Complex Numbers in Engineering Math
Complex Numbers in Engineering Math
Part I
Linear Algebra and Differential Equations
eNote 1
Complex Numbers
In this eNote we introduce and investigate the set of numbers C , the complex numbers. Since
C is considered to be an extension of R , the eNote presumes general knowledge of the real
numbers, including the elementary real functions such as the trigonometric functions and the
natural exponential function. Finally elementary knowledge of vectors in the plane is taken for
granted.
1.1 Introduction
In the two examples above the right-hand sides were positive. When considering the
equation
x2 = k , k ∈ R
we must be more careful; here everything depends on the sign of k . If k ≥ 0, the
equation has the solutions √ √
x = k and x = − k ,
eNote 1 1.2 COMPLEX NUMBERS INTRODUCED AS PAIRS OF REAL NUMBERS 2
√ √
since k 2 = k and (− k ) 2 = k . But if k < 0 the equation has no solutions, since real
numbers with negative squares do not exist.
But now we ask ourselves the question, is it possible to imagine a set of numbers larger
than the set of real numbers; a set that includes all the real numbers but in addition also
includes solutions to an equation like
x2 = −1?
The equation should then in analogy to the equations above have two solutions
√ √
x = −1 and x = − −1 .
Let us
√ be bold and assume2that this is in fact possible. We then choose to call this number
i = −1. The equation x = −1 then has two solutions, viz.
x = i and x = −i
since, if we assume that the usual rules of algebra hold,
√ 2 √ √
i2 = −1 = −1 and (−i)2 = (−1 · −1 ) 2 = (−1)2 (− −1)2 = −1 .
As we just mentioned, we make the further demand on the hypothetical number i , that
one must be able to use the same algebraic rules that apply to the real numbers. We
must e.g. be able to multiply i by a real number b and add this to another real number
a. In this way a new kind of number z of the type
z = a + ib , ( a, b) ∈ R2
emerges.
Below we describe how these ambitions about a larger set of numbers can be fulfilled.
We look at how the structure of the set of numbers should be and at which rules apply.
We call this set of numbers the complex numbers and use the symbol C . R must be a
proper subset of C — that is, C contains all of R together with the new numbers which
fulfill the above ambitions that are impossible in R. As we have already hinted C must
be two-dimensional in the sense that a complex number contains two real numbers, a and
b.
where a and b are real numbers and i is the new imaginary number that satisfies i2 = −1 .
This form is very practical in computation with complex numbers. But we have not re-
ally clarified the meaning of the expression (1-1). For what is the meaning of a product
like ib , and what does the addition a + ib mean?
A satisfactory way of introducing the complex numbers is as the set of pairs of real num-
bers ( a, b). In this section we will show how in this set we can define arithmetic opera-
tions (addition, subtraction, multiplication and division) that fulfill the ordinary arith-
metic rules for real numbers. This will turn out to give full credit to the form (1-1).
C = {( a, b) | a, b ∈ R} (1-2)
As the symbol for an arbitrary complex number we will use the letter z .
Example 1.2
First we introduce the arithmetic rule for the addition of complex numbers. Then sub-
traction as a special form of addition.
eNote 1 1.2 COMPLEX NUMBERS INTRODUCED AS PAIRS OF REAL NUMBERS 4
z1 + z2 = ( a, b) + (c, d) = ( a + c, b + d) . (1-3)
For the two complex numbers z1 = (2, 7) and z2 = (4, −3) we have:
The complex number (0, 0) is neutral with respect to addition, since for every complex
number z = ( a, b) we have:
z + (0, 0) = ( a, b) + (0, 0) = ( a + 0, b + 0) = ( a, b) = z .
It is evident that (0, 0) is the only complex number that is neutral with respect to addi-
tion.
For every complex number z there exists an additive inverse (also calld opposite number)
denoted −z, which, when added to z, gives (0, 0). The complex number z = ( a, b) has
the additive inverse −z = (− a, −b), since
It is clear that (− a, −b) is the only additive inverse for z = ( a, b), so the notation −z
is well-defined. By use of this, subtraction of complex numbers can be introduced as a
special form of addition.
eNote 1 1.2 COMPLEX NUMBERS INTRODUCED AS PAIRS OF REAL NUMBERS 5
z1 − z2 = z1 + (−z2 ) . (1-4)
Let us for two arbitrary complex numbers z1 = ( a, b) and z2 = (c, d) calculate the
difference z1 − z2 using definition 1.5:
z1 − z2 = ( a, b) + (−c, −d) = ( a + (−c), b + (−d)) = ( a − c, b − d) .
This gives the simple formula
z1 − z2 = ( a − c, b − d) . (1-5)
For the two complex numbers z1 = (5, 2) and z2 = (4, −3) we have:
z1 − z2 = (5 − 4, 2 − (−3)) = (1, 5) .
While addition and subtraction appear to be simple and natural, multiplication and
division of complex numbers appear to be more odd. Later we shall see that all the
four arithmetic rules have geometrical equivalents in the so-called complex plane that
constitutes the graphical representation of the complex numbers. But first we must
accept the definitions at their face value. First we give the definition of multiplication.
Then follows the definition of division as a special form of multiplication.
For the two complex numbers z1 = (2, 3) and z2 = (1, −4) we have:
The complex number (1, 0) is neutral with respect to multiplication, since for every com-
plex number z = ( a, b) we have that:
z · (1, 0) = ( a, b) · (1, 0) = ( a · 1 − b · 0 , a · 0 + b · 1) = ( a, b) = z .
It is clear that (1, 0) is the only complex number that is neutral with respect to multipli-
cation.
For every complex number z apart from (0, 0) there exists a unique reciprocal number
that when multiplied by the given number gives (1, 0). It is denoted 1z . The complex
number ( a, b) has the reciprocal number
1 a b
= ,− 2 , (1-7)
z a2 + b2 a + b2
since
a2 b2
a b ab ba
( a, b) · 2 2
,− 2 = + ,− 2 + 2 = (1, 0) .
a +b a + b2 2
a +b 2 2
a +b 2 a +b 2 a + b2
Exercise 1.9
Show that every complex number z 6= (0, 0) has exactly one reciprocal number.
By the use of reciprocal numbers we can now introduce division as a special form of
multiplication.
eNote 1 1.2 COMPLEX NUMBERS INTRODUCED AS PAIRS OF REAL NUMBERS 7
z1 1
The quotient is defined as the product of z1 and the reciprocal number for z2 :
z2 z2
z1 1
= z1 · . (1-8)
z2 z2
Let us for two arbitrary complex numbers z1 = ( a, b) and z2 = (c, d) 6= (0, 0) compute
z
the quotient 1 from the Definition 1.10:
z2
ac + bd bc − ad
1 c d
z1 · = ( a, b) 2 ,− 2 = , .
z2 c + d2 c + d2 c2 + d2 c2 + d2
ac + bd bc − ad
z1
= , . (1-9)
z2 c2 + d2 c2 + d2
1·3+2·4 2·3−1·4
z1 11 2
= , = , .
z2 32 + 42 32 + 42 25 25
We end this section by showing that the complex numbers, with the above arithmetic
operations, fulfill the computational rules known from the real numbers.
eNote 1 1.2 COMPLEX NUMBERS INTRODUCED AS PAIRS OF REAL NUMBERS 8
Proof
Let us look at property 1, the commutative rule. Given two complex numbers z1 = ( a, b) and
z2 = (c, d). We see that
z1 + z2 = ( a + c, b + d) = (c + a, d + b) = z2 + z1 .
To establish the second equality sign we have used that for both the first and the second
coordinates the commutative rule for addition of real numbers applies. By this it is seen that
the commutative rule also applies to complex numbers.
In the proof of the properties 2, 5, 6 and 9 we similarly use the fact that the corresponding
rules apply to the real numbers. The details are left to the reader. For the properties 3, 4, 7
and 8 we refer to treatment above in this section.
eNote 1 1.3 COMPLEX NUMBERS IN RECTANGULAR FORM 9
(0,b) (a,b)
(0,1)
X
(0,0) (1,0) (a,0)
Since to every ordered pair of real numbers corresponds a unique point in the ( x, y)-
plane and vice versa, C can be considered to be the set of points in the ( x, y)-plane.
Figure 1.1 shows six points in the ( x, y)-plane, i.e. six complex numbers.
First we identify all complex numbers of the type ( a, 0), i.e. the numbers that lie on the
x-axis, with the corresponding real number a . In particular the number (0, 0) is written
as 0 and the number (1, 0) as 1 . Note that this will not be in conflict with the arithmetic
rules for complex numbers and the ordinary rules for real numbers, since
( a, 0) + (b, 0) = ( a + b, 0 + 0) = ( a + b, 0)
and
( a, 0) · (b, 0) = ( a · b − 0 · 0 , a · 0 + 0 · b) = ( ab, 0) .
In this way the x-axis can be seen as an ordinary real number axis and is called the real
axis. In this way the real numbers can be seen as a subset of the complex numbers. That
the y-axis is called the imaginary axis is connected to the extraordinary properties of the
complex number i which we now introduce and investigate.
eNote 1 1.3 COMPLEX NUMBERS IN RECTANGULAR FORM 10
A decisive motivation for the introduction of complex numbers was the wish
for a set of numbers that contained the solution to the equation
x 2 = −1 .
z = a + i · b = a + ib . (1-10)
This way of writing the complex number is called the rectangular form of z .
Proof
The proof consists of simple manipulations in which we use the new way of writing numbers
of this type.
( a, b) = ( a, 0) + (0, b) = ( a, 0) + (0, 1) · (b, 0) = a + i b .
eNote 1 1.3 COMPLEX NUMBERS IN RECTANGULAR FORM 11
Imaginary Axis
Real Axis
Figure 1.2: Six complex numbers in rectangular form in the complex number plane
0 + z = z and 1z = z .
0z = 0 .
1
Let us now consider all complex numbers of the type (0, b) . Since
(0, b) = 0 + ib = ib ,
i can be understood as the unit of the y-axis, and therefore we refer to i as the imaginary
unit. From this comes the name the imaginary axis for the y-axis.
In Figure 1.2 we see an update of the situation from Figure 1.1, where numbers are given
in their rectangular form.
All real numbers are complex but not all complex numbers are real!
eNote 1 1.3 COMPLEX NUMBERS IN RECTANGULAR FORM 12
In the following example it is shown how multiplication can be carried out through
ordinary computation with the rectangular form of the factors.
We compute the product of two complex numbers given in rectangular form z1 = a + ib and
z2 = c + id :
Exercise 1.17
Prove that the following rule for real numbers — the so-called zero rule — also applies to
complex numbers: "‘A product is 0 if and only if at least one of factors is 0 ."’
eNote 1 1.3 COMPLEX NUMBERS IN RECTANGULAR FORM 13
1. z1 = z , z2 = z · z , z3 = z · z · z etc.
2. By definition z0 = 1 .
1
3. Finally we put z−n = .
zn
It is easily shown that the usual rules for computations with integer powers of real
numbers also apply for integer powers of complex numbers:
We end this section by introducing the concepts real part and imaginary part of complex
numbers.
The expression rectangular form refers to the position of the number in the com-
plex number plane, where Re(z) is the number’s perpendicular drop point on
the real axis, and Im(z) its perpendicular drop point on the imaginary axis. In
short the real part is the first coordinate of the number while the imaginary
part is the second coordinate of the number.
Note that every complex number z can be written in rectangular form like this:
z = Re(z) + i Im(z) .
z1 = 3 − 2i , z2 = i5 , z3 = 25 + i .
Find the real part and the imaginary part of each number.
Re(z1 ) = 3 , Im(z1 ) = −2
Re(z2 ) = 0 , Im(z2 ) = 5
Re(z3 ) = 25 , Im(z3 ) = 1
Two complex numbers in rectangular form are equal if and only if both their
real parts and imaginary parts are equal.
eNote 1 1.4 CONJUGATION OF COMPLEX NUMBERS 15
z = a − ib . (1-13)
Conjugating a complex number corresponds to reflecting the number in the real axis as
shown in Figure 1.3.
It is obvious that the conjugate number of a conjugate number is the original number:
z = z. (1-14)
Furthermore the following useful formula for the product of complex number and its
conjugate applies:
z · z = | z |2 (1-15)
In the following method we show a smart way of finding the rectangular form of a
fraction when the denominator is not real: we use the fact that the product of a number
z = a + ib and its conjugate z = a − ib is always a real number, cf. (1-15).
z z( a − ib) z( a − ib)
= = 2 .
a + ib ( a + ib)( a − ib) a + b2
An example:
2−i (2 − i)(1 − i) 1 − 3i 1 − 3i 1 3
= = 2 2
= = − i.
1+i (1 + i)(1 − i) 1 +1 2 2 2
In conjugation in connection with the four ordinary arithmetic operations the following
rules apply.
1. z1 + z2 = z1 + z2
2. z1 − z2 = z1 − z2
3. z1 · z2 = z1 · z2
Proof
The proof is carried out by simple transformation using the rectangular form of the numbers.
As an example we show the first formula. Suppose that z1 = a1 + ib1 and z2 = a2 + ib2 . Then:
Finally we note that all complex numbers on the real axis are identical with their con-
jugate number and that they are the only complex numbers that fulfill this condition.
Therefore we can state a criterion for whether a given number in a set of complex num-
bers is real:
Proof
z = z ⇔ a − ib = a + ib ⇔ 2ib = 0 ⇔ b = 0 ⇔ z ∈ AR .
eNote 1 1.5 POLAR COORDINATES 18
The obvious way of stating a point (or a position vector) in an ordinary ( x, y)-coordinate
system is by the point’s rectangular, i.e. orthogonal, coordinates ( a, b). In many situations
it is, however, useful to be able to determine a point by its polar coordinates, consisting of
its distance to (0, 0) together with its direction angle from the x-axis to its position vector.
The direction angle is then positive if it is measured counter-clockwise and negative if
measured clockwise.
Analogously, we now introduce polar coordinates for complex numbers. Let us first be
absolutely clear about the orientation of the complex number plane.
Re
0
The ingredients in the polar coordinates of complex numbers are (as mentioned above)
its distance to (0, 0) called the absolute value, and its direction angle called the argument.
We now introduce these two quantities.
eNote 1 1.5 POLAR COORDINATES 19
Suppose z 6= 0 . Every angle from the positive part of the real axis to the position
vector for z is called an argument for z and is denoted arg(z) . The angle is positive
or negative relative to the orientation of the complex number plane.
Im
z
|z|
arg(z)
Re
0
The pair
| z | , arg(z)
of the absolute value of z and an argument for z will collectively be called the polar
coordinates of the number.
Note that the argument for a number z is not unique. If you add 2π to an
arbitrary argument for z, you get a new valid direction angle for z and there-
fore a valid argument. Therefore a complex number has infinitely many argu-
ments corresponding to turning an integer number of times extra clockwise or
counter-clockwise in order to reach the same point again.
You can always choose an argument for z that lies in the interval from −π to π. Tradi-
tionally this argument is given a preferential position. It is called the principal value of
the argument.
eNote 1 1.5 POLAR COORDINATES 20
arg(z) ∈ ] − π, π ] .
Two complex numbers are equal if and only if both their absolute values and
the principal arguments are equal.
Im
‐2+2i 2+2i
3π
4
π
π 4
‐2 Re
_π
4
_ 3π
4
‐2‐2i 2‐2i
The figure shows five complex numbers, four of which lie on the lines through (0, 0) bisecting
the four quadrants. We read:
eNote 1 1.5 POLAR COORDINATES 21
π
• 2 + 2i has the principal argument 4 ,
i|z|sin(v) z=a+ib
|z| b
i
v a
0 1 |z|cos(v)
1. The rectangular form is computed from the polar coordinates like this:
2. The absolute value is computed from the rectangular form like this :
p
| z | = a2 + b2 . (1-18)
a b
cos(v) = and sin(v) = . (1-19)
|z| |z|
When z is drawn in the first quadrant it is evident that the computational rules
(1-17) and (1-19) are derived from well-known formulas for cosine and sine to
acute angles in right-angled triangles and (1-18) from the theorem of Pythago-
ras. By using the same formulas it can be shown that the introduced methods
are valid regardless of the quadrant in which z lies.
eNote 1 1.5 POLAR COORDINATES 23
√
Find the polar coordinates for the number z = − 3 + i .
Im
z = ‐ 3 +i
|z|
v
Re
0
We use the rules in Method 1.29. Initially we identify the real and the imaginary parts of z
as √
a = − 3 and b = 1.
First we determine the absolute value:
p q √ √
| z | = a + b = (− 3) 2 + 12 = 3 + 1 = 2 .
2 2
5π 5π
Since only v = satisfies both equations, we see that Arg(z) = .
6 6
eNote 1 1.5 POLAR COORDINATES 24
We end this section with the important product rules for absolute values and arguments.
| z1 · z2 | = | z1 | · | z2 | . (1-20)
Corollary 1.32
The absolute value for the quotient of two complex numbers z1 and z2 where z2 6= 0
is found by
z1 |z |
= 1 . (1-21)
z2 | z2 |
The absolute value of the nth power of a complex number z is for every n ∈ Z given
by
| z1 n | = | z1 | n . (1-22)
Exercise 1.33
Write down in words what the formulas (1-20), (1-21) and (1-22) say and prove them.
eNote 1 1.6 GEOMETRIC UNDERSTANDING OF THE FOUR COMPUTATIONAL
OPERATIONS 25
Corollary 1.35
Given two complex numbers z1 6= 0 and z2 6= 0 .Then:
Exercise 1.36
The first exact description of the complex numbers was given by the Norwegian sur-
veyor Caspar Wessel in 1796 . Wessel introduced complex numbers as line segments
with given lengths and directions, that is what we now call vectors in the plane. There-
fore computations with complex numbers were geometric operations carried out on
eNote 1 1.6 GEOMETRIC UNDERSTANDING OF THE FOUR COMPUTATIONAL
OPERATIONS 26
z2=c+id
z1+z2=(a+c)+i( b+d)
i
0 1
z1=a+ib
vectors. In the following we recollect the ideas in the definition of Wessel. It is easy
to see the equivalence between the algebraic and geometric representations of addition
and subtraction — it is more demanding to understand the equivalence when it comes
to multiplication and division.
The position vector for z1 + z2 is the sum of the position vectors for z1 and
z2 . (See Figure 1.4).
Proof
eNote 1 1.6 GEOMETRIC UNDERSTANDING OF THE FOUR COMPUTATIONAL
OPERATIONS 27
z1=a+ib
z1-z2=(a-c)+i(b-d)
z2=c+id
i
0 1
-z2=-c-id
While in the investigation of geometrical addition (and subtraction) we have used the
rectangular form of complex numbers, in the treatment of geometric multiplication (and
division) we shall need their polar coordinates.
Proof
First part of the theorem appears from Theorem 1.31 while the second part is evident from
Theorem 1.34.
Im
2π
z2 3
11π
12
z1 π
z1 z2
4
1/2 1 2 Re
1 π
and 2, 2π
Two complex numbers z1 and z2 are given by the polar coordinates 2, 4 3 , re-
spectively. (Figure 1.6,)
We compute the product of z1 and z2 by the use of their absolute values and arguments:
1
| z1 z2 | = | z1 | | z2 | = ·2 = 1
2
π 2π 11π
arg(z1 z2 ) = arg(z1 ) + arg(z2 ) = + = .
4 3 12
Thus the product z1 z2 is the complex number that has the absolute value 1 and the argument
11π
.
12
z1
u
z1
z2
z2 u-v
v
2 3 6 Re
The numbers z1 and z2 are given by |z1 | = 6 with arg(z1 ) = u and |z2 | = 2 with arg(z2 ) = v
z1
respectively. Then can be determined as
z2
z1 6 z
= = 3 and arg 1 = u − v .
z2 2 z2
1. e0 = 1 ,
In this section we will introduce a particularly useful extension of the real exponential
function to a complex exponential function, that turns out to follow the same rules of
computation as its real counterpart.
eNote 1 1.7 THE COMPLEX EXPONENTIAL FUNCTION 30
where e (about 2.7182818 . . . ) is base for the real natural exponential function.
we see that the complex exponential function is everywhere on the real axis identical to
the real exponential function. Therefore we do not risk a contradiction when we in the
following allow (and often use) the way of writing
Im
ez
ex
y
0
Re
We now consider the complex number ez where z is an arbitrary complex number with
the rectangular form z = x + iy . Then (by use of Theorem 1.31) we see that
The polar coordinates for z = x + iy are then (ex , y) , which is illustrated in Figure 1.8.
For the trigonometric functions cos( x ) and sin( x ) we know that for every integer p
cos( x + p2π ) = cos( x ) and sin( x + p2π ) = sin( x ) . If the graph for cos( x ) or sin( x ) is
displaced by an arbitrary multiple of 2π , it will be mapped onto itself. Therefore the
functions are called periodic having a period of 2π .
A similar phenomenon is seen for the complex exponential function. It has the imaginary
period i 2π . This is closely connected to the periodicity of the trigonometric functions
as can be seen in the proof of the following theorem.
ez+ip2π = ez . (1-27)
Proof
Then:
= ez .
In the following example the periodicity of the complex exponential function is illus-
trated.
eNote 1 1.7 THE COMPLEX EXPONENTIAL FUNCTION 32
First we write z in rectangular form: z = x + iy . In Example 1.30 we found that the right-hand
side in (1-28) has the absolute value |z| = 2 and the principal argument v = 5π
6 . Since the left-
hand and the right-hand sides must have the same absolute value and the same argument,
apart from an arbitrary multiple of 2π , we get
√
| ez | = | − 3 + i | ⇔ ex = 2 ⇔ x = ln(2)
√ 5π
arg(ez ) = arg(− 3 + i) ⇔ y = v + p2π = + p2π , p ∈ Z .
6
All solutions for (1-28) are then
5π
z = x + iy = ln(2) + i + p2π , p ∈ Z .
6
We end this section by stating and proving the rule of computations mentioned in the
introduction and known from the real exponential function.
1. e0 = 1
Proof
Point 1 in the theorem that e0 = 1, follows from the fact that the complex exponential function
is identical with the real exponential function on the real axis, cf. (1-24).
eNote 1 1.8 THE EXPONENTIAL FORM OF COMPLEX NUMBERS 33
In point 2 we set z1 = x1 + iy1 and z2 = x2 + iy2 . From the set of polar coordinates and 1.38
we get:
ez1 · ez2 = (e x1 , y 1 ) · (e x2 , y 2 ) = (e x1 · e x1 , y 1 + y 2 ) = (e x1 + x2 , y 1 + y 2 )
= e(x1 +x2 )+i(y1 +y2 ) = e(x1 +iy1 )+(x2 +iy2 )
= ez1 + z2 .
In point 3 we set z = x + iy and with the use of sets of polar coordinates and the repeated
use of Theorem 1.38 we get:
Exercise 1.45
Let v be an arbitrary real number. If we substitute the pure imaginary number iv into
the complex exponential function we get from the Definition 1.41:
Im
eiv isin(v)
v
Re
cos(v) 0 1
The two most-used ways of writing complex numbers both in pure and applied math-
ematics are the rectangular form (as is frequently used above) and the exponential form.
In the exponential form the polar coordinates of the number (absolute value and argu-
ment), in connection with the complex exponential function. Since the polar coordinates
appear explicitly in this form, it is also called the polar form.
where v is an argument for z . This way of writing is called the exponential form (or
the polar form) of the number.
eNote 1 1.8 THE EXPONENTIAL FORM OF COMPLEX NUMBERS 35
Proof
Let v be an argument for the complex number z 6= 0 , and put r = |z| . We show that reiv has
the same absolute value and argument as z , and thus the two numbers are identical:
1.
reiv = |r | eiv = r .
We now give an example of multiplication following Method 1.48; cf. Example 1.39.
1 πi 3π
z1 = e 4 and z2 = 2 e 2 i .
2
The product of the numbers is found in exponential form as
1 π i 3π 1 π 3π π 3π 7π
z1 z2 = e 4 2 e 2 i = ( · 2)e 4 i+ 2 i = 1ei( 4 + 2 ) = e 4 i .
2 2
eNote 1 1.8 THE EXPONENTIAL FORM OF COMPLEX NUMBERS 36
Exercise 1.50
In the following we will show how so-called binomial equations can be solved by the use
of the exponential form. A binomial equation is an equation with two terms in the form
zn = w , (1-32)
where w ∈ C and n ∈ N . Binomial equations are described in more detail in eNote 2
about polynomials.
First we show an example of the solution of a binomial equation by use of the exponen-
tial form and then we formulate the general method.
The idea is that we write both z and the right-hand side in exponential form.
If z has the exponential form z = seiu , then the equation’s left-hand side can be computed as
The right-hand side is also written in exponential form. The absolute value r of the right-hand
side is found by
√ q √
r = | − 8 + 8 3 i | = (−8)2 + (8 3)2 = 16 .
The argument v of the right-hand side satisfies
√ √
−8 1 8 3 3
cos(v) = = − and sin(v) = = .
16 2 16 2
By use of the two equations the principal argument of the right-hand side can be determined
to be
√ 2π
v = arg(−8 + 8 3i) = ,
3
eNote 1 1.8 THE EXPONENTIAL FORM OF COMPLEX NUMBERS 37
We now substitute (1-34) and (1-35) into (1-33) in order to replace the right- and left-hand
side with the exponential counterparts
2π
s4 ei4u = 16e 3 i .
Since the absolute value of the left-hand side must be equal to absolute value of the right-
hand side we get √
4
s4 = 16 ⇔ s = 16 = 2 .
2π
The argument of the left-hand side 4u and the argument of the right-hand side 3 must be
equal apart from a multiple of 2π . Thus
2π π π
4u = + p2π ⇔ u = + p , p ∈ Z .
3 6 2
These infinitely many arguments correspond, as we have seen earlier, to only four half-lines
from (0, 0) determined by the arguments obtained by putting p = 0, p = 1, p = 2 and p =
3 . For any other value of p the corresponding half-line will be identical to one of the four
mentioned above. E.g. the half-line corresponding to p = 4 has the argument
π π π
u= + 4 = + 2π ,
6 2 6
i.e. the same half-line that corresponds to p = 0 , since the difference in argument is a whole
revolution, that is 2π .
Therefore the given equation (1-33) has exactly four solutions that lie on the four mentioned
half-lines and that are separated the distance s = 2 from 0 . Stated in exponential form:
π π
z = 2 ei( 6 + p 2 ) , p = 0 , 1 , 2 , 3 .
All solutions to a binomial equation lie on a circle with the centre at 0 and radius equal
to the absolute value of the right-hand side. The connecting lines between 0 and the
solutions divide the circle into equal angles. This is illustrated in Figure 1.10 which
shows the solutions to the equation of the fourth degree from Example 1.51.
eNote 1 1.8 THE EXPONENTIAL FORM OF COMPLEX NUMBERS 38
Im
z1
z0
‐2
Re
0 2
z2
z3
√
Figure 1.10: The four solutions for z4 = −8 + 8 3i
The method in Example 1.51 we now generalize in the following theorem. The theorem
is proved in eNote 2 about polynomials.
Let r be an arbitrary positive real number. Show by use of Theorem 1.52 that the binomial
quadratic equation
z2 = −r
has the two solutions √ √
z0 = i r and z1 = −i r.
eNote 1 1.9 LINEAR AND QUADRATIC EQUATIONS 39
Let a and b be complex numbers with a 6= 0 . A complex linear equation of the form
az = b
in analogy with the corresponding real linear equation has exactly one solution
b
z= .
a
With a and b in rectangular form, the solution is easily found in rectangular form, as
shown in the following example.
The equation
(1 − i) z = (5 + 2i)
has the solution
5 + 2i (5 + 2i)(1 + i) 3 + 7i 3 7
z= = = = + i.
1−i (1 − i)(1 + i) 2 2 2
Also in the solution of complex quadratic equations we use a formula that corresponds
to the well-known solution formula for real quadratic equations. This is given in the
following theorem that is proved in eNote 2 about polynomials.
eNote 1 1.10 COMPLEX FUNCTIONS OF A REAL VARIABLE 40
az2 + bz + c = 0 (1-38)
−b
If in particular D = 0 , we find z0 = z1 = .
2a
Concrete examples of the application of the theorem can be found in Section 30.5.2 in
eNote 2 about polynomials.
In this section we use the theory of the so-called epsilon functions for the introduction of dif-
ferentiability. The material is a bit more advanced than previously and knowledge about epsilon
functions from eNote 3 (see Section 3.4) may prove advantageous. Furthermore the reader should
be familiar with the rules of differentiation of ordinary real functions.
f : t 7→ ect , t ∈ R , (1-40)
where c is a given complex number. This type of function has many uses in pure and
applied mathematics. A main purpose of this section is to give a closer description
of these. They are examples of the so-called complex functions of a real variable. Our
investigation starts off in a wider sense with this broader class of functions. I.a. we
show how concepts such as differentiability and derivatives can be introduced. Then
we give a fuller treatment of functions of the type in (1-40).
eNote 1 1.10 COMPLEX FUNCTIONS OF A REAL VARIABLE 41
The notation f : R 7→ C tells us the function f uses a variable in the real num-
ber space, but ends up with a result in the complex number space. Consider
e.g. the function f (t) = eit . At the real number t = π4 we get the complex
function value
√ √
π π
i
π π 2 2
f = e 4 = cos + i sin = + i.
4 4 4 2 2
1. e(0) = 0, and
2. |e(t)| → 0 for t → 0 .
eNote 1 1.10 COMPLEX FUNCTIONS OF A REAL VARIABLE 42
Note that if e is an epsilon function, then it follows directly from the Definition
1.57 that for every t0 ∈ R:
| e(t − t0 ) | → 0 for t → t0 .
In the following example a pair of complex epsilon functions of a real variable are
shown.
The function
t 7→ i sin(t) , t ∈ R
is an epsilon function. This is true because requirement 1 in definition 1.57 is fulfilled by
i sin(0) = i · 0 = 0
and requirement 2 by
We are now ready to introduce the concept of differentiability for complex functions of a
real variable.
eNote 1 1.10 COMPLEX FUNCTIONS OF A REAL VARIABLE 43
Theorem 1.60
For a function f : R 7→ C with the rectangular form f (t) = g(t) + ih(t) and a
complex number c with the rectangular form c = a + ib :
f is differentiable at t0 ∈ R with
f 0 ( t0 ) = c ,
Proof
First suppose that f is differentiable at t0 and f 0 (t0 ) = a + ib, where a, b ∈ R . Then there
exists an epsilon function e such that f for every t can be written in the form
We rewrite both the left- and the right-hand side into their rectangular form:
g(t) + ih(t) =
g(t0 ) + ih(t0 ) + a(t − t0 ) + ib(t − t0 ) + Re(e(t − t0 )(t − t0 )) + iIm(e(t − t0 )(t − t0 )) =
( g(t0 ) + a(t − t0 ) + Re(e(t − t0 ))(t − t0 )) + i(h(t0 ) + b(t − t0 ) + Im(e(t − t0 ))(t − t0 )) .
g(t) = g(t0 ) + a(t − t0 ) + Re(e(t − t0 ))(t − t0 ) and h(t) = h(t0 ) + b(t − t0 ) + Im(e(t − t0 ))(t − t0 ) .
In order to conclude that g and h are differentiable at t0 with g0 (t0 ) = a and h0 (t0 ) = b , it
only remains for us to show that Re(e) and Im(e) are real epsilon functions. This follows
from
By the expression
f (t) = t + it2
a function f : R 7→ C is defined. Since the real part of f has the derivative 1 and the
imaginary part of f the derivative 2t we obtain from Theorem 1.60:
f 0 (t) = 1 + i2t , t ∈ R .
eNote 1 1.10 COMPLEX FUNCTIONS OF A REAL VARIABLE 45
Since cos0 (t) = − sin(t) and sin0 (t) = cos(t) , it is seen from Theorem 1.60 that
Proof
Let f 1 (t) = g1 (t) + i h1 (t) and f 2 (t) = g2 (t) + i h2 (t), where g1 , h1 , g2 and h2 are differentiable
real functions. Furthermore let c = a + ib be an arbitrary complex number in rectangular
form.
We then get from Theorem 1.60 and by the use of computational rules for derivatives for real
functions:
( f 1 + f 2 ) 0 ( t ) = ( g1 + g2 ) 0 ( t ) + i ( h 1 + h 2 ) 0 ( t )
= g10 (t) + g20 (t) + i h10 (t) + h20 (t)
= f 10 (t) + f 20 (t) .
We get from Theorem 1.60 and by the use of computational rules for derivatives for real
functions:
( c · f 1 ) 0 ( t ) = ( a g1 − b h 1 ) 0 ( t ) + i ( a h 1 + b g1 ) 0 ( t )
= a g10 (t) − b h10 (t) + i a h10 (t) + b g10 (t)
= c · f 10 (t) .
Exercise 1.64
Show that if f 1 and f 2 are differentiable complex functions of a real variable, then the function
f 1 − f 2 is differentiable with the derivative
We now return to functions of the type (1-40). First we give a useful theorem about their
conjugation.
eNote 1 1.10 COMPLEX FUNCTIONS OF A REAL VARIABLE 47
Theorem 1.65
For an arbitrary complex number c and every real number t:
ect = ec t . (1-46)
Proof
Let c = a + ib be the rectangular form of c . We then get by the use of Definition 1.41 and the
rules of computation for conjugation in Theorem 1.23:
ect = eat+ibt
= e at (cos(bt) + i sin(bt))
= e at (cos(bt) + i sin(bt))
= e at (cos(bt) − i sin(bt))
= e at (cos(−bt) + i sin(−bt))
= e at−ibt
= ect .
f : x 7→ ekx , x ∈ R ,
f 0 ( x ) = k f ( x ) = kekx . (1-47)
We end this eNote by showing that the complex exponential function of a real variable
satisfies a quite similar rule of differentiation.
eNote 1 1.10 COMPLEX FUNCTIONS OF A REAL VARIABLE 48
Proof
ect = eat+ibt
= e at (cos(bt) + i sin(bt))
= e at cos(bt) + i e at sin(bt) .
Thus we have
g0 (t) = aeat cos(bt) − eat b sin(bt) and h0 (t) = aeat sin(bt) + eat b cos(bt) ,
we now get
If c in Theorem 1.66 is real, (1-49) naturally only expresses the ordinary differentiation
of the real exponential function as in (1-47), as expected.
eNote 2 49
eNote 2
In this eNote complex polynomials of one variable are introduced. An elementary knowledge of
complex numbers is a prerequisite, and knowlege of real polynomials of one real variable is
recommended.
2.1 Introduction
Knowledge about the roots of polynomials is the main road to understanding their prop-
erties and efficient usage, and is therefore a major subject in the following. But first we
introduce some general properties.
eNote 2 2.1 INTRODUCTION 50
Definition 2.1
By a polynomial of degree n we understand a function that can be written in the form
P ( z ) = a n z n + a n −1 z n −1 + · · · + a 1 z + a 0 (2-1)
where a0 , a1 , . . . , an are complex constants with an 6= 0 , and z is a complex vari-
able.
If you multiply a polynomial by a constant, or when you add, subtract, multiply and
compose polynomials with each other, you get a new polynomial. This polynomial can
be simplified by gathering terms of the same degree and written in the form (2.1).
P ( z ) = a n z n + a n −1 z n −1 + · · · + a 1 z + a 0 . (2-2)
bn − 1 = a n (2-3)
bk = ak+1 + z0 · bk+1 for k = n−2, . . . , 0 , (2-4)
P ( z ) = ( z − z0 ) Q ( z ) (2-5)
Proof
Let the polynomial P be given as in the theorem, and let α be an arbitrary number. Consider
an arbitrary (n − 1)-degree polynomial
It is seen that the polynomials (z − α) Q(z) and P(z) have the same representation if we in
succession write the bk -coefficients for Q as given in (2-3) and (2-4), and if at the same time
the following is valid:
−αb0 = a0 ⇔ b0 α = − a0 .
We investigate whether this condition is satisfied by using (2-3) and (2-4) in the opposite
eNote 2 2.2 THE ROOTS OF POLYNOMIALS 53
order:
b0 α = ( a1 + αb1 )α = b1 α2 + a1 α
= ( a2 + αb2 )α2 + a1 α = b2 α3 + a2 α2 + a1 α
..
.
= bn − 1 α n + a n − 1 α n − 1 + · · · + a 2 α 2 + a 1 α
= a n α n + a n −1 α n −1 + · · · + a 2 α 2 + a 1 α = − a 0
⇔ P ( α ) = a n α n + a n −1 α n −1 + · · · + a 2 α 2 + a 1 α + a 0 = 0 .
It is seen that the condition is only satisfied if and only if α is a root of P . By this the proof is
complete.
Given the polynomial P(z) = 2z4 − 12z3 + 19z2 − 6z + 9 . It is seen that 3 is a root
since P(3) = 0 . Determine a third-degree polynomial Q such that
P ( z ) = ( z − 3) Q ( z ) .
b3 = a4 = 2
b2 = a3 + 3b3 = −12 + 3 · 2 = −6
b1 = a2 + 3b2 = 19 + 3 · (−6) = 1
b0 = a1 + 3b1 = −6 + 3 · 1 = −3 .
We conclude that
Q(z) = 2z3 − 6z2 + z − 3
so
P(z) = (z − 3) (2z3 − 6z2 + z − 3) .
When a polynomial P with the root z0 is written in the form P(z) = (z − z0 ) Q1 (z) ,
where Q1 is a polynomial, it is possible that z0 is also a root of Q1 . Then Q1 can
eNote 2 2.2 THE ROOTS OF POLYNOMIALS 54
Proof
Assume that α is a root of P , and that (contrary to the statement in the theorem) there exist
two different factorisations
P ( z ) = ( z − α )r R ( z ) = ( z − α ) s S ( z )
where r > s , and R(z) and S(z) are polynomials of which α is not a root. We then get
(z − α)r R(z) − (z − α)s S(z) = (z − α)s (z − α)k R(z) − S(z) = 0 , for all z ∈ C
Since both the left-hand and the right-hand sides are continuous functions, they must have
the same value at z = α . From this we get that
eNote 2 2.2 THE ROOTS OF POLYNOMIALS 55
Example 2.9
where 3 is a root. But 3 is also a root of the factor 2z3 − 6z2 + z − 3 . By using the theorem
of descent, Theorem 2.6, on this polynomial we get
Now we have started a process of descent! How far can we get along this way? To con-
tinue this investigation we will need a fundamental result, viz the Fundamental Theorem.
A decisive reason for the introduction of complex numbers is that every (complex) poly-
nomial has a root in the set of complex numbers. This result was proven by the math-
ematician Gauss in his ph.d.-dissertation from 1799 . The proof of the theorem is de-
manding, and Gauss strove all his life to refine his proof more. Four versions of the
proof by Gauss exist, so there is no doubt that he put a lot of emphasis on this theorem.
Here we take the liberty to state Gauss’ result without proof:
The polynomial P(z) = z2 + 1 has no roots within the set of real numbers. But
within the set of complex numbers it has two roots i and −i because
The road from the fundamental theorem of algebra until full knowledge of the number
of roots is not long. We only have to develop the ideas put forward in the theorem of
descent futher.
P ( z ) = ( z − α1 ) Q1 ( z ) (2-7)
Q1 ( z ) = ( z − α2 ) Q2 ( z )
• First all the n numbers α1 , . . . , αn that are listed in (2-8), are roots of P since
substitution into the formula gives the value 0 .
• The second thing we notice is that P cannot have other roots than the n given
ones. That there cannot be more roots is easily seen as follows: If an arbitrary
number α 6= αk , k = 1, . . . , n , is inserted in place of z in (2-8), all factors on the
right-hand side of (2-8) will be different from zero. Hence their product will also
be different form zero. Therefore P(α) 6= 0 , and α is not a root of P .
• The last thing we notice in (2-8), is that the roots are not necessarily different. If
z1 , z2 , . . . , z p are the p different roots of P , and mk is the multiplicity of zk , k =
1, . . . , p , then the completely factorized form (2-8) can be simplified as follows
P ( z ) = a n ( z − z 1 ) m1 ( z − z 2 ) m2 · · · ( z − z p ) m p (2-9)
m1 + m2 · · · + m p = n .
eNote 2 2.2 THE ROOTS OF POLYNOMIALS 57
According to the preceding arguments we can now present the fundamental theorem of
algebra in the extended form.
where α and β are roots of P . If α 6= β , P has two different roots, both with algebraic
multiplicity 1 . If α = β , P has only one root with the algebraic multiplicity 2 . The roots are
then denoted a double root.
P ( z ) = 7( z − 1)2 ( z + 4)3 ( z − 5) .
We see that P has three different roots: 1, −4 and 5 with the algebraic multiplicities 2 , 3
and 1 , respectively.
We notice that the sum of the algebraic multiplicities is 6 which equals the degree of P in
concordance with the Fundamental Theorem of Algebra — Version 2.
P has only one root z = 0 . The algebraic multiplicity of this root is 3. One says that 0 is a
eNote 2 2.3 IDENTICAL POLYNOMIALS 58
Two polynomials P and Q are equal (as functions of z) if P(z) = Q(z) for all z . But
what does it take for two polynomials to be equal? Is it possible that a fourth-degree and
a fifth-degree polynomial take on the same value for all variables as long as you choose
the right coefficients? This is not the case as is seen from the following theorem.
Proof
We consider two arbitrary polynomials P og Q . If they are of the same degree, and all
the coefficients for terms of the same degree are equal, they must have the same value for
all variables and hence they are identical. This proves the first direction of the theorem of
identity.
Assume hereafter that P og Q are identical as functions of z, but that not all coefficients for
terms of the same degree from the two polynomials are equal. We assume further that P has
the degree n and Q the degree m where n ≥ m . Let ak be the coefficients for P and let bk
be the coefficients for Q , and consider the difference polynomial
where we for the case n > m put bk = 0 for m < k ≤ n . We note that the 0-degree
coefficient ( a0 − b0 ) cannot be the only coefficient of R(z) that is different from 0, since this
would make P(0) − Q(0) = ( a0 − b0 ) 6= 0 which contradicts that P and Q are identical as
functions. Therefore the degree of R is greater than or equal to 1. On the other hand (2-10)
shows that the degree of R at the most is n . Now let zk , k = 1, . . . , n + 1 , be n + 1 different
eNote 2 2.4 POLYNOMIAL EQUATIONS 59
This contradicts the fundamental theorem of algebra – version 2, Theorem 2.11: R cannot
have a number of roots that is higher than its degree. The assumption, that not all coefficients
of terms of the same degree from P and Q are equal, must therefore be wrong. From this
it also follows that P and Q have the same degree. By this the second part of the identity
theorem is proven.
The equation
3 z2 − z + 4 = a z2 + b z + c
is satisfied for all z exactly when a = 3, b = −1 og c = 4 .
In the following section we treat methods of finding roots of certain types of polynomi-
als.
From the fundamental theorem of algebra, Theorem 2.10, we know that every polyno-
mial of degree greater than or equal to 1 has roots. Moreover, in the extended version,
Theorem 2.11, it is maintained that for every polynomial the degree is equal to the num-
ber of roots if the roots are counted with multiplicity. But the theorem is a theoretical
theorem of existence that does not help in finding the roots.
In the following methods for finding the roots of simple polynomials are introduced.
eNote 2 2.4 POLYNOMIAL EQUATIONS 60
But let us keep the level of ambition (safely) low, because in the beginning of the 170 th
century the Norwegian algebraicist Abel showed that one cannot establish general meth-
ods for finding the roots of arbitrary polynomials of degree larger than four!
For polynomials of higher degree than four a number of smart tricks exist by which one
can successfully find a single root. Hereafter one descends to a polynomial of lower de-
gree — and successively decends to a polynomial of fourth degree or lower for which
one can find the remaining roots.
Let us at the outset maintain that when you want to find the roots of a polynomial
P(z) , you should solve the corresponding polynomial equation P(z) = 0 . As a simple
illustration we can look at the root of an arbitrary first-degree polynomial:
P(z) = az + b .
az + b = 0 .
b
this is not difficult. It has the solution z0 = − which therefore is a root of P(z) .
a
P(z) = (1 − i) z − (5 + 2i) .
(1 − i) z − (5 + 2i) = 0 ⇔ (1 − i) z = (5 + 2i) .
3 7
Hence the equation has the solution z0 = + i that also is the root of P.
2 2
eNote 2 2.4 POLYNOMIAL EQUATIONS 61
For binomial equations an explicit solution formula exists, which we present in the fol-
lowing theorem.
w = |w| eiv .
Proof
p v 2π
For every p ∈ {0, 1, . . . , n − 1} z p = n
|w| ei( n + p n ) is a solution to (2-11), since
q n
n i( nv + p 2π )
(z p ) = n
| w |e n = |w| ei(v+ p 2π ) = |w| eiv = w.
It is seen that the n solutions viewed as points in the complex plane all lie on a circle with
centre at z = 0, radius n |w| and a consecutive angular distance of 2π
p
n . In other words the
eNote 2 2.4 POLYNOMIAL EQUATIONS 62
connecting lines between z = 0 and the solutions divide the circle in n angles of the same
size.
From this it follows that all n solutions are mutually different. That there are no more solu-
tions is a consequence of the fundamental theorem of algebra – version 2, Theorem 2.11. By
this the theorem is proven.
In the next examples we will consider some important special cases of binomial equa-
tions.
We consider a complex number in the exponential form w = |w| eiv . It follows from (2-12)
that the quadratic equation
z2 = w
has two solutions q q
v v
z0 = |w| ei 2 and z1 = − | w | ei 2 .
z2 = −r
Sometimes the method used in Example 2.21 can be hard to carry out. In the following
example we show an alternative method.
eNote 2 2.4 POLYNOMIAL EQUATIONS 63
Since we expect the solution to be complex we put z = x + iy where x and y are real numbers.
