Calculus of Vector Functions
Calculus of Vector Functions
RICHARD E. WILLIAMSON
RICHARD H. CROWELL
HALE F. TROTTER
Digitized by the Internet Archive
in 2010
[Link]
Calculus
of
Vector Functions
Third Edition
Calculus
of
Vector Functions
RICHARD E. WILLIAMSON
Department of Mathematics
Dartmouth College
RICHARD H. CROWELL
Department of Mathematics
Dartmouth College
HALE F. TROTTER
Department of Mathematics
Princeton University
10 9 8 7
R. E. W.
R. H. C.
H. F. T.
Possible Courses
of Study
We have tried to organize the text so that each section leads naturally
to the next, but it is by no means necessary to include all sections or to
take them up in strictly consecutive order. In particular, it is not necessary
to complete the linear algebra before starting on the calculus. (Students
who have already taken linear algebra can, of course, start with Chapter 3
and use the first two chapters for reference and review.) Everything
essential for Chapters 3 through 7 is contained in the first seven sections
of Chapter (An exception is the use of facts about dimension in the
1.
until the section on change of variable in multiple integrals and the last
three sections on vector analysis, where they are really needed. On the
other hand, determinants can be used in practice for inverting small
(2-by-2 or 3-by-3) matrices or solving small systems of linear equations,
so that some may prefer to take them up earlier and postpone (or omit)
the row-reduction method given in the first section of Chapter 2. At the
cost of avoiding a few higher dimensional exercises later on, one may
even restrict discussion of matrix inversion to the trivial 2-by-2 case.
Other changes of order, such as taking up multiple integration early
in the course, should also cause no difficulty.
The sequences of section numbers listed below are minimal in the sense
that they contain the prerequisite material for later entries, but do not
contain everything that might be desirable in a typical course. Additional
sections can be added to make up a full one-year course suitable for par-
ticular needs. Experience has shown that for ordinary class use two
meetings should be used to cover an average section. An occasional short
Possible Courses of Study
section merits only one meeting, and some longer ones, in order to be
covered entirely, require three.
Ch. 1
Contents
Introduction, 1
1 Vectors, 4
2 Geometric interpretations, 7
3 Matrices, 15
4 Linear functions, 25
5 Dot products, 37
6 Euclidean geometry, 46
7 Determinants, 51
8 Determinant expansions, 67
2 Some applications, 88
9 Eigenvectors, 167
10 Coordinates, 179
8 Differentials, 250
Appendix, 579
1 Introduction, 579
Index, 611
Calculus
of
Vector Functions
Introduction
y = x2 + 3
yields a real number y for any real number x and so defines a function/
from % to %
with (for instance) /(0) 3,/(-2) 7,/(V3) =
6, etc. = =
In this book we are concerned with functions of several variables whose
values may be real numbers or, more generally, may be w-tuples of real
numbers. For example, a pair of formulas
Jl = V^i + x\ -f X3
y2 = xrx 2 + 5x 3
g(l,2,3) = (Vl4,17),
g(3, 2, 1) = (Vl4, 11).
We shall use 51" to stand for the set of all rt-tuples of real numbers.
1
(Jl is thus the same as '.R .) The domain of a function is on which it
the set
is defined, and the range or image is the set of values assumed by the
Introduction
/
31" ft"
ft 2 ft 1 defined by a formula
f(x, y) = z.
context in which they will not be confused with the usual notations for
matrices and numbers. The printing of a word in boldface type indicates
that the word is being defined at that point in the text.
1
SECTION 1
(x t +yu x +y t % ,. . . ,x n +_y„).
(rxj, rx 2 , . . . , rx n ).
x + y = (2, 6, -2, 6)
and
3x = (6, -3,0,9).
Sec. 1 Vectors
1 rx + sx = (r + s)x.
2. rx + ry = r(x + y).
3. r(sx) = (rs)x.
4. x + y = y + x.
5. (x + y) + z = x + (y + z).
6. x + = x.
7. x + (-x) = 0.
in Chapter 2, but for the present "vector" may be taken to mean "element
of :K"" for some n. Numbers arc sometimes called scalars when emphasis
on the distinction between numbers and vectors is wanted. In physics, for
example, mass and energy may be referred to as scalar quantities in
distinction to vector quantities such as velocity or momentum. The term
scalar multiple is synonymous with what we have called numerical
multiple.
The vectors
(1,0,. ..,0)
(0, 1 , 0, . . . , 0)
= (0,
en . . . , 0, 1)
x = xfr + . . . + x„e„.
Because every element of ,'R" can be so simply represented in this way, the
set of vectors {e l5 . . . , e„} is called the natural basis for 3l n . For example,
in 'Ji
3
we have (1 , 2, — 7) = + 2e — 7e The
ei 2 3. entries in
x=(x ,..., x n 1 )
EXERCISES
1. Given x = (3, — 1, 0), y (0, 1, 5), and z = (2, 5, — 1) compute 3x, y -I- z,
3. Show that no choice of numbers a and b can make ax by (3, 0, 0), -'
where x and y are as in Problem 1. For what value(s) of c (if any) can the
equation ax by = (3, 0, c) be satisfied?
I
Sec. 2 Geometric Interpretations
4. Write out proofs for (a) law 3 and (b) law 4 on page 5, giving precise justifica-
tion for each step.
5. Verify that the set C[a, b] of all continuous real-valued functions defined on
the interval a <x<b is a vector space, with addition and numerical
multiplication defined by (/ \- g)(x) = f(x) + g(x) and (rf)(x) = rf{x).
6. Prove that the representation of a vector x in Jl" in terms of the natural basis
is unique. That is, show that if
-*i e i + • • • + x ne n y^ + . . . + yn e n ,
then xk = yk for k = 1 n.
SECTION 2
Geometric representations of Jl 1
as a line, of 5l 2
as a plane, and of .II
3 GEOMETRIC
as 3-dimensional space may be obtained by using coordinates. To represent INTERPRETATIONS
1
'Si one must first specify a point on the line to be called the origin,
as a line,
a unit of distance, and a direction on the line to be called positive. (The
opposite direction is then called negative.) Then a positive number x
corresponds to the point which is a distance x in the positive direction
from the origin. A negative number x corresponds to the point which is a
distance |.y| from the origin in the negative direction. The number zero of
course corresponds to the origin. The number line is most often thought
of as horizontal with the positive direction to the right. With this standard
convention, we obtain the familiar Fig. 1, in which the arrow indicates
the positive direction.
Figure 1
first axis is
p x and whose projection on the second axis is /?.,. The (per-
pendicular) projection of a point/? on a line L is defined as the foot of the
Vectors and Linearity Chap. 1
The conventional choice is to take the first axis horizontal with the
positive direction to the right, and the second axis vertical with the
positive direction upwards. This leads to the usual picture shown in Fig. 2.
gives points p x ,p 2 and/? 3 on the three axes. Then the point/?, correspond-
,
ing to (jtx x 2 x 3 ), is the one whose projections on the three axes are p x p 2
, , , ,
and p 3 .
/>(-3,2)
Figure 2 Figure 3
Figure 4
This rule can be applied to any pair of vectors, whereas the parallelogram
law does not (strictly speaking) apply if the arrows representing u and v
lie in the same straight line.
If we write w for u + v, then v =w— u. Figure 6 (which is simply
Fig. 5 relabeled) illustrates the useful fact that the difference of two vectors
w and u is represented by the arrow from the end of u to the end of w,
appropriately translated.
v(translated) w - u (translated)
Figure 5 Figure 6
ing an has the same direction as the arrow representing u and is a times as
long. For a negative number b, b\\ points in exactly the opposite direction
to u and is \b\ times as long. (See Exercise 6(b).)
The question of whether to represent a vector geometrically as a
10 I 'velars and Linearity Chap. 1
Sec. 2 Geometric Interpretations 11
multiples r(l, 1, 1) to get a line through the origin. Then the set of all
Figure 9
plane, we require that x x and x 2 not lie on the same line. Another way to
state this condition is that no multiple of x x should equal a multiple of
x 2 except for the zero multiples. More generally, we say that a set of
vectors
Xj, x2 , . . . ,x n
c1 x 1 + . . . + cnxn =
12 Vectors and Linearity Chap. 1
for some numbers c lt c n , then all the c's must be zero. If on the
. . . ,
Ci—c +c = 2 3
2ci + c — 3c = 0. 2 3
If we now let cz have any nonzero value, we can then solve for c x and c 2 .
Ci — c2 =— 1
2c i + c2 = 3,
and we solve the two equations. Adding them gives 3c x = 2, or c x = §,
while subtracting the second from two times the first gives — 3c = 2 —5,
or Co = f Thus .
X = C&i + . . . + c„x n ,
3
Example 4. The natural basis for .'K is the set of vectors
then (<?!, c2 , c3 ) = (0, 0, 0). On the other hand, any vector (x, y, z)
3
in !il can be written
(x, y, z) = xe x + ye + 2 ze 3 .
EXERCISES
1. For each pair u, v of vectors in Jl 2 given below, draw the arrows representing
u, v, u -f v, u — v, and u + 2v.
u2 + .vv
2
. [Ans. w = (— f, f).]
(b) Let u lt Vj, and u 2 be as in part (a), but take v 2 = (2, —2). (Note that v 2
is then a numerical multiple of v^) Sketch the lines u l | t\ 1 and
u2 + s\.,. Show algebraically that the lines do not intersect.
5. Show that in :R", the lines represented by sux t u and /v t + v are the same
if and only if both v t and v ()
— u are numerical multiples of u l .
u2 + v
2
14 I ectors and Linearity Chap. 1
OVB and UPA are congruent. From this deduce that OK and UP are
parallel and of equal length. Then OVPU is a parallelogram, and the
parallelogram law follows,
(b) In Figs. 7(a) and 7(b) the points U, V, and W are constructed to have
coordinates as shown. Prove that the triangles OUP, OVQ, and OWR
are similar. Show that the angles HOP, VOQ, and WOR are therefore
equal and that the lengths OK and OW^arc proportional to the length
of Oil as stated in the text.
8. (a) Show if u and v are two distinct vectors, then the vectors ru +
that
(1 —form a line through the points corresponding to u and v.
r)v
n
(b) If x x and x 2 are two vectors in 3i then the set of all vectors rx x +
,
(1 — /)x 2 where ,< t < 1, is the line segment joining \ x and x 2 A set .
5" in R" is convex if, whenever S contains two points, it also contains the
line segment joining them. Prove that the intersection of any collection
of convex sets is convex.
9. Represent the following lines in the form rx x [ x , where / runs over all real
3
(a) The line in J\ parallel to (1, 2, 0) and passing through the point (1,1, 1).
2
(b) The line in 3i joining the points (1, 0) and (0, 1).
3
(c) The line in :R joining the points (1, 0, 0) and (0, 0, 1).
3
10. Represent the following planes in 3l in the form ux 1 + vx 2 I
x„, where
// and v run over all real numbers. Sketch each plane.
(a) The plane parallel to the vectors (1, 1,0) and (0, 1, 1) and passing
through the origin.
(b) The plane parallel to the vectors e t and ea and passing through the point
(0, 0, 1).
(c) The plane passing through the three points (1, 0, 0), (0, 1, 0), and
(0, 0, 1).
11. Determine whether each of the following sets of vectors is linearly dependent
or linearly independent.
12. Determine whether the first vector below is in the set spanned by the set S.
In each case give a geometric interpretation: the first point does (or does
not) lie on the plane (or line) spanned by S.
14. (a) Prove that two nonzero vectors x and y are linearly dependent if and
only if they lie on the same line through the origin,
(b) Prove that three nonzero vectors x, y, and z are linearly dependent if
and only if they all lie on the same plane through the origin.
J>!
= 2x x + 3x 2 — 4x 3
y2 = xl — x2 + 2x 3 ,
fiveexamples above have shapes 3-by-2, 2-by-3, 2-by-2, l-by-3, and 4-by-l.
Note that the number of rows always comes before the number of columns.
The 1-by-/? matrices are called n-dimensional row vectors, and /?-by-l
j a 21 a 22 a 23 a 2i ) or A = {au ) t
x 2 i, x nl , .
matrix with the same shape and ijth entry equal to a u + b u For example
.
2 3\ /l -1\ /3 2^
4/ \0 0/ \0
example
Sec. 3 Matrices 17
(The proofs are just the same as when X, Y, and Z are all in .'ft".) In other
words, according to the definition in Section 2, for any fixed m and n, the
set of in -by- n matrices forms a vector space with the operations of addition
z2 = 5jj + 2y 2 ,
and
yt = 2x t + 3x 2 + 4x 3
yi = X\ -^2 + ^X 3 ,
3 -l\ (2 3 4
and B =
5 2/ \l -1 2
za = 3(2*! + 3x 2 + 4*3) — Ui — x + 2x 2 3)
zx = (3 • 2 - 1 l)x, + (3 •
3 - 1 •
(-1)> 2 + (3 -4 - 1 • 2)x 3
z2 = (5 • 2 + 2 •
1)*! + (5 • 3 +2 • (- l))x 2 + (5 4 + 2 •
2)x3
'3-2-1-1 3 •
3 — 1 -
(— 1) 3 -
4 — 1
-
2\ (5 10 10
C=
X5-2 + 2-1 5-3 + 2-(-l) 5-4 + 2-2/ \12 13 24
A = (a u a 2 , . . . , a k) and B =
are row and column vectors of the same dimension. Then the product AB
is defined to be thenumber a-Jjx + a 2 b 2 + + a k b k Now let A be an . . . .
m-by-k matrix and B be an k-by-n matrix. (It is important that the number
of columns of A be equal to the number of rows of B.) Then each row of
A is a A'-dimensional row vector and each column of B is a ^-dimensional
column vector. We define the matrix product AB as the m-by-n matrix
whose ijth entry is the product (in the sense just defined) of the z'th row of
A and they'th column of B. The product AB always has the same number
of rows as A and the same number of columns as B. For instance, in our
example
'3 -1\ /2 3 4\ / 5 10 10^
12 13 24
the entry in the second row and third column of the result is obtained by the
calculation
You should check that the other entries in the product can be obtained by
the rule stated above. Schematically the entries in a matrix product are
found by the mechanism illustrated below. The process is sometimes
called row-by-column multiplication of matrices.
* * * *
* * *
* * * 3 ***[*]*
* * *
* * * *
3.1 Theorem
1. (A + B)C = AC + BC.
2. A(B + C) = AB + AC.
3. {tA)B = t{AB) = A(tB).
4. A(BC) = (AB)C.
According to the last law, it makes sense to talk of the product of three
matrices and simply write ABC, since the result is independent of how the
factors are grouped. In fact this 3-term associative law implies that the
result of multiplying together any finite sequence of matrices is inde-
pendent of how they are grouped. Not all the laws that hold for multi-
plication of numbers hold for multiplication of matrices. In particular,
the value of a matrix product depends on the order of the factors, and
AB is usually different from BA. It is also possible for the product of two
matrices to be a zero matrix, without either of the factors being zero.
Exercise 5 at the end of this section illustrates these points.
