1
the plane. Given P1 (1, −1, 0) and P2 (−2, −1, 3) on the lines, let u = P1 P2 = 0 . Compute
3
−1 1
−7 1
u1 = proj n u = 14
3 = 2
−3
−2 2
√ √
The required distance is ku1 k = 21 1 + 9 + 4 = 12 14.
Now let A = A(1 + s, −1 + s, s) and B = B(2 + 3t, −1 + t, 3) be the points on the two lines
−
→ 1 + 3t − s
that are closest together. Then AB = t −s is orthogonal to both direction vectors d1 =
3−s
1 3 −
→ −→
1 and d2 = 1 . By Theorem 4.2.3 this means d1 · AB = 0 = d2 · AB, giving equations
1 0
4t − 3s = −4, 10t − 4s = −3. The solution is t = 12 , s = 2, so the points are A = A(3, 1, 2) and
B = B 72 , − 12 , 3 .
√
d. Analogous to (b). The distance is 66 , and the points are A( 19 1 37 13
3 , 2, 3 ) and B = B 6 , 6 , 0 .
26. b. Position the cube with one vertex at the origin and sides along the positive
axes. Assume
a
each side has length a and consider the diagonal with direction d = a . The face diagonals
a
a a 0
that do not meet d are: ± −a , ± 0 and ± a , and all are orthogonal to d (the dot
0 −a −a
product is 0).
34. b. The sum of the squares of the lengths of the diagonals equals the sum of the squares of the
lengths of the four sides.
1
4. b. The area of the triangle is the area of the parallelogram ABCD. By Theorem 4.3.4,
2
1 − → − → 2 4 0
1 1
Area of triangle = 2 AB × AC = 2 1 × 2 =2 0 =0
−1 −2 0
−
→ −
→
Hence AB and AC are parallel.
√
d. Analogous to (b). Area = 5.
−4
5. b. We have u × v = 5 so w · (u × v) = −7. The volume is |w · (u × v)| = |−7| = 7 by
1
Theorem 4.3.5.
6. b. The line through P0 perpendicular to the plane has direction vector n, and so has vector
T
equation p = p0 + tn where p = x y z . If P(x, y, z) also lies in the plane, then
n · p = ax + by + cz = d. Using p = p0 + tn we find
d = n · p = n · p0 + t(n · n) = n · p0 + t knk2
d−n·p0 d−n·p0
Hence t = 2 , so p = p0 + 2 n. Finally, the distance from P0 to the plane is
knk knk
−→
d−n·p0
|d−n·p0 |
PP0 = kp − p0 k = 2 n = knk
knk
4.4. Linear Operators on R3 71
10. The points A, B and C are all on one line if and only if the parallelogram they determine has area
−
→ − → −
→ − →
zero. Since this area is kAB × ACk, this happens if and only if AB × AC = 0.
12. If u and v are perpendicular, Theorem 4.3.4 shows that ku × vk = kuk kvk. Moreover, if w is
perpendicular to both u and v, it is parallel to u × v so w · (u × v) = ± kwk ku × vk because the angle
between them is either 0 or π . Finally, the rectangular parallepiped has volume
v1 w1 u1
16. b. Let v = v2 ,w= w2 and u = u2 . Compute
v3 w3 u3
v · [(u × v) + (v × w) + (w × u)] = v · (u × v) + v · (v × w) + v · (w × u)
v1 w1 u1
= 0 + 0 + det v2 w2 u2
v3 w3 u3
These determinants are equal because each can be obtained from the other by two column
interchanges. The result follows because (v − w) · x = v · x − w · x for any vector x.
22. If v1 and v2 are vectors of points in the planes (so v1 · n = d1 and v2 · n = d2 ), the distance is the
length of the projection of v2 − v1 along n; that is
(v2 −v1 )·n |(v −v )·n| |d −d |
k proj n (v2 − v1 )k = knk2
n = 2 knk1 = 2knk 1
h i
−3 4
d. By inspection, A = 15 4 3 ; by the formulas precedinging Theorem 4.4.2, this is the matrix
of reflection in y = 2x.
h √ i
√1 − 3
f. By inspection, A = 2 1
3 1
; by Theorem 2.6.4 this is the matrix of rotation through π3 .
h i
1 1 m
2. b. For any slope m, projection on the line y = mx has matrix (see the discussion 1+m2 m m2
preceding
h h 4.4.2).iHence the projections on the lines y = x and y = −x have matrices
i Theorem
1 1 1
2 1 1 and 2 −11 −11 , respectively, so the first followed by the second has matrix (note
1
the order) h i h i h i
1 1 −1 1 1 1 1 0 0
2 −1 1 2 1 1 = 4 0 0 =0
6. Denote the rotation by RL, θ . Here the rotation takes place about the y-axis, so RL, θ (j)h= j. In the xz-i
θ − sin θ
plane the effect of RL, θ is to rotate counterclockwise through θ , and this has matrix cos sin θ
h i h i h i h i cos θ
θ sin θ
Theorem 2.6.4. So, in the xz-plane, RL, θ 10 = cos sin θ and RL, θ 01 = −cos θ . Hence
cos θ − sin θ
RL, θ (i) = 0 and RL, θ (k) = 0 . Finally, the matrix of
sin θ cos θ
cos θ 0 − sin θ
RL, θ is RL, θ (i) RL, θ (j) RL, θ (k) = 0 1 0 .
sin θ 0 cos θ
T
9. a. Write v = x y . h i h i h ih i
v·d ax+by a 1 a2 x + aby 1 a2 ab x
Then PL (v) = proj d v = d= = a2 +b abx + b2 y = a2 +b b2 .
kdk2 a2 +b2 b 2 2 ab y
4.5. An Application to Computer Graphics 73
h 2 i
1 a ab
Hence the matrix of PL is a2 +b 2 ab b2 . Note that if the line L has slope m this retrieves
h i
1 1 m
the formula 1+m 2 m m2 preceding Theorem 4.4.2. However the present matrix works for
h i
vertical lines, where d = 10 .
1. b. Translate to the origin, rotate and then translate back. As in Example 4.5.1, we compute
" √2 − √2 0 # " 1 0 −1 # " 0 6 5 1 3 #
1 0 1 √2 √2
0 1 2 2 2 0 1 −2 0 0 3 3 9
2 2 0
0 0 1 0 0 1 0 0 1 1 1 1 1 1
√ √ √ √ √
√2+2 7√2 + 2 3√2 + 2 −√ 2 + 2 −5√ 2 + 2
1
= 2
−3 2 + 4 3 2+4 5 2+4 2+4 9 2+4
2 2 2 2 2
i h
0
5. b. The line has a point w = , so we translate by −w, then reflect in y = 2x, and then trans-
1
h i
late back by w. The line y = 2x has matrix 51 −34 43 . Thus the matrix (for homogeneous
coordinates) is
1 0 0 −3 4 0 1 0 0 −3 4 −4
1 1
0 1 1
5
4 3 0 0 1 −1 =5 4 3 2
0 0 1 0 0 5 0 0 1 0 0 5
1 −3 4 −4 1 9
1 1 9 18
Hence for w = 4 we get 5
4 3 2 4 = 5
18 . Hence the point is P 5, 5 .
1 0 0 5 1 5
4. Let p and w be the velocities of the airplane and the wind. Then kpk = 100 knots and kwk = 75
knots and the resulting actual velocity of the airplane is v = w + p. Since w and p are orthogonal.
Pythagoras’ theorem gives kvk2 = kwk2 + kpk2 = 752 + 1002 = 252 (32 + 42 ) = 252 · 52 . Hence
kvk = 25 · 5 = 125 knots. The angle θ satisfies cos θ = kwk 75 ◦
kvk = 125 = 0.6 so θ = 0.93 radians or 53 .
T
6. Let v = x y denote the velocity of the boat in the water. If c is the current velocity then
c = (0, −5) because it flows south at 5 knots. We want to h choose
i v so that the resulting actual
z
velocity w of the boat has easterly direction. Thus w = 0 for some z. Now w = v + c so
h i h i h
z x 0
i h
x
i p √
= + = . Hence z = x and y = 5. Finally, 13 = kvk = x 2 + y2 = x2 + 25
0 y −5 y−5
2
T
gives x = 144, x = ±12. But x > 0 as w heads east, so x = 12. Thus he steers v = 12 5 , and
the resulting actual speed is kwk = z = 12 knots.
5. The Vector Space Rn
10. Since ai xi is in span {xi } for each i, Theorem 5.1.1 shows that span {ai xi } ⊆ span {xi }. Since
xi = a−1
i (ai xi ) is in span {ai xi }, we get span {xi } ⊆ span {ai xi }, again by Theorem 5.1.1.
75
76 The Vector Space Rn
20. If U is a subspace then S2 and S3 certainly hold. Conversely, suppose that S2 and S3 hold. It is here
that we need the condition that U is nonempty. Because we can then choose some x in U , and so
0 = 0x is in U by S3. So U is a subspace.
22. b. First, 0 is in U +W because 0 = 0 + 0 (and 0 is in both U and W ). Now suppose that P and Q
are both in U +W , say p = x1 + y1 and q = x2 + y2 where x1 and x2 are in U , and y1 and y2
are in W . Hence
1. b. Yes. The matrix with these vectors as columns has determinant −2 6= 0, so Theorem 5.2.3
applies.
d. No. (1, 1, 0, 0) − (1, 0, 1, 0) + (0, 0, 1, 1) − (0, 1, 0, 1) = (0, 0, 0, 0) is a nontrivial linear
combination that vanishes.
2. b. Yes. If a(x + y) + b(y + z) + c(z + x) = 0 then (a + c)x + (a + b)y + (b + c)z = 0. Since we are
assuming that {x, y, z} is independent, this means a + c = 0, a + b = 0, b + c = 0. The only
solution is a = b = c = 0.
d. No. (x + y) − (y + z) + (z + w) − (w + x) = 0 is a nontrivial linear combination that vanishes.
3. b. Write x1 = (2, 1, 0, −1), x2 = (−1, 1, 1, 1), x3 = (2, 7, 4, 1), and write U = span {x1 , x2 , x3 }.
Observe that x3 = 3x1 + 4x2 so U = {x1 , x2 }. This is a basis because {x1 , x2 } is independent,
so the dimension is 2.
d. Write x1 = (−2, 0, 3, 1), x2 = (1, 2, −1, 0), x3 = (−2, 8, 5, 3), x4 = (−1, 2, 2, 1) and write
U = span {x1 , x2 , x3 , x4 }. Then x3 = 3x1 +4x2 and x4 = x1 +x2 so the space is span {x1 , x2 }.
As this is independent, it is a basis so the dimension is 2.
4. b. (a +b, a −b, b, a) = a(1, 1, 0, 1) +b(1, −1, 1, 0) so U = span {(1, 1, 0, 1), (1, −1, 1, 0)}.
This is a basis so dim U = 2.
d. (a − b, b + c, a, b + c) = a(1, 0, 1, 0) + b(−1, 1, 0, 1) + c(0, 1, 0, 1). Hence U =
span {(1, 0, 1, 0), (−1, 1, 0, 1), (0, 1, 0, 1)}. This is a basis so dim U = 3.
f. If a + b = c + d then a = −b + c + d. Hence U = {(−b + c + d, b, c, d) | b, c, d in R} so
U = span {(−1, 1, 0, 0), (1, 0, 1, 0), (1, 0, 0, 1)}. This is a basis so dim U = 3.
5.2. Independence and Dimension 77
5.3 Orthogonality
n o
1. b. √1 (1,
1, 1), √1 (4, 1, −5), √1 (3, −3, 1) where in each case we divide by the norm of
3 42 14
the vector.
e1 · e2 = 1 + 0 − 1 = 0, e1 · e3 = 2 + 0 − 2 = 0, e2 · e3 = 2 − 4 + 2 = 0
so {e1 , e2 , e3 } is orthogonal and hence a basis of R3 . If x = (a, b, c), Theorem 5.3.6 gives
x·e1
x= e1 + x·e22 e2 + x·e32 e3 = a−c
2 e1 +
a+4b+c
18 e2 +
2a−b+2c
e3
ke1 k2 ke2 k ke3 k 9
e1 · e2 = 1 − 1 + 0 = 0, e1 · e3 = 1 + 1 − 2 = 0, and e2 · e3 = 1 − 1 + 0 = 0
Hence {e1 , e2 , e3 } is orthogonal and hence is a basis of R3 . If x = (a, b, c), Theorem 5.3.6
gives
x = x·e12 e1 + x·e22 e2 + x·e32 e3 = a+b+c a−b
3 e1 + 2 e2 +
a+b−2c
6 e3
ke1 k ke2 k ke3 k
4. b. If e1 = (2, −1, 0, 3) and e2 = (2, 1, −2, −1) then {e1 , e2 } is orthogonal because e1 · e2 =
4 − 1 + 0 − 3 = 0. Hence {e1 , e2 } is an orthogonal basis of the space U it spans. If x =
(14, 1, −8, 5) is in U , Theorem 5.3.6 gives
x·e1
x= e1 + x·e22 e2 = 42 40
14 e1 + 10 e2 = 3e1 + 4e2
ke1 k2 ke2 k
We check that these are indeed equal. [We shall see in Section 8.1 that in any case,
x·e1 x·e2
x− 2 e1 + 2 e2 is orthogonal to every vector in U .]
ke1 k ke2 k
5. b. The condition that (a, b, c, d) is orthogonal to each of the other three vectors gives the fol-
lowing equations for a, b, c, and d.
a − c + d = 0
2a + b + c − d = 0
a − 3b + c = 0
Solving we get:
1 0 −1 1 0 1 0 −1 1 0
2 1 1 −1 0 → 0 1 3 −3 0
1 −3 1 0 0 1 −3 2 −1 0
" 1
#
1 0 0 0
1 0 −1 1 0 11
→ 0 1 3 −3 0 → 0 1 0 3
− 11 0
0 0 11 −10 0 0 0 1 10
− 11 0
7. b. False. For example, if x = (1, 0) and y = (0, 1) in R2 , then {x, y} is orthogonal but x + y =
(1, 1) is not orthogonal to x.
d. True. Let x and y be distinct vectors in the larger set. Then either both are xi ’s, both are yi ’s, or
one is an xi and one is a yi . In the first two cases x · y = 0 because {xi } and {y j } are orthogonal
sets; in the last case x · y = 0 by the given condition.
f. True. Every pair of distinct vectors in {x} are orthogonal (there are no such pairs). As x 6= 0,
this shows that {x} is an orthogonal set.
11. b. Take x = (1, 1, 1)qand y = (r1 , r2 , r3 ). Then |x · y| ≤ kxk kyk by Theorem 5.3.2; that is
√
|r1 + r2 + r3 | ≤ 3 r12 + r22 + r32 . Squaring both sides gives
− 21 1
" 2 −1 1
# " 2 −1 1
# 1 2
−2 1 1 0 0 2 0 0 1
1. b. 4 −2 3 → 0 0 1 → 0 0 0
.
−6 3 0 0 0 3 0 0 0
n T T o
Hence, rank A = 2 and 1 − 12 1
2 , 0 0 1 is a basis of row A. Thus
n T T o
2 −1 1 , 0 0 1 is also a basis of row A. Since the leading 1’s are in columns
1 and 3, columns 1 and 3 of A are a basis of col A.
h i h i h i
d. −31 −62 −13 −23 → 01 02 −10 73 → 10 20 −10 31 .
n T T o
Hence, rank A = 2 and 1 2 −1 3 , 0 0 0 1 is a basis of row A. Since the
leading 1’s are in columns 1 and 4, columns 1 and 4 of A are a basis of col A.
2. b. Apply the gaussian algorithm to the matrix with these vectors as rows:
" 1 −1
# " 2 5
# 1 1 1 0 0 0 1 1 0 0 0
3 1 4 2 7 0 −2 2 5 1 0 1 −1 − 25 − 21
1 1 0 0 0 → 0 −2 4 2 7 → 0 0 1 − 32 3
5 1 6 7 8 0 −4 6 7 8 0 0 0 0 0
n T T T o
Hence, 1 1 0 0 0 , 0 2 −2 −5 −1 , 0 0 2 −3 6 is a basis
of U (where we have cleared fractions using scalar multiples).
d. Write these columns as the rows of the following matrix:
" 1 5 −6 # " 1 5 −6 # " 1 5 −6
# " 1 5 −6
#
2 6 −8 0 −4 4 0 1 −1 0 1 −1
3 7 −10 → 0 −8 8 → 0 0 24 → 0 0 1
4 8 12 0 −12 36 0 0 0 0 0 0
1 0 0
Hence, 5 , 1 , 0 is a basis of U .
−6 −1 1
7. b. The null space of A is the set of columns x such that Ax = 0. Applying gaussian elimination to
the augmented matrix gives:
" 3 5 5 2 0 0 # " 1 0 2 2 1 0 #
1 0 2 2 1 0 0 5 −1 −4 −3 0
1 1 1 −2 −2 0 → 0 1 −1 −4 −3 0
−2 0 −4 −4 −2 0 0 0 0 0 0 0
1 0 2 2 1 0 1 0 0 −6 −5 0
" # " #
0 1 −1 −4 −3 0 0 1 0 0 0 0
→ 0 0 4 16 12 0 → 0 0 1 4 3 0
0 0 0 0 0 0 0 0 0 0 0 0
6s + 5t
0
Hence, the set of solutions is null A = −4s − 3t | s, t in R = span B where
s
t
6 5
0 0
B = −4 , −3 . Since B is independent, it is the required basis of null A. We have
1 0
0 1
r = rank A = 3 by the above reduction, so n − r = 5 − 3 = 2. This is the dimension of null A,
as Theorem 5.4.3 asserts.
8. A. To compute rank A, let R = r1 r2 · · · rn .
b. Since A is m × n, dim ( null A) = n − rank
Then A = CR = r1C r2C · · · rnC by block multiplication, so
col A = span {r1C, r2C, . . . , rnC} = span {C}
because some ri 6= 0. Hence rank A = 1, so dim ( null A) = n − rank A = n − 1.
b. Let A = c1 · · · cn where c j is the jth column of A; we must show that {c1 , . . . , cn } is
9.
T
independent. Suppose that x1 c1 + · · · + xn cn = 0, xi in R. If we write x = x1 · · · xn ,
this reads Ax = 0 by Definition 2.5. But then x is in null A, and null A = 0 by hypothesis. So
x = 0, that is each xi = 0. This shows that {c1 , . . . , cn } is independent.
10. b. If A2 = 0 then A(Ax) = 0 for all x in Rn , that is {Ax | x in Rn } ⊆ null A. But col A = {Ax | x
in Rn }, so this shows that col A ⊆ null A. If we write r = rank A, taking dimensions gives
r = dim ( col A) ≤ dim ( null A) = n − r by Theorem 5.4.3. It follows that 2r ≤ n; that is r ≤ n2 .