If we can find x and y, then we have found the solutions for z. Therefore we have z2 =
( x + iy)2 = x2 − y2 + 2xyi and we see that (2-13) is equivalent to
x2 − y2 + 2xyi = 8 − 6i .
Since a complex equation is true exactly when both the real parts and the imaginary parts of
the right-hand and the left-hand sides of the equation are identical, (2-13) is equivalent to
−6 3
If we put y = = − in x2 − y2 = 8 , and put x2 = u , we get a quadratic equation that
2x x
can be solved:
2
2 3 9
x − − = 8 ⇔ x2 − =8⇔
x x2
2 9
x − 2 x2 = 8x2 ⇔ x4 − 9 = 8x2 ⇔ x4 − 8x2 − 9 = 0 ⇔
x
u2 − 8u − 9 = 0 ⇔ u = 9 or u = −1 .
The equation x2 = u = 9 has the solutions x1 = 3 and x2 = −3, while the equation
x2 = u = −1 has no solution, since x and y are real numbers. If we put x1 = 3 respective
x2 = −3 in (2-14), we get the corresponding y-values y1 = −1 and y2 = 1 .
From this we conclude that the given equation (2-13) has the roots
For the solution of quadratic equations we state below the formula that corresponds
to the well-known solution formula for real quadratic equations. There is a single de-
viation, viz. we do not compute the square-root of the discriminant since we in this
theorem do not presuppose knowledge of square-roots of complex numbers.
az2 + bz + c = 0 , a 6= 0 (2-15)
− b − w0 − b + w0
z1 = og z2 = (2-16)
2a 2a
where w0 is a solution to the binomial equation of the second degree w2 = D .
−b
If in particular D = 0 , we have that z1 = z2 = .
2a
eNote 2 2.4 POLYNOMIAL EQUATIONS 65
Proof
Example 2.25 Real Quadratic Equation with a Positive Value of the Dis-
criminant
2z2 + 5z − 3 = 0 .
D = 52 − 4 · 2 · (−3) = 49 .
−5 + 7 1 −5 − 7
z1 = = and z2 = = −3 . (2-17)
2·2 2 2·2
z2 − 2z + 5 = 0 .
D = (−2)2 − 4 · 1 · 5 = −16 .
According to Example 2.22 the solution to the binomial equation of the second degree w2 =
D = −16 is given by w0 = 4i . Now the solutions can be computed as:
−(−2) + 4i −(−2) − 4i
z1 = = 1 + 2i and z2 = = 1 − 2i . (2-18)
2·1 2·1
z2 − (1 + i)z − 2 + 2i = 0 . (2-19)
First we identify the coefficients: a = 1, b = −(1 + i), c = −2 + 2i , and we find the discrimi-
nant:
D = (−(1 + i))2 − 4 · 1 · (−2 + 2i) = 8 − 6i .
From Example 2.23 we know that the solution to the binomial equation w2 = D = 8 − 6i is
w0 = 3 − i . From this we find the solution to (2-19) as
From antiquity geometrical methods for the solution of (real) quadratic equations are
known. But not until A.D. 800 did algebraic solution formulae became known, through
the work (in Arabic) of the Persian mathematician Muhammad ibn Musa al-Khwarismes
famous book al-Jabr. In the West the name al-Khwarisme became the well-known word
algorithm, while the book title became algebra.
Three centuries later history repeated itself. Around A.D. 1100 another Persian math-
ematician (and poet) Omar Khayyám gave exact methods on how to find solutions to
real equations of the third and fourth degree by use of advanced geometrical methods.
As an example he solved the equation x3 + 200x = 20x2 + 2000 by intersecting a circle
with a hyperbola the equations of which he could derive from the equation of third de-
gree.
Omar Khayyám did not think it possible to draw up algebraic formulae for solutions to
equations of degree greater than two. He was proven wrong by the Italian Gerolamo
Cardano who in the 16th century published formulae for the solution of Equations of
the third and fourth degree.
Khayyáms methods and Cardanos formulae are beyond the scope of this eNote. Here
we only give — see the previous Example 2.9 and the following Example 2.28 — a few
examples by use of the “method of descent”, Theorem 2.6, on how one can find all so-
lutions to equations of degree greater that two if one in advance knows or can guess a
sufficient number of the solutions.
z3 − 3z2 + 7z − 5 = 0 .
It is easily guessed that 1 is a solution. By use of the algorithm of descent one easily gets the
factorization:
z3 − 3z2 + 7z − 5 = (z − 1)(z2 − 2z + 5) = 0 .
We know that 1 is a solution, the remaining solutions are found by solving the quadratic
equation
z2 − 2z + 5 = 0 ,
which, according to Example 2.26, has the solutions 1 + 2i and 1 − 2i .
eNote 2 2.5 REAL POLYNOMIALS 68
The theory that has been unfolded in the previous section applies to all polynomials
with complex coefficients. In this section we present two theorems that only apply to
polynomials with real coefficients — that is the subset called the real polynomials. The
first theorem shows that non-real roots always appear in pairs.
Proof
Let
P ( z ) = a n z n + a n −1 z n −1 + · · · + a 1 z + a 0
be a real polynomial. By use of the arithmetic rules for conjugation of the sum and prod-
uct of complex numbers (see eNote 1 about complex numbers) with the condition that all
coefficients are real, we get
P ( z ) = a n z n + a n −1 z n −1 + · · · + a 1 z + a 0
= a n z n + a n −1 z n −1 + · · · + a 1 z + a 0
= P(z) .
If z0 is a root of P , we get
P ( z0 ) = 0 = 0 = P ( z0 )
from which it is seen that z0 is also a root. Thus the theorem is proven.
eNote 2 2.5 REAL POLYNOMIALS 69
has the root 2 − 3i . Determine all roots of P , and write P in a complete factorized
form.
We see that all the three coefficients in P are real. Therefore the conjugate of the given root
2 + 3i is also a root of P . Since P is a quadratic polynomial, there are no more roots.
From Theorem 2.29 we know that complex roots always are present in conjugated pairs.
This leads to the following theorem:
Given that a real polynomial of seventh degree P has the roots 1, i, 1 + 2i as well
as the double root −2 , and that the coefficient to its term of the highest degree is
a7 = 5 . Write P as a product of real linear and real quadratic polynomials without
real roots.
We use the fact that the conjugates of the complex roots are also roots and write P in its
complete factorized form:
Two pairs of factors correspond to conjugated roots. When we multiply these we obtain the
form we wanted:
P(z) = 5 (z − 1)(z2 + 1)(z2 − 2z + 5)(z − 2)2 .
eNote 3
Elementary Functions
In this eNote we will both repeat some of the basic properties for a selection of the (from high
school) well-known functions f ( x ) of one real variable x, and introduce some new functions,
which typically occur in a variety of applications. The basic questions concerning any function
are usually the following: How, and for which values of x, is the function defined? Which
values for f ( x ) do we get when we apply the functions to the x-elements in the domain? Is the
function continuous? What is the derivative f 0 ( x ) of the function – if it exists? As a new
concept, we will introduce a vast class of functions, the epsilon functions, which are denoted
by the common symbol ε( x ) and which we will use generally in order to describe continuity and
differentiability – also of functions of more variables, which we introduce in the following
eNotes.
In the description of a real function f ( x ) both the real numbers x where the function is
defined and the values that are obtained by applying the function on the domain are
stated. The Domain we denote D ( f ) and the range, or image, we denote R( f ).
Here are domains and the corresponding ranges for some well-known functions.
Figure 3.1: The well-known exponential function e x = exp( x ) and the natural loga-
rithmic function ln( x ). The red circles on the negative x-axis and at 0 indicate that the
logarithmic function is not defined on ] − ∞, 0].
eNote 3 3.1 DOMAIN AND RANGE 73
The function f 8 ( x ) in Example 3.1 is defined using | x |, which denotes the ab-
solute value of x, i.e.
x>0, for x > 0
|x| = 0, for x = 0 (3-2)
− x > 0 , for x < 0 .
The function
sin( x )
f ( x ) = tan( x ) = (3-3)
cos( x )
has the domain D ( f ) = R \ A, A denoting those real numbers x for which cos( x ) = 0, cos( x )
being the denominator, i.e.
Figure 3.2: The graphs for the functions tan( x ) and cot( x ).
eNote 3 3.1 DOMAIN AND RANGE 74
Exercise 3.3
cos( x )
g( x ) = cot( x ) = (3-5)
sin( x )
Determine the domain for g( x ) and state it in the same way as above for tan( x ), see Figure
3.2.
A function f ( x ) that is not defined for all real numbers can easily be extended to a func-
tion fb( x ), which has D ( fb) = R. One way of doing this is by the use of a curly bracket
in the following way:
Definition 3.4
Given a function f ( x ) with D ( f ) 6= R. We then define the 0-extension of f ( x ) by:
f ( x ) , for x ∈ D ( f )
f (x) = (3-6)
for x ∈ R \ D ( f ) .
b
0,
It is evident that depending on the application one can seal and extend the do-
main for f ( x ) in many other ways than choosing the constant 0 as the value for
the extended function at the points where the original function is not defined.
Naturally, the Range R( fb) for the 0-extended function is the original range for
f ( x ) united with 0, i.e. R( fb) = R( f ) ∪ {0} .
Hereafter we will assume – unless otherwise stated – that the functions we consider are
defined for all R possibly by extension as above.
eNote 3 3.2 EPSILON FUNCTIONS 75
We introduce a special class of functions, which we will use in order to define the im-
portant concept of continuity.
ε 1 (x) = x
ε 2 (x) = |x|
(3-8)
ε 3 ( x ) = ln(1 + x )
ε 4 ( x ) = sin( x ) .
The quality ’to be an epsilon function’ is rather stable: The product of an ep-
silon function and an arbitrary other function that only has to be bounded is
also an epsilon function. The sum and the product of two epsilon functions are
again epsilon functions. The absolute value of an epsilon function is an epsilon
function.
eNote 3 3.3 CONTINUOUS FUNCTIONS 76
Functions that are 0 in other places than x = 0 can also be epsilon functions:
Exercise 3.7
Show that the 0-extension fb8 ( x ) of the function f 8 ( x ) = | x |/x is not an epsilon function. Hint:
If we choose k = 10 then clearly there does not exist a value of K such that
1 1
| f 8 ( x )| = | | x |/x | = 1 < , for all x with |x| < . (3-9)
10 K
Draw the graph for fb8 ( x ). This cannot be drawn without ’lifting the pencil from the paper’!
Exercise 3.8
Show that the 0-extension of the function f ( x ) = sin(1/x ) is not an epsilon function.
f ( x ) = f ( x0 ) + ε f ( x − x0 ) . (3-10)
Note that even though it is clear what the epsilon function precisely is in the
definition 3.9, viz. f ( x ) − f ( x0 ), then the only property in which we are in-
terested is the following: ε f ( x − x0 ) → 0 for x → x0 such that f ( x ) → f ( x0 )
for x → x0 , that is precisely as we know the concept of continuity from high
school!
Exercise 3.10
According to the above, all epsilon functions are continuous at x0 = 0 (with the value 0 at
x0 = 0). Construct an epsilon function that is not continuous at any of the points x0 = 1/n
where n = 1, 2, 3, 4, · · · .
Exercise 3.11
Show that the 0-extension fb( x ) of the function f ( x ) = | x − 7|/( x − 7) is not continuous on
R.
f ( x ) = f ( x0 ) + a · ( x − x0 ) + ( x − x0 ) · ε f ( x − x0 ) . (3-11)
f ( x ) = f ( x0 ) + f 0 ( x0 ) · ( x − x0 ) + ( x − x0 ) · ε f ( x − x0 ) . (3-12)
d
f 0 (x) = f (x) . (3-13)
dx
We will show that there is only one value of a that fulfills Equation (3-11). Assume
that two different values, a1 and a2 both fulfill (3-11) possibly with two different
epsilon functions:
f ( x ) = f ( x0 ) + a1 · ( x − x0 ) + ( x − x0 ) · ε 1 ( x − x0 )
(3-14)
f ( x ) = f ( x0 ) + a2 · ( x − x0 ) + ( x − x0 ) · ε 2 ( x − x0 ) .
By subtracting (3-14) from the uppermost equation we get:
0 = 0 + ( a1 − a2 ) · ( x − x0 ) + ( x − x0 ) · (ε 1 ( x − x0 ) − ε 2 ( x − x0 )) , (3-15)
such that
a2 − a1 = ε 1 ( x − x0 ) − ε 2 ( x − x0 ) (3-16)
for all x 6= x0 – and clearly this cannot be true; the right hand side tends towards
0 when x tends towards x0 ! Therefore the above assumption, i.e. that a1 6= a2 , is
eNote 3 3.4 DIFFERENTIABLE FUNCTIONS 79
wrong. The two constants a1 and a2 must be equal, and this is what we should
realize.
The definition above is quite equivalent to the one we know from high school.
If we first subtract f ( x0 ) from both sides of the equality sign in Equation (3-12)
and then divide by ( x − x0 ) we get
f ( x ) − f ( x0 )
= f 0 ( x0 ) + ε f ( x − x0 ) → f 0 ( x0 ) for x → x0 , (3-17)
x − x0
i.e. the well-known limit value for the quotient between the increment in the
function f ( x ) − f ( x0 ) and the x-increment x − x0 . The reason why we do not
apply this known definition of f 0 ( x0 ) is simply that for functions of more vari-
ables the quotient does not make sense – but more about this in a later eNote.
Proof
We have that
f ( x ) = f ( x0 ) + f 0 ( x0 ) · ( x − x0 ) + ( x − x0 ) ε f ( x − x0 )
(3-18)
= f ( x0 ) + f 0 ( x0 ) · ( x − x0 ) + ( x − x0 ) ε f ( x − x0 ) ,
and since the function in the square brackets on the right hand side is an epsilon function of
( x − x0 ) then f ( x ) is continuous at x0 .
But the opposite is not valid – here is an example that shows this:
eNote 3 3.4 DIFFERENTIABLE FUNCTIONS 80
f ( x ) = f ( x0 ) + a · ( x − x0 ) + ( x − x0 ) ε f ( x − x0 ). (3-19)
Definition 3.16
The first degree approximating polynomial for f ( x ) expanded about the point x0 is
defined by:
P1,x0 ( x ) = f ( x0 ) + f 0 ( x0 ) · ( x − x0 ) . (3-22)
Note that P1,x0 ( x ) really is a first degree polynomial in x. The graph for the
function P1,x0 ( x ) is the tangent to the graph for f ( x ) at the point ( x0 , f ( x0 )),
see Figure 3.3. The equation for the tangent is y = P1,x0 ( x ), thus y =
f ( x0 ) + f 0 ( x0 ) · ( x − x0 ). The slope of the tangent is clearly α = f 0 ( x0 ) and
the tangent intersects the y-axis at the point (0, f ( x0 ) − x0 · f 0 ( x0 )). Later we
will find out how we can approximate with polynomials of higher degree n,
i.e. polynomials that are then denoted Pn,x0 ( x ).
Even though this formula is rather well known from high school we shall give a short
sketch of a proof – to illustrate the use of epsilon functions.
eNote 3 3.4 DIFFERENTIABLE FUNCTIONS 82
Proof
f ( x ) = f ( x0 ) + f 0 ( x0 ) · ( x − x0 ) + ( x − x0 ) ε f ( x − x0 )
(3-24)
g ( x ) = g ( x0 ) + g 0 ( x0 ) · ( x − x0 ) + ( x − x0 ) ε g ( x − x0 ) ,
h( x ) = f ( x ) · g( x )
(3-25)
= f ( x0 ) · g( x0 ) + ( f 0 ( x0 ) · g( x0 ) + f ( x0 ) · g0 ( x0 )) · ( x − x0 ) + ( x − x0 )ε h ( x − x0 ) ,
where we have used ( x − x0 )ε h ( x − x0 ) as short for the remaining part of the product sum.
Furthermore any of the addends in the remaining part contains the factor ( x − x0 )2 or the
product of ( x − x0 ) with an epsilon function and therefore can be written in the stated form.
But then the product formula follows directly from the factor in front of ( x − x0 ) in Equation
(3-25):
h 0 ( x0 ) = f 0 ( x0 ) · g ( x0 ) + f ( x0 ) · g 0 ( x0 ) . (3-26)
The following differentiation rule is also well known from high school:
Exercise 3.19
Use the epsilon function argument in the same way as in the differentiation rule for a product
to show Equation 3.18.
h0 ( x0 ) = f 0 ( g( x0 )) · g0 ( x0 ) (3-28)
Proof
We exploit that the two functions f ( x ) and g( x ) are differentiable. In particular g( x ) is dif-
ferentiable at x0 :
g( x ) = g( x0 ) + g0 ( x0 )( x − x0 ) + ( x − x0 ) · ε g ( x − x0 ) , (3-29)
and the function f (u) is differentiable at u0 = g( x0 ):
f (u) = f (u0 ) + f 0 (u0 )(u − u0 ) + (u − u0 ) · ε f (u − u0 ) . (3-30)
From this we get, setting u = g( x ) and u0 = g( x0 ):
h( x ) = f ( g( x ))
= f ( g( x0 )) + f 0 ( g( x0 ))( g( x ) − g( x0 ) + ( g( x ) − g( x0 ) · ε f ( g( x ) − g( x0 )
= h( x0 ) + f 0 ( g( x0 ))( g0 ( x0 )( x − x0 ) + ( x − x0 ) · ε g ( x − x0 )) (3-31)
0
+ ( g ( x0 )( x − x0 ) + ( x − x0 ) · ε g ( x − x0 )) · ε f ( g( x ) − g( x0 )
= h( x0 ) + f 0 ( g( x0 )) g0 ( x0 ) · ( x − x0 ) + ( x − x0 ) · ε h ( x − x0 ) ,
from which we directly read that h0 ( x0 ) = f 0 ( g( x0 )) g0 ( x0 ) – because this is exactly the unique
coefficient of ( x − x0 ) in the above expression.
eNote 3 3.5 INVERSE FUNCTIONS 84
Exercise 3.21
f 0 ( g( x0 )) · ε g ( x − x0 ) + ( g0 ( x0 ) + ·ε g ( x − x0 )) · ε f ( g( x ) − g( x0 )) (3-32)
is an epsilon function, which we accordingly can call (and have called) ε h ( x − x0 ). Consider
why this is entirely OK.
Exercise 3.22
Find the derivatives of the following functions for every x-value in their respective domains:
f 1 ( x ) = ( x2 + 1) · sin( x )
f 2 ( x ) = sin( x )/( x2 + 1) (3-33)
2
f 3 ( x ) = sin( x + 1) .
The exponential function exp( x ) and the logarithmic function ln( x ) are inverse func-
tions to each other – as is well known the following is valid:
Note that even though exp( x ) is defined for all x, the inverse function ln( x ) is
only defined for x > 0 – and vice versa (!).
inverse function g( x ) maps the interval B one-to-one onto the interval A such that:
√
f ( g( x )) = ( x )2 = x for x ∈ B = [0, ∞[
√ (3-35)
g( f ( x )) = x2 = x for x ∈ A = [0, ∞[ .
f ◦−1 ( f ( x )) = x for x ∈ A ⊂ D( f )
◦−1
(3-37)
f(f ( x )) = x for x ∈ B ⊂ D ( f ◦−1 ) .
We use here the symbol f ◦−1 ( x ) in order to avoid confusion with ( f ( x ))−1 =
1/ f ( x ). However the reader should note that the standard notation is simply
f −1 for the inverse function. The graph for the inverse function g( x ) = f ◦−1 ( x )
to a function f ( x ) can be obtained by mirroring the graph for f ( x ) in the diag-
onal in the first quadrant in the ( x, y)-coordinate system – i.e. the line with the
equation y = x – see Figure 3.4.
eNote 3 3.5 INVERSE FUNCTIONS 86
Figure 3.4: The graph for a function f ( x ) and the graph for the inverse function g( x ).
It is valid that g( x ) = f ◦−1 ( x ) and f ( x ) = g◦−1 ( x ), but they each have their own
definition intervals.
1
( f ◦−1 )0 ( x0 ) = (3-38)
f 0 ( f ◦−1 ( x 0 ))
eNote 3 3.6 HYPERBOLIC FUNCTIONS 87
Proof
h( x ) = f ( f ◦−1 ( x )) = x , (3-39)
The names cosh( x ) and sinh( x ) (often spoken as “cosh” and “sinsh”) look like cos( x )
and sin( x ), but the functions are very different, as we shall demonstrate below.
Yet there are also fundamental structural similarities between the two pairs of functions
and this is what motivates the names. In the system of differential equations for cos( x )
eNote 3 3.6 HYPERBOLIC FUNCTIONS 88
and sin( x ) only a single minus sign separates this from (3-41):
In addition (again with the decisive minus sign as the only difference) the following
simple analogy to the well-known and often used relation cos2 ( x ) + sin2 ( x ) = 1 applies:
Proof
Make the derivative with respect to x on both sides of the equation (3-43) and conclude that
cosh2 ( x ) − sinh2 ( x ) is a constant. Finally use the initial conditions.
Exercise 3.27
Show directly from the system of differential equations (3-41) that the two ”new” functions
are in fact not so new:
e x + e− x
cosh( x ) = , D (cosh) = R , R(cosh) = [1, ∞[
2 (3-44)
e x − e− x
sinh( x ) = , D (sinh) = R , R(sinh) = ] − ∞, ∞[
2
eNote 3 3.6 HYPERBOLIC FUNCTIONS 89
Exercise 3.28
Exercise 3.29
The graph for the function f ( x ) = cosh( x ) looks a lot like a parabola, viz. the graph for the
function g( x ) = 1 + ( x2 /2) when we plot both functions on a suitably small interval around
x0 = 0. Try this! If we instead plot the two graphs in very large x-interval, we learn that
the two functions have very different graphical behaviours. Try this, i.e. try to plot both
functions on the interval [−50, 50]. Comment upon and explain the qualitative differences.
Similarly compare the two functions sinh( x ) and x + ( x3 /6) in the same way.
It is natural and useful to define hyperbolic analogies to tan( x ) and cot( x ). This is done
as follows:
eNote 3 3.6 HYPERBOLIC FUNCTIONS 90
sinh( x ) e2x − 1
tanh( x ) = = 2x , D (tanh) = R , R(tanh) = ] − 1, 1[
cosh( x ) e +1
(3-46)
cosh( x ) e2x + 1
coth( x ) = = 2x , D (coth) = R − {0} ,
sinh( x ) e −1
R(coth) = ] − ∞, −1[ ∪ ]1, ∞[ .
The derivatives of cosh( x ) and of sinh( x ) are already given by the defining system in
(3-41).
d
cosh( x ) = sinh( x )
dx
d
sinh( x ) = cosh( x )
dx
d 1 (3-47)
tanh( x ) = = 1 − tanh2 ( x )
dx cosh2 ( x )
d −1
coth( x ) = 2
= 1 − coth2 ( x ) .
dx sinh ( x )
eNote 3 3.7 THE AREA FUNCTIONS 91
Exercise 3.31
Show the last two expressions for the derivatives for tanh( x ) and coth( x ) in (3-47) by the use
of the differentiation rule in Theorem 3.18.
The inverse functions to the hyperbolic functions are called area functions and are
named cosh◦−1 ( x ) = arcosh( x ), sinh◦−1 ( x ) = arsinh( x ), tanh◦−1 ( x ) = artanh( x ), and
coth◦−1 ( x ) = arcoth( x ), respectively.
Since the functions cosh( x ), sinh( x ), tanh( x ), and coth( x ) all can be expressed in terms
of exponential functions it is no surprise that the inverse functions and their derivatives
can be expressed by logarithmic functions. We gather the information here:
p
arcosh( x ) = ln( x + x2 − 1) for x ∈ [1, ∞[
p
arsinh( x ) = ln( x + x2 + 1) for x ∈ R
1 1+x (3-48)
artanh( x ) = ln for x ∈ ] − 1, 1[
2 1−x
x−1
1
arcoth( x ) = ln for x ∈ ] − ∞, 1[ ∪ ]1, ∞[ .
2 x+1
d 1
arcosh( x ) =√ for x ∈]1, ∞[
dx x2 − 1
d 1
arsinh( x ) =√ for x ∈ R
dx x2 + 1 (3-49)
d 1
artanh( x ) = for x ∈ ] − 1, 1[
dx 1 − x2
d 1
arcoth( x ) = for x ∈ ] − ∞ 1[ ∪ ]1, ∞[ .
dx 1 − x2
The inverse functions to the trigonometric functions are a bit more complicated. As
mentioned earlier here we must choose for each trigonometric function an interval
eNote 3 3.8 THE ARC FUNCTIONS 92
where the function in question is monotonic. In return, once we have chosen such an
interval, it is clear how the inverse function should be defined and how it should then
be differentiated. The inverse functions to cos( x ), sin( x ), tan( x ), and cot( x ) are usu-
ally written arccos( x ), arcsin( x ), arctan( x ), and arccot( x ), respectively; their names are
arccosine, arcsine, arctangent, and arccotangent. As above we gather the results here:
Note that the derivatives for arccos( x ) and arcsin( x ) are not defined at x0 = 1
or at x0 = −1. This is partly because, if the function we consider is only defined
on a bounded interval then we cannot say that the function is differentiable
at the end-points of the interval. Moreover the formulas for arccos0 ( x ) and
arcsin0 ( x ) show that they are not defined at x0 = 1 or x0 = −1; these values
give 0 in the denominators.
eNote 3 3.8 THE ARC FUNCTIONS 93
Exercise 3.32
Use a suitable modification of arctan( x ) in order to determine a new differentiable (and hence
continuous) function f ( x ) that looks like the 0-extension of | x |/x (which is neither continuous
nor differentiable), i.e. we want a function f ( x ) with the following properties: 1 > f ( x ) >
0.999 for x > 0.001 while −0.999 > f ( x ) > −1 for x < −0.001. See Figure 3.10. Hint: Try to
plot arctan(1000x ).
eNote 3 3.8 THE ARC FUNCTIONS 94
Figure 3.9: Arccosine and arcsine. Again the red circles indicate that the arc-functions
are not defined outside the interval [−1, 1]. Similarly the green circular disks indicate
that the arc-functions are defined at the end-points x = 1 and x = −1.
3.9 Summary
We have treated some of the fundamental properties of some well-known and some
not so well-known functions. How are they defined, what are their domains, are they
continuous, are they differentiable, and if so what are their derivatives?
f ( x ) = f ( x0 ) + f 0 ( x0 )( x − x0 ) + ( x − x0 )ε f ( x − x0 ) .
d
( f ( x ) · g( x )) = f 0 ( x ) · g( x ) + f ( x ) · g0 ( x ) . (3-53)
dx
f 0 (x) f ( x ) · g0 ( x ) f 0 ( x ) · g( x ) − f ( x ) · g0 ( x )
d f (x)
= − = . (3-54)
dx g( x ) g( x ) g2 ( x ) g2 ( x )
d
f ( g( x )) = f 0 ( g( x )) · g0 ( x ) . (3-55)
dx
eNote 4
In eNotes ?? and ?? it is shown how functions of one and two variables can be approximated by
first-degree polynomials at every (development) point and that the graphs for the approximating
first-degree polynomial are exactly the tangents and the tangent planes, respectively, for the
corresponding graphs of the functions. In this eNote we will show how the functions can be
approximated even better by polynomials of higher degree, so if the approximation to a function
is sufficiently good then one can use and continue the computations with the approximation
polynomial in place of the function itself and hope for a sufficiently small error. But what does it
mean that the approximation and the error are sufficiently good and sufficiently small? And
how does this depend on the degree of the approximating polynomial? You will find the answers
to these questions in this eNote.
(Updated: 22.09.2021 David Brander).
First we consider functions f ( x ) of one variable x on an open interval of the real num-
bers. We will also assume that the functions can be differentiated an arbitrary number of
times, that is, all the derivatives exist for every x in the interval: f 0 ( x0 ), f 00 ( x0 ), f 000 ( x0 ),
f (4) ( x0 ), f (5) ( x0 ), etc. where f (4) ( x0 ) means the 4th derivative of f ( x ) in x0 . These
higher order derivatives we will use in the construction of (the coefficients to) the ap-
proximating polynomials.
eNote 4 4.1 HIGHER ORDER DERIVATIVES 97
Definition 4.1
If a function f ( x ) can be differentiated an abitrary number of times at every point x
in a given open interval I we say that the function is smooth on the interval I.
Note that
f (n) ( x ) = n · (n − 1) · (n − 2) · · · 2 · 1 = n! , (4-2)
Where n! (n factorial) is the short way of writing the product of the natu-
ral numbers from and including 1 to and including n, cf. Table 4.2 where
n! appears as 2! = 2, 3! = 6, 4! = 24, 5! = 120. Note: by definition 0! = 1,
so n! is well-defined for non-negative integers.
A function f ( x ) can e.g. be given as an integral (that in this case can be expressed by the
ordinary elementary functions):
Z x
2
f (x) = e−t dt . (4-5)
0
But we can easily find the higher order derivatives of the function for every x:
2 2 2 2
f 0 ( x ) = e− x , f 00 ( x ) = −2 · x · e− x , f 000 ( x ) = −2 · e− x + 4 · x2 · e− x etc. (4-6)
eNote 4 4.2 APPROXIMATIONS BY POLYNOMIALS 99
We assume that a function f ( x ) is given as a solution to a differential equation with the initial
conditions at x0 :
where q( x ) is a given smooth function of x. Again we can fairly easily find the higher order
derivatives of the function at x0 by using the initial conditions directly and by differentiating
the differential equation. We get the following from the initial conditions and from the differ-
ential equation itself:
f 0 ( x0 ) = −3 , f 00 ( x0 ) = q( x0 ) − 3 f 0 ( x0 ) − 7 f ( x0 ) = q( x0 ) + 2 . (4-8)
The third (and the higher-order) derivatives of f ( x ) we then obtain by differentiating both
sides of the differential equation. E.g. by differentiating once we get:
f 000 ( x ) + 3 f 00 ( x ) + 7 f 0 ( x ) = q0 ( x ) , (4-9)
f 000 ( x0 ) = q0 ( x0 ) − 3 f 00 ( x0 ) − 7 f 0 ( x0 )
= q0 ( x0 ) − 3 · (q( x0 ) + 2) − 7 · (−3) (4-10)
= q0 ( x0 ) − 3q( x0 ) + 15 .
The point of the following to find the polynomial of degree n (e.g. the second-degree
polynomial) that best approximates a given smooth function f ( x ) at and around a given
x0 in the domain of the function D ( f ).
f ( x ) = a0 + a1 · ( x − x0 ) + a2 · ( x − x0 )2 + R2,x0 ( x ) , (4-11)
where a0 , a1 , and a2 are suitable constants that are to be chosen so that the remainder
function also known as the Lagrange remainder term R2,x0 ( x ) is as small as possible at
and around x0 . The remainder function we can express by f ( x ) and the polynomial we
are testing:
R2,x0 ( x ) = f ( x ) − a0 − a1 · ( x − x0 ) − a2 · ( x − x0 )2 , (4-12)
eNote 4 4.2 APPROXIMATIONS BY POLYNOMIALS 100
and it is this function that should be as close as possible to 0 when x is close to x0 such
that the difference between the function f ( x ) and the second-degree polynomial be-
comes as small as possible – at least in the vicinity of x0 .
The next natural requirement is that the graph of the remainder function has horizontal
gradient at x0 such that the tangent to the remainder function then is identical to the x
axis:
0
R2,x 0
( x0 ) = 0 such that f 0 ( x0 ) = a1 , (4-14)
by which a1 is determined.
If similarly we had wished to find an approximating n’th degree polynomial for the
same function f ( x ) we would have found:
f 0 ( x0 ) f ( n ) ( x0 )
f ( x ) = f ( x0 ) + · ( x − x0 ) + · · · + · ( x − x0 )n + Rn,x0 ( x ) , (4-18)
1! n!
where the remainder function Rn,x0 ( x ) is a smooth function that satisfies all the require-
ments:
(n)
Rn,x0 ( x0 ) = R0n,x0 ( x0 ) = · · · = Rn,x0 ( x0 ) = 0 . (4-19)
eNote 4 4.2 APPROXIMATIONS BY POLYNOMIALS 101
At this point it is reasonable to expect, on one hand, that these requirements on the re-
mainder functions can be satisfied; on the other, that the remainder function itself must
’appear like’ and be as small as a power of ( x − x0 ) close to x0 .
f (n+1) (ξ ( x ))
Rn,x0 ( x ) = · ( x − x 0 ) n +1 , (4-20)
( n + 1) !
The other way is the following one, that contains an epsilon function:
Rn,x0 ( x ) = ( x − x0 )n · ε f ( x − x0 ) , (4-21)
Proof
We will content ourselves by proving the first statement (4-20) in the simplest case, viz. for
n = 0, i.e. the following : On the interval between (a fixed) x and x0 we can always find a
value ξ such that the following applies:
f 0 (ξ )
R0,x0 ( x ) = f ( x ) − f ( x0 ) = · ( x − x0 ) . (4-22)
(1) !
But this is only a form of the mean value theorem: If a smooth function has values f ( a) and
f (b), respectively, at the end points of an interval [ a, b], then the graph for f ( x ) has at some
position a tangent that is parallel to the line segment connecting the two points ( a, f ( a)) and
(b, f (b)), see Figure 4.1.
f ( n +1) ( ξ )
The other statement (4-21) follows from the first (4-20) by observing that ( n +1) !
· ( x − x 0 ) n +1
f ( n +1) ( ξ )
is an epsilon function of ( x − x0 ), since ( n +1) !
is bounded and since ( x − x0 )n+1 is itself an
epsilon function.
eNote 4 4.2 APPROXIMATIONS BY POLYNOMIALS 102
Figure 4.1: Two points on the blue graph curve for a function are connected with a line
segment (red). The mean value theorem then says that at least one position exists (in
the case shown, exactly two positions, marked in green) on the curve between the two
given points where the slope f 0 ( x ) for the tangent (black) to the curve is exactly the
same as the slope of the straight line segment.
f 0 ( x0 ) f ( n ) ( x0 )
Pn,x0 ( x ) = f ( x0 ) + · ( x − x0 ) + · · · + · ( x − x0 ) n (4-23)
1! n!
is called the approximating polynomial of nth degree for the function f ( x ) with
development point x0 .
To sum up:
eNote 4 4.2 APPROXIMATIONS BY POLYNOMIALS 103
f (n+1) (ξ ( x ))
Rn,x0 ( x ) = · ( x − x 0 ) n +1 for the ξ ( x ) between x and x0
( n + 1) !
(4-25)
and
Rn,x0 ( x ) = ( x − x0 )n · ε f ( x − x0 ) .
f 0 ( x0 ) f ( n ) ( x0 )
f ( x ) = f ( x0 ) + · ( x − x0 ) + · · · + · ( x − x0 ) n + ( x − x0 ) n · ε f ( x − x0 ) ,
1! n!
where ε f ( x − x0 ) denotes an epsilon function of ( x − x0 ), i.e. ε f ( x − x0 ) → 0 for
x → x0 .
One might be led to believe that every polynomial is its own approximating polynomial
because every polynomial must be the best approximation to itself. Here is an example that
shows that this is not that simple. We look at the third-degree polynomial
f ( x ) = 1 + x + x2 + x3 . (4-26)
eNote 4 4.2 APPROXIMATIONS BY POLYNOMIALS 104
The polynomial f ( x ) has the following quite different approximating polynomials - depen-
dent on the choice of development point x0 and degree of development n:
P7,x0 =0 ( x ) = 1 + x + x2 + x3
P3,x0 =0 ( x ) = 1 + x + x2 + x3
P2,x0 =0 ( x ) = 1 + x + x2
P1,x0 =0 ( x ) = 1 + x
P0,x0 =0 ( x ) = 1
P7,x0 =1 ( x ) = 1 + x + x2 + x3
P3,x0 =1 ( x ) = 1 + x + x2 + x3
P2,x0 =1 ( x ) = 2 − 2 · x + 4 · x2 (4-27)
P1,x0 =1 ( x ) = −2 + 6 · x
P0,x0 =1 ( x ) = 4
P7,x0 =7 ( x ) = 1 + x + x2 + x3
P3,x0 =7 ( x ) = 1 + x + x2 + x3
P2,x0 =7 ( x ) = 344 − 146 · x + 22 · x2
P1,x0 =7 ( x ) = −734 + 162 · x
P0,x0 =7 ( x ) = 400 .
For the function f ( x ) = 1 + x + x2 + x3 we consider the following two splittings into approx-
imating polynomials and corresponding remainder functions:
where the two approximating polynomials P2,x0 =1 ( x ) and P1,x0 =7 ( x ) already are stated in
example 4.9. Determine the two remainder functions R2,x0 =1 ( x ) and R1,x0 =7 ( x ) expressed in
both of the two ways shown in 4-25: For each of the two remainder functions the respective
expressions for ξ ( x ) and for ε( x − x0 ) are stated.
eNote 4 4.3 CONTINUOUS EXTENSIONS 105
Here are some often-used functions with their respective approximating polynomials (and
corresponding remainder functions expressed by epsilon functions) with the common devel-
opment point x0 = 0 and arbitrarily high degree:
x2 xn
ex = 1 + x + +···+ + xn · ε( x )
2! n!
2 x4 x2n
ex = 1 + x2 + +···+ + x2n · ε( x )
2! n!
x2 x4 x2n
cos( x ) = 1− + + · · · + (−1)n · + x2n · ε( x )
2! 4! (2n)!
x3 x5 x2n+1
sin( x ) = x− + + · · · + (−1)n · + x2n+1 · ε( x )
3! 5! (2n + 1)! (4-29)
x2 xn
ln(1 + x ) = x− + · · · + (−1)n−1 · + xn · ε( x )
2 n!
x2 xn
ln(1 − x ) = −x − −···− − xn · ε( x )
2 n!
1
= 1 − x + x2 − x3 + · · · + (−1)n−1 · x n−1 + x n · ε( x )
1+x
1
= 1 + x + x 2 + x 3 + · · · + x n −1 + x n · ε ( x )
1−x
Note that in Taylor’s Limit formula we always end with an epsilon function
and with the power of x that is precisely the same as the last power used in the
preceding approximating polynomial.
A direct application of Taylor’s limit formula appears in the determination of limit val-
ues for those quotients f ( x )/g( x ) where both the functions, i.e. the numerator f ( x ) and
the denominator g( x ), tend towards 0 for x tending towards 0. What happens to the
quotient as x tends towards 0? We illustrate with a number of examples. Note that even
though the numerator function and the denominator function both are continuous at 0,
the quotient needs not be continuous.
sin( x ) x + x1 · ε( x )
= = 1 + ε( x ) → 1 for x→0 . (4-32)
x x
sin( x ) x − 3!1 x3 + x3 · ε( x ) 1 x
= = − + x · ε( x ) , (4-33)
x2 x2 x 3!
that has no limit value for x → 0. Therefore a continuous extension does not exist in this case.
sin( x2 ) sin(u)
→1 for x→0 because →1 for u→0 . (4-34)
x2 u
1 3
sin( x ) − x x− 3! x + x3 · ε( x ) − x x
2
= 2
= − + x · ε( x ) → 0 for x → 0 . (4-35)
x x 3!
1 3
sin( x ) − x x− 3! x + x3 · ε( x ) − x 1 1
3
= 3
= − + ·ε( x ) → − for x→0 . (4-36)
x x 3! 6
Figure 4.2: The function f ( x ) = sin( x )/x (blue) together with the numerator function
sin( x ) (red) and the denominator function x (also red). The function f ( x ) is continuous
at x = 0 exactly when we use the value f (1) = 1.
How large is the error committed by using the approximating polynomial (which it is
easy to compute) instead of the function itself (that can be difficult to compute) on a
given (typically small) interval around the development point? The remainder function
can of course give the answer to this question. We give here a couple of examples
that show how the remainder function can be used for such error estimations for given
functions.
Figure 4.3: The function f ( x ) = ln( x ) from Example 4.14 (blue), the approximating
first-degree polynomial (black) with development point x0 = 1 and the corresponding
3 5 as the difference between f ( x ) and the approximat-
remainder function (red) illustrated
ing polynomial on the interval 4 , 4 . To the right is shown the figure around the point
(1, 0) close-up.
The logarithmic function ln( x ) is defined for positive values of x. We approximate with
the approximating first-degree polynomial with the development point at x0 = 1 and will
estimate the remainder term on a suitably small interval around x0 = 1, i.e. the starting point
is the following:
3 5
f ( x ) = ln( x ) , x0 = 1 , n = 1 , x ∈ , . (4-38)
4 4
According to Taylor’s formula with the remainder function we have - using the development
point x0 = 1 where f (1) = 0 and f 0 (1) = 1 and using f 00 ( x ) = −1/x2 for all x in the domain:
eNote 4 4.4 ESTIMATION OF THE REMAINDER FUNCTIONS 109
f 0 (1) f 00 (ξ ) 1
f ( x ) = ln( x ) = ln(1) + ( x − 1) + · ( x − 1)2 = x − 1 − · ( x − 1)2 (4-39)
1! 2! 2 · ξ2
1
P1,x0 =1 ( x ) = x − 1 , and R1,x0 =1 ( x ) = − · ( x − 1)2 . (4-40)
2 · ξ2
The absolute value of the remainder function on the given interval can now be evaluated for all
x in the given interval - even if we do not know very much about the position of ξ in the
interval apart from the fact that ξ lies between x and 1:
We have 2
1 1 1
| R1,x0 =1 ( x )| = | − 2
· ( x − 1)2 | ≤ | · | . (4-41)
2·ξ 2 · ξ2 4
Here the minus sign has been removed because we only look at the absolute value and we
have also used that ( x − 1)2 clearly is largest (with the value (1/4)2 ) for x = 3/4 and for
x = 5/4 in the interval. In addition ξ is smallest and thus (1/ξ )2 largest on the interval for
ξ = 3/4. (Note that here we do not use the fact of ξ lying between x and 1 - we simply use
the fact of ξ lying in the interval!) I.e.
1 1 1
| R1,x0 =1 ( x )| ≤ | |≤| | = 18 , (4-42)
32 · ξ 2 3 2
32 · 4
One may well wonder why the remainder function estimation of such a simple
function as f ( x ) = ln( x ) in Example 4.14 should be so complicated, when it is
evident to everybody (!) that the red remainder function in that case assumes
its largest numerical (absolute) value at one of the end points of the actual
interval, see Figure 4.3 – a statement, moreover, which we can prove by a quite
ordinary function investigation.
0 d 1
R1,x 0 =1
(x) = (ln( x ) − ( x − 1)) = − 1 , (4-44)
dx x
that is less than 0 precisely for x > 1 (such that R1,x0 =1 ( x ) to the right of x = 1
is negative and decreasing from the value 0 at x = 1) and greater than 0 for
x < 1 (such that R1,x0 =1 ( x ) to the left of x = 1 is negative and increasing
towards the value 0 at x = 1). But the problem is that we in principle do not
know what the value of ln( x ) in fact is – neither at x = 3/4 nor at x = 5/4
unless we use Maple or some other tool as help. The remainder function
estimate uses only the defined properties of f ( x ) = ln( x ), i.e. f 0 ( x ) = 1/x and
f (1) = 0 and the estimation gives the values (also at the end points of the
interval) with a (numerical) error of at most 1/18 in this case.
With the ordinary function analysis we get a somewhat better estimate of the
remainder function – but only because we beforehand can estimate the func-
tion value at the end points.
Figure 4.4: The function f ( x ) from Example 4.15 (blue), the approximating first- and
third-degree polynomials (black) with development points x0 = 0 and the correspond-
ing remainder functions (red) on the interval [−1, 1].
Given the function from example 4.4, i.e. the function satisfies the following differential
equation with initial conditions:
where we have assumed that the right-hand side of the equation is q( x ) = x2 and that the
development point is x0 = 0. By this we now get:
We have
f 00 (0) 2 f 000 (0) 3
f ( x ) = f (0) + f 0 (0) · x + ·x + · x + x3 · ε( x )
2 6 (4-49)
5 3 2 3
=1 − 3 · x + x + · x + x · ε ( x ) ,
2
such that the approximating third-degree polynomial for f ( x ) with development point x0 =
0 is
5
P3,x0 =0 ( x ) = 1 − 3 · x + x2 + · x3 . (4-50)
2
eNote 4 4.5 FUNCTIONAL INVESTIGATIONS 112
Figure 4.5: The function f ( x ) from Example 4.16 (blue), the approximating first-,
second-, and third-degree polynomials (black) with the development point x0 = 0. The
corresponding respective remainder functions (red) are illustrated as the differences be-
tween f ( x ) and the approximating polynomials.
Note that P3,x0 =0 ( x ) satisfies the initial conditions in (4-47) but the polynomial P3,x0 =0 ( x ) is
not a solution to the differential equation itself!
A very important property of continuous functions is the following, which means one
can control how large and how small values a continuous function can assume on an
interval, as long as the interval is sufficiently nice:
eNote 4 4.5 FUNCTIONAL INVESTIGATIONS 113
Then the range for the function f ( x ) on the interval I is also a bounded, closed and
connected interval [ A , B ] ⊂ R, thus denoted:
R( f | I ) = f ( I ) = { f ( x ) | x ∈ I } = [ A , B ] , (4-51)
where the possibility that A = B is allowed and this happens precisely when f ( x ) is
constant on the whole interval I.
A well-known and important task is to find the global maximum and minimum values
for given functions f ( x ) on given intervals and to determine the x-values for which
these maximum and minimum values are assumed, that is, the minimum and maximum
points. To solve this task the following is an invaluable help – see Figure 4.6:
Proof
f ( x ) = f ( x0 ) + f 0 ( x0 ) · ( x − x0 ) + ( x − x0 ) · ε f ( x − x0 )
(4-52)
= f ( x0 ) + ( x − x0 ) · ( f 0 ( x0 ) + ε f ( x − x0 )) .