The laws stated above are easily proved by writing out what they mean,
using the definitions of the operations, and then applying the associative,
distributive, and commutative laws of arithmetic. Number 4 is the most
complicated to prove, and we give its proof in full below. The other
proofs are left as exercises.
To prove thatA(BC) = (AB)C, let A, B, and C have respective shapes
p-by-q, q-by-r, and r-by-s. Let U = BC and V = AB. (Then U has
shape q-by-s, and V has shape p-by-r.) We have to show that AU = VC.
The //'th element of AU is (by definition of matrix multiplication) equal
Q T
AU
k=l
ik ll kj-
HIT Ay'th
\
is
1=1
. (/th
r Tin \
sum
l<l<r
l<k<Q
. . . 1
that has l's on its main diagonal and zeros elsewhere is called an identity
matrix. It has the property that
IA = A, BI=B
for any matrices A, B such that the products are defined. Thus it is an
identity element for matrix multiplication just as the number 1 is an identity
for multiplication of numbers. There an n x n identity matrix for every
is
and according
T, 1 - U DC
to the definition this shows that
K
1 2\ /l 2\~ l
=
[1-2
is invertible and that I
3 l) \3 7/ \-3 1
Sec. 3 Matrices 21
Many matrices, on the other hand, are not invertible. No zero matrix can
have an inverse, and several less obvious examples are given in the
exercises.
It is usually not easy to tell whether a large matrix is invertible, and it
can take a lot of work to compute its inverse if it has one. Determinants
can be used to give a formula for the inverse of a matrix (Theorem 8.3),
and a more effective way to compute inverses is given in Section 1 of
Chapter 2. Two-by-two matrices and a few other easy cases are discussed
in Exercises 9 to 13 of this section. The rule for finding the inverse of a
2-by-2 matrix is as follows.
bY 1
-b\
3 2
-
a
=— H d i /
if ad-bc^O.
ad - bc\-
,
\c d! c a
-1
1 3\ (2 -3
,-1 2/ Hi 1
1 3\ /* —\\ /* -*\ / 1 3
-1 2/U \J U \JY
so that we have indeed found the inverse.
Two important properties of invertible matrices are easily proved
directly from the definition. The first one ensures that, no matter how an
inverse to a matrix A is computed, the resulting matrix A~ x is always the
same.
3.3 Theorem
Proof. Suppose there are two matrices, B and C, such that both
AB = BA = / and AC = CA = /.
3.4 Theorem
If A =A X A2 . . . A n and all of A
,
x , . . . , AH are invertible, then A is
Problems 7 and 8 show that matrix operations with diagonal matrices are
particularly simple.
EXERCISES
-2
C =
Sec. 3 Matrices 23
2. Show that for any matrix A and zero matrices of appropriate shapes,
AO =0 and OA = 0.
(a) AX = B + Y. (d) CX + DY = 0.
A(B + C) = AB + AC
for matrix multiplication.
1 2\ 2 6
5. Let U= V= . Compute UV and VU. Are they the
2-4/ \1 3/
same? It is possible for the product of two matrices to be zero without
either factor being zero?
6. Let X
Let D - diag(l,2, 3)
b\
10. (a) Let A = I
J
be a 2-by-2 matrix with ad # be, and let A~ x be given
2 1/' \l 0/ \1 3,
zero.
b\
12. Show that if ad = be, then I I is not invertible.
(a) A = (b) A =
(
4 3^
Ans. (a)
li
(Note that is the only number whose cube is 0. Part (b) of this problem
thus illustrates another difference between the arithmetic of numbers and of
matrices.)
(a) Show that if A = /or -/, then A2 = /, where /is an identity matrix of
any dimension.
(a b\- (\ 0\
(b) Show that I I = I I if a2 + be = 1 ; so the equation
Sec. 4 Linear Functions 25
SECTION 4
The product of an m-by-n matrix and ^-dimensional column vector LINEAR FUNCTIONS
(n-by-l matrix) is an w-dimensional column vector. An w-tuple in %
n
tion such as
fyi
) =
2 3
~4
\yj
~ (
\l -1 2)
jx = 2x + x 3x 2 — 4x 3
y<i = Xi X2 ~r 2X3,
as may be seen by simply writing out the result of the matrix multiplication.
Thus the vector function described by a matrix amounts to multiplication
of a domain vector by the matrix.
Functions given by matrix multiplication can be characterized by some
very simple properties. Note that the definitions given below apply to
functions between any two vector spaces, although we are at present
concerned only with the spaces 31".
C
hold for all vectors x, y in \J, and all numbers r.
4.1 Theorem
g is linear,
= *ig(ei) + • • • + x n g(e„).
By the definition of A we have
g(x)
a 21 x x + . . . + «o„.v„
^/,„ l
.v, + . . . + fl,„, r v,
7
g(x) = /fx.
The proof given above actually shows how to construct the matrix
corresponding to a linear function. This construction is important and
we summarize it as a theorem for emphasis.
Sec. 4 Linear Functions 27
4.2 Theorem
b d]\o)
~ \b)' \b d)\\) \d t
By Theorem 4.2,
2
28 I 'ectors and Linearity Chap. 1
m = Kg
*•> - ("$
(a) (b)
Figure 10
The geometrical effect of/ is illustrated in Fig. 11, in which C is the unit
/*i\ /"i\
= =
image of
W under/ then x x
\uj and .v 2 Ui.z . Hence, if
\W2/
I is in
Theorem 4.2 says that, f/is a linear function from :K" to 3t"\ then the
;
jih column of the matrix of/is/(e ). Because we can write any vector x in
;
x = c^ + . . . + c„e„,
Figure 11
Sec. 4 Linear Functions 29
Since x is an arbitrary vector in the domain of/, the last equation expresses
4.3 Theorem
/(
= "yi + ^2
J
is linear (why?), and Theorem 4.2 says that the matrix of/has as columns
the two vectors
- J, and = y.
/(J) /(J)
For example, if
then
[\
VI
'
1
and indeed the range of/is spanned by the columns of the 3-by-2 matrix.
= ty x + v .
30 Vectors and Linearity Chap. J
4.4 Theorem
is certainly one-to-one. For only one vector is carried into any other
vector by the rotation. Alternatively, the zero vector is the only vector
carried into the zero vector. In algebraic terms, the rotation is pro\ed in
Example 1 to be representable by
73
1
- x — —
n/3
v
Then Theorem 4.4 shows that the one-to-one property of/ is equivalent
to the equations
-'
2
"
1
v/ 3
Sec. 4 Linear Functions 31
having only the solution (x,y) = (0,0). This of course can be verified
directly by solving the equations.
If/ and g are any two functions (not necessarily linear) such that the
range space off is the same as the domain space of g we define the com-
positiong °f to be the function obtained by applying first /and then g.
More explicitly, if x is in the domain of/and f(x) is in the domain ofg,
then g °/(x) is defined as g(f(x)). If /(x) or g(f(x)) is not defined, then
g °/(x) is not defined.
Composition of linear functions lies behind matrix multiplication. In
introducing the concept of matrix multiplication, we considered a function
from .'H 2 to :K 2 given by
*\ = 3j>i - y2
?2 = 5)>i + 2y 2
3 2
and another from :K to .'K given by
yx = 2x + x 3x 2 + 4.y 3
JAj = Xi X2 -\- ^v 3 .
We then computed that the composition of the two functions was given by
the formulas
z, = 5*! + 10x 2 + \0x 3
/ 5 10 10\
of the composite function as the product of the matrices
\12 13 24/
'3 -l\/2 3 4\
5 2/ \ 1 — 1 2,
for the original functions. The following theorem states the important
fact that the composition of linear functions is always given by matrix
multiplication.
4.5 Theorem
B =
M e 2 =&*l)
gi*i)
Figure 12
v5
34 Vectors and Linearity Chap. 1
4.6 Theorem
0/ l-i/
«w>-(_3-(_;
/0 -1\
Therefore ^(x) = I x for all x and
(2\ /0 -l\/2\ / 1\ /0
which agrees with the geometric description off. See Fig. 13.
Linear and affine functions are often called linear and affine transforma-
tions, though we more often use the term function in this book.
Sec. 4 Linear Functions 35
'
(b) What matrix corresponds to reflection in the line through the origin
135 counterclockwise from the horizontal?
(c) Compute the product of the matrices in (a) and (b) and interpret the
result geometrically.
Am. (b) "
')
(
\ - 1 0/
6. (a) Find the 2-by-2 matrix .\f r corresponding to reflection in the line-
through the origin at an angle x from the horizontal. Check your
result against Exercise 5 for a 45 . a 135 . What is M\l
(b) Let -'
be another angle and compute the product M M«. X
Show that this
represents a rotation, and identify the angle o( rotation. When does
MM x p
Mp MS
7. (a) Show that
and (
8. Show that a function /from one vector space to another is affine if and only if
for all numbers rand all x, y in the domain off. [Hint. Consider the function
g(x) f(x) /(0).]
9. Let / be a linear function from a vector space 'i to a vector space 10. Show
that if x, x„ are linearly independent vectors in U and /is one-to-one,
then the vectors /(x,) /'(*„) are linearly independent in 10. [Hint. If
/ is one-to-one, and f(x) 0, then x 0.]
3 3
10. (a) Prove that if /is a linear function from :K to ft and /is one-to-one,
then the image / (I.) of a line /. by f is also a line.
(b) Show by example that, if the linear function of part (a) fails to be one-to-
one, then the image of a line by / mav reduce to a point.
(c) Show that if L ]
and I.
z
are parallel lines and /'is a linear function, then
/(/.]) and I (L,,) are parallel lines, provided that / is one-to-one.
11. Show that the composition of two affine functions from a space into itself
is affine. Suppose f(x) Ax b, ;mxi (\ d. Suppose*/ g)(x)
13. Show that an affine function A is one-to-one if and only if A(x) = A(0)
always implies x = 0.
To allow the full application of vector ideas to Euclidean geometry, we DOT PRODUCTS
must have a means of introducing concepts such as length, angle, and
perpendicularity. We will show in this section that all these concepts can
be defined if we introduce a new operation on vectors. If x (jcx xn) =
and y = (y\, ,yn) are vectors in Jl", we define the dot product or
. . .
x-y =x y +
1 1 . . . +xnyn .
Symmetry x •
y = y x. •
Additivity: (x -j-
y) • z = x z + y • • z.
3
the origin to the point with coordinates i \,. x 2 x3 ). ,
In :K the Pythagorean
theorem gives us a simple formula for the distance (see Fig. 14). Letting
|x| stand for the length of the sector x. we have
|x|- x\ \
x\ \
x\. (I)
Thus we see that the length of the vector x can be expressed in terms of the
dot product as \ x • x |x|. Note that we use the same symbol for the
length of a vector as for the absolute value of a number. Indeed, if we
think of a number as a one-coordinate vector, its length is its absolute
value. In :H" we define the length of a vector by the same formula that
works in 3l 3 |x| : \ x • x.
Next we would like to express the angle between two nonzero vectors
in :H". The usual convention is to take this angle 6 to be in the interval
: tt (see Exercise 1). The solution to the problem is provided by
the following theorem.
x y
Figure 14
5.2 Theorem
x-y
cos
W |y|
Proof. Let us apply the law of cosines to the triangle shown in Fij
|x — y|
2
= |x|
2
+ |y|
2
— 2 |x| |y| cos 6,
(x — y) • (x — y) = x • x -f y •
y — 2 |x| |y| cos 6.
Hence,
2x •
y = 2 |x| |y| cos 6,
yy= (-i) 2
+ i
2
= 2,
and
xy = -1+3 = 2.
Hence
x •
y = 0. (2)
have to write down three conditions, two for the perpendicularity require-
ments, using (2), and a third condition to assure length 2:
Xj -f 2x 2 + 3.v 3 = 0,
X, - X3 = 0,
x\ + x\ + x\ = 4.
These equations have the pair of solutions x = ±(vf, — vf, V§).
If n is a unit vector, that is, a vector of length 1 , then the dot product
n • x is called the coordinate of x in the direction of n. The geometric
interpretation of n • x is shown in Fig. 17. For since cos — (n • x)/|x|,
it follows that n • x is either the length of the perpendicular projection on
the line containing n, or else its negative. The vector (n • x)n is called the
component of x in the direction of n and is sometimes denoted x n .
Figure 17
Figure 18
)
we first find the unit vector in that direction. Since |(1, 1, 1)| = \ 3, the
(F . n)n = (W .(^A) , 2) n
V3 \3 3 3/
Any vector space on which there is defined a product with the properties
5. 1 is an inner product space. Thus %n is an inner product space,
called
and some other examples are given in Problems 8 and 9. Inner products
in spaces other than Jl" are used in this book only in Chapter 2 Section 4
and Chapter 5 Section 5.
In terms of the inner product we can always define the length or
norm of a vector by
|x| = VX • X.
The proofs of the first two are easy and are left for the reader to check.
The proof of the third is harder and will be taken up later, though we
remark here on its geometric significance, illustrated by Fig. 19.
The fact that length has been defined in terms of an inner product
leads to some properties of length that are not derivable from those
already listed in 5.3. First we prove the following.
Figure 19
Proof. We assume first that x and y are unit vectors, that is, that
or
xy < 1.
|x-y|
<1,
|x| |y|
cos 6 = ^^- .
|x| |y|
we get
|x + y|
2
= + y) (x + y) = |x| + 2x
(x •
2 •
y + |y|
2
<|x| + 2|x||y| +
2
= (|x| + |y|
2
|y|)
2
,
fa n a 12
A =
a\ 2 + a\ 2 + . . . =
and a lx a 12 + a 21 a 22 +
1 = 0. We form the matrix
. . .
A 1
, called the transpose of A, by reflecting A across its main diagonal. Thus
'«n 21 • •
a 12 a 22 .
5.5 Theorem
I 1 V3
\ 2 2
and so A~ x =A 1
.
The second of the above two square matrices is obtained from the first
by interchanging rows and columns. When two matrices are so related
we still say that one is the transpose of the other, even if they are not
square. Thus if
1 2 3
A =
4 5 6
then
T 4\
2 5
3 6,
Obviously,
(A 1
)
1
= A for any matrix A.
EXERCISES
1. Show that the natural basis vectors satisfy e,- • e,- = 1, and et
- • e, = if
i *].
2. (a) Find the angle between the vectors (1, 1, 1) and (1, 0, 1).
(b) Find a vector of length 1 perpendicular to both vectors in part (a).
4. Prove that the dot product has the properties listed in 5.1.
6. Find the coordinate of the vector x = (1, —1,2): (a) in the direction of
n = (1/V3, l/v/3, 1/V3) and (b) in the direction of the nonunit vector
(1,1, 3). (c) What is the component of x in the direction of n?
7. (a) Prove that any vector in 31" and n is a unit vector, then x can be
if x is
written as x = y +
z, where y is a multiple of n and z is perpendicular to
n. [Hint. Take y
to be the component of x in the direction of n.]
(b) Show that the vectors y and z of part (a) are uniquely determined. The
vector z so determined is called the component of x perpendicular to n
Sec. 5 Dot Products 45
8. (a) Consider the vector space C[0, 1] consisting of all continuous real-
valued functions defined on the interval <x<
[The sum of/ and
1.