12. We have rank (A) = dim [ col (A)] and rank (AT ) = dim [ row (AT )]. Let {c1 , c2 , . . . , ck } be a basis
of col(A); it suffices to show that {cT1 , cT2 , . . . , cTk } is a basis of row (AT ). But if t1 cT1 + t2 cT2 +
· · · + tk cTk = 0, t j in R, then (taking transposes) t1 c1 + t2 c2 + · · · + tk ck = 0 so each t j = 0. Hence
{cT1 , cT2 , . . . , cTk } is independent. Given v in row (AT ) then vT is in col (A), say vT = s1 c1 + s2 c2 +
· · · + sk ck , s j in R. Hence v = s1 cT1 + s2 cT2 + · · · + sk cTk so {cT1 , cT2 , . . . , cTk } spans row (AT ), as
required.
15. b. Let {c1 , . . . , cr } be a basis of col A where r = rank A. Since Ax = b has no solution, b is not
in col A = span {c1 , · · · , cr } by Exercise 12. It follows that {c1 , . . . , cr , b} is independent
[If a1 c1 + · · · + ar cr + ab = 0 then a = 0 (since b is not in col A), whence each ai = 0 by the
independence of the ci ]. Hence, it suffices to show that col [A, B] = span {c1 , · · · , cr , b}. It
is clear that b is in col [A, b], and each c j is in col [A, b] because it is a linear combination of
columns of A (and so those of [A, b]). Hence
span {c1 , . . . , cr , b} ⊆ col [A, b]
82 The Vector Space Rn
On the other hand, each column x in col [A, b] is a linear combination of b and the columns
of A. Since these columns are themselves linear combinations of the c j , so x is a linear combi-
nation of b and the c j . That is, x is in span {c1 , . . . , cr , b}.
1. b. det A = −5, det B = −1 (so A and B are not similar). However, tr A = 2 = tr B, and rank A =
2 = rank B (both are invertible).
d. tr A = 5, tr B = 4 (so A and B are not similar). However, det A = 7 = det B, so rank A = 2 =
rank B (both are invertible).
f. tr A = −5 = tr B; det A = 0 = det B; however rank A = 2, rank B = 1 (so A and B are not
similar).
3. b. We have A ∼ B, say B = P−1 AP. Hence B−1 = (P−1 AP)−1 = P−1 A−1 (P−1 )−1 , so A−1 ∼ B−1
because P−1 is invertible.
x−3 0 −6
4. b. cA (x) = 0 x+3 0 = (x + 3)(x2 − 5x − 24) = (x + 3)2 (x − 8). So the eigenvalues are
−5 0 x−2
λ1 = −3, λ2 = 8. To find the associated eigenvectors:
−6 0 −6 1 0 1 −1 0
λ1 = −3: 0 0 0 → 0 0 0 ; basic eigenvectors 0 , 1 .
−5 0 −5 0 0 06 1 0
5 0 −6 1 0 −5 6
λ2 = 8: 0 11 0 → 0 1 0 ; basic eigenvector 0 .
−5 0 6 0 0 0 5
−1 0 6
Since 0 , 1 , 0 is a basis of eigenvectors, A is diagonalizable and
1
0 5
−1 0 6 −3 0 0
P= 0 1 0 −1
will satisfy P AP = 0 −3 0 .
1 0 5 0 0 8
x−4 0 0 0 0 0 1 0 0
2
d. cA (x) = 0 x − 2 −2 = (x − 4) (x + 1). For λ = 4, 0 2 −2 → 0 1 −1 ;
−3 x−1 2 −3
2 3 0 0 0
0
E1 = 1 . Hence A is not diagonalizable by Theorem 5.5.6 because the dimension of E4 (A) =
1
1 while the eigenvalue 4 has multiplicity 2.
9. b. Let the diagonal entries of A all equal λ . If A is diagonalizable then P−1 AP = λ I by Theorem
5.5.3 for some invertible matrix P. Hence A = P(λ I)P−1 = λ (PIP−1) = λ I.
10. b. Let P−1 AP = D = diag {λ1 , λ2 , . . . , λn }. Since A and D are similar matrices, they have the
same trace by Theorem 5.5.1. That is
tr A = tr (P−1 AP) = tr D = λ1 + λ2 + · · · + λn
12. b. TP (A)TP (B) = (P−1 AP)(P−1 BP) = P−1 AIBP = P−1 ABP = TP (AB)
5.6. Best Approximation and Least Squares 83
13. b. Assume that A is diagonalizable, say A ∼ D where D is diagonal, say D = diag (λ1 , . . . , λn )
where λ1 , · · · , λn are the eigenvalues of A. But A and AT have the same eigenvalues (Example
3.3.5) so AT ∼ D also. Hence A ∼ D ∼ AT , so A ∼ AT as required.
17. b. We use Theorem 5.5.7. The characteristic polynomial of B is computed by first adding rows 2
and 3 to row 1. For convenience, write s = a + b + c, k = a2 + b2 + c2 − (ab + ac + bc).
x−c −a −b x−s x−s x−s x−s 0 0
cB (x) = −a x−b −c = −a x−b −c = −a x + (a − b) a−c
−b −c x−a −b −c x−a −b b−c x − (a − b)
= (x − s) x2 − (a − b)2 − (a − c)(b − c)
= (x − s)(x2 − k)
√ √
Hence, the eigenvalues of B are s, k and − k. These must be real by Theorem 5.5.7, so
k ≥ 0. Thus a2 + b2 + c2 ≥ ab + ac + bc.
20. b. To compute cA (x) = det (xI − A), add x times column 2 to column 1, and expand along row 1:
x −1 0 0 ··· 0 0
0 x −1 0 ··· 0 0
0 0 x −1 ··· 0 0
.. .. .. .. .. ..
cA (x) = . . . . . .
0 0 0 0 ··· −1 0
0 0 0 0 ··· x −1
−r0 −r1 −r2 −r3 ··· −rk−2 x − rk−1
0 −1 0 0 ··· 0 0
x2 x −1 0 ··· 0 0
0 0 x −1 ··· 0 0
. . . . . .
= .. .. .. .. .. ..
0 0 0 0 ··· −1 0
0 0 0 0 x −1
−r0 − r1 x −r1 −r2 −r3 −rk−2 x − rk−1
This matrix has the same form as xI − A, so repeat this procedure. It leads to the given expres-
sion for det (xI − A).
3 1 1 6
" # " #
x 26 −2 12
2 3 −1 1
1. b. Here A = 2 −1 1 ,b= 0 ,x= y . Hence, AT A = −2 20 −12 .
z 12 −12 12
3 −3 3 8
84 The Vector Space Rn
Of course this can be found more efficiently using gaussian elimination on the normal equa-
tions for z.
i 1 2 i 4
" # " #
h h i h h i
2. b. Here M T M = 12 14 17 18 1 4
1 7 = 4 21
21 133 , M Ty = 1 1 1 1
2 4 7 8
3
2 = 10
42 . We
1 8 1
solve the normal equation (M T M)A = M T y by inverting M T M:
h ih i h i h i
133 −21
A = (M T M)−1 M T y = 91
1
−21 4
10
42 = 1
91 · 448
−42 = 1
13
64
−6
64 6
Hence the best fitting line has equation y = 13 − 13 x.
4
d. Analogous to (b). The best fitting line is y = − 10 − 17
10 x.
3. b. Now
" 1 −2 4
#
1 1 1 1 4 5 29
T 1 0 0
M M= −2 0 3 4
1 3 9 = 5 29 83
4 0 9 16 29 83 353
1 4 16
" 1
#
1 1 1 1 6
T 0
M y= −2 0 3 4
2 = 16
4 0 9 16 70
3
We use (MM T )−1 to solve the normal equations even though it is more efficient to solve them
by gaussian elimination.
3348 642 −426 6 540 .127
T −1 T 1 1
A = (M M) (M y) = 4248 642 571 −187 16 = 4248 −102 = −.024
−426 −187 91 70 822 .194
Hence the best fitting quadratic has equation y = .127 − .024x + .194x2 .
" 1 # " 0 02 20 # " 0 0 1
#
1 1 12 21 1 1 2
4. b. In the notation of Theorem 5.6.3: y = ,M= 5 2 22 22 = 2 4 4 .
10 3 32 23 3 9 8
14 36 34 230 0 −92 115 0 −46
T T −1 1 1
Hence, M M = 36 98 90 , and (M M) = 92 0 34 −36 = 46
0 17 −18 .
34 90 85 −92 −36 76 −46 −18 38
Thus, the (unique) solution to the normal equation is
115 0 −46 41 −23
T −1 T 1 1
z = (M M) M y = 46 0 17 −18 111 = 46
33
−46 −18 38 103 30
1 2 x
The best fitting function is thus 46 [−23x + 33x + 30(2) ].
5.6. Best Approximation and Least Squares 85
1
(−1)2 sin − π2
2 1 1 1 −1
1 1 02 sin(0) 1 0 0
5. b. Here y = 5
, M =
1 22 sin(π )
=
1 4 0
. Hence
32 sin 32π
9 1 1 9 −1
4 14 0 −24 7 35
T T −1 1
M M= 14 98 −10 and (M M) = 2
7 −2 −10
0 −10 2 35 −10 −49
Hence the estimate for g comes from − 12 g = −4.87, g = 9.74 (the true value of g is 9.81).
Now fit s = a + bt + ct 2 . In this case
1 1 1 1 1 1 3 6 14
T
M M= 1 2 3 1 2 4 = 6 14 36
1 4 9 1 3 9 14 36 98
1 1 1 95 231
MT y = 1 2 3 80 = 423
1 4 9 56 919
Hence " #
101
76 −84 20 231 404
A = (M T M)−1 (M T y) = 1
4
−84 98 −24 423 = 1
4
−6 = − 32
20 −24 6 919 −18 − 92
5 195 80 62
" #
195 7825 3150 2390
AT A = 80 3150 1320 1008
62 2390 1008 796
1035720 −16032 10080 −45300
" #
−16032 416 −632 800
(AT A)−1 = 1
50160 10080 −632 2600 −2180
−45300 800 −2180 3950
Hence
s2z = 1
n−1 ∑(zi − z)2 = n−1
1
∑[(a + bxi) − (a + bx)]2 = n−1
1
∑ b2 (xi − x)2 = b2s2x
√
The result follows because b2 = |b|.
1. b. No: S5 fails 1(x, y, z) = (1x, 0, 1z) = (x, 0, z) 6= (x, y, z) for all (x, y, z) in V . Note that the
other nine axioms do hold.
d. No: S4 and S5 fail: S5 fails because 1(x, y, z) = (2x, 2y, 2z) 6= (x, y, z); and S4 fails because
a[b(x, y, z)] = a(2bx, 2by, 2bz) = (4abx, 4aby, 4abz) 6= (2abx, 2aby, 2abz) = ab(x, y, z).
Note that the eight other axioms hold.
2. b. No: A1 fails — for example (x3 + x + 1) + (−x3 + x + 1) = 2x + 2 is not in the set.
d. No: A1 and S1 both fail. For example x + x2 and 2x are not in the set. Hence none of the other
axioms make sense. h i h i
f. Yes. First verify A1 and S1. Suppose A = ac db and B = xz y
w are in V , so a + c = b + d
h i
+x b+y
and x + z = y + w. Then A + B = ac + z d+w is in V because
(a + x) + (c + z) = (a + c) + (x + z) = (b + d) + (y + w) = (b + y) + (d + w)
h i
Also rA = ra rb
rc rd is in V for all r in R because ra + rc = r(a + c) = r(b + d) = rb + rd.
A2, A3,
h S2,iS3, S4, S5. These hold for matrices in general.
A4. 00 00 is in V and so serves as the zero of V .
h i h i
A5. Given A = ac db with a+c = b+d, then −A = −a −b
−c −d is also in V because −a−c =
−(a + c) = −(b + d) = −b − d. So −A is the negative of A in V .
h. Yes. The vector space axioms are the basic laws of arithmetic.
j. No. S4 and S5 fail. For S4, a(b(x, y)) = a(bx, −by) = (abx, aby), and this need not equal
ab(x, y) = (abx, −aby); as to S5, 1(x, y) = (x, −y) 6= (x, y) if y 6= 0.
Note that the other axioms do hold here:
A1, A2, A3, A4 and A5 hold because they hold in R2 .
S1 is clear; S2 and S3 hold because they hold in R2 .
l. No. S3 fails: Given f : R → R and a, b in R, we have
These need not be equal: for example, if f is the function defined by f (x) = x2 ;
Then f (ax + bx) = (ax + bx)2 need not equal (ax)2 + (bx)2 = f (ax) + f (bx).
89
90 Vector Spaces
Note that the other axioms hold. A1-A4 hold by Example 6.1.7 as we are using pointwise
addition.
S2. a( f + g)(x) = ( f + g)(ax) definition of scalar multiplication in V
= f (ax) + g(ax) definition of pointwise addition
= (a f )(x) + (ag)(x) definition of scalar multiplication in V
= (a f + ag)(x) definition of pointwise addition
As this is true for all x, a( f + g) = a f + ag.
S4. [a(b f )](x) = (b f )(ax) = f [b(ax)] = f [(ba)x] = [(ba) f ](x) = [ab f ](x) for all x,
so a(b f ) = (ab) f .
S5. (1 f )(x) = f (1x) = f (x) for all x, so 1 f = f .
n. No. S4, S5 fail: a ∗ (b ∗ X ) = a ∗ (bX T ) = a(bX T )T = abX T T = abX , while (ab) ∗ X = abX T .
These need not be equal. Similarly: 1 ∗ X = 1X T = X T need not equal X .
Note that the other axioms do hold:
A1-A5. These hold for matrix addition generally.
S1. a ∗ X = aX T is in V .
S2. a ∗ (X +Y ) = a(X +Y )T = a(X T +Y T ) = aX T + aY T = a ∗ X + a ∗Y .
S3 (a + b) ∗ X = (a + b)X T = aX T + bX T = a ∗ X + b ∗ X .
3(3y + v − u) − 2y = u
7y = 4u − 3v
6.1. Examples and Basic Properties 91
y = 47 u − 73 v
a + c = 0, b + c = 0, b + c = 0, a − c = 0
13. b. We proceed by induction on n (see Appendix A). The case n = 1 is clear. If the equation holds
for some n ≥ 1, we have
15. c. Since a 6= 0, a−1 exists in R. Hence av = aw gives a−1 av = a−1 aw; that is 1v = 1w, that is
v = w.
Alternatively: av = aw gives av − aw = 0, so a(v − w) = 0. As a 6= 0, it follows that v − w = 0
by Theorem 6.1.3, that is v = w.
92 Vector Spaces
1. b. Yes. U is a subset of P3 because xg(x) has degree one more than the degree of g(x). Clearly
0 = x · 0 is in U . Given u = xg(x) and v = xh(x) in U (where g(x) and b(x) are in P2 ) we have
(a + a1 ) + (b + b1) = (a + b) + (a1 + b1 )
= (c + d) + (c1 + d1)
= (c + c1 ) + (d + d1 )
h i
ka kb
ku = kc kd is in U because ka + kb = k(a + b) = k(c + d) = kc + kd.
d. Yes. Here 0 is in U as 0B = 0. If A and A1 are in U then AB = 0 and A1 B = 0, so (A + A1 )B =
AB + A1 B = 0 + 0 = 0 and (kA)B = k(AB) = k0 = 0 for all k in R. This shows that A + A1 and
kA are also in U .
h i h i
1 0 0 0
f. No. U is not closed under addition. In fact, A = 0 0 and A1 = 0 1 are both in U , but
h i
A + A1 = 10 01 is not in U .
3. b. No. U is not closed under addition. For example if f and g are defined by f (x) = x + 1 and
g(x) = x2 + 1, then f and g are in U but f + g is not in U because ( f + g)(0) = f (0) + g(0) =
1 + 1 = 2.
d. No. U is not closed under scalar multiplication. For example, if f is defined by f (x) = x, then
f is in U but (−1) f is not in U (for example [(−1) f ]( 12 ) = − 21 so is not in U ).
f. Yes. 0 is in U because 0(x + y) = 0 = 0 + 0 = 0(x) + 0(y) for all x and y in [0, 1]. If f and g
are in U then, for all k in R:
( f + g)(x + y) = f (x + y) + g(x + y)
= ( f (x) + f (y)) + (g(x) + g(y))
6.2. Subspaces and Spanning Sets 93
x = 13 (3x) is in U
1 = (1 + x) − x is in U
x2 = 12 [(1 + 2x2 ) − 1] is in U
11. b. The vectors u−v = 1u+(−1)v, u+v, and w are all in span {u, v, w} so span {u − v, u + w, w} ⊆
span {u, v, w} by Theorem 6.2.2. The other inclusion also follows from Theorem 6.2.2 be-
cause
u = (u + w) − w
v = −(u − v) + (u + w) − w
w=w
14. No. For example (1, 1, 0) is not even in span {(1, 2, 0), (1, 1, 1)}. Indeed (1, 1, 0) = s(1, 2, 0) +
t(1, 1, 1) requires that s + t = 1, 2s + t = 1, t = 0, and this has no solution.
18. Write W = span {u, v2 , . . . , vn }. Since u is in V we have W ⊆ V . But the fact that a1 6= 0 means
1 a2 an
v1 = a1 u − a1 v2 − · · · − a1 vn
22. If U is a subspace then, u1 + au2 is in U for any ui in U and a in R by the subspace test. Conversely,
assume that this condition holds for U . Then, in the subspace test, conditions (2) and (3) hold for
U (because 1v = v for all v in V ), so it remains to show that 0 is in U . This is where we use
the assumption that U is nonempty because, if u is any vector in U then u + (−1)u is in U by
assumption, that is 0 ∈ U .
2r + s = 0
xr + t = 0
r + s + 3t = 0
nh i h io
1 1 1 0
Hence U = span B where B = −1 0 , 0 1 . But B is independent here because
h i h i h i
s −11 10 + t 10 01 = 00 00 means s + t = 0, s = 0, −s = 0, t = 0, so s = t = 0. Thus B
is a basis of U , so dim U = 2.
n h i h i o h i
1 1 0 1 x y
d. Write U = A | A −1 0 = −1 1 A . If A = z w then A is in U if and only if
h ih i h ih i h i h i
x y 1 1 0 1 x y x−y x z w
z w −1 0 = −1 1 z w ; that is z−w z = z−x w−y .
This holds if and only if z = x − y and x = w; that is
h i h i h i
x y 1 0 0 1
A = x − y x = x 1 1 + y −1 0
96 Vector Spaces
nh i h io h i
1 0 0 1 1 0
Thus U = span B where B = 1 1 , −1 0 . But B is independent because s 1 1 +
h i h i
0 1 0 0
t −1 0 = 0 0 implies s = t = 0. Hence B is a basis of U , so dim U = 2.
h i h i h i
x y x+z y+w x y
8. b. If X = z the condition AX = X is
w = 0 0 and this holds if and only
z w
h i h i h i
if z = w = 0. Hence X = 0x 0y = x 10 00 + y 00 10 . So U = span B where B =
nh i h io
1 0 0 1
0 0 , 0 0 . As B is independent, it is a basis of U , so dim U = 2.