Maximum and minimum values for the function f ( x ), x ∈ I, i.e. A and B in the
range [ A, B] for f ( x ) restricted to I, are found by finding and comparing the function
values at the following points:
1. Interval end points (the boundary points a and b for the interval I).
2. Exception points, i.e. the points in the open interval ] a, b[ where the function is
not differentiable.
3. The stationary points, i.e. all the points x0 in the open interval ] a, b[ where
f 0 ( x0 ) = 0.
eNote 4 4.5 FUNCTIONAL INVESTIGATIONS 115
With this method of investigation we not only find the global maximum and
minimum values but also the x-values in I for which the global maximum and
the global minimum are assumed i.e. maximum and minimum points in the
actual interval.
See Figure 4.6, where we only consider the function on the interval I = [−1.5, 2.0]. There are
two exception points where the function is not differentiable: x0 = 0 and x0 = 1. There is
one stationary point in ] − 1.5, 2.0[ where f 0 ( x0 ) = 0 viz. x0 = −1. And finally there are two
boundary points (the interval end points x0 = −1.5 and x0 = 2) that need to be investigated.
Therefore we have the following candidates for global maximum and minimum values for f
on I:
x0 = −1.5 −1 0 1 2
(4-54)
f ( x0 ) = 0.75 0.5 1.5 0 1
In conclusion we read from this that the maximum value for f ( x ) is B = 1.5 which is assumed
at the maximum point x0 = 0. The minimum value is A = 0, assumed at the minimum point
x0 = 1. There are no other maximum or minimum points for f on I.
eNote 4 4.5 FUNCTIONAL INVESTIGATIONS 116
Figure 4.6: The continuous function f ( x ) from example 4.21 (blue). On the graph we
have marked (in red) the 5 points that need to be investigated particularly in order to
determine the range for f in the interval [−1.5, 2], cf. Method 4.20.
If the function we want to investigate is smooth at its stationary points then we can
qualify the Method 4.20 even better, since the approximating polynomial of degree 2
with development point at the stationary point can help in the decision whether the
value of f ( x ) at the stationary point is a candidate to be a maximum value or a minimum
value.
eNote 4 4.5 FUNCTIONAL INVESTIGATIONS 117
Exercise 4.24
Prove Lemma 4.23 by using Taylor’s limit formula with the approximating second-degree
polynomial for f ( x ) and with the development point x0 . Remember that x0 is a stationary
point, such that f 0 ( x0 ) = 0.
is shown in Figure 4.6. On the interval I = [−1.5, 2.0] the function has the proper local
minimum values 0.5 and 0 in the respective proper local minimum points x0 = −1 and x0 = 1
and the function has a proper local maximum value 1.5 at the proper local maximum point
x0 = 0. If we extend the interval to J = [−7, 7] and note that the function values by definition
are constant outside the interval I we get the new local maximum values 0.75 and 1 for f on
J – not one of them is a proper local maximum value. All x0 ∈] − 7, −1.5] and all x0 ∈ [2, 7[
are local maximum points for f on J but not one of them is a proper local maximum point. All
x0 in the open interval x0 ∈] − 7, −1.5[ and all x0 in the open interval x0 ∈]2, 7[ in addition also
local minimum points for f ( x ) i J but not one of them is a proper local minimum point.
eNote 4 4.5 FUNCTIONAL INVESTIGATIONS 118
Figure 4.7: Proper local maxima and proper local minima for the function from Example
4.26 are here indicated on the graph for the function. We note: The local maximum and
minimum points for the function are the x-coordinates of the graph points shown in red,
and the local maximum and minimum values for the function are the y-coordinates of
the graph-points shown in red.
The function f ( x ) Z x
f (x) = cos(t2 ) dt (4-56)
0
has stationary points at those values of x0 satisfying:
π
f 0 ( x0 ) = cos( x02 ) = 0 , dvs. x02 = +p·π where p is an integer . (4-57)
2
Since we also have that
f 00 ( x ) = −2 · x · sin( x2 ) , (4-58)
such that at the stated stationary points it applies
f 00 ( x0 ) = −2 · x0 · (−1) p . (4-59)
From this it follows – via Lemma 4.23 – that every other stationary point x0 along the x-axis
is a proper local maximum point for f ( x ) and the other points proper local minimum points.
See Figure 4.7. In Figure 4.8 are shown graphs (parabolas) for a pair of the approximat-
ing second-degree polynomials for f ( x ) with the development points at chosen stationary
points.
eNote 4 4.5 FUNCTIONAL INVESTIGATIONS 119
Figure 4.8: The graph for the function in Example 4.26 and two approximating parabolas
with development points in two stationary points, which are a proper local minimum
point and a proper local maximum point for f ( x ).
As stated in Lemma 4.23 one cannot from f 0 ( x0 ) = f 00 ( x0 ) = 0 decide whether the function
has a local maximum or minimum at x0 . This is shown in the three simple functions in Figure
4.9 with all the clarity one could wish for: f 1 ( x ) = x4 , f 2 ( x ) = − x4 and f 3 ( x ) = x3 . All three
functions have a stationary point at x0 = 0 and all have f 00 ( x0 ) = 0, but f 1 ( x ) has a proper
local minimum point at 0, f 2 ( x ) has a proper local maximum point at 0, and f 3 ( x ) has neither
a local minimum point nor a local maximum point at 0.
eNote 4 4.5 FUNCTIONAL INVESTIGATIONS 120
4.6 Summary
In this eNote we have studied how one can approximate smooth functions using poly-
nomials.
where the polynomial and the remainder function in Taylor’s limit formula are
written like this:
f 0 ( x0 ) f ( n ) ( x0 )
f ( x ) = f ( x0 ) + · ( x − x0 ) + · · · + · ( x − x0 ) n + ( x − x0 ) n · ε f ( x − x0 ) ,
1! n!
with ε f ( x − x0 ) denoting an epsilon function of ( x − x0 ), i.e. ε f ( x − x0 ) → 0 for
x → x0 .
• Taylor’s limit formula can be used to find the continuous extension of quotients
of functions by finding (if possible) their limit values for x → x0 where x0 are
the values where the numerator function is 0 such that the quotient at the starting
point is not defined at x0 :
sin( x ) x + x1 · ε( x )
= = 1 + ε( x ) → 1 for x→0 . (4-61)
x x
• Estimation of the remainder function gives an upper bound for the largest numer-
ical difference between a given function and the approximating polynomial of a
suitable degree and with a suitable development point on a given interval of in-
vestigation. Such an estimation can also be made for functions that are possibly
only ”known” via a differential equation or as a non-elementary integral:
1 3 5
| ln( x ) − ( x − 1)| ≤ for all x ∈ , . (4-62)
18 4 4
eNote 5
This eNote is about the real number space Rn and the complex number space Cn , which are
essential building blocks in Linear Algebra.
Rn is the symbol for the set of all n-tuples that contain n real elements. For example,
(1, 4, 5) and (1, 5, 4)
are two different 3-tuples that belong to R3 . Similarly Cn is the symbol for the set of all
n-tuples which contains n complex elements, e.g.
(1 + 2i, 0, 3i, 1, 1) and (1, 2, 3, 4, 5)
eNote 5 5.1 NUMBER SPACES 123
are two different 5-tuples that belong to C5 . Formally we write Ln in set notation as:
Ln = {( a1 , a2 , ..., an ) | ai ∈ L} . (5-1)
Definition 5.2
Let ( a1 , a2 , ..., an ) and (b1 , b2 , ..., bn ) be two elements of Ln and let k be a number in L
(a scalar). The sum of the two n-tuples is defined by
Rn with the operations (5-2) and (5-3) is called the n-dimensional real number space.
Similarly, Cn with the operations (5-2) and (5-3), the n-dimensional complex number space.
i · (2 + i, 4) = (−1 + 2i, 4 i ) .
eNote 5 5.1 NUMBER SPACES 124
As a short notation for n-tuples we often use small bold letters, we write e.g.
a = (3, 2, 1) or b = (b1 , b2 , ..., bn ) .
For the n-tuple (0, 0, ..., 0), which is called the zero element of Ln , we use the notion
0 = (0, 0, ..., 0) .
When more complicated computational exercises in the number spaces are called for,
there is a need for the following arithmetic rules.
1. a + b = b + a (addition is commutative)
2. (a + b) + c = a + (b + c) (addition is associative)
7. k1 (a + b) = k1 a + k1 b (distributive rule)
Proof
Concerning rule 4: Given two vectors a = ( a1 , ... an ) and b = (b1 , ..., bn ) . Then
a + b = ( a1 + b1 , ..., an + bn ) = 0 ⇔ b1 = − a1 , ..., bn = − an .
From this we deduce that a has an opposite vector −a given by −a = (− a1 , ..., − an ) . More-
over, this vector is unique.
The other rules are proved by calculating the left and right hand side of the equations and
then comparing the two results.
eNote 5 5.1 NUMBER SPACES 125
From the proof of rule 4 in theorem 5.5 it is evident that for an arbitrary n-tuple
a: −a = (−1)a .
Exercise 5.6
a − b = a + (−b) . (5-4)
ka = 0 ⇔ k = 0 or a = 0 . (5-5)
eNote 5 5.1 NUMBER SPACES 126
eNote 6
a1 · x1 + a2 · x2 + . . . + a n · x n = b . (6-1)
The numbers a1 , a2 , . . . , an are called the coefficients and the number b is, in this con-
text, called the right hand side. The coefficients and the right hand side are considered
known in contrast to the unknowns. The equation is called homogeneous if b = 0, else
inhomogeneous.
eNote 6 6.1 LINEAR EQUATIONS 128
a1 · x1 + a2 · x2 + . . . + a n · x n = b . (6-2)
By the general solution or just the solution set we understand the set of all solutions
to the equation.
An example of a linear equation is the equation for a straight line in the ( x, y)-plane:
y = 2x +5. (6-3)
Here y is isolated on the left hand side and the coefficients 2 and 5 have well known geomet-
rical interpretations. But the equation could also be written
−2 x1 + 1 x2 = 5 (6-4)
where x and y are substituted by the more general names for unknowns, x1 and x2 , and the
equation is of the form (6-1).
The solution set for the equation (6-3) is of course the coordinate set for all points on the line
- by substitution they will satisfy the equation in contrast to all other points!
If all the coefficients of the equation are 0 but the right hand side is non-zero, the equation is
an inconsistent equation, that is, an equation without a solution. An example is the equation
When you investigate linear equations, you can use the usual rule of conversion for
equations: The set of solutions for the equation is not changed if you add the same
number to both sides of the equality sign, and you do not change the solution set if you
multiply both sides of the equality sign by a non-zero constant.
All linear equations that are not inconsistent and which contain more than one solution,
have infinitely many solutions. The following example shows how the solution set in
this case can be written.
2 x1 − x2 + 4 x3 = 5 . (6-7)
First we isolate x1 :
5
x1 = 2 + 21 x2 − 2 x3 . (6-8)
To every choice of x2 and x3 corresponds exactly one x1 . For example, if we set x2 = 1 and
x3 = 4, then x1 = −5. This means that the 3-tuple (−5, 1, 4) is a solution. Therefore we
can consider x2 and x3 free parameters that together determine the value of x1 . Therefore we
rename x2 and x3 to the parameter names s and t, respectively: s = x2 and t = x3 . Then x1
can be expressed as:
x1 = 25 + 12 x2 − 2 x3 = 25 + 12 s − 2 t . (6-9)
Now we can write the general solution to (6-7) in the following standard parameter form:
5 1
−2
x1 2 2
x = x2 = 0 + s · 1 + t · 0 with s, t ∈ L . (6-10)
x3 0 0 1
Note that the parameter form of the middle equation x2 = 0 + s · 1 + t · 0 only expresses the
renaming x2 → s. Similarly, the last equation only expresses the renaming x3 → t.
eNote 6 6.2 A SYSTEM OF LINEAR EQUATIONS 130
The system has m rows, each of which contains an equation. The n unknowns, denoted
x1 , x2 , . . . xn , are present in each of the m equations (unless some of the coefficients
are zero, and we choose not to write down the zero terms). The coefficient of x j in the
equation in row number i is denoted aij . The system is termed homogeneous if all the m
right hand sides bi are equal to 0, otherwise inhomogeneous.
By the general solution or just the solution set we understand the set of all solutions
to the system. A single solution is often termed a particular solution.
eNote 6 6.3 THE COEFFICIENT MATRIX AND THE AUGMENTED MATRIX 131
A homogeneous system of linear equations consisting of two equations with four unknowns
is given by:
x1 + x2 + 2 x3 + x4 = 0
(6-13)
2 x1 − x2 − x3 + x4 = 0
We investigate whether the two 4-tuples x = (1, 1, 2, −6) and y = (3, 0, 1, −5) are particular
solutions to the equations (6-13). Substituting x into the left hand side of the system we get
1+1+2·2−6 = 0
(6-14)
2 · 1 − 1 − 2 − 6 = −7
Because the left hand side is equal to the given right hand side 0 in the first of these equations,
x is only a solution to the first of the two equations. Therefore x is not a solution to the system.
Substituiting y we get
3+0+2·1−5 = 0
(6-15)
2·3−0−1−5 = 0
Since in both equations the left hand side is equal to the right hand side 0 , y is a solution to
both of the equations. Therefore y is a particular solution to the system.
The solution set to a system of linear equations is the intersection of the solu-
tion sets for all the equations comprising the system.
has two rows and three columns. The six elements are termed the elements of the ma-
trix. The diagonal of the matrix consists of the elements with equal row and column
numbers. In M the diagonal consists of the elements 1 and 3.
eNote 6 6.3 THE COEFFICIENT MATRIX AND THE AUGMENTED MATRIX 132
By the coefficient matrix A to the system of linear equations (6-11) we understand the ma-
trix whose first row consists of the coefficients in the first equation, whose second row
consists of the coefficients in the second equation, etc. In short, the following matrix
with m rows and n columns:
a11 a12 · · · a1n
21 a22 · · · a2n
a
A = . .. .. (6-17)
.. . .
am1 am2 · · · amn
The augmented matrix T of the system is constructed by adding a new column to the
coefficient matrix consisting of the right hand sides bi of the system. Thus T consists of
m rows and n + 1 columns. If we collect the right hand sides bi into a column vector b,
which we denote the right hand side of the system, T is composed as follows, where the
vertical line symbolizes the equality sign of the system:
a11 a12 · · · a1n b1
a21 a22 · · · a2n b2
T = A b = . .. .. .. (6-18)
.. . . .
am1 am2 · · · amn bm
The vertical line in front of the last column in (6-18) has only the didactical
function to create a clear representation of the augmented matrix. One can
chose to leave out the line if in a given context this does not lead to misunder-
standings.
Example 6.8 Coefficient Matrix, Right Hand Side and Augmented Matrix
− x2 + x3 = 2
2x1 + 4x2 − 2x3 = 2 (6-19)
3x1 + 4x2 + x3 = 9
we have
0 −1 0 −1
1 2 1 2
A = 2 4 −2 , b = 2 and T = 2 4 −2 2 (6-20)
3 4 1 9 3 4 1 9
Notice that the 0 that is placed in the top left position in A and T, denotes that the coefficient
of x1 in the uppermost row of the system is 0.
eNote 6 6.4 ROW REDUCTION OF SYSTEMS OF LINEAR EQUATIONS 133
The clever thing about a coefficient matrix (and an augmented matrix) is that
we do not need to write down the unknowns. The unique position of the
coefficients in the matrix means that we are sure of which of the unknowns
any single particular coefficient belongs to. Thus we have removed redundant
symbols!
Systems or linear equations can be reduced, that is, made simpler using a method called
Gaussian elimination. The method has several versions, and the special variant used in
these eNotes goes by the name Gauss-Jordan elimination . The algebraic basis for all
variants is that you can reshape a system of linear equations by so-called row operations
without thereby changing the solution set for the system. When a system of equations
is reduced as much as possible it is usually easy to read it and to evaluate the solution
set.
ro3 : To a given equation add one of the other equations multiplied by a constant.
Here we introduce a short notation for each of the three row operations:
ro1 : Ri ↔ R j : The equation in row i is swapped with the equation in row j.
ro2 : k · Ri : The equation in row i is multiplied by k.
ro3 : R j + k · Ri : Add the equation in row i, multiplied by k, to the equation in row j.
An example of ro1 : Consider the system of equations below to the left. We swap two equa-
tions in the two rows thus performing R1 ↔ R2 .
x1 + 2x2 = −3 x1 + x2 = 0
→ (6-21)
x1 + x2 = 0 x1 + 2x2 = −3
The system to the right has the same solution set as the system on the left.
An example of ro2 : Consider the system of equations below to the left. We multiply the
equation in the second row by 5, thus performing 5 · R2 :
x1 + 2x2 = −3 x1 + 2x2 = −3
→ (6-22)
x1 + x2 = 0 5 x1 + 5 x2 = 0
The system to the right has the same solution set as the system on the left.
An example of ro3 : Consider the system of equations below to the left. To the equation in the
second row we add the equation in the first row multiplied by 2, thus performing R2 + 2 · R1 :
x1 + 2x2 = −3 x1 + 2x2 = −3
→ (6-23)
x1 + x2 = 0 3x1 + 5x2 = −6
The system to the right has the same solution set as the system on the left.
The arrow, →, which is used in the three examples indicates that one or more row
operations have taken place.
Proof
The first part of the proof of 6.9 is simple: Since the solution set of a system of equations is
equal to the intersection F of the solution sets for the various equations comprising the system,
F is not altered by the order of the equations being changed. Therefore ro1 is allowed.
Since the solution set of a given equation is not altered when the equation is multiplied by
a constant k 6= 0, F will not be altered if one of the equations is replaced by the equation
multiplied by a constant different from 0. Therefore ro2 is allowed.
we perform an arbitrary row operation of the type ro3 in the following way: An arbitrary
equation L1 (x) = b1 is multiplied by an arbitrary number k and is then added to an arbitrary
different equation L2 (x) = b2 . This produces a new equation L3 (x) = b3 where
We now show that the system of equations B that emerges as a result of replacing L2 (x) = b2
in A by L3 (x) = b3 has the same solution set as A, and that ro3 thus is allowed. First, assume
that x0 is an arbitrary solution to A . Then it follows from the transformation rules for a linear
equation that
k L1 (x0 ) = k b1
and further that
L2 (x0 ) + k L1 (x0 ) = b2 + k b1 .
From this it follows that L3 (x0 ) = b3 , and that x0 is a solution to B. Assume vice versa that
x1 is an arbitrary solution to B . Then it follows that
−k L1 (x1 ) = −k b1
Corollary 6.11
The solution set of a system of linear equations is not altered if the system is trans-
formed an arbitrary number of times, in any order, by the three row operations.
We are now ready to use the three row operations for the row reduction of systems
of linear equations. In the following example we follow the principles of Gauss-Jordan
elimination, and a complete description of the method follows in subsection 6.5.
eNote 6 6.4 ROW REDUCTION OF SYSTEMS OF LINEAR EQUATIONS 136
Consider below, to the left, a system of linear equations, consisting of three equations with
the three unknowns x1 , x2 and x3 . On the right the augmented matrix for the system is written:
− x2 + x3 = 2 0 −1
1 2
The purpose of reduction is to achieve, by means of row operations, the following situation:
x1 is the only remaining part on left hand side of the upper equation , x2 is the only one on the
left hand side of the middle equation and x3 is the only one on the left hand side of the lower
equation. If this is possible then the system of equations is not only reduced but also solved!
This is achieved in a series of steps taken in accordance with the Gauss-Jordan algorithm.
Simultaneously we look at the effect the row operations have on the augmented matrix.
First we aim to have the topmost equation comprise x1 , and to have the coefficient of this x1
be 1. This can be achieved in two steps. We swap the two top equations and multiply the
equation now in the top row by 21 . That is,
1
R1 ↔ R2 and · R1 :
2
x1 + 2x2 − x3 = 1
1 2 −1 1
(6-25)
− x2 + x3 = 2 0 −1 1 2
3x1 + 4x2 + x3 = 9 3 4 1 9
Now we remove all other occurrences of x1 . In this example it is only one occurrence, i.e. in
row 3. This is achieved as follows: we multiply the equation in row 1 by the number −3 and
add the product to the equation in row 3, in short
R3 − 3 · R1 :
x1 + 2x2 − x3 = 1
1 2 −1 1
(6-26)
− x2 + x3 = 2 0 −1 1 2
−2x2 + 4x3 = 6 0 −2 4 6
We have now achieved that x1 only appears in row 1 . There it must stay! The work on x1
is finished. This corresponds to the fact that at the top of the first column of the augmented
matrix there is 1 and directly below it only 0’s. This means that work on the first column is
finished !
The next transformations aim at ensuring that the unknown x2 will be represented only in
row 2 and nowhere else. First we make sure that the coefficient of x2 in row 2 switches
eNote 6 6.4 ROW REDUCTION OF SYSTEMS OF LINEAR EQUATIONS 137
(−1) · R2 :
x1 + 2x2 − x3 = 1
1 2 −1 1
(6-27)
x2 − x3 = −2 0 1 −1 −2
−2x2 + 4x3 = 6 0 −2 4 6
We now remove the occurrences of x2 from row 1 and row 3 with the operations
R1 − 2 · R2 and R3 + 2 · R2 :
x1 + x3 = 5
1 0 1 5
(6-28)
x2 − x3 = −2 0 1 −1 −2
2x3 = 2 0 0 2 2
Now the work with x2 is finished, which corresponds to the fact that in row 2 in the aug-
mented matrix the number in the second column is 1, all the other numbers in the second
column being 0. This column must not be altered by subsequent operations.
Finally we wish that the unknown x3 is represented in row 3 by the coefficient 1 and that x3
is removed from row 1 and row 2. This can be accomplished in two steps. First
1
· R3 :
2
x1 + x3 = 5
1 0 1 5
(6-29)
x2 − x3 = −2 0 1 −1 −2
x3 = 1 0 0 1 1
Then
R1 − R3 and R2 + R3 :
x1 = 4
1 0 0 4
(6-30)
x2 = −1 0 1 0 −1
x3 = 1 0 0 1 1
Now x3 only appears in row 3. This corresponds to the fact that in column 3 in the third row
of the augmented matrix we have 1, each of the other elements in the column being 0. We
have now completed a total reduction of the system, and from this we can conclude that there
exists exactly one solution to the system viz :
Let us remember what a solution is: an n-tuple that satisfies all the equations in
the system! Let us prove that formula (6-31) actually is a solution to equation
(6-24):
−(−1) + 1 = 2
2 · 4 + 4 · (−1) − 2 · 1 = 2
3 · 4 + 4 · (−1) + 1 = 9
As expected all three equations are satisfied!
In (6-30) after the row operations the augmented matrix of the system of linear equations
has achieved a form of special beauty with three so-called leading 1’s in the diagonal
and zeros everywhere else. We say that the transformed matrix is in reduced row echelon
form. It is not always possible to get the simple representation shown in (6-30). Some-
times the leading 1 in the next row is found more than one column to the right, as one
moves down. The somewhat complex definition follows below.
1. The first number in a row that is not 0, is a 1. This is called the leading 1 or the
pivot of the row.
2. In two consecutive rows which both contain a pivot, the upper row’s leading
1 is further to the left than the leading 1 in the following row.
4. Any rows with only 0’s are placed at the bottom of the matrix.
eNote 6 6.4 ROW REDUCTION OF SYSTEMS OF LINEAR EQUATIONS 139
The three matrices shown are all in row reduced echelon form. In A all the leading 1’s are
nicely placed in the diagonal. B has only leading two leading 1’s and you have to go two steps
to the right to go from the first to the second step. In C there is only one leading 1.
Example 6.15
None of the following four matrices is in reduced row echelon form because each violates
exactly one of the rules in the definition 6.13 – which, is left to reader to figure out!
1 1 0 0 0 0 1 0 0 1 0 0
A = 0 1 0 , B = 1 2 0 , C = 0 2 1 and D = 0 0 1 . (6-33)
0 0 1 0 0 1 0 0 0 0 1 0
Note the following important theorem about the relationship between a matrix on the
one hand, and the reduced row echelon form of the same matrix produced through the
use of row operations, on the other.
The unique reduced row echelon form a given matrix M can be transformed into
this way is termed the reduced row echelon form, and given the symbol rref(M).
eNote 6 6.4 ROW REDUCTION OF SYSTEMS OF LINEAR EQUATIONS 140
Proof
We use the following model for the six matrices that are introduced in the course of the proof:
f1 f2
A ←− M −→ B
↓ (6-34)
f1 f2
A1 ←− M1 −→ B1
Suppose a matrix M has been transformed, by two different series of row operations f 1 and
f 2 , into two different reduced row echelon forms A and B . Let column number k be the first
column of A and B where the two matrices differ from one another. We form a new matrix M1
from M in the following way. First we remove all the columns in M whose column numbers
are larger than k. Then we remove just the columns in M whose column numbers are less
than k , and have the same column numbers as a column in A (and thus B ) which does not
contain a leading 1.
Now we transform M1 by the series of row operations f 1 and f 2 , and the resulting matrices
formed hereby are called A1 and B1 , respectively. Then A1 necessarily will be the same matrix
that would result if we remove all the columns from A, similar to those we took away from M
to produce M1 . And the same relationship exists between B1 and B. A1 and B1 will therefore
have a leading 1 in the diagonal of all columns apart from the last, which is the first column
where the two matrices are different from one another. In this last column there are two
possibilities: Either one of the matrices has a leading 1 in this column or neither of them has.
An example of how the situation in the first case could be is:
1 0 0 1 0 0
A1 = 0 1 0 B1 = 0 1 2 (6-35)
0 0 1 0 0 0
We now interpret M1 as the augmented matrix for a system of linear equations L . Both A1
and B1 will then represent a totally reduced system of equations with the same solution set
as L . However, this leads to a contradiction since one of the totally reduced systems is seen
to be inconsistent due to one of the equations now being invalid and the other will have just
one solution. We can therefore rule out that one of A1 and B1 contains a leading 1 in the last
column.
We now investigate the other possibility, that neither of A1 and B1 contains a leading 1 in the
last column. The situation could then be like this:
1 0 1 1 0 1
A1 = 0 1 3 B1 = 0 1 2 (6-36)
0 0 0 0 0 0
Both the totally reduced system of equations as represented by A1 , and that which is
represented by B1 , will in this case have exactly one solution. But when the last column
eNote 6 6.4 ROW REDUCTION OF SYSTEMS OF LINEAR EQUATIONS 141
is different in the two matrices the solution for A1 ’s system of equations will be different
from the solution for B1 ’s system of equations, whereby we again have ended up in a
contradiction.
We conclude that the assumption that M might be transformed into two different reduced
row echelon forms cannot be true. Hence, to M corresponds a unique reduced row echelon
form: rref(M).
From Theorem 6.16 it is relatively easy to obtain the next result about matrices that can
transformed into each other through row operations:
Corollary 6.17
If a matrix M has been transformed by an arbitrary sequence of row operations into
the matrix N, then
Proof
Let s be a sequence of row operations that transforms the matrix M to the matrix N, and
let t be a sequence of row operations that transforms the the matrix N to rref(N). Then the
sequence of row operations consisting of s followed by t , transform M to rref(N). But since
M in accordance with 6.16 has a unique reduced row echelon form, rref(M) must be equal to
rref(N).
If, in the preceding corollary, we interpret M and N as the augmented matrices for two
systems of linear equations, then it follows directly from definition (6.13) that:
eNote 6 6.5 GAUSS-JORDAN ELIMINATION 142
Corollary 6.18
If two systems of linear equations can be transformed into one another by the use of
row operations, then they are identical in the reduced row echelon form (apart from
possible trivial equations).
We are now able to precisely introduce the method of elimination that is applied in these
eNotes.
When you are in the process of reducing systems of linear equations, you are
free to deviate from the Gauss-Jordan method if it is convenient in the situation
at hand. If you have achieved a reduced row echelon form by using other
sequences of row operations, it is the same form that would have been obtained
by using the Gauss-Jordan method strictly. This follows from corollary 6.18.
In Example 6.12 it was possible to read the solution from the totally reduced system of
linear equations. In the following main example the situation is a bit more complicated
owing to the fact that the system has infinitely many solutions.
eNote 6 6.5 GAUSS-JORDAN ELIMINATION 143
We want to reduce the following system of four linear equations in five unknowns:
1 3 2 4 5 9
2 6 4 3 5 3
T =
3 8
(6-39)
6 7 6 5
4 14 8 10 22 32
Below we reduce the system using three row operations. This we will do by only looking at
the transformations of the augmented matrix!
R2 − 2 · R1 , R3 − 3 · R1 and R4 − 4 · R1 :
1 3 2 4 5 9
0 0 0 −5 −5 −15 (6-40)
0 −1 0 −5 −9 −22
0 2 0 −6 2 −4
Now we have completed the treatment of the first column, because we have a leading 1 in
the first row and only 0’s on the other entries in the column.
R2 ↔ R3 and (−1) · R2 :
1 3 2 4 5 9
0 1 0 5 9 22 (6-41)
0 0 0 −5 −5 −15
0 2 0 −6 2 −4
R1 − 3 · R2 and R4 − 2 · R2 :
1 0 2 −11 −22 −57
0 1 0 5 9 22 (6-42)
0 0 0 −5 −5 −15
0 0 0 −16 −16 −48
The work on the second column is now completed. Now a deviation from the standard
situation follows, where leading 1’s are established in the diagonal, because it is not possible
to produce a leading 1 as the third element in the third row. We are not allowed to swap
eNote 6 6.5 GAUSS-JORDAN ELIMINATION 144
row 1 and row 3, because by doing so the first column would be changed in conflict with
the principle that the treatment of the first column is complete. This means that we have
also completed the treatment of the third column (the number 2 in the top row cannot be
removed). To continue the reduction we move on to the fourth element in row three, where
it is possible to provide a leading 1.
− 15 · R3 :
1 0 2 −11 −22 −57
0 1 0 5 9 22 (6-43)
0 0 0 1 1 3
0 0 0 −16 −16 −48
R1 + 11 · R3 , R2 − 5 · R3 and R4 + 16 · R3 :
1 0 2 0 −11 −24
0 1 0 0 4 7 (6-44)
0 0 0 1 1 3
0 0 0 0 0 0
Now the Gauss-Jordan elimination has ended and we can write the totally reduced system
of equations:
1x1 + 0x2 + 2x3 + 0x4 − 11x5 = −24
0x1 + 1x2 + 0x3 + 0x4 + 4x5 = 7
(6-45)
0x1 + 0x2 + 0x3 + 1x4 + 1x5 = 3
0x1 + 0x2 + 0x3 + 0x4 + 0x5 = 0
First, we note that the original system of equations has actually been reduced (made easier)
by the fact that many of the coefficients of the equation system are replaced by 0’s. But
moreover the system with four equations can now be replaced by a system consisting of
only three equations. The last row is indeed a trivial equation that has the whole L5 as
its solution set. Therefore, the solution set of the system system will not change if the last
equation is omitted in the reduced system (since the intersection of the solutions sets of all
four equations equals that of the solution sets from the first three equations alone). Quite
simply , we can therefore write the totally reduced system of equations as:
In the example 6.20 a system of linear equations consisting of 4 equations with 5 un-
knowns has been totally reduced, see equation (6-46). Only three equations are left,
because the trivial equation 0x1 + 0x2 + 0x3 + 0x4 + 0x5 = 0 has been left out since it
only expresses the fact that 0 = 0. That the reduced system of equations contains a
trivial equation means that the reduced row echelon form of the the augmented matrix
contains a 0-row, as in equation (6-44). This leads to the following definition.
From the definition 6.21 and corollary 6.18 together with corollary 6.17 we obtain:
The rank gives the least possible number of non-trivial equations that a system
of equations can be transformed into using row operations. You can never
transform a system of linear equations through row operations in such a way
that it will contain fewer non-trivial equations than it does when it is totally
reduced. This is a consequence of theorem 6.22.
A matrix M with 3 rows and 4 columns is brought into the reduced row echelon form as
follows:
7 −2
3 1 1 0 3 0
M = −1 −3 3 1 → rref(M) = 0 1 −2 0 (6-47)
2 3 0 −3 0 0 0 1
eNote 6 6.6 THE CONCEPT OF RANK 146
A matrix N with 5 rows and 3 columns is brought into reduced row echelon form like this:
2 2 1 1 0 0
−2 −5 −4 0 1 0
N = 3 1 −7 → rref(N) = 0 0 1 (6-48)
2 −1 −8 0 0 0
3 1 −7 0 0 0
Since rref(N) contains three rows that are not 0-rows, ρ(N) = 3.
If we interpret M and N as augmented matrices for linear systems of equations we see that for
both coefficient matrices the rank is 2, this is less than the ranks of the augmented matrices.
We now investigate the relationship between rank and the number of rows and columns.
First we notice that from the definition of 6.21 it follows that the rank of a matrix can
never be larger than the number of matrix rows.
In Example 6.23 the rank of M equals the number of rows in M, while the rank of N is
less than the number of rows in N.
Analogously the rank of a matrix cannot be larger than the number of columns. The
rank is in fact equal to the number of leading 1’s in the reduced row echelon form . And
if the echelon form of the matrix contains more leading 1’s than there are columns, then
there must be at least one column containing more than one leading 1. But this contra-
dicts condition number 3 in the definition 6.13.
In the example 6.23 the rank of M is less than the number of columns in M, while the
rank of N equals the number of columns in N.
Sometimes it is possible to write down the solution set for a system of linear equations
immediately when the corresponding augmented matrix is brought into its reduced
echelon form. This applies when the system has no solution or when the system has
exactly one solution. If the system has infinitely many solutions, work is needed in
order to be able to characterize the solution set. This can be achieved by writing the
solution in a standard parametric form. The concept of rank proves well suited to give
an instructive overview of the classes of solution sets.
The augmented matrix T for a system of linear equations has the same number of rows
as the coefficient matrix A but one column more, which contains the right hand sides
of the equations. There are two possibilities. Either ρ(T) = ρ(A), or ρ(T) = ρ(A) +
1, corresponding to the fact that the last column in rref(T) contains a leading 1. The
consequence of the last possibility is investigated in Example 6.25.
The augmented matrix for a system of linear equations consisting of three equations in two
unknowns is brought into reduced row echelon form
1 −2 0
rref(T) = 0 0 1 (6-50)
0 0 0
The system is thereby reduced to
x1 − 2x2 = 0
0x1 + 0x2 = 1 (6-51)
0x1 + 0x2 = 0
Notice that the equation in the second row is inconsistent and thus has no solutions. Because
the solution set for the system is the intersection of the solution sets for all the equations, the
system has no solutions at all.
Let us look at the reduced row echelon form of the coefficient matrix
1 −2
rref(A) = 0 0 (6-52)
0 0
eNote 6 6.7 FROM REDUCED ROW ECHELON FORM TO THE SOLUTION SET 148
We note that ρ(A) = 1. This is less than ρ(T) = 2, and this is exactly due to the inconsistency
of the equation in the reduced system of equations.
If rref(T) has a row of the form 0 0 · · · 0 1 , then the system has no solu-
tions.
Exercise 6.27
Determine the reduced rwo echelon form of the augmented matrix for the following system
of linear equations, and determine the solution set for the system.
x1 + x2 + 2x3 + x4 = 1
(6-54)
−2x1 − 2x2 − 4x3 − 2x4 = 3
Let n denote the number of unknowns in a given system of linear equations. Then by
the way the coefficient matrices are formed there must be n columns in A.
Further we assume that ρ(A) = n. Then rref(A) contains exactly n leading 1’s. There-
fore the leading 1’s must be placed in the diagonal in rref(A), and all other elements of
rref(A) are zero.
eNote 6 6.7 FROM REDUCED ROW ECHELON FORM TO THE SOLUTION SET 149
Finally we assume that in the given example ρ(A) = ρ(T). Then the solution set can
be read directly from rref(T). The first row in rref(T) will correspond to an equation
where the first unknown has the coefficient 1 while all the other unknowns have the co-
efficient 0. Therefore the value of the first unknown is equal to to the last element in the
first row (the right hand side). Similarly with the other rows, row number i corresponds
to an equation where unknown number i is the only unknown, and therefore its value
is equal to the last element in row number i. Since each unknown there corresponds
to exactly one value, and since ρ(A) = ρ(T) we are certain that there is no inconsis-
tent equation in the given system of equations. Thus the given system of equations has
exactly one solution.
The augmented matrix for a system of linear equations consisting of three equations in two
unknowns has been brought onto the reduced row echelon form
1 0 −3
rref(T) = 0 1 5 (6-55)
0 0 0
Consider the reduced row echelon form of the coefficient matrix for the system
1 0
rref(A) = 0 1 (6-56)
0 0
This has a leading 1 in each column and 0 in all other entries. We further note that ρ(A) =
ρ(T) = 2.
1x1 + 0x2 = −3
0x1 + 1x2 = 5 (6-57)
0x1 + 0x2 = 0
which shows that this system of equations has exactly one solution x = ( x1 , x2 ) = (−3, 5).
The argument given just before the example proves the following theorem:
eNote 6 6.7 FROM REDUCED ROW ECHELON FORM TO THE SOLUTION SET 150
then the system has exactly one solution, and this can be immediately read from
rref(T).
We are now ready to resume the discussion of our main example 6.20, a system of linear
equations with 5 unknowns, for which we found the totally reduced system of equations
consisting of 3 non-trivial equations. Let us now find the solution set and investigate its
properties!
In the example 6.20 the augmented matrix T for a system of linear equations with 4 equations
in 5 unknowns was reduced to
1 0 2 0 −11 −24
0 1 0 0 4 7
rref(T) =
0 0 0 1
(6-59)
1 3
0 0 0 0 0 0
We see that ρ(A) = ρ(T) = 3, i.e. less than 5, the number of unknowns.
With geometry-inspired wording we term the vector (−24, 7, 0, 3, 0) the initial point of the
solution set and the two vectors (−2, 0, 1, 0, 0) and (11, −4, 0, −1, 1) its directional vectors.
Letting x0 , v1 and v2 denote the initial point, and the directional vectors, respectively, we can
write the parametric representation in this way:
x = x0 + t1 v1 + t2 v2 hvor t1 , t2 ∈ L. (6-64)
Since the solution set has two free parameters corresponding to two directional vectors, we
say that it has a double -infinity of solutions.
Let us, inspired by example 6.30, formulate a general method for changing the solution
set to standard parametic form from the totally reduced system of equations:
eNote 6 6.7 FROM REDUCED ROW ECHELON FORM TO THE SOLUTION SET 152
The solution set of the system is brought into standard parametric form in this way:
1. We find rref(T) and from this we write the totally reduced system of equations
(as is done in (6-60)).
3. In this way we have isolated k different unknowns on the left hand side of the
total system. The other (n − k ) unknowns, that are placed on the right hand
side are renamed the parameter names t1 , t2 , . . . , tn−k .
where the vector x0 denotes the initial point of the parameter representation,
while v1 , v2 , . . . , vn−k are its directional vectors (as is done in (6-63)).
Notice that the numbers t1 , t2 , . . . , tn−k can be chosen freely. Regardless of the choice
equation (6-66) will be a valid solution. Therefore they are called free parameters.
Solution sets in which some of the unknowns have definite values are possible. In the
following example the free parameter only influences one of the unknowns. The other
two are locked:
eNote 6 6.8 ON THE NUMBER OF SOLUTIONS 153
We see that ρ(A) = ρ(T) = 2 < n = 3. Accordingly we have one free parameter. We write
the solution set as:
−2
x1 0
x = x2 =
0 +t 1
(6-68)
x3 5 0
where t is a scalar that can be chosen freely.
Then the system has infinitely many solutions that can be written in standard pa-
rameter form with an initial point and (n − k ) directional vectors.
a1 · x + b1 · y = c1
a2 · x + b2 · y = c2 (6-70)
a3 · x + b3 · y = c3
We have previously emphasized that the solution set for a system of equations is the
intersection of the solution sets for each of the equations in the system. Let us now in-
terpret the given system of equations as equations for three straight lines in a coordinate
eNote 6 6.8 ON THE NUMBER OF SOLUTIONS 154
system in the plane. Then the solution set corresponds to a set of points that are common
to all the three lines. In order to answer the question about “number” of solutions we
draw the different situations in Figure 6.1. In situation 2 two of the lines are parallel,
Figure 6.1: The six possible structures of the solutions for three linear equations in two
unknowns.
and in situation 3 all three lines are parallel. Therefore there are no points that are part
of all three lines in the situations 1, 2 and 3. In situation 5 two of the lines are identical
(the blue and the red line coincide in the purple line). Hence there is exactly one com-
mon point in the situations 4 and 5. In the situation 6 all the three lines coincide (giving
the black line). Therefore in this situation there are infinitely many common points.
The example with three equations in two unknowns illustrates the following theorem
which follows from our study of the solution sets in the previous section, see the theo-
rems 6.26, 6.29 and 6.33:
In this section we dig a little deeper into the question about the structure of the solution
set for systems of linear equations. It is particularly important to observe the corre-
spondence between the solution set for an inhomogeneous system of equations and the
corresponding homogeneous system of equations. We start by investigating the homogenous
system.
In the following theorem we describe an important property of the structure of the so-
lution set for homogeneous systems of linear equations.
Let Lhom denote the solution set of a homogeneous system of linear equations. Then
there exists at least one solution to the system, namely the zero or trivial solution. If
x = ( x1 , x2 , . . . x n ) and y = ( y1 , y2 , . . . y n ) (6-72)
are two arbitrary solutions, and k is an arbitrary scalar then both the sum
x + y = ( x1 + y1 , x2 + y2 , . . . x n + y n ) (6-73)
Proof
An obvious property of the system (6-71) is that ρ(A) = ρ(T) (because the right hand side
consists of only zeros). Therefor the system has at least one solution - it follows from theorem
6.29. We can also immediately find a solution, viz. the zero vector, 0 ∈ Ln . That this is a
solution is evident when we replace all the unknowns in the system with the number 0, then
the system consists of m equations of the form 0 = 0.
Apart from this the theorem comprises two parts that are proved separately:
1. If
ai1 x1 + ai2 x2 + · · · + ain xn = 0 for every i = 1, 2, . . . , m (6-75)
and
ai1 y1 + ai2 y2 + · · · + ain yn = 0 for every i = 1, 2, . . . , m (6-76)
then by addition of the two equations and a following factorization with respect to the
coeficients we get
2. If
ai1 x1 + ai2 x2 + · · · + ain xn = 0 for every i = 1, 2, . . . , m (6-78)
and k is an arbitrary scalar, then by multiplying both sides of the equation by k and a
following factorization with respect to the coefficients we get
Remark 6.36
If you take an arbitrary number of solutions from Lhom , multiply these by arbitrary
constants and add the products then the so-called linear combination of solutions
also is a solution. This is a consequence of theorem 6.35.
eNote 6 6.9 THE LINEAR STRUCTURE OF THE SOLUTION SET 157
and the corresponding homogeneous system of linear equations, by which we mean the equa-
tions (6-80) after all the right hand sides bi have been replaced by 0. The solution set
for the inhomogeneous system of equations is called Linhom and the solution set for the
corresponding homogeneous system of equations is called Lhom .
In other words
Linhom = x = x0 + y y ∈ Lhom . (6-81)
or in short
Linhom = x0 + Lhom . (6-82)
Proof
Note that the theorem contains two propositions. One is that the sum of x0 and an arbitrary
vector from Lhom belongs to Linhom . The other is that an arbitrary vector from Linhom can be
written as the sum of x0 and a vector from Lhom . We prove the two propositions separately:
Since
ai1 x01 + ai2 x02 + · · · + ain x0n = bi for any i = 1, 2, . . . , m (6-84)
and
ai1 y1 + ai2 y2 + · · · + ain yn = 0 for any i = 1, 2, . . . , m (6-85)
then by addition of the two equations and a following factorization with respect to the
coeficients we get
2. Assume x ∈ Linhom . We want to show that a vector y ∈ Lhom exists that fulfills
x = x0 + y. (6-87)
and
ai1 x01 + ai2 x02 + · · · + ain x0n = bi for any i = 1, 2, . . . , m (6-89)
When we subtract the lower equation from the upper, we get after factorization
which shows that the vector y defined by y = x − x0 , belongs to Lhom and satisfies:
x = x0 + y.
eNote 7 159
eNote 7
This eNote introduces matrices and arithmetic operations for matrices and deduces the relevant
arithmetic rules. Math knowledge comparable to that of a Danish gymnasium (high school)
graduate is the only requirement for benefitting from this note, but it is a good idea to acquaint
oneself with the number space Rn that is described in eNote 5 The Number Spaces.
A matrix is characterized by the number of rows and columns, and the matrix M above
is therefore called a 2 × 3 matrix. The matrix M is said to contain 2 · 3 = 6 elements. In
addition to rows and columns a number of further concepts are connected. In order to
describe these we write a general matrix, here called A, as:
a11 a12 . . . a1n
a
21 a22 . . . a2n
A = . .. .. (7-2)
.. . .
am1 am2 . . . amn
The matrix A has m rows and n columns, and this can indicated by writing Am×n or the
m × n matrix A. The matrix A is also said to be of the type m × n.
Two m × n-matrices A and B are called equal if the elements in each matrix are equal,
eNote 7 7.1 MATRIX SUM AND THE PRODUCT OF A MATRIX BY A SCALAR 160
A matrix with a single column (n = 1), is called a column matrix. Similarly a matrix with
only one row (m = 1), a row matrix.
A matrix with the same number of row and columns (m = n), is called a square matrix.
Square matrices are investigated in depth in eNote 8 Square Matrices.
If all the elements in an m × n-matrix are real numbers, the matrix is called a real matrix.
The set of these matrices is denoted Rm×n .