<f,g)=\f(x)g{.x)dx
x * y = Xiji + 2x 2y 2
is an inner product,
(b) With length defined by jx] „.
= (x * x) 1/2 , sketch the set of points
satisfying |x|„. = 1.
10. Show that if |x| = |y| in an inner product space, then x | y is perpendicular
to x - y.
/'cos 9 —sin 6
ysin 6 cos 6
13. (a) Show that if A is a 2-by-2 orthogonal matrix, then \A\\ = |x| for all
vectors x in R2 .
15. Let
2-1
and B
3-5 2
Compute AB, BA, A B\ and B A l l l
.
H>. (a) I or any two matrices A and //. show that it' /)# is defined, so is Z?'/F,
. and that IV A' (Mi)'
(b) Slum that if I in imciliblc, then so is .1'. and that (/f) '
(/J
]
)'.
si ( HON (,
II (I 11)1 \\ In this section we develop some facts and formulas about lines and planes
GEOMETRY in :ii- and 3l 3 . Some of the ideas will reappear in Section 1 of Chapter 3.
2 3
Recall that if x, is a nonzero vector in 5l or :l\ , then the set of all
numerical multiples tx }
is a line L passing through the origin, and any
line parallel to L„ consists of the set of all points /x, f-
x , where x ()
is some
fixed vector. An alternative way to say the same thing is: a line is the
range of a function of the form f{t) tx x„, where x, }
0.
parallel toL, and the points /(a 2 a,) a, make up L itself. For example, |
when I \vc get a, and when / we get a 2 See Fig. 20. Alternatively,
. I .
Figure 20
3
To determine a plane in .'fi , we take two noncollinear vectors x, and x 2
(that is, so that neither is a multiple of the other) and consider all points
ux + t'x 2 where u and v are numbers. These points form a plane P
x
through the origin. A parallel plane will then consist of all points u\ + l
g 3
saying that a plane is the range of a function 'M- > .'H , where g(u, v)
passing through them, they must not lie on a line. (Then the vectors a 3 — a,
of all points
w = iv(a 3 - aO + r(a 2 - a,) + a,.
Sec. 6 Euclidean Geometry 47
Figure 21
p • (x -x = ) (1)
p •
Xi = U l U 2 Vi — UiU 3 V 2 +u 2 u3v x —u 2 u x v3 + u u v — U U V = 0.
3 t 2 3 2 X
In other words, the given plane consists of all points (x, y, z) with co-
ordinates satisfying that equation.
The simplest way to sketch the plane satisfying an equation like (3)
is to pick three points with simple coordinates that satisfy say (0, 0, f), it,
(0, \ , 0), and (— f , 0, 0). Then locate the points relative to coordinate
axes and sketch the plane using the three points as references. We have
purposely chosen in our example points that lie on the coordinate axes.
See Fig. 23.
Figure 23
If p is a nonzero vector in 2
ft and x is a point in ft
2
we can determine ,
This is one of the several forms for the equation of a line in the xv-plane.
The slope of the line is evidently —p /p 2 1
.
6.1 Theorem
But n • n = 1, so we obtain t +n •
y = c, from which the theorem
follows.
Hence the distance is f . Notice that the equation of the plane could also
be written 3x — 4z = —3 and, in normalized form, (f)x — (|)z = — f.
EXERCISES
1. Write a formula in the form/(r) = tin x + x for a line containing x and
parallel to \ v In each case determine whether the point a lies on the line,
that is, whether a is in the range off.
5. Find an equation for the line in :ft 2 that is perpendicular to the line tx x + x„
and passes through the point y 2 .
8. For each of the points and planes or lines listed below, find the distance
from the point to the plane or line.
9. (a) Verify that the cross-product of x x and x 2 [formula (2) in the text] is
perpendicular to x 2 .
(b) Find a representation for the line perpendicular to the plane consisting
of all points w(l, 2, 1) + v( — 1, 0, 1) + (1,1,1) and passing through the
origin.
2x + 3y + z = 1
(b) Do the same as in part (a) for the general equation p • (x —x) = 0,
when p ^0. [Hint. If p ^=0, then p has a nonzero coordinate.]
x = (x • e^e! + (x • e 2 )e 2 + (x • e 3 )e3.
(b) If the vector x in part (a) is a unit vector, that is, a vector u of length 1
show that u • e^ = cos a,, where a, is the angle between u and e,. The
coordinates cos a t
are called the direction cosines of u relative to the
natural basis vectors e,. If x is any nonzero vector, the direction
cosines of x are defined to be the direction cosines of the unit vector
x/|x|.
(c) Find the direction cosines of (1, 2, 1).
(d) Show that parts (a) and (b) generalize to &n .
12. Let u and v be points in ft". Show that the point ^u + ^v is the midpoint of
the line segment joining u and v.
3
13. Let u and v be noncollinear vectors in :R . To find a vector x perpendicular
to both u and v, we solve the equations u • x = and v • x =0, that is,
Sec. 7 Determinants 51
solve
u xx + u 2y + u3z =
(*)
v xx + v 2y + v3z = 0,
(u x v 2 - u^Jx = (u 2 v 3 - u3 v 2)z.
(x, y, z) = {u 2 v 3 - u3 v 2 u 3 v 1
,
- u x v3 u x v 2
,
- u 2 v x ).
Then
-9
5, An
4 2
-5 -6^
a 23 = 0,
-3 4)
=
(4), &la =
det A = a.
= a n det y4 n — a 12 det A l2 + . . .
— (— \)"a ln det A ln .
In words, the formula says that det A is the sum with alternating signs, of
the elements of the row of A, each multiplied by
first the determinant of
its corresponding minor. For this reason the numbers
(_ 1)2+1 det /
) =40
\ 4 2/
The factor (
— 1)'>J associates plus and minus signs with det A n according
to the pattern
/+- + -••
+ - + • •
+ - + • •
Sec. 7 Determinants 53
Example 2.
1 2\
(a) det [
= 1 (4) - 2(3) =4-6= -2.
la b\
(c) det I = ad — be.
®Z\X\ I #22-^2
= '"
2°
equation
(a 22 a n - a 12 a 2l )x l = (a 22 r x - a 12 r 2 ).
is
I
W
the result of replacing the
"J
is the matrix ol coefficients
first column of A by
,-
see in Theorem 8.4, a similar result holds for systems of // linear equations
in /; unknowns, for all values of n.
Since our definition of determinants by 7.1 is inductive, most of the
proofs have the same character. That is, to prove a theorem about
determinants of all square matrices, we verify it for 1 -by- 1 (or in some
cases. 2-by-2) matrices, and also show that if it is true for (// - l)-by-
(// 1) matrices, then it holds for //-by-" matrices. In the proofs, we
give only the argument for going from step // I to step //. The reader
should verify the propositions directly for 1 -by- 1 and 2-by-2 matrices;
the verification is in all cases quite trivial. A, B, and C will always denote
n-by-n matrices. We write a, for the /lh column of a matrix A. If A has //
7.2 Theorem
detB=i(-l) tn />,,detB u .
A-=l
7.3 Corollary
Example 3. Let
1 2 3\ / 1 6 3N
1 2 41, 5=1-164
12/ \ 3 2,
=0+4-3=1
det B= - 6(-2 - 0) +
(1)(12 - 12) 3(-3 + 0)
7.4 Theorem
while
Cij det C = 1;
a Xj det C +
1? fe
1;
- det CXi
= a t j det Au + 6 l3 det - 2? l3 .
Hence
fc+1
detC=i(-l) c [Link] 1 ,.
= det A+ det B.
56 Vectors and Linearity Chap. 1
Example 4. Let
1
Sec. 7 Determinants 57
(Xi bx cx
x2 b2 c2
x3 b3 c3
lb 2 c2
\
lx 2 c2
=x x det — bi det
\
-f c x det
\b 3 cj \x 3 c3 J \x 3 u3
= Xiibfa - b 3c 2 ) - - x c + c^xfo - x b
b x (x 2 c 3 3 2) 3 2)
7.6 Lemma
If B is obtained from A by exchanging two adjacent columns, then
det B = — det A.
Proof. Suppose A and B are the same, except that a ; = b J+1 and
a J+1 =.bj. For k #y or j -f 1, we have b lk Blk = = a lk and det
— detA lk by the inductive hypothesis, so Blk = (— l) k+1 b lk det
— (— l) u det A lk On
fc+1
tf the other hand b Xj = a lj+1 and Bu =
.
a lj+1 detA lj+1 Similarly (-\)> +2 b lj+1 det B 1J+1 = (-l) 3+1fl„ det A
.
v .
Then
det = (1)(8 - 5) - (3)(-4 - 3) + (-2)(l0 - (-12))
A
= 3 + 21 - 44 = -20
det B = (1)(5 - 8) - (-2)(10 - (-12)) + (3)(-4 - 3)
= -3 + 44 - 21 = 20
= -det A
,
1.1 Theorem
by Lemma 7.6 each step changes the sign of the determinant. Since
2k + 1 is an odd number, det B = — det A.
7.8 Theorem
/3 0\
,
Example 6. Multiplication by gives a function Jl 2 -> 51 2 which
J
1 2
• Z \
For another example, consider the function g given by the matrix I
which has determinant 1. Its effect is illustrated in Fig. 24(b). The unit
square is mapped into a parallelogram with the same base and altitude,
so the area remains unchanged. The composition g of multiplies areas
by 6 [since f(S) has 6 times the area of S and g(f(S)) has the same area
as/(5)]. The matrix of g °/is given by the matrix product
Sec. 7 Determinants 59
(0,2)
(0,1)
60 Vectors and Linearity Chap. 1
If A and B are any two square matrices of the same size, then
det (AB) = (det^)(det5).
Proof. Let
L(Xj, . .
.
, x„) — det A det (x ls . . . , xn) — det (Ax Y , . . . , Ax n ),
where x l5 . . . , x„ are vectors in 51". Clearly L is linear as a func-
tion of each vector Xj. Furthermore, L(e, , . . . , e, ) = for any
set {e, , . . . , e, } of natural basis vectors. The reason is that if
Lib,, . .
. , b„) = I.(i><A, . . . , jU»e<)
Hence,
det A det B- det /*J? = det A det (b lt . .
.
, bj
- det (Ab u . . . , Ab n )
= L(bls . . . , b„) = 0.
Note that the proof just given uses only Theorems 7.5, 7.7, and 7.8
and does not use Theorem 7.9. (The point is important because the
product rule is used in the proof in Section 7 of the Appendix, on which
the proof of Theorem 7.9 depends.)
Sec. 7 Determinants 61
The natural unit of area in 3l 2 is given by the unit square with edges
(1,0) and (0, 1), and the natural unit of volume in iR 3 is given by the unit
cube with edges (1 0, 0), (0, 0), (0, 0, 1). In general we take the unit of
, 1 ,
volume in 31" to be that of the cube whose edges are the natural basis
vectors e l5 e„ that form the columns of the n-by-n identity matrix.
. . . ,
tn satisfies the condition < tt < 1. The resulting set of points is called
the parallelepiped determined by its edges x x x n If we choose only , . . . , .
Figure 25
7.11 Theorem
Let a l5 a 2 ,
. . . , a n be n vectors in 31". Then the volume of the
parallelepiped with edges a l5 . . . , a„ is |det (a l5 . . . , a„)|.
the factor |det (a l5 a„)|, and the cube has unit volume, the
. . . ,
r x cos X
r2 cos 0;
Example 7. Let a x , a .. The vectors have
r x sin X
r 2 sin 2
Figure 26
r x cos L r2 cos 2
det /^(cos X sin 2
— sin 6 X cos 2)
/*! sin d 1 r 2 sin 2
= r x r 2 sin (0 2 — X)
Figure 27
with coordinates
«2
64 Vectors and Linearity Chap. I
Figure 28
ordinary volume is very much like the relation between directed distance
on a line and ordinary distance. Indeed, oriented volume may be con-
sidered a generalization of directed distance, and we use the idea in
Chapter 7, Section 7.
EXERCISES
1. Find AB, BA, and the determinants of A, B, AB, and BA when
/l -2
(a) A [Arts, det AB = -14.]
,3 1
'2
| 3
v
4
(a)
66 Vectors and Linearity Chap. 1
3. Show that if D is the diagonal matrix diag (/-,,..., r„), then det D =
/,;_, . . . rn .
6. Apply the product rule to show that, if A is invertible, then det A # and
(det A l
) (det/4)- 1 .
7. Let /4 be an m-by-m matrix and B an n-by-n matrix. Consider the (//; /;)-
IA Ov
by-(/w «) matrix which has A in the upper left corner, B in the
\0 B/
lower right corner, and zeros elsewhere. Show that its determinant is equal
to (det /l)(det B). [Suggestion. First consider the case where one of A or B
is an identity matrix, and derive the general result from the product rule.]
10. Find the volume, and the area of each side, of the parallelepiped with edges
(1, 1, 0), (0, 1, 2), (-3, 5, -1).
[Ans. volume = 17, areas = Vl66> V66, 3.]
13. Find a representation for a line perpendicular to (2, 1, 3) and (0, 2, —1),
and passing through (1, 1, 1).
3
14. Let P be a parallelogram determined by two vectors in :K . Let Px P v and
, ,
A 2 (P Z ).
15. (a) Verify by direct coordinate computation that |u x v|
2
= |u|
2
|v|
2 -
2
(u • v) .
(b) Use the result of part (a) to show that |u x v| = |u| |v| sin 0, where 6 is
(c) Show that |u| |v| sin is the area of the parallelogram with edges u and v.
16. The complex numbers can be extended to the quaternion algebra JC, which
is a four-dimensional vector space with natural basis {1, i,j, k). Thus a
R
(a) Show that the quaternion product of two elements of S is not necessarily
in S.
(b) Define a product on S by first forming the quaternion product and then
replacing its real part by zero. Show that the resulting product is the
same as the cross-product in R3 .
a = fliUj + o2u 2 + fl
3u3
b = b^ + b 2u 2
c = c^.]
SECTION 8
det^=i(-l) m a 1;.det,4 iy ,
where, in general, A i}
; denotes the (« — l)-by-(« — 1) minor corresponding
to a tj . In the present section we prove more general formulas of the same
kind. These formulas, which apply to any n-by-n matrix A, are
Formula 8.1 holds for each integer j between Formula 1 and n, while
8. 1 R holds for each integer Formula 8.
i between 1 and n. (For / = 1 , 1
postpone the proof of the formulas to the end of the section, first showing
some of their consequences.
Example 1. Let
12 3 4\
A = I 5 6 7
\8 9 0/
/3 4\ (2 4\ (2 3
-5 det +6 det - 7 det
\9 0/ \8 0/ \8 9
IS 6\ (2 3\ (2 3
4 det - 7 det + det
\8 9/ \8 9/ \5 6
= (4)(-3) - (7)(-6)
= -12 + 42 = 30.