10. b. If the
common column sum is m, V has the form
a q r
V= b p s | a, b, p, q, r, s, m in R = span B where
m−a−b m− p−q m−r−s
0 0 0 1 0 0 0 0 0 0 1 0 0 0 0 0 0 1 0 0 0
B= 0 0 0 , 0 0 0 , 1 0 0 , 0 0 0 , 0 1 0 , 0 0 0 , 0 0 1 .
1 1 1 −1 0 0 −1 0 0 0 −1 0 0 −1 0 0 0 −1 0 0 −1
The set B is independent (a linear combination using coefficients a, b, p, q, r, s, and m yields
the matrix in V , and this is 0 if and only if a = b = p = q = r = s = m = 0). Hence B is a basis
of B, so dim V = 7.
= span B
where B = (x2 − x), x(x2 − x), x2 (x2 − x), x3 (x2 − x) . We claim that B is independent. For
if a(x2 − x) + bx(x2 − x) + cx2 (x2 − x) + dx3 (x2 − x) = 0 then (a + bx + cx2 + dx3 )(x2 − x) = 0,
whence a + bx + cx2 + dx3 = 0 by the hint in (a). Thus a = b = c = d = 0. [This also follows
by comparing coefficients.] Thus B is a basis of V , so dim V = 4.
12. b. No. If P3 = span { f1 (x), f2 (x), f3 (x), f4 (x)} where fi (0) = 0 for each i, then each polynomial
p(x) in P3 is a linear combination
for every p(x) in P3 . This is not the case, so no such basis of P3 can exist. [Indeed, no such
spanning set of P3 can exist.]
nh i h i h i h io
1 0 1 1 1 0 0 1
d. No. B = 0 1 , 0 1 , , is a basis of invertible matrices.
h i h 1 i1 h1 1 i h i h i
Independent: r 10 01 + s 10 11 + t 11 01 + u 01 11 = 00 00 gives
r + s + t = 0, s + u = 0, t + u = 0, r + s + t + u = 0. The only solution is r = s = t = u = 0.
6.3. Linear Independence and Dimension 97
h i h i h i
0 1 1 1 1 0
Spanning: 0 0 = 0 1 − 0 1 is in span B
h i h i h i
0 0 1 0 1 0
1 0 = 1 1 − 0 1 is in span B
h i h i h i h i
0 0 0 1 0 1 0 0
0 1 = 1 1 − 0 0 − 1 0 is in span B
h i h i
h i
1 0 1 0 0 0
0 0 = 0 1 − 0 1 is in span B
nh i h i h i h io
0 1
Hence M22 = span 0 0 , 01 0
0 , 0
0
0
1 , 1 0
0 0 ⊆ span B. Clearly span B ⊆
M22 .
f. Yes. Indeed, 0u + 0v + 0w = 0 for any u, v, w, independent or not!
h. Yes. If su + t(u + v) = 0 then (s + t)u + tv = 0, so s + t = 0 and t = 0 (because {u, v} is
independent). Thus s = t = 0.
j. Yes. If su + tv = 0 then su + tv + 0w = 0, so s = t = 0 (because {u, v, w} is independent).
This shows that {u, v} is independent.
l. Yes. Since {u, v, w} is independent, the vector u + v + w is not zero. Hence {u + v + w} is
independent (see Example 5.2.5).
n. Yes. If I is a set of independent vectors, then |I| ≤ n by the fundamental theorem because V
contains a spanning set of n vectors (any basis).
15. If a linear combination of the vectors in the subset vanishes, it is a linear combination of the vectors
in the larger set (take the coefficients outside the subset to be zero). Since it still vanishes, all the
coefficients are zero because the larger set is independent.
19. We have su′ + tv = s(au + bv) + t(cu + dv) = (sa + tc)u + (sb + td)v. Since {u, v} is independent,
we have
su′ + tv′ = 0 if and only if sa
h + tc i=h 0 and h + td
i sb i =0
a c s 0
if and only if b d t = 0
h ih i h i h i h i
′ ′ a c
Hence {u , v } is independent if and only if b d s
t = 0 implies st = 00 .
0
1. b. B = {(1, 0, 0), (0, 1, 0), (0, 1, 1)} is independent as r(1, 0, 0) + s(0, 1, 0) + t(0, 1, 1) =
(0, 0, 0) implies r = 0, s + t = 0, t = 0, whence r = s = t = 0. Hence B is a basis by Theorem
6.4.3 because dim R3 = 3.
d. B = 1, x, x2 − x + 1 is independent because r1 + sx + t(x2 − x − 1) = 0 implies r − t = 0,
6. b. {4, 4x, 4x2 , 4x3 } is such a basis. There is no basis of P3 consisting of polynomials have the
property that their coefficients sum to zero. For if it did then every polynomial in P3 would have
this property (since sums and scalar multiples of such polynomials have the same property).
8. b. Yes, four vectors can span R3 — say any basis together with any other vector.
No, four vectors in R3 cannot be independent by the fundamental theorem (Theorem 6.3.2)
because R3 is spanned by 3 vectors ( dim R3 = 3).
10. We have det A = 0 if and only if A is not invertible. This holds if and only if the rows of A are
dependent by Theorem 5.2.3. This in turn holds if and only if some row is a linear combination of
the rest by the dependent lemma (Lemma 6.4.3).
11. b. No. Take X = {(0, 1), (1, 0)} and D = {(0, 1), (1, 0), (1, 1)}. Then D is dependent, but its
subset X is independent.
d. Yes. This is follows from Exercise 15 Section 6.3 (solution above).
18. b. The two-dimensional subspaces of R3 are the planes through the origin, and the one-dimensional
subspaces are the lines through the origin. Hence part (a) asserts that if U and W are distinct
planes through the origin, then U ∩W is a line through the origin.
23. b. Let vn denote the sequence with 1 in the nth coordinate and zeros elsewhere. Thus v0 =
(1, 0, 0, . . . ), v1 = (0, 1, 0, . . . ) etc. Then a0 v0 +a1 v1 +· · ·+an vn = (a0 , a1 , . . . , an , 0, 0, . . . )
so a0 v0 + a1 v1 + · · · + an vn = 0 implies a0 = a1 = · · · = an = 0. Thus {v0 , v1 , . . . , vn } is an
independent set of n+1 vectors. Since n is arbitrary, dim V cannot be finite by the fundamental
theorem.
25. b. Observe that Ru = {su | s in R}. Hence Ru + Rv = {su + tv | s in R, t in R} is the set of all
linear combinations of u and v. But this is the definition of span {u, v}.
100 Vector Spaces
2. b. f (0) (x) = f (x) = x3 + x + 1, so f (1) (x) = 3x2 + 1, f (2) (x) = 6x, f (3) (x) = 6. Hence, Taylor’s
theorem gives
f (2) (1) f (3) (1)
f (x) = f (0) (1) + f (1) (1)(x − 1) + 2
2! (x − 1) + 3! (x − 1)
3
d. f (0) (x) = f (x) = x3 − 3x2 + 3x, f (1) (x) = 3x2 − 6x + 3, f (2) (x) = 6x − 6, f (3) (x) = 6. Hence,
Taylor’s theorem gives
f (2) (1) f (3) (1)
f (x) = f (0) (1) + f (1) (1)(x − 1) + 2
2! (x − 1) + 3! (x − 1)
3
0
= 1 + 0(x − 1) + 2! (x − 1)2 + 1(x − 1) 3
= 1 + (x − 1)3
Given f (x) = x2 + x + 1:
10. b. If r(x − a)2 + s(x − a)(x − b) + t(x − b)2 = 0, then taking x = a gives t(a − b)2 = 0, so t =
0 because a 6= b; and taking x = b gives r(b − a)2 = 0, so r = 0. Thus, we are left with
s(x −a)(x − b) = 0. If x is any number except a, b, this implies s = 0. Thus
B = (x − a)2 , (x − a)(x − b), (x − b)2 is independent in P2 ; since dim P2 = 3, B is a basis.
11. b. Have Un = { f (x) in Pn | f (a) = 0 = f (b)}. Let {p1 (x), . . . , pn−1 (x)} be a basis of Pn−2 ; it
suffices to show that
is a basis of Un . Clearly B ⊆ Un .
Independent: Let s1 (x − a)(x − b)p1 (x) + · · · + sn−1 (x − a)(x − b)pn−1 (x) = 0. Then (x −
a)(x − b)[s1 p1 (x) + · · · + sn−1 pn−1 (x)] = 0, so (by the hint) s1 p1 (x) + · · · + sn−1 pn−1 (x) = 0.
Thus s1 = s2 = · · · = sn−1 = 0.
6.6. An Application to Differential Equations 101
Spanning: Given f (x) in Pn with f (a) = 0, we have f (x) = (x − a)g(x) for some polynomial
g(x) in Pn−1 by the factor theorem. But 0 = f (b) = (b − a)g(b) so (as b 6= a) g(b) = 0. Then
g(x) = (x − b)h(x) with h(x) = r1 p1 (x) + · · · + rn−1 pn−1 (x), ri in R, whence
f (x) = (x − a)g(x)
= (x − a)(x − b)g(x)
= (x − a)(x − b)[r1 p1 (x) + · · · + rn−1 pn−1 (x)]
= r1 (x − a)(x − b)p1 (x) + · · · + rn−1 (x − a)(x − b)pn−1 (x)
1. b. By Theorem 6.6.1, f (x) = ce−x for some constant c. We have 1 = f (1) = ce−1 , so c = e. Thus
f (x) = e1−x .
d. The characteristic polynomial is x2 + x − 6 = (x − 2)(x + 3). Hence f (x) = ce2x + de−3x for
1
some c, d. We have 0 = f (0) = c+d and 1 = f (1) = ce2 +de−3 . Hence, d = −c and c = e2 −e −3
e2x −e−3x
so f (x) = e2 −e−3
.
f. The characteristic polynomial is x2 − 4x + 4 = (x − 2)2 . Hence, f (x) = ce2x + dxe2x = (c +
dx)e2x for some c, d. We have 2 = f (0) = c and 0 = f (−1) = (c − d)e−2 . Thus c = d = 2 and
f (x) = 2(1 + x)e2x .
h. The characteristic polynomial is x2 − a2 = (x − a)(x + a), so (as a 6= −a) f (x) = ceax + de−ax
for some c, d. We have 1 = f (0) = c + d and 0 = f (1) = cea + de−a . Thus d = 1 − c and
1
c = 1−e 2a whence
eax −ea(2−x)
f (x) = cax + (1 − c)e−ax = 1−e2a
j. The characteristic polynomial is x2 + 4x + 5. The roots are λ = −2 ± i, so
f (x) = e−2x (c sin x + d cos x) for some real c and d.
We have 0 = f (0) = d and 1 = f π2 = e−π (c). Hence f (x) = eπ −2x sin x.
4. b. If f (x) = g(x) + 2 then f ′ + f = 2 becomes g′ + g = 0, whence g(x) = ce−x for some c. Thus
f (x) = ce−x + 2 for some constant c.
3
5. b. If f (x) = − x3 then f ′ (x) = −x2 and f ′′ (x) = −2x, so
f ′′ (x) + f ′ (x) − 6 f (x) = −2x − x2 + 2x3
3
Hence, f (x) = −x3 is a particular solution. Now, if h = h(x) is any solution, write g(x) =
3
h(x) − f (x) = h(x) + x3 . Then
g′′ + g′ − 6g = (h′ + h′ − 6h) − ( f ′′ + f ′ − 6 f ) = 0
So, to find g, the characteristic polynomial is x2 + x − 6 = (x − 2)(x + 3). Hence we have
g(x) = ce−3x + de2x , where c and d are constants, so
3
h(x) = ce−3x + de2x − x3
102 Vector Spaces
3 ln(1/2)
6. b. The general solution is m(t) = 10( 45 )1/3 . Hence 10( 45 )t/3 = 5 so t = ln(4/5) = 9.32 hours.
7. b. If m = m(t) is the mass at time t, then the rate m′ (t) of decay is proportional to m(t), that is
m′ (t) = km(t) for some k. Thus, m′ −km = 0 so m = cekt for some constant c. Since m(0) = 10,
1/3
we obtain c = 10, whence m(t) = 10ekt . Also, 8 = m(3) = 10e3k so e3k = 45 , ek = 45 ,
t/3
m(t) = 10(ek )t = 10 45
.
2π 2π
9. In Example 6.6.4, we found that the period of oscillation is √ . Hence √ = 30 so we obtain
k k
π 2
k = 15 = 0.044.
Then y21 + y22 + · · · + y2m = yT y = (Ax)T (Ax) = xT AT Ax = xT 0 = 0. Since the yi are real numbers,
this implies that y1 = y2 = · · · = ym = 0; that is y = 0, that is Ax = 0, that is x is in null A.
7. Linear Transformations
T (x + y) = (x + y) · z = x · z + y · z = T (x) + T (y)
T (rx) = (rx) · z = r(x · z) = rT (x)
103
104 Linear Transformations
h i h i h i
1 1 1
d. Since we know the action of T on −1 and 1 , it suffices to express −7 as a linear
combination of these vectors.
h i h i h i
1 1 1
−7 =r −1 +s 1
In fact, we can find the action of T on any vector a + bx + cx2 in the same way. Observe that
a + bx + cx2 = (a − 2b + 2c) · 1 + (b − c)(x + 2) + c(x2 + x)
for any a, b and c, so
T (a + bx + cx2 ) = (a − 2b + 2c)T (1) + (b − c)T (x + 2) + cT (x2 + x)
= (a − 2b + 2c) · 5 + (b − c) · 1 + c · 0
= 5a − 9b + 9c
= 13 (x − y, 3y, x − y)
nh i h i h i h io h i
1 0
d. Since B = 0 0 , 01 1
0 , 1
1
0
0 , 0
0
0
1 is a basis of M22 , every vector a
c
b
d is
a linear combination
h i h i h i h i h i
a b 1 0 0 1 1 0 0 0
c d =r 0 0 +s 1 0 +t 1 0 +u 0 1
−3s − t −3 1
(" # ) (" # " #)
7s − t 7 1
Hence ker TA = s s, t in R = span 1 , 0 . These vectors are inde-
t 0 −1
pendent so nullity of TA = dim ( ker TA ) = 2. Next
im TA = Ax | x in R4
( " r # )
2 1 −1 3
s
= 1 0 3 1
t r, s, t, u in R
1 1 −4 2
y
2 1 −1 3
= r 1 +s 0 +t 3 +u 1 r, s, t, u in R
1 1 −4 2
Thus im TA = col A as is true in general. Hence dim ( im TA ) = dim ( col A) = rank A, and we
can compute this by carrying A to row-echelon form:
2 1 −1 3 1 0 3 1
1 0 3 1 → 0 1 −7 1
1 1 −4 2 0 0 0 0
Thus dim ( im TA ) = rank A = 2. However, we want a basis of col A, and we obtain this by
writing the columns of A as rows and carrying the resulting matrix (it is AT ) to row-echelon
form: " 2 1 1 # " 1 0 1 # " 1 0 1 #
1 0 1 0 1 −1 0 1 −1
−1 3 −4 → 0 3 −3 → 0 0 0
3 1 2 0 1 −1 0 0 0
1 0
Hence, by Lemma 5.4.2, 0 , 1 is a basis of im TA = col A. Of course this once
1 −1
again shows that rank TA = dim ( col A) = 2.
d. ker TA = {x | Ax = 0} so, as in (b), we use gaussian elimination:
" 2 1 0 # " 1 −1 3 # " 1 0 1
#
1 −1 3 0 3 −6 0 1 −2
1 2 −3 → 0 3 −6 → 0 0 0
0 3 −6 0 3 −6 0 0 0
−t −1
Hence, ker TA = 2t t in R = span 2 . Thus the nullity of TA is dim ( ker TA ) =
t 1
1. As in (b), im TA = col A and we find a basis by doing gaussian elimination on AT :
2 1 1 0 1 −1 2 3 1 0 1 1
1 −1 2 3 → 0 3 −3 −6 → 0 1 −1 −2
0 3 −3 −6 0 3 −3 −6 0 0 0 0
1 0
(" # " #)
0 1
Hence, im TA = col A = span 1 , −1 , so rank TA = dim ( im TA ) = 2.
1 −2
7.2. Kernel and Image of a Linear Transformation 107
{T (1, 0, 0), T (0, 1, 0)} = {(1, 2, 0, 3), (1, −1, −3, 0)}
8. b. If T is onto, let v be any vector in V . Then v = T (r1 , . . . , rn ) for some (r1 , . . . , rn ) in Rn ; that
is v = r1 v1 + · · · + rn vn is in span {v1 , . . . , vn }. Thus V = span {v1 , . . . , vn }. Conversely, if
V = span {v1 , . . . , vn }, let v be any vector in V . Then v is in span {v1 , . . . , vn } so r1 , . . . , rn
exist in R such that
v = r1 v1 + · · · + rn vn = T (r1 , . . . , rn )
Thus T is onto.
10. The trace map T : M22 → R is linear (Example 7.1.2) and it is onto (for example,
r = tr [ diag (r, 0, . . . , 0)] = T [ diag (r, 0, . . . , 0)] for any r in R). Hence the dimension theorem
gives dim ( ker T ) = dim Mnn − dim ( im T ) = n2 − dim (R) = n2 − 1.
12. Define TA : Rn → Rm and TB : Rn → Rk by TA (x) = Ax and TB (x) = Bx for all x in Rn . Then the
given condition means ker TA ⊆ ker TB , so dim ( ker TA ) ≤ dim ( ker TB ). Hence
20. If we can find an onto linear transformation T : Mnn → Mnn with ker T = U and im T = V , then
we are done by the dimension theorem. The condition ker T = U suggests that we define T by
T (A) = A − AT for all A in Mnn . By Example 7.2.3, T is linear, ker T = U , and im T = V . This is
what we wanted.
22. Fix a column y 6= 0 in Rn , and define T : Mmn → Rm by T (A) = Ay for all A in Mmn . This is linear
and ker T = U , so the dimension theorem gives
29. b. Choose a basis {u1 , . . . , um } of U and (by Theorem 6.4.1) let {u1 , . . . , um , . . . , un } be a basis
of V . By Theorem 7.1.3, there is a linear transformation S : V → V such that
S(ui ) = ui if 1 ≤ i ≤ m
S(ui ) = 0 if i > m
110 Linear Transformations
v = r1 u1 + · · · + rm um + · · · + rn un
so
6. b. No inverse. For example T (1, −1, 1, −1) = (0, 0, 0, 0) so (1, −1, 1, −1) is a nonzero vector
in ker T . Hence T is not one-to-one, and so has no inverse.