It is possible to add two matrices if they are of the same type. You then add the elements
with the same row and column numbers and in this way form a new matrix of the same
type. Similarly you can multiply any matrix by a scalar (a number), this is done by
multiplying all the elements by the scalar. The matrix in which all elements are equal to
0 is called the zero matrix regardless of the type, and is denoted 0 or possibly 0m×n . In
these notes, all other matrices are called proper matrices.
eNote 7 7.1 MATRIX SUM AND THE PRODUCT OF A MATRIX BY A SCALAR 161
The sum is only defined when the matrices are of the same type.
The product of the matrix A by the scalar k is written kA or Ak and is defined as:
k · a11 k · a12 . . . k · a1n
k·a
21 k · a22 . . . k · a2n
kA = Ak = . (7-5)
. .
.. .. ..
k · am1 k · am2 . . . k · amn
The opposite matrix −A (additive inverse) to a matrix A is defined by the matrix that
results when all the elements in A are multiplied by −1 . It is seen that −A = (−1)A .
The matrices are both of the type 2 × 2. We wish to determine a third and fourth matrix
eNote 7 7.1 MATRIX SUM AND THE PRODUCT OF A MATRIX BY A SCALAR 162
C = 4A and D = 2A + B. This can be done through the use of the definition 7.1.
4 −1 4 · 4 4 · (−1) 16 −4
C = 4A = 4 · = =
8 0 4·8 4·0 32 0
(7-7)
8 −2 −4 3 4 1
D = 2A + B = + 1 =
16 0 9 2 25 21
In the following theorem we summarize the arithmetic rules that are valid for sums of
matrices and multiplication by a scalar.
Theorem 7.3 Arithmetic Rules for the Matrix Sum and the Product by a
Scalar
For arbitrary matrices A, B and C in Rm×n and likewise arbitrary real numbers k1
and k2 the following arithmetic rules are valid:
1. A+B = B+A Addition is commutative
2. (A + B) + C = A + (B + C) Addition is associative
3. A+0 = A 0 is a neutral matrix for addition in Rm×n
4. A + (−A) = 0 Every matrices in Rm×n has an opposite matrix
5. k 1 ( k 2 A) = ( k 1 k 2 )A Product of a matrix by scalars is associative
6. ( k 1 + k 2 )A = k 1 A + k 2 A
The distributive rules are valid
7. k 1 (A + B) = k 1 A + k 1 B
8. 1A = A The scalar 1 is neutral in the product by a matrix
The arithmetic rules in Theorem 7.3 can be proved by applying the ordinary arithmetic
rules for real numbers. The method is demonstrated for two of the rules in the following
example.
plus the constants k1 and k2 . We now try by way of example to show the distributive rules in
eNote 7 7.1 MATRIX SUM AND THE PRODUCT OF A MATRIX BY A SCALAR 163
If you take a11 , a12 , a21 anda22 outside the parentheses in each of the elements in the last ex-
pression, it is seen that (k1 + k2 )A = k1 A + k2 A in this case. The operation of taking the
a-elements outside the parentheses is exactly equivalent to be using the distributive rule for
the real numbers.
The second distributive rule is demonstrated for given matrices and constants:
a11 + b11 a12 + b12 k1 ( a11 + b11 ) k1 ( a12 + b12 )
k 1 (A + B) = k 1 =
a21 + b21 a22 + b22 k1 ( a21 + b21 ) k1 ( a22 + b22 )
(7-10)
k1 a11 k1 a12 k1 b11 k1 b12 k1 a11 + k1 b11 k1 a12 + k1 b12
k1 A + k1 B = + =
k1 a21 k1 a22 k1 b21 k1 b22 k1 a21 + k1 b21 k1 a22 + k1 b22
If k1 is taken outside of the parenthesis in each of the elements in the matrix in the last ex-
pression it is seen that the second distributive rule also is valid in this case: k1 (A + B) =
k1 A + k1 B. The distributive rule for real numbers is again used for each element.
Note that the zero matrix in Rm×n is the only matrix Rm×n that is neutral with
respect to addition, and that −A is the only solution to the equation A + X = 0.
In this subsection we describe the multiplication of a matrix with a vector and then the
multiplication of matrix by another matrix.
Notice that the square brackets around the column vectors can be removed just
like that! This can be done in all dealings with matrices, where double square
brackets occur. It is always the innermost brackets that are removed. In this
way there is no difference between the two expressions. The last expression is
always preferred, because it is the easier to read.
We now define the product of a matrix and a vector, in which the matrix has as many
columns as the vector has elements:
eNote 7 7.2 MATRIX-VECTOR PRODUCTS AND MATRIX-MATRIX PRODUCTS 165
The result is a column vector with m elements. It is the sum of the products of
the k’th column in the matrix and the k’th element in the column vector for all k =
1, 2, . . . , n.
It is necessary that there are as many columns in the matrix as there are rows in the
column vector, here n.
Notice the order in the matrix-vector product: first matrix, then vector! It is not
a vector-matrix product so to speak. The number of rows and columns will not
match in the other configuration unless the matrix is of the type 1 × 1.
It is seen that the result (in both cases) is a column vector with as many rows as there are
rows in A.
Form the matrix-vector product A with x in the equation Ax = b, when it is given that
a11 a12 a13 x1 b1
A = a21 a22 a23 , x = x2 and b = b2 (7-20)
a31 a32 a33 x3 b3
Is this something you have seen before? From where does it come?
The matrix-matrix product or just the matrix product of A and B is defined like this:
AB = A b1 b2 . . . b p = Ab1 Ab2 . . . Ab p (7-21)
The result is matrix of type m × p. The k’th column in the resulting matrix is a
matrix-vector product of the first matrix (here A) and the k’th column vector in the
last matrix (here B), cf. definition 7.7.
There must be as many columns in the first matrix as there are rows in the last
matrix.
We wish to form the matrix-matrix product of A and B. This is done by use of definition 7.10.
4 5 −8 4 5 3 4 5 3
AB =
1 2 2 1 2 9 1 2 −9
(7-23)
4 · (−8) + 5 · 2 4 · 3 + 5 · 9 4 · 3 + 5 · (−9) −22 57 −33
= =
−8 + 2 · 2 3+2·9 3 + 2 · (−9) −4 21 −15
NB: It is not possible to form the matrix-matrix product BA, because there are not as many
columns in B as there are rows in A (3 6= 2).
Because the two matrices are square matrices of the same type both matrix-matrix products
AB and BA can be calculated. We use the definition 7.10.
3 2 4 3 2 4
AB =
−5 1 −1 −5 1 0
3 · 4 + 2 · (−1) 3·4+2·0 10 12
= =
−5 · 4 + 1 · (−1) −5 · 4 + 1 · 0 −21 −20
(7-25)
4 4 3 4 4 2
BA =
−1 0 −5 −1 0 1
4 · 3 + 4 · (−5) 4 · 2 + 4 −8 12
= =
−1 · 3 −1 · 2 −3 −2
Here we summarize the arithmetic rules that apply to matrix-matrix products and ma-
trix sums. Because the matrix-vector product is a special case of the matrix-matrix prod-
uct, the rules also apply for matrix-vector products.
eNote 7 7.2 MATRIX-VECTOR PRODUCTS AND MATRIX-MATRIX PRODUCTS 168
4 −5
1 2 −3 −2 −1
A= , B= and C = 2 1 (7-26)
3 4 0 0 7
1 −3
We see that A(BC) = (AB)C, and therefore it doesn’t matter which of the matrix products
AB and BC we calculate first. This is valid for all matrices (although not proved here).
As is done in example 7.14 we can demonstrate the other arithmetic rules. By writing
eNote 7 7.3 TRANSPOSE OF A MATRIX 169
down carefully the formula for each element of a matrix in the final product, in terms of
the elements of the other matrices, one can prove the rules properly.
Demonstrate the first arithmetic rule in Theorem 7.13 with two real matrices A2×2 and B2×2
and the constant k.
A> is ’A transpose’. In addition you have that (A> )> = A. Here is a useful arithmetic
rule for the transpose of a matrix-matrix product.
9 1
0 1 6
Given the two matrices A = and B = 1 0 . Then
7 −3 2
−6 3
9 1
0 1 6 0 1 6
AB = 1 0
7 −3 2 7 −3 2
−6 3 (7-31)
1·1−6·6 6·3 −35 18
= =
7·9−3·1−2·6 7·1+2·3 48 13
We now try to form the matrix-matrix product B> A> and we find
0 7
9 1 −6
A> = 1 −3 and B> = (7-32)
1 0 3
6 2
and then
0 7
9 1 −6 9 1 −6
B> A> = 1 −3
1 0 3 1 0 3
6 2 (7-33)
1·1−6·6 9·7−1·3−6·2 −35 48
= =
3·6 1·7+3·2 18 13
The two results look identical:
>
−35 18 −35 48
= ⇔ (AB)> = B> A> , (7-34)
48 13 18 13
7.4 Summary
• Matrices are arrays characterized by the number of columns and rows, determining
the type of the matrix. An entry in the matrix is called an element.
• The type of a matrix is denoted as: Am×n . The matrix A has m rows and n columns.
• Matrices can be added if they are of the same type. This is done by adding corre-
sponding elements.
• The matrix-vector product, of Am×n with the vector v with n elements, is defined
as:
v1
v2
Am×n v = a1 a2 . . . an . = a1 v1 + a2 v2 + . . . + an vn , (7-36)
..
vn
• The matrix-matrix product (or just the matrix product) is defined as a series of
matrix-vector products:
Am×n Bn× p = A b1 b2 . . . b p = Ab1 Ab2 . . . Ab p (7-37)
• More arithmetic rules for matrix sums, matrix products and matrix-scalar prod-
ucts are found in Theorem 7.3 and Theorem 7.13.
eNote 8
Square Matrices
In this eNote we explore the basic characteristics of the set of square matrices and introduce the
notion of the inverse of certain square matrices. We presume that the reader has a knowledge of
basic matrix operations, see e.g. eNote 7, Matrices and Matrix Algebra.
Square matrices are simply matrices with equal number of rows and columns, that is they
are of the type n × n. This note will introduce some of the basic operations that apply
to square matrices.
The elements a11 , a22 , . . . , ann are said to be placed on the main diagonal or just the
diagonal of A.
A square matrix D, the non-zero elements of which lie exclusively on the main diagonal,
is termed a diagonal matrix, and one can write D = diag( a11 , a22 , . . . , ann ).
A symmetric matrix A is a square matrix that is equal to its own transpose, thus A = A> .
The square matrix with 1’s in the main diagonal and zeroes elsewhere, is called the
eNote 8 173
identity matrix regardless of the number of rows and columns. The identity matrix is
here denoted E, (more commonly in the literature as I). Accordingly
1 0 ··· 0
0 1 ··· 0
E = En × n = . . . (8-2)
.
.. .. . . ..
0 0 ... 1
Proof
Suppose another matrix D, satisfies the same relations as E, that is AD = DA = A for all
n × n matrices A. This arbitrary matrix A could be the identity matrix, combining the two
equations we get : D = ED = DE = E.
Since E = D there is no other matrix than the identity matrix E that can be a neutral element
for the matrix product.
The identity matrix can be regarded as the ”1 for matrices": A scalar is not
altered by multiplication by 1, likewise a matrix is not altered by the matrix
product of the matrix with the identity matrix of the same type.
As is evident from the following it is often crucial for the manipulation of square matri-
ces whether they have full rank or not. Therefore we now introduce a special concept
eNote 8 8.1 INVERSE MATRIX 174
to express this.
A square matrix is called singular if it not of full rank, that is ρ(An×n ) < n.
In order to determine the “reciprocal matrix” to a matrix A, termed the inverse matrix,
a matrix equation similar to a · x = 1 for a scalar:
AX = XA = E (8-4)
The unknown X is a matrix. If there is a solution X , it is denoted A−1 and is called the
inverse matrix to A. Hence we wish to find a certain matrix called A−1 , for which the
matrix product of A with this matrix yields the identity matrix.
It is not all square matrices that possess an inverse. This is postulated in the following
theorem.
The inverse matrix is uniquely determined by the solution of the matrix equation
AX = E, where X is the unknown.
In the following method it is explained, how the matrix equation described above (8-4)
is solved, and thus how the inverse of an invertible matrix is found.
3. In the elimination process the identity matrix is finally formed on the left hand
side of the vertical line, while
the solution (theinverse of A) can be read on the
right hand side: rref(T) = E X = E A−1 .
We wish to find the inverse matrix A−1 to the matrix A, given in this way:
−16 9 −10
A = 9 −5 6 (8-6)
2 −1 1
This can be done using method 8.4. First the augmented matrix is formed:
−16 9 −10 1 0 0
T = 9 −5 6 0 1 0 (8-7)
2 −1 1 0 0 1
Now we form the leading 1 in the first row: First the row operation R1 + R2 and then R1 +
4 · R3 . This yields
−7 4 −4 1 1 0
1 0 0 1 1 4
9 −5 6 0 1 0 → 9 −5 6 0 1 0 (8-8)
2 −1 1 0 0 1 2 −1 1 0 0 1
Then the numbers in the 1st column of the 2nd and 3rd row are eliminated: R2 − 9 · R1 and
eNote 8 8.1 INVERSE MATRIX 176
R3 − 2 · R1 . Furthermore the 2nd and 3rd rows are swapped: R2 ↔ R3 . We then get
1 0 0 1 1 4 1 0 0 1 1 4
0 −5 6 −9 −8 −36 → 0 −1 1 −2 −2 −7 (8-9)
0 −1 1 −2 −2 −7 0 −5 6 −9 −8 −36
Now we change the sign in row 2: (−1) · R2 and then we eliminate the number in the 2nd
column of the 3rd row: R3 + 5 · R2 .
1 0 0 1 1 4 1 0 0 1 1 4
0 1 −1 2 2 7 → 0 1 −1 2 2 7 (8-10)
0 −5 6 −9 −8 −36 0 0 1 1 2 −1
We see that ρ(A) = ρ(T) = 3, thus A is of full rank, and therefore one can read the inverse to
A on the right hand side of the vertical line:
1 1 4
A−1 = 3 4 6 (8-12)
1 2 −1
Notice that the left hand side of the augmented matrix is the identity matrix. It is
so to speak ”moved” from the right to the left hand side of the equality signs (the
vertical line).
−16 9 −10
1 1 4
AA−1 = 9 −5 6 3 4 6
2 −1 1 1 2 −1
−16 9 −10 −16 9 −10 −16 9 −10
1 1 4
= 9 −5 6 3 9 −5 6 4 9 −5 6 6 (8-13)
2 −1 1 1 2 −1 1 2 2 −1 1 −1
−16 + 27 − 10 −16 + 36 − 20 −64 + 54 + 10
1 0 0
= 9 − 15 + 6 9 − 20 + 12 36 − 30 − 6 = 0 1
0 = E
2−3+1 2−4+2 8−6−1 0 0 1
It is true! By use of the same procedure it is seen that A−1 A = E is also true.
eNote 8 8.1 INVERSE MATRIX 177
As can be seen in the next example, the inverse can be used in the solution of matrix
equations with square matrices. In matrix equations one can interchange terms and mul-
tiply by scalars in order to isolate the unknown just as one would do in ordinary scalar
equations. Moreover one can multiply all terms by matrices – this can be done either
from right or the left on all terms in the equation, yielding different results.
AX = B − CX (8-14)
where
−4 2 −1 −12 7 −9
0 1 0
A = 9 5 −5 , B = 8 −12 5 and C = 0 −10 11 (8-15)
2 0 7 5 0 0 0 −1 −6
First the equation is reduced as far as possible, see e.g. Theorem 7.13, without using the
values:
AX = B − CX ⇔ AX + CX = B − CX + CX ⇔ (A + C)X = B (8-16)
Since X is the unknown we try to isolate this matrix totally. If (A + C) is an invertible matrix,
one can multiply by the inverse to (A + C) from the left on both sides of the equality sign.
Thus:
(A + C)−1 (A + C)X = (A + C)−1 B ⇔ EX = X = (A + C)−1 B , (8-17)
because (A + C)−1 (A + C) = E according to the definition of inverse matrices. We now form
A + C and determine whether the matrix is invertible:
The inverse of this matrix is already determined in Example 8.5, and this part of the procedure
is therefor skipped. X is determined as:
−1
−16 9 −10
0 1 0
X = ( A + C ) −1 B = 9 −5 6 8 −12 5
2 −1 1 5 0 0
(8-19)
28 −11 5
1 1 4 0 1 0
= 3 4
6 8 −12 5 = 62 −45 20
1 2 −1 5 0 0 11 −23 10
eNote 8 8.1 INVERSE MATRIX 178
In the further investigation of the invertibility of the transpose or the inverse of an in-
vertible matrix plus the invertibility of the product of two or more invertible matrices
we will need the following corollary, which is stated without proof (see eNote 9, in
particular, Theorem 9.20 for one way to prove it).
2. The product A B of two square matrices is invertible if and only if both A and B
are invertible.
(A−1 ) −1 = A (8-20)
3. In matrix equations we can multiply all terms by the inverse of a matrix. This
can be done either from the right or the left hand side on both sides of the
equality sign:
4. The inverse of a matrix product of two matrices is equal to the product of the
corresponding inverse matrices in reverse order:
All the arithmetic rules in theorem 8.8 are easily proven by checking.
Below one of the rules is tested in an example. The arithmetic rule in equation (8-22)
has already been used in example 8.6.
We wish to test the last arithmetic rule in theorem 8.8, viz. that (AB)−1 = B−1 A−1 . First A−1
and B−1 are determined by use of method 8.4.
1
1 0 − 52
2 4 1 0 1 2 2 0 1
A E = → → 3 1 (8-25)
6 10 0 1 0 −2 −3 1 0 1 2 −2
Similarly with B:
1 1 1 0
1 1 1 0 1 0 3 −1
B E = → → (8-26)
2 3 0 1 0 1 −2 1 0 1 −2 1
Since we have obtained the identity matrix on the left hand side of the vertical line in both
cases , we get
5
−1 −2 1 −1 3 −1
A = 3 1 and B = (8-27)
2 −2 −2 1
B−1 A−1 is determined:
− 25 7
−1 −1 3 −1 3 −1 1 −9 2
B A = 3 = 13 (8-28)
−2 1 2 −2 1 − 21 2 − 52
On the other side of the equality sign in the arithmetic rule we first calculate AB:
2 4 1 2 4 1 10 14
AB = = (8-29)
6 10 2 6 10 3 26 36
Finally we arrive at
7
−9
(AB)−1 = 13
2 , (8-31)
2 − 52
Comparison of equations (8-28) and (8-31) yields the identity: (AB)−1 = B−1 A−1 .
eNote 8 8.2 POWERS OF MATRICES 180
a) Determine (BA)−1 .
b) Show that AB is not invertible and therefore one cannot determine (AB)−1 .
c) Is it possible to decide whether (AB)−1 exists after you have tried to determine A−1
and B−1 ? If yes, how?
We have now seen how the inverse of an invertible matrix is determined and we say that
it has the power −1. Similarly we define arbitrary integer powers of square matrices.
eNote 8 8.2 POWERS OF MATRICES 181
Furthermore for an arbitrary invertible square matrix B the negative powers are
defined:
n times
z }| {
−n −1 n −1 −1 −1
B = (B ) = B B · · · B , for n ∈ N (8-35)
If A is invertible, these arithmetic rules are also valid for negative integers a and b.
Below is an example of two (simple) matrices that possess some funny characteristics.
The characteristics are not typical for matrices!
.. ..
. .
A−3 = (AA2 )−1 = A−1 = A A−2 = (A2 )−1 = E
A−1 = A A0 = E (8-39)
A1 = A A2 = E
.. ..
. .
Thus all odd powers of A give A, while even powers give the identity matrix:
B2 is determined:
2 2 1 2 1 0 0
B = = =0 (8-41)
−4 −2 −4 −2 0 0
According to the same arithmetic rule B is singular. Then it follows
B0 = E , B1 = B , B2 = 0 , Bn = 0 for n ≥ 2 (8-42)
eNote 8 8.3 SUMMARY 183
8.3 Summary
• Square matrices are matrices where the number of rows equals the number of
columns.
• The unit matrix E is a square matrix with the number one in the diagonal and
zeros elsewhere:
1 0 ··· 0
0 1 ··· 0
E = En × n = . (8-43)
.. . . ..
.. . . .
0 0 ... 1
• If a square matrix has full rank, it is called regular, otherwise it is called singular.
• A square matrix, the entries of which are all zero except for those on the diagonal,
is called a diagonal matrix.
• A square matrix, that is equal to the transpose of itself, is called a symmetric ma-
trix.
• For a regular matrix A there exists a unique inverse, denoted A−1 , satisfying:
• Rules of computation with square and inverse matrices exist, see Theorem 8.8.
• Powers of suare matrices are defined, see Definition 8.12. In addition some arith-
metic rules exist.
• Inverse matrices are e.g. used in connection with change of basis and the eigen-
value problem. Moreover the determinant of a square matrix is defined in eNote 9,
Determinants.
eNote 9 184
eNote 9
Determinants
In this eNote we look at square matrices; that is they are of type n × n for n ≥ 2, see eNote 8.
It is an advantage but not indispensable to have knowledge about the concept of a determinant
for (2 × 2)-matrices in advance. The matrix algebra from eNote 7 is assumed known (sum,
product, transpose and inverse of matrices, plus the general solution method for systems of
linear equations from eNote 6.
The determinant is a well defined function of the total of n2 numbers, that constitute the
elements of an (n × n)-matrix.
In order to define – and then calculate – the value of the determinant of an (n × n)-
matrix directly from the n2 elements in each of the matrices we need two things: First the
well-known formula for the determinant of (2 × 2)-matrices (see the definition 9.1 be-
low) and secondly a method to cut up an arbitrary (n × n)-matrix into (2 × 2)-matrices
eNote 9 9.2 DETERMINANTS OF (2 × 2)−MATRICES 185
and thereby define and calculate arbitrary determinants from the determinants of these
(2 × 2)-matrices.
Remember that the inverse matrix A−1 of a invertible matrix A has the characteristic property
that A−1 · A = A · A−1 = E. Show directly from (9-1) and (9-2), that the inverse matrix A−1
to a (2 × 2)−matrix A can be expressed in the following way (when det(A) 6= 0) :
−1 1 a22 − a12
A = . (9-3)
det(A) − a21 a11
For allsquare matrices a number of basic arithmetic rules are valid; they are presented in
theorem 9.20 below. Check the three first equations in theorem 9.20 for (2 × 2)-matrices A
and B. Use direct calculation of both sides of the equations using (9-2).
9.3 Submatrices
eNote 9 9.3 SUBMATRICES 186
The corresponding determinants are determinants of (2 × 2)−matrices and each of these can
be calculated directly from the definition 9.1 above:
b 11 ) = −7 , det(A
det(A b 12 ) = 1 , det(Ab 13 ) = 5 ,
b 21 ) = −3 , det(A
det(A b 22 ) = 0 , det(Ab 23 ) = 0 , (9-6)
det(A
b 31 ) = 1 , det(Ab 32 ) = −1 , det(A
b 33 ) = −2 .
eNote 9 9.4 INDUCTIVE DEFINITION OF DETERMINANTS 187
The determinant of a 3 × 3 matrix can now be defined from the determinants of 3 of the
9 submatrices, and generally: The determinant of an n × n matrix is defined by the use
of the determinants of the n submatrices that belong to a (freely chosen) row r in the
following way, which naturally is called expansion along the r-th row:
We here and subsequently use the following short notation for the sum and
products of many terms, e.g. n given real numbers c1 , c2 , . . . , cn−1 , cn :
n
c 1 + c 2 + · · · + c n −1 + c n = ∑ ci , and
i =1
n
(9-8)
c 1 · c 2 · · · · · c n −1 · c n = ∏ ci .
i =1
We will use Definition 9.6 directly in order to calculate the determinant of the matrix A that
is given in example 9.5. We choose r = 1 and we thus need three determinants of the sub-
matrices, det(Ab 11 ) = −7, det(A
b 12 ) = 1, and det(A
b 13 ) = 5, which we calculated already in
example 9.5 above:
n
det(A) = ∑ (−1)1+ j a1j det(Ab 1j )
j =1
(9-9)
= (−1)1+1 · 0 · det(A
b 11 ) + (−1)1+2 · 2 · det(A
b 12 ) + (−1)1+3 · 1 · det(A
b 13 )
= 0−2+5 = 3 .
eNote 9 9.4 INDUCTIVE DEFINITION OF DETERMINANTS 188
Show by direct calculation that we obtain the same value for the determinant by use of one
of the other two rows for the expansion of the determinant in example 9.5.
As is already hinted with the definitions and as shown in the concrete case of the matrix
in example 9.5, it doesn’t matter which row (or column) defines the expansion:
Show by direct calculation that we get the same value for the determinant in 9.5 by using
expansion along any of the three columns in A.
It is of course wisest to expand along a row (or a column) that contains many
0’s.
Use the above instructions and results to find the determinants of each of the following ma-
trices:
0 0 2 1 5 3
0 2 1 0 5
0 2 7 1 1 3 2 0 2 0 1 3 2 2 1
1 3 0 2 0 0 5 1 1 4
, 0 5 1 0 1 , . (9-11)
0 0 1 0
1 0 0 0 0 0
0 0 1 0 0
0 5 8 1 0 0 1 0 0 0
5 2 7 1 9
0 5 2 7 1 9
If there are many 0’s in a matrix then it is much easier to calculate its deter-
minant! Especially if all the elements in a row (or a column) are 0 except one
element then it is clearly wisest to expand along that row (or column). And we
are allowed to ’obtain’ a lot of 0’s by application of the well-known row opera-
tions, if you keep record of the constants used for divisions and how often you
swap rows. See theorem 9.16 and example 9.17 below.
We collect some of the most important tools that are often used for the calculation and
inspection of the matrix determinants.
It is not difficult to prove the following theorem, e.g. by expansion first along the first
column or the first row, after which the pattern shows:
eNote 9 9.5 COMPUTATIONAL PROPERTIES OF DETERMINANTS 190
Determine the determinant of the (n × n)− bi-diagonal matrix with arbitrarily given values
µ1 , . . . , µn in the bi-diagonal and 0’s elsewhere:
· · · 0 µ1
0
0 · · · µ2 0
M = bidiag(µ1 , µ2 , · · · , µn ) = .. . (9-14)
.. . ..
. .. . .
µn ... 0 0
General matrices (including square matrices), as known from eNote 6, can be reduced
to reduced row echelon form by the use of row operations. If you keep an eye on what
happens in every step in this reduction then the determinant of the matrix can be read
eNote 9 9.5 COMPUTATIONAL PROPERTIES OF DETERMINANTS 191
directly from the process. The determinant of a matrix behaves ’nicely’ even if you
perform row operations on the matrix:
As indicated above it follows from these properties of the determinant function that the
well-known reduction of a given matrix A to the reduced row echelon form, rref(A),
through row operations as described in eNote 6, in fact comprises a totally explicit calcu-
lation of the determinant of A. We illustrate with a simple example:
We reduce A1 to the reduced row echelon form in the usual way by Gauss–Jordan row op-
erations and all the time we keep an eye on what happens to the determinant by using the
rules in 9.16 (and possibly by checking the results by direct calculations):
1 3 2
A2 = 0 2 1 (9-17)
0 5 1
Operation: 12 R2 , row 2 is multiplied by 1
2 : The determinant is multiplied by 1
2 :
1 1
det(A3 ) = det(A2 ) = − det(A1 ) : (9-18)
2 2
1 3 2
A3 = 0 1 1/2 (9-19)
0 5 1
Operation: R1 − 3R2 : The determinant is unchanged:
1 1
det(A4 ) = det(A3 ) = det(A2 ) = − det(A1 ) : (9-20)
2 2
1 0 1/2
A4 = 0 1 1/2 (9-21)
0 5 1
Operation: R3 − 5R2 : The determinant is unchanged:
1 1
det(A5 ) = det(A4 ) = det(A3 ) = det(A2 ) = − det(A1 ) : (9-22)
2 2
1 0 1/2
A5 = 0 1 1/2 (9-23)
0 0 −3/2
Now the determinant is the product of the elements in the diagonal because all the elements
below the diagonal are 0, see theorem 9.13. All in all we therefore have:
3 1 1
− = det(A5 ) = det(A4 ) = det(A3 ) = det(A2 ) = − det(A1 ) : (9-24)
2 2 2
From this we obtain directly – by reading ’backwards’:
1 3
− det(A1 ) = − , (9-25)
2 2
such that
det(A1 ) = 3 . (9-26)
In addition we have the following relation between the rank and the determinant of a
eNote 9 9.5 COMPUTATIONAL PROPERTIES OF DETERMINANTS 193
If a matrix contains a variable, a parameter, then the determinant of the matrix is a func-
tion of this parameter; in the applications of matrix-algebra it is often crucial to be able
to find the zeroes of this function – exactly because the corresponding matrix is singular
for those values of the parameter, and hence there might not be a (unique) solution to
the corresponding system of linear equations with the matrix as the coefficient matrix.
3. Find the rank of A for a ∈ {−4, −3, −2, −1, 0, 1, 2, 3, 4} . What does the rank have to
do with the roots of the determinant?
1. det(A) = det(A> )
Exercise 9.21
Prove the last 3 equations in theorem 9.20 by the use of det(AB) = det(A) det(B).
Exercise 9.22 The Determinant of a Sum is not the Sum of the Determi-
nants
Show by the most simple example, that the determinant-function det() is not additive. That
is, find two (n × n)−matrices A and B such that
2 −1
3 4 4 0
A = 1 a 2 and B = −5 3 −1 . (9-29)
2 3 3 0 1 a
3. Determine those values of a for which A is invertible and find for these values of a the
expression for det(A−1 ).
Cramer’s method for solving such systems of equations is a direct method. Essentially it
consists of calculating suitable determinants of matrices constructed from A and b and
then writing down the solution directly from the calculated determinants.
Ax = b , (9-30)
1
xj = det(A†bj ) , (9-31)
det(A)
If in particular we let A be the same matrix as above and now let b = (1, 3, 2), then we get by
substitution of b in (9-33) and then computing the relevant determinants:
1 2 1
det(A†b1 ) = det 3 3 2 = 4
2 5 1
0 1 1
det(A†b2 ) = det 1 3 2 = 1 (9-34)
0 2 1
0 2 1
b
det(A†3 ) = det 1 3 3 = 1 .
0 5 2
Since we also know det(A) = 3 we have now constructed the solution to the system of
equations Ax = b, through (9-31):
1 1 1 4 1 1
x = ( x1 , x2 , x3 ) = · 4, · 1, · 1 = , , . (9-35)
3 3 3 3 3 3
2. Determine A−1 and use it directly for the solution of the system of equations.
eNote 9 9.6 ADVANCED: CRAMER’S SOLUTION METHOD 197
3. Solve the system of equations by reduction of the augmented matrix to the reduced
row echelon form as in eNote 2 followed by a reading of the solution.
In order to show what is actually going on in Cramer’s solution formula we first define
the adjoint matrix for a matrix A:
(−1)1+1 det(A
b 11 ) · b 1n ) >
· (−1)1+n det(A
· · · ·
(9-36)
adj(A) =
· · · ·
(−1) n + 1 det(A
b n 1) · · (−1)n+n det(A
b n n)
In other words: The element in entry ( j, i ) in the adjoint matrix adj(A) is the sign-
modified determinant of the (i, j) submatrix, that is: (−1)i+ j det(A
b ij ). Notice the use
of the transpose in (9-36).
such that the inverse matrix to A (which exists precisely if det(A) 6= 0) can be found in the
following way:
1
A−1 = adj(A) . (9-41)
det(A)
Hint: The exercise it not easy. It is recommended to practice on a (2 × 2)-matrix. The zeroes of
the identity matrix in equation (9-40) are obtained by using the property that the determinant
of a matrix with two identical columns is 0.
Proof
1
x = A−1 b = adj(A)b , (9-42)
det(A)
α11 · α1n
x1 b1
· = 1
· · · · . (9-43)
det(A)
xn αn1 · αnn bn
1
= det(A†bj ) ,
det(A)
where we in the establishment of the last equality sign have used that
n
∑ (−1)i+ j bi det(Ab ij ) (9-45)
i =1
eNote 9 9.6 ADVANCED: CRAMER’S SOLUTION METHOD 199
is exactly the expansion of det(A†bj ) along column number j, that is the expansion along the
b-column in det(A†bj ), see the definition in equation (9-10).
eNote 9 9.7 SUMMARY 200
9.7 Summary
• The determinant of a square matrix with real elements is a real number that is cal-
culated from the n2 elements in the matrix, either by expansion along a row or
a column or through inspection of the Gauss-Jordan reduction process to the re-
duced row echelon form. When expanding along a matrix row or column the in-
telligent choice for it is a row or column in the matrix with many 0-elements. The
expansion along row number r takes place inductively after the following formula
that expresses the determinant as a sum of ’smaller’ determinants (with suitable
signs), see definition 9.6:
n
det(A) = ∑ (−1)r+ j arj det(Ab rj ) , (9-46)
j =1
where Ab rj is the submatrix that emerges by deleting row r and column j from the
matrix A, see definition 9.4.
• There exist convenient arithmetic rules for the calculation of determinants of prod-
ucts of matrices, determinants of the inverse matrix, and determinants of the trans-
pose of a matrix. See Theorem 9.20. The most important arithmetic rules are the
product-formula
det(A · B) = det(A) · det(B)
and the transpose-formula
det(A) = det(A> ) .
• Cramer’s solution formula gives the direct way (through computations of deter-
minants) to the solution of a inhomogeneous system of linear equations with a
invertible coefficient matrix, see Theorem 9.24. If the system of equations is
Ax = b , (9-47)
where A†bj denotes the matrix that emerges by replacing column j in the matrix A
with b.
eNote 9 9.8 ADVANCED: CHARACTERISTIC POLYNOMIAL 201
The material in this subsection naturally belongs to this eNote about determinants due
to the involved calculations, but it is only later, when solving the so-called eigenvalue
problem, that we will find the characteristic polynomialsreally useful.
Fo a given square matrix A we define the corresponding characteristic matrix and the
corresponding characteristic polynomial in the following way:
Given a (3 × 3)−matrix A by
3 −2 0
A = 0 1 0. (9-50)
1 −1 2
Then
3 − λ −2
0
KA (λ) = A − λ E = 0 1−λ 0 . (9-51)
1 −1 2 − λ
The corresponding characteristic polynomial for the matrix A is then the following real
polynomial to the variable λ:
eNote 9 9.8 ADVANCED: CHARACTERISTIC POLYNOMIAL 202
With A as in example 9.31 we get the following characteristic polynomial for A by expansion
of the characteristic matrix along the last column:
3 − λ −2
0
KA (λ) = det 0 1−λ 0
1 −1 2 − λ
(9-53)
3+3 3 − λ −2
= (−1) (2 − λ) det
0 1−λ
= (2 − λ)(3 − λ)(1 − λ) .
Give the reasons, why the characteristic polynomial KA (λ) for an (n × n)−matrix A is a
polynomial in λ of the degree n.
eNote 9 9.8 ADVANCED: CHARACTERISTIC POLYNOMIAL 203
Determine the characteristic polynomials for the following matrices and find all real roots in
each of the polynomials:
1 2
A1 = , A2 = diag( a1 , a2 , a3 ) , A3 = bidiag(b1 , b2 , b3 ) . (9-54)
3 4
Construct two (4 × 4)-matrices A and B such that one has only real roots in the corresponding
characteristic polynomial, and such that the other has no real roots in the corresponding
characteristic polynomial.
eNote 10 204
eNote 10
Geometric Vectors
The purpose of this note is to give an introduction to geometric vectors in the plane and
3-dimensional space, aiming at the introduction of a series of methods that manifest themselves
in the general theory of vector spaces. The key concepts are linear independence and linear
dependence, plus basis and coordinates. The note assumes knowledge of elementary geometry in
the plane and 3-space, of systems of linear equations as described in eNote 6 and of matrix
algebra as described in eNote 7.
Geometric vectors can be applied in parallel displacement in the plane and 3-space. In Figure
10.1 the line segment CD is constructed from the line segment AB as follows: all points of
AB are displaced by the vector u. In the same way the line segment EF emerges from AB by
→ → → → →
parallel displacement by the vector v. AB = CD = EF but notice that e.g. AB 6= FE.
In what follows we assume that a unit line segment has been chosen, that is a line segment
eNote 10 205
u
F
v
A
that has the length 1. By |v| we understand the length of the vector v as the proportion-
ality factor with respect to the unit line segment, that is, a real number. All vectors of
the same length as the unit line segment are called unit vectors.
For practical reasons a particular vector that has length 0 and which has no direction
is introduced. It is called the zero vector and is written 0. For every point A we put
→
AA = 0. Any vector that is not the zero vector is called a proper vector.
For every proper vector v we define the opposite vector −v as the vector that has the same
→ →
length as v , but the opposite direction. If v = AB, then BA = −v . For the zero vector
we put −0 = 0 .
It is often practical to use a common initial point when different vectors are to be rep-
resented by arrows. We choose a fixed point O which we term the origin, and consider
those representations of the vectors that have O as the initial point. Vectors represented
in this way are called position vectors, because every given vector v has a unique point
→
(position) P that satisifies v =OP. Conversely, every point Q corresponds to a unique
→
vector u such that OQ = u.
eNote 10 10.1 ADDITION AND MULTIPLICATION BY A SCALAR 206
By the angle between two proper vectors in the plane we understand the unique angle be-
tween their representations radiating from O , in the interval [ 0; π ] . If a vector v in the
plane is turned the angle π/2 counter-clockwise, a new vector emerges that is called
v’s hat vector, it is denoted v
b.
By the the angle between two proper vectors in 3-space we understand the angle between
their representations radiating from O in the plane that contains their representations.
It makes good and useful sense “to add vectors”, taking account of the vectors’ lengths
and directions. Therefore in the following we can introduce some arithmetic operations
for geometric vectors. First it concerns two linear operations, addition of vectors and
multiplication of a vector by a scalar (a real number). Later we will consider three ways
of multiplying vectors, viz. the dot product, and for vectors in 3-space the cross product
and the scalar triple product.
R P
v u+v
u
O Q
eNote 10 10.1 ADDITION AND MULTIPLICATION BY A SCALAR 207
In physics you talk about the ”parallelogram of forces": If the object O is influ-
enced by the forces u and v, the resulting force can be determined as the vector
sum u + v, the direction of which gives the direction of the resulting force, and
the length of which gives the magnitude of the resulting force. If in particular
u and v are of the same length, but have opposite directions, the resulting force
is equal to the 0-vector.
• If k > 0, then kv = vk is the vector that has the same direction as v and which
is k times as long as v.
• If k = 0, then kv = 0.
• If k < 0, then kv = vk is the vector that has the opposite direction of v and which
is −k = | k | as long as v.
v
(-1)v
2v
(−1)u = −u .
From the definition 10.3 the zero rule follows immediately for geometric vec-
tors:
kv = 0 ⇔ k = 0 or v = 0 .
Given a vector a and a line segment of length k, we wish to construct the vector ka.
Q
ka
O 1 k
Figure: Multiplication of a vector by an arbitrary scalar
→
First the position vector OQ= a is marked. Then with O as the initial point we draw a line
which is used as a ruler and which is not parallel to a, and where the numbers 1 and k are
marked. The triangle OkP is drawn so it is congruent with the triangle O1Q. Since the two
→
triangles are similar it must be true that ka =OP.
eNote 10 10.1 ADDITION AND MULTIPLICATION BY A SCALAR 209
Exercise 10.6
Given two parallel vectors a and b and a ruler line. How can you using a pair of compasses
and the ruler line construct a line segment of the length k given that b = ka.
Exercise 10.7
1
Given the proper vector v and a ruler line. Draw the vector |v|
v.
Parametric representations for straight lines in the plane or 3-space are written using proper
vectors. Below we first give an example of a line through the origin and then an example
of a line not passing through the origin.
Given a straight line l through the origin, we wish to write the points on the line using a
parametric representation:
tr P
R
r
O
→
A point R on l different from the origin is chosen. The vector r =OR is called a direction
vector for l . For every point P on l corresponds exactly one real number t that satisfies
→
OP= tr. Conversely, to every real number t corresponds exactly one point P on l so that
→
OP= tr . As t traverses the real numbers from -∞ to +∞, P will traverse all of l in the
direction determined by r. Then
→
{ P | OP= tr where t ∈ R }
is a parametric representation of l .
eNote 10 10.1 ADDITION AND MULTIPLICATION BY A SCALAR 210
The line m does not go through the origin. We wish to describe the points on m by use of a
parametric representation:
tr
r
B
R
m P
b
→
First an initial point B on m is chosen, and we put b =OB. A point R ∈ m different from
→
B is chosen. The vector r = BR is then a directional vector for m . To every point P on m
→
corresponds exactly one real number t that fulfils OP= b + tr. Conversely, to every number t
→
exactly one point P on m corresponds so that OP= b + tr. When t traverses the real numbers
from -∞ to +∞, P will traverse all of m in the direction determined by r. Then
→
{ P | OP= b + tr where t ∈ R }
Parametric representations can also be used for the description of line segments. This is
the subject of the following exercise.
Exercise 10.10
Consider the situation in example 10.9. Draw the oriented line segment with the parametric
representation
→
{ P | OP= b + tr, where t ∈ [ −1; 2 ] } .
eNote 10 10.1 ADDITION AND MULTIPLICATION BY A SCALAR 211
Exercise 10.11
Given two (different) points A and B . Describe with a parametric representation the oriented
line segment from A to B .
We will need more advanced arithmetic rules for addition of geometric vectors and
multiplication of geometric vectors by scalars than the ones we have given in the exam-
ples above. These are sketched in the following theorem and afterwards we will discuss
examples of how they can be justified on the basis of already defined arithmetic opera-
tions and theorems known from elementary geometry.
The arithmetic rules in Theorem 10.12 can be illustrated and proven using geometric
constructions. Let us as an example take the first rule, the commutative rule. Here we
just have to look at the figure in the definition 10.2 , where u + v is constructed. If we
construct v + u, we will displace the line segment OQ with v and consider the emerg-
ing line segment RP2 . It must be true that the parallelogram OQPR is identical to the
parallelogram OQP2 R and hence P2 = P and u + v = v + u.
In the following two exercises the reader is asked to explain two of the other arithmetic
rules.
eNote 10 10.1 ADDITION AND MULTIPLICATION BY A SCALAR 212
Exercise 10.13
ka
kb
a b
a+b
O
k(a+b)
Exercise 10.14
For a given vector u it is obvious that the opposite vector −u is the only vector that
satisfies the equation u + x = 0 . For two arbitrary vectors u and v it is also obvious
that exactly one vector exists that satisfies the equation u + x = v , viz. the vector
x = v + (−u) which is illustrated in Figure 10.2.
x v
-u u
v − u = v + (−u) . (10-1)
v 1
Division by a scalar : = · v ; k 6= 0
k k
k1 v1 + k2 v2 + . . . + k n vn
If all the coefficients k1 , · · · , k n are equal to 0, the linear combination is called im-
proper, or trivial, but if at the least one of the coefficients is different from 0, it is
proper, or non-trivial.
eNote 10 10.2 LINEAR COMBINATIONS 214
d
3c
c b
O a O 2a
-b
In the diagram, to the left the vectors a, b and c are drawn. On the figure to the right we
have constructed the linear combination d = 2a − b + 3c.
Exercise 10.18
There are given in the plane the vectors u, v, s and t, plus the parallelogram A, see diagram.
A
s t
O u
4. Determine real numbers a, b, c and d such that A can be described by the parametric
eNote 10 10.3 LINEAR DEPENDENCE AND LINEAR INDEPENDENCE 215
representation
→
A = { P OP= xu + yv with x ∈ [ a; b ] and y ∈ [ c; d ]} .
If two vectors have representations on the same straight line, one says that they are
linearly dependent. It is evident that two proper vectors are linearly dependent if they
are parallel; otherwise they are linearly independent. We can formulate it as follows: Two
vectors u and v are linearly dependent if the one can be obtained from the other by
multiplication by a scalar different from 0, if e.g. there exists a number k 6= 0 such that
v = ku .
We wish to generalize this original meaning of the concepts of linear dependence and
independence such that the concepts can be used for an arbitrary set of vectors.
If none of the vectors can be written as a linear combination of the others, the set is
called linearly independent.
NB: A set that only consists of one vector is called linearly dependent if the vector is
the 0-vector, otherwise linearly independent.
eNote 10 10.3 LINEAR DEPENDENCE AND LINEAR INDEPENDENCE 216
In the plane are given three sets of vectors (u, v), (r, s) and (a, b, c) , as shown.
c
u
v
b
a
r
s
The set (u, v) is linearly dependent since for this example we have
u = −2v .
b = a−c.
Exercise 10.21
Explain that three vectors in 3-space are linearly dependent if and only if they have represen-
tations lying in the same plane. What are the conditions three vectors must fulfill in order to
be linearly independent?
Exercise 10.22
Consider (intuitively) what is the maximum number of vectors a set of vectors in the plane
can comprise, if the set is to be linearly independent. The same question in 3-space.
When investigate whether or not a given set of vectors is linearly independent or lin-
early dependent, the definition 10.19 does not give a practical procedure. It might be
eNote 10 10.3 LINEAR DEPENDENCE AND LINEAR INDEPENDENCE 217
easier to use the theorem that follows below. This theorem is based on the fact that a
set of vectors is linearly dependent if and only if the 0-vector can be written as a proper
linear combination of the vectors. Assume – as a prerequisite to the theorem – that the
set (a, b, c) is linearly dependent because
c = 2a − 3b.
2a − 3b − c = 0 .
Conversely assume that the 0-vector is a proper linear combination of the vectors u, v
og w like this:
2u − 2v + 3w = 0 .