2= (-iy+^det^,,
t' l
where x 1 , . . . , x n may be any set of n numbers. From 8.1 we see that this
is equal to a certain determinant; in fact it is the expansion by the yth
column of the matrix obtained from A by replacing column with
they'th
(Xj, . . . , -Y,,). Now consider what happens if we x n equal
take xlt . . . ,
8.2 Theorem
For any A7-by-/? matrix A,
i(-D'
!
det Au
:
v/ i\f+; a a
(det /I if k = i
The number (— \)
i+i det Au is called the ijth cofactor of the matrix A.
We shall abbreviate and write A for the matrix with entries a ti
it as a u .
the y'th column of A) and the kth column of A. That is, it is the sum
n
^ a tj a lk . By Theorem 8.2 this is det A if y = k, and otherwise. Hence
«=i
A'A is equal to (det A)I, a numerical multiple of the identity matrix. A
similar calculation using 8.2 shows that AA is also equal to (det A)I. l
8.3 Theorem
12 3 4\
A = j
5 6 7
\8 9 0,
70 Vectors and Linearity Chap. 1
To obtain the matrix of cofactors, insert the factors ( — l) i+i , changing the
sign of every second entry and giving
63
Sec. 8 Determinant Expansions 71
the other hand, A(A~ l \>) = (AA'^b = b. In other words, the equations
have a unique solution, and it is /l _1 b. They'th entry in the column vector
A~ x
\t is the matrix product of the y'th row of A~* and the vector b. If
det A 7^ 0, we may express the elements of A' 1
in terms of cofactors of A
and obtain
for this product. From 8.1 this may be recognized as (det A)' 1 det B U)
where B U) is the result of replacing they'th column of A by b. We have
proved:
det B U)
x< = ,
det A
xx —2x 2 +4x 3 = 1
2x x +3x 2 —x = 3 3.
We have
1 -2
-1 1
2 3
1 1
-1 2
2 3
72 Vectors and Linearity Chap. I
We have not made any assertions in this section about what happens
if the determinant of the coefficient matrix of a system of equations is zero.
It an easy consequence of the product rule that a matrix with zero
is
Since 8.1 withy = 1 is exactly like 7.1 except that it refers to the first
column instead of the first row of a matrix, it is clear that transposing a
matrix (which just exchanges the roles of rows and columns) should not
affect the value of the determinant. The formal statement and proof
follow.
8.5 Theorem
of the matrices. We shall not always bother to write out these corre-
sponding theorems, but shall refer to the "row" version of a numbered
statement by using the same number with an R after it. (We have already
numbered 8.1R and 8.2R to conform to this convention.) For example,
Theorem 7.2R would read: If B is obtained from A by multiplying some
row by the number r, then det B = r det A.
Theorem 8.5 implies that, from any theorem about determinants that
involves columns of matrices, we can derive a corresponding theorem
involving rows instead, by applying the given theorem to the transposes
of the matrices. In particular, 8.1R is a consequence of 8.1 and 8.5. We
shall not bother to write out the row versions of the other statements
but may refer to them by the original statement number with an R after it.
8.6 Theorem
-4
The third column of C is equal to the third column of A plus 2 times the
first column. As in Example 5, Section 7, det A = —20. Then det C =
—20. As a check,
(b) Let
A =
Sec. 8 Determinant Expansions 75
8.7 Lemma
If the first column of the matrix A is e8 , then
3 =1
column of A is e l5 then a u = 1, while for > 1 the first
If the first /'
holds for the case i = 1 For / > 1 we need to use the inductive
.
det ,4 = 2(-l)> +1 a l3 .
det Ar
3 =2
Each minor A Xi has zir_x for its first column. (Removing the top entry
from the vector e f in 31" gives e z _! in 3l n_1 .) By the inductive
hypothesis,
3=2
By Formula 7.1 the right side of this equation is (— l) i+1 det i?,
Proof of 8.1. We first prove 8.1 for the special case j = 1. The first
+ flni det ( e n, a 2 , . . . , aj
= |(-ir a det^ l
il 1.
j=i
76 Vectors and Linearity
Thus we obtain
= i(-l)*"a„det4„
as was to be proved.
EXERCISES
1. Using appropriate cases of 8.1 or 8.1R, express each of the following as a
linear function of the jc's.
(a)
(b)
Sec. 8 Determinant Expansions 77
(a) Ix + 6y = 5
'
6x + 5y = -3.
(b) 2x +y =
3y + z = \
Az + x = 2. M«5. x = --£ 2 S -I
(c) x x + x 2 + x3 + x4 = — 1
x — x2
1 + 2x4 =
->x 1 — X3 == *
x2 — xt = 0.
/"'
78 Vectors and Linearity Chap. 1
(b) Consider tridiagonal matrices in which the entries on the diagonal all
have the value 2 and the entries next to the diagonal all have the value 1.
Let dn be the determinant of an n-by-n matrix of this type. Find a formula
for dn [Suggestion. Start out by seeing what happens for n = 2, 3, 4.]
.
2
Linear Algebra
SECTION 1
a u Xi + . . . + a ln x n = bx
1.1 ...
a mlX l + • • • + a mn X n = t>m
where the a's and fs are given and the ;c's are to be determined. The whole
system can be written in matrix form as Ax = b, where A is the m-by-n
coefficient matrix with entries o i3 b is a column vector in %
,
m and x is a
,
79
80 Linear Algebra Chap. 2
Example 1.
3x + \2y + 9z = 3
2x + 5y + 4z = 4
-x + 3y + 2z = -5.
Multiply the first equation by \, which makes the coefficient of* equal to 1
and gives
x+ Ay + 3z = 1
2x+ 5y + 4z = 4
-x + 3y + 2z - -5.
Add (—2) times the first equation to the second, and replace the second
equation by the result. This makes the coefficient of x in the second
equation equal to and gives
x+ 4y + 3z = 1
- ly - 2z = 2
—x + 3y + 2z = —5.
Add the first equation to the third, and replace the third equation by the
result, to get
x + Ay + 3z = 1
- 3y - 2z = 2
ly + 5z = -4.
x + Ay + 3z = 1
y+&= —§
ly + 5z = -A.
Sec. 1 Linear Equations, Inverse Matrices 81
Add (—4) times the second equation to the first, and (—7) times the
second equation to the third, to get
x ~r ~zZ = ~3—
y + §z = -I
17
3^
—
— 3- 2
x ~r 3^ = "3"
y + |z= -f
z= 2.
Add (—5) times the third equation to the first and (— f) times the third
equation to the second to get
x= 3
y = -2
z= 2.
Clearly, this sytem has just one solution, namely, the column vector
have found a solution for them. This verification of course does not rule
out the theoretical possibility that the original equations might have other
solutions as well. In fact the final system is equivalent to the original
system and has the same set of solutions. The same is true for any pair of
systems where one is obtained from the other by steps such as were used
in this example. Before we can prove this, we must first state exactly
-1
1 |x = I -2
v
1/ \ 2,
which, since the matrix on the left is the identity matrix, simply amounts
to saying that x is equal to the vector on the right.
The theorem that justifies our method of solving linear equations is as
follows.
Sec. 1 Linear Equations, Inverse Matrices 83
1.2 Theorem
-3
Add (— V) times the row to the second row, and then add 3 times the
first
first row to the third row to produce zeros in the second and third entries
of the first column, and obtain
'1 -2 -3\ / 2^
1 5|x= j
-6
,0 -1 -5/ \ 6;
Add 2 times the second row to the first and then add 1 times the second
84 Linear Algebra Chap. 2
1 7\ /-ION
1 5 )x = I -6
,0 0/ \ Oy
x +7z=-10
y + 5z = -6
Ox + Oy + Oz = 0.
The third equation is satisfied for any values of x, y, z. The first two
equations may be rewritten as x — — 10 — Iz and y = —6 — 5z. Thus,
for any value of z,
is a solution, and every solution has this form for some value of z. We
have now described the set of solutions of
and by Theorem 1.2 we know that this is the same as the set of solutions
of the matrix equation we started with.
1 -2 -3\ /2\
\ -2 —^ x= 7
-3 5 4/ \2,
Sec. 1
2. Carrying out
85
ION
2y
/, 7^
Whatever x is, the third row in the product 10 1 5 |x will be zero
— 0—
\0 0/
because the third row of the left factor is Thus no value of x can give
zero.
a column vector with 2 in the third row, and we conclude that the equation
has no solution.
equations, we obtain
x + 7z=-10
y + 5z = -6
Ox + Oy + Oz = 2.
The last equation obviously cannot be satisfied for any values of x, y, z.)
1.3 Theorem
'2 4
A =
We start with
4
-3 -1,
Add —2 times the second row to the first and —1 times the second row
to the third to get
'0 4 8\ /l -2 0\
1
Sec. 1 Linear Equations, Inverse Matrices 87
Multiply the first row by \ and then add 3 times the first row to the third
to get
'0 1 2\ n -\ ON
o I, I 1
Multiply the third row by — 1 and then add —2 times the third row to the
first to get
'0 1 0\
1 I.
1,
v
1 0\ /
oio, 1
.0 1/ \-|
The last matrix on the right is A~ x , as may be verified by multiplying by A.
EXERCISES
1. Solve the following systems of equations:
'
/
88 Linear Algebra Chap. 2
1 2
(a)
5 6
Ans.
SECTION 2
SOME In this section we shall look at some applications of vectors and linear
APPLICATIONS equations. The selection of examples is made so as to avoid technical
complications from the fields of application. Several of our examples
involve the notion of a network, which we define to be a finite collection
of points, or nodes, some of which may be joined to some others by line
segments. It is theoretically unimportant whether a network is visualized
as lying in 2-dimensional space or 3-dimensional space; we choose which-
ever is pictorially more convenient. Some networks are illustrated in
Fig. 1.
Sec. 2 Some Applications 89
Figure 1
(»< - v t)
(1)
where c u is the current flowing from node to nodey, r a is the resistance /'
of the connection between nodes i andy, and vt and vt are the values of the
electrical potential at nodes
andy. (The appropriate units of measure-
/'
Figure 1(a) shows a circuit with four nodes and five segments, with
the resistance of each segment indicated beside it. Suppose an external
power source is connected at nodes 1 to 4 to maintain values v 1 = 12
and v4 = 0. Since node 2 has no external connection, the current flowing
in must balance the current flowing out, so that if signs are taken into
account, the sum of the currents out of node 2 must be zero. Using (1),
we get the equation
which are the reciprocals of the resistances in the lines joining node 2 to
the others. A similar equation will hold at any node that does not have an
external connection. Thus at node 3 we get
+ fy8
-v 2 2.
^(12 — 6.22) = 0.96. The total current flowing from node 1 into the rest
of the network is then 2.33 + 0.96 = 3.29, which must of course be equal
to the current flowing into node 1 from the external source.
-»-f,
(a) (b)
Figure 2
,
then by definition
r =f + 1 f2 +f = 3 (l,2).
But suppose we are given the directions of the three force vectors
and are asked to find constants of proportionality that will produce a
given resultant, say, r = (— 1, — 1). In other words, suppose we want to
find nonnegative numbers clt c 2 c 3 such that
,
Cjfa + c 2 f2 + c 3 f3 = (— 1, — 1).
(A negative value for some one of the c's would reverse the direction of
the corresponding force.) This vector equation is equivalent to the system
of equations we get by substituting the given vectors (1) that determine
the force directions. We find
Ci
IMaMiH-i) (3>
—c +x 4c 2 — 2c 3 = —1,
3c +x 3c 2 — 4c 3 = — 1.
C\+ c2 = 1
with magnitudes
IcAl = fVlO, |c 2 f2 |
= 2, |c 3 f3 |
= 2^5.
the zero vector. We would then have replaced the vector (— 1 , — 1) on the
right side of Equation (2) by (0, 0), and solved the new system in a similar
way.
O l«5
(a) (b)
Figure 3
Pi = £ + (£)/>2
Similarly, because going to a 4 does not occur in the events we are watching,
Pi = (¥)Pi + i\)Pz
Pa = (*)/>•
Pi - (i)/>2 = a
-(£)/>! +/>2-(i)/>3 =
— (h)P2 + Pz =
Sec. 2 Some Applications 93
Pi = f» Pz — t> />3
= t-
It appears that, the nearer we the more likely we are to get to a 5
start to cr
5 ,
(a)
Figure 4
any joint has the effect of closing off the external pipe at that joint.) We
assume that the inflow at any joint must equal the outflow. Thus at the
upper left corner in Fig. 4(a) we find t1 r1 + r 2 while at the lower left = ,
riH
94 Linear Algebra Chap. 2
1 1 o\
-1 0-1 1
0-1 1
0-1 1
The vector equation shows that there is a linear relation between t and r,
/i + r2 =0
— »i — r« + r5 =0
—r 2 + r3 =0
r6 + r< = 0.
A simple check shows that these assignments for r x through r 6 will satisfy the
above system. It follows that there are infinitely many different pipe flows
that will produce zero external flow at each joint. Similarly, if we have
a solution r of the original system (4) for some given vector t on the right
side, then by the superposition principle any of the solutions r can be
added to r to give a new solution r + r; we have
A(t + r) = Ar + At
= +t= t.
I
"
'f(x) dx ^ uf(a - h) -:-
if (a) + wf(a + h)
Ja-h
where it, v, and w are constants. If the formula is to be correct for all
polynomials of degree less than or equal to 2, it must in particular be
correct for the polynomials f (x) = U /i(-v ) = fix) = x 2 Each
x, ar>d .
have
l+
f (x) dx = 2h and fQ (a - h) =f (a) =/„(a + h)
I -A
so u +v+ u' = 2//. Using /^.v) = .r and 2 (.\)
= .y
2
similarly gives the
equations
(a — h)u + av + (a + h)w = 2ah
and
(a 2 - h2 - 2ah)u + a2 v + (a 2 + h2 + 2a/2)tv = 2a 2 /? + f/?
3
.
These equations are easily solved (see Exercise 13) and give the result
u = vv = \h, v = ^h.
We have obtained a rule which is correct for the particular polynomials
f ,fi, and/ Its correctness for any quadratic polynomial follows readily.
2 .
Let us write E(f) for the error committed when the rule is applied to a
general function/, so
a+k
E(f) - \
Ja—h
fW ^ - \hf(a - h) - if (a) - \f{a + h).
It is easy to see that £ is a linear function from the vector space of poly-
nomials to JR.
1
, and of course E(f )
E{f^) E(f2 ) = = = by construction.
If/is any quadratic polynomial, so f(x) px 2 qx = + + r, then/= pf2 +
qf x
-^
f and by linearity
EXERCISES
Figure 1(b) shows an electrical network with the resistance (in ohms) of each
edge marked on Suppose an external power supply maintains node A at
it.
2. The edges and vertices of a 3-dimensional cube form a network with 8 nodes
and 12 edges. Suppose each edge is a wire of resistance ohm, and that two 1
at the other vertices, and the current flowing in the external connections if
(a) the two vertices with external connections are at opposite corners of the
cube
(b) they are at the two ends of an edge
(c) they are at opposite corners of a face of the cube.