7.3. Isomorphisms and Composition 111
h i h i
d. T is one-to-one because T ac db = 00 00 implies a + 2c = 0 = 3c − a and b + 2d = 0 =
3d − b, whence a = b = c = d = 0. Thus T is an isomorphism by Theorem 7.3.3. If
h i h i h i h i h i
T −1V ac db = xz wy , then ac db = T xz wy = x3z+−2zx 3w y + 2w
− y . Thus
x + 2z = a
−x + 3z = c
y + 2w = b
−y + 3w = d
7. b. T 2 (x, y) = T [T (x, y)] = T (ky − x, y) = (ky − (ky − x), y) = (x, y) = 1R2 (x, y). Since this
holds for all (x, y) in R2 , it shows that T 2 = 1R2 . This means that T −1 = T .
d. It is a routine verification that A2 = I. Hence
holds for all x in M22 . This means that T 2 = 1M22 , and hence that T −1 = T .
16. Theorem 7.2.5 shows that {T (e1 ), T (e2 ), . . . , T (er )} is a basis of im T . Write
U = span {e1 , . . . , er }. Then B = {e1 , . . . , er } is a basis of U , and T : U → im T carries B to the
basis {T (e1 ), . . . , T (er )}. Thus T : U → im T is itself an isomorphism. Note that T : V → W may
not be an isomorphism, but restricting T to the subspace U of V does result in an isomorphism in
this case.
(x, y) ⊕ (x1 , y1 ) = (x + x1 , y + y1 + 1)
a ⊙ (x, y) = (ax, ay + a − 1)
T : V → R2 by T (x, y) = (x, y + 1)
T [(x, y) ⊕ (x1 , y1 )] = T (x + x1 , y + y1 + 1)
= (x + x1 , (y + y1 + 1) + 1)
= (x, y + 1) + (x1 , y1 + 1)
= T (x, y) + T (x1 , y1 )
7.3. Isomorphisms and Composition 113
Moreover T is one-to-one because T (x, y) = (0, 0) means x = 0 = y + 1, so (x, y) = (0, 1), the
zero vector of V . (Alternatively, T (x, y) = T (x1 , y1 ) implies (x, y + 1) = (x1 , y1 + 1), whence
x = x1 , y = y1 .) As T is clearly onto R2 , it is an isomorphism.
26. b. If p(x) is in ker T , then p(x) = −xp′ (x). If we write p(x) = a0 + a1 x + · · · + an xn , this becomes
28. b. If T = SR, then every vector T (v) in im T has the form T (v) = S[R(v)], whence im T ⊆ im S.
Since R is invertible, S = T R−1 implies im S ⊆ im T , so im S = im T .
Conversely, assume that im S = im T . The dimension theorem gives
a1 g1 + · · · + ar gr + br+1 fr+1 + · · · + bn fn = 0
114 Linear Transformations
apply T to get
R(gi ) = ei for i = 1, 2, . . . , r
R(f j ) = e j for j = r + 1, . . . , n
Then, S is an isomorphism (by Theorem 7.3.1) and we claim that T ST = T . We verify this by
showing that T ST and T agree on the basis {e1 , . . . , er , . . . , en } of V (and invoking Theorem 7.1.2).
Hence, {[1), [2n ), [(−3)n )} is a basis of the space of all solutions to the recurrence. The
general solution is thus,
[xn ) = a[1) + b[2n) + c[(−3)n)
where a, b and c are constants. The requirement that x0 = 1, x1 = 2, x2 = 1 determines a, b,
and c. We have
xn = a + b2n + c(−3)n
for all n ≥ 0. So taking n = 0, 1, 2 gives
a + b + c = x0 = 1
a + 2b − 3c = x1 = 2
a + 4b + 9c = x2 = 1
15 8 3
The solution is a = 20 , b= 20 , c = − 20 , so
1 n+3
xn = 20 (15 + 2 + (−3)n+1 ) n≥0
p(x) = x3 − 3x + 2 = (x − 1)2 (x + 2)
As 1 is a double root of p(x), [1n ) = [1) and [n1n ) = [n) are solutions to the recurrence by
Theorem 7.5.3. Similarly, [(−2)n ) is a solution, so {[1), [n), [(−2)n)} is a basis for the space
of solutions by Theorem 7.5.4. The required sequence has the form
a + c = x0 = 1
a + b − 2c = x1 = −1
a + 2b + 4c = x2 = 1
The solution is a = 95 , b = − 69 , c = 49 , so
1
5 − 6n + (−2)n+2
xn = 9 n≥0
Hence, [1n) = [1), [n1n) = [n) and [n2 1n ) = [n2 ) are solutions and so [1), [n), [n2 ) is a basis
a = x0 = 1
a + b + c = x1 = −1
a + 2b + 4c = x2 = 1
xn = 1 − 4n + 2n2 n≥0
4. b. The recurrence xn+4 = −xn+2 + 2xn+3 has r0 = 0 as there is no term xn . If we write yn = xn+2 ,
the recurrence becomes
yn+2 = −yn + 2yn+1
Now the associated polynomial is x2 − 2x + 1 = (x − 1)2 so basis sequences for the solution
space for yn are [1n ) = [1, 1, 1, 1, . . . ) and [n1n ) = [0, 1, 2, 3, . . . ). As yn = xn+2 , corre-
sponding basis sequences for xn are [0, 0, 1, 1, 1, 1, . . . ) and [0, 0, 0, 1, 2, 3, . . . ). Also,
[1, 0, 0, 0, 0, 0, . . . ) and [0, 1, 0, 0, 0, 0, . . . ) are solutions for xn , so these four sequences
form a basis for the solution space for xn .
7. The sequence has length 2 and associated polynomial x2 + 1. The roots are nonreal: λ1 = i and
λ2 = −i. Hence, by Remark 2,
[in + (−i)n ) = [2, 0, −2, 0, 2, 0, −2, 0, . . . ) and [i(in − (−i)n )) = [0, −2, 0, 2, 0, −2, 0, 2, . . . )
are solutions. They are independent as is easily verified, so they are a basis for the space of solutions.
8. Orthogonality
e1 = x1 = (2, 1)
e2 = x2 − x2 ·e12 e1
ke1 k
= (1, 2) − 45 (2, 1)
= 15 {(5, 10) − (8, 4)}
= 35 (−1, 2)
In hand calculations, {(2, 1), (−1, 2)} may be a more convenient orthogonal basis.
d. If x1 = (0, 1, 1), x2 = (1, 1, 1), x3 = (1, −2, 2) then
e1 = x1 = (0, 1, 1)
e2 = x2 − x2 ·e12 e1 = (1, 1, 1) − 22 (0, 1, 1) = (1, 0, 0)
ke1 k
x3 ·e1
e 3 = x3 − e1 − x3 ·e22 e2 = (1, −2, 2) − 02 (0, 1, 1) − 11 (1, 0, 0) = (0, −2, 2)
ke1 k2 ke2 k
2. b. Write e1 = (3, −1, 2) and e2 = (2, 0, −3). Then {e1 , e2 } is orthogonal and so is an orthogonal
basis of U = span {e1 , e2 }. Now x = (2, 1, 6) so take
x·e1
x1 = projU x = e1 + x·e22 e2
ke1 k2 ke2 k
17 14
= 14 (3, −1, 2) − 13 (2, 0, −3)
1
= 182 (271, −221, 1030)
1
Then x2 = x − x1 = 182 (93, 402, 62). As a check: x2 is orthogonal to both e1 and e2 (and so
is in U ⊥ ).
d. If e1 = (1, 1, 1, 1), e2 = (1, 1, −1, −1), e3 = (1, −1, 1, −1) and x = (2, 0, 1, 6), then
{e1 , e2 , e3 } is orthogonal so take
x1 = projU x = x·e1
e1 + x·e22 e2 + x·e32 e3
ke1 k2 ke2 k ke3 k
= 94 (1, 1, 1, 1) − 45 (1, 1, −1, −1) − 34 (1, −1, 1, −1)
1
= 4 (1, 7, 11, 17)
Then, x2 = x − x1 = 41 (7, −7, −7, 7) = 74 (1, −1, −1, 1). Check: x2 is orthogonal to each ei ,
hence x2 is in U ⊥.
117
118 Orthogonality
c. projU x = − 15 3
14 (1, 0, 2, −3) + 70 (4, 7, 1, 2) =
3
10 (−3, 1, −7, 11).
4. b. U = span {(1, −1, 0), (−1, 0, 1)} but this basis is not orthogonal. By Gram-Schmidt:
e1 = (1, −1, 0)
e2 = (−1, 0, 1) − (−1, 0, 1)·(1, −1,
2
0)
(1, −1, 0) = − 12 (1, 1, −2)
k(1, −1, 0)k
So we use U = span {(1, −1, 0), (1, 1, −2)}. Then the vector x1 in U closest to x = (2, 1, 0)
is
x1 = projU x = 2−1+0
2 (1, −1, 0) +
2+1+0
6 (1, 1, −2) = (1, 0, −1)
d. The given basis of U is not orthogonal. The Gram-Schmidt algorithm gives
e1 = (1, −1, 0, 1)
e2 = (1, 1, 0, 0) = (1, 1, 0, 0) − 03 e1 = (1, 1, 0, 0)
e3 = (1, 1, 0, 1) − 31 (1, −1, 0, 1) − 22 (1, 1, 0, 0) = 13 (−1, 1, 0, 2)
Given x = (2, 0, 3, 1), we get (using e′3 = (−1, 1, 0, 2) for convenience)
projU x = 33 (1, −1, 0, 1) + 22 (1, 1, 0, 0) + 06 (−1, 1, 0, 2) = (2, 0, 0, 1).
h i h i h i
1 −1 2 1 1 −1 2 1 1 0 −1 1
5. b. Here A = 1 0 −1 1 → 0 1 −3 0 → 0 1 −3 0 . Hence, AxT = 0 has x =
(s − t, 3s, s, t) = s(1, 3, 1, 0) + t(−1, 0, 0, 1).
Thus U ⊥ = span {(1, 3, 1, 0), (−1, 0, 0, 1)}.
8. If x = projU x then x is in U by Theorem 8.1.3. Conversely, if x is in U , let {f1 , . . . , fm } be an
orthogonal basis of U . Then the expansion theorem (applied to the space U ) gives x = ∑i x·fi2 fi =
kfi k
projU x by the definition of the projection.
10. Let {f1 , . . . , fm } be an orthonormal basis of U . If x is in U then, since kfi k = 1 for each i, so
x = (x · f1 )f1 + · · · + (x · fm )fm = projU x by the expansion theorem (applied to the space U ).
yT
1
..
14. If {y1 , . . . , ym } is a basis of U ⊥, take A = . . Then Ax = 0 if and only if yTi x = 0 for each
yTm
0
i; if and only if yi · x = 0 for each i; if and only if x is in (U ⊥ )⊥ = U ⊥⊥ = U . This shows that
U = {x in Rn | Ax = 0}.
8.2. Orthogonal Diagonalization 119
is orthogonal.
√ √
h. Each row has length 4 + 36 + 9 = 49 = 7. Hence
" 2 6 3
# " #
7 7 −7 2 6 −3
1
3
7
2
7
6
7 = 7
3 2 6
− 67 3
7
2
7 −6 3 2
is orthogonal.
2. Let P be orthogonal, so P−1 = PT . If P is upper triangular, so also is P−1 , so P−1 = PT is both upper
triangular (P−1 ) and lower triangular PT ). Hence, P−1 = PT is diagonal, whence P = (P−1 )−1 is
diagonal. In particular, P is symmetric so P−1 = PT = P. Thus P2 = I. Since P is diagonal, this
implies that all diagonal entries are ±1.
x−3 0 −7
d. cA (x) = 0 x−5 0 = (x − 5)(x2 − 6x − 40) = (x − 5)(x + 4)(x − 10). Hence the eigen-
−7 0 x−3
λ1 = 5, λ2= 10,
values are λ3 = −4.
2 −7 0 1 0 0 0
λ1 = 5: 0 0 0 → 0 0 1 ; E5 (A) = span .
1
−7 2 0
0 0 0
0
7 0 −7 1 0 −1 1
λ2 = 10: 0 5 0 → 0 1 0 ; E10 (A) = span 0 .
−7 0 7 0 0 0
1
−7 0 −7 1 0 1 1
λ3 = −4: 0 −9 0 → 0 1 0 ; E−4 (A) = span 0 .
−7 0 −7 0 0 0 −1
Note that the three eigenvectors are pairwise orthogonal (as Theorem 8.2.4 asserts). Normal-
izing them gives an orthogonal matrix
" #
0 √12 √1
2
1 √0 1 1
P = 1 01 0
1
= √2 2 0 0
0 √ − √2 0 1 −1
2
5 0 0
Then P−1 = PT and PT AP = 0 10 0 .
0 0 −4
x−5 2 4 x−9 0 9−x x−9 0 0
f. cA (x) = 2 x−8 2 = 2 x−8 2 = 2 x−8 4
4 2 x−5 4 2 x−5 4 2 x−1
= (x − 9) x −2 8 x −4 1
= (x − 9)(x2 − 9x) = x(x − 9)2 .
The eigenvalues
are λ1 =0, λ2 = 9.
−5 2 4 1 −4 1 1 −4 1 1 0 −1 2
λ1 = 0 : 2 −8 2 → 0 −18 9 → 0 1 − 21 → ; E0 (A) = span . 0 1 − 21 1
2 −5
4
0 1 18 −9 0 0 0
0 0 0 2
4 2 4 1 2 1 −1 −1
λ2 = 9: 2 1 2 → 0 0 0 ; E9 (A) = span 0 . , 2
4 2 4 0 0 0 1 0
−1
However, these are not orthogonal and the Gram-Schmidt algorithm replaces 2 with Z2 =
0
" 2 √−1 √1 # " √ #
2 2 −3 1
1 3 2 3 2 √
1
−4 . Hence P = 1
3 0 −4
√
3 2 = √
√
2 0 −4 is orthogonal and satisfies PT AP =
1 2 √1 1 3 2
3
√ 2 2 3 1
2 3 2
0 0 0
0 9 0 .
0 0 9
−2 1
We note in passing that and 2 2 are another orthogonal basis of E9 (A), so Q =
1
−2
2 −2 1 0 0 0
1
3
1 2 2 also satisfies QT AQ = 0 9 0 .
2 1 −2 0 0 9
−3 −5 1 −1 −3 −5 1 −1 1 −1 −3 −5
" # " # " #
−5 −3 −1 1 −8 −8 0 0 0 −8 −8 −16
λ1 = 0 : 1 −1 −3 −5 → 1 −1 −3 −5 → 0 −16 −24 −40
−1 1 −5 −3 0 0 −8 −8 0 0 1 1
1 0 −2 −3 1 0 0 −1 1
" # " # (" #)
0 1 1 2 0 1 0 1 −1
→ 0 0 1 1 → 0 0 1 1 ; E0 (A) = span −1
0 0 1 1 0 0 0 0 1
−5 −7
1 −1 1 −1 −7 −5 1 0 −3 −2
" # " # " #
−7 −1 −5 1 0 −12 −48 −36
λ2 = −4 : −1 −7 −5
1 → 0 −12 −36 −24 → 00 10 −14 −13
−1
" 1 10 −5 0
−7
1
# 0 0 −12 −12
(" −1 #) 0 0 1 1
→ 00 01 10 −11 ; E−4 (A) = span −1
1
0 0 0 0
" 5 −5 1 −1
# " 1 −1 5 −5
# 1 " 1 −1 0 0 #
−5 5 −1
λ3 = 8 : 1 −1 5 −5
1
→ 00 00 −24 24 −24
24
→ 00 00 10 −10 ;
−1 1 −5 5
(" 1
# " 00 #)0 0 0 0 0 0 0
1
E8 (A) = span 0 , 01
0 1
1 1 √1
√
2 − 2 0
1 −1 2 0
" 0 0 0 0
2
#
1 1 1 √
− √ 0 −1 1 2 0 −4
Hence, P = − 21 −21 02 √1 = 12 −1 −1 0 √2 gives PT AP = 00 0
0
8
0
0 .
2 2 2 √
1 1 √1 1 1 0 2 0 0 0 8
0
2 2 2
x −a 0
x −c −a 0
6. cA (x) = −a x −c =x −c x +a −c x = x(x2 − c2 ) − a2 x = x(x2 − k2 ).
0 −c x
Hence cA (x) = x(x − k)(x + k), where k2 = a2 + c2 , so the eigenvalues are λ1 = 0, λ2 = k, λ3 = −k.
They areall distinct (k6= 0, and
a 6= 0 or c 6= 0) so the eigenspaces
are all one dimensional.
0 −a 0 c 0 c
λ1 = 0 : −a 0 −c 0 = 0 ; E0 (A) = span 0
0 −c
0
−a 0 −a
k −a 0 a 0 a
λ2 = k : −a k −c k = 0 ; Ek (A) = span k
0 −c k c 0 c
−k −a 0 a 0 a
λ3 = −k : −a −k −c −k = 0 ; E−k (A) = span −k
0 −c −k c 0 c
√ √
These eigenvalues
√ are orthogonal
and have length, k, 2k, 2k respectively. Hence,
c 2 a a 0 0 0
P= √1 0√ k −k is orthogonal and PT AP = 0 k 0 .
2k −a 2 c c 0 0 −k
h i
10. Similar to Example 8.2.6, q has matrix A = 12 −22 with eigenvalues λ1 = −3 and λ2 = 2 and
h i h i h i
−1 2 1 −1 2
corresponding eigenvectors x1 = and x2 = 1 respectively. Hence P = √ is
h i 2 5 2 1
Since A = PDP−1 , we obtain A2 = (PDP−1 )2 = PD2 P−1 = PIP−1 = I. Since A is symmetric, this
proves (a).
13. b. Let A and B be orthogonally similar, say B = PT AP where PT = P−1 . Then B2 = PT APPT AP =
PT AIAP = PT A2 P. Hence A2 and B2 are orthogonally similar.
15. Assume that (Ax) · y = x · Ay for all columns x and y; we must show that AT = A. We have (Ax) · y =
xT AT y and x · Ay = xT Ay, so the given condition asserts that
Since (∗) shows that AT and A have the same (i, j)-entry for each i and j. In other words, AT = A.
Note that the same argument shows that if A and B are matrices with the property that xT By = xT Ay
for all columns x and y, then B = A.
h i h i
cos θ sin θ cos θ sin θ
18. b. If P = − sin θ cos θ and Q = sin θ − cos θ then P and Q are orthogonal matrices, det P = 1
and det Q = −1. (We note that every 2 × 2 orthogonal matrix has the form of P or Q for some
θ .)
d. Since P is orthogonal, PT = P−1 . Hence
Hence, if I − P is invertible, then det (I − P) 6= 0 so this gives det PT = (−1)n ; that is det P =
(−1)n , contrary to assumption.