Then we have (e.g.) that
2 2
w = − u+ v
3 3
and hence the vectors are linearly dependent.
k1 v1 + k2 v2 + · · · + k n vn = 0 (10-2)
Proof
Assume that the set (v1 , v2 , . . . , vn ) is linearly dependent, and let vi be a vector that can be
written as a linear combination of the other vectors. We reorder (if necessary) the set, such
that i = 1, following which v1 can be written in the form
v1 = k 2 v2 + · · · + k n v n ⇔ v1 − k 2 v2 − · · · − k n v n = 0 . (10-3)
The 0-vector is hereby written in the form (10-2), in which not all the coefficients are 0 , be-
cause the coefficient to v1 is 1 .
eNote 10 10.3 LINEAR DEPENDENCE AND LINEAR INDEPENDENCE 218
Conversely, assume that the set is written in the form (10-2), and let k i 6= 0 . We reorder (if
necessary) the set such that i = 1 following which we have
k2 kn
k1 v1 = −k2 v2 − · · · − k n vn ⇔ v1 = − v2 − · · · − vn . (10-4)
k1 k1
From this we see that the set is linearly independent.
Every set of vectors containing the zero vector is linearly dependent. Consider e.g. the set
(u, v, 0, w). It is obvious that the zero-vector can be written as the other three vectors:
0 = 0u + 0v + 0w ,
where the zerovector is written as a linear combination of the other vectors in the set.
Parametric representations for planes in 3-space is written using two linearly independent
vectors. Below we first give an example of a plane through the origin, then an example
of a plane that does not contain the origin.
Given a plane in 3-space through the origin as shown. We wish to describe the points in the
plane by a parametric representation.
R
v
O
P
u
In the given plane we choose two points Q and R, both not the origin, and that do not lie
→ →
on a common line through the origin. The vectors u =OQ and v =OR will then be linearly
independent, and are called direction vectors of the plane. For every point P in the plane we
→
have exactly one pair of numbers (s, t) such that OP= su + tv . Conversely, for every pair of
→
real numbers (s, t) exists exactly one point P in the plane that satisfies OP= su + tv . Then
→
{ P | OP= su + tv ; (s, t) ∈ R2 }
A plane in 3-space does not contain the origin. We wish to describe the plane using a para-
metric representation.
R
P
v
B u
Q
→
First we choose an initial point B in the plane, and we put b =OB. Then we choose two
→ →
linearly independent direction vectors u = BQ and v = BR where Q and R belong to the
plane. To every point P in the plane corresponds exactly one pair of real numbers (s, t), such
that → → →
OP=OB + BP= b + su + tv .
Conversely, to every pair of real numbers (s, t) corresponds exactly one point P in the plane
as given by this vector equation. Then
→
{ P | OP= b + su + tv ; (s, t) ∈ R2 }
eNote 10 10.4 THE STANDARD BASES IN THE PLANE AND SPACE 220
Exercise 10.27
Give a parametric representation for the parallelogram A lying in the plane shown:
v
A B
u
In analytic geometry one shows how numbers and equations can describe geometric ob-
jects and phenomena including vectors. Here the concept of coordinates is decisive. It
is about how we determine the position of the geometric objects in 3-space and relative
to one another using numbers and tuples of numbers. To do so we need to choose a
number of vectors which we appoint as basis vectors. The basis vectors are ordered,
that is they are given a distinct order, and thus they constitute a basis. When a basis is
given all the vectors can be described using coordinates, which we assemble in so called
coordinate vectors. How this whole procedure takes place we first explain for the stan-
dard bases in the plane and 3-space. Later we show that often it is useful to use other
bases than the standard bases and how the coordinates of a vector in different bases are
related.
eNote 10 10.4 THE STANDARD BASES IN THE PLANE AND SPACE 221
j
i
X
O 1
v = xi + yj.
The coefficients x and y in the linear combination are called v’s coordinates with respect
to the basis e, or for short v’s e-coordinates, and they are assembled in a coordinate
vector as follows:
x
ev = .
y
eNote 10 10.4 THE STANDARD BASES IN THE PLANE AND SPACE 222
a
yj
j
X
x xi O i
• When i, j and k are drawn from a chosen point, and we view i and j from the
endpoint of k, then i turns into j, when i is turned by the angle π2 counter-
clockwise.
k
j 1
i O Y
1
v = xi + yj + zk.
The coefficients x, y and z in the linear combination are called v’s coordinates with re-
spect to the basis, or in short v’s e-coordinates, and they are assembled in a coordinate
vectodr as follows:
x
ev = y .
z
P
k a
y
i j Y
O zk
xi+yj
x
Q
X
eNote 10 10.5 ARBITRARY BASES FOR THE PLANE AND SPACE 225
If two linearly independent vectors in the plane are given, it is possible to write every
other vector as a linear combination of the two given vectors. In Figure 10.3 we consider
e.g. the two linearly independent vectors a1 and a2 plus two other vectors u and v: in
the plane
v u
a2
O a1
We see that u = 1a1 + 2a2 and v = −2a1 + 2a2 . These linear combinations are unique
because u and v cannot be written as a linear combination of a1 and a2 using any other
coefficients than those written. Similarly, any other vector in the plane can be written as
a linear combination of a1 and a2 , and our term for this is that the two vectors span the
whole plane.
This makes it possible to generalise the concept of a basis. Instead of a standard basis
we can choose to use the set of vectors (a1 , a2 ) as a basis for the vectors in the plane.
If we call the basis a , we say that the coefficients in the linear combinations above are
coordinates for u and v, respectively, with respect to a basis a, which is written like this:
1 −2
au = and a v = . (10-5)
2 2
For the set of geometric vectors in 3-space we proceed in a similar way. Given three
linearly independent vectors, then every vector in 3-space can be written as a unique-
linear combination of the three given vectors. They span all of 3-space. Therefore we
can choose three vectors as a basis for the vectors in 3-space and express an arbitrary
vector in 3-space by coordinates with respect to this basis. A method for determination
of the coordinates is shown in Figure 10.4, where we are given an a-basis (a1 , a2 , a3 ) plus
eNote 10 10.5 ARBITRARY BASES FOR THE PLANE AND SPACE 226
a3
a2
O
Q
a1
an arbitrary vector u. Through the endpoint P for u a line parallel to a3 is drawn, and
the point of intersection of this line and the plane that contains a1 and a2 , is denoted Q.
→
Two numbers k1 and k2 exist such that OQ= k1 a1 + k2 a2 because (a1 , a2 ) constitutes a
basis in the plane that contains a1 and a2 . Furthermore there exists a number k3 such
→ →
that QP= k3 a3 since QP and a3 are parallel. But then we have
→ →
u =OQ + QP= k1 a1 + k2 a2 + k3 a3 .
u thereby has the coordinate set (k1 , k2 , k3 ) with respect to basis a.
In 3-space three linearly independent vectors a1 , a2 and a3 are given as shown in the Figure.
a1
a3
O
a2
then u has the coordinates (3, 1, 2) with respect to the basis a given by (a1 , a2 , a3 ) which we
write in short as
3
au = 1 .
(10-7)
2
We gather the above considerations about arbitrary bases in the following more formal
definition:
• By a basis a for the geometric vectors in the plane we will understand an ar-
bitrary ordered set of two linear independent vectors (a1 , a2 ). Let an arbitrary
vector u be determined by the linear combination u = xa1 + ya2 . The coeffi-
cients x and y are called u’s coordinates with respect to the basis a, or shorter u’s
a-coordinates, and they are assembled in a coordinate vector as follows:
x
au = . (10-8)
y
The coordinate set of a given vector will change when we change the basis. This crucial
point is the subject of the following exercise.
eNote 10 10.5 ARBITRARY BASES FOR THE PLANE AND SPACE 228
Exercise 10.36
j a2
O i
a1
In the diagram, we are given the standard basis e = (i, j) in the plane plus another basis
a = (a1 , a2 ).
1. A vector u has the coordinates (5, −1) with respect to basis e. Determine u’s
a-coordinates.
2. A vector v has the coordinates (−1, −2) with respect to basis a. Determine v’s
e-coordinates.
eNote 10 10.6 VECTOR CALCULATIONS USING COORDINATES 229
Exercise 10.37
c
O
a d
2. It is also evident from the figure that (a, b, d) is a basis, let us call it n. Determine the
coordinate vector n c.
3. Draw, with the origin as the initial point, the vector u that has the m-coordinates
2
mu = 1 .
1
When you have chosen a basis for geometric vectors in the plane (or in 3-space), then
all vectors can be described and determined using their coordinates with respect to the
chosen basis. For the two arithmetic operations, addition and multiplication by a scalar,
that were introduced previously in this eNote by geometrical construction, we get a
particularly practical alternative. Instead of geometrical constructions we can carry out
calculations with the coordinates that correspond to the chosen basis.
We illustrate this with an example in the plane with a basis a given by (a1 , a2 ) plus two
eNote 10 10.6 VECTOR CALCULATIONS USING COORDINATES 230
vectors u and v drawn from O, see Figure 10.5. The exercise is to determine the vector
b = 2u − v, and we will do this in two different ways.
v
a2 u
b (4,2)
O a1
Method 2 (algebraic): We read the coordinates for u and v and carry out the arithmetic
operations directly on the coordinates:
1 −2 4
a b = 2 a u −a v = 2 − = . (10-10)
2 2 2
Now b can be drawn directly from its coordinates (4, 2) with respect to basis a.
1. a (u + v) = a u + a v
2. a (ku) = k a u
In other words: the coordinates for a vector sum are obtained by adding the coordi-
nates for the summands. And the coordinates for a vector multiplied by a number
are the coordinates of the vector multiplied by that number.
Proof
We carry through the proof for the set of geometric vectors in 3-space. Suppose the coordi-
nates for u and v with respect to the chosen basis a are given by
u1 v1
a u = u 2
and a v = v 2 .
(10-11)
u3 v3
We then have
u = u1 a1 + u2 a2 + u3 a3 og v = v1 a1 + v2 a2 + v3 a3 (10-12)
and accordingly, through the application of the commutative, associative and distributive
arithmetic rules, see Theorem10.12,
u + v = (u1 a1 + u2 a2 + u3 a3 ) + (v1 a1 + v2 a2 + v3 a3 )
(10-13)
= (u1 + v1 )a1 + (u2 + v2 )a2 + (u3 + v3 )a3
which yields
u1 + v1 u1 v1
a (u + v) = u2 + v2 = u2 + v2 = a u + a v
(10-14)
u3 + v3 u3 v3
so that now the first part of the proof is complete. In the second part of the proof we again
use a distributive arithmetic rule, see Theorem 10.12:
The three plane vectors a, b and c have the following coordinate vectors with respect to a
chosen basis v:
1 0 5
va = , vb = and v c = . (10-17)
2 1 −1
Problem: Determine the coordinate vector d = a − 2b + 3c with respect to basis v.
Solution:
vd = v (a − 2b + 3c)
= v (a + (−2)b + 3c)
= v a + v (−2b) + v (3c)
= va − 2 vb + 3 vc
1 0 5 16
= −2 +3 = .
2 1 −1 −3
Here the third equality sign is obtained using the first part of Theorem 10.38 and the fourth
equality sign from the second part of that theorem.
R
v
O
P
u
In accordance with Example 10.25, the plane through the origin shown in the diagram has
the parametric representation
→
{ P | OP= su + tv ; (s, t) ∈ R2 }. (10-18)
The parametric representation (10-18) can then be written in coordinate form like this:
x u1 v1
y = s u2 + t v2 (10-19)
z u3 v3
→
where a OP= ( x, y, z) and (s, t) ∈ R2 .
R
P
v
B u
Q
In accordance with Example 10.26 the plane through the origin shown in the diagram has the
parametric representation
→
{ P | OP= b + su + tv ; (s, t) ∈ R2 }. (10-20)
Suppose that in 3-space we are given a basis a and that
b1 u1 v1
a b = b2 , a u = u2
and a v = v2 .
b3 u3 v3
eNote 10 10.7 VECTOR EQUATIONS AND MATRIX ALGEBRA 234
The parametric representation (10-18) can then be written in coordinate form like this:
x b1 u1 v1
y = b2 + s u2 + t v2 (10-21)
z b3 u3 v3
→
where a OP= ( x, y, z) and (s, t) ∈ R2
A large number of vector-related problems are best solved by resorting to vector equa-
tions. If we wish to solve these equations using the vector coordinates in a given ba-
sis, we get systems of linear equations. The problems can then be solved using matrix
methods that follow in eNote 6. This subsection gives examples of this and sums up
this approach by introducing the coordinate matrix concept in the final Exercise 10.45.
In 3-space are given a basis a and three vectors u, v and p which have the coordinates with
respect to the basis a given by:
2 1 0
au = 1 , av = 4
and a p = 7 .
5 3 1
Problem: Investigate whether p is a linear combination of u and v.
Solution: We will investigate whether we can find coefficients k1 , k2 , such that
k1 u + k2 v = p .
We arrange the corresponding coordinate vector equation
2 1 0
k 1 1 + k 2 4 = 7
5 3 1
which is equivalent to the following system of equations
2k1 + k2 = 0
k1 + 4k2 = 7 (10-22)
5k1 + 3k2 = 1
eNote 10 10.7 VECTOR EQUATIONS AND MATRIX ALGEBRA 235
We consider the augmented matrix T for the system of equations and give (without details)
the reduced row echelon form of the matrix:
1 0 −1
2 1 0
T = 1 4 7 → rref(T) = 0 1 2 (10-23)
5 3 1 0 0 0
We see that the system of equations has exactly one solution, k1 = −1 and k2 = 2, meaning
that
−1u + 2v = p .
In 3-space are given a basis v and three vectors a, b and c which with respect to this basis
have the coordinates
5 1 2
v a = 1 , v b = 0 and v c = 3 .
3 4 1
Problem: Investigate whether the set of vectors (a, b, c) is linearly dependent.
Solution: Following theorem 10.23 we can investigate whether there exists a proper linear
combination
k1 a + k2 b + k3 c = 0 .
We look at the corresponding coordinate vector equation
5 1 2 0
k1 1 + k2 0 + k3 3 = 0
3 4 1 0
5k1 + k2 + 2k3 = 0
k1 + 3k3 = 0 (10-24)
3k1 + 4k2 + k3 = 0
We arrange the augmented matrix T of the system of equations and give (without details) the
reduced row echelon form of the matrix:
5 1 2 0 1 0 0 0
T = 1 0 3 0 → rref(T) = 0 1 0 0 (10-25)
3 4 1 0 0 0 1 0
We see that the system of equations only have the zero solution k1 = 0, k2 = 0 and k3 = 0. The
set of vectors (a, b, c) is therefore linearly independent. Therefore you may choose (a, b, c) as
a new basis for the set of vectors in 3-space.
eNote 10 10.7 VECTOR EQUATIONS AND MATRIX ALGEBRA 236
In the following example we continue the discussion of the relation between coordinates
and change of basis from exercise 10.36
j a2
O i
a1
In the diagram we are given a standard basis e= (i, j) and another basis a= (a1 , a2 ). When the
basis is changed, the coordinates of any given vector are changed. Here we give a systematic
method for expressing the change in coordinates using a matrix-vector product. First we read
the e-coordinates of the vectors in basis a:
1 1
e a1 = and e a2 = . (10-26)
−2 1
v1
1. Problem: Suppose a vector v has the set of coordinates a v = . Determine the
v2
e-coordinates of v.
ev = M · av (10-27)
v1
2. Problem: Suppose a vector v has the set of coordinates e v = . Determine the
v2
a-coordinates of v.
eNote 10 10.7 VECTOR EQUATIONS AND MATRIX ALGEBRA 237
Solution: We multiply from the left on both sides of 10-27 with the inverse matrix to M
and get a-coordinates of v expressed by the matrix-vector product:
av = M−1 · e v (10-28)
Exercise 10.45
By a coordinate matrix with respect to a given basis a for a set of vectors me mean the matrix
that is formed by combining the vector’s a-coordinate columns to form a matrix.
Describe the matrix T in example 10.42 and 10.43 and the matrix M in 10.44 as coordinate
matrices.
eNote 10 10.8 THEOREMS ABOUT VECTORS IN A STANDARD BASIS 238
In this subsection we work with standard coordinate systems, both in the plane and
in 3-space. We introduce two different multiplications between vectors, the dot product
which is defined both in the plane and in 3-space, and the cross product that is only
defined in 3-space. We look at geometric applications of these types of multiplication
and at geometrical interpretations of determinants.
a · b = a1 · b1 + a2 · b2 . (10-29)
a · b = a1 · b1 + a2 · b2 + a3 · b3 . (10-30)
For the dot product between two vectors the following rules of calculation apply.
eNote 10 10.8 THEOREMS ABOUT VECTORS IN A STANDARD BASIS 239
1. a · b = b · a (commutative rule)
2. a · (b + c) = a · b + a · c (associative rule)
3. (ka) · b = a · (kb) = k (a · b)
4. a · a = |a|2
Proof
The Rules 1, 2, 3 follow from a simple coordinate calculation. Rule 4 follows from the
Pythagorean Theorem, and Rule 5 is a direct consequence of Rules 1, 2 and 4.
In the following three theorems we look at geometric applications of the dot product.
Proof
eNote 10 10.8 THEOREMS ABOUT VECTORS IN A STANDARD BASIS 240
b v
O a
The following fact concerns the angle between two vectors, see Figure 10.6.
Proof
The theorem can be proved using the cosine relation. In carrying out the proof one needs
Rule 5 in theorem 10.48. The details are left for the reader.
1. a · b = 0 ⇔ angle(a, b) = π
2
The following theorems are dedicated to orthogonal projections. In Figure 10.7 two
vectors a and b in the plane or 3-space are drawn from the origin.
O v a P
proj(b,a)
Consider P, the foot of the perpendicular from b’s endpoint to the line containing a. By
→
the orthogonal projection of b onto a we mean the vector OP, denoted proj(b, a).
|a · b|
|proj(b, a)| = (10-33)
|a|
eNote 10 10.8 THEOREMS ABOUT VECTORS IN A STANDARD BASIS 242
Proof
Using a known theorem about right angled triangles plus Theorem 10.51 we get
|a · b|
|proj(b, a)| = | cos(v)| |b| = .
|a|
Proof
If a and b are orthogonal the theorem is true since the projection in that case is the zero
vector. Conversely, let sign(a · b) denote the sign of a · b. We have that sign(a · b) is positive
exactly when a and proj(b, a) have the same direction and negative exactly when they have
the opposite direction. Therefore we get
a a·b
proj(b, a) = sign(a · b) · |proj(b, a)| = a,
| a| |a|2
a
where we have used Theorem 10.53, and the fact that |a|
is a unit vector pointing in the
direction of a.
eNote 10 10.8 THEOREMS ABOUT VECTORS IN A STANDARD BASIS 243
The area of a triangle is known to be half the base times its height. We can choose the
length |a| of a as the base. And the height in the triangle is
|b · b
a|
|b| sin(θ ) = , (10-35)
|b
a|
where θ is the angle between the two vectors a and b, and where b a denotes the hat
vector in the plane to a, that is in coordinates we have b
a = (− a2 , a1 ). Hence the area is:
1
Area(4( p, a, b)) = |b · ba|
2
1
= | a1 b2 − a2 b1 |
2
1 a1 b1
= | | (10-36)
2 a2 b2
1 a1 b1
= | det |
2 a2 b2
1
= | det ( [a b] ) | .
2
eNote 10 10.8 THEOREMS ABOUT VECTORS IN A STANDARD BASIS 244
The cross product of two vectors and the scalar triple product of three vectors are intro-
duced using determinants:
The cross product has a geometric significance. Consider Figure 10.9 and the following
theorem:
eNote 10 10.8 THEOREMS ABOUT VECTORS IN A STANDARD BASIS 245
axb
b
p v
a
2. |a × b| = 2 · Area(4( p, a, b)) .
3. The vector set (a, b, a × b) follows the right hand rule: seen from the tip of a × b
the direction from a to b is counter-clockwise.
[a, b, c] = (a × b) · c
= (c1 ( a2 b3 − a3 b2 ) + c2 ( a3 b1 − a1 b3 ) + c3 ( a1 b2 − a2 b1 )
a1 b1 c1
(10-38)
= det a2 b2 c2
a3 b3 c3
= det ([e a e b e c]) .
eNote 10 10.8 THEOREMS ABOUT VECTORS IN A STANDARD BASIS 246
From elementary Euclidean space geometry we know that the volume of a tetrahe-
dron is one third of the area of the base times the height. Consider the tetrahedron
= ( p, a, b, c) spanned by the vectors a, b and c drawn from the point p, in Figure
10.10. The area of the base, 4( p, a, b) has been determined in the second part of Theo-
rem 10.57.
The height can then be determined as the scalar product of the third edge vector c with
a unit vector, perpendicular to the base triangle.
But a × b is exactly perpendicular to the base triangle (because the cross product is
perpendicular to the edge vectors of the base triangle, see part 2 of Theorem (10.57), so
eNote 10 10.8 THEOREMS ABOUT VECTORS IN A STANDARD BASIS 247
we use this:
1 (a × b) · c
Vol(( p, a, b, c)) = | Area(4( p, a, b)) |
3 |a × b| (10-39)
1
= |(a × b) · c|
6
where we have used part 2 of Theorem 10.57.
By comparing this to the definition of scalar triple product, see 10.58, we now get the
volume of a tetrahedron written in ’determinant-form’:
1
Vol(( p, a, b, c)) = | det ([a b c]) | . (10-40)
6
A tetrahedron has the volume 0, is collapsed, exactly when the determinant in (10-40) is
0, and this occurs exactly when one of the vectors can be written as a linear combination
of the two others (why is that?).
Exercise 10.61
A = [a b] . (10-41)
Show that the determinant of A is 0 if and only if the column vectors a and b are linearly
dependent in R2 .
eNote 10 10.8 THEOREMS ABOUT VECTORS IN A STANDARD BASIS 248
Exercise 10.62
A = [a b c] . (10-42)
Show, that the determinant of A is 0 if and only if the column vectors a, b and c constitute a
linearly dependent set of vectors in R3 .
Exercise 10.63
Use the geometric interpretations of the determinant above to show the following
Hadamard’s inequality for (2 × 2)−matrices and for (3 × 3)−matrices (in fact the inequality
is true for all square matrices):
!
n n
(det(A))2 ≤ ∏ ∑ a2ij . (10-43)
j i
eNote 11
In this eNote a general theory is presented for all mathematical sets where addition and
multiplication by a scalar are defined and which satisfy the same arithmetic rules as geometric
vectors in the plane and in 3-space. Using the concepts of a basis and coordinates, it is shown
how one can simplify and standardize the solution of problems that are common to all these sets,
which are called vector spaces. Knowledge of eNote 10 about geometric vectors is an advantage
as is knowledge about the solution sets for systems of linear equations, see eNote 6. Finally,
elementary matrix algebra and a couple of important results about determinants are required
(see eNotes 7 and 8).
The vector concept originates in the geometry of the plane and space where it denotes a
pair consisting of a length and a direction. Vectors can be represented by a line segment
with orientation (an arrow) following which it is possible to define two geometric op-
erations: addition of vectors and multiplication of vectors by numbers (scalar). For the use
in more complicated arithmetic operations one proves eight arithmetic rules concerning
the two arithmetic operations.
In many other sets of mathematical objects one has a need for defining addition of the
objects and multiplication of an object by a scalar. The number spaces Rn and Cn and
the set of matrices Rm×n are good examples, see eNote 5 and eNote 6, respectively. The
remarkable thing is, that the arithmetic rules for addition and multiplication by a scalar,
eNote 11 11.1 GENERALIZATION OF THE CONCEPT OF A VECTOR 250
that are possible to prove within each of these sets, are the same as the arithmetic rules
for geometric vectors in the plane and in space! Therefore one says: Let us make a theory
that applies to all the sets where addition and multiplication by a scalar can be defined
and where all the eight arithmetic rules known from geometry are valid. By this one
carries out a generalization of the concept of geometric vectors, and every set that obeys
the conditions of the theory is therefore called a vector space.
In eNote 10 about geometric vectors it is demonstrated how one can introduce a basis
for the vectors following which the vectors are determined by their coordinates with re-
spect to this basis. The advantage of this is that one can replace the geometric vector
calculation by calculation with the coordinates for the vector. It turns out that it is also
possible to transfer the concepts of basis and coordinates to many other sets of mathe-
matical objects that have addition and multiplication by a scalar.
In the following, when we investigate vector spaces in the abstract sense, it means that
we look at which concepts, theorems and methods follow from the common arithmetic
rules, as we ignore the concrete meaning of addition and multiplication by a scalar has
in the sets of concrete objects where they are introduced. By this one obtains general
methods for every set of the kind described above. The application in any particular vec-
tor space then calls for interpretation in the context of the results obtained. The approach
is called the axiomatic method. Concerning all this we now give the abstract definition of
vector spaces.
eNote 11 11.1 GENERALIZATION OF THE CONCEPT OF A VECTOR 251
I. Addition that from two elements a and b in V forms the sum a + b that also
belongs to V.
II. Multiplication by a scalar that from any a ∈ V and any scalar k ∈ L forms a
product ka or ak that also belongs to V.
V is called a vector space and the elements of V vectors if the following eight
arithmetic rules are valid:
1. a+b = b+a Addition is commutative
2. (a + b) + c = a + (b + c) Addition is associative
3. a+0 = a In V there exists 0 that is neutral wrt. addition
4. a + (−a) = 0 For every a ∈ V there is an opposite object −a ∈ V
5. k 1 ( k 2 a) = ( k 1 k 2 )a Product by a scalar is associative
6. ( k 1 + k 2 )a = k 1 a + k 2 a
The distributive rule applies
7. k 1 (a + b) = k 1 a + k 1 b
8. 1a = a The scalar 1 is neutral in products with vectors
If L in the definition 11.1 stands for R then V is a vector space over the real
numbers. This means that the scalar k (only) can be an arbitrary real number.
Similarly one talks about V as a vector space over the complex numbers if L
stands for C , where k can be an arbitrary complex number.
The requirements I and II in the definition 11.1, that the results of addition and
of multiplication by a scalar in itself must be an element in V, are called the
stability requirements. In other words V must be stable with respect to the
two arithmetic operations.
The set of geometric vectors in the plane and the set of geometric vectors in space are
naturally the most obvious examples of vector spaces, since the eight arithmetic rules in
the definition 11.1 are constructed from the corresponding rules for geometric vectors.
But let us check the stability requirements. Is the sum of two vectors in the plane itself
a vector in the plane? And is a vector in the plane multiplied by a number in itself a
eNote 11 11.1 GENERALIZATION OF THE CONCEPT OF A VECTOR 252
vector in the plane? From the definition of the two arithmetic operations (see Definition
10.2 and Definition 10.3), the answer is obviously yes, therefor the set of vectors in the
plane is a vector space. Similarly we see that the set of vectors in 3-space is a vector
space.
Proof
First part:
Let 01 and 02 be two elements in V both neutral with respect to addition. Then:
01 = 01 + 02 = 02 + 01 = 02 ,
where we have used the fact that addition is commutative. There is only one 0-vector: 0 .
Second part:
Let a1 , a2 ∈ V be two opposite elements for a ∈ V . Then:
a1 = a1 + 0 = a1 + ( a + a2 ) = ( a + a1 ) + a2 = 0 + a2 = a2 ,
where we have used the fact that addition is both commutative and associative. Hence there
is for a only one opposite vector −a .
Exercise 11.4
Prove that the following variant of the zero-rule applies to any vector space:
ka = 0 ⇔ k = 0 or a = 0 . (11-2)
For two arbitrary natural numbers m and n, Rm×n (that is, the set of real m × n-matrices) is a
vector space. Similarly Cm×n (that is, the set of complex m × n-matrices) is a vector space
Consider e.g. R2×3 . If we add two matrices of this type we get a new matrix of the same type,
and if we multiply a 2 × 3-matrix by a number, we also get a new 2 × 3-matrix (see Definition
7.1). Thus the stability requirements are satisfied. That R2×3 in addition satisfies the eight
arithmetic rules, is apparent from Theorem 7.3.
Exercise 11.7
Explain that for every natural number n the number space Ln is a vector space. Remember
to think about the case n = 1 !
In the following two examples we shall see that the geometrically inspired vector space
theory, surprisingly, can be applied to well known sets of functions. Mathematic histori-
ans have in this connection talked about the geometrization of mathematical analysis!
where the coefficients a0 , a1 , · · · an are arbitrary real numbers. Addition of two polynomials
in Pn (R) is defined by pairwise addition of coefficients belonging to the same degree of the
variable, and multiplication of a polynomial in Pn (R) by a number k is defined as the multi-
plication of every coefficient with k. As an example of the two arithmetic operations we look
at two polynomials from P3 (R):
P( x ) = 1 − 2x + x3 = 1 − 2x + 0x2 + 1x3
and
Q( x ) = 2 + 2x − 4x2 = 2 + 2x − 4x2 + 0x3 .
R( x ) = (1 + 2) + (−2 + 2) x + (0 − 4) x2 + (1 + 0) x3 = 3 − 4x2 + x3
− P ( x ) = − a0 − a1 x − · · · − a n x n .
In the same way we show that polynomial P : C 7→ C of at most n’th degree, which we
denote by Pn (C) , is a vector space.
Exercise 11.9
Explain that P(R), that is the set of real polynomials, is a vector space.
eNote 11 11.2 LINEAR COMBINATIONS AND SPAN 255
n( x ) = (k · f )( x ) = k · f ( x ) for every x ∈ I .
We will now justify that C0 ( I ), with the introduced operations of calculations, is a vector
space. Since f + g and k · f are continuous functions, we see that C0 ( I ) satisfies the two
stability requirements. Moreover: there exists a function that acts as the zero vector, viz. the
zero function, that is, the function that has the value 0 for all x ∈ I, and the opposite vector
to f ∈ C0 ( I ) is the vector (−1) f that is also written − f , and which for all x ∈ I has the value
− f ( x ). Now it is obvious that C0 ( I ) with the introduced operations of calculation satisfies
all eight rules in definition 11.1, and C0 ( I ) is thus a vector space.
k1 v1 + k2 v2 + . . . + k p v p
If all the k1 , · · · , k p are equal to 0, the linear combination is called improper, or trivial,
but if at least one of the scalars is different from 0, it is called proper or non-trivial.
eNote 11 11.2 LINEAR COMBINATIONS AND SPAN 256
In the definition 11.11 only one linear combination is mentioned. In many circumstances
it is of interest to consider the total set of possible linear combinations of given vectors.
The set is called the span of the vectors. Consider e.g. a plane in space, through the
origin and containing the position vectors for two non-parallel vectors u and v. The
plane can be considered the span of the two vectors since the position vectors
→
OP= k1 u + k2 v
”run through” all points P in the plane when k1 and k2 take on all conceivable real
values, see Figure 11.1.
R
v
O
P
u
Two geometric vectors u and v are linearly dependent if they are parallel, that is if there
exists a number k, such that v = ku. More generally an arbitrary set of vectors are
linearly dependent if one of the vectors is a linear combination of the others. We wish
to transfer this concept to vector space theory:
NB: If the set of vectors only consists of a single vector, the set is called linearly
dependent if it consists of the 0-vector, and otherwise linearly independent.
Any set of vectors containing the zero vector, is linearly dependent! Consider e.g. the set
{u, v, 0, w}), here the zero vector can trivially be written as a linear combination of the three
other vectors in the set:
0 = 0u + 0v + 0w ,
where the zero vector is written as a linear combination of the other vectors in the set.
eNote 11 11.3 LINEAR DEPENDENCE AND LINEAR INDEPENDENCE 258
C = 3A − 2B .
In contrast the set consisting of A and B is linearly independent, because these two vectors
are not ”parallel’, since a number k obviously does not exist such that B = kA. Similarly with
the sets {A, C} and {B, C} .
When you investigate whether a set of vectors is linearly dependent, use of the defi-
nition 11.14 provokes the question which of the vectors is a linear combination of the
others. Where should we begin the investigation? The dilemma can be avoided if by-
passing the definition we instead use the following theorem:
k1 v1 + k2 v2 + · · · + k p v p = 0 . (11-8)
Proof
Assume first that {v1 , v2 , . . . , v p } are linearly dependent, then one can be written as a linear
combination of the others, e.g.
v1 = k 2 v2 + k 3 v3 + · · · + k p v p . (11-9)
eNote 11 11.4 BASIS AND DIMENSION OF A VECTOR SPACE 259
Conversely, assume that the zero-vector is written as a proper linear combination of the set
of vectors, where one of the coefficients, for example the v1 coefficient k1 , is different from 0
(the same argument works for any of other coefficient). Then we have
k2 kp
k1 v1 + k2 v2 + · · · + k p v p = 0 ⇔ v1 = (−1) v2 + · · · + (−1) v p . (11-11)
k1 k1
Thus v1 is written as a linear combination of the other vectors and the proof is complete.
In the number space R4 the vectors a = (1, 3, 0, 2), b = (−1, 9, 0, 4) and c = (2, 0, 0, 1) are
given. Since
3a − b − 2c = 0
the zero vector is written as a non-trivial linear combination of the three vectors. Thus they
are linearly dependent.
A compelling argument for the introduction of a basis in a vector space is that all vectors
in the vector space then can be written using coordinates. In a later section it is shown
how problems of calculation can be simplified and standardized with vectors when we
use coordinates. But in this section we will discuss the requirements that a basis should
satisfy and investigate the consequences of these requirements.
A basis for a vector space consists of certain number of vectors, usually written in a
definite order. A decisive task for the basis vectors is that they should span the vector
space, but more precisely we want this task to be done with as few vectors as possible.
In this case it turns out that all vectors in the vector space can be written uniquely as a
linear combination of the basis vectors. And it is exactly the coefficients in the unique
linear combination we will use as coordinates.
eNote 11 11.4 BASIS AND DIMENSION OF A VECTOR SPACE 260
Let us start out from some characteristic properties about bases for geometric vectors in
the plane.
v
a2
O
a1
a3
Figure 11.2: Coordinate system in the plane with the basis (a1 , a2 )
Consider the vector set {a1 , a2 , a3 } in Figure 11.2. There is no doubt that any other
vector in the plane can be written as a linear combination of the three vectors. But the
linear combination is not unique, for example the vector v can be written in these two
ways:
The problem is that the a-vectors are not linearly independent, for example a3 = −a1 −
a2 . But if we remove one of the vectors, e.g. a3 , the set is linearly independent, and
there is only one way of writing the linear combination
v = 1a1 + 2a2 .
We can summarize the characteristic properties of a basis for the geometric vectors in
the plane thus:
2. any basis must contain exactly two vectors (if there are more than two, they are
linearly dependent, if there are less than two they do not span the plane), and
These properties can be transferred to other vector spaces. We embark on this now, and
we start by the general definition of a basis.
eNote 11 11.4 BASIS AND DIMENSION OF A VECTOR SPACE 261
1. {v1 , v2 , . . . , vn } spans V .
When we discuss coordinates later, it will be necessary to consider the basis elements
to have a define order, and so we will write them as an ordered set, denoted by using
parentheses: (v1 , v2 , . . . , vn ).
Here we should stop and check that the definition 11.19 does in fact satisfy our require-
ments of uniqueness of a basis. This is established in the following theorem.
Proof
We give the idea in the proof by looking at a vector space V that has a basis consisting of
three basis vectors (a, b, c) and assume that v is an arbitrary vector in V that in two ways
can be written as a linear combination of the basis vectors. We can then write two equations
v = k1 a + k2 b + k3 c
(11-12)
v = k4 a + k5 b + k6 c
By subtracting the lower equation from the upper equation in (11-12) we get the equation
0 = ( k 1 − k 4 )a + ( k 2 − k 5 )b + ( k 3 − k 6 )c . (11-13)
Since a, b and c are linearly independent, the zero vector can only be written as an improper
linear combination of these, therefore all coefficients in (11-12) are equal to 0, yielding k1 = k4 ,
k2 = k5 and k3 = k6 . But then the two ways v has been written as linear combinations of the
basis vectors, is in reality the same, there is only one way!
We now return to the fact that every basis for geometric vectors in the plane contains
two linearly independent basis vectors, and that similarly for geometric vectors in space
the basis must consist of three linearly independent basis vectors. It turns out that the
fixed number of basis vectors is a property of all vector spaces with a basis, and this
makes it possible to talk about the dimension of a vector space that has a basis. To prove
the property we need a lemma.
Lemma 11.21
If a vector space V has a basis consisting of n basis vectors then every set from V
that contains more than n vectors will be linearly dependent.
Proof
To get a grasp of the proof’s underlying idea, consider a vector space V that has a basis
consisting of two vectors (a, b), and investigate three arbitrary vectors c, d and e from V. We
prove that the three vectors necessarily must be linearly independent.
c = c1 a + c2 b
d = d1 a + d2 b (11-14)
e = e1 a + e2 b
Consider further the zero vector written as the following linear combination
x1 c + x2 d + x3 e = 0 , (11-15)
( x 1 c 1 + x 2 d 1 + x 3 e1 ) a + ( x 1 c 2 + x 2 d 2 + x 3 e2 ) b = 0 . (11-16)
Since the zero vector only can be obtained as a linear combination of a and b, if every coeffi-
cient is equal to 0, (11-16) is equivalent to the following system of equations
c 1 x 1 + d 1 x 2 + e1 x 3 = 0
(11-17)
c 2 x 1 + d 2 x 2 + e2 x 3 = 0
eNote 11 11.4 BASIS AND DIMENSION OF A VECTOR SPACE 263
This is a homogeneous system of linear equations in which the number of equations is less
than the number of unknowns. Therefore the system of equations has infinitely many solu-
tions, which means that (11-16) not only is obtainable with x1 = 0, x2 = 0 and x3 = 0. Thus
we have shown that the ordered set (c, d, e) is linearly dependent.
In general: Assume that the basis V consists of n vectors, and that m vectors from V where
m > n are given. By following the same procedure as above a homogeneous system of lin-
ear equations emerges with n equations in m unknowns that, because m > n , similarly has
infinitely many solutions. By this it is shown that the m vectors are linearly dependent.
Proof
Assume that V has two bases with different numbers of big(asis vectors. We denote the basis
with the least number of basis vectors a and the one with largest number b. According to
Lemma 11.21 the b-basis vectors must be linearly dependent, and this contradicts that they
form a basis. The assumption that V can have two bases with different numbers of basis
vectors, must therefore be untrue.
That the number of basis vectors according to theorem 11.22 is a property of vector spaces
with a basis, motivates the introduction of the concept of dimension:
eNote 11 11.4 BASIS AND DIMENSION OF A VECTOR SPACE 264
dim(V ) = n . (11-18)
Remark: There are vector spaces that do not have a finite basis, see Section 11.7.2 below.
Luckily the definition 11.23 confirms an intuitive feeling that the set of geometric vectors
in the plane has the dimension two and that the set of geometric vectors in space has the
dimension three!
We put e1 = (1, 0, 0, 0), e2 = (0, 1, 0, 0), e3 = (0, 0, 1, 0) and e4 = (0, 0, 0, 1) and conclude
using (11-19) that the ordered set (e1 , e2 , e3 , e4 ) spans L4 .
Since we can see that none of the vectors can be written as a linear combination of the others,
the set is linearly independent, and (e1 , e2 , e3 , e4 ) is thereby a basis for L4 . This particular
basis is called standard basis for L4 . Since the number of basis vectors in the standard e-basis
is four, dim(L4 ) = 4 .
This can immediately be generalized to Ln : For every n the set (e1 , e2 , . . . , en ) where
is a basis for Ln . This is called standard basis for Ln . It is noticed that dim(Ln ) = n .
eNote 11 11.4 BASIS AND DIMENSION OF A VECTOR SPACE 265
By standard basis for the vector space R2×3 or C2×3 , we understand the matrix set
1 0 0 0 1 0
0 0 0
, ,..., (11-20)
0 0 0 0 0 0 0 0 1
Similarly we define a standard basis for an arbitrary matrix space Rm×n and for an arbitrary
matrix space Cm×n .
Exercise 11.27
Explain that the matrix set, which in Example 11.26 is referred to as the standard basis for
R2×3 , is in fact a basis for this vector space.
In the vector space P2 (R) of real polynomials of at most 2nd degree, the ordered set (1, x, x2 )
is a basis. This is demonstrated in the following way.
P ( x ) = a0 · 1 + a1 · x + a2 · x 2 ,
a0 · 1 + a1 · x + a2 · x2 = 0 for every x
according to the identity theorem for polynomials is only satisfied if all the coefficients
a0 , a1 and a2 are equal to 0 .
A monomial is a polynomial with only one term. Hence, the ordered set (1, x, x2 ) is called the
monomial basis for P2 (R), and dim( P2 (R)) = 3 .
For every n the ordered set (1, x, x2 , . . . , x n ) is a basis for Pn (R), and is called the monomial
basis for Pn (R). Therefore we have that dim( Pn (R)) = n + 1 .
Similarly the ordered set (1, z, z2 , . . . , zn ) is a basis for Pn (C), it is called monomial basis for
Pn (C). Therefore we have that dim( Pn (C)) = n + 1 .
eNote 11 11.4 BASIS AND DIMENSION OF A VECTOR SPACE 266
In the set of plane geometric vectors one can choose any pair of two linearly independent
vectors as basis. Similarly in 3-space any set of three linear independent vectors is a
basis. We end the section by transferring this to general n-dimensional vector spaces:
Proof
Exercise 11.30
Two geometric vectors a = (1, −2, 1) and b = (2, −2, 0) in 3-space are given. Determine a
vector c such that the ordered set (a, b, c) is a basis for the set of space vectors.
Exercise 11.31
Explain why the ordered set (A, B, C) is a linearly independent set, and complement the set
with a 2×2 matrix D such that (A, B, C, D) is a basis for R2×2 .
eNote 11 11.5 VECTOR CALCULATIONS USING COORDINATES 267
Coordinates are closely connected to the concept of a basis. When a basis is chosen
for a vector space, any vector in the vector space can be described with the help of its
co-ordinates with respect to the chosen basis. By this we get a particularly practical al-
ternative to the calculation operations, addition and multiplication by a scalar, which
originally are defined from the ’anatomy’ of the specific vector space. Instead of car-
rying out these particularly defined operations we can implement number calculations
with the coordinates that correspond to the chosen basis. In addition it turns out that
we can simplify and standardize the solution of typical problems that are common to
all vector spaces. But first we give a formal introduction of coordinates with respect to
a chosen basis.
x = x1 a1 + x2 a2 + · · · + xn an . (11-22)
The coefficients x1 , x2 , . . . , xn in (11-22) are denoted x’s coordinates with respect to the
basis a, or x’s a-coordinates, and they are gathered in a coordinate vector as follows:
x1
x2
a x = .. . (11-23)
.
xn
eNote 11 11.5 VECTOR CALCULATIONS USING COORDINATES 268
In the number space R3 a basis a is given by ((0, 0, 1), (1, 2, 0), (1, −1, 1)). Furthermore the
vector v = (7, 2, 6) is given. Since
we see that
2
av = 3 .
4
The vector (7, 2, 6) therefore has the a-coordinates (2, 3, 4) .
1. a (u + v) = a u + a v
2. a (ku) = k a u
In other words: The coordinates for a vector sum are obtained by adding the coor-
dinates for the vectors, and the coordinates for a vector multiplied by a number are
the coordinates of the vector multiplied by the number.
Proof
See the proof for the corresponding theorem for geometric vectors in 3-space, Theorem 10.38.
The proof for the general case is obtained as a simple extension.
eNote 11 11.6 ON THE USE OF COORDINATE MATRICES 269
We now carry out a vector calculation using coordinates. The example is not particularly
mathematically interesting, but we carry it out in detail in order to demonstrate the technique
of Theorem 11.34.
P( x ) = 3 − 2x + x2 .
When we embark on problems with vectors and use their coordinates with respect to a
given basis it often leads to a system of linear equations which we then solve by matrix
calculations. One matrix is of particular importance, viz. the matrix that is formed by
gathering the coordinate columns of more vectors in a coordinate matrix:
tors is given, then the a-coordinate matrix is formed by gathering the a-coordinate
columns in the given order to form an m×n matrix.
By way of example consider a set of three vectors in R2 : ((1, 2), (3, 4), (5, 6)) . The
coordinate matrix of the set with respect to the standard e-basis for R2 is the 2×3-
matrix
1 3 5
.
2 4 6
We will now show how coordinate matrices emerge in series of examples which we,
for the sake of variation, take from different vector spaces. The methods can directly
be used on other types of vector spaces, and after each example the method is demon-
strated in a concentrated and general form.
It is important for your own understanding of the theory of vector spaces that
you practice and realize how coordinate matrices emerge in reality when you
start on typical problems.
a1 = (1, 1, 1, 1)
a2 = (1, 0, 0, 1)
a3 = (2, 3, 1, 4)
b = (2, −2, 0, 1)
x1 a1 + x2 a2 + x3 a3 = b . (11-24)
eNote 11 11.6 ON THE USE OF COORDINATE MATRICES 271
x1 + x2 + 2x3 = 2
x1 + 3x3 = −2
x1 + x3 = 0
x1 + x2 + 4x3 = 1
We form the augmented matrix of the system of equations and give (without further
details) its reduced row echelon form
1 1 2 2 1 0 0 0
1 0 3 −2 0 1 0 0
T = ⇒ rref(T) = . (11-25)
1 0 1 0 0 0 1 0
1 1 4 1 0 0 0 1
From (11-25) it is seen that the rank of the coefficient matrix of the system of equations
is 3, while the rank of the augmented matrix is 4. The system of equations has therefore
no solutions. This means that (11-24) cannot be solved. We conclude
b∈
/ span{a1 , a2 , a3 } .