3. (a) Suppose that three forces acting at the origin in ft 3 have the same
directions as (1, 0, 0), (1,1, 0), and (1,1, 1). Find magnitudes for the
forces acting in these directions so that the resultant force vector will be
(-1,2,4).
(b) Can any force vector be the resultant of forces acting in directions
specified in part (a)?
5. Ifforcesactin Jl 2 in the directions of (2, 1), (2, 2), and (-3, -1), show that
magnitudes can be assigned so that the system is in equilibrium.
6. Suppose that a random walk traverses the paths shown in Fig. 3(a). What is
7. (a) Suppose that a particle traces a random walk on the paths shown in
Fig. 3(b). Letting pk be the probability of going from b k to b 6 without
going through b b fmd ,
pk for k = 1,2, 3, 4.
(b) How is the result of part (a) modified if b t and the path leading to it are
eliminated altogether?
(c) How is the result of part (a) modified if a new path is introduced between
6 4 and A 6 ?
\0 1 3/
12. (a) Let the external flow vector in Fig. 4(b) be given by t = (1, 1, 2, 4).
Show that there is more than one consistent internal flow vector r, and
find two of them,
(b) Let the external flow vector in Fig. 4(b) be given by t = (1,0, 1, 1).
13. Carry out the solution of the equations for u, v, w given in Example 5
of the text. [Suggestion: begin by subtracting a times the first equation
from the second and a 2 times the first equation from the third.]
a+Sh
f(x) dx = tf(a) + uf{a + h) + vf(a + 2A) + wf(a + 3/0
I
Ja
SECTION 3
In this section we are concerned with formalizing the method of solving THEORY OF LINEAR
linear equations that has been illustrated in the two preceding sections, EQUATIONS
and in proving that it always works. We begin by giving a precise definition
for the "noninterference" property discussed informally after Example 3
3
Linear Algebra Chap. 2
3.1 (a) Every column containing a leading entry is zero except for the
leading entry.
(b) Every leading entry is 1.
1
Sec. 3 Theory of Linear Equations 99
3.3 Theorem
n
way satisfies Ax = b. The /th entry in the product is 2 a ikxk- We
a:=i
and by assumption b is then also zero. If the /th row is not zero,
t
Example 2. Consider
W 1 5 0\
= .12
3 0.
A , bx = ^ ,
b2 =
^0 1,
The third row of A is zero and the third entry of \i x is not; so according to
Theorem 3.3 the equation Ax = b x has no solution. On the other hand,
the third entry of b 2 is zero. The pivotal columns of A are the first, third,
and fifth, with associated rows the second, first, and fourth. The proof
of the theorem shows that, if we construct
100 Linear Algebra Chap. 2
by making the first, third, and fifth entries equal to the second, first, and
fourth entries of b 2 and making the other rows zero, then Ax will be equal
,
The next problem is how to tell whether an equation that has a solution
has more than one. Theorem 3.4 and its corollary show that the question
can be reduced to a special case, and theorem 3.5 deals with this special
case.
3.4 Theorem
Corollary
homogeneous equation.
3.5 Theorem
and
+ s
are solutions of
violated. If the column contains the leading entry r for the /th row,
_1
multiplying the /th row by r will make the leading entry 1. (Since r was
3.6 Theorem
-2
Column 1 does not satisfy 3.1(a), but has a leading entry of in the first 1
1 3 -2 0\
3-2 0\
(1
1
Adding 2 times row 2 to row 1 and (—2) times row 2 to row 3 clears out
the other entries in column 3 to give
1
104 Linear Algebra Chap. 2
3.7 Theorem
3.8 Theorem
We thus see that it is sensible to apply the method used for computing
method fails, then one obtains a reduced matrix with at least one non-
pivotal column and, hence, by Theorem 3.5, a nonzero x such that Ax = 0,
which demonstrates that A is not invertible.
We conclude by observing that every elementary operation, when
applied to a column vector in %
n
is a linear function, called an elementary
,
on the left by some n-by-n elementary matrix. (The precise forms of such
matrices are described in Exercise 7 at the end of this section.) This fact
enables us to prove the next theorem.
3.9 Theorem
f(x) = Ax
can be expressed as a composition of elementary transformations.
EXERCISES
1. Determine which of the following matrices are reduced. For those that are
not, state exactly how they violate the conditions. For those that are
reduced, list the pivotal columns and their associated rows.
1
Linear Algebra Chap. 2
3. Let
5. Show that if a square «-by-« matrix is reduced and has no all-zero row, then
every row and column contains n — 1 zeros and one 1. Hence show that it
(a) Denote by D t
(r) a matrix which is the same as the identity matrix
except for having r in place of 1 in the /th diagonal entry. Show that the
Sec. 3 Theory of Linear Equations 107
matrix product
D (r)M
f
(b) Denote a matrix with 1 for its //th entry and 0's elsewhere by Eu . For
example, for 3-by-3 matrices,
(I + rEvT1 = (/ - rE ),
i} D&T X
= />,(;] ,
and Tj = Tu .
8. Prove that if there are more unknowns than there are equations in a linear
system, then the system has either no solutions or infinitely many.
are any n + 1 numbers, then there is exactly one polynomial of degree < n
such that/?O ) = b ,p(x n ) = b n [Hint. Show that the problem leads
, . . . .
11. A reduced matrix in which the first pivotal column (starting from the left)
is associated with the first row, the second pivotal column is associated with
SECTION 4
VECTOR SPACES, In the earlier parts of the book we have restricted ourselves to vectors in
SUBSPACES, n
'Ji . In this section we consider more general vector spaces, though some
DIMENSION
of the ideas have already been introduced with ii\
2
and Jl 3 as the main
examples.
Recall that a vector x is a linear combination of the vectors xXl . . . , xn
if there are numbers rlt . . . ,r„ such that
x = r1 x 1 + ..'.+ rn \ n .
Example 1. Let
l
= (1,0,0), x2 = (0,1,0), x3 = (0, 0, 1), k = 0,1,1).
Then y = (2, 2, 0) is a linear combination of x 1 and x 2 because it is
s; so y = rx + sx is impossible. 2 3
If a set of vectors lies in a plane through the origin in 3-space, then every
linear combination of them lies in the same plane —
x and y recall that if
are in a plane through the origin, the parallelogram rule makes x + y
a vector in the same plane. Any numerical multiple of x lies in the same
plane because it lies in the line containing x. Any linear combination of
x 1} . . . up by multiplications and additions, and if the vectors
, x„ is built
Xj, x„ lie in a plane, so do all linear combinations of them.
. . . ,
Similarly, if x x x n all lie in one line through the origin (so they
are all multiples of some one vector), any linear combination of them
lies in the same line.
(a) The set of all vectors in %n with first entry equal to is a subspace,
since any linear combination of such vectors will also have a for its first
entry.
.
(b)
C
For any vector space U,
C
U itself is a subspace. The term proper
subspace is often used to refer to those subspaces that are not the whole
space. In any vector space the zero vector forms a subspace all by itself.
(c) For any linear function 17 — > ID, the set JC of vectors x in U
with/(x) = is a subspace of T) called the null space off. If x l5 . . . , x fc
k k
are in JV and x = 2 r t x i> tnen /( x ) = 2 r if( x i) = 0> because / is
linear and all the/(x,) are 0. Hence x is also in JV, and so JV is a subspace
C
of \J. In particular, the set of vectors (x, y, z) in 'A 3 such that
x + 2y + 3z = 0,
or equivalently,
(1
3
is a subspace because it is the null space of the linear function from 'Ji to
Jl defined by the preceding l-by-3 matrix.
the range off means that there are vectors \ lf x k in the domain of / . . . ,
=f(r1x 1 + . . + rk x k
. ).
5 =x
)0
is a subspace because it is the range of the linear function just defined by
the above 3-by-2 matrix.
n
vectors x x x which are in the span of S, that x = £ s a u 3 f° r
, . . . , fc ,
kin \
is,
n
t
3=1
some vectors u_, in S. Then x = ^ rA ^= s^uA = £ ^, where /
;
=
k = i l \ 3 1 / = 3 1
1 rx+ sx = (r + s)x.
2. rx + ry = r(x -f y).
3. r(sx) = (rs)x.
4. x + y = y + x.
5. X + y) + z = x + (y +
( z).
4.1 Theorem
c
Example 3. (a) Let l) consist of all continuous real-valued functions
of a Define/ + g and rf in the obvious way as the functions
real variable.
whose values for any number x are/(x) + g(x) and rf(x), respectively.
(Of course, we are using the theorems that/+ g and //are continuous if
Sec. 4 Vector Spaces, Subspaces, Dimension 111
/ and g are.) It is easy to verify that the laws for a vector space are
satisfied.
C
(b) Let P be the subspace of U consisting of all polynomials, i.e., all
of P.
The description of lines, planes, and ordinary space as 1-, 2-, and 3-
Example 4. (a) The four vectors x = (2, 0, 0), x = (0, —2, 0), 1 2
rx + sx + rx =
x only if r — s = = 0.
2 3 t
for its z'th entry and for all other entries, form a basis for 3\". Verification
that the e, are in fact linearly independent and span Jl" is left to the reader.
112 Linear Algebra Chap. 2
4.2 Theorem
appear on the right side. By substituting the right side for x A in any .
4.3 Theorem
any two vectors y x and y 2 in the space, one must be a multiple of the
Sec. 4 Vector Spaces, Subspaces, Dimension 113
that the y's are dependent, then we will have shown that the state-
ment of the theorem holds for a spanning set of n elements, and the
inductive proof will be complete. Each of the vectors yu y n+1 . . .
,
n + 1 • If the /i + 1 numbers a a are all zero, then the y's all lie in the
space spanned by the /; — vectors x 2
1 x w thus by the inductive , . . . , ;
j =2
But since not all the r's are zero, this implies that y x , . . .
, y n+1 are
dependent, as we wanted to show.
4.4 Theorem
both sets are independent, and both are spanning sets, Theorem 4.3
implies that k < n and n < k.
The dimension of a vector space that has a finite spanning set is the
number of elements in any basis for the space. (The dimension of the
space consisting of the zero vector alone is defined to be 0.) We write
dim (1)) for the dimension of the vector space C U. Note that Theorem 4.2
114 Linear Algebra Chap. 2
guarantees the existence of a basis, and that Theorem 4.4 guarantees that
the dimension does not depend on which basis is taken.
(c) The space P of all polynomials (Example 4(b)) does not have any
finite spanning set. If it did have one with k elements, then the fact that
1, x, x2 , . . .x k are k -\-
, 1 linearly independent elements of P would
contradict Theorem 4.3.
dimensional.
Theorem 4.2 asserts that we can get a basis from a finite spanning set by
deletingsome of its members. The next theorem shows that, in a finite-
dimensional space, we can get a basis from a linearly independent set by
adding vectors to it.
4.5 Theorem
must show that all the r's are 0. If rk+l were not 0, we could write
x A.^ x = — (/'i//'i + i)x 1 — ... — (rk lrk+1 )x k which is impossible because
. ,
4.6 Theorem
4.7 Theorem
u l5 . . . (Theorem
, ur for 17. 4.6 guarantees the possibility of this
construction.) Then dim (JV) = k and dim (17) =k+ r. Let wx =
/(Ux), . . . , w =/(u
r r ).
We claim that wl9 . . . , wr is a basis for W,
which implies that dim (ID) = r and proves the theorem. It is
k r
x = 2 a y + 2 Mi. t i
and then
y = I aj(jd + 2
=
t' l = Z l
WW = + 2 6,-w,,
2 =1
i=l
fljVx + • • • + ak y k — ^i u i — • • — b Tu r = 0. Since v l5 . . . \k u 1, ,
. .
.
u r are linearly independent, this is possible only if all the 6's (and all
the a's) are zero, which shows that the w's are linearly independent.
^+ b,
where C
U is a linear subspace and b is a single vector. The dimension of an
c
affine subspace is of course defined to be the dimension of Vf, and a
1-dimensional affine subspace is usually called aline, while a 2-dimensional
3
affine subspace is called a plane. Examples in ,'K are shown in Fig. 5(b).
Two affine subspaces are parallel if they are translates of one another.
Two of the most important ways of describing subspaces, namely, as
range or null space of a linear function, provide standard ways of describ-
ing lines and planes. The parametric representation of planes discussed in
Sec. 4 Vector Spaces, Subspaces, Dimension 1 17
L(u, v) — ux r + vx 2
A(u, v) = ux 1 + vx 2 + x
L(x - x ) = 0.
Methods for solving such equations are discussed in the earlier sections
of this chapter. If x = (x, y) is 2-dimensional then equation (1) would
take the form
ax + by = c,
a xx + b xy+ cx z = dx
a.,x + b y +
2 c 2z = d2
x — y + 2z = 1
x +y + 3z =
118 Linear Algebra Chap. 2
2x + 5z = 1
x +y + 3z = 0.
Multiplying the first equation by I and subtracting from the second gives
x + \z = J
We can represent the set of all solutions of this pair of equations para-
metrically as an affine subspace by setting z — t. We find
X = + 2"/ 2
/V — — It
2»
— A2
Z = /
EXERCISES
1. Which of the following subsets of Jl 3 are subspaces ? In each case either show
that the subset is a subspace or find some linear combination of elements of
the subset that is not in the subset.
2. Let x x = (1, 2, 3), x2 = (-1, 2, 1), x3 = (1, 1, 1), and x 4 = (1, 1, 0).
5. Part (d) shows that the union of two subspaces is not always
of Exercise 1
a subspace. Show that the union of two subspaces is a subspace if and only
if one of them is contained in the other.
6. Show that the range of a linear function may be a proper subspace of its
range space.
8. For any two subsets A and 3i of a vector space, let .-t — 3i be the set of all
9. Show that if is the only element in the intersection of two subspaces S, 73,
then
dim (S + TJ) = dim (S) + dim (73)
[Hint. Show that a basis for S together with a basis for 13 gives a basis for
s + t;.]
[Hint. Start with a basis for S n 13 and extend it (Theorem 4.6) to a basis
for S and a basis for 13.]
1
1 -3 2
Find a basis for the null space of/, and one for the range of/. Verify that
Theorem 4.7 holds.
12. Describe the solution set of each of the equations or systems of equations
below as an affine subspace, that is, as a translate by a specified vector of a
linear subspace with a specified basis.
(a) x +y = 1.
(b) 2jc + y =1
2x - 3y + z = 2.
(c) x + 2y + 3z = 10
Ax + 5y + 6z = 1
(a) x + 2y
(b)
(c)
SECTION 5
LINEAR FUNCTIONS In Chapter 1 we saw that matrices obey some of the same rules for
addition and multiplication that ordinary numbers do. Furthermore,
we have seen that given a linear function /from :K" to %m , there is an
m-by-/7 matrix A such that/(x) = Ax for all x in J{". Using these facts
we could prove that linear functions from 'A" to %m obey algebraic rules
just as their matrices do. However, it turns out that these same rules
apply to a wider class of linear functions and not just those representable
by matrices. For this reason we shall prove the rules in general form and
then apply them in Section 6 to a systematic analysis of some linear
differential equations.