21. By the definition of matrix multiplication, the (i, j)-entry of AAT is ri · r j . This is zero if i 6= j,
and equals kri k2 if i = j. Hence, AAT = D = diag (kr1 k2 , kr2 k2 , . . . , krnk2 ). Since D is invertible
(kri k2 6= 0 for each i), it follows that A is invertible and, since row i of AT is a1i a2i · · · a ji · · · ani
. . .
kr11k2 0 ··· 0
.. .. ..
−1 T −1
··· ··· 0 kr21k2 ··· 0
A =A D = a 1i ··· a ji ··· a ni .
.. .
.. .. .
..
. .. .. . .
. ··· . ··· . 0 0 ··· 1
krn k2
a ji
Thus, the (i, j)-entry of A−1 is 2 .
kr j k
23. b. Observe first that I − A and I + A commute, whence I − A and (I + A)−1 commute. Moreover,
T −1
(I + A)−1 = (I + A)T = (I T + AT )−1 = (I − A)−1 . Hence,
8.4 QR-Factorization
h i h i
2 1
1. b. The columns of A are c1 = 1and c2 = 1 . First apply the Gram-Schmidt algorithm
h i
f1 = c1 = 21
h i h i h i
c2 ·f1 1 2 − 51
f2 = c2 − f = − 53 =
kf1 k2 1 1 1 2
5
Then A = QR.
T T
d. The columns of A are c1 = 1 −1 0 1 , c2 = 1 0 1 −1 and
T
c3 = 0 1 1 0 . Apply the Gram-Schmidt algorithm
T
f1 = c1 = 1 −1 0 1
T 0 T
f2 = c2 − c2 ·f12 f1 = 1 0 1 −1
− 3 F1 = 1 0 1 −1
kf1 k
c3 ·f1
f3 = c3 − f1 − c3 ·f22 f2
kf1 k2 kf2 k
T −1 T T
− 31
= 0 1 1 0 − 3 1 −1 0 1 1 0 1 −1
T
= 23 0 1 1 1
Normalize
1
T
√1
Q1 = kf1 k f1 = 1 −1 0 1
3
1 1
T
Q2 = kf2 k f2 = √ 1 0 1 −1
3
1 1
T
Q3 = kf3 k f3 = √
3
0 1 1 1
1 1 0
" #
−1 0 1
√1
Hence Q = q1 q2 q3 = 3 0 1 1 has orthonormal columns. We obtain R from
1 −1 1
equation (8.5) preceding Theorem 8.4.1:
√ −1
" √1
# " #
kf1 k c2 · q1 c3 · q1 3 √
3 0 −1
√3 3
R= 0 kf2 k c3 · q2 = 0 3 √1
3 = √1 0 3 1
0 0 kf3 k √2
3
0 0 0 0 2
3
Then A = QR.
8.5. Computing Eigenvalues 125
h i
2
These are approaching (scalar multiples of) the dominant eigenvector −1 . The Rayleigh
x ·x
quotients are rk = k k+1 , k = 0, 1, 2, . . . , so r0 = 1, r1 = 3.29, r2 = 4.23, r3 = 3.94. These
kxk k2
are approaching the dominant eigenvalue 4.
h i
− 3 −1
√
d. A = 31 10 ; cA (x) = x−1 x = x2 − 3x − 1, so the eigenvalues are λ1 = 21 (3 + 13),
√ √
λ2 = 12 (3 − 13). Thus the dominant eigenvalue is λ1 = 21 (3 + 13). Since λ1 λ2 = −1 and
λ1 + λ2 = 3, we get h i h i
λ1 − 3 −1 1 −λ1
−1 λ1 → 0 0
h i h i
so a dominant eigenvector is λ11 . We start with x0 = 11 .
Then xk+1 = Axk , k = 0, 1, . . . gives
h i h i h i h i
x1 = 41 , x2 = 134 , x3 = 43 13 , x 4 = 142
43
h i h i
λ1 3.302776
These are approaching scalar multiples of the dominant eigenvector 1 = 1 . The
xk ·xk+1
Rayleigh quotients are rk = :
kxk k2
h i h i h i
1 33 1 1 33 1 109 3
A 2 = R 1 Q1 = = Q2 R2 where Q2 = √1090 √1
10 1 −3 1 −33 , R2 = 1090 0 10 .
h i h i
360 1
1
A3 = R2 Q2 = 109 1 −33 = 3.302752 0.009174
0.009174 −0.302752 .
The diagonal entries already approximate λ1 and λ2 to 4 decimal places.
4. b. t σ1 , . . . , t σr .
8. b. First AT A = In so ΣA = In .
√1
1 1
1 0 √1 1 1
A = 2 1 −1
0 1 2 −1 1
1 −1 √1 −1 1
= √12
1 1 2 1 1
−1 0
=
0 1
9. b.
A=F
1 1 1 1
" #
h ih i
1 3 4 20 0 0 0 1 1 −1 1 −1
= 5 4 −3 0 10 0 0 2 1 1 −1 −1
1 −1 1 −1
2. b. Not orthogonal: h(i, −i, 2 + i), (i, i, 2 − i)i = i(−i) + (−i)(−i) + (2 + i)(2 + i) = 3 + 4i
d. Orthogonal: h(4 + 4i, 2 + i, 2i), (−1 + i, 2, 3 − 2i)i = (4 + 4i)(−1 − i) + (2 + i)2 + (2i)(3 +
2i) = (−8i) + (4 + 2i) + (−4 + 6i) = 0.
v + w = ((v + v′ ) + (w + w′ ), (v + v′ ) − 2(w + w′ ), (v + v′ )) is in U
zv = (zv + zw, zv − 2zw, zv) is in U
0 = (0 + 0, 0 − 20, 0) is in U
Hence U is a subspace.
h i h i
4 3−i x−4 −3 + i
8. b. A = 3+i , cA (x) = −3 = x2 − 5x − 6 = (x + 1)(x − 6).
1
h −i x−1 i h i h i
−5 −3 + i 3+i 2 −2
Eigenvectors for λ1 = −1: −3 − i → 0 ; an eigenvector is x1 = .
h 1
i h 0
i h 3+i i
Eigenvectors for λ2 = 6: −32− i −35+ i → 02 −30+ i ; an eigenvector is x2 = 3 −2 i .
√ 1
h
−2 3−i
i
As x1 and x2 are orthogonal and kx1 k = kx2 k = 14, U = √ is unitary and
h i 14 3 + i 2
U H AU = −10 06 .
h i
d. A = 1 −2 i 1 +3 i ; cA (x) = −1 x−2
+i
−1 − i
x−3 = x2 − 5x + 4 = (x − 1)(x − 4).
h i h i h i
Eigenvectors for λ1 = 1: −1−1+ i −1−2− i → 10 1 +0 i ; an eigenvector is x1 = 1−1 +i
.
h i h i h i
Eigenvectors for λ2 = 4: −12+ i −11− i → −10+ i 01 ; an eigenvector is x2 = 1 −1 i .
√ h
+i
i
Since x1 and x2 are orthogonal and kx1 k = kx2 k = 3, U = √13 1−1 1
1−i is unitary and
h i
U H AU = 10 04 .
1 0 0 x−1 0 0
f. A = 0 1 1 + i ; cA (x) = 0 x−1 −1 − i = (x − 1)(x2 − 3x) = (x − 1)x(x − 3).
0 1−i −1 + i
2
0
x − 2
0 0 0 0 1 0 1
Eigenvectors for λ1 = 1: 0 0 −1 − i → 0 0 1 ; an eigenvector is x1 = 0 .
0 −1 + i
−1 0 0 0 0
−1 0 0 1 0 0 0
If λ2 = 0: 0 −1 −1 − i → 0 1 1 + i ; an eigenvector is x2 = 1 + i .
0 −1 + i −i 0 0 −1
0
2 0 0 1 0 0 0
Eigenvectors for λ3 = 3: 0 2 −1 − i → 0 −1 + i 1 ; an eigenvector is x3 = 1 .
0 −1 + i 1 0 0 0 1−i
√
√
3 0 0
Since {x1 , x2 , x3 } is orthogonal and kx2 k = kx3 k = 3, U = √1 0 1+i 1 is orthog-
3 0 −1 1−i
1 0 0
onal and U ∗AU = 0 0 0 .
0 0 3
10. b. (1) If z = (z1 , z2 , . . . , zn ) then kzk2 = |z1 |2 + |z2 |2 + · · · + |zn |2 . Thus kzk = 0 if and only if
|z1 | = · · · = |zn | = 0, if and only if z = (0, 0, . . . , 0).
(2) By Theorem 8.7.1, we have hλ z, wi = λ hz, wi and hz, λ wi = λ hz, wi. Hence
b. If A is hermitian then A = AT . If A = ai j , the (k, k)-entry of A is akk , and the (k, k)-entry of
11.
AT is akk . Thus, A = AT implies that akk = akk for each k; that is akk is real.
AH = 12 (Z H + Z HH ) = 21 (Z H + Z) = A
BH = 12 (Z H − Z HH ) = 21 (Z H − Z) = −B
16. b. If U is unitary, then U −1 = U H . We must show that U −1 is unitary, that is (U −1 )−1 = (U −1)H .
But
(U −1)−1 = U = (U H )H = (U −1 )H
h i h i h i
1 i 1 −i T i −1
18. b. If V = −i 0 then V is hermitian because V = i 0 = V , but iV = 1 0 is not
hermitian (it has a nonreal entry on the main diagonal).
h i h i
0 1 a b
21. b. Given A = −1 0 , let U = c d be invertible and real, and assume that U −1AU =
h i h i
λ µ λ µ
0 ν . Thus, AU = U 0 ν so
h i h i
c d aλ aµ + bν
−a −b = cλ cµ + d ν
1. b. The elements with inverses are 1, 3, 7, 9: 1 and 9 are self-inverse; 3 and 7 are inverses of each
other. As for the rest, 2 · 5 = 4 · 5 = 6 · 5 = 8 · 5 = 0 in Z10 so 2, 5, 4, 6 and 8 do not have
inverses in Z10 .
d. The powers of 2 computed in Z10 are: 2, 4, 8, 16 = 6, 32 = 2, . . . , so the sequence repeats: 2,
4, 8, 16, 2, 4, 8, 16, . . . .
3. b. We want a number a in Z19 such that 11a = 1. We could try all 19 elements in Z19 , the one
that works is a = 7. However the euclidean algorithm is a systematic method for finding a. As
in Example 8.8.2, first divide 19 by 11 to get
19 = 1 · 11 + 8
130 Orthogonality
1 = 3 − 1 · 2 = 3 − (8 − 2 · 3) = 3 · 3 − 8
= 3(11 − 1 · 8) − 8 = 3 · 11 − 4 · 8
= 3 · 11 − 4(19 − 1 · 11) = 7 · 11 − 4 · 19
b. Workinghin Z7 , weihavehdet A = −1 −1
6. i 15h− 24 =i 1 + 4 = 5 6= 0, so A exists. Since 5 = 3 in Z7 ,
A−1 = 3 −43 −65 = 3 33 15 = 22 31 .
7. b. Gaussian elimination works over any field F in the same way that we have been using it over
R. In this case we have F = Z7 , and we reduce the augmented matrix of the system as follows.
We have 5 · 3 = 1 in Z7 so the first step in the reduction is to multiply row 1 by 5 in Z7 :
h i h i h i h i h i
3 1 4 3 1 5 6 1 1 5 6 1 1 5 6 1 1 0 5 3
4 3 1 1 → 4 3 1 1 → 0 4 5 4 → 0 1 3 1 → 0 1 3 1
Hence x and y are the leading variables, and the non-leading variable z is assigned as a param-
eter, say z = t. Then, exactly as in the real case, we obtain x = 3 + 2t, y = 1 + 4t, z = t where t
is arbitrary in Z7 .
10. b. The minimum weight of C is 5, so it detects 4 errors and corrects 2 errors by Theorem 8.8.5.
11. b. The linear (5, 2)-code {00000, 01110, 10011, 11101} has minimum weight 3 so it corrects 1
error by Theorem 8.8.5.
13. b. C = {00000, 10110, 01101, 11011} is a linear (5, 2)-code of minimal weight 3, so it corrects
single errors.
h i
u
14. b. G = 1 u where u is any nonzero vector in the code. H = In−1 .
8.9. An Application to Quadratic Forms 131
Finally, q = 3y21 − y22 , the index of q is 1 (the number of positive eigenvalues) and the rank of q
is 2 (the number of nonzero eigenvalues).
7 4 4
T
d. q = x Ax where A = 4 1 −8 . To find cA (x), subtract row 2 from row 3:
4 −8 1
x−7 −4 −4 x−7 −4 −4
cA (x) = −4 x−1 8 = −4 x−1 8
−4 8 x−1 0 −x + 9 x−9
x−7 −8 −4
= −4 x+7 8 = (x − 9)2 (x + 9)
0 0 x−9
2 −4 −4 1 −2 −2
2 2
λ1 = 9: −4 →8 8 ; orthogonal eigenvectors
0 0 2 ,
0 −1 .
−4 8 8 0 0 −1
0
2
−16
−4 −4 4 1 1 4 0 2 −1
λ2 = −9: −48 → 0 −9 9 → 0 1 −1 ; eigenvector
−10 2 . These
−4
−10 8 0 9 −9 0 0 0
2
2 2 −1
eigenvectors are orthogonal and each has length 3. Hence, P = 31 2 −1 2 is orthog-
−1 2 2
9 0 0
onal and PT AP = 0 9 0 . Thus
0 0 −9
2 2 −1 x1 2x1 + 2x2 − x3
T 1 1
y=P x= 3
2 −1 2 x2 = 3
2x1 − x2 + 2x3
−1 2 2 x3 −x1 + 2x2 + 2x3
so
y1 = 31 [2x1 + 2x2 − x3 ]
y2 = 13 [2x1 − x2 + 2x3 ]
y3 = 13 [−x1 + 2x2 + 2x3 ]
will give q = 9y21 + 9y22 − 9y23 . The index of q is 2 and the rank of q is 3.
132 Orthogonality
5 −2 −4
f. q = xT Ax where A = −2 8 −2 . To find cA (x), subtract row 3 from row 1:
−4 −2 5
x−5 2 4 x−9 0 −x + 9
cA (x) = 2 x−8 2 = 2 x−8 2
4 2 x−5 4 2 x−5
x−9 0 0
= 2 x−8 4 = x(x − 9)2
4 2 x−1
4 2 4 2 1 2 −2 1
λ1 = 9: →
2 1 2 ; orthogonal eigenvectors are
0 0 0 2 and 2 .
4 2 4 0 0 0 1 −2
−5 2
4 1 −4 1 1 0 −1 2
λ2 = 0: 2 −8
2 → 0 −18 9 → 0 2 −1 ; an eigenvector is 1 .
4 −5
2 0 18 −9 0 0 0 2
−2 1 2
These eigenvectors are orthogonal and each has length 3. Hence P = 13 2 2 1 is or-
1 −2 2
9 0 0
T
thogonal and P AP = 0 9 0 . If
0 0 0
−2 2 1 x1
T 1
y=P x= 3
1 2 −2 x2
2 1 2 x3
then
y1 = 13 (−2x1 + 2x2 + x3 )
y2 = 13 (x1 + 2x2 − 2x3 )
y3 = 13 (2x1 + x2 + 2x3 )
1 1 0 1 1 0 1 0 1 −1
λ1 = 2: 1 2 −1 → 0 1 −1 → 0 1; an eigenvector is
−1 . 1
0 −1 1 0 −1 1 0 0 0 1
0 1 0 1 1 −1 1 0 −1 1
λ2 = 1: 1 1 −1 → 0 1 0 → 0 1 0 ; an eigenvector is 0 .
0 −1 0 0 0 0 0 0 0 1
−2 1 0 1 −1 −1 1 0 1 1
λ3 = −1: 1 −1 −1 → 0 −1 −2 → 0 1 2 ; an eigenvector is 2 .
0 −1 −2 0 −1 −2 0 0 0 −1
Hence,
√ √
− √13 √1 √1
" #
2 6 − 2 3 1
√
P= √1
3
0 √2 = √1 2 0 2
6 √ √
6
√1 √1 − √16 2 3 −1
3 2
8.9. An Application to Quadratic Forms 133
2 0 0
is orthogonal and PT AP = 0 1 0 . If
0 0 −1
√ √ √
−√ 2 2 √2 x1
y=P x= T √1 3 0 3 x2
6 1 2 −1 x3
then
y1 = √1 (−x1 + x2 + x3 )
3
1
y2 = √ (x1 + x3 )
2
y3 = √1 (x1 + 2x2 − x3 )
6
2 2
Such an angle 2θ exists because √ 2a−c 2 + √ 2b 2 = 1.
b +(a−c) b (a−c)
134 Orthogonality
x1 1 2 −2
xT Ax + Bx =
7. b. The equation is 7 where x = x2 ,A= 2 3 0 , B = 5 0 −6 .
x3 −2 0 3
t −1 −2 2 t −1 −2 2 t −1 −4 2
cA (x) = −2 t −3 0 = −2 t −3 0 = −2 t −3 0
2 0 t −3 0 t −3 t −3 0 0 t −3
2 −2 2 1 0 0 0
λ1 = 3: −2 0 0 → 0 1 −1 ; an eigenvector is 1 .
2 0 0
0 0 0
1
4 −2 2 −2 2 0 1 0 1 1
λ2 = 5: −2 2 0 → 0 2 2 → 0 1 1 ; an eigenvector is 1 .
−1
2 0 2 0 2 2
0 0 0
−2 −2 2 1 1 −1 1 0 −2 2
λ3 = −1: −2 −4 0 → 0 −2 −2 → 0 1 1 ; an eigenvector is −1 .
2 0 −4 0 −2 −2 0 0 0 1
√
√1 √2
" #
0 3 6 0 2 2
3 0 0
√ √
Hence, P = √1
2
− √13 − √16 = √1 3 2 −1 T
satisfies P AP = 0 5 0 . If
6 √ √ 0 0 −1
√1 + √13 √1 3 − 2 1
2 6
√
y1 √ 3(x2 + x3 )
y= y2 T
=P x= √1 2(x1 + x2 − x3 )
y3 6 2x1 − x2 + x3
then
y1 = √1 (x2 + x3 )
2
1
y2 = √ (x1 + x2 − x3 )
3
y3 = √1 (2x1 − x2 + x3 )
6
Hence
Thus
0 0 1 0
= 0 1 0 0
0 0 0 1
137
138 Change of Basis
1 2
" #
5 3
4. b. MDB (T ) = CD [T (1, 1)] CD [T (1, 0)] CD (1, 5, 4, 1) CD (2, 3, 0, 1)
= = 4 0 .