NB: In general there can be none, one or infinitely many ways a vector can be written
as linear combinations of the others.
eNote 11 11.6 ON THE USE OF COORDINATE MATRICES 272
Solution: We use theorem 11.17 and try to find three real numbers x1 , x2 of x3 that are
not all equal to 0, but which satisfy
0 0 0
x1 A + x2 B + x3 C = . (11-27)
0 0 0
1 2 −1 0
0 1 −2 0
3 0 9 0
x1 + x2 + x3
=
0
0
0 0
2 3 0 0
2 1 4 0
That is equivalent to the homogeneous system of linear equations with the augmented
matrix that here is written together with reduced row echelon form (details are omitted):
1 2 −1 0 1 0 3 0
0
1 −2 0
0 1 −2 0
3 0 9 0 0 0 0 0
T =
0 ⇒ rref(T) =
. (11-28)
0 0 0 0 0 0 0
2 3 0 0 0 0 0 0
2 1 4 0 0 0 0 0
From (11-28) we see that both the coefficient matrix and the augmented matrix have the
rank 2, and since the number of unknowns is larger, viz. 3, we conclude that Equation
(11-27) has infinitely many solutions , see Theorem 6.33. Hence the three matrices are
linearly dependent. For instance, from rref(T) one can derive that
0 0 0
−3A + 2B + C = .
0 0 0
eNote 11 11.6 ON THE USE OF COORDINATE MATRICES 273
NB: Since the system of equations is homogeneous, there will be either one solution
or infinitely many solutions. If the rank of the coordinate matrix is equal to p, there
is one solution, and this solution must be the zero solution, and the p vectors are
therefore linearly independent. If the rank of the coordinate matrix is less than p,
there are infinitely many solutions, including non-zero solutions, and the p vectors
are therefore linearly dependent.
In an n-dimensional vector space we require n basis vectors, see theorem 11.22. When
one has asked whether a given set of vectors can be a basis, one can immediately con-
clude that this is not the case if the number of vectors in the set is not equal to n. But
if there are n vectors in the set according to theorem 11.29 we need only investigate
whether the set is linear independent, and for this we already have method 11.38. How-
ever we can in an interesting way develop the method further by using the determinant
of the coordinate matrix of the vector set!
Let us e.g. investigate whether the polynomials
P1 ( x ) = 1 + 2x2 , P2 ( x ) = 2 − x + x2 of P3 ( x ) = 2x + x2
form a basis for P2 (R). Since dim( P2 (R)) = 3, the number of polynomials is compatible
with being a basis. In order to investigate whether they also are linearly independent,
we use their coordinate vectors with respect to the monomial basis and consider the
equation:
1 2 0 0
x1 0 + x2 −1 + x3 2 = 0 .
2 1 1 0
The vectors are linearly independent if and only if the only solution is the trivial so-
lution x1 = x2 = x3 = 0 . The equation is equivalent to a homogeneous system of
linear equations consisting of 3 equations in 3 unknowns. The coefficient matrix and
eNote 11 11.6 ON THE USE OF COORDINATE MATRICES 274
As for every homogeneous system of linear equations the right hand side of the aug-
mented matrix consists of only 0’s, therefore ρ(A) = ρ(T), and thus solutions do exist.
There is one solution exactly when ρ(A) is equal to the number of unknowns, that is
3. And this solution must be the zero solution x1 = x2 = x3 = 0 , since Lhom always
contains the zero solution.
Here we can use that A is a square matrix and thus has a determinant. A has full rank
exactly when it is invertible, that is when det(A) 6= 0.
Since a calculation shows that det(A) = 5 we conclude that P1 ( x ), P2 ( x ), P3 ( x ) con-
stitutes a basis for P2 (R) .
An important technical problem for the advanced use of linear algebra is to be able to
calculate new coordinates for a vector when a new basis is chosen. In this context a
particular change of basis matrix plays an important role. We now demonstrate how basis
matrices emerge.
eNote 11 11.6 ON THE USE OF COORDINATE MATRICES 275
In a 3-dimensional vector space V a basis a is given. We now choose a new basis b that
is determined by the a-coordinates of the basis vectors:
1 1 2
a b1 = 1 , a b2 = 0
and a b3 = 3 .
1 2 0
Problem 1: Determine the a-coordinates for a vector v given by the b-coordinates as:
5
b v = −4 . (11-30)
−1
which we below first convert to an a-coordinate vector equation, re-writing the right
hand side as a matrix-vector product, before finally computing the result:
1 1 2
av = 5 1 − 4 0 − 1 3
1 2 0
1 1 2 5 −1
= 1 0 3 −4 = 2 .
1 2 0 −1 −3
Notice that the 3×3-matrix in the last equation is the coordinate matrix for the b-basis
vectors with respect to basis a . It plays an important role, since we apparently can
determine the a-coordinates for v by multiplying b-coordinate vector for v on the left by
this matrix! Therefore the matrix is given the name change of basis matrix. The property of
this matrix is that it translates b-coordinates to a-coordinates, and it is given the symbol
a Mb . The coordinate change relation can then be written in this convenient way
av = a Mb b v . (11-31)
Solution: Since a Mb is the coordinate matrix for a basis, it is invertible, and thus has an
inverse matrix. We therefore use the coordinate change relation (11-31) as follows:
au = a Mb b u ⇔
−1
a Mb au = a Mb −1 a Mb b u ⇔
bu = a Mb −1 a u ⇔
−1
1 1 2 1 11
bu = 1 0 3
2 = −4 .
1 2 0 3 −3
bv = a Mb − 1 a v .
b Ma = (a Mb )−1 .
11.7 Subspaces
Often you encounter that a subset of a vector space is itself a vector space. In Figure 11.3
→ →
are depicted two position vectors OP and OQ that span the plane F :
eNote 11 11.7 SUBSPACES 277
F Q
→ →
Since span{OP, OQ} can be considered to be a (2-dimensional) vector space in its own
right, it is named a subspace of the (3-dimensional) vector space of position vectors in
space.
When one must check whether a subset is a subspace, one only has to check whether
the stability requirements are satisfied:
Proof
Since U satisfies the two stability requirements in 11.1, it only remains to show that U also
satisfies the eight arithmetic rules in the definition. But this is evident since all vectors in U
are also vectors in V where the rules apply.
where a and b are arbitrary real numbers. We try to add two matrices of the type (11-33)
1 2 3 4 4 6
+ =
2 1 4 3 6 4
in both cases the resulting matrix is of type (11-33) and it is obvious that this would also apply
had we used other examples. Therefore M1 satisfies the stability requirements for a vector
space. Thus it follows from theorem 11.42 that M1 is a subspace of R2×2 .
Further remark that M1 is spanned by two linear independent 2×2 matrices since
a b 1 0 0 1
=a +b .
b a 0 1 1 0
where a and b are arbitrary real numbers. We try to add two matrices of the type (11-34)
1 2 2 3 3 5
+ = .
2 0 6 0 8 0
Since 8 6= 3 · 5, this matrix is not of the type (11-34). Therefore M2 is not stable under linear
combinations, and cannot be a subspace.
Proof
The stability requirements are satisfied because 1) the sum of two linear combinations of the
p vectors in itself is a linear combination of them and 2) a linear combination of the p vectors
multiplied by a scalar in itself is a linear combination of them. The rest follows from Theorem
11.42.
The solution set Lhom for a homogeneous system of linear equations with n unknowns
is always a subspace of the number space Rn and the dimension of the subspace is the
same as the number of free parameters in Lhom . We show an example of this below.
eNote 11 11.7 SUBSPACES 280
x1 + 2 x3 − 11 x5 = 0
x2 + 4 x5 = 0
x4 + x5 = 0
We see that Lhom is a span of two vectors in R5 . Then it is according to theorem 11.45
a subspace of R5 . Since the two vectors evidently are linearly independent, Lhom is a 2-
dimensional subspace of R5 , with a basis
(−2, 0, 1, 0, 0) , (11, −4, 0, −1, 1) .
In the following example we will establish a method for how one can determine a basis
for a subspace that is spanned by a number of given vectors in a subspace.
U = span {v1 , v2 , v3 , v4 } .
Let b = (b1 , b2 , b3 ) be an arbitrary vector in U. We thus assume that the following vector
equation has a solution:
x1 v1 + x2 v2 + x3 v3 + x4 v4 = b . (11-36)
By substitution of the five vectors into (11-36), it is seen that (11-36) is equivalent to an
eNote 11 11.7 SUBSPACES 281
Here c1 is placeholder for the number that b1 has been transformed into following the
row operations leading to the reduced row echelon form rref(T). Similarly for c2 . Re-
mark that b3 after the row operations must be transformed into 0, or else ( x1 , x2 , x3 , x4 )
could not be a solution as we have assumed.
But it is in particular the leading 1’s in rref(T) on which we focus! They show that v1
and v2 span all of U, and that v1 and v2 are linear independent. We can convince our-
selves of both by considering equation (11-36) again.
First: Suppose we had only asked whether v1 and v2 span all of U. Then we should
have omitted the terms with v3 and v4 from (11-36), and then we would have obtained:
1 0 c1
rref(T2 ) = 0 1 c2
0 0 0
Secondly: Suppose we had asked whether v1 and v2 are linearly independent. Then we
should have omitted the terms with v3 and v4 from (11-36), and put b = 0. And then
we would have got:
1 0 0
rref(T3 ) = 0 1 0
0 0 0
That shows that the zero vector can only be written as a linear combination of v1 and v2
if both of the coefficients x1 and x2 are 0. And thus we show that v1 and v2 are linearly
independent. In total we have shown that (v1 , v2 ) is a basis for U.
The conclusion is that a basis for U can be singled out by the leading 1’s in rref(T), see
(11-37). The right hand side in rref(T) was meant to serve our argument but its con-
tribution is now unnecessary. Therefore we can summarize the result as the following
method:
eNote 11 11.7 SUBSPACES 282
If in the i’th column in (11-38) there are no leading 1’s, then vi is deleted from the
set (v1 , v2 , . . . , v p ) . The set reduced in this way is a basis for U .
Since the number of leading 1’s in (11-38) is equal to the number of basis vectors in
the chosen basis for U , it follows that
Dim(U ) = ρ v v
a 1 a 2 . . . v
a p . (11-39)
Before we end this eNote, that has cultivated the use of bases and coordinates, we must
admit that not all vector spaces have a basis. Viz. there exist infinite-dimensional vector
spaces.
This we can see through the following example:
All polynomials in the vector space Pn (R) are continuous functions, therefore Pn (R) is an
n+1 dimensional subspace of the vector space C0 (R) of all real continuous functions. Now
consider P(R) , the set of all real polynomials, that for the same reason is also a subspace of
C0 (R). But P(R) must be infinite-dimensional, since it has Pn (R), for every n, as a subspace.
For the same reason C0 (R) must also be infinite-dimensional.
eNote 11 11.7 SUBSPACES 283
Exercise 11.49
By C1 (R) is understood the set of all differentiable functions, with R as their domain, and
with continuous derivatives in R.
eNote 12
Linear Transformations
This eNote investigates an important type of transformation (or map) between vector spaces,
viz. linear transformations. It is shown that the kernel and the range for linear transformations
are subspaces of the domain and the codomain, respectively. When the domain and the
codomain have finite dimensions and a basis has been chosen for each, questions about linear
maps can be standardized. In that case a linear transformation can be expressed as a product
between a so-called standard matrix for the transformation and the coordinates of the vectors
that we want to map. Since standard matrices depend on the chosen bases, we describe how the
standard matrices are changed when one of the bases or both are replaced. The prerequisite for
the eNote is knowledge about systems of linear equations, see eNote 6, about matrix algebra, see
eNote 7 and about vector spaces, see eNote 10.
A map (also known as a function) is a rule f that for every element in a set A attaches an
element in a set B, and the rule is written f : A → B . A is called the domain and B the
codomain.
CPR-numbering is a map from the set of citizens in Denmark into R10 . Note that there
is a 10-times infinity of elements in the codomain R10 , so luckily we only need a small
subset, about five million! The elements in R10 that in a given instant are in use are the
range for the CPR-map.
Elementary functions of the type f : R → R . are simple maps. The meaning of the
eNote 12 12.1 ABOUT MAPS 285
15
1 2
y = f (x) =
12
x −2. (12-1)
2
11
Here the function has the form of a calculation procedure: Square the number, multiply
the result by one half and subtract 2. Elementary
10
functions have a great advantage in
that their graph { ( x, y) | y = f ( x ) } can be9 drawn to give a particular overview of the
map (Figure 12.1). 8
7
Y
6
X
10 8 6 4 2 0 2 4 6
1. Determine the zeros of f . This means9 we must find all x for which f ( x ) = 0. In
the example above the answer is x = −2 and x = 2.
2. Solve for a given b the equation f ( x ) = b . For b = 6 there are in the example two
solutions: x = −4 and x = 4 .
3. Determine the range for f . We must find all those b for which the equation f ( x ) =
b has a solution. In the example the range is [ −2; ∞ [.
In this eNote we look at domains, codomains and ranges that are vector spaces. A map
f : V → W attaches to every vector x in the domain V a vector y = f (x) in the codomain
W . All the vectors in W that are images of vectors in V together constitute the range.
eNote 12 12.2 EXAMPLES OF LINEAR MAPS IN THE PLANE 286
Then e.g.
1 0 2 1 0 2 1 0
5 0
g = 0 3 =
.
0 3 0 0 3 0 0 9
2 0
We investigate in the following a map f that has the geometric vectors in the plane as
both the domain and codomain. For a given geometric vector x we will by x̂ understand
its hat vector, i.e. x rotated π/2 counter-clockwise. Consider the map f given by
y = f (x) = 2 x̂ . (12-3)
To every vector in the plane there is attached its hat vector multiplied (extended) by 2.
In Figure 12.2 two vectors u and v and their images f (u) and f (v) are drawn.
f(u)
f(v)
v
u
O
Figure 12.2 gives rise to a couple of interesting questions: How is the vector sum u + v
mapped? More precisely: How is the image vector f (u + v) related to the two image
vectors f (u) and f (v)? And what is the relation between the image vectors f (ku) and
f (u) , when k is a given real number?
eNote 12 12.2 EXAMPLES OF LINEAR MAPS IN THE PLANE 287
f(u+v)
f(ku)
f(u)
f(u)
f(v)
v
u+v
u
u
O ku
O
+ v = û + v̂.
1. u[
c = kû .
2. ku
f (u + v) = 2u[
+ v = 2(û + v̂) = 2û + 2v̂
= f (u) + f (v)
f (ku) = 2ku
c = 2kû = k (2û)
= k f (u)
eNote 12 12.2 EXAMPLES OF LINEAR MAPS IN THE PLANE 288
Exercise 12.2
f(v)=3v
v
O
Scaling of vectors
Exercise 12.3
In the plane a line l through the origin is given. A map f 2 reflects vectors drawn from the
origin in l :
l
O
f(v)
Reflection of a vector
Exercise 12.4
A map f 3 turns vectors drawn from the origin the angle t about the origin counterclockwise:
f(v)
t
v
Rotation of a vector
All maps mentioned in this section are linear, because they satisfy (12-4) . We now turn
to a general treatment of linear mappings between vector spaces.
L1 : f (u + v) = f (u) + f (v) .
L2 : f (ku) = k f (u) .
The image of a linear combination becomes in a very simple way a linear com-
bination of the images of the vectors that are part of the given linear combina-
tion:
f ( x1 , x2 ) = (0, x1 , x2 , x1 + x2 ) . (12-7)
R2 and R4 are vector spaces and we investigate whether f is a linear map. First we test the
left hand side and the right hand side of L1 with the vectors (1, 2) and (3, 4):
The investigatíon suggests that f is linear. This is now shown generally. First we test L1 :
f ( ( x1 , x2 ) + (y1 , y2 ) ) = f ( x1 + y1 , x2 + y2 ) = (0, x1 + y1 , x2 + y2 , x1 + x2 + y1 + y2 ) .
f ( x1 , x2 ) + f (y1 , y2 ) = (0, x1 , x2 , x1 + x2 ) + (0, y1 , y2 , y1 + y2 )
= (0, x1 + y1 , x2 + y2 , x1 + x2 + y1 + y2 ) .
Then we test L2 :
f ( k · ( x1 , x2 ) ) = f (k · x1 , k · x2 ) = (0, k · x1 , k · x2 , k · x1 + k · x2 ) .
k · f ( x1 , x2 ) = k · (0, x1 , x2 , x1 + x2 ) = (0, k · x1 , k · x2 , k · x1 + k · x2 ) .
That this map is not linear, can be shown by finding an example where either L1 or L2 is not
valid. Below we give an example of a matrix X that does not satisfy g(2X) = 2 g(X) :
1 0 0 2 0 0 2 0 0 2 0
4 0
g 2 =g = 0 0 =
.
0 0 0 0 0 0 0 0 0 0 0
0 0
But
1 0 0 1 0
1 0 0 1 0 2 0
2g =2 0 0 =2
= .
0 0 0 0 0 0 0 0 0 0
0 0
Therefore g does not satisfy the linearity requirements L2 , hence g is not linear.
f P ( x ) = P 0 (1) .
(12-9)
For every second degree polynomial the slope of the tangent at x = 1 is attached. An arbi-
trary second degree polynomial P can be written as P( x ) = ax2 + bx + c , where a, b and c are
real constants. Since P0 ( x ) = 2ax + b we have:
f P( x ) = 2a + b .
f P1 ( x ) + P2 ( x ) = f ( a1 + a2 ) x2 + (b1 + b2 ) x + (c1 + c2 )
= 2( a1 + a2 ) + (b1 + b2 )
= (2a1 + b1 ) + (2a2 + b2 )
= f P1 ( x ) + f P2 ( x ) .
Furthermore for every real number k and every second degree polynomial P( x ):
f k · P( x ) = f k · ax2 + k · bx + k · c
= (2k · a + k · b) = k · (2a + b)
= k · f P( x ) .
eNote 12 12.4 KERNEL AND RANGE 292
It is hereby shown that f satisfies the linearity conditions L1 and L2 , and that f thus is a linear
map.
Exercise 12.9
By C ∞ (R) we understand the vector space consisting of all functions f : R → R that can be
differentiated an arbitrary number of times. One example (among infinitely many) is the sine
function. Consider the map D : C ∞ (R) → C ∞ (R) that to a function f ( x ) ∈ C ∞ (R) assigns its
derivative:
D f (x) = f 0 (x) .
The zeros of an elementary function f : R → R are all the real numbers x that satisfy
f ( x ) = 0 . The corresponding concept for linear maps is called the kernel. The range
of an elementary function f : R → R are all the real numbers b for each of which a
real number x exists such that f ( x ) = b . The corresponding concept for linear maps is
also called the range or image. The kernel is a subspace of the domain and the range is a
subspace of the codomain. This is now shown.
Proof
1) First, the kernel is not empty, as f (0) = 0 by linearity. So we just need to prove that the
kernel of f satisfies the stability requirements, see Theorem 11.42. Assume that x1 ∈ V and
x2 ∈ V, and that k is an arbitrary scalar. Since (using L1 ):
f (kx1 ) = k f (x1 ) = k 0 = 0 ,
the kernel of f is also stable with respect to multiplication by a scalar. In total we had shown
that the kernel of f is a subspace of V .
2) The range f (V ) is non-empty, as it contains the zero vector. We now show that it satisfies
the stability requirements. Suppose that b1 ∈ f (V ) and b2 ∈ f (V ), and that k is an arbitrary
scalar. There exist, according to the definition, see (12.10), vectors x1 ∈ V and x2 ∈ V that
satisfy f (x1 ) = b1 and f (x2 ) = b2 . We need to show that there exists an x ∈ V such that
f (x) = b1 + b2 . There is, namely x = x1 + x2 , since
Hereby it is shown that f (V ) is stable with respect to addition. We will, in a similar way,
show that there exists an x ∈ V such that f (x) = kb1 . Here we choose x = kx1 , then
from which it appears that f (V ) is stable with respect to multiplication by a scalar. In total
we have shown that f (V )is a subspace of W.
eNote 12 12.4 KERNEL AND RANGE 294
But why is it so interesting that the kernel and the range of a linear map are subspaces?
The answer is that it becomes simpler to describe them when we know that they possess
vector space properties and we thereby in advance know their structure. It is particu-
larly elegant when we can determine the kernel and the range by giving a basis for
them. This we will try in the next two examples.
There is an entire plane of vectors in the space, that by insertion into the expression
for f give the image 0. This basis yields all of them.
Note, that it is not x1 , x2 and x3 , we are looking for, as we usually do in such a system
of equations. Rather it is b1 and b2 of the right hand side, which we will determine
exactly in those cases, when solutions exist! Because when the system has solution of a
particular right hand side, then this right-hand side must be in the image space that
we are looking for.
This is a system of linear equations consisting of two equations in three unknowns. The
corresponding augmented matrix is
1 2 1 b1 1 2 1 b1
T= → rref(T) =
−1 −2 −1 b2 0 0 0 b1 + b2
O 1 X
In example 12.8 it was shown that the map f : P2 (R) → R given by the rule
f P ( x ) = P 0 (1) .
(12-15)
is linear. The kernel of f consists of all second degree polynomials that satisfy P0 (1) = 0 .
The graphs for a couple of these are shown in Figure 12.4:
In eNote 6 the relation bewteen the solution set for an inhomogeneous system of linear
equations and the corresponding homogeneous linear system of equations is presented
in Theorem 6.37 (the structural theorem). We now show that a corresponding relation
exists for all linear equations.
or in short
Linhom = x0 + ker( f ) . (12-18)
Proof
The theorem contains two assertions. The one is that the sum of x0 and an arbitrary vector
from the ker( f ) belongs to Linhom . The other is that an arbitrary vector from Linhom can be
written as the sum of x0 and a vector from ker( f ) . We prove the two assertions separately:
eNote 12 12.5 MAPPING MATRIX 297
x2 − x0 = x1 ⇔ x2 = x0 + x1 (12-21)
whereby we have stated x2 in the form wanted. The proof is hereby complete.
Exercise 12.15
Consider the map D : C ∞ (R) → C ∞ (R) from Exercise 12.9 that to the function f ∈ C ∞ (R)
relates its derivative:
D f (x) = f 0 (x) .
D( f ( x )) = x2
All linear maps from a finite dimensional domain V to a finite dimensional codomain
W can be described by a mapping matrix. This is the subject of this subsection. The
prerequisite is only that a basis for both V and W is chosen, and that we turn from vector
calculation to calculation using the coordinates with respect to the chosen bases. The
great advantage by this setup is that we can construct general methods of calculation
valid for all linear maps between finite dimensional vector spaces. We return to this
subject, see section 12.6. First we turn to mapping matrix construction.
eNote 12 12.5 MAPPING MATRIX 298
y = f (x) = A x . (12-22)
Using the matrix product computation rules from Theorem 7.13, we obtain for every
choice of x1 , x2 ∈ Rn and every scalar k:
The formula:
y1 1 2 x1 + 2x2
y2 = 3 4 x1 = 3x1 + 4x2
x2
y3 5 6 5x1 + 6x2
defines a particular linear map from the vector space R2 to the vector space R3 .
But also the opposite is true: Every linear map between finite-dimensional vector spaces
can be written as a matrix-vector product in the form (12-22) if we replace x and y with
their coordinates with respect to a chosen basis for the domain and codomain, respec-
tively. This we show in the following.
For V a basis a is chosen and for W a basis c. This means that a given vector x ∈ V
can be written as a unique linear combination of the a-basis vectors and that the image
y = f (x) can be written as a unique linear combination of the c-basis vectors:
x = x1 a1 + x2 a2 + . . . + xn an and y = y1 c1 + y2 c2 + · · · + ym cm .
This means that ( x1 , x2 , . . . , xn ) is the set of coordinates for x with respect to the a-basis,
and that (y1 , y2 , . . . , ym ) is the set of coordinates for y with respect to the c-basis.
eNote 12 12.5 MAPPING MATRIX 299
x
y = f(x)
We now pose the question: How can we describe the relation between the a-coordinate
vector for the vector x ∈ V and the c-coordinate vector for the image vector y? In other
words we are looking for the relation between:
y1 x1
y2
x2
cy = .. and a x = .. .
. .
ym xn
This we develop through the following rewritings where we first, using L1 and L2 , get
y written as a linear combination of the images of the a-vectos.
y = f (x)
= f ( x1 a1 + x2 a2 + · · · + xn an )
= x1 f (a1 ) + x2 f (a2 ) + · · · + xn f (an ) .
Hereafter we can investigate the coordinate vector for y with respect to the c-basis, while
we first use the cooordinate theorem, see Theorem 11.34, and thereafter the definition
on matrix-vector product, see Definition 7.7.
c y = c x1 f (a1 ) + x2 f (a2 ) + · · · + xn f (an )
= x1 c f (a1 ) + x2 c f (a2 ) + · · · + xn c f (an )
= c f ( a1 ) c f ( a2 ) · · · c f ( a n ) a x .
eNote 12 12.5 MAPPING MATRIX 300
The matrix c f (a1 ) c f (a2 ) · · · c f (an ) in the last equation is called the mapping ma-
trix for f with respect to the a-basis for V and the c-basis for W.
Thus we have achieved this important result: The coordinate vector c y can be found
by multiplying the coordinate vector a x on the left by the mapping matrix. We now
summarize the results in the following.
The mapping matrix for f thus consists of the coordinate vectors with respect to the
basis c of the images of the n basis vectors in basis a.
The main task for a mapping matrix is of course to determine the images in W of the
vectors in V, and this is justified in the following theorem which summarizes the inves-
tigations above.
eNote 12 12.5 MAPPING MATRIX 301
cy = c Fa a x (12-24)
where c Fa is the mapping matrix for f with respect to the basis a of V and the
basis c of W .
cy = c Ga a x (12-25)
where c Ga ∈ Lm×n . Then g is linear and c Ga is the mapping matrix for g with
respect to the basis a of V and basis c of W .
Below are three examples of the construction and elementary use of mapping matrices.
f(x)
x
j
v
i
O
Rotation of plane vectors drawn from the origin is a simple example of a linear map, see
Exercise 12.4. Let v be an arbitrary angle, and let f be the linear mapping that rotates an
arbitrary vector the angle v about the origin counterclockwise, (see the figure above).
We wish to determine the mapping matrix for f with respect to the standard basis for vectors
in the plane. Therefore we need the images of the basis vectors i and j :
j f(i)
v
f(j) v
O i
It is seen that f (i) = (cos(v), sin(v)) and f (j) = (− sin(v), cos(v)). Therefore the mapping
matrix we are looking for is
cos(v) − sin(v)
e Fe = .
sin(v) cos(v)
The coordinates for the image y = f (x) of a given vector x are thus given by the formula:
y1 cos(v) − sin(v) x1
= .
y2 sin(v) cos(v) x2
We wish to find the image under f of the vector v = a1 + 2a2 + a3 ∈ V using the mapping
matrix c Fa . The mapping matrix is easily constructed since we already from (12-26) know
the images of the basis vectors in V :
3 6 −3
F
c a = c f ( a 1 ) c f ( a2 ) c f ( a 3 ) =
1 −2 1
eNote 12 12.5 MAPPING MATRIX 303
Since v has the set of coordinates (1, 2, 1) with respect to basis a, we find the coordinate vector
for f (v) like this:
1
3 6 −3 12
c f (v) = c Fa a v = 2 = .
1 −2 1 −2
1
Hence we have found f (v) = 12c1 − 2c2 .
Let us determine the mapping matrix for f with respect to the standard basis e of R4 and the
standard basis e of R3 . First we find the images of the four basis vectors in R4 using the rule
(12-27):
1 2
f (1, 0, 0, 0) = 2 , f (0, 1, 0, 0) = −1 ,
1 −3
0 1
f (0, 0, 1, 0) = 2 , f (0, 0, 0, 1) = −1 .
2 −2
We can now construct the mapping matrix for f :
1 2 0 1
e Fe = 2 −1 2 −1 . (12-28)
1 −3 2 −2
We wish to find the image y = f (x) of the vector x = (1, 1, 1, 1). At our disposal we have of
course the rule (12-27), but we choose to find the image using the mapping matrix:
1
1 2 0 1 4
1
e y = e Fe e x = 2 −1 2 −1 = 2 .
1
1 −3 2 −2 −2
1
Thus we have found that y = f (1, 1, 1, 1) = (4, 2, −2) .
eNote 12 12.6 ON THE USE OF MAPPING MATRICES 304
Exercise 12.22
In the plane is given a customary (O, i, j)-coordinate system. Reflection of position vectors
about the line y = 12 x is a linear map, let us call it s .
Determine s(i) and s(j), construct the mapping matrix e Se for s and determine an expression
k for the reflection of an arbitrary position vector v with the coordinates (v1 , v2 ) with respect
to the standard basis. The figure below contains some hints for the determination of s(i).
Proceed similarly with s(j) .
Y
y =½ x
s(i) P
j
t
t X
O i E
s(j)
The mapping matrix tool has a wide range of applications. It allows us to translate
questions about linear maps between vector spaces to questions about matrices and
coordinate vectors that allow immediate calculations. The methods only require that
bases in each of the vector spaces be chosen, and that the mapping matrix that belongs
to the two bases has been formed. In this way we can reduce problems as diverse as
that of finding polynomials with certain properties, finding the result of a geometrical
construction and finding the solution of differential equations, to problems that can be
solved through the use of matrix algebra.
3-dimensional vector space with chosen basis c = (c1 , c2 , c3 ) . The mapping matrix for f
is:
1 3 −1 8
c Fa = 2 0 4 −2 . (12-29)
1 −1 3 −4
To obtain the kernel of f you must find all the x ∈ V that are mapped to 0 ∈ W. That is
you solve the vector equation
f (x) = 0 .
This equation is according to the Theorem 12.18 equivalent to the matrix equation
= c0
c Fa a x
x1
1 3 −1 8 0
x2
⇔ 2
0 4 −2 = 0
x3
1 −1 3 −4 0
x4
that corresponds to the homogeneous system of linear equations with the augmented
matrix:
1 3 −1 8 0 1 0 2 −1 0
T = 2 0 4 −2 0 → rref(T) = 0 1 −1 3 0.
1 −1 3 −4 0 0 0 0 0 0
It is seen that the solution set is spanned by two linear independent vectors: (−2, 1, 1, 0)
and (1, −3, 0, 1). Let v1 and v2 be the two vectors in V that are determined by the a-
coordinates like this:
How can you decide whether a vector b ∈ W belongs to the image for a given linear
map? The question is whether (at least) one x ∈ V exists that is mapped to b . And
the question can be extended to how to determine all x ∈ V with this property that is
mapped in b .
Since the rank of the augmented matrix is larger than the rank of the coefficient matrix,
the inhomogeneous system of equations has no solutions. We have thus found a vector
in W that has no “original vector” in V.
f (x) = b
can be solved using the inhomogeneous system of linear equations that has the aug-
mented matrix
T = c Fa c b
If solutions exist and x0 is one of these solutions the whole solution set can be written
as:
x0 + ker( f ) .
f (x) = Ax
Above we have found that the image space for a linear map is a subspace of the codomain,
see theorem 12.11. How can this subspace be delimited and characterized?
Again we consider the linear map f : V → W that is represented by the mapping ma-
trix (12-29). Since the basis (a1 , a2 , a3 , a4 ) for V is chosen we can write all the vectors in
V at once:
x = x1 a1 + x2 a2 + x3 a3 + x4 a4 ,
where we imagine that x1 , x2 , x3 og x4 run through all conceivable combinations of real
values. But then all images in W of vectors in V can be written as:
f (x) = f ( x1 a1 + x2 a2 + x3 a3 + x4 a4 )
= x1 f (a1 ) + x2 f (a2 ) + x3 f (a3 ) + x4 f (a4 ) ,
where we have used L1 og L2 , and where we continue to imagine that x1 , x2 , x3 and x4
run through all conceivable combinations of real values. But then:
f (V ) = span { f (a1 ), f (a2 ), f (a3 ), f (a4 ) } .
The image space is thus spanned by the images of the a-basis vectors! But then we can
(according to Method 11.47Method in eNote 11) determine a basis for the image space
by finding the leading 1’s in the reduced row echelon form of
c f(a1 ) c f(a2 ) c f(a3 ) c f(a4 ) .
This is the mapping matrix for f with respect to the chosen bases
1 3 −1 8
c Fa = 2 0 4 −2
1 −1 3 −4
that by Gauss-Jordan elimination is reduced to
1 0 2 −1
rref(c Fa ) = 0 1 −1 3.
0 0 0 0
To the two leading 1’s in rref(c Fa ) correspond the first two columns in c Fa . We thus
conclude:
rref(c Fa ) (12-31)
in the following way: If there is no leading 1 in the i’th column in (12-31) then f (a i )
is removed from the vector set f (a1 ), f (a2 ), . . . , f (a n ) . After this thinning the
vector set constitutes a basis for f (V ) .
Since the number of leading 1’s in (12-31) is equal to the number of basis vectors in
the chosen basis for f (V ) it follows that
dim( f (V )) = ρ (c Fa ) . (12-32)
In the method of the preceding section 12.23 we found the following expression for the
dimension of the kernel of a linear map f : V → W :
dim( f (V )) = ρ (c Fa ) . (12-34)
By combining (12-33) and (12-34) a remarkably simple relationship between the domain,
the kernel and the image space for a linear map is achieved:
The image space for a linear map can never have a higher dimension than the
domain.
If the kernel only consists of the 0-vector, the image space keeps the dimension
of the domain.
If the kernel has the dimension p > 0, then p dimensions disappear through the
map.
Exercise 12.27
A linear map f : R3 → R3 has, with respect to the standard basis for R3 , the mapping matrix
1 2 1
e Fe = 2 4 0 .
3 6 0
It is stated that the kernel of f has the dimension 1. Find by mental calculation, a basis for
f (V ) .
eNote 12 311
12.8 CHANGE IN THE MAPPING MATRIX WHEN THE BASIS IS CHANGED
Exercise 12.28
In 3-space a standard (O, i, j, k)-coordinate system is given. The map p projects position
vectors down into ( x, y)-plane in space:
v
k
O Y
i j
p(v)
Show that p is linear and construct the mapping matrix e Pe for p. Determine a basis for the
kernel and the image space of the projection. Check that the Dimension Theorem is fulfilled.
In eNote 11 it is shown how the coordinates of a vector change when the basis for the
vector space is changed, see method 11.40. We begin this section by repeating the most
important points and showing two examples.
Assume that in V an a-basis (a1 , a2 , . . . , an ) is given, and that a new b-basis (b1 , b2 , . . . , bn )
is chosen in V. If a vector x has the b-coordinate vector b x , then its a-coordinate vector
can be calculated as the matrix vector-product
av = a Mb b v (12-35)
We now show two examples of the use of (12-36). In the first example the “new” coor-
dinates are given following which the “old” are calculated. In the second example it is
vice versa: the “old” are known, and the “new” are determined.
2 −3
1 1
and a Mb = 0 −2
bx = 2 3 . (12-37)
3 −1 1 −1
Then we get
2 −3 −4
1 1
a x = a Mb b x = 0 −2 3 2 = 5 . (12-38)
−1 1 −1 3 −2
In a 2-dimensional vector space W a basis c = (c1 , c2 ) is given, following which a new basis
d is chosen consisting of the vectors
d1 = 2c1 + c2 and d2 = c1 + c2 .
Then we get
1 −1 10 4
d y = d Mc c y = = . (12-40)
−1 2 6 2
eNote 12 313
12.8 CHANGE IN THE MAPPING MATRIX WHEN THE BASIS IS CHANGED
We now continue to consider how a mapping matrix is changed when the basis for the
domain or the codomain is changed.
For two vector spaces V and W with finite dimension the mapping matrix for a linear
map f : V → W can only be constructed when a basis for V and a basis for W are cho-
sen. By using the mapping matrix symbol c Fa we show the foundation to be the pair of
given bases a of V and c of W.
Often one wishes to change the basis of V or the basis of W. In the first case the co-
ordinates for those vectors x ∈ V will change while the coordinates for their images
y = f (x) are unchanged; in the second case it is the other way round with the x coor-
dinates remaining unchanged while the image coordinates change. If the bases of both
Vand W are changed then the coordinates for both x and y = f (x) change.
In this section we construct methods for finding the new mapping matrix for f , when
we change the basis for either the domain, the codomain or both. First we show how
a vector’s coordinates change when the basis for the domain is changed (as in detail in
Method 11.40 in eNote 11.)
x
y = f(x)
In Figure 12.6 a linear map f : V → W is given that, with respect to basis a of V and
basis c of W, has the mapping matrix c Fa . We change the basis for V from basis a to
basis b. The mapping matrix for f now has the symbol c Fb . Let us find it. The equation
y = f (x)
cy = c Fa a x = c Fa (a Mb b x) = (c Fa a Mb ) b x .
From this we deduce that the mapping matrix for f with respect to the basis b of V and
basis c of W is formed by a matrix product:
c Fb = c Fa a Mb . (12-41)
We consider the 3-dimensional vector space V that is treated in Example 12.29 and the 2-
dimensional vector space W that is treated in Example 12.30. A linear map
f : V → W is given by the mapping matrix:
9 12 7
c Fa = .
6 8 5
Solution: We try two different ways. 1) We use a-coordinates for x as found in (12-37):
−4
9 12 7 10
c y = c Fa a x = 5 = .
6 8 5 6
−2
2 −3
1
9 12 7 2 1 2
c Fb = c Fa a Mb = 0 −2 3 = .
6 8 5 1 1 1
−1 1 −1
x
y = f(x)
In Figure 12.7 a linear map f : V → W is given that, with respect to the basis a of V and
basis c of W has a mapping matrix c Fa . We change the basis for W from basis c to basis
d . The mapping matrix for f now has the symbol d Fa . Let us find it. The equation
y = f (x)
cy = c Fa a x
that is equivalent to
d Mc c y = d Mc (c Fa a x)
from which we get that
dy = (d Mc c Fa ) a x .
From this we deduce that the mapping matrix for f with respect to the a-basis for V and
the d-basis for W is formed by a matrix product:
d Fa = d Mc c Fa . (12-42)
eNote 12 316
12.8 CHANGE IN THE MAPPING MATRIX WHEN THE BASIS IS CHANGED
We consider the 3-dimensional vector space V that is treated in Example 12.29 and the 2-
dimensional vector space W that is treated in Example 12.30. A linear map
f : V → W is given by the mapping matrix:
9 12 7
c Fa = .
6 8 5
Problem: Given the vector x = −4a1 + 5a2 − 2a3 . Determine the image y = f (x) as a linear
combination of d1 and d2 .
−4
9 12 7 10
c y = c Fa a x = 5 =
.
6 8 5 6
−2
−4
3 4 2 4
d y = d Fa a x = 5 = .
3 4 3 2
−2
In Figure 12.8 a linear map f : V → W is given that, with respect to the basis a for V
and basis c for W, has the mapping matrix c Fa . We change the basis for V from basis
a to basis b, and for W from basis c to basis d . The mapping matrix for f now has the
eNote 12 317
12.8 CHANGE IN THE MAPPING MATRIX WHEN THE BASIS IS CHANGED
x
y = f(x)
From here we deduce that the mapping matrix for f with respect to b-basis of V and
d-basis of W is formed by a matrix product:
d Fb = d Mc c Fa a Mb . (12-43)
We consider the 3-dimensional vector space V that is treated in example 12.29, and the 2-
dimensional vector space W that is treated in example 12.30. A linear map f : V → W is
given by the mapping matrix:
9 12 7
F
c a = .
6 8 5
eNote 12 318
12.8 CHANGE IN THE MAPPING MATRIX WHEN THE BASIS IS CHANGED
Problem: Given the vector x = b1 + 2b2 + 3b3 . Determine y = f (x) as a linear combination
of d1 and d2 .
2 −3
1
1 −1 9 12 7 1 0 1
d Fb = d Mc c Fa a Mb = 0 −2 3 = .
−1 2 6 8 5 0 1 0
−1 1 −1
Then we can directly use the given b-coordinates and directly read the d-coordinates:
1
1 0 1 4
d y = d Fb b x = 2 = .
0 1 0 2
3
The change of basis in this example turns out to be rather practical. With the new mapping
matrix d Fb it is much easier to calculate the image vector: You just add the first and the third
coordinates of the given vector and keep the second coordinate!
We gather the results concerning change of basis in the subsections above in the follow-
ing method:
eNote 12 319
12.8 CHANGE IN THE MAPPING MATRIX WHEN THE BASIS IS CHANGED
c Fb = c Fa a Mb . (12-44)
d Fa = (c Md )−1 c Fa = d Mc c Fa . (12-45)
d Fb = (c Md )−1 c Fa a Mb = d Mc c Fa a Mb . (12-46)
eNote 13
This note introduces the concepts of eigenvalues and eigenvectors for linear maps in arbitrary
general vector spaces and then delves deeply into eigenvalues and eigenvectors of square
matrices. Therefore the note is based on knowledge about general vector spaces, see eNote 11, on
knowledge about algebra with matrices, see eNote 7 and eNote 8, and on knowledge about linear
maps see eNote 12.
13.1.1 Introduction
f : V → V, (13-1)
that is, linear maps where the domain and the codomain are the same vector space. This
gives rise to a special phenomenon, that a vector can be equal to its image vector:
f (v) = v . (13-2)
Vectors of this type are called fixed points of the map f . More generally we are looking
for eigenvectors, that is vectors that are proportional to their image vectors. In this
eNote 13 13.1 THE EIGENVALUE PROBLEM FOR LINEAR MAPS 321
connection one talks about the eigenvalue problem: to find a scalar λ and a proper (i.e.
non-zero) vector v satisfying the vector equation:
f (v) = λv . (13-3)
If λ is a scalar and v a proper vector satisfying 13-3 the proportionality factor λ is called
an eigenvalue of f and v an eigenvector corresponding to λ. Let us, for example, take
a linear map f : G3 → G3 , that is, a linear map of the set of space vectors into itself,
mapping three given vectors as shown in Figure 13.1.
c
f(c)
a
f(a)
f(b)
To solve eigenvalue problems for linear maps is one of the most critical problems in
engineering applications of linear algebra. This is closely connected to the fact that a
linear map whose mapping matrix with respect to a given basis is a diagonal matrix is
particularly simple to comprehend and work with. And here the nice rule, that if one
chooses a basis consisting of eigenvectors for the map, then the mapping matrix auto-
matically becomes a diagonal matrix.
In the following example we illustrate these points using linear maps in the plane.
eNote 13 13.1 THE EIGENVALUE PROBLEM FOR LINEAR MAPS 322
The vector space of vectors in the plane has the symbol G2 (R) . We consider a linear map
of the set of plane vectors into itself, that with respect to a given basis (a1 , a2 ) has the follow-
ing diagonal matrix as its mapping matrix:
2 0
a Fa = . (13-5)
0 3
Since
2 0 1 2 1
a f ( a1 ) = = = 2·
0 3 0 0 0
and
2 0 0 0 0
a f ( a2 ) = = = 3·
0 3 1 3 1
we have that f (a1 ) = 2a1 and f (a2 ) = 3a2 . Both basis vectors are thus eigenvectors for f ,
because a1 corresponds to the eigenvalue 2 and a2 corresponds to the eigenvalue 3 . The
eigenvalues are the diagonal elements in a Fa .
By the map the x1 -coordinate is multiplied by the eigenvalue 2, while the x2 -coordinate is
multiplied by the eigenvalue 3. Geometrically this means that through the map all of the
plane “is stretched” first by the factor 2 in the direction a1 and then by the factor 3 in the
direction a2 , see the effect on an arbitrarily chosen vector x in the figure A:
f(x)
a2 x
O a1
12
In Figure B we have chosen the standard basis (i, j) and illustrate how the linear map g that
has the mapping matrix
2 0
e Ge = ,
10
0 3
maps the “blue house” into the “red house” by stretching all position vectors in the blue
house by the factor 2 in the horizontal8 direction and by the factor 3 in the vertical direction.
j
5 O i 5 10
Figure B: The blue house is stretched 2in the horizontal direction by the factor 2 and vertically
by the factor 3.
14
4
We now investigate another map h , the mapping matrix of which, with respect to the stan-
dard basis, is not a diagonal matrix:
12
6
7/3 2/3
H
e e = .
10
1/3 8/3
Here it is not possible to decide directly
8 whether the map is composed of two stretchings in
two given directions. And the mapping of the blue house by h as shown in the figure below
8
does not give a clue directly:
10
j
5 i 5 10
2 Figure C: House
6
eNote 13 13.1 THE EIGENVALUE PROBLEM FOR LINEAR MAPS 324
But it is actually also possible in the case of h to choose a basis consisting of two linearly
independent eigenvectors for h . Let b1 be given by the e-coordinates (2, −1) and b2 by the
e-coordinates (1, 1) . Then we find that
7/3 2/3 2 4 2
e h ( b1 ) = = = 2·
1/3 8/3 −1 −2 −1
and
7/3 2/3 1 3 1
e h (b2 ) = = = 3· .
1/3 8/3 1 3 1
In other words, h(b1 ) = 2b1 and h(b2 ) = 3b2 . We see that b1 and b2 are eigenvectors for
h , and when we choose (b1 , b2 ) as basis, the mapping matrix for h with respect to this basis
12
takes the form:
2 0
b Gb = .
10
0 3
Surprisingly it thus shows that the mapping matrix for h also can be written in the form
(13-5). The map h is also8 composed of two stretchings with the factors 2 and 3. Only the
stretching directions are now determined by the eigenvectors b1 and b2 . This is more evident
if we map a new blue house whose principal lines are parallel to the b-basis vectors:
6
b2
5 O b 5 10
1
Figure D: The blue house is stretched by the factor 2 and the factor 3, respectively, in the
6
directions of the eigenvectors
Thus we have illustrated: 8 If you can find two linearly independent eigenvectors for a linear
map in the plane it is possible:
10
1. to write its mapping matrix in diagonal form by choosing the eigenvectors as basis
12
2. to describe the map as stretchings in the directions of the eigenvectors with the corre-
sponding eigenvectors as stretching factors.
eNote 13 13.1 THE EIGENVALUE PROBLEM FOR LINEAR MAPS 325
The eigenvalue problem for a linear map is briefly about answering the question: do any
proper vectors, each with its image vector proportional to the vector itself, exist. The
short answer to this is that this cannot be answered in general, it depends on the partic-
ular map. In the following we try to pinpoint what can actually be said generally about
the eigenvalue problem.
f (v) = λv , (13-6)
If, in Definition 13.2, it were not required to find a proper vector that satisfies
f (v) = λv , then every scalar λ would be an eigenvalue, since for any scalar λ
f (0) = λ 0 is valid. On the other hand, for a given eigenvalue, it is a matter
of convention whether or not to say that the zero vector is also a correspond-
ing eigenvector. Most commonly, the zero vector is not considered to be an
eigenvector.