We
begin by describing the operations of addition, scalar multiplica-
tion,and composition of functions. Let / and g be functions with the
same domain and having the same vector space as range space. Then the
function/ + g is the sum of/ and g defined by
(f+g)(x)=f(x) + g{x)
Sec. 5 Linear Functions 121
r/(x) = r(/(x)).
We have already defined the composition of functions in Chapter 1,
but we repeat the definition here. We require now that the range of/ be
contained in the domain of g. Then g°f is the composition of/ and g
and is defined by
/ g
Example 1. Suppose 5l 2 > % 2
and .'R
2
3t 2 are given by
and
'Q-C-K -3C
x\ _ /2x+y\ 12 \\lx
v+ O= / +
3x + 2y\ /3 2\/x
2x + 2j/ ~\2 2/V
Also, 3/ is given by
'•'CM'C
; x -X
.
5.1 Theorem
(f + g)(x)=f(x) + g(x)
= Ax + Bx = (A + B)x.
The first equality holds by the definition of/+ g, the second by the
relation of matrix to function, and the last by the general rule
AC + BC — (A + B)C for matrix multiplication and addition.
To prove statement 2 we write, similarly,
where the last equality follows from the rule B(AC) = (BA)C for
matrix multiplication.
Statement 3 is just the result of Theorem 4.5, Chapter 1.
'Q-Q- 'C)-t:
;:)-(:) 'Ch:
Then the matrices of/ and g are
2 -1\ /l
and
3 0/ \0 -1
Sec. 5 Linear Functions 123
2 -«_/, 0j_/0 -,
3 0/ \0 -1/ \3 2
^2 -1\/1 0\ /2 1
_
,3 0/\0 -lj \3
5.2 Theorem
(1) (f + g)°h=foh+goh
(2)fo(g + h)=fog + fo h
(3) (rf)og = r(fog)=fo(rg)
(4) ho (gof) = (hog) of.
ho(gof)(x) = h(gof(x))
= Kg(A*))Y
Similarly,
(hog)of(x)=hog(f(x))
= h(g(f(*)))-
124 Linear Algebra Chap. 2
The results of the two computations are the same, so the formulas we
started with must be the same. Since x is arbitrary, h ° (g of) =
(h o g) of.
(0O)
Example 3. Let C (Jl) be the vector space of infinitely often differen-
tiable real-valued functions y of the real variable x. If we let D stand for
differentiation with respect to x, then
Dy=y',
and D is a function with domain C <oo)
(3l). The familiar rules about
differentiation stating that (/ + g)' =/' + g' and (cf)' = cf imply that D
is linear. Because the meaning will always be clear, we can omit the little
D y=y",
2
D*y=y'".
We can even define
D°y =y
for occasional convenience. Combining powers of D by addition and
numerical multiplication leads to formulas like
D -D-
2
2, (D + 1)(Z> - 2),
and applying the rules (1) and (2) of Theorem 5.2 we get
(D + l)(D - 2) =D -D- 2
2.
for everyx in the domain of/. Applying/ to both sides of this equation
gives/°/ _1 (/(x)) =/(x); therefore, we also have
/o/-l(y) = y,
for every y =/(x) in the image of/. We leave as an exercise the proof
that:
Sec. 5 Linear Functions 125
x
5.3 If/ is linear, then so is/ .
Finally, the row reduction method of the first section of this chapter is
an efficient way to find the inverse of an invertible matrix, that is, to find
A -1 such that
A- 1 A = AA^ = 1
/,
f(x) = Ax,
-1
always has an inverse function/ given by
defined by
fx\
g(x)
126 Linear Algebra Chap. 2
)= x, for in C.
w
j
which is not invertible because it isn't even a square matrix. The difficulty
_1 2
is that g is not defined on all of ,'il , but only on the subspace C.
have a name: the rank of a linear function/is the dimension of the range of
/. Thus the linear function
f =
\y) \0 3/\
(2x\
2
has its rank equal to 2 because its range is all of 'J\ . The function
1 0\
\o oj\j
2
has its rank equal to 1 because its range is just the x-axis in Si For a .
useful criterion.
5.6 Theorem
Let/be a linear function with matrix A. Then the rank of/is equal
to the number of linearly independent columns in A.
1 7^
,2 14/
Then subtracting 2 times the first row from the third row gives
1 7\
15.
v
0/
5.7 Theorem
5.8 Theorem
Let %n —>% n
be linear, with square matrix A. Then/ has an
inverse if and only if A is invertible.
ABx =/o/-i(x) = x
EXERCISES
1. Suppose that/ and g are linear functions such that
1\ /1\ /0\ /l
'o-i- A.
y-n -CHI
Sec. 5 Linear Functions 129
(a)/.
(bU-
(c)f + g.
(£)2f-g.
(e)<?o/.
(Ofo(f + g).
Linear Algebra Chap. 2
(0 1 1 2
(a) (b)
U 0) 2 4y
<o o : 2
(c) | 1
I
2
,1 Oy v4
9. Let °U and '10 be finite dimensional vector spaces and let/be a linear function
with domain 1J and with range equal to 10. (Thus dim (1D) = rank (/).)
(a) Show that V contains a largest subspace S such that/, when restricted to
S, becomes one-to-one from StoW. [Hint. Let 91 be the null space of/
with basis n 1 n k Extend this basis to a basis for 1) by adding
, . . . , .
vectors s l5 . . . , s t .]
(b) Show that dim (S) = rank (/), where S is the subspace of part (a).
10. If/ is a function and y an element of the image of/, then the subset S of
is
11. What is the simplest way to find the rank of a diagonal matrix
diag («!,.. .,«„)?
SECTION 6
DIFFERENTIAL In this section we look at some vector space ideas that arise in studying
OPERATORS differential equations. The equations we shall treat will be like the
following:
/ - ly = (1)
y — 7>y = ex (2)
y = ce rx , (5)
because the c will cancel on both sides. In fact, Equation (5) gives the
most general solution to (4), for observe that we can write (5) in the form
e~ rx
v — c.
(e- rxy)' =
or, using the product rule for derivatives,
rewritten. But now we can reverse these steps, supposing that y is some
solution. We start with
y - ry =
rx
and then multiply by e" to get
(e-ry)' = 0.
e~ TX
y = c,
y - 3y = ex ,
£-3x,/ 2e~ 3x v = e~ 2x .
(e
-3Xy)' — e ~2x
for the most general solution. It is easy to verify directly, of course, that
we have indeed found some solutions, one for each value of c. What we
have shown additionally is that any solution must be of the form — \ex +
ce 3x .
(D + 2)y = Dy + 2y
= / + 2y
(D 2 - \)y = D' y -y z
= D(Dy) -y=y"-y.
An is that D acts as a linear function on y;
important observation
the term linear operator sometimes used to avoid possible confusion
is
over the fact that y itself is a function of x (though not necessarily a linear
one). To see that D acts linearly, all we have to do is recall the familiar
Sec. 6 Differential Operators 133
properties of differentiation:
D{)\ + y = Dy, + Dy
s) 2
These two equations express the linearity of D. From the fact that
compositions of linear functions are linear it follows that the operators
D D32
, , and in general Dn are also linear. Because numerical multiplica-
tion is a linear operation and because the sum of linear operations is
D +
2
a, D + aD +
2
b, (D + s)(D +
are all linear operators, with the respective interpretations
+ a)y = y" + ay
(D 2
(D + s){D + t) =D + 2
(s + t)D + st,
Z) a - = (D-
1 \)(D + 1)
(D 2 + aD + %= 0.
134 Linear Algebra Chap. 2
Our method of solution will be to try to factor the operator into factors
of the form (D + s) and (D + /), and then apply the exponential multiplier
method of Examples and 2 repeatedly.
1
(Z)
2
+ 5D + 6)y = 0.
(D 2 + 5D + 6) = (D + 3)(D + 2);
(D + 2)y = u
for the moment, we substitute u into the previous equation and arrive at
(D + 3)w = 0.
Therefore
e 3x u = cx ,
for some constant clf and so
u = c x e~~ 3x .
Recall now that we have temporarily set (D + 2)y = it. We then have
(D + 2)y = c x e~
3x
.
tx
Multiply this last equation by e to get
e^Dy + 2e 2xy = c x e~ x
or
D(e 2xy) = c x e~ x .
e^y = —c x e~ x + c2
or
y = — c e~ 3x + x c 2 e~ 2x .
Sec. 6 Differential Operators 135
Since the constants c x and c 2 are arbitrary anyway, we can change the
sign on the first one to get
y = c t e~ 3x + c^er 21
e
ax
(D + a)y = D(e axy).
6.1 Theorem
(D - r,){D - r2) . . . (D - r n )y = 0,
(D - ri )(D - r 2 )y = 0,
we set
(D - r 2 )y = u.
e~ riX Du — r x e~ T ^ x u =
or
D{e~ rix u) = 0.
e -TiX U _ Ci
Now
(D - r 2 )y = cx e ^x
r
.
136 Linear Algebra Chap. 2
We multiply by e r- r
to get
D{e- r **y) = Cl e
(r >- r * )x
(6)
gives
Cl
e-*z*y = gin-*)*
+ C2
cl
y = e
r
r, — r..
D{e~ rt*y) = ev
Now integration gives
g-rtxy = CiX _|_
Ca
or
y = c 1 xe TiX + c 2 e T2X ,
(D - rjiij = 0.
Substitution of the general solution for u^ into (7) gives a new equa-
tion which we split up by setting
(D-r )...(D- 3 r n )y = u2 .
(D — r 2 )u 2 = «l9
D n + a^D"- + 1
. . . + ax D+ a (8)
y
(n)
+ a n_ iy ^-v + . . . + a iy '
+ a y = 0.
Notice that, to get the polynomial from the equation, we replace y
ik)
by D k
and that the term a D° = a Finding a
y corresponds to a . factor-
ization for the polynomial depends on knowing its roots, for if the
polynomial (8) has roots r u r n then it has the factored form
. . . , ,
(D - ri )(D -r )...(D- 2 r n ).
Z> 3 - 4D + 2
AD,
D(D 2 - AD + 4) = D(D - 2)
2
The roots are and 2, where 2 is a repeated root. Thus the general
solution to the equation is a linear combination of e 0x , e 2x and xe Zx
, ,
Linear Algebra Chap. 2
and so
y = cx + c 2 e 2x + c 3 xe 2*
(D - rO . . . (D - r n )y =
is a vector subspace of C (n)
, because JV is the null space of the linear
operator
L = (D - rx) . . . (D - rn)
6.2 Theorem
(D-r )...(D- y r n )y = 0.
differential equations
{D - rk )y = uk_ x
D(e~ TkXy) =
are of the form
e -r k Xy _ Cf
Example 5. Given
y" + 2y' +y = e 3x ,
140 Linear Algebra Chap. 2
(D + 1)« = e Zx .
Multiplication by e x gives
e x Du + ex u = e ix
or
D(e*w) = e 4x .
or
u = £e 3x + c^ - *.
Since (Z) + l)_y = w, we have
(D + 1)7 = ie
3x
+ c^.
Again multiplying by ex , we get
e x Dy + e xy = le
ix
+ cx
or
Then
or
j = ^e * + 3
c x xe~ x + c 2 e _x .
In the above example the solution breaks naturally into a sum of two
parts, y h and y9 :
yh = c x xe~ x + c 2 e~ x ,
L(y) =
associated with L(y) = f. The function _y„ is called a particular solution of
Uy) =f
because it is just that: a particular solution, though not the most general
one. In fact, we get y p by setting c l = c2 — in the general solution. This
breakup of the solution into two parts is an example of a general fact about
linear functions, a fact that is used in solving systems of linear algebraic
equations. The principle is important enough, and at the same time simple
enough, that we state it here also.
Sec. 6 Differential Operators 141
6.3 Theorem
Proof. Suppose that L(y) =/and that also L(y p ) = f. Then, since
L is linear,
L(y - yp = ) L(y) - L(y p )
The method of Example 5 can always be used to find the most general
solution to the equation L{y) =/of the form
y = Oi7i + a 2y 2
equation.
.
(D + \)
2
y = e 3x
had the general solution
TVe
3x
+ Cyxe- X + c 2 e~ x .
we would not have to start all over again, but would only have to find a
particular solution for
(D + \fy = 1.
(D + \fy = e 3x + e~*
(D + \fy = e~ x .
EXERCISES
1. For each of the following differential equations, find an appropriate
exponential multiplier and then solve by integrating both sides of the
modified equation
e $p(x) dx
d
(
e lv(x) dx 7
y\
dx
Sec. 6 Differential Operators 143
3. Use the result of Problem 2 to find an exponential multiplier for each of the
following differential equations. Then solve the equation.
(b)
£ + xy = x. (d) / + y = 0.
4. Write each of the following differential equations in operator form, e.g.,
(D 2 + 2D t 1)/ = 0, and then factor the operator into factors of order 1.
5. Sketch the graph of each function of x given below. Then find a differential
equation of which each one is a solution, the equation being of the form
y" + ay' + by = 0.
(b) e* + e- x . (d) 2e
2x - 3e
3x
.
x y"2
+ (x 3 + x)y + (x
2
- \)y =
can be written in operator form as
x)y-0.
K)< \(D
y(x) = cx e
tx
+ c 2 e~
tx
y(x) =d 1
cos x + d2 sin x.
(c) Verify directly that cos x and sin x are solutions of v" +y = 0.
(D - ri)...(D -r„)/=0.
By applying the operators (Z) — r,) to both sides of the given Equation (1),
it follows that y must also be a solution of the higher-order homogeneous
equation
(Z> - /i) . . . (D - r n )(D
2
+ aD + b)y = 0. (2)
Since we have a routine for writing down the most general solution yg of
this last equation, we can find a particular solution y p of Equation (1) by
substituting the general solution of (2) into it and seeing what conditions
save duplication, we can first eliminate from yg all terms that already occur
in the homogeneous part, yh of the solution of (1). Linear independence of
,
for y v The
. general solution of (1) is then y = yh + y v Find . the general
solutions of the following differential equations.
x
(c) y" + 2/ + y = e .
yh = C^iO) + C 2 U 2 (x).
(a) Verify that substitution of Equations (4), (6), and (7) of Exercise 11
into
/ I ay' +by =f
produces Equation (8).
(b) Verify that Equations (5) and (8) of Exercise 1 1 have the solution given
by Equation (9).
x(t) and its acceleration is x(t). From physical principles it can be shown
that, theoretically, x satisfies the differential equation
mx + kx = 0,
(a) Show that if the initial displacement is x(0) = and the initial velocity
Vcfn
x(t) = -7- (1 - e- kt ' m ).
k
(b) Show that the displacement x(t) has an upper bound equal to v^mjk.
What is the effect of increasing the viscosity constant or of increasing
the mass of the pellet?
(c) Show that the velocity of the pellet decreases to zero as t increases.