1 1
h i
b
We have v = (a, b) = b(1, 1) + (a − b)(1, 0) so CB (v) = a − b . Hence,
1 2 2a − b
" # " #
h i
5 3 b 3a + 2b
CD [T (v)] = MDB (T )CB (v) = 4 0 a−b = 4b
1 1 a
h i a
h i
1 1 1 −1 1 a+b−c
CD [T (v)] = MDB (T )CB (v) = 2 1 1 1
b = 2 a+b+c
c
T S
5. b. Have R3 → R4 → R2 . Let B, D, E be the standard bases. Then
MED (S) = CE [S(1, 0, 0, 0)] CE [S(0, 1, 0, 0)] CE [S(0, 0, 1, 0)] CE [S(0, 0, 0, 1)]
= CE (1, 0) CE (1, 0) CE (0, 1) CE (0, −1)
9.1. The Matrix of a Linear Transformation 139
h i
1 1 0 0
= 0 0 1 −1
MDB (T ) = CD [T (1, 0, 0)] CD [T (0, 1, 0)] CD [T (0, 0, 1)]
= CD (1, 0, 1, −1) CD (1, 1, 0, 1) CD (0, 1, 1, 0)
" 1 1 0 #
0 1 1
= 1 0 1
−1 1 0
T S
d. Have R3 → P2 → R2 with bases B = {(1, 0, 0), (0, 1, 0), (0, 0, 1)}, D = 1, x, x2 ,
MDB (T ) = CD [T (1, 0, 0)] CD [T (0, 1, 0)] CD [T (0, 0, 1)]
= CD (1 − x) CD (−1 + x2 ) CD (x)
1 −1 0
= −1 0 1
0 1 0
MEB (ST ) = CE [ST (1, 0, 0)] CE [ST (0, 1, 0)] CE [ST (0, 0, 1)]
= CE (2, 0) CE (−1, 1) CE (−1, 0)
h i
2 −1 −1
= 0 1 0
7. b. MDB (T ) = CD [T (1, 0, 0)] CD [T (0, 1, 0)] CD [T (0,
0, 1)]
= CD (0, 1, 1) CD (1, 0, 1) CD (1, 1, 0)
0 1 1
= 1 0 1
1 1 0
If T −1 (a, b, c) = (x, y, z) then (a, b, c) = T (x, y, z) = (y + z, x + z, x + y). Hence, y + z = a,
x + z = b, x + y = c. The solution is
T −1 (a, b, c) = (x, y, z) = 12 (b + c − a, a + c − b, a + b − c)
Hence,
MBD (T −1 ) = CB [T −1 (1, 0, 0)] CB [T −1 (0, 1, 0)] CB [T −1 (0, 0, 1)]
= CB − 12 , 12 , 21 CB 12 , − 12 , 12 CB 12 , 21 , − 12
−1 1 1
1
=2 1 −1 1
1 1 −1
If v = (a, b, c, d) then
1 −1 0 0 a−b
" #" a
# " #
0 1 −1 0 b−c
CB [T −1 (v)] = MDB (T −1 )CD (v) = 0 0 1 0
b
c = c
0 0 0 1 d d
9.1. The Matrix of a Linear Transformation 141
12. Since D = {T (e1 ), . . . , T (en )}, we have CD T (e j ) = C j = column j of In . Hence,
MDB (T ) = CD [T (e1 )] CD [T (e2 )] · · · CD [T (en )]
= C1 C2 · · · Cn = In
16. b. Define T : Pn → Rn+1 by T [p(x)] = (p(a0 ), p(a1 ), . . . , p(an )), where a0 , . . . , an are fixed
distinct real numbers. If B = {1, x, . . . , xn } and D ⊆ Rn+1 is the standard basis,
Since the ai are distinct, this matrix has nonzero determinant by Theorem 3.2.7. Hence, T is
an isomorphism by Theorem 9.1.4.
R S, T
20. d. Assume that V → W → U . Recall that the sum S + T : W → U of two operators is defined by
(S + T ) (w) = S(w) + T (w) for all w in W . Hence, for v in V :
22. b. If T is in X10, then T (v) = 0 for all v in X1 . As X ⊆ X1 , this implies that T (v) = 0 for all v in
X ; that is T is in X 0. Hence, X10 ⊆ X 0 .
Hence, Sv = T as required.
E j (v) = E j (r1 b1 + r2 b2 + · · · + r j b j + · · · + rn bn )
= r1 E j (b1 ) + r2 E j (b2 ) + · · · + r j E j (b j ) + · · · + rn E j (bn )
= r1 · 0 + r2 · 0 + · · · + r j · 1 + · · · + rn · 0
= rj
− 23 1
−1 2
−3 −2 1
1
(x2 )
1. b. PD←B = CD (x) CD (1 + x) CD = 1 1 0 = 2
2 2 0 because
0 0 1 0 0 2
x = − 32 · 2 + 1(x + 3) + 0(x2 − 1)
1 + x = (−1) · 2 + 1(x + 3) + 0(x2 − 1)
x2 = 12 · 2 + 0(x + 3) + 1(x2 − 1)
9.2. Operators and Similarity 143
Given v = 1 + x + x2 , we have
− 12
0
CB (v) = 1 and CD (v) = 1
1 1
as expected.
1 1 −1
(1 + x + x2 ) (−1 + x2 )
4. b. PB←D = CB CB (1 − x) CB = 1 −1 0
1 0 1
1 1 1 1
2
PD←B = CD (1) CD (x) CD (x ) = 3 1 −2 1 because
−1 −1 2
1 2 2
1= 3 (1 + x + x ) + (1 − x) − (−1 + x )
1 2 2
x= 3 (1 + x + x ) − 2(1 − x) − (−1 + x )
x2 = 1 2 2
3 (1 + x + x ) + (1 − x) + 2(−1 + x )
The fact that PD←B = (PB←D )−1 is verified by multiplying these matrices. Next:
1 0 1
2 2
PE←D = CE (1 + x + x ) CE (1 − x) CE (−1 + x ) = 1 −1 0
0 1 −1
0 0 1
PE←B = CE (1) CE (x) CE (x2 ) = 0 1 0
1 0 0
where we note the order of the vectors in E = x2 , x, 1 . Finally, matrix multiplication verifies
5. b. Let B = {(1, 2, −1), (2, 3, 0), (1, 0, 2)} be the basis formed by the transposes of the columns
of A. Since D is the standard basis:
1 2 1
PD←B = CD (1, 2, −1) CD (2, 3, 0) CD (1, 0, 2) = 2 3 0 =A
−1 0 2
because
b. Since B0 = 1, x, x2 , we have
7.
1 1 0
(1 − x2 ) (2x + x2 )
P = PB0 ←B = CB0 CB0 (1 + x) CB0 = 0 1 2
−1 0 1
Finally
CB [T (1 − x2 )] CB [T (1 + x)] CB [T (2x + x2 )]
MB (T ) =
= CB (1 − x) CB (2 + x + x2 ) CB (2 + 3x + x2 )
−2 −3 −1
= 3 5 3
−2 −2 0
because
The verification that P−1 MB0 (T )P = MB (T ) is equivalent to checking that MB0 (T )P = PMB (T ),
and so can be seen by matrix multiplication.
h ih ih i h ih i h i
8. b. P−1 AP = −75 −23 29 −12
70 −29
3 2
7 5 = 5 −2
7 −3
3 2
7 5 = 1 0
0 −1 .
nh i h io
3
Let B = 7 , 25 consist of the columns of P. These are eigenvectors of A correspond-
ing to the eigenvalues 1, −1 respectively. Hence,
h h i h i i h h i h i i h i
MB (TA ) = CB TA 37 CB TA 25 = CB 37 CB −2 −5 = 1
0 −1
0
−3 −5
Hence, cT (x) = cMB (T ) (x) = x−2 x−3 = x2 − 6x − 1. Note that the calculation is easy
because B is the standard basis, but any basis could be used.
d. Use the basis B = 1, x, x2 of P2 and compute
= CB (1 + x − 2x2 ) CB (1 − 2x + x2 ) CB (−2 + x)
9.2. Operators and Similarity 145
1 1 −2
= 1 −2 1
−2 1 0
Hence,
x−1 −1 2 x−1 −1 2
cT (x) = cMB (T ) (x) = −1 x+2 −1 = −1 x+2 −1
2 −1 x −x + 3 0 x−2
= x3 + x2 − 8x − 3
nh i h i h i h io
1 0 0 1 0 0 0 0
f. Use B = 0 0 , 0 0 , 1 0 , 0 1 and compute
h
n h io n h io n h io n h io i
MB (T ) = CB T 10 00 CB T 00 10 CB T 01 0
0 CB T 0
0
0
1
h h i h i h i h i i
= CB 11 00 CB 00 11 CB −1 0
−1 0 CB 0
0 −1
−1
" 1 0 −1 0 #
0 1 0 −1
= 1 0 −1 0
0 1 0 −1
Hence,
x−1 0 1 0
0 x−1 0 1
cT (x) = cMB (T ) (x) = −1 0 x+1 0
0 −1 0 x+1
x−1 0 1 0 x−1 1
= (x − 1) 0 x+1 0 + −1 0 0 = x4
−1 0 x+1 0 −1 x+1
12. Assume that A and B are both n ×n and that null A = null B. Define TA : Rn → Rn by TA (x) = Ax for
all x in Rn ; similarly for TB . Then let T = TA and S = TB . Then ker S = null B and ker T = null A
so, by Exercise 28 Section 7.3 there is an isomorphism R : Rn → Rn such that T = RS. If B0 is the
standard basis of Rn , we have
where U = MB0 (R). This is what we wanted because U is invertible by Theorem 9.1.4.
Conversely, assume that A = U B with U invertible. If x is in null A then Ax = 0, so U Bx = 0,
whence Bx = 0 (because U is invertible), that is x is in null B. In other words null A ⊆ null B. But
B = U −1 A so null B ⊆ null A by the same argument. Hence null A = null B.
16. b. We verify first that S is linear. Showing S(w+v) = S(w)+S(v) means showing that MB (Tw+v ) =
MB (Tw ) + MB (Tv ). If B = {b1 , b2 } then column j of MB (Tw+v ) is
because CB is linear. This is column j of MB (Tw ) + MB (Tv ), which shows that S(w + v) =
S(w) + S(v). A similar argument shows that MB (Taw ) = aMB (Tw ), so S(aw) = aS(w), and
hence that S is linear.
To see that S is one-to-one, let S(w) = 0; by Theorem 7.2.2 we must show that w = 0. We
have MB (Tw ) = S(w) = 0 so, comparing jth columns, we see that CB [Tw (b j )] = CB [wb j ] = 0 for
146 Change of Basis
3. b. Given v in S(U ), we must show that T (v) is also in S(U ). We have v = S(u) for some u in U .
As ST = T S, we compute:
As T (u) is in U (because U is T -invariant), this shows that T (v) = S[T (u)] is in S(U ).
6. Suppose that a subspace U of V is T -invariant for every linear operator T : V → V ; we must show
that either U = 0 or U = V . Assume that U 6= 0; we must show that U = V . Choose u 6= 0 in U , and
(by Theorem 6.4.1) extend {u} to a basis {u, e2 , . . . , en } of V . Now let v be any vector in V . Then
(by Theorem 7.1.3) there is a linear transformation T : V → V such that T (u) = v and T (ei ) = 0
for each i. Then v = T (u) lies in U because U is T -invariant. As v was an arbitrary vector in V , it
follows that V = U .
[Remark: The only place we used the hypothesis that V is finite dimensional is in extending {u} to
a basis of V . In fact, this is true for any vector space, even of infinite dimension.]
B = 1 − 2x2 , x + x2 , x2
9.3. Invariant Subspaces and Direct Sums 147
where the last column is because T (x2 ) = 1 + x + 2x2 = (1 − 2x2 ) + (x + x2 ) + 3(x2 ). Finally,
x−3 1 −1
x−3 1
cT (x) = −3 x −1 = (x − 3) −3 x = (x − 3)(x2 − 3x + 3)
0 0 x−3
X E is in U because (X E)E = X E 2 = X E
X − X E is in W because (X − X E)E = X E − X E 2 = X E − X E = 0.
These are the subspaces of even and odd polynomials in P3 , respectively, and have bases
B1 = {1, x2 } and B2 = {x, x3 }. Hence, use the ordered basis B = {1, x2 , x, x3 } of P3 . Then
h i h i
MB (T ) = MB10(T ) MB 0(T ) = I02 −I0 2
2
23. b. Given v, T [v − T (v)] = T (v) − T 2 (v) = T (v) − T (v) = 0, so v − T (v) lies in ker T . Hence
v = (v − T (v)) + T (v) is in ker T + im T for all v, that is V = ker T + im T . If v lies in
ker T ∩ im T , write v = T (w), w in V . Then 0 = T (v) = T 2 (w) = T (w) = v, so ker T ∩ im T =
0.
25. b. We first verify that T 2 = T . Given (a, b, c) in R3 , we have
Then we claim that R3 = U1 ⊕U2 . To show R3 = U1 +U2 , observe that v = T (v) + [v − T (v)]
for each v in R3 , and T (v) is in U1 [because T [T (v)] = T 2 (v) = T (v)] while v − T (v) is in U2
[because T [v − T (v)] = T (v) − T 2 (v) = 0]. Finally we show that U1 ∩U2 = {0}. For if v is in
U1 ∩U2 then T (v) = v and T (v) = 0 so certainly v = 0.
Next, we show that U1 and U2 are T -invariant. If v is in U1 then T (v) is also in U1 because
T [T (v)] = T 2 (v) = T (v). Similarly U2 is T -invariant because, if v is in U2 , that is T (v) = 0,
then T [T (v)] = T 2 (v) = R(v) = 0; that is T (v) is also in U2 .
It is clear that T (a, b, c) = (a, b, c) if and only if b = 0; that is U1 = {(a, 0, c) | b, c in R},
so B1 = {(1, 0, 0), (0, 0, 1)} is a basis of U1 . Since T (v) = v for all v in U1 the restriction of
T to U1 is the identity transformation on U1 , and so has matrix I2 .
Similarly, T (a, b, c) = (0, 0, 0) holds if and only if a = −2b and c = −4b for some b, so
U1 = R(2, −1, 4) and B2 = {(2, −1, 4)} is a basis of U2 . Clearly the restriction of T to U2 is
the zero transformation, and so has matrix 02 — a 1 × 1 matrix.
Finally then, B = B1 ∪ B2 = {(1, 0, 0), (0, h0, 1), (2, −1, 4)} 3
i ish a basisiof R (since we have
shown that R3 = U1 ⊕U2 ), so T has matrix MB10(T ) MB 0(T ) = I02 001 .
2
29. b. We have T f2, z [v] = T f , z [T f , z (v)] = T f , z [ f (v)z] = f [ f (v)z]z = f (v) f (z)z. This expression
equals T f , z (v) = f (v)z for all v if and only if
f (v)(z − f (z)z) = 0
150 Change of Basis
for all v. Since f 6= 0, f (v) 6= 0 for some v, so this holds if and only if
z = f (z)z
2. If h , i denotes the inner product on V , then hu1 , u2 i is a real number for all u1 and u2 in U .
Moreover, the axioms P1 − P5 hold for the space U because they hold for V and U is a subset of V .
So h , i is an inner product for the vector space U .
π
3. b. k f k2 = −ππ cos2 x dx = −ππ 12 [1 + cos(2x)] dx = 21 x + 12 sin(2x) −π = π . Hence fˆ = √1π f is a
R R
unit vector.
h i h ih i
d. kvk2 = hv, vi = vT −11 −12 v = [ 3 −1 ] −11 −12 3
−1 = 17.
h i
Hence kvk 1
v = √117 −13 is a unit vector in this space.
√
4. b. d(u, v) = ku − vk = k(1, 2, −1, 2) − (2, 1, −1, 3)k = k(−1, 1, 0, −1)k = 3.
151
152 Inner Product Spaces
d. k f − gk2 = −ππ (1 − cos x)2 dx = −ππ 32 − 2 cos x + 21 cos(2x) dx because we have cos2 (x) =
R R
1 2 3 1
π
2 [1 + cos(2x)]. Hence k f − gk = 2 x − 2 sin(x) + 4 sin(2x) −π
= 32 [π − (−π )] = 3π . Hence
√
d( f , g) = 3π .
8. The space Dn uses pointwise addition and scalar multiplication:
( f + g)(k) = f (k) + g(k) and (r f )(k) = r f (k)
for all k = 1, 2, . . . , n.
P1. h f , gi = f (1)g(1) + f (2)g(2) + · · · + f (n)g(n) is real.
P2. h f , gi = f (1)g(1) + f (2)g(2) + · · ·+ f (n)g(n)) = g(1) f (1) + g(2) f (2) + · · ·+ g(n) f (n)
= hg, f i
P3. h f + h, gi = ( f + h)(1)g(1) + ( f + h)(2)g(2) + · · · + ( f + h)(n)g(n)
= [ f (1) + h(1)]g(1) + [ f (2) + h(2)]g(2) + · · ·+ [ f (n) + h(n)]g(n)
= [ f (1)g(1) + f (2)g(2) + · · ·+ f (n)g(n)] + [h(1)g(1) + h(2)g(2) + · · ·+ h(n)g(n)]
= h f , gi + hh, gi
P4. hr f , gi = (r f )(1)g(1) + (r f )(2)g(2) + · · ·+ (r f )(n)g(n)
= [r f (1)]g(1) + [r f (2)]g(2) + · · ·+ [r f (n)]g(n)
= r[ f (1)g(1) + f (2)g(2) + · · ·+ f (n)g(n)] = rh f , gi
P5. h f , f i = f (1)2 + f (2)2 + · · · + f (n)2 ≥ 0 for all f . If h f , f i = 0 then
f (1) = f (2) = · · · = f (n) = 0 (as the f (k) are real numbers) so f = 0
12. b. We need only verify P5. [P1 −h P4 ihold for any symmetric matrix A by (the discussion preced-
ing) Theorem 10.1.2.] If v = vv12 :
h 5 −3 i h v1 i
hv, vi = vT Av = v1 v2
−3 2 v2
Thus, hv, vi ≥ 0 for all v; and hv, vi = 0 if and only if 5v1 − 3v2 = 0 = v2 ; that is if and only
if v1 = v2 = 0 (i.e. v = 0). So P5 holds.
h i
d. As in (b), consider v = vv12 .
h 3 4 i h v1 i
hv, vi = v1 v2 4 6 v2
Thus, hv, vi ≥ 0 for all v; and (v, v) = 0 if and only if 3v1 + 4v2 = 0 = v2 ; that is if and only
if v = 0. Hence P5 holds. The other axioms hold because A is symmetric.