If a linear map f has one eigenvector v , then it has infinitely many eigenvectors. This
is a simple consequence of the following theorem.
Proof
Let f : V → V be a linear map of the vector space V into itself, and assume that λ is an
eigenvalue of f . Obviously Eλ is not empty, since it contains the zero vector. We shall show
that the it satisfies the two stability requirements for subspaces, see Theorem 11.42. Let k be
an arbitrary scalar, and let u and v be two arbitrary elements of Eλ . Then the following is
valid :
f (u + v) = f (u) + f (v) = λu + λv = λ(u + v) .
Thus the vector sum u + v ∈ Eλ and thus we have shown that Eλ satisfies the stability re-
quirement with respect to addition. Furthermore the following is valid:
Thus we have shown stabilit with respect to multiplication by a scalar. Together we have
shown that Eλ is a subspace of the domain.
In the following example we consider a linear map that has two eigenvalues, both with
the geometric multiplicity 1 .
eNote 13 13.1 THE EIGENVALUE PROBLEM FOR LINEAR MAPS 327
In the plane a straight line through the origin is drawn. By s we denote the linear map that
maps a vector v, drawn from the origin, in its reflection s(v) in m :
v
m
s(b)
O
a
s(v)
s(a)
b
s (a) = a = 1 · a
We now draw a straight line n through the origin, perpendicular to m . Let b be an arbitrary
proper vector lying on n. Since
s(b) = −b = (−1) · b ,
That not all linear maps have eigenvalues and thus eigenvectors is evident from the
following example.
eNote 13 13.1 THE EIGENVALUE PROBLEM FOR LINEAR MAPS 328
Example 13.6
Let us investigate the eigenvalue problem for the linear map f : G2 → G2 that to every proper
vector v in the plane assigns its hat vector:
f (v) = v
b.
Since a proper vector v never can be proportional (parallel) to its hat vector, then for any
scalar λ we have
b 6= λv .
v
Therefore eigenvalues and eigenvectors for f do not exist.
From the following exercise we see that the dimension of an eigenspace can be greater
than 1.
Exercise 13.7
In space an ordinary (O, i, j, k)-coordinate system is given. All vectors are drawn from the
origin. The map p projects vectors down onto the ( X, Y )-plane in space:
v
k
O Y
i j
p(v)
It is shown in Exercise 12.28 that p is linear. Determine all eigenvalues and the eigenspaces
that correspond to the eigenvalues, solely by mental calculation (ponder).
eNote 13 13.1 THE EIGENVALUE PROBLEM FOR LINEAR MAPS 329
f ( x (t)) = x 0 (t) .
x 0 (t) = λx (t)
The following corollary gives an important result for linear maps of a vector space into
itself. It is valid even if the vector space considered is of infinite dimension.
Corollary 13.9
Let f : V → V be a linear map of a vector space V into itself, and assume
2. that some of the eigenspaces are chosen, and within each of the chosen
eigenspaces some linearly independent vectors are chosen,
3. and that all the so chosen vectors are consolidated in a single set of vectors v .
Proof
Let f : V → V be a linear map, and let v be a set of vectors that are put together according to
points 1. to 3. in Corollary 13.9. We shall prove that v is linearly independent. The flow of
the proof is that we assume the opposite, that is, v is linearly dependent, and show that this
leads to a contradiction.
First we delete vectors from v to get a basis for span{v}. There must be at least one vector in
v that does not correspond to the basis. We choose one of these, let us call it x. Now we write
x as a linear combination of the basis vectors, in doing so we leave out the trivial terms, i.e.
those with the coefficient 0:
x = k 1 v1 + · · · + k m v m (13-7)
We term the eigenvalue that corresponds to x λ, and the eigenvalues corresponding to vi λi .
From (13-7) we can obtain an expression for λx in two different ways, partly by multiplying
(13-7) by λ, partly by finding the image by f of the right and left hand side in (13-7):
λx = λk1 v1 + · · · + λk m vm
λx = λ1 k1 v1 + · · · + λm k m vm
If all the coefficients to the vectors on the right hand side of (13-8) are equal to zero, then
λ = λi for all i = 1, 2, . . . , m. But then x and all the basis vectors vi that are chosen form the
same eigenspace, and therefore they should collectively be linearly independent, this is how
they are chosen. This contradicts that x is a linear combination of the basis vectors.
Therefore at least one of the coefficients in (13-8) must be different from 0. But then the zero
vector is written as a proper linear combination of the basis vectors. This contradicts the
requirement that a basis is linearly independent.
Conclusion: the assumption that v is a linearly independent set of vectors, necessarily leads
to a contradiction. Therefore v is linearly independent.
A linear map f : V → V has three eigenvalues λ1 , λ2 and λ3 that have the geometric multi-
plicities 2, 1 and 3 , respectively. The set of vectors (a1 , a2 ) is a basis for Eλ1 , (b) is a basis for
eNote 13 13.1 THE EIGENVALUE PROBLEM FOR LINEAR MAPS 331
Eλ2 , and (c1 , c2 , c3 ) is a basis for Eλ3 . Then it follows from corollary 13.9 that any selection of
the six basis vectors is a linearly independent set of vectors.
Corollary 13.9 is useful because it leads directly to the following important results:
4. The sum of the geometric multiplicities of eigenvalues for f can at the most be
n.
5. If and only if the sum of the geometric multiplicities of the eigenvalues for f is
equal to n, a basis for V exists consisting of eigenvectors for f .
Exercise 13.12
The first point in 13.11 is a simple special case of Corollary 13.9 and therefore follows directly
from the corollary. The second point can be proved like this:
Assume that a linear map has k different eigenvalues. We choose a proper vector from each of the k
eigenspaces. The set of the k chosen vectors is then (in accordance with the corollary 13.9) linearly
independent, and k must therefore be less than or equal to the dimension of the vector space (see
Corollary 11.21).
Similarly, show how the last three points in Theorem 13.11 follow from Corollary 13.9.
1. The mapping matrix v Fv for f with respect to v is a diagonal matrix if and only if
v is an eigenbasis for V with respect to f .
2. Assume that v is an eigenbasis for V with respect to f . Let Λ denote the diagonal
matrix that is the mapping matrix for f with respect to v . The order of the diago-
nal elements in Λ is then determined from the basis like this: The basis vector vi
corresponds to the eigenvalue λi that is in the i’th column in Λ .
The proof of this theorem can be found in eNote 14 (see Theorem 14.7).
eNote 13 13.1 THE EIGENVALUE PROBLEM FOR LINEAR MAPS 333
Let us again consider the situation in example 13.5, where we considered the map s that
reflects vectors drawn from the origin in the line m :
v
m
s(b)
O
a
s(v)
s(a)
b
Reflection about m.
In the example 13.6 we found that the map, which maps a vector in the plane onto its hat
vector, has no eigenvalues. Therefore there is no eigenbasis for the map, and therefore it
cannot be described by a diagonal matrix for this map.
Since:
f (i, 1) = (−1, i ) = i (i, 1) and f (−i, 1) = (−1, −i ) = (−i )(−i, 1) ,
it is seen that i is an eigenvalue of f with a corresponding eigenvector (i, 1), and that −i is an
eigenvalue of f with a corresponding eigenvector (−i, 1) .
Since (i, 1) and (−i, 1) are linearly independent, (i, 1), (−i, 1) is an eigenbasis for C2 with
respect to f . The mapping matrix for f with respect to this basis is in accordance with Theo-
rem 13.14
i 0
.
0 −i
Exercise 13.18
Consider once more the situation in Example 13.7. Choose two different eigenbases (bases
consisting of eigenvectors for p ), and determine in each of the two cases the diagonal matrix
that will become the mapping matrix for p with respect to the chosen basis.
When a linear map f : V → V maps an n-dimensional vector space V into the vec-
tor space itself the mapping matrix for f with respect to the arbitrarily chosen basis a
becomes a square matrix. The eigenvalue problem f (v) = λv is the equivalent of the
matrix equation:
a Fa ·a v = λ · a v . (13-9)
Thus we can formulate an eigenvalue problem for square matrices generally, that is
without necessarily having to think about a square matrix as a mapping matrix. We will
standardize the method, when eigenvalues and eigenvectors for square matrices are to
be determined. At the same time, due to (13-9), we get methods for finding eigenvalues
and eigenvectors for all linear maps of a vector space into itself, that can be described
by mapping matrices.
First we define what is to be understood by the eigenvalue problem for a square matrix.
eNote 13 13.2 THE EIGENVALUE PROBLEM FOR SQUARE MATRICES 335
A v = λv . (13-10)
We wish to investigate whether v1 = (2, 3), v2 = (4, 4) and v3 = (2, −1) are eigenvectors for
A given by
4 −2
A= (13-11)
3 −1
For this we write the eigenvalue problem, as stated in Definition 13.19.
4 −2 2 2
Av1 = = = 1 · v1
3 −1 3 3
4 −2 4 8
Av2 = = = 2 · v2 (13-12)
3 −1 4 8
4 −2 2 10
Av3 = = 6= λ · v3 .
3 −1 −1 7
From this we see that v1 and v2 are eigenvectors for A. v1 corresponding to the eigenvalue 1,
and v2 corresponding to the eigenvalue 2.
For the use in the following investigations we make some important comments to Defi-
nition 13.19 .
First we note that even if the square matrix A in Definition 13.19 is real, one is often
interested not only in the real solutions to (13-10), but more generally complex solutions.
In other words we seek a scalar λ ∈ C and a vector v ∈ Cn , satisfying (13-10).
If one is only interested in real solutions to the eigenvalue problem for real square ma-
trices, one can alternatively see the left hand side of (13-10) as a real map f : Rn → Rn
given by:
f (v) = A v .
Of course, this map is linear, too. We get the following version of Theorem 13.23:
In the light of Theorem 13.23 and Theorem 13.24 we now introduce the concept eigen-
vector space, compare with Definition 13.4.
The subspace of all the eigenvectors that correspond to λ is termed the eigenvector
space (or in short the eigenspace) corresponding to λ and is termed Eλ .
Now we have sketched the structural framework for the eigenvalue problem for square
matrices, and we continue in the following two subsections by investigating in an ele-
mentary way, how one can begin to find eigenvalues and eigenvectors for square matri-
ces.
First we put λv onto the left hand of the equality sign, and then v “is placed outside a
pair of brackets”. This is possible because v = E v where E is the identity matrix:
KA (λ) = (A − λE)
Since it is a homogeneous system of linear equations that we have to solve we have two
possibilities for the structure of the solution. Either the characteristic matrix is invertible,
and the the only solution is v = 0 . Or the matrix is singular, and then infinitely many
solutions v exist. But since Definition 13.19 requires that v must be a proper vector, that
is a vector different from the zero vector, the characteristic matrix must be singular. To
investigate whether this is true, we take the determinant of the square matrix. This is
zero exactly when the matrix is singular:
Note that the left hand side in (13-15) is a polynomial in the variable λ. The polynomial
is given a special symbol:
The equation that results when the characteristic polynomial is set equal to zero
By the use of the method for calculating the determinant we see that the characteristic
polynomial is always an n’th degree polynomial. See also the following examples. The
main point is that the roots in the characteristic polynomial (solutions to the character
equation) are the eigenvalues of the matrix, because the eigenvalues precisely satisfy
that the characteristic matrix is singular.
eNote 13 13.2 THE EIGENVALUE PROBLEM FOR SQUARE MATRICES 339
Then:
2. The roots of the characteristic polynomial (the solutions to the characteristic equa-
tion) are all the eigenvalues of A .
(A − λE)v = 0 , (13-22)
that was achieved in (13-14). Since the eigenvalues are now known, the homogeneous
system of linear equations corresponding to (13-22) can be solved with respect to the n
unknowns v1 , ..., vn that are the elements in v = (v1 , ..., vn ) . We just have to substitute
the eigenvalues one after one. As mentioned above, the characteristic matrix is singular
when the substituted λ is an eigenvalue. Therefore infinitely many solutions to the
system of equations exist. Finding these corresponds to finding all eigenvectors v that
correspond to λ .
(A − λE)v = 0 , (13-24)
Method 13.28 is unfolded in the following three examples that also show a way of char-
acterizing the set of eigenvectors corresponding to a given eigenvalue, in the light of
Theorem 13.23 and Theorem 13.24.
From this we see that v2 = (1, 1) is an eigenvector corresponding to the eigenvalue λ2 . All
real eigenvectors corresponding to λ2 can be written as
1
v = t· , t ∈ R. (13-34)
1
This is a one-dimensional subspace in R2 that can also be written as:
E3 = span{(1, 1)} . (13-35)
We will now check our understanding: When v1 = (−1, 1) is mapped by A, will the image
vector only be a scaling (change of length) of v1 ?
2 1 −1 −1
Av1 = = = v1 . (13-36)
1 2 1 1
It is true! It is also obvious that the eigenvalue is 1 .
Now we check v2 :
2 1 1 3
Av2 = = = 3 · v2 . (13-37)
1 2 1 3
v2 is also as expected an eigenvector and the eigenvalue is 3.
eNote 13 13.2 THE EIGENVALUE PROBLEM FOR SQUARE MATRICES 343
that has no real eigenvalues. But we found two complex eigenvalues, λ1 = 1 + 2i and λ2 =
1 − 2i .
Exercise 13.32
A = 0 −1 6 . (13-53)
0 1 4
2. λ is said to have the geometric multiplicity m when the dimension of the eigen-
vector space corresponding to λ is m. This is termed gm(λ) = m. In other words:
dim( Eλ ) = gm(λ) .
We do not always have am(λ) = gm(λ) . This is dealt with in Theorem 13.34.
The following theorem has some important properties concerning algebraic and geo-
metric multiplicity of eigenvalues of square matrices, cf. Theorem 13.11.
1. A has at the most n different real eigenvalues, and also the sum of algebraic
multiplicities of the real eigenvalues is at the most n .
2. A has at the most n different complex eigenvalues, but the sum of the algebraic
multiplicities of the complex eigenvalues is equal to n .
That is, the geometric multiplicity of an eigenvalue will at the least be equal to
1, it will be less than or equal to the algebraic multiplicity of the eigenvalue,
which in turn will be less than or equal to the number of rows and columns in
A.
eNote 13 13.2 THE EIGENVALUE PROBLEM FOR SQUARE MATRICES 347
Exercise 13.35
Check that all three points in Theorem 13.34 are valid for the eigenvalues and eigenvectors
in example 13.31.
Points 1 and 2 follow directly from the theory of polynomials. The characteristic poly-
nomial for a real n × n-matrix A is an n’th degree polynomial, and it has at the most n
different roots, counting both real and complex ones. Furthermore the sum of the mul-
tiplicities of the real roots is at the most n , whereas the sum of the multiplicities to the
complex roots is equal to n .
We have previously shown that for every linear map of an n-dimensional vector space
into itself the sum of the geometric multiplicities of the eigenvalues for f can at the
most be n , see Theorem 13.11. Note that this can be deduced directly from the state-
ments about multiplicities in Theorem 13.34.
As something new and interesting it is postulated in point 3 that the geometric multi-
plicity of a single eigenvalue can be less than the algebraic multiplicity. This is demon-
strated in the following summarizing Example 13.36. Furthermore the geometric mul-
tiplicity of a single eigenvalue cannot be greater than the algebraic one. The proof of
point 3 in Theorem 13.34 is left out.
A = −3 1 5 (13-55)
1 −4 6
The eigenvalues of A are determined:
−9 − λ
10 0
det −3 1−λ 5 = −λ3 − 2λ2 + 7λ − 4 = −(λ + 4)(λ − 1)2 = 0 . (13-56)
1 −4 6 − λ
From the factorization in front of the last equality sign we get that A has two different eigen-
values: λ1 = −4 and λ2 = 1. Moreover am(−4) = 1 and am(1) = 2, as can be seen from the
factorization.
eNote 13 13.2 THE EIGENVALUE PROBLEM FOR SQUARE MATRICES 348
−9 − (−4)
10 0 0
−3 1 − (−4) 5 0 →
1 −4 6 − (−4) 0
(13-57)
1 −2 0 1 −2
0 0 0
0 −1 5 0 → 0
1 −5 0
0 −2 10 0 0 0 0 0
There are two non-trivial equations: v1 − 2v2 = 0 and v2 − 5v3 = 0. If we put v3 equal to the
free parameter we see that all real eigenvectors corresponding to λ1 can be stated as
We have that gm(−4) = dim( E−4 ) = 1, and that an eigenvector to λ1 is v1 = (10, 5, 1). It is
seen that gm(−4) = am(−4).
Similarly for λ2 = 1:
−9 − 1
10 0 0
−3 1−1 5 0 →
1 −4 6 − 1 0
(13-59)
1 −1 −1
0 0 1 0 0
0 −3 5 0 → 0 3 −5 0
0 −3 5 0 0 0 0 0
Again we have two non-trivial equations: v1 − v2 = 0 and 3v2 − 5v3 = 0. If we put v3 = 3s
we see that all to λ2 corresponding real eigenvectors can be stated as
This gives the following results: gm(1) = dim( E1 ) = 1 and that an eigenvector to λ2 = λ3 is
v2 = (5, 5, 3). Furthermore it is seen that gm(1) < am(1).
−1 4
B= . (13-61)
−2 3
eNote 13 13.2 THE EIGENVALUE PROBLEM FOR SQUARE MATRICES 349
From Example 13.30 in order to make more precise some special phenomena for square,
real matrices when their eigenvalue problems are studied in a complex framework.
Proof
The first part of Theorem 13.37 follows from the theory of polynomials. The characteristic
polynomial of a square, real matrix is a polynomial with real coefficients. The roots of such a
polynomial come in conjugate pairs.
By the trace of a square matrix we understand the sum of the diagonal elements. The
trace of B is thus −1 + 3 = 2 . Now notice that the sum of the eigenvalues of B is
(1 − i ) + (1 + i ) = 2 , that is equal to the trace of B . This is also a general phenomenon,
which we state without proof:
eNote 13 13.2 THE EIGENVALUE PROBLEM FOR SQUARE MATRICES 350
Exercise 13.39
In Example 13.31 we found that the characteristic polynomial for the matrix
6 3 12
A = 4 −5 4
−4 −1 −10
has the double root −6 and the single root 3 . Prove that Theorem 13.38 is valid in this case.
eNote 14 351
eNote 14
In this eNote it is explained how certain square matrices can be diagonalized by the use of
eigenvectors. Therefore it is presumed that you know how to determine eigenvalues and
eigenvectors for a square matrix and furthermore that you know about algebraic and geometric
multiplicity.
b Fb = (a Mb )−1 · a Fa · a Mb (14-1)
where a Mb = a b1 a b2 · · · a bn is the change of basis matrix that shifts from b to a
coordinates.
It is of special interest if a basis v consisting of eigenvectors for f can be found. Viz. let
a be an arbitrary basis for V and a Fa the corresponding mapping matrix for f . Further-
more let v be an eigenvector basis for V with respect to f . From Theorem 13.14 in eNote
13 it appears that the mapping matrix for f with respect to the v-basis is a diagonal ma-
trix Λ in which the diagonal elements are the eigenvalues of f . If V denotes the change
of basis matrix that shifts from v-coordinates to the a-coordinate vectors, according to
(14-1)Λ will appear as
Λ = V−1 · a Fa · V . (14-2)
eNote 14 14.1 SIMILAR MATRICES 352
Naturally formula 14-1 and formula 14-2 inspire questions that take their starting point
in square matrices: Which conditions should be satisfied in order for two given square
matrices to be interpreted as mapping matrices for the same linear map with respect to
two different bases? And which conditions should a square matrix satisfy in order to
be a mapping matrix for a linear map that in another basis has a diagonal matrix as a
mapping matrix? First we study these questions in a pure matrix algebra context and
return in the last subsection to the mapping viewpoint. For this purpose we now intro-
duce the concept similar matrices.
B = M−1 A M . (14-3)
2 3 8 21
Given the matrices A = and B = .
3 −4 −3 −10
2 3 − 1 2 −3
The matrix M = is invertible and has the inverse matrix M = .
1 2 −1 2
Exercise 14.4
Proof
Let M be an invertible matrix that satisfies B = M−1 AM and let, as usual, E denote the
identity matrix of the same size as the three given matrices. Then:
Thus it is shown that the two matrices have the same characteristic polynomial and thus the
same eigenvalues with the same corresponding algebraic multiplicities. Moreover, that they
have the same eigenvalues appears from Theorem 14.13 which is given below: When A and
eNote 14 14.2 MATRIX DIAGONALIZATION 354
B can represent the same linear map f with respect to different bases they have identical
eigenvalues, viz. the eigenvalues of f .
But the eigenvalues also do have the same geometric multiplicities. This follows from the fact
that the eigenspaces for A and B with respect to any of the eigenvalues can be interpreted as
two different coordinate representations of the same eigenspace, viz. the eigenspace for f
with respect to the said eigenvalue.
Note that Theorem 14.5 says that two similar matrices have the same eigenval-
ues, but not vice versa: that two matrices, which have the same eigenvalues,
are similar. There is a difference and only the first statement is true.
Two similar matrices A and B have the same eigenvalues, but an eigenvector
for the one is not generally and eigenvector for the other. But if v is an eigen-
vector for A corresponding to the eigenvalue λ then M−1 v is an eigenvector
for B corresponding to the eigenvalue λ, where M is the invertible matrix that
satisfies B = M−1 AM . Viz.:
Exercise 14.6
Explain that two square n × n-matrices are similar, if they have identical eigenvalues with the
same corresponding geometric multiplicities and that the sum of the geometric multiplicities
is n .
Now we will ask the question whether or not an arbitrary square matrix A can be
diagonalized by a similarity transformation. Therefore we form the equation
V−1 A V = Λ,
where V is an invertible matrix and Λ is a diagonal matrix. Below we prove that
the equation has exactly one solution if the columns of V are linearly independent
eigenvectors for A , and the diagonal elements in Λ are the eigenvalues of A written
such that the i-th column of V is an eigenvector corresponding to the eigenvalue for
the i-th column in Λ .
where
Λ = diag(λ1 , λ2 , . . . , λn ) .
V = v1 v2 · · · vn and (14-10)
Proof
0 λ2 · · · 0
⇔ A v1 v2 · · · vn = v1 v2 · · · vn . .. . . . (14-12)
.. . . ..
0 0 · · · λn
⇔ AV = VΛ
Now all the eigenvectors are inserted (vertically one after the other) in the matrix V in the
same order as that of the eigenvalues in the diagonal of the matrix Λ that outside the diag-
onal contains only zeroes. Since the eigenvectors are linearly independent the matrix V is
invertible. Therefore the inverse V−1 exists, and we multiply by this from the left on both
sides of the equality sign:
Thus the first part of the theorem is proved. Suppose on the contrary that A can be diagonal-
ized by a similarity transformation. Then an invertible matrix V = v1 v2 · · · vn and a
diagonal matrix Λ = diag(λ1 , λ2 , . . . λn ) exist such that
V−1 AV = Λ . (14-14)
If we now repeat the transformations in the first part of the proof only now in the opposite
order, it is seen that (14-14) is the equivalent of the following n equations:
eNote 14 14.2 MATRIX DIAGONALIZATION 357
The following theorem can be of great help when one investigates whether matrices can
be diagonalized by similarity in different contexts . The main result is already given in
Theorem 14.7, but here we refine the conditions by drawing upon previously proven
theorems about the eigenvalue problem for linear maps and matrices.
Proof
Ad. 1. If a proper eigenvector from each of the n eigenspaces is chosen, it follows from Corol-
lary 13.9 that the collected set of n eigenvectors is linearly independent. Therefore, according
to Theorem 14.7, A can be diagonalized by similarity transformation.
Ad. 2: If a basis from each of the eigenspaces is chosen, then the collected set of the chosen
n eigenvectors according to Corollary 13.9 is linearly independent. Therefore, according to
Theorem 14.7 A can be diagonalized by similarity transformation .
Ad. 3: If the sum of the geometric multiplicities is less than n, n linearly independent eigen-
vectors for A do not exist. Therefore, according to Theorem 14.7 A cannot be diagonalized
by similarity transformation.
Ad. 4: Since according to Theorem 13.34 point 1, the sum of the algebraic multiplicity is less
than or equal to n, and since according to the same theorem point 2 for every eigenvalue
λ gm(λ) ≤ am(λ) , the sum of the geometric multiplicities cannot become n, if one of the
geometric multiplicities is less than its algebraic one. Therefore, according to what has just
been proved, A cannot be diagonalized by similarity transformation.
eNote 14 14.3 COMPLEX DIAGONALIZATION 358
In the following examples we will see how to investigate in practice whether diagonal-
ization by similarity transformation is possible and, if so, carry it through.
Example 14.9
−2 −3 1
λ1 0 0 4 0 0
Λ = 0 λ1 0 = 0 4 0 and V = v1 v2 v3 = 0
1 2 . (14-17)
0 0 λ2 0 0 15 1 0 2
What we so far have said about similar matrices is generally valid for square, complex
matrices. Therefore the basic equation for diagonalization by similarity transformation:
V−1 A V = Λ,
will be understood in the broadest sense, where the matrices A, V and Λ are complex
n × n-matrices. Until now we have limited ourselves to real examples, that is examples
where it has been possible to satisfy the basic equation (14.3) with real matrices. We will
in the following look upon a special situation that is typical in technical applications of
diagonalization: For a given real n × n matrix A we seek an invertible matrix M and
a diagonal matrix Λ satisfying the basic equation in a broad context where M and Λ
possibly are complex (not real) n × n matrices.
eNote 14 14.3 COMPLEX DIAGONALIZATION 359
The following example shows a real 3 × 3 matrix that cannot be diagonalized (with only
non-complex entries in the diagonal) because its characteristic polynomial only has one
real root. On the other hand it can be diagonalized in a complex sense.
1 −2 + i −2 − i
λ1 0 0 2 0 0
Λ = 0 λ2 0 = 0 −i 0 and V = v1 v2 v3 = 0
i −i .
0 0 λ3 0 0 i 0 1 1
(14-19)
The next example shows a real, square matrix that cannot be diagonalized either in a
real or in a complex way.
A = 1 4 1 , (14-20)
0 0 3
and A has the eigenvalues λ1 = 3 and λ3 = 5. The eigenvalue 3 has the algebraic multiplicity
2, but only one linearly independent eigenvector can be chosen, e.g. v1 = (1, −1, 0). Thus the
eigenvalue has the geometric multiplicity 1. Therefore, according to Theorem 14.7, it is not
possible to diagonalize A by similarity transformation.
eNote 14 14.4 DIAGONALIZATION OF LINEAR MAPS 360
Exercise 14.12
2. All corresponding linearly independent eigenvectors and thus the geometric multiplic-
ities of the eigenvectors.
In the introduction to this eNote we asked the question: What conditions should be
satisfied so that two given square matrices can be interpreted as mapping matrices for
the same linear map with respect to two different bases? The answer is simple:
Exercise 14.14
In the introduction we also asked the question: Which conditions should a square ma-
trix satisfy in order to be a mapping matrix for a linear map that in another basis has
eNote 14 14.4 DIAGONALIZATION OF LINEAR MAPS 361
a diagonal matrix as a mapping matrix? The answer appears from Theorem 14.7 com-
bined with Theorem 14.13: the matrix must have n linearly independent eigenvectors.
A linear map f : P1 (R) → P1 (R) is given by the following mapping matrix with respect to
the standard monomial basis m:
−17 −21
F
m m = (14-22)
14 18
This means that f (1) = −17 + 14x and f ( x ) = −21 + 18x. We wish to investigate whether a
(real) eigenbasis for f can be found and if so, how the mapping matrix looks with respect to
this basis, and what the basis vectors are.
It is already now possible to confirm that a real eigenbasis for f exists since 2 = dim( P2 (R)),
viz. λ1 = −3 and λ2 = 4 each with the algebraic multiplicity 1. Eigenvectors corresponding
to λ1 are determined:
1 23 0
−17 + 3 −21 0
→ . (14-24)
14 18 + 3 0 0 0 0
This yields an eigenvector m v1 = (−3, 2), if the free parameter is put equal to 2. Similarly we
get the other eigenvector:
−17 − 4 −21 0 1 1 0
→ . (14-25)
14 18 − 4 0 0 0 0
This yields an eigenvector m v2 = (−1, 1), if the free parameter is put equal to 1.
Thus a real eigenbasis v for f , given by the basis vectors m v1 and m v2 , exists. We then get
−3 −1 −3 0
m Mv = and v Fv = (14-26)
2 1 0 4
The basis consists of the vectors v1 = −3 + 2x and v2 = −1 + x and the map is “simple” with
respect to this basis.
eNote 14 14.4 DIAGONALIZATION OF LINEAR MAPS 362
It is true!
eNote 15 363
eNote 15
Symmetric Matrices
In this eNote we will consider one of the most used results from linear algebra – the so-called
spectral theorem for symmetric matrices. In short it says that all symmetric matrices can be
diagonalized by a similarity transformation – that is, by change of basis with a suitable
substitution matrix.
The introduction of these concepts and the corresponding method were given in eNotes 10, 13
and 14, which therefore is a necessary basis for the present eNote.
Precisely in that eNote it became clear that not all matrices can be diagonalized.
Diagonalization requires a sufficiently large number of eigenvalues (the algebraic multiplicities
add up to be as large as possible) and that the corresponding eigenvector spaces actually span all
of the vector space (the geometric multiplicities add up to be as large as possible). It is these
properties we will consider here, but now for symmetric matrices, which turn out to satisfy the
conditions and actually more: the eigenvectors we use in the resulting substitution matrix can
be chosen pairwise orthogonal, such that the new basis is the result of a rotation of the old
standard basis in Rn .
In order to be able to discuss and apply the spectral theorem most effectively we must first
introduce a natural scalar product for vectors in Rn in such a way that we will be able to
measure angles and lengths in all dimensions. We do this by generalizing the well-known
standard scalar product from R2 and R3 . As indicated above we will in particular use bases
consisting of pairwise orthogonal vectors in order to formulate the spectral theorem, understand
it and what use we can make of this important theorem.
In the vector space Rn we introduce an inner product, i.e. a scalar product that is a
natural generalization of the well-known scalar product from plane geometry and space
geometry, see eNote 10.
Then we define the scalar product, the inner product, (also called the dot product) of
the two vectors in the following way:
n
a·b = a1 b1 + a2 b2 + · · · an bn = ∑ a i bi . (15-2)
i =1
b1
·
>
a · b = ea · eb = a1 · · · an ·
·
(15-3)
·
bn
For the scalar product introduced above the following arithmetic rules apply:
eNote 15 15.1 SCALAR PRODUCT 365
A main point about the introduction of a scalar product is that we can now talk about
the lengths of the vectors in (Rn , ·):
The length of a is also called the norm of a with respect to the scalar product in
(Rn , ·). A vector a is called a proper vector if |a| > 0 .
eNote 15 15.1 SCALAR PRODUCT 366
Finally it follows from Definition 15.1 and Definition 15.3 that for a ∈ (Rn , ·)
and an arbitrary real number k we have that
|a · b| ≤ | a | | b | . (15-11)
Proof
If b = 0 , both sides of (15-11) are equal to 0 and the inequality is thereby satisfied. We now
assume that b is a proper vector.
1
We put k = b · b and e = √ b . It then follows from (15-6) that
k
1 1 1
e · e = ( √ b) · ( √ b) = (b · b) = 1
k k k
|a · e| ≤ | a | (15-12)
Since it follows from (15-8) that (a − te) · (a − te) = 0 if and only if (a − te) = 0 , we see that
|a · e| = | a | if and only if a and e are linearly dependent. The proof is hereby complete.
From the Cauchy-Schwarz inequality follows the triangle inequality that is a general-
ization of the well-known theorem from elementary plane geometry, that a side in a
triangle is always less than or equal to the sum of the other sides:
Exercise 15.6
a·b
−1 ≤ ≤1 . (15-14)
|a| · |b|
Therefore the angle between two vectors in (Rn , ·) can be introduced as follows:
a·b
cos(θ ) = . (15-15)
|a| · |b|
If a · b = 0 we say that the two proper vectors are orthogonal or perpendicular with
respect to each other. This occurs exactly when cos(θ ) = 0, that is, when θ = π/2.
Definition 15.8
A square matrix A is symmetric if it is equal to its own transpose
A = A> , (15-16)
What is the relation between symmetric matrices and the scalar product? This we con-
sider here:
eNote 15 15.2 SYMMETRIC MATRICES AND THE SCALAR PRODUCT 369
Theorem 15.9
Let v and w denote two vectors in the vector space (Rn , ·) with scalar product intro-
duced above. If A is an arbitrary (n × n)−matrix then
(A v) ·w = v· A> w . (15-17)
Proof
We use the fact that the scalar product can be expressed as a matrix product:
(A v) · w = (A v) > · w
= v> A> · w
(15-18)
= v> · A> w
= v · A> w .
Theorem 15.10
A matrix A is a symmetric (n × n)−matrix if and only if
(A v) ·w = v· (A w) (15-19)
Proof
If A is symmetric then we have that A = A> and therefore Equation (15-19) follows directly
from Equation (15-17). Conversely, if we assume that (15-19) applies for all v and w, we
eNote 15 15.2 SYMMETRIC MATRICES AND THE SCALAR PRODUCT 370
will prove that A is symmetric. But this follows easily just by choosing suitable vectors, e.g.
v = e2 = (0, 1, 0, ..., 0) and w = e3 = (0, 0, 1, ..., 0) and substitute these into (15-19) as seen
below. Note that A ei is the ith column vector in A.
(A e2 ) · e3 = a23
= e2 · (A e3 )
(15-20)
= (A e3 ) ·e2
= a32 ,
such that a23 = a32 . Quite similarly for all other choices of indices i and j we get that aij = a j i
– and this is what we had set out to prove.
A basis a in (Rn , ·) consists (as is known from eNote 11) of n linearly independent vec-
tors (a1 , ..., an ). If in addition the vectors are pairwise orthogonal and have length 1 with
respect to the scalar product, then (a1 , ..., an ) is an orthonormal basis for (Rn , ·) :
Definition 15.11
A basis a = (a1 , ..., an ) is an orthonormal basis if
1 for i = j ,
ai · a j = (15-21)
0 for i 6= j .
Exercise 15.12
Show that if n vectors (a1 , ..., an ) in (Rn , ·) satisfy Equation (15-21) then a = (a1 , ..., an ) is
automatically a basis for (Rn , ·), i.e. the vectors are linearly independent and span all of
(Rn , ·).
eNote 15 15.2 SYMMETRIC MATRICES AND THE SCALAR PRODUCT 371
Exercise 15.13
Show that the following 3 vectors (a1 , a2 , a3 ) constitute an orthonormal basis for (R3 , ·) for
any given value of θ ∈ R :
a1 = (cos(θ ), 0, − sin(θ ))
a2 = (0, 1, 0) (15-22)
a3 = (sin(θ ), 0, cos(θ )) .
If we put the vectors from an orthonormal basis into a matrix as columns we get an
orthogonal matrix:
Definition 15.14
An (n × n)−matrix A is said to be orthogonal if the column vectors in A constitute an
orthonormal basis for (Rn , ·), that is if the column vectors are pairwise orthogonal
and all have length 1 – as is also expressed in Equation (15-21).
Note that orthogonal matrices alternatively (and maybe also more descriptively)
could be called orthonormal, since the columns in the matrix are not only pair-
wise orthogonal but also normalized such that they all have length 1. We will
follow international tradition and call the matrices orthogonal.
Theorem 15.15
An (n × n)−matrix Q is orthogonal if and only if
which is equivalent to
Q> = Q−1 . (15-24)
eNote 15 15.2 SYMMETRIC MATRICES AND THE SCALAR PRODUCT 372
Proof
See eNote 7 about the computation of the matrix product and then compare with the condi-
tion for orthogonality of the column vectors in Q (Equation (15-21)).
Theorem 15.16
An n × n matrix A is orthogonal if and only if the linear mapping f : Rn → Rn given
by f (x) = Ax preserves the scalar product, i.e.:
Exercise 15.17
| det(A)| = 1 . (15-25)
Show that this condition is not sufficient, thus matrices exist that satisfy this determinant-
condition but that are not orthogonal.
Definition 15.18
An orthogonal matrix Q is called special orthogonal or positive orthogonal if
det(Q) = 1 and it is called negative orthogonal if det(Q) = −1.
In the literature, orthogonal matrices with determinant 1 are called special orthogonal,
eNote 15 15.3 GRAM–SCHMIDT ORTHONORMALIZATION 373
Exercise 15.19
2. The next vector v2 in the basis v is now chosen in span{u1 , u2 } but such that
at the same time v2 is orthogonal to v1 , i.e. v2 ·v1 = 0; finally this vector is
normalized. First we construct an auxiliary vector w2 .
w2 = u2 − (u2 · v1 ) v1
w2 (15-28)
v2 = .
|w2 |
Note that w2 (and therefore also v2 ) then being orthogonal to v1 :
w2 · v1 = (u2 − (u2 · v1 ) v1 ) · v1
= u2 · v1 − (u2 · v1 ) v1 · v1
= u2 · v1 − (u2 · v1 ) |v1 |2 (15-29)
= u2 · v1 − (u2 · v1 )
=0 .
The constructed v-vectors span the same subspace U as the given linearly indepen-
dent u-vectors, U = span{u1 , · · ·, u p } = span{v1 , · · ·, v p } and v = (v1 , · · ·, v p )
constituting an orthonormal basis for U.
eNote 15 15.3 GRAM–SCHMIDT ORTHONORMALIZATION 375
Example 15.21
In (R4 , ·) we will by the use of the Gram–Schmidt orthonormalization method find an or-
thonormal basis v = (v1 , v2 , v3 ) for the 3−dimensional subspace U that is spanned by the
three given linearly independent (!) vectors having the following coordinates with respect to
the standard e-basis in R4 :
We construct the new basis vectors with respect to the standard e-basis in R4 by working
through the orthonormalization procedure. There are 3 ’steps’ since there are in this example
3 linearly independent vectors in U :
1.
u1 1
v1 = = (2, 2, 4, 1) . (15-32)
|u1 | 5
2.
w2 = u2 − (u2 · v1 ) v1 = u2 + 5v1 = (2, 2, −1, −4)
w2 1 (15-33)
v2 = = (2, 2, −1, −4) .
| w2 | 5
3.
w3 = u3 − (u3 · v1 ) v1 − (u3 · v2 ) v2 = u3 − 5v1 − 5v2 = (1, −1, 0, 0)
w3 1 (15-34)
v3 = = √ (1, −1, 0, 0) .
| w3 | 2
Thus we have constructed an orthonormal basis for the subspace U consisting of those vec-
tors that with respect to the standard basis have the coordinates:
1 1 1
v1 = · (2, 2, 4, 1) , v2 = · (2, 2, −1, −4) , v3 = √ · (1, −1, 0, 0) .
5 5 2
We can check that this is really an orthonormal basis by posing the vectors as columns in a
matrix, which then is of the type (4 × 3). Like this:
√
2/5 2/5 1/√2
2/5 2/5 −1/ 2
V= (15-35)
4/5 −1/5 0
1/5 −4/5 0
The matrix V cannot be an orthogonal matrix (because of the type), but nevertheless V can
satisfy the following equation, which shows that the three new basis vectors indeed are
eNote 15 15.4 THE ORTHOGONAL COMPLEMENT TO A SUBSPACE 376
Exercise 15.22
In (R4 , ·) the following vectors are given with respect to the standard basis e:
We let U denote the subspace in (R4 , ·) that is spanned by the four given vectors, that is
U = span{u1 , u2 , u3 , u4 } . (15-37)
1. Show that u = (u1 , u2 , u3 ) is a basis for U and find coordinates for u4 with respect to
this basis.
Example 15.23
In (R3 , ·) a given first unit vector v1 is required for the new orthonormal basis v= (v1 , v2 , v3 )
and the task is to find the two other vectors in the basis. Let us assume that the given vector is
v1 = (3, 0, 4)/5. We see immediately that e.g. v2 = (0, 1, 0) is a unit vector that is orthogonal
to v1 . A last vector for the orthonormal basis can then be found directly using the cross
product: v3 = v1 × v2 = 51 · (−4, 0, 3).
Definition 15.24
The orthogonal complement to a subspace U of (Rn , ·) is denoted U ⊥ and consists
of all vectors in (Rn , ·) that are orthogonal to all vectors in U:
Theorem 15.25
The orthogonal complement U ⊥ to a given p−dimensional subspace U of (Rn , ·) is
itself a subspace in (Rn , ·) and it has dimension dim(U ⊥ ) = n − p .
Proof
Example 15.26
Exercise 15.27
Determine the orthogonal complement to the subspace U = span{u1 , u2 , u3 } in (R4 , ·), when
the spanning vectors are given by their respective coordinates with respect to the standard
basis e in R4 as:
We will now start to formulate the spectral theorem and start with the following non-
trivial observation about symmetric matrices:
Theorem 15.28
Let A denote a symmetric (n × n)−matrix. Then the characteristic polynomial
KA (λ) for A has exactly n real roots (counted with multiplicity):
λ1 ≥ λ2 ≥ · · · ≥ λ n . (15-40)
If e.g. {7, 3, 3, 2, 2, 2, 1} are the roots of KA (λ) for a (7 × 7)−matrix A, then these
roots must be represented with their respective multiplicity in the eigenvalue-list:
λ1 = 7 ≥ λ2 = 3 ≥ λ3 = 3 ≥ λ4 = 2 ≥ λ5 = 2 ≥ λ6 = 2 ≥ λ7 = 1 .
Since Theorem 15.28 expresses a decisive property about symmetric matrices, we will
here give a proof of the theorem:
eNote 15 15.5 THE SPECTRAL THEOREM FOR SYMMETRIC MATRICES 379
Proof
From the fundamental theorem of algebra we know that KA (λ) has exactly n complex roots
- but we do not know whether the roots are real; this is what we will prove. So we let α + i β
be a complex root of KA (λ) and we will then show that β = 0. Note that α and β naturally
both are real numbers.
Therefore we have
det (A − (α + i β)E) = 0 , (15-41)
and thus also that
det (A − (α + i β)E) · det (A − (α − i β)E) = 0 (15-42)
such that
det ((A − (α + i β)E) · (A − (α − i β)E)) = 0
(15-43)
det (A − α E)2 + β2 E = 0 .
The last equation yields that the rank of the real matrix (A − α E)2 + β2 E is less than n;
this now means (see eNote 6) that proper real solutions x to the corresponding system of
equations must exist.
(A − α E)2 + β2 E x = 0 . (15-44)
Let us choose such a proper real solution v to (15-44) with |v| > 0. Using the assumption that
A (and therefore A − αE also) is assumed to be symmetric, we have:
2 2
0= ( A − α E ) + β E v ·v
= (A − α E)2 v · v + β2 (v · v)
(15-45)
2 2
= ((A − α E) v) · ((A − α E) v) + β |v|
= | (A − α E) v|2 + β2 |v|2 .
Since |v| > 0 we are bound to conclude that β = 0, because all terms in the last expression
are non-negative. And this is what we had to prove.
Exercise 15.29
Where was it exactly that we actually used the symmetry of A in the above proof?
eNote 15 15.5 THE SPECTRAL THEOREM FOR SYMMETRIC MATRICES 380
If two eigenvalues λi and λ j for a symmetric matrix are different, then the two corre-
sponding eigenvector spaces are orthogonal, Eλi ⊥ Eλ j in the following sense:
Theorem 15.30
Let A be a symmetric matrix and let λ1 and λ2 be two different eigenvalues for A
and let v1 and v2 denote two corresponding eigenvectors. Then v1 · v2 = 0, i.e. they
are orthogonal.
Proof
and since λ1 6= λ2 we therefore get the following conclusion: v1 ·v2 = 0, and this is what we
had to prove.
We can now formulate one of the most widely applied results for symmetric matrices,
the spectral theorem for symmetric matrices that, with good reason, is also called the
theorem about diagonalization of symmetric matrices:
eNote 15 15.5 THE SPECTRAL THEOREM FOR SYMMETRIC MATRICES 381
Theorem 15.31
Let A denote a symmetric (n × n)−matrix. Then a special orthogonal matrix Q exists
such that
Λ = Q−1 AQ = Q> AQ is a diagonal matrix . (15-47)
I.e. that a real symmetric matrix can be diagonalized by application of a positive
orthogonal substitution, see eNote 14.
The diagonal matrix can be constructed very simply from the n real eigenvalues
λ1 ≥ λ2 ≥ · · · ≥ λn of A as:
λ1 0 · 0
0 λ2 · 0
Λ = diag(λ1 , λ2 , ..., λn ) =
·
, (15-48)
· · ·
0 0 · λn
Q = [v1 v2 · · · vn ] , (15-49)
3. The resulting matrix Q has determinant 1 (if not then multiply one of the cho-
sen eigenvectors by −1 to flip the sign of the determinant)
That this is so follows from the results and remarks – we go through a series of
enlightening examples below.
eNote 15 15.6 EXAMPLES OF DIAGONALIZATION 382
Here are some typical examples that show how one diagonalizes some small symmetric
matrices, i.e. symmetric matrices of type (2 × 2) or type (3 × 3):
2 −2
1
A = −2 5 −2 . (15-50)
1 −2 2
We will determine a special orthogonal matrix Q such that Q−1 AQ is a diagonal matrix:
2 − λ −2
1
KA (λ) = det −2 5 − λ −2 = (λ − 1)2 · (7 − λ) , (15-52)
1 −2 2 − λ
so A has the eigenvalues λ1 = 7, λ2 = 1, and λ3 = 1. Because of this we already know
through Theorem 15.31 that it is possible to construct a positive orthogonal matrix Q such
that
7 0 0
Q−1 AQ = diag(7, 1, 1) = 0 1 0 . (15-53)
0 0 1
The rest of the problem now consists in finding the eigenvectors for A that can be used as
columns in the orthogonal matrix Q.