(d) Show that the acceleration of the pellet is negative for t > 0. What is
ay + pz =/(/)
yy + 6z = g(t)
as a special case when a, b, c, and d are all zero. The method of elimination
can be used to solve the differential system as well as the algebraic. To find
y{t) and z(t), operate on both sides of the first equation by (dD + d),and
on the second equation by (bD + ji). Then subtract one equation from the
other. The resulting equation can be solved for y(t) since it does not contain
z. Next substitute the general solution y(t) into one of the equations and
solve that for z(t); to determine possible relations between the constants of
integration, it may be necessary to substitute y(t) and z(t) into the other
given equation. Sometimes simplifications in this procedure can be made.
Sec. 7 Complex Vector Spaces 147
(a) Use the method just described to find the general solution of the system
(£> + \)y + z = 0.
3y + (D - \)z = 0.
(b) Determine the constants in the solution of part (a) so that the initial
(Z> + \)y + Dz =
Dy - (D - \)z = t.
(d) Determine the constants in the solution of part (c) so that the initial
15. Suppose that two 100-gallon tanks of salt solution have concentrations (in
pounds per gallon) of salt y(t) and z(t), respectively. Suppose that solution
is flowing from the v-tank to the z-tank at a rate of 1 gallon per minute, and
from the z-tank to the j-tank at a rate of 4 gallons per minute, and that the
overflow from the j'-tank goes down the drain, while the z-tank is kept full
by the addition of fresh water. We assume that each tank is kept thoroughly
mixed at all times.
(a) Show that y and z satisfy a system of differential equations of the type
discussed in Exercise 14. [Hint. Express Dy and Dz each as a linear
combination of y and z.]
(b) Find the general solution of the system found in part (a), and then
determine the constants in it to be consistent with initial concentrations
j(0) = ^ and z(0) = i,
(c) Draw the graphs of the particular solutions y{t) and z(t) found in part
SECTION 7
In the earlier parts of this chapter we have always understood the vector COMPLEX VECTOR
space operation of numerical multiplication to mean multiplication by a SPACES
real number. However, we can replace real numbers by complex numbers,
and let the definition of vector space and linear function remain otherwise
the same. Then, all the theorems we have proved for real vector spaces
are still true relative to the complex numbers. To prove this, all we have
to do is observe that the only properties of real numbers that we used in
proving theorems about an abstract vector space are properties that are
shared by complex numbers. Theorems involving inner products are
another matter, which we shall discuss at the end of the section. As
motivation for considering complex vector spaces, we shall explain how
the extensionis the key to further development of the study of differential
(a- •
x') i (v y')
IV
Figure 7
y '.l'-
\x + iy\ = v/x
z
+ y
2
X r iy
and corresponds to the length of a vector in 'Ji
2
. (See Fig. 8.)
are related by
y
x + iy = \x <y\ + i
x + iy\ \x + iy\)'
iy\
Because \x + iy\ = V* + y 2 2
, the numbers xj\Jx 2 +y 2
and yjs/x 2 +y 2
r|(cos i sin 8)
Figure 9
number has infinitely many polar angles, each pair differing by an integer
multiple of 277.
Now if z and z' are complex numbers with polar angles 6 and 0', we
can write their product in polar form as follows:
= \z\ \z'\ ((cos 6 cos 6' — sin 6 sin 6') + / (cos 6 sin d' + sin 6 cos 6'))
In the last step we have used the addition formulas for cosine and
sine. The result of the computation is a number in polar form
having absolute value \z\ \z'\ and polar angle d + 6'. We conclude
that the absolute value of a product of complex numbers z and z'
is the product of their absolute values
and that if z and z' have polar angles 0(z) and O(z'), then zz'
has a polar angle 0(z) + d(z'). These facts are illustrated in
Fig. 10. Figure 10
150 Linear Algebra Chap. 2
2
Thus e'° is a complex number with |<"°| \ sin cos- I , and with
polar angle d. Since polar angles are added when complex numbers are
multiplied, we have
1 = a- = 7°.
These last equations are justifications for using the exponential notation;
the function behaves like the real exponential, for which e°e° = e
6+e '
z = \z\e
mz
\
and its conjugate is
2 = \z\e~
mz \
ax ax clx
and
|
(Ma) - iv(x))dx = |
m(x) clx + i fi-(A) dx.
—d e
ix
= —d (cos x +
/ , •
i sin x)
dx dx
= —sin x f i cos x
= /(cos x + / sin x) = ie
tx
.
Sec. 7 Complex Vector Spaces 151
In short, we have
4- e
ix
= ie
ix
.
dx
Similarly
e
ix
dx = - e ix + c,
J i
and compute
7.1 A. e
<a+ifi)x
= (a + ip)e
la+ifi)x
dx
and
7.2 r e <«+«-/»* rf x = —— 1
e
<«-M/»*
+ Cj a + i/5 ^ 0.
J a + //5
(D 2 + aD + b)y =
when the factored operator
(D - ri )(D - r2) = i)
2
- (r, + r 2)D + r^
contains complex numbers rx and r2 . We shall see that the usual tech-
(£»
2
+ \)y = 0,
and factor D + 2
1 to get
(D - i)(D + i)y = 0.
152 Linear Algebra Chap. 2
Then set
(D + i)y = u, (1)
and try to solve
(D - i)u =
for //. As in the real case, we multiply by a factor designed to make the
left side the derivative of a product. The same multiplier rule suggests
that the correct factor is e~ ix so we write
,
er ix {D - 0" =
or, since D(e~ ix
u) = e~ ix
{D — i)u,
D(e~ ix u) = 0.
e~ ix
u = cx or u — cx eix .
(D + i)y = cxe*,
Integrating gives
e
ix
y = \ Cl e 2ix + c2
or
y =~ cx e
ix
+ c 2 e~
ix
.
2i
y = c x e ix + c 2 e~ ix
d +
=— x id*
= —d,
— id 2
c, and Co .
Sec. 7 Complex Vector Spaces 153
c1 x 1 + . . . + c nx n =
implies that all the c's are zero whenever they are chosen from the complex
numbers, then the same implication certainly holds if the c's are only
chosen from the real numbers. However, the converse statement is not
true, as the following example shows.
7.3 Theorem
Theorem 7.3 is not usually applied directly in the above form because
operators such as
P(D) = D + aD + 2
b,
which occur in practice, most often have real numbers for the coefficients
a and b. This implies that, in the factorization
D + aD +
2
b = (D - r{){D - r 2 ),
qTX qTX
Cl e
rx
+ c2 e = c e (a+mx + c e ~
fx
x 2
(x ill)x
e ax sin /3x are easier to interpret geometrically than are the complex
exponentials that gave rise to them. Hence the solutions are often written
using the trigonometric form
y = dx e™ cos j8x + d 2 e*
x
sin /3x.
y" + 2/ + 2y =
has characteristic polynomial
D + 2D +
2
2.
(2>_(_1_/))(2)_(_1 +f))y = 0.
The complex exponential solutions are
e (-l-i)x £ {-l+i)x_
(-i-i)x {
- 1+i)x
Cie _|_ c 2e
or
dx e~ x cos x + d.2 e~
x sin x.
7.4 Theorem
{D n + a n _ x D n -i + . . . +aD+ x a )j = 0,
where the # are real, is a vector space JL of dimension n relative to
fc
l ax l ax
x e cos /5.x, x e sin /?x,
where a and ft
are real.
Proof. We know from Theorem 7.3 that the complex solutions of the
above differential equation are linear combinations of functions of
the form x lerx where either r is real or else there is a companion
,
solution x e
rx
We have seen that any solution, and in particular
l
.
l x l ax
x e" cos fix, x e sin fix. (3)
c + cx x + . . . + c nx
n
— for all x
implies that the polynomial has more than n roots. This is possible only
if all the coefficients are zero.
|cx|
2
= (ex, ex) = c 2 (x, x) = c 2 |x| 2 .
But |cx| 2 and |x| 2 are both real numbers, while in general c 2 the square of a ,
complex number, is not real. To get around this difficulty we require that
(x, y) be conjugate symmetric,
then |z| turns out to have the three properties of length listed in 5.3 of
Section 5, Chapter 1.
P(x) = 2c ke
ik *
k=-n
defined for— n < x < tt with complex coefficients ck . We can define a
complex inner product on 8„ by
Ipl = (P, Pf
2
\l/2
2
\p(x)\ dxj .
real case and is left as Exercise at the end of this section. The differ-
1 1
ence between the real and complex proofs here illustrates the fact that,
because a complex inner product has somewhat different properties from
a real inner product, we cannot expect theorems involving inner products
to extend without change from real to complex vector spaces. Complex
inner products are used in this chapter in the exercises following Sections
8 and 9.
EXERCISES
1. For each of the following complex numbers z, find z and \z\. Then write z
in polar form and find 1/z.
2 + /
(b) -1 + 2/. (d)
i
2. Verify that, for complex numbers zlt z 2 , and z 3 the distributive law ,
z-Sztf^ = {z x z 2 )z z and zx + (z 2 + z 3) = (z 2 + z2 ) + z3 ,
all hold.
4. Prove directly from the definitions of conjugate and absolute values that
z^ = z^ and \z z = x t \
\z x \
|z 2 |.
to k\
fc
e
into two infinite series, and use the result to justify defining e' by
e
tB
= cos 6 + i sin d.
(a) (£>
2
+ 1)7 = 1. (d) y' +y = \.
mx + kx + hx = 0,
(b) Assuming m, h, and k all equal to 1 , find the solution of the differential
equation satisfying x(0) = and x(0) = 3.
(c) If in part (b) we change to k =2, but leave the other conditions the
same, find the corresponding solution to the differential equation.
(d) What is the maximum displacement from the initial position under the
conditions of part (b)? Show that the oscillation tends to zero as t
increases.
(e) Sketch the graph of the displacement function under the conditions of
part (c). What is the maximum displacement and at what time does it
occur?
9. (a) Show that {a cos ex + b sin ex), where a, b, and c are real numbers,
can be written in the form r cos (ex — 0), where r = Va 2 + b 2 and
is an angle such that cos d = a\r and sin Q = bjr.
(b) The result of part (a) is useful because it shows that a linear combination
of cos ex and sin ex has a graph which is the graph of cos ex shifted by a
suitable phase angle, 8, and multiplied by an amplitude, r. Sketch the
graph of
1 1
- cos 2x -i 7= sin 2x
2 V3
by first finding r and an appropriate 0.
ax x
10. Show directly that e cos Px and e' sin fix are linearly independent
relative to the real numbers by using the formulas
gift*
+ e -iP*
i0x _e -ifix
rx TX
together with the fact that e and e are linearly independent relative to the
complex numbers when r j= f.
|<z,w>|<|z||w|
:
so that
(e-
i0
z, w) - |<z,w)|.
12. Show that C", the set of w-tuples of complex numbers, has dimension //
relative to the complex numbers and dimension 2a? relative to the real
numbers.
SECTION 8
ORTHONORMAL BASES If 1) is a vector space with an inner product, and T) has a basis
Ui,U 2 , . . . , u„,
then it is often desirable that the basis be an orthonormal set. This means
that distinct vectors u (
and u ; are perpendicular:
u. -u ; = 0, i^j, (1)
X = CjUj + • • + CjU; + • • • + c nu n ,
X • U; = CiUj • IT, + . . . + Ci U j • Uj + . . . + c n \x n u;
= + • • + c, + . . . + 0,
so that
C, = X • u, (3)
2
Example 1. The vectors (1,0) and (0, 1) in 'J\ form an orthonormal
basis. So does the pair of vectors (l/\'2, l/\/2) and (— 1/V2, 1/V2). In
terms of the latter basis we can write
(x y) '
= c +c
i-Jrji} 'hr2-jl}
s/2 J2i
-
*-*»>(*# fe a
5 -
(1 ' 2) = + 72'^'72)'
7i(vi'^)
Example 2. The set of functions defined for — <x <
tt tt by
n
T(x) = 2 (o n cos kx + foj. sin /ex).
(/g)^ 1 fV(x)g(x)Jx,
77 J-s-
and with respect to this inner product, the above set turns out to be
orthonormal. In fact, computation shows that
o,
- I cos kx cos Ix dx =
— I sin kx sin Ix dx
Linear Algebra Chap. 2
The fact that orthonormal bases are simpler to work with suggests
the following question: Given a basis for a vector space with an inner
product, is there some way to find an orthonormal basis? The answer is
8.1 Theorem
CiUj + . . . + cm + . . . + cnu n = 0.
111 = —
|Xx|
.
Then lu^ = Ix^/lxJ = 1. Now pick another of the x's, say x 2 and form ,
its projection on alt that is, form the vector (x 2 • u 1 )u 1 . If the vectors were
ordinary geometric vectors, the relationship between xlt x 2 , and
ux would be somewhat as shown in Fig. 1 1.
y2 = x2 - (x 2 -u,)uls
y2 • Uj = x2 • ux - (x 2 •
UiHii! • u2)
Figure 11 = x2 • ux — x2 • ux = 0.
Sec. 8 Orthonormal Bases 163
u2
"
= —
|y,l
.
The vector y 2 cannot be zero because, by its definition, that would imply
and u x were linearly dependent.
that x 2
Having found u x and u 2 we choose x 3 and form ,
its projection on the
subspace of T) spanned by u t and u 2 By definition, . this is the vector
p = (x3 • U^Ui + (x 3 • u 2 )u 2 .
y3 = x3 - ( x3 • "i)ui - ( X3 ' u 2 )u 2 .
y3 • ux = (x 3 • ux) - (x 3 • ux) = 0.
Similarly
y3 • u2 = (x 3 • u2) - (x 3 • u2) = 0.
u3 = —y3
|y 3 l
.
Once again y 3 =
would imply linear dependence of x 3 ul5 and u 2 ,
.
But because the subspace spanned by u 2 and u 2 is the same as the one
spanned by x x and x 2 this would imply linear dependence of the x's.
,
y m=x J+1
- (x ;+1 • ujux (X; + l ' U,K,
8.2
;+l
ly^+ii
The vector
(x • ujux + . . .
+ (x •
11,011,
y2 = x - (x2 2
• uju!
= (i,o,-i)- (=±)(-L, — *)
V6/V6 V6 %/*/
= (1, 0, -1) + (|, -i, J) = (*, -i, -|).
Then
2 ~ ~
|y.l V66/36
(7 '-''- 4) -
=vk
Thus the plane P can be represented as all linear combinations
WXj + vx 2
or all linear combinations
su x + /u 2 .
The relationship between the two pairs of vectors is shown in Fig. 13.
Figure 13
Sec. 8 Orthonormal Bases 165
product by
x x
u,(x) =
(x, x r x ax
= y/ix.
Next set
>'
3 (x)
= x
2
— (x
2
, u 1 (x))w 1 (x) - (x
2
, u 2 (x))u 2 (x)
(£5*)^-
Then
» 3 (x) = l\l/2
2 2
'
(x -i) Jx
= V¥(x 2 -i) = 2
Vl(3x -l).
Of course, if we start with two bases for the same vector space "U and
apply the Gram-Schmidt process to them, we will in general get different
8.3 Theorem
= v fc
• u; - = r it j= 1 k- 1,
yk ' U A:
~ rk-
Thus yk = {yk • u^u^.. Since both the u's and the v's have length 1,
we must have
=
.