10.1. Inner Products and Norms 153
h i
a11 a12
13. b. If A = a21 , then ai j is the coefficient of vi w j in hv, wi. Here a11 = 1, a12 = −1 = a21 ,
a22
h i
1 −1
and a22 = 2. Thus, A = −1 2 . Note that a12 = a21 , so A is symmetric.
1 0 −2
d. As in (b): A = 0 2 0 .
−2 0 5
14. As in the hint, write hx, yi = xT Ay. Since A is symmetric, this satisfies axioms P1, P2, P3 and P4
for an inner product on Rn —(and only P2 requires that A be symmetric). Then it follows that
Hence hx, yi = 0 for all x and y in Rn . But if e j denotes column j of In , then hei , e j i = eTi Ae j is the
(i, j)-entry of A. It follows that A = 0.
16. b. hu − 2v − w, 3w − vi = 3hu, wi − 6hv, wi − 3hw, wi − hu, vi + 2hv, vi + hw, vi
= 3hu, wi − 5hv, wi − 3 kwk2 − hu, vi + 2 kvk2
= 3 · 0 − 5 · 3 − 3 · 3 − (−1) + 2 · 4
= −15
P2 P3 P2
20. (1) hu, v + wi = hv + w, ui = hv, ui + hw, ui = hu, vi + hu, wi
P2 P4 P2
(2) hv, rwi = hrw, vi = rhw, vi = rhv, wi
(1)
(3) By (1): hv, 0i = hv, 0+0i = hv, 0i + hv, 0i. Hence hv, 0i = 0. Now h0, vi = 0 by P2.
(4) If v = 0 then hv, vi = h0, 0i = 0 by (3). If hv, vi = 0 then it is impossible that v 6= 0 by P5, so
v = 0.
22. b. h3u − 4v, 5u + vi = 15hu, ui + 3hu, vi − 20hv, ui − 4hv, vi
= 15 kuk2 − 17hu, vi − 4 kvk2
d. ku + vk2 = hu + v, u + vi = hu, ui + u, vi + hv, ui + 4hv, vi
= kuk2 + 2hu, vi + kvk2
26. b. Here
W = w | w in R3 and v · w = 0
= {(x, y, z) | x − y + 2z = 0}
= {(s, s + 2t, t) | s, t in R}
= span B
where B = {(1, 1, 0), (0, 2, 1)}. Then B is the desired basis because B is independent
[In fact, if s(1, 1, 0) + t(0, 2, 1) = (s, s + 2t, t) = (0, 0, 0) then s = t = 0].
28. Write u = v − w; we show that u = 0. We are given that
= r1 hu, v1 ) + · · · + rn hu, vn i
= r1 · 0 + · · · + rn · 0
=0
Thus, kuk = 0, so u = 0.
29. b. If u = (cos θ , sin θ ) in R2 (with the dot product), then kuk = 1. If v = (x, y) the Schwarz
inequality (Theorem 10.1.4) gives
hu, vi2 ≤ kuk2 kvk2 ≤ 1 · kvk2 = kvk2
This is what we wanted.
2. b. Write b1 = (1, 1, 1), b2 = (1, −1, 1), b3 = (1, 1, 0). Note that in the Gram-Schmidt algorithm
we may multiply each ei by a nonzero constant and not change the subsequent ei . This avoids
fractions.
f1 = b1 = (1, 1, 1)
hb2 , e1 i
f2 = b2 − f1
ke1 k2
= (1, −1, 1) − 46 (1, 1, 1)
= 13 (1, −5, 1); use e3 = (1, −5, 1) with no loss of generality
hb3 , f1 i hb3 , f2 i
f3 = b3 − 2 f1 − f2
ke1 k ke2 k2
So the orthogonal basis is {(1, 1, 1), (1, −5, 1), (3, 0, −2)}.
Dh i h ′ ′ iE
a b
3. b. Note that c d , ac′ db′ = aa′ + bb′ + cc′ + dd ′ . For convenience write
h i h i h i h i
1 1 1 0 1 0 1 0
b1 = 0 1 , b2 = 1 1 , b3 = 0 1 , b4 = 0 0 . Then:
h i
1 1
f1 = b1 = 0 1
f2 = b2 − hb2 , f21 i f1
kf1 k
h i h i h i
−2
= 11 01 − 23 1
0
1
1 = 1
3
1
3 1
h i
1 −2
For the rest of the algorithm, use f2 = 3 1 , the result is the same.
hb , f i hb4 , f2 i hb , f i
f4 = b4 − 4 21 f1 − f2 − 4 23 f3
kf1 k kf2 k2 kf3 k
h i h i h i h i
−2 −2
= 10 00 − 31 10 1
1
1
− 15 1
3 1
1
− 10 1
−2 1
h i
1 1 0
= 2 0 −1
h i
1 0
Use f4 = 0 −1 for convenience. Hence, finally, the Gram-Schmidt algorithm gives the
nh i h i h i h io
1 1 1 −2 1 −2 1 0
orthogonal basis 0 1 , 3 1 , −2 1 , 0 −1 .
156 Inner Product Spaces
4. b. f1 = 1
hx, f1 i 2
f2 = x − 2 f1 = x − 2 · 1 = x − 1
kf1 k
2 2
f3 = x − hx , f21 i f1 − hx , f22 i f2
2 = x2 − 8/3 4/3 2 2
2 · 1 − 2/3 · (x − 1) = x − 2x + 3 .
kf1 k kf2 k
is in U ⊥ if and only if
6. b. x y z w
x+y = x y z w · 1 1 0 0 = 0
Thus y = −x and
U⊥ =
x −x z w | x, z, w in R
= span 1 −1 0 0 , 0 0 1 0 , 0 0 0 1
Thus a = 2s + t, b = −3s, c = −2t where s and t are in R, so p(x) = (2s + t) − 3sx − 2tx2 .
Hence, U ⊥ = span 2 − 3x, 1 − 2x2 and dim U ⊥ = 2, dim U = 1.
a b
f. is in U if and only if
c d
Dh i h iE
a b 1 1
0= c d , 0 0 = a+b
Dh i h iE
0= a b
c d , 11 00 = a+c
Dh i h iE
0= a b
c d , 11 01 = a+c+d
nh i o nh io
a −a 1 −1
The solution d = 0, b = c = −a, so U ⊥ = −a 0 a in R = span −1 0 . Thus
dim U ⊥ = 1 and dim U = 3.
h i h i h i
7. b. Write b1 = 10 01 , b2 = 11 −11 , and b3 = 10 10 . Then {b1 , b2 , b3 } is independent but
not orthogonal. The Gram-Schmidt algorithm gives
h i
f1 = b1 = 10 01
h i h i h i
f2 = b2 − hb2 , f21 i f1 = 11 −11 − 02 10 01 = 11 −11
kf1 k
hb3 , f1 i hb , f i
f3 = b3 − f − 3 22 f2
kf1 k2 1 kf2 k
h i h i h i
1 1 1 1 0 2 1 1
= 0 0 −2 0 1 − 4 1 −1
h i
= 12 −10 10
10.2. Orthogonal Sets of Vectors 157
h i h i
0 1 2 1
If E3′ = then E1 , E2 , E3′ is an orthogonal basis of U . If A =
−1 0 3 2 then
hA, E1 i hA, E2 i hA, E3′ i ′
projU A = 2 E1 + 2 E2 + 2E
kE1 k kE2 k kE3′ k 3
h i h i h i
4 1 0 4 1 1 −2 0 1
= 2 0 1 + 4 1 −1 + 2 −1 0
h i
3 0
= 2 1
We use U = span {1, 5 − 3x2 }. Then Theorem 10.2.8 asserts that the closest vector in U to x
is
hx, 1i hx, 5−3x2 i 2 3 −12 2 3 2
projU x = 2 1+ 2 (5 − 3x ) = 3 + 78 (5 − 3x ) = 13 (1 + 2x )
k1k k5−3x2 k
Here, for example hx, 5 − 3x2 i = 0(5) + 1(2) + 2(−7) = −12, and the other calculations are
similar.
R1
9. b. {1, 2x − 1} is an orthogonal basis of U because h1, 2x − 1i = 0 (2x − 1)dx = 0. Thus
hx2 +1, 1i hx2 +1, 2x−1i
projU (x2 + 1) = 2 1+ (2x − 1)
k1k k2x−1k2
3/4 1/6
= 1 1 + 1/3 (2x − 1)
= x + 56
Hence, x2 + 1 = (x + 56 ) + (x2 − x + 16 ) is the required decomposition. Check: x2 − x + 16 is in
U ⊥ because
Z 1
2 1
x2 − x + 16 dx = 0
hx − x + 6 , 1i =
0
Z 1
hx2 − x + 16 , 2x − 1i = x2 − x + 16 (2x − 1)dx = 0
0
11. b. We have hv + w, v − wi = hv, vi − hv, wi + hw, vi − hu, ui = kvk2 − kwk2 . But this means
that hv + w, v − ui = 0 if and only if kvk = kwk. This is what we wanted.
14. b. If v is in U ⊥ then hv, ui = 0 for all u in U . In particular, hv, ui i = 0 for 1 ≤ i ≤ n, so v is in
{u1 , . . . , um }⊥ . This shows that U ⊥ ⊆ {u1 , . . . , um }⊥ . Conversely, if v is in {u1 , . . . , um }⊥
then hv, ui i = 0 for each i. If u is in U , write u = r1 u1 + · · · + rm um , ri in R. Then
hv, ui = hv, r1 u1 + · · · + rm um i
= r1 hv, u1 i + · · · + rm hv, um i
= r1 · 0 + · · · + rm · 0
=0
As u was arbitrary in U , this shows that v is in U ⊥; that is {u1 , . . . , um }⊥ ⊆ U ⊥ .
158 Inner Product Spaces
18. b. Write e1 = (3, −2, 5) and e2 = (−1, 1, 1), write B = {e1 , e2 }, and write U = span B. Then
B is orthogonal and so is an orthogonal basis of U . Thus if v = (−5, 4, −3) then
v·e1
projU v = e1 + v·e22 e2
ke1 k2 ke2 k
−38 6
= 38 (3, −2, 5) + 3 (−1, 1, 1)
= (−5, 4, −3)
=v
20. b. CE (bi ) is column i of P. Since CE (bi ) ·CE (b j ) = hbi , b j i by (a), the result follows.
23. b. Let V be an inner product space, and let U be a subspace of V . If U = span {f1 , . . . , fm }, then
m m 2
projU v = ∑ hv,kfi kfi2i fi by Theorem 10.2.7 so k projU vk2 = ∑ hv,kfifki2i by Pythagoras’ theorem.
i=1 i=1
So it suffices to show that k projU vk2 ≤ kvk2 .
Given v in V , write v = u + w where u = projU v is in U and w is in U ⊥. Since u and w are
orthogonal, Pythagoras’ theorem (again) gives
h i
0 −1
T (E2 ) = 0 1 = −E2 + E4
h i
1 0
T (E3 ) = 2 0 = E1 + 2E3
h i
0 1
T (E4 ) = 0 2 = E2 + 2E4
Hence,
MB (T ) = CB [T (E1 )] CB [T (E2 )] CB [T (E3 )] CB [T (E4 )]
" −1 0 1 0 #
0 −1 0 1
= 1 0 2 0
0 1 0 2
4. b. If T is symmetric then hv, T (w)i = hT (v), wi holds for all v and w in V . Given r in R:
hv, (rT )(w)i = hv, rT (w)i = rhv, T (w)i = rhT (v), wi = hrT (v), wi = h(rT )(v), wi
5. b. If E = {e1 = (1, 0, 0), e2 = (0, 1, 0), e3 = (0, 0, 1)} is the standard basis of R3 :
ME (T ) = CE [T (e1 )] CE [T (e2 )] CE [T (e3 )]
= CE (7, −1, 0) CE (−1, 7, 0) CE (0, 0, 2)
7 −1 0
= −1 7 0
0 0 2
x−7 1 0
Thus, cT (x) = 1 x−7 0 = (x − 6)(x − 8)(x − 2) so the eigenvalues are λ1 = 6, λ2 = 8,
0 0 x−2
and
λ3 = 2, (real as
MB0 (T ) issymmetric).
Corresponding (orthogonal) eigenvectors are x1 =
1 1 0
1 , x2 = −1 , and x3 = 0 , so
0 0 1
1 1 0
√1 1 , √1 −1 , 0
2 0 2 0 1
h i
is an orthonormal basis of eigenvectors of ME (T ). These vectors are equal to CE √1 (1, 1, 0) ,
h i 2
1
CE 2 (1, −1, 0) , and CE [(0, 0, 1)] respectively, so
√
n o
√1 (1, 1, 0), √1 (1, −1, 0), (0, 0, 1)
2 2
d. If B0 = 1, x, x2 then
x+1 0 −1
Hence, cT (x) = 0 = x(x−3)(x+2) so the (real) eigenvalues are λ1 = 3, λ2 = 0,
x−3 0
−1 0 x+1
0 1 1
λ3 = −2. Corresponding (orthogonal) eigenvectors are x1 = 1 , x2 = 0 , x3 = 0 ,
0 1 −1
0 1 1
so 1 , √1 0 , √1 0 is an orthonormal basis of eigenvectors of MB0 (T ). These
0 2 1 2 −1
h i h i
have the form CB0 (x), CB0 √1 (1 + x2 ) , and CB0 √1 (1 − x2 ) , respectively, so
2 2
n o
x, √1 (1 + x2 ), √1 (1 − x2 )
2 2
is an orthonormal basis of eigenvectors of T .
h i
7. b. Write A = ac db and compute:
h h i h i h i h i i
MB (T ) = CB T 10 00 CB T 01 00 CB T 0
0
1
0 CB T 0
0
0
1
h h i h i h i h i i
a 0 b 0 0 a 0 b
= CB c 0 CB d 0 CB 0 c CB 0 d
" a b 0 0 #
c d 0 0
= 0 0 a b
0 0 c d
h i
A 0
= 0 A
Hence,
nh i h io
xI 0 A 0
cT (x) = det [xI − MB (T )] = det − 0 xI 0 A
h i
= det xI −0 A xI −0 A = det (xI − A) · det (xI − A) = [cA (x)]2
12. (2) We prove that (1) ⇒ (2). If B = {f1 , . . . , fn } is an orthonormal basis of V , then MB (T ) = [ai j ]
where ai j = hfi , T (f j )i by Theorem 10.3.2. If (1) holds then a ji = hf j , T (fi )i = −hT (f j ), fi i =
−hfi , T (f j )i = −ai j . Hence [MV (T )]T = −MV (T ), proving (2).
14. c. We have
MB (T ′ ) = CB [T ′ (f1 )] CB [T ′ (f2 )] · · · CB [T ′ (fn )]
Hence, column j of MB (T ′ ) is
hf j , T (f1 )i
hf j , T (f2 )i
CB (T ′ (f j )) = ..
.
hf j , T (fn )i
by the definition of T ′ . Hence the (i, j)-entry of MB (T ′ ) is hf j , T (fi )i. But this is the ( j, i)-
entry of MB (T ) by Theorem 10.3.2. Thus, MB (T ′ ) is the transpose of MB (T ).
10.4. Isometries 161
10.4 Isometries
h i h i h ih i h i
2. b. We have T ab = −a −b = −1 0
0 −1
a
b so T has matrix −1 0
0 −1 , which is orthogonal.
Hence T is an isometry, and det T = 1 so T is a rotation by Theorem 10.4.4.
h In fact,i T
cos θ − sin θ
is counterclockwise rotation through π . (Rotation through θ has matrix sin θ cos θ by
Theorem 2.6.4; see also the discussion following Theorem 10.4.3). This can also be seen
directly from the diagram.
y
a
b
a −a
T =
b −b
h i h i h ih ii h
a −b 0 −1 a 0 −1
d. We have T b = = −a so T has matrix
−1 0 b . This is orthogonal, −1 0
so T is an isometry. Moreover, det T = −1 so T is a reflection by Theorem 10.4.4. In fact,
T is reflection in the line y = −x by Theorem 2.6.5. This can also be seen directly from the
diagram.
y
a
b
a −b
T = y = −x
b −a
Hence, det T = 1 so T is a rotation. Indeed, (the discussion following) Theorem 10.4.3 shows
that T is a rotation through an angle θ where cos θ = √1 , sin θ = √1 ; that is θ = π4 .
2 2
√ √ √
a
1 √3c − a 1
−1
√ 0 3 a
1
−1
√ 0 3
3. b. T b = 2
3a + c = 2
3 0 1 b , so T has matrix 2
3 0 1 . Thus,
c 2b 0 2 0 c 0 2 0
√ √
3 3
x + 12 0 − 2 x+ 1
0 − √
√ √2 2
x+ 1
− 3
= (x − 1) x2 + 32 x + 1
cT (x) = − 23 x − 21 = − 23 0 x2 − 12 = √2
− 23
2
x − 12
2
0 −1 x 0 −1 x
Hence,
1we are in (1) of Table 10.1 so T is a rotation about the line Re with direction vector
√
e = √3 , where e is an eigenvector corresponding to the eigenvalue 1.
3
a a 1 0 0 a 1 0 0
d. T b = −b = 0 −1 0 b , so T has matrix 0 −1 0 . This is orthogonal,
c −c 0 0 −1 c 0 0 −1
2
isometry. Since cT (x) = (x − 1)(x + 1) , we are in case (4) of Table 10.1. Then
so Tis an
1
e = 0 is an eigenvector corresponding to 1, so T is a rotation of π about the line Re with
0
direction vector e, that is the x-axis.
a +c
a√ 1 0
√ 1 a 1 0
√ 1
1 1 √1
f. T b = 2 − 2b = 2
√ √ 0 − 2 0 b , so T has matrix 2
0 − 2 0 . Hence,
c c−a −1 0 1 c −1 0 1
x − √12 0 − √12 √
x − √1 − √1
cT (x) = 0 x+1 0 = (x + 1) √1
2 2
x − √12
= (x + 1)(x2 − 2x + 1)
√1 0 x − √1 2
2 2
0
Thus we are in case (2) of Table 10.1. Now e = 1 is an eigenvector corresponding to the
0
3π
eigenvalue −1, so T is rotation (of 4 ) about the line Re (the y-axis) followed by a reflection
in the plane (Re)⊥ — the xz-plane.
6. Let T be an arbitrary isometry, and let a be a real number. If aT is an isometry then Theorem 10.4.2
gives
kvk = k(aT )(v)k = ka(T (v))k = |a| kT (v)k = |a|kvk holds for all v.
Thus |a| = 1 so, since a is real, a = ±1. Conversely, if a = ±1 then |a| = 1 so we have k(aT )(v)k =
|a| kT (v)k = 1 kT (v)k = kvk for all v. Hence aT is an isometry by Theorem 10.4.2.
12. b. Assume that S = Su ◦ T where u is in V and T is an isometry of V . Since T is onto (by Theorem
10.4.2), let u = T (w) where w ∈ V . Then for any v ∈ V , we have
The integrations involved in the computation of the Fourier coefficients are omitted in 1(b), 1(d), and 2(b).
π
1. b. f5 = 2 − π4 cos x + cos323x + cos525x
π
+ sin x − sin22x + sin33x − sin44x + sin55x − π2 cos x + cos323x + cos525x
d. f5 = 4
2
2. b. π − π8 cos 2x
22 −1
+ cos 4x
42 −1
+ cos 6x
62 −1
4. We use the formula that cos(θ ± φ ) = cos θ cos φ ∓ sin θ sin φ , so that 2 cos θ cos φ = cos(θ −
φ ) cos(θ + φ ). Hence:
Z π Z π
1
cos(kx) cos(ℓx)dx = 2 {cos[(k − ℓ)x] + cos[(k + ℓ)x]} dx
0 h0
sin[(k−ℓ)x] π
i
1 sin[(k+ℓ)x]
= 2 k+ℓ + (k−ℓ)
0
= 0 if k 6= ℓ.
11. Canonical Forms
2 .
Hence, {p11 , p12 } is a basis of null (−I −A). We now expand this to a basis of null (−I − A)
However, (−I − A)2 = 0 so null (−I − A)2 = R3 . Hence,
this case, we expand {p11 , p12 }
in
0
to any basis {p11 , p12 , p13 } of R3 , say by taking p13 = 0 . Hence
1
1 0 0 −1 0 1
−1
P= p11 p12 p13 = 1 1 0 satisfies P AP = 0 −1 0
1 −3 1 0 0 −1
as may be verified.
x+3 1 0 x+3 1 0 x+3 1 0
d. cA (x) = −4 x+1 −3 = −4 x+1 −3 = −4 x−2 −3 = (x − 1)2 (x + 2).
−4 2 x−4 0 −x + 1 x−1 0 0 x−1
Hence λ1 = 1, λ3 = −2, and we are in case k = 2 of the triangulation algorithm.
4 1 0 4 1 0 −1
I − A = −4 2 −3 → 0 3 −3 ; p11 = 4
−4 2 −3 0 0 0 4
12 6 −3 4 2 −1 −1 0
12
(I − A) = −12 −6 3 → 0 0 0 ; p11 = 4 , p12 = 1
−12 −6 3 0 0 0 4 2
Thus, null (I − A)2 = span {p11 , p12 }. As dim Gλ1 (A) = 2 in this case (by Lemma 11.1.1),
we have Gλ1 (A) = span {p11 , p12 }. However, it is instructive to continue the process:
4 2 −1
2
(I − A) = 3 −4 −2 1
−4 −2 1
whence
4 2 −1
3
(I − A) = 9 −4 −2 1 = 3(I − A)2
−4 −2 1
165
166 Canonical Forms
This continues tohgive (I −i A)4 = 3h2 (I − A)2i, . . . , and in general (I − A)k = 3k−2 (I − A)2 for
k ≥ 2. Thus null (I − A)k = null (I − A)2 for all k ≥ 2, so
h i
Gλ1 (A) = null (I − A)2 = span {p11 , p12 }
Hence, null [−2I − A] = span {p21 }. We need go no further with this as {p11 , p12 , p21 } is a
basis of R3 . Hence
−1 0 −1 1 1 0
−1
P = p11 p12 p21 = 4 1 1 satisfies P AP = 0 1 0
4 2 1 0 0 −2
as may be verified.
f. To evaluate cA (x), we begin by adding column 4 to column 1:
x+3 −6 −3 −2 x+1 −6 −3 −2 x+1 −6 −3 −2
2 x−3 −2 −2 0 x−3 −2 −2 0 x−3 −2 −2
cA (x) = 1 −3 x −1 = 0 −3 x −1 = 0 −3 x −1
1 −1 −2 x x+1 −1 −2 x 0 5 1 x+2
x−3 −2 −2 x−3 −2 0 x−3 −2 0
= (x + 1) −3 x −1 = (x + 1) −3 x −x − 1 = (x + 1) 2 x+1 0
5 1 x+2 5 1 x+1 5 1 x+1
x−3 −2
= (x + 1)1 2 x+1 = (x + 1)2 (x − 1)2
Hence, λ1 = −1, λ2 = 1 and we are in case k = 2 of the triangulation algorithm. We omit the
details of the row reductions:
" 2 −6 −3 −2 # " 1 0 0 −1 # " 1 #
2 −4 −2 −2 0 1 0 0 0
−I − A = 1 −3 −1 −1 → 0 0 1 0 ; p11 = 0
1 −1 −2 −1 0 0 0 0 1
−13 23 13 13 1 0 −1 −1 1 1
" # " # " # " #
−8 12 8 8 0 1 0 0 0 0
(−I − A)2 = −6 10 6 6 → 0 0 0 0 ; p11 0 , p12 = 1
−3 5 3 3 0 0 0 0 1 0
We have dim Gλ1 (A) = 2 as λ1 = −1 has multiplicity 2 in cA (x), so Gλ1 (A) = span {p11 , p12 }.
Turning to λ2 = 1:
" 4 −6 −3 −2 # " 1 0 0 −5 # " 5 #
2 −2 −2 −2 0 1 0 −2 2
I −A = 1 −3 1 −1 → 0 0 1 −2 ; p21 = 2
1 −1 −2 1 0 0 0 0 1
−1 −1 1 5 1 1 0 0 5 1
" # " # " # " #
0 0 0 0 0 0 1 0 2 −1
(I − A)2 = −2 −2 2 −6 → 0 0 0 1 ; p21 = 2 , p22 = 0
1 1 −5 3 0 0 0 0 1 0
Hence, Gλ2 (A) = span {p21 , p22 } using Lemma 11.1.1. Finally, then
−1
" 1 1 5 1 # " 1 0 0
#
0 0 2 −1 0 −1 0 0
P = p11 p12 p21 p22 = 0 1 2 0 gives P−1 AP =
0 0 1 −2
1 0 1 0 0 0 0 1
as may be verified.
11.2. Jordan Canonical Form 167
4. Let B be any basis of V and write A = MB (T ). Then cT (x) = cA (x) and this is a polynomial: cT (x) =
a0 + a1 x + · · · + an xn for some ai in R. Now recall that MB : L(V , V ) → Mnn is an isomorphism
of vector spaces (Exercise 9.1.26) with the additional property that MB (T k ) = MB (T )k for k ≥ 1
(Theorem 9.2.1). With this we get
MB [cT (T )] = MB [a0 1V + a1 T + · · · + an T n ]
= a0 MB (1V ) + a1 MB (T ) + · · · + an MB (T )n
= a 0 I + a 1 A + · · · + a n An
= cA (A)
=0
(a + bi)(2 − i) = (a − bi + 1)(1 + i)
(2a + b) + (2b − a)i = (a + 1 + b) + (a + 1 − b)i
169
170 Complex Numbers
√
d. If x = re−iθ then x4 = 64 becomes r4 e4iθ = 64ei·0 . Hence r4 = 64 (whence r = 2 2) and
4θ = 0 + 2kπ ; θ = k π2 , k = 0, 1, 2, 3. The roots are
√ √
2 2e0i = 2 2 (k = 0)
√ √
2 2eπ i/2 = 2 2i (k = 1)
√ √
2 2eπ i = −2 2 (k = 2)
√ √
2 2e3π i/2 = −2 2i (k = 3)
6. b. The quadratic is (x−u)(x−u) = x2 −(u +u)x+u u = x2 −4x+13. The other root is u = 2 +3i.
d. The quadratic is (x−u)(x−u) = x2 −(u +u)x+u u = x2 −6x+25. The other root is u = 3 +4i.
10. b. Taking x = u = −2: x2 +ix −(4 −2i) = 4 −2i −4 +2i = 0. If v is the other root then u +v = −i
(i is the coefficient of x) so v = −u − i = 2 − i.
d. Taking x = u = −2 + i: (−2 + i)2 = 3(1 − i)(−2 + i) − 5i
= (3 − ri) + 3(−1 + 3i) − 5i
= 0.
If v is the other root then u + v = −3(1 − i), so v = −3(1 − i) − u = −1 + 2i.
h i √ √
b. x2 − x + (1 − i) = 0 gives x = 21 1 ± 1 − 4(1 − i) = 21 1 ± −3 + 4i . Write w = −3 + 4i
p
11.
so w2 = −3 + 4i. If w = a + bi then w2 = (a2 − b2 ) + (2ab)i, so a2 − b2 = −3, 2ab = 4. Thus
b = 2a , a2 − a42 = −3, a4 + 3a2 − 4 = 0, (a2 + 4)(a2 − 1) = 0, a = ±1, b = ±2, w = ±(1 + 2i).
Finally the roots are 12 [1 ± w] = 1 + i, −i.
h i √
d. x − 3(1 − i)x − 5i = 0 gives x = 2 3(1 − i) ± 9(1 − i) + 20i = 12 3(1 − i) ± 2i . If w =
1
p
2
2
√
2i then w2 = 2i. Write w = a + bi so (a2 − b2 ) + 2abi = 2i. Hence a2 = b2 and ab = 1; the
solution is a = b = ±1 so w = ±(1 + i). Thus the roots are x = 12 (3(1 − i) ± w) = 2 − i, 1 − 2i.
12. b. |z − 1| = 2 means that the distance from z to 1 is 2. Thus the graph is the circle, radius 2, center
at 1.
d. If z = x + yi, then z = −z becomes x + yi = −x + yi. This holds if and only if x = 0; that is if
and only if z = yi. Hence the graph is the imaginary axis.
f. If z = x + yi, then im z = m · re z becomes y = mx. This is the line through the origin with slope
m.
√ √
f. |−6 + 6i| = 6 1 + 1 = 6 2 and cos ϕ = 6√6 2 = √12 . Thus ϕ = π4 so θ = 34π ; whence −6 + 6i =
√
6 2e−3π i/4 .
√
19. b. e7π i/3 = e(π /3+2π )i = eπ i/3 = cos π3 + i sin π3 = 12 + 23 i
√ −π i/4 √ −π
−π
√ 1 1
d. 2e = 2 cos 4 + i sin 4 = 2 √ − i = 1−i
√
2 2
√ −2π i/6 √ √ √ √
= 2 3 cos −3π + i sin −3π = 2 3 12 − 23 i = 3 − 3i
f. 2 3e
√
20. b. (1 + 3i)−4 = (2eπ i/3 )−4 = 2−4 e−4π i/3
1
= 16 [cos(−4
π /3) + i sin(−4π /3)]
√
= 16 − 2 + 23 i
1 1
√
= − 321
+ 323 i
h√ i10 √ √
d. (1 − i)10 = 2e−π i/4 = ( 2)10 e−5π i/2 = ( 2)10 e(−π /2−2π )i
√
= ( 2)10 e−π i/2 = 25 cos −2π + i sin −2π
= 32(0 − i) = −32i
√
h √ i5
9 5
f. ( 3 − i) (2 − 2i) = 2e − π i/6 9
] 2 2e −π i/4
√
= 29 e−3π i/2 (2 2)5 e−5π i/4
√ √
= 29 (i)25 ( 2)4 2 − √12 + √12 i
= 216 i(−1 + i)
= −216 (1 + i)
√ √
23. b. Write z = reiθ . Then z4 = 2( 3i − 1) becomes r4 e4iθ = 4e2π i/3 . Hence r4 = 4, so r = 2, and
4θ = 23π + 2π k; that is
θ = π6 + π2 k k = 0, 1, 2, 3
The roots are
√ √ √3 1 √2 √
2eπ i/6 =
2 2 + 2i = 2 3+i
√ 4π i/6 √ √
3
√
2
√
1
2e = 2 − 2 + 2 i = 2 −1 + 3i
√ 7π i/6 √ √3 1 √ √
2e = 2 − 2 − 2 i = − 22 3+i
√ 10π i/6 √ 1 √3 √ √
2e = 2 2 − 2 i = − 22 −1 + 3i
172 Complex Numbers
k=2
k=1
π /6
k=3
k=4
d. Write z = reiθ . Then z6 = −64 becomes r6 e6iθ = 64eπ i . Hence r6 = 64, so r = 2, and 6θ =
π + 2π k; that is θ = π6 + π3 k where k = 0, 1, 2, 3, 4, 5. The roots are thus z = 2eπ /6+π /3k for
these values of k. In cartesian form they are
k √0 1 √2 √3 4 √5
z 3 + i 2i − 3 + i − 3 − i −2i 3−i
26. b. Each point on the unit circle has polar form eiθ for some angle θ . As the n points are equally
spaced, the angle between consecutive points is 2nπ . Suppose the first point into the first quad-
rant is z0 = eα i . Write w = e2π i/n . If the points are labeled z1 , z2 , z3 , . . . , zn around the unit
circle, they have polar form
z1 = eα i
z2 = e(α +2π /n)i = eα i e2π i/n = z1 w
z3 = e[α +2(2π /n)]i = eα i e4π i/n = z1 w2
z4 = e[α +3(2π /n)]i = eα i e6π i/n = z1 w3
..
.
zn = e[α +(n−1)(2π /n)]i = eai e2(n−1)π i/n = z1 wn−1
z4 z3
z2
2π
n 2π
n
2π
n z1
α
173
z1 + z2 + · · · + zn = z1 (1 + w + · · · + wn−1 ) (∗)
n
Now wn = e2π i/n = e2π i = 1 so
0 = 1 − wn = (1 − w)(1 + w + w2 + · · · + wn−1 )
z1 + z2 + · · · + zn = z1 · 0 = 0
B. Proofs
1. b. (1). We are to prove that if the statement “m is even and n is odd ” is true then the statement
“m + n is odd ” is also true.
If m is even and n is odd, they have the form m = 2p and n = 2q + 1, where p and q are
integers. But then m + n = 2(p + q) + 1 is odd, as required.
(2). The converse is false. It states that if m + n is odd then m is even and n is odd; and a
counterexample is m = 1, n = 2.
d. (1). We are to prove that if the statement “x2 − 5x + 6 = 0” is true then the statement “x = 2 or
x = 3” is also true.
Observe first that x2 −5x+6 = (x−2)(x−3). So if x is a number satisfying x2 −5x+6 = 0
then (x − 2)(x − 3) − 0 so either x = 2 or x = 3. [Note that we are using an important fact
about real numbers: If the product of two real numbers is zero then one of them is zero.]
(2). The converse is true. It states that if x = 2 or x = 3 then x satisfies the equation x2 −5x+6 =
0. This is indeed the case as both x = 2 or x = 3 satisfy this equation.
2. b. The implication here is p ⇒ q where p is the statement “n is any odd integer ”, and q is the
statement “n2 = 8k + 1 for some integer k”. We are asked to either prove this implication or
give a counterexample.
This implication is true. If p is true then n is odd, say n = 2t + 1 for some integer t. Then
n2 = (2t)2 + 2(2t) + 1 = 4t(t + 1) + 1. But t(t + 1) is even (because t is either even or odd),
say t(t + 1) = 2k where k is an integer. Hence n2 = 4t(t + 1) + 1 = 4(2k) + 1, as required.
3. b. The implication here is p ⇒ q where p is the statement “n + m = 25, where n and m are
integers ”, and q is the statement “one of m and n is greater than 12” is also true. We are asked
to either prove this implication by the method of contradiction, or give a counterexample.
The implication is true. To prove it by contradiction, we assume that the conclusion q is false,
and look for a contradiction. In this case assuming that q is false means both n ≤ 12 and
m ≤ 12. But then n + m ≤ 24, contradicting the hypothesis that n + m = 25. So the statement
is true by the method of proof by contradiction.
The converse is false. It states that q ⇒ p, that is if one of m and n is greater than 12 then
n + m = 25. But n = 13 and m = 13 is a counterexample.
d. The implication here is p ⇒ q where p is the statement “mn is even, where n and m are
integers ”, and q is the statement “m is even or n is even ”. We are asked to either prove this
implication by the method of contradiction, or give a counterexample.
This implication is true. To prove it by contradiction, we assume that the conclusion q is false,
and look for a contradiction. In this case assuming that q is false means that m and n are both
odd. But then mn is odd (if either were even the product would be even). This contradicts the
hypothesis, so the statement is true by the method of proof by contradiction.
The converse is true. It states that if m or n is even then mn is even, and this is true (if m or n
is a multiple of 2, then mn is a multiple of 2).
175
176 Proofs
5. b. At first glance the statement does not appear to be an implication. But another way to say it is
that if the statement “n ≥ 2” is true then the statement “n3 ≥ 2n ” is also true.
This is not true. In fact, n = 10 is a counterexample because 103 = 1000 while 210 = 1024. It
is worth noting that the statement n3 ≥ 2n does hold for 2 ≤ n < 9.
C. Mathematical Induction
1
The second last term on the left side is n(n+1)
so we can use Sn :
h i
1 1 1 1 1 1 1
1·2 + 2·3 + · · · + (n+1)(n+2) = 1·2 + 2·3 + · · · + n(n+1) + (n+1)(n+2)
n 1
= n+1 + (n+1)(n+2)
n(n+2)+1
= (n+1)(n+2)
(n+1)2
= (n+1)(b+2)
n+1
= n+2
177
178 Mathematical Induction
which is clearly a multiple of 3. Hence Sn+1 is true, and so Sn is true for every n by induction.
20. Look at the first few values: B1 = 1, B2 = 5, B3 = 23, B4 = 119, . . . . If these are compared to the
factorials: 1! = 1, 2! = 4, 3! = 6, 4! = 24, 5! = 120, . . . , it is clear that Bn = (n + 1)! − 1 holds for
n = 1, 2, 3, 4 and 5. So it seems a reasonable conjecture that
22. b. If we know that Sn ⇒ Sn+8 then it is enough to verify that S1 , S2 , S3 , S4 , S5 , S6 , S7 , and S8 are
all true. Then
S1 ⇒ S9 ⇒ S17 ⇒ S25 ⇒ · · ·
S2 ⇒ S10 ⇒ S18 ⇒ S26 ⇒ · · ·
S3 ⇒ S11 ⇒ S19 ⇒ S27 ⇒ · · ·
S4 ⇒ S12 ⇒ S20 ⇒ S28 ⇒ · · ·
S5 ⇒ S13 ⇒ S21 ⇒ S29 ⇒ · · ·
S6 ⇒ S14 ⇒ S22 ⇒ S30 ⇒ · · ·
S7 ⇒ S15 ⇒ S23 ⇒ S31 ⇒ · · ·
S8 ⇒ S16 ⇒ S24 ⇒ S32 ⇒ · · ·
Clearly each Sn will appear in this array, and so will be true.
179
ADAPTED FORMATIVE ONLINE COURSE COURSE LOGISTICS
OPEN TEXT ASSESSMENT SUPPLEMENTS & SUPPORT
a d v a n c i n g l e a r n i n g
LYRYX.COM