,
|v**u*l I-
It follows that yk • uk = ± 1 ; so v fc
= ±uk .
Similarly, the pairs {e 1 and {x l5 x 2 } span the xv-plane, and the com- , e2}
plete bases both span all of 3l 3 It follows from Theorem 8.3 that applying .
e,
.•'e,
Figure 14
Sec. 9 Eigenvectors 167
EXERCISES
1. (a) Find a vector (x, y, z) such that the triple of vectors (1, 1, 1), ( — 1, \, \),
and (x, y, z) forms an orthogonal basis for Jl 3 .
2. The vectors (1,1,1) and (1, 2, 1) span a plane Pin R 3 . Find an orthonormal
basis for P by using the Gram-Schmidt process.
3. The three vectors (1 2,, 1), ( -10, 0), and (0, 1,0,2) form a basis for a
, 1 , 1 ,
f{x)g{x) dx,
</".*>
f
find an orthonormal basis for P.,. [Hint. One basis for P2 consists of
{I,*,* 2 }.]
5. Show that the application of the Gram-Schmidt process to the three vectors
Xj, x 2 x3
, in Example 5 of the text gives the triple e l5 — e2 , e3 .
7. Let C[ — 7T, tt] be the vector space of complex-valued functions /(x) defined
for — tt < x < n. Let {f,g) be defined for/and £ in C[ — n, n] by
!\,m = n.
0, m j= n.
(c) Show that the vector subspaces of C[~n, tt] spanned by the following
two sets are the same:
SECTION 9
In this section we shall find a natural way to associate a basis for a vector EIGENVECTORS
t
space 1) with a given linear function /from 1? to \5. Suppose that there
168 Linear Algebra Chap. 2
f(x) = Ax.
Thus
x\ / 1 1 \ I x\ j +y x
1 1\/1\ /l
= 3
4 l/\2/ \2
and that
:>-i-i
2
That is, the vector (1,2) in 'Ji is an eigenvector corresponding to the
eigenvalue and the vector (1 —2) is an eigenvector corresponding to the
3, ,
Before discussing how to find eigenvectors, we shall see why they are
[Link] that 'V is a vector space, /is a linear function from C
U to
HI, and suppose that °0 has a basis {x ]5 x 2 , . . . , x„} consisting of eigen-
vectors of/, that is,
/(x fc ) = 4x ft ,
k= 1,2, ... ,«,
X = CjXx + . . . + c n\ n .
=
Then, using the linearity of/ and the fact that the x's are eigenvectors,
we have
/(x) = Cl /(Xl ) + . . .+ c n f(x n )
= c^Xj + . . . + cn X n x n .
(n \ n
9.1
6=1 / fc=i
where
9.2 /(x,) = 4x fc
, k=\,2,...,n.
zoo U/
Q+ „/( _]
we get
Figure 15 shows the effect of/ on each of the two eigenvectors x x and
x2 . It follows that the image /(x) of any vector x can be constructed by
170 Linear Algebra Chap. 2
MX | • VXj
Figure 15
1. Stretching away from the line through x 2 along the lines parallel to
For this analysis to work, it was essential that the eigenvectors of/ span
all of 3l 2 .
^4 l/\j>
A0 ,
4 l!\y kJ\y
or
(1-/1) 1 x
(1)
4 (i-X)l\y
:
It is clear that if the foregoing 2-by-2 matrix has an inverse, then the only
solutions are and y — 0. Hence we must try to find values of X for
x =
which the matrix fails to have an inverse. This will occur precisely when
/(1-A) 1 \
det = 0,
\ 4 (1 - X)l
and
- (1J0-C
X = 3: -2x +y=
X = - 1 2x + y = 0.
It follows that there are many solutions; but all we need is one for each
eigenvalue. We choose for simplicity
X = 3
\yl \2
X= -1
\yl \-2
though any nonzero numerical multiple of either vector would do.
9.4 (A - VK = 0.
\
Because Equations 9.3 and 9.4 are expressed in terms of the matrix A off,
we sometimes and eigenvalues of the matrix rather
refer to eigenvectors
than the function/. Of course Equation 9.4 is just Equation 9.2 in matrix
form; but Equation 9.2 also applies to linear functions that are not
representable by matrices.
Example 4. This example will show one way in which a linear function
can fail to have eigenvectors that provide a suitable basis for the vector
2
space on which/acts. Consider the linear function/from J{ to :K 2 defined
for fixed 0, < < 2tt, by
x\ /cos — s'mO\/x
/
yl \sin cos 01 \y
This function carries
1 /cos
into
0/ \sin
and
0\ /-sin 6
into
1/ \ cos0
(cos 6 — X —sin \
=
sin 6 cos 6 — XI
or
X2 - 2X cos 6 + 1 = 0.
This quadratic equation for X has only complex conjugate roots of the
form
cos 6 ± i sin 6 = e
±i6
.
2
Since we are looking for nonzero vectors x in III satisfying
f(x) = e
±ie
x,
/(x) •
y = x ./(y)
c
for all x and y in l). In particular, if/is a linear function from 51" to 31",
it has a matrix A = (a i}), and we can write the symmetry equation as
Ax '
y = x • Ay.
<a b'
A=
\b c
a b\ / Vl \ />'A _ lax, + bx 2 \ U)
b c!\xj \yj \bx1 + cxj \y 2 l
= ax^i + b(x 2 y 1 + x 1 y 2) + cx 2 y 2 .
Similarly, we compute
f
xA la b\ly 1 \_(xA(ayx + by %
Since the two dot-products are equal, we conclude that /is symmetric.
9.5 Theorem
C
Let /be a symmetric linear function from HJ to \J, where °\3 is a
vector space with an inner product. If x x and x 2 are eigenvectors of
/corresponding to distinct eigenvalues X x and A 2 then x 2 and x 2 are
,
orthogonal.
and
Xi -/(x 2 ) = xx • (A 2 x 2 ) = A 2 (Xi • x 2 ).
A 1 (x 1 • x2) = X (x x2 1
•
2 ).
T 2\
U\-X) 2 \
det = 0,
\ 2 (1 - X)l
formation /of Example 1, even though they have the same eigenvalues.)
Equation 9.4 for the eigenvectors has two interpretations, depending on
which eigenvalue is used. We have
1 J)0 -C
1= -1:
X= 3: -2x + = 0, 2y
A=-l: 2x+2y = 0.
We find solutions
X= 3: (x,y)=(l,l),
The vectors Xj = (1, 1), x2 = (1, — 1) are clearly orthogonal, and they
We can now take advantage of the fact that ii! and u 2 are eigenvectors
and also that they form an orthonormal set. The latter fact enables us to
express any vector in Jl 2 as a linear combination of u t and u 2 very simply.
From the previous section we have
X = (X • Uj)^ + (x • u 2 )u 2
for any x in 5l
2
. Now using the fact that u x and u 2 are eigenvectors of the
linear function g, we can write
g(\) = (x •
uOsCuO + (x • u 2 )g(u 2 ).
g(x) = 3(x •
u^Ux - (x u 2 )u 2 .
Thus the action of g has a geometric description like that given for fin
Example 2 of this section. The only difference is that the stretching and
reflection takes place along different lines: perpendicular lines in the case
of g, nonperpendicular lines in the case off.
EXERCISES
1. The linear function /from Jl
2
to R2 with matrix
'1 12\
2. Find all
;) n< (::)•
following matrices.
4\ /l 0\
(b)
10/ (d) 2 1
\0 1 2/
4. Show that is an eigenvalue of a linear function /if and only if/ is not
one-to-one.
176 Linear Algebra Chap. 2
f(x) for x in 31. Then the differential operator D 2 acts linearly from C (G0) (^)
to C (00) (3l).
(a) Show that for any real number A the functions cos ?.x and sin Ax are
eigenvectors of D 2
corresponding to the eigenvalue —A 2 .
(b) Let C (x, [0, tt] be the subspace of C (x, (#) consisting of functions/such
=0. Show if D
2
that /(0) =/(tt) that is restricted to acting on
C (x) [0, 77], then its only eigenfunctions are of the form sin kx, corre-
sponding to A = —A: 2 , where k is an integer.
'1 V
\0 1,
9. (a) Find the eigenvalues and a corresponding pair of eigenvectors for the
/from R 2 to 31 2 having matrix
linear function
(b) Show that the eigenvectors of the function /in part (a) form an orthog-
2
onal basis for 3l , and use this fact to give a geometric description of the
2
action of/ on 31 .
(c) Generalize the results you found for (a) and (b) to any linear function/
from 31" to 31" having a diagonal matrix diag (a u a 2 , . . . , c„).
2
10. Find the eigenvalues of the function^ on 3l with matrix
1 2^
,1 1,
2
Show that the corresponding eigenvectors span 3l and describe the action
ofg.
Sec. 9 Eigenvectors 111
11. Let C 2 be the complex vector space of pairs z = (zx z 2) of complex numbers, ,
z • w = z 1 w1 + z2 w2 .
/'cos 6 —sin 6\
v
sin 8 cos 8}
V
X = e
i{
X=e-
(b) Show that if sin 6=0, then every nonzero vector in C 2 is an eigen-
vector of the function fin part (a).
(c) Show that the eigenvectors in part (a) form a basis for C2 .
(d) Show that the eigenvectors in part (a) are orthogonal with respect to the
complex inner product in C2 .
12. Explain in geometric terms why a rotation about the origin in ft 2 through
an angle 8, < 8 < 2tt, has no eigenvectors in ft 2 unless 8 = or 8 = -n.
13. Suppose that 'W is a complex vector space with a complex inner product, so
that <z, w) = (w, z). Then a linear function / from 10 to 10 is called
Hermitian symmetric if </(z), w) = <z,/(w)>, for z, w in 10.
(a) Show that if /'is Hermitian symmetric and has A for a complex eigen-
value, then A must actually be real.
(b) Show that if 10 has complex dimension 2 and / is given by a 2-by-2
complex matrix
y\
(a) Let
(a c
b d
Show that the vector equation
x' = Ax
is the same as the system of first-order differential equations
x = ax + cy
y = bx + dy
'1 2\
v2 1,
find solutions of the vector differential equation x' = As. in the form
x = e^i'c! + e^'cj.
(Z> - r x ){D - r 2 )y =
is equivalent to the system of first-order equations
x = rx x
y' =x + r 2 x,
(e) Show that the characteristic equation of the matrix A found in part (d)
is the same as the characteristic equation of the operator (D — r^) x
(D — r 2 ) as defined in Section 6.
15. (a) Show that the functions cos kx and sin kx are eigenvectors of the
differential operator D2 acting on C (cc) What are the corresponding
.
eigenvalues?
(b) If we restrict the operator D 2
to real-valued functions defined for
—" < x < n, we can define an inner product by
</><?>=
L f(x)g(x)dx.
Show that the eigenvectors cos kx and sin Ix are orthogonal with respect
to this inner product for k, I = 1,2, 3, ... . Then use Theorem 8.1 to
(c) Show that the functions e ikx and e~ lkx are eigenvectors of the differential
operator D2 acting on complex-valued functions. What are the corre-
sponding eigenvalues?
(d) Using the complex inner product
SECTION 10
10.1 Theorem
then
^Vi + . . . + s n\ n = x = w + r n\ n ,
fa - *iK + • • • + (r n - s n )v n = 0.
Since the v's are independent, the numbers rt — st must all be zero,
and so (s x , . . . , s n ) is the same as (r ls . . . , r n ).
\ / \ A (b) In the plane, let Vj be the vector of unit length along the
q ~ *"*
horizontal axis, and v 2 the vector of unit length at an angle 60°
counterclockwise from v x . Let x be the vector of unit length 60°
Figure 16 beyond v 2 shown , in Fig. 16.
By geometry, CB is parallel to OA and OC is parallel to AB
(why?); so OABC is a parallelogram and v = 2 x + Vj. Hence x = —v + 1
v2 and the coordinates of x with respect to this basis are (—1, 1).
known for each vector of a basis, then the function is known completely.
This follows from the fact that any vector x can be expressed as a linear
combination r x \x + • • • + r n\ n of the vectors in a basis. Then/(x) is the
combination ^/(Vi) + . . . + rn
f(\ n) of the vectors /(vj with the same
ID, respectively. The possibility that 1J and 10 are the same space is not
ruled out. The matrix off with respect to the bases V and is defined to W
be the m-by-n matrix whose y'th column is the coordinate vector of/(v ) ;
with respect to the basis W. Thus if we denote by A = (a^) the matrix off
relative to Kand W, then the entries in they'th column of A appear as the
coefficients a tj in the equation
The matrix A contains all the information needed to evaluate f(x) for
any x. For if x is represented as
/(x)=i>,/(v,)
n m
=i
f=l
(i<wW
\j=l I
to be the same bases. We do this in all the examples for which HJ is the
same as ID.
0/ \ 1/
itself, and it may be considered as a linear function from °U to C
U. We have
-1 -1
1
(0 -1 N
1 1
nomial into the 4-tuple (1, 1, 1, 1); it takes the polynomial x into the
1
4-tuple (1, —1,2,3); and it takes x z into (1, 1,4,9). Its matrix with
182 Linear Algebra Chap. 2
10.2 Theorem
Note that if °\J = ft", the coordinate map c for the natural basis is the
identity function c(x) = x.
all of 'Ui, is called an isomorphism, and two spaces such that there exists an
isomorphism between them are said to be isomorphic. We have shown that
coordinate maps are isomorphisms, which proves:
10.3 Theorem
Any statement true for one space (provided it can be formulated entirely
in terms of addition and numerical multiplication) can be carried over by
an isomorphism to give a true corresponding statement for an isomorphic
space. For example, if TJ —^-> ID is an isomorphism, then an equation
n n
y =2
i=l
a xi
i
ls true m ^ if an d only if
f(y) =^
t'=l
a if(*i) is true in ID.
with respect to the bases V and W. To prove the assertion, note that the
10.4 Theorem
(C W o g o Cy
1
) °(c v ofo Cr]) = CW o
(g of) O c'r],
The special case in which 1L, CU, and ID are all the same space with the
same basis is particularly important. If °0 —f-+ T) has matrix A, then
/o/has matrix A 2 f °f °f has
,
matrix A3 , etc.
A) 1 0'
A=10 2
\0 0/
It is easy to compute that
'0 2"
\0 0/
1 2 0\
1 1 3y
3 2
with respect to the natural bases in 'J\ and ll\ . Consider the bases
in 3t 3 and
•) -(;;
in Si 2
. To find the matrix off with respect to (v l5 v 2 v 3 ) , and (w x , w 2 ), we
compute , ,
5w
'(
'Hio) = -
/(v 2 ) = ( 1=8*!- 3w 2 , /(v3) = ( )
= - 4w +
i
3w 2-
'5 8 -4
\0 -3 3
X
Sec. 10 Coordinates 185
c.v( w) = (c x ° cv
1
) ° c v (v/)
The inverse matrix gives the x's in terms of the v's and corresponds to
x
cv o cy . In the example in the preceding paragraph, the matrices giving
V and W in terms of the natural bases in 'Si
2