Eigenvectors for A corresponding to the eigenvalue λ1 = 7 are found by solving the homo-
geneous system of equations that has the coefficient matrix
−5 −2
1
KA (7) = A − 7E = −2 −2 −2 , (15-54)
1 −2 −5
which by suitable row operations is seen to have
1 0 −1
The eigenvector solutions to the corresponding homogeneous system of equations are seen
to be
u = t · (1, −2, 1) , t ∈ R , (15-56)
√
such that E7 = span{(1, −2, 1)}. The normalized eigenvector v1 = (1/ 6) · (1, −2, 1) is
therefore an orthonormal basis for E7 (and it can also be used as the first column vector in the
wanted Q: √
1/√6 ∗ ∗
Q = −2/√6 ∗ ∗ . (15-57)
1/ 6 ∗ ∗
We know from Theorem 15.31 that the two last columns are found by similarly determining
all eigenvectors E1 belonging to the eigenvalue λ2 = λ3 = 1 and then choosing two
orthonormal eigenvectors from E1 .
Finally we investigate whether the chosen eigenvectors give a positive orthogonal matrix.
Since
1 −1 1
Q−1 AQ = Q> AQ
√ √ √ √ √ √
1/√6 −2/ 6 1/√6 2 −2 1 1/√6 −1/ 2 −1/√3
= −1/√2 √0 1/√2 · −2 5 −2 · −2/√6 √0 −1/√3
−1/ 3 −1/ 3 −1/ 3 1 −2 3 1/ 6 1/ 2 −1/ 3
7 0 0
= 0 1 0 ,
0 0 1
(15-66)
which we wanted to show.
We should finally remark here that, since we are in three dimensions, instead of using Gram–
Schmidt orthonormalization for the determination of v3 we could have used the cross prod-
uct v1 × v2 (see 15.23):
√
v3 = v1 × v2 = (1/ 3) · (−1, −1, −1) . (15-67)
We will determine a special orthogonal matrix Q such that Q−1 AQ is a diagonal matrix:
The eigenvectors for A corresponding to the eigenvalue λ1 = 20 are found by solving the
homogeneous system of equations having the coefficient matrix
−9 −12
KA (20) = A − 20E = , (15-72)
−12 −16
which, through suitable row operations, is shown to have the equivalent reduced matrix:
3 4
rref(KA (20)) = . (15-73)
0 0
The eigenvector solutions to the corresponding homogeneous system of equations are found
to be
u = t · (4, −3) , t ∈ R , (15-74)
such that E20 = span{(4, −3)}. The normalized eigenvector v1 = (1/5) · (4, −3) is therefore
an orthonormal basis for E20 (and it can therefore be used as the first column vector in the
wanted Q:
4/5 ∗
Q= . (15-75)
−3/5 ∗
The last column in Q is an eigenvector corresponding to the second eigenvalue λ2 = −5
and can therefore be found from the general solution E−5 to the homogeneous system of
equations having the coefficient matrix
16 −12
KA (−5) = A − 5 · E = , (15-76)
−12 9
but since we know that the wanted eigenvector is orthogonal to the eigenvector v1 we can just
use a vector perpendicular to the first eigenvector, v2 = (1/5) · (3, 4), evidently a unit vector,
that is orthogonal to v1 . It is easy to check that v2 is an eigenvector for A corresponding to
the eigenvalue −5 :
16 −12 3 0
KA (−5) · v2 = · = . (15-77)
−12 9 4 0
eNote 15 15.6 EXAMPLES OF DIAGONALIZATION 386
This matrix has the determinant det (Q) = 1 > 0, so Q is a positive orthogonal substitution
matrix satisfying that Q−1 AQ is a diagonal matrix:
Q−1 AQ = Q> AQ
4/5 −3/5 11 −12 4/5 3/5
= · ·
3/5 4/5 −12 4 −3/5 4/5
(15-79)
20 0
=
0 −5
= diag(20, −5) = Λ .
2 0 −1
The eigenvector solutions to the corresponding homogeneous system of equations are found
to be
u3 = t · (1, 2, 2) , t ∈ R , (15-86)
such that E3 = span{(1, 2, 2)}. The normalized eigenvector v1 = (1/3) · (1, 2, 2) is therefore
an orthonormal basis for E3 so it can be used as the third column vector in the wanted Q; note
that we have just found the eigenvector space to the third eigenvalue on the list of eigenvalues
for A :
∗ ∗ 1/3
Q = ∗ ∗ 2/3 . (15-87)
∗ ∗ 2/3
We know from Theorem 15.31 that the two last columns are found by similarly determining
the eigenvector space E6 corresponding to eigenvalue λ2 = 6, and the eigenvector space E9
corresponding to the eigenvalue λ1 = 9.
For λ2 = 6 we have:
1 −2
0
KA (6) = A − 6 · E = −2 0 −2 , (15-88)
0 −2 −1
which by suitable row operations is found to have the following equivalent reduced matrix:
1 0 1
rref(KA (6)) = 0 2 1 . (15-89)
0 0 0
The eigenvector solutions to the corresponding homogeneous system of equations are found
to be
u2 = t · (−2, −1, 2) , t ∈ R , (15-90)
so that E6 = span{(−2, −1, 2)}. The normalized eigenvector v2 = (1/3) · (−2, −1, 2) is
therefore an orthonormal basis for E6 (and it can therefore be used as the second column vector
in the wanted Q:
∗ −2/3 1/3
to both v3 and v2 , so we can use v1 = v2 × v3 = (1/3) · (−2, 2, −1), and then we finally get
This matrix is positive orthogonal since det(Q) = 1 > 0, and therefore we have determined
a positive orthogonal matrix Q that diagonalizes A to the diagonal matrix Λ. This is easily
proved by direct computation:
Q−1 AQ = Q> AQ
−2/3 2/3 −1/3 7 −2 −2/3 −2/3 1/3
0
= −2/3 −1/3 2/3 · −2 6 −2 · 2/3 −1/3 2/3
1/3 2/3 2/3 0 −2 5 −1/3 2/3 2/3
(15-93)
9 0 0
= 0 6 0
0 0 3
= diag(9, 6, 3) = Λ .
In the light of the above examples it is clear that if only we can construct all orthogonal
(2 × 2)- and (3 × 3)-matrices Q (or for that matter (n × n)-matrices), then we can produce
all symmetric (2 × 2)- and (3 × 3)-matrices A as A = Q · Λ · Q> . We only have to choose
the wanted eigenvalues in the diagonal for Λ.
Every special orthogonal 2 × 2-matrix has the following form, which shows that it is a
rotation given by a rotation angle ϕ :
cos( ϕ) − sin( ϕ)
Q= , (15-94)
sin( ϕ) cos( ϕ)
where ϕ is an angle in the interval [−π, π ]. Note that the column vectors are orthogonal
and both have length 1. Furthermore the determinant det(Q) = 1, so Q is special
orthogonal.
eNote 15 15.7 CONTROLLED CONSTRUCTION OF SYMMETRIC MATRICES 389
Exercise 15.35
Prove the statement that every special orthogonal matrix can be stated in the form (15-94) for
a suitable choice of the rotation angle ϕ.
If ϕ > 0 then Q rotates vectors in the positive direction, i.e. counter-clockwise; if ϕ < 0
then Q rotates vectors in the negative direction, i.e. clockwise.
A rotation about a coordinate axis, i.e. a rotation by a given angle about one of the
coordinate axes, is produced with one of the following special orthogonal matrices:
1 0 0
Rx (u) = 0 cos(u) − sin(u)
0 sin(u) cos(u)
cos(v) 0 sin(v)
Ry ( v ) = 0 1 0 (15-95)
− sin(v) 0 cos(v)
cos(w) − sin(w) 0
Rz (w) = sin(w) cos(w) 0 ,
0 0 1
eNote 15 15.7 CONTROLLED CONSTRUCTION OF SYMMETRIC MATRICES 390
Exercise 15.38
Show by direct calculation that the three axis-rotation matrices and every product of axis-
rotation matrices really are special orthogonal matrices, i.e. they satisfy R−1 = R> and
det(R) = 1.
Exercise 15.39
Find the image vectors of every one of the given vectors a, b, and c by use of the given
mapping matrices Qi :
The combination of rotations about the coordinate axes by given rotation angles u, v,
and w about the x −axis, y−axis, and z−axis is found by computing the matrix product
of the three corresponding rotation matrices.
Here is the complete general expression for the matrix product for all values of u, v and
w:
cos(w) cos(v) − sin(w) cos(u) − cos(w) sin(v) sin(u) sin(w) sin(u) − cos(w) sin(v) cos(u)
= sin(w) cos(v) cos(w) cos(u) − sin(w) sin(v) sin(u) − cos(w) sin(u) − sin(w) sin(v) cos(u) .
sin(v) cos(v) sin(u) cos(v) cos(u)
In other words: the effect of every rotation matrix can be realized by three con-
secutive rotations about the coordinate axes – with the rotation angles u, v, and w,
respectively, as given in the above matrix product.
When a given special orthogonal matrix R is given (with its matrix ele-
ments rij ), it is not difficult to find these axis rotation angles. As is evident
from the above matrix product we have e.g. that sin(v) = r31 such that
v = arcsin(r31 ) or v = π − arcsin(r31 ), and cos(w) cos(v) = r11 such that
w = arccos(r11 / cos(v)) or v = − arccos(r31 / cos(v)), if only cos(v) 6= 0 i.e. if
only v 6= ±π/2.
Exercise 15.41
Show that if v = π/2 or v = −π/2 then there exist many values of u and w giving the same
R(u, v, w). I.e. not all angle values are uniquely determined in the interval ] − π, π ] for
every given rotation matrix R.
Exercise 15.42
Show that if R is a rotation matrix (a positive orthogonal matrix) then R> is also a rotation
matrix, and vice versa: if R> is a rotation matrix then R is also a rotation matrix.
Exercise 15.43
Show that if R1 and R2 are rotation matrices then R1 · R2 and R2 · R1 are also rotation matrices.
Give examples that show that R1 · R2 is not necesarily the same rotation matrix as R2 · R1 .
eNote 15 15.8 STRUCTURE OF ROTATION MATRICES 392
As mentioned above (Exercise 15.35), every 2 × 2 special orthogonal matrix has the
form:
cos φ − sin φ
Q= .
sin φ cos φ
This is a rotation of the plane anticlockwise by the angle φ. The angle φ is related to the
eigenvalues of Q:
Exercise 15.44
λ1 = eiφ , λ2 = e−iφ .
How about the 3 × 3 case? We already remarked that any 3 × 3 special orthogonal matrix
can be written as a composition of rotations about the three coordinate matrices: Q =
Rz (w) · Ry (v) · Rx (u). But is Q itself a rotation about some axis (i.e. some line through
the origin)? We can prove this is so, by examining the eigenvalues and eigenvectors of
Q.
Theorem 15.45
The eigenvalues of any orthogonal matrix all have absolute value 1.
|λ|2 = λλ̄ = 1.
Theorem 15.46
Let Q be a 3 × 3 special orthogonal matrix, i.e. QT Q = E, and det Q = 1. Then the
eigenvalues are:
λ1 = 1, λ2 = eiφ , λ3 = e−iφ ,
for some φ ∈] − π, π ].
Proof. Q is a real matrix, so all eigenvalues are either real or come in complex conjugate
pairs. There are 3 of them, because Q is a 3 × 3 matrix, so the characteristic polynomial
has degree 3. Hence there is at least one real eigenvalue:
λ1 ∈ R.
Case 2: λ1 is real and the other two are complex conjugate, λ3 = λ̄2 , so:
where we used that |λ2 | = 1. Any complex number λ with absolute value 1 is of the
form eiφ , where φ = Arg(λ), so this gives the claimed form of λ1 , λ2 and λ3 .
We can also say something about the eigenvectors corresponding to the eigenvalues.
Theorem 15.47
Let Q be a special orthogonal matrix, and denote the eigenvalues as in Theorem
15.46. If the eigenvalues are not all real, i.e. Im(λ2 ) 6= 0, then the eigenvectors
corresponding to λ2 and λ3 are necessarily of the form:
where x and y are respectively the real and imaginary parts of v2 , and
v1 ·x = v1 ·y = 0
Proof. We have Qv2 = λ2 v2 , and Qv̄2 = λ̄2 v̄2 . So clearly a third eigenvector, corre-
sponding to λ̄2 , is v3 = v̄2 . Using QT Q = E, we have
If v2T v2 6= 0, then we can divide by this number to get λ22 = 1. But λ2 = a + bi, with
b 6= 0, so this would mean: 1 = λ22 = a2 − b2 + 2iab. The imaginary part is: ab = 0,
which implies that a = 0 and hence λ22 = −b2 , which cannot be equal to 1. Hence:
v2T v2 = 0.
v1T v2 = 1 · λ2 · v1T v2 ,
Now we can give a precise description of the geometric effect of a 3 × 3 rotation matrix:
Theorem 15.48
Let Q be a 3 × 3 special orthogonal matrix, and λ1 = 1, λ2 = eiφ , λ3 = e−iφ be its
eigenvalues, with corresponding eigenvectors v1 , v2 and v3 = v̄2 . Then:
v1 Im v2 Re v2
u1 = , u2 = , u3 = ,
|v1 | |Im v2 | |Re v2 |
where v2 is an eigenvector for λ2 = eiφ . The mapping matrix for f with respect
to this basis is:
1 0 0
u f u = 0 cos φ − sin φ .
0 sin φ cos φ
Proof. Statement 1 follows from statements 2 and 3, since these represent rotations by
angle φ around the v1 axis.
This mapping matrix is precisely the matrix of a rotation by angle φ around the v1 axis
(compare u f u with the matrix Rx (u) discussed earlier).
For statement 3, the special case that λ2 is real, if λ2 = λ3 = 1, then φ = 0 and Q is the
identity matrix, which can be regarded as a rotation by angle 0 around any axis.
Finally, for the case λ2 = λ3 = −1, briefly: let E1 = span{v1 }. Choose any orthonormal
basis for the orthogonal complement E1⊥ . Using this, one can show that the restricition of
f to E1⊥ is a 2 × 2 rotation matrix with a repeated eigenvalue −1. This means it is minus
the identity matrix on E1⊥ , i.e. a rotation by angle π, from which the claim follows.
Example 15.49
The axis of rotation for a 3 × 3 rotation matrix is sometimes called the Euler axis. Let’s find
the Euler axis, and the rotation angle for the special orthogonal matrix:
−2 −2 1
1
Q= 2 −1 2 ,
3
−1 2 2
So the axis of rotation is the line spanned by v1 = (0, 1, 2), and the angle of rotation is:
√ !
5
φ = Arg(λ2 ) = − arctan +π
2
We can set:
−1
0 0
1 1
u1 = v1 = √ 1 , u2 = Imv2 = 0 , u3 = Rev2 = √ −2 ,
5 2 0 5 1
Exercise 15.50
√ √
0 − 2 2
√
Find the axis and angle of rotation for the rotation matrix: Q = 12 2 1 1 .
√
− 2 1 1
Conversely, we can construct a matrix that rotates by any desired angle around any
desired axis:
Example 15.51
Problem: Construct the matrix for the linear map f : R3 → R3 that rotates 3-space around
the axis spanned by the vector a = (1, 1, 0) anti-clockwise by the angle π/2.
Solution: Choose any orthonormal basis (u1 , u2 , u3 ) where u1 points in the direction of a. For
example:
−1
1 0
1 1
u1 = √ 1 , u2 = √ 1 , u3 = 0 .
2 0 2 0 1
eNote 15 15.8 STRUCTURE OF ROTATION MATRICES 398
We have chosen them such that det([u1 , u2 , u3 ]) = 1. This means that the orientation of
space is preserved by this change of basis, so we know that the rotation from the following
construction will be anti-clockwise around the axis.
The matrix with respect to the u-basis that rotates anti-clockwise around the u1 -axis by the
angle π/2 is:
1 0 0 1 0 0
u f u = 0 cos( π/2) − sin( π/2) = 0 0 −1 .
0 sin(π/2) cos(π/2) 0 1 0
The change of basis matrix from u to the standard e-basis is:
1 −1 0
1
e Mu = [u1 , u2 , u3 ] = √ 1 1 0 ,
2 0 0 √2
Note: for the vectors u1 and u2 , it would have made no difference what choice we make as
long as they are orthogonal to a, and orthonormal. If we rotate them in the plane orthogonal
to a, this rotation will cancel in the formula e Mu u f u e MuT .
Exercise 15.52
Find an orthogonal matrix Q that, in the standard e-basis for R3 , represents a rotation about
the axis spanned by a = (1, 1, 1) by an angle π/2.
eNote 15 15.9 REDUCTION OF QUADRATIC POLYNOMIALS 399
Definition 15.53
Let A be a symmetric (n × n)-matrix and let ( x1 , x2 , · · · , xn ) denote the coordinates
for an arbitrary vector x in (R, ·) with respect to the standard basis e in Rn .
x1
x2
PA (x) = PA ( x1 , x2 , · · · , xn ) = x1 x2 · · x n ·A·
·
·
(15-98)
xn
n n
= ∑∑ aij · xi · x j ,
i =1 j =1
We will now see how the spectral theorem can be used for the description of every
quadratic form by use of the eigenvalues for the matrix that represents the quadratic
form.
The reduction in the theorem means that the new expression does not contain
any product terms of the type xi · x j for i 6= j.
Proof
Let x be an arbitrary vector in Rn . Then we have the following set of coordinates for x, partly
with respect to the standard e-basis and partly with respect to the new basis v
ex = ( x1 , · · · , x n ) ,
(15-104)
vx = ( xe1 , · · · , xen ) .
eNote 15 15.9 REDUCTION OF QUADRATIC POLYNOMIALS 401
Then
PA (x) = PA ( x1 , · · · , xn )
x1
·
= x1 · · · xn · A ·
·
xn
x1
·
= x 1 · · · x n · Q · Λ · Q−1 ·
·
xn
x1
> ·
x1 · · · x n · Q · Λ ·
= Q · ·
(15-105)
xn
xe1
·
· · · xen · Λ ·
= xe1 ·
xen
λ1 0 · · ·
0 xe1
0 λ2 · · ·
0
·
= xe1 · · · xen · . .. . . ·
..
.. . . . ·
0 0 . . . λn xen
Note that the matrix that represents the quadratic form in Example 15.54,
Equation (15-102), is not much different from the Hessian Matrix H f ( x, y) for
f ( x, y), which is also a constant matrix, because f ( x, y) is a second degree poly-
nomial. See eNote ??. In fact we observe that:
1
A= · H f ( x, y) , (15-106)
2
and this is no coincidence.
eNote 15 15.9 REDUCTION OF QUADRATIC POLYNOMIALS 402
Lemma 15.56
Let f ( x1 , x2 , · · · , xn ) denote an arbitrary quadratic polynomial without linear and
constant terms. Then f ( x1 , x2 , · · · , xn ) can be expressed as a quadratic form in ex-
actly one way – i.e. there exists exactly one symmetric matrix A such that:
f (x) = f ( x1 , x2 , · · · , xn ) = PA ( x1 , x2 , · · · , xn ) . (15-107)
Proof
We limit ourselves to the case n = 2 and refer the analysis to functions of two variables in
eNote ??: If f ( x, y) is a polynomial in two variables without linear (and constant) terms, i.e.
a quadratic form in (R2 , ·), then the wanted A-matrix is exactly the (constant) Hesse-matrix
for f ( x, y).
00 ( x, y, z ) =
where we explicitly have used the symmetry of the Hessian matrix, e.g. f zx
00 ( x, y, z ).
f xz
f ( x, y, z) = x2 + 3 · y2 + z2 − 8 · x · y + 4 · y · z . (15-111)
1 −4 0
x
PA ( x, y, z) = x y z · −4 3 2 · y
0 2 1 z
x−4·y
= x y z · 3·y−4·x+2·z
(15-113)
z+2·y
= x · ( x − 4 · y ) + y · (3 · y − 4 · x + 2 · z ) + z · ( z + 2 · y )
= x 2 + 3 · y2 + z2 − 8 · x · y + 4 · y · z
= f ( x, y, z) .
As is shown in Section ?? in eNote ?? the signs of the eigenvalues for the Hessian matrix
play a decisive role when we analyse and inspect a smooth function f ( x, y) at and about
a stationary point. And since it is again the very same Hessian matrix that appears in the
present context we will here tie a pair of definitions to this sign-discussion – now for the
general (n × n) Hessian matrices, and thus also for general quadratic forms represented
by symmetric matrices A :
eNote 15 15.9 REDUCTION OF QUADRATIC POLYNOMIALS 404
We now formulate an intuitively reasonable result that relates this ”definiteness” to the
values which the quadratic polynomial PA (x) assumes for different x ∈ Rn .
Proof
We refer to Theorem 15.55 and from that we can use the reduced expression for the quadratic
form:
PA ( x1 , · · · , xn ) = PeΛ ( xe1 , · · · , xen ) = λ1 · xe12 + · · · + λn · xen2 , (15-114)
from which it is clear to see that since A is positive definite we get λi > 0 for all i = 1, · · · , n
and then PA (x) > 0 for all x 6= 0, which corresponds to the fact that none of the sets of
coordinates for x can be (0, · · · , 0).
Similar theorems can be formulated for negative definite and indefinite matrices, and
eNote 15 15.10 REDUCTION OF QUADRATIC POLYNOMIALS 405
f ( x, y) = 11 · x2 + 4 · y2 − 24 · x · y − 20 · x + 40 · y − 60 (15-115)
The part of the polynomial that can be described by a quadratic form is now
PA ( x, y) = 11 · x2 + 4 · y2 − 24 · x · y , (15-116)
where
11 −12
A= . (15-117)
−12 4
Exactly this matrix is diagonalized by a positive orthogonal substitution Q in Example 15.32:
The eigenvalues for A are λ1 = 20 and λ2 = −5 and
4/5 3/5 cos( ϕ) − sin( ϕ)
Q= = , where ϕ = − arcsin(3/5) . (15-118)
−3/5 4/5 sin( ϕ) cos( ϕ)
The change of coordinates xe, ye consequently is a rotation of the standard coordinate system
by an angle of − arcsin(3/5).
We use the reduction theorem 15.55 and get that the quadratic form PA ( x, y) in the new
coordinates has the following reduced expression:
By introducing the reduced expression for the quadratic form in the polynomial f ( x, y) we
get:
f ( x, y) = 20 · xe2 − 5 · ye2 + (−20 · x + 40 · y − 60) , (15-120)
eNote 15 15.10 REDUCTION OF QUADRATIC POLYNOMIALS 406
where all that remains is to express the last parenthesis by using the new coordinates. This
is done using the substitution matrix Q. We have the linear relation between the coordinates
( x, y) and ( xe, ye):
x xe 4/5 3/5 xe
= Q· = · (15-121)
y ye −3/5 4/5 ye
so that:
1
· (4 · xe + 3 · ye)
x=
5 (15-122)
1
y = · (−3 · xe + 4 · ye) .
5
We substitute these rewritings of x and y in (15-120) and get:
Thus we have reduced the expression for f ( x, y) to the following expression in new coordi-
nates xe and ye, that appears by a suitable rotation of the standard coordinate system:
f ( x, y) = 11 · x2 + 4 · y2 − 24 · x · y − 20 · x + 40 · y − 60
= 20 · xe2 − 5 · ye2 − 40 · xe + 20 · ye − 60 (15-124)
= fe( xe, ye) .
Note again that the reduction in Example 15.60 results in the reduced quadratic
polynomial fe( xe, ye) not containing any product terms of the form xe · ye. This
reduction technique and the output of the large work becomes somewhat more
clear when we consider quadratic polynomials in three variables.
In Example 15.34 we have diagonalized the matrix A that represents the quadratic form in
the following quadratic polynomial in three variables:
f ( x, y, z) = 7 · x2 + 6 · y2 + 5 · z2 − 4 · x · y − 4 · y · z − 2 · x + 20 · y − 10 · z − 18 . (15-125)
This polynomial is reduced to the following quadratic polynomial in the new variables ob-
tained using the same directives as in Example 15.60:
The substitution matrix Q can be factorized to a product of axis-rotation matrices like this:
Q = Rz ( w ) · Ry ( v ) · R x ( u ) , (15-128)
By rotation of the coordinate system and by using the new coordinates xe, ye, and e z we ob-
tain a reduction of the polynomial f ( x, y, z) to the end that the polynomial f ( xe, ye, e
e z) does
not contain product terms while f ( x, y, z) contains two product terms, with x · y and y · z,
respectively.
eNote 15 15.11 SUMMARY 408
15.11 Summary
The main result in this eNote is that symmetric (n × n)-matrices are precisely those
matrices that can be diagonalized by a special orthogonal change of basis matrix Q.
We have used this theorem for the reduction of quadratic polynomials in n variables –
though particularly for n = 2 and n = 3.
|a · b| ≤ | a | | b | , (15-132)
and the equality sign applies if and only if a and b are linearly dependent.
• The angle θ ∈ [0, π ] between two proper vectors a and b in (Rn , ·) is determined
by
a·b
cos(θ ) = . (15-133)
|a| · |b|
• Two proper vectors a and b in (Rn , ·) are orthogonal if a · b = 0.
• A matrix Q is orthogonal if the column vectors are pairwise orthogonal and each
has length 1 with respect to the scalar product introduced. This corresponds ex-
actly to
Q> · Q = E (15-134)
or equivalently:
Q−1 = Q> . (15-135)
eNote 15 15.11 SUMMARY 409
• Every special orthogonal matrix Q is change of basis matrix that rotates the coordinate-
system. It can for n = 3 be factorized in three axis-rotation matrices:
Q = Rz ( w ) · Ry ( v ) · R x ( u ) , (15-137)
and such that the reduced quadratic form PeΛ ( xe, ye, e
z) does not contain any product
term of the type xe · ye, xe · e
z, or ye · e
z:
z) = λ1 · xe2 + λ2 · ye2 + λ3 · e
PeΛ ( xe, ye, e z2 , (15-139)
eNote 16
In this eNote we first give a short introduction to differential equations in general and then the
main subject is a special type of differential equation the so-called first order differential
equations. The eNote is based on knowledge of special functions, differential and integral
calculus and linear maps.
One says that a differential equation has the order n if it contains the nth derivative of
the unknown function, but no derivatives of order higher than n . The unknown func-
tion is in this eNote denoted by x or x (t) , if the name of the independent variable t is
important in the context.
The equation has order 3, since the highest number of times the unknown function x
is differentiated in the equation is 3. A solution to the equation is a function x0 which,
inserted into the equation, makes it true. If we, for example, want to investigate whether
the function
x 0 ( t ) = et + t + 2 , t ∈ R
x0000 (t) − 2x00 (t) + x0 (t) = (et + t + 2)000 − 2(et + t + 2)0 + ((et + t + 2))
= et − 2 (et + 1 ) + et + t + 2
= t,
x0 is a solution.
This eNote is about an important type of first order differential equation, the so-called
linear differential equations. In order to be able to investigate these precisely we first
express them in a standard way.
Definition 16.1
By a first order linear differential equation we understand a differential equation that
can be brought into the standard form
The equation is called homogeneous if q(t) = 0 for all t . Otherwise it is called inho-
mogeneous.
eNote 16 16.2 INTRODUCTION TO FIRST ORDER LINEAR DIFFERENTIAL
EQUATIONS 412
Let I be an open interval in R . Consider the first order linear differential equation
In order to bring this into standard form we first have to add 8t2 to both sides of the equation,
since on the left-hand side only terms containing the unknown function x (t) must appear.
Then we divide both sides by t, since the coefficient of x 0 (t) must be 1 in the standard form.
To avoid division by 0, we must assume that t is either greater or less than zero: let us choose
the first:
2 10
x 0 (t) + x (t) = 8t − , t > 0 . (16-5)
t t
2
Now the differential equation is in the standard form with p(t) = and
t
10
q(t) = 8t − , t > 0 .
t
1 2
x 0 (t) + t = 0, t∈R
2
is not homogeneous.
eNote 16 16.2 INTRODUCTION TO FIRST ORDER LINEAR DIFFERENTIAL
EQUATIONS 413
x 0 (t) + 2x (t) = 30 + 8t , t ∈ R.
is a solution.
From Exercise 16.5 it appears that a differential equation can have more than one solu-
tion. We will in what follows investigate in more detail the question about the number
of solutions. In order to understand precisely what is meant by a first order differential
equation being linear and what this means for its solution set, we will need the follow-
ing lemma.
Lemma 16.6
Let p be a continuous function defined on an open interval I in R . Then the map
f : C1 ( I ) → C0 ( I ) given by
is linear.
Proof
We will show that f satisfies the two linearity requirements L1 and L2 . Let x1 , x2 ∈ C1 ( I )
(i.e. the two functions are arbitrary differentiable functions with continuous derivatives on
I ), and let k ∈ R . That f satisfies L1 appears from
f (kx1 (t)) = (kx10 (t)) + p(t)(kx1 (t)) = k ( x10 (t) + p(t) x1 (t))
= k f ( x1 (t)) .
From Lemma 16.6 we can deduce important properties for the solution set for first order
linear differential equations. First we introduce convenient notations for the solution
sets that we will treat.
1. If the equation is homogeneous (i.e. q(t) is the 0-function), then the solution
set is a vector subspace of C1 ( I ) .
Proof
1. Lhom is equal to ker( f ) . Since the kernel for every linear map is a subspace of the
domain, Lhom is a subspace of C1 ( I ) .
2. Since the equation f ( x (t)) = x 0 (t) + p(t) x (t) = q(t) is linear, the structure theorem
follows directly from the general structure theorem for linear equations (see eNote 12,
Theorem 12.14).
3. The superposition principle follows from the fact that f satisfies the linearity require-
ment L1 . Assume that f ( x1 (t)) = q1 (t) and f ( x2 (t)) = q2 (t) . Then
When we call a first order differential equation of the form (16-2) linear, it is – as shown
above – closely related to the fact that its left-hand side represents a linear map, and that
its solution set therefore has the unique properties of Theorem 16.7. In the following
example we juggle with the properties in order to decide whether a given differential
equation is not linear.
Here we will show that one, in different ways, can demonstrate that the differential equation
is not linear.
1. We can show directly that f does not satisfy the linearity conditions. To show this we
can test L2 , e.g. with k = 2 . We compute the two sides in L2 :
where the right-hand side is only the 0-function when x (t) is the 0-function. Since L2
applies for all x (t) ∈ C1 (R) , L2 is not satisfied. Therefore the equation is nonlinear.
2. The solution set to the corresponding homogeneous equation is not a subspace. E.g.
it does not satisfy the stability requirement with repsect to multiplication by a scalar
which we can show as follows:
1
The function x0 (t) = − is a solution to the homogeneous eqution because
t
1 1
x00 (t) − ( x0 (t))2 = 2
− 2 = 0.
t t
2
But 2 · x0 (t) = − is not, because
t
2 4 2
(2 · x0 (t))0 − (2 · x0 (t))2 = 2
− 2 = − 2 6= 0 .
t t t
3. The solution set does not satisfy the superposition principle. E.g. we see that
1 1 2
f − = 0 and f = − 2 , while
t t t
1 1 2
f − + = 0 6= 0 − 2 .
t t t
It follows from the structure theorem that homogeneous equations play a special role
for linear differential equations. Therefore we treat them separately in the next section.
eNote 16 417
16.3 HOMOGENEOUS FIRST ORDER LINEAR DIFFERENTIAL EQUATIONS
We now establish a solution formula for homogeneous first order linear equations.
is then given by
x ( t ) = c e− P ( t ) , t ∈ I (16-12)
where c is an arbitrary real number.
Proof
The theorem follows from the fact that the derivative of a function g(t) is zero on an interval
if and only if that function is constant. We apply this to the function g(t) = x (t)e P(t) . Using
the chain rule and the product rule for differentiation we have:
0
x (t)e P(t)) = e P(t) x 0 ( t ) + p ( t ) e P(t) x ( t )
= e P(t) x 0 ( t ) + p ( t ) x ( t ) .
Since e P(t) 6= 0, the above expression is zero if and only if the equation (16-11) holds. That is,
the differential equation (16-11) is equivalent to the equation:
0
x (t)e P(t)) = 0.
As mentioned, this is equivalent to the statement:
x (t)e P(t) = c, (16-13)
where c is some real constant, i.e. that x (t) = ce− P(t) . This shows that not only is ce− P(t) a
solution, for any constant c, but that any solution to (16-11) must be of this form, since it must
satisfy Equation (16-13) for some c.
eNote 16 418
16.3 HOMOGENEOUS FIRST ORDER LINEAR DIFFERENTIAL EQUATIONS
We already know that the solution set is a subspace of C1 ( I ) . From the formula
(16-12) we now know that the subspace is 1-dimensional, and that the function
e− P(t) is a basis for the solution set.
Remark 16.10
Theorem 16.9 is also valid if p is a continuous complex-valued function, with the
slight modification that the arbitrary constant c is now a complex constant. The
proof is exactly the same, because the product rule is the same for complex-valued
functions of t, and, writing P(t) = u(t) + iv(t), one finds that the derivative of e P(t)
is still P0 (t)e P(t) . Finally, by separating the function into real and imaginary parts,
one again finds that the derivative of a complex-valued function is zero if and only
if the function is equal to a complex constant.
Exercise 16.11
In Theorem 16.9 an arbitrary antiderivative P(t) for p(t) is used. Explain why it is immate-
rial to the solution set which antiderivative you use when you apply the theorem.
We see that that the coefficient function p(t) = cos(t) . An antiderivative for p(t) is
P(t) = sin(t) . Then the general solution can be written as
Now that we know how to find the general solution for homogeneous first order linear
differential equations, it is about time to look at the inhomogeneous ones. If you already
know or can guess a particular solution to the inhomogeneous equation, it is obvious
to use the structure theorem, see Theorem 16.7. This is demonstrated in the following
examples.
It is easily seen that x0 (t) = 1 is a particular solution. Then we solve the corresponding
homogeneous differential equation
Using symbols from Theorem 16.9 we have p(t) = t that has the antiderivative
1 2
P(t) = t .
2
The general solution therefore consists of the following functions where c is an arbitrary real
number:
1 2
x (t) = ce− 2 t , t ∈ R . (16-18)
In short: n o
1 2
Lhom = ce− 2 t , t∈R c∈R . (16-19)
Now we can establish the general solution to the inhomogeneous differential equation using
the structure theorem as:
n o
1 2
Linhom = x0 (t) + Lhom = 1 + ce− 2 t , t ∈ R c ∈ R .
eNote 16 16.4 INHOMOGENEOUS EQUATIONS SOLVED BY THE GUESS METHOD420
First let us try to guess a particular solution. Since the right-hand side is first degree poly-
nomial, one can – with the given left-hand side, where you only differentiate and multiply
by 2 – assume that a first degree polynomial could be a solution. Therefore we try to insert
an arbitrary first degree polynomial x0 (t) = b + at in the left-hand side of the differential
equation:
x00 (t) + 2x0 (t) = (b + at)0 + 2(b + at) = a + 2b + 2at .
a + 2b + 2at = 30 + 8t
a + 2b = 30 and 2a = 8 ⇔ a = 4 and b = 13 .
x0 (t) = 13 + 4t , t ∈ R .
Using symbols from Theorem 16.9 we have p(t) = 2 that has the antiderivative P(t) = 2t .
Therefore the general solution consists of the following functions where c is an arbitrary real
number:
x (t) = ce−2t , t ∈ R . (16-22)
In short:
ce−2t , t ∈ R c ∈ R .
Lhom = (16-23)
Now its is possible to establish the general solution to the inhomogeneous differential equa-
tion using the structure theorem:
First let us try to guess a particular solution. Since the right-hand side consists of constant
plus a sine function with the angular frequency 2, it is obvious to guess a solution the type
Since the set (cos(2t), sin(2t), 1) is linearly independent, this equation is satisfied exactly
when
2 1
2b + a = 0 , b − 2a − 1 = 0 and k = 1 ⇔ a = − , b = and k = 1 .
5 5
2 1
x0 ( t ) = 1 − cos(2t) + sin(2t) , t ∈ R .
5 5
we get the general solution to the given inhomogeneous differential equation by use of the
structure theorem:
As demonstrated in the three previous examples it makes sense to use the guess method
in the inhomogeneous cases, when you already know a particular solution or easily can
eNote 16 16.5 THE GENERAL SOLUTION FORMULA 422
find one. It only requires that you can find an antiderivative P(t) for the coefficient
function p(t) .
Otherwise if you do not have an immediate particular solution, you must use the gen-
eral solution formula (see below) instead. Here you get rid of the guesswork, but you
must find two antiderivatives, one is P(t) as above, while the other often is somewhat
more difficult (if not impossible) to find, since you must integrate a product of func-
tions. In the following section we establish the general solution formula and discuss the
said problems.
Now we consider the general first order linear differential equation in the standard form
x 0 ( t ) + p ( t ) x ( t ) = q ( t ), t ∈ I, (16-27)
We can determine the general solution using the following general formula.
x 0 ( t ) + p ( t ) x ( t ) = q ( t ), t∈I (16-28)
Proof
The second term in the solution formula (16-29) we identify as Lhom . If we can show that the
first term is a particular solution to the differential equation, then it follows from the struc-
ture theorem that the solution formula is the general solution to the differential equation.
eNote 16 16.5 THE GENERAL SOLUTION FORMULA 423
First we must of course ask ourselves whether the indefinite integral that is part of the solu-
tion formula even exists. It does! See a detailed reasoning for this in the proof of the existence
and uniqueness Theorem 16.24. That the first term
Z
x 0 ( t ) = e− P ( t ) eP(t) q(t)dt
is a particular solution we show by testing. We insert the term in left-hand side of the differ-
ential equation and see that the result is equal to the right-hand side.
Z 0 Z
x00 (t) + p ( t ) x0 ( t ) = e − P(t)
e P(t)
q(t) dt + p ( t ) e− P ( t ) eP(t) q(t) dt
Z Z
= − p ( t )e− P ( t ) eP(t) q(t) dt + e− P(t) eP(t) q(t) + p(t)e− P(t) eP(t) q(t)dt
= q(t) .
Remark 16.17
Using Remark 16.10, it is straightforward to show that Theorem 16.16 is also valid
if p(t) and q(t) are continuous complex-valued functions, with the modification that
the arbitrary constant c is a complex constant.
If one inserts q(t) = 0 in the general solution formula (16-29), the first term
disappears, and what is left is the second term that is the formula (16-12) for
homogeneous equation. Therefore the formula (16-29) is a "‘general formula"’
that covers both the homogeneous and the inhomogeneous case.
Exercise 16.18
Z
The solution formula (16-29) includes the indefinite integral eP(t) q(t)dt , that represents an
arbitrary antiderivative of e P(t) q(t) . Explain why it does not matter to the solution set which
antiderivative you choose to use, when you apply the formula.
Now we give a few examples using the general solution formula. Since it contains
eNote 16 16.5 THE GENERAL SOLUTION FORMULA 424
an indefinite integral of a product of functions you will often need integration by parts,
which the second example demonstrates.
2 10
x 0 (t) + x (t) = 8t − , t > 0. (16-30)
t t
2 10
With the symbols in the general solution formula we have p(t) = t and q(t) = 8t − t . An
antiderivative for p(t) is given by:
P(t) = 2 ln t . (16-31)
We then have
−2 ) 1
e− P(t) = e−2 ln t = eln(t = t −2 = . (16-32)
t2
From this it follows that eP(t) = t2 . Now we use the general solution formula:
Z
− P(t)
x (t) = e eP(t) q(t) dt + ce− P(t)
1 10 1
Z
2
= t 8t − dt + c 2
t2 t t
1 1
Z
= 2 (8t3 − 10t) dt + c 2 (16-33)
t t
1
= 2 2t4 − 5t2 + c
t
c
x (t) = 2t2 − 5 + , t > 0.
t2
The general solution consists of these functions where c is an arbitrary real number. In short:
n c o
Linhom = x (t) = 2t2 − 5 + 2 , t > 0 c ∈ R . (16-34)
t
We then have
1
e− P(t) = eln t = t and eP(t) = e− ln t = (eln t )−1 = . (16-37)
t
Now we use the general solution formula::
Z
x ( t ) = e− P ( t ) eP(t) q(t) dt + ce− P(t)
1 2
Z
=t t sin(2t) dt + ct
Z t
= t t sin(2t) dt + ct .
Now we perform an intermediate computation where we use integration by parts to find the
antiderivative.
1 1
Z Z
t sin(2t)dt = − t cos(2t) − − cos(2t) dt
2 2
1 1
Z
= − t cos(2t) + cos(2t) dt
2 2
1 1
= − t cos(2t) + sin(2t) .
2 4
And return to the computation
Z
x (t) = t t sin(2t) dt + ct
1 1
= t − t cos(2t) + sin(2t) + ct
2 4
1 1
x (t) = − t2 cos(2t) + t sin(2t) + ct t > 0 .
2 4
The general solution consists of these functions where c is an arbitrary real number. In short:
Until now we have considered the general solution to the differential equation. Often
one is interested in a particular solution that for a given value of t assumes a desired
functional value, a so-called initial value problem. We treat this in the next section.
If we need a solution to the equation that for a given value of t assumes a desired
functional value, the following questions arise: 1) Is there even a solution that satisfies
the desired properties and 2) If yes, how many solutions are there? Before we answer
these question generally, we consider a couple of examples.
In the Example 16.13 we found the general solution to the differential equation
viz.
1 2
x (t) = 1 + ce− 2 t , t ∈ R
where c is an arbitrary real number.
Now we will find the solution x0 (t) that satisfies the initial value condition x0 (0) = 3 . This
is done by insertion of the initial value in the general solution, whereby we determine c :
1 2
x0 (0) = 1 + ce− 2 ·0 = 1 + c = 3 ⇔ c = 2 . (16-41)
Therefore the conditioned solution function to the differential equation is given by
1 2
x0 (t) = 1 + 2e− 2 t , t ∈ R . (16-42)
The figure below shows the graphs for the seven solutions that correspond to initial value
conditions x0 (0) = b where b ∈ {−3, −2, −1, 0, 1, 2, 3} . The solution we just found is the
uppermost. The others are found in a similar way.
eNote 16 16.6 INITIAL VALUE PROBLEMS 427
2 10
x 0 (t) + x (t) = 8t − , t > 0, (16-43)
t t
viz.
c
x (t) = 2t2 − 5 + , t>0
t2
where c is an arbitrary real number.
Now we will find the particular solution x0 (t) that satisfies the initial value condition
x0 (1) = 2 . It is done by insertion of initial value in the general solution, whereby we de-
termine c :
c
x 0 ( 1 ) = 2 · 12 − 5 + 2 2 − 5 + c = 2 ⇔ c = 5 . (16-44)
1
Therefore the conditioned solution function to the differential equation is given by
5
x0 (t) = 2t2 − 5 + , t > 0. (16-45)
t2
The figure below shows the graphs for the seven solutions that correspond to initial value
conditions x0 (0) = b where b ∈ {−4, −3, −2, −1, 0, 1, 2} . The solution we just found is the
uppermost. The others are found in a similar way.
eNote 16 16.6 INITIAL VALUE PROBLEMS 428
viz.
2 1
x (t) = 1 − cos(2t) + sin(2t) + ce−t , t ≥ 0. (16-47)
5 5
Here we show a series of solutions with the initial values from -1 to 3 for t = 0 :
The figure indicates that all solutions approach a periodic oscillation when t → ∞ . That this
is the case is seen from the general solution of the differential equation where the fourth term
ce−t regardless of the choice for c is negligible due to the negative exponent. The first three
terms constitute the the stationary response.
In the three preceding examples we did not have any difficulties in finding a solution
to the differential equation that satisfied a given initial condition. In fact we saw that,
for each of the initial value conditions considered, exactly one solution that satisfied the
condition exists. That this applies in general we show in the following theorem.
eNote 16 16.6 INITIAL VALUE PROBLEMS 429
x 0 ( t ) + p ( t ) x ( t ) = q ( t ), t∈I (16-48)
where I is an open interval and p(t) and q(t) are continuous functions on I .
Then: for every number pair (t0 , b) exactly one (particular) solution x0 (t) to the
differential equation exists that satisfies the inital value condition
x0 ( t0 ) = b . (16-49)
Proof
From Theorem 16.16 we know that the set of solutions to the differential equation (16-48) is
given by Z
x ( t ) = e− P ( t ) eP(t) q(t)dt + ce− P(t) (16-50)
Let us first investigate the indefinite integral that is included in the formula. Does it exist?
This is equivalent to asking: does an antiderivative for the function under the integration
sign exist? We must start with p(t) . Since it is continuous, it has an antiderivative which
we call P(t) . Being an antiderivative, P(t) is differentiable and thus continuous. Since the
exponential function is also continuous the composite function eP(t) is continuous. Finally
since q(t) is continuous, the product eP(t) q(t) is continuous.
By this we have shown that the function under the integration sign is continuous. Therefore
it has an antiderivative, in fact infinitely many antiderivatives that only differ from each other
by constants. We choose an arbitrary antiderivative and call it F (t) . Now we can reformulate
the solution formula as
x (t) = e− P(t) F (t) + ceP(t) (16-51)
where c is an arbitrary real number. Then we insert the initial value condition:
where we first multiplied by eP(t0 ) on both sides of the equality sign and then isolated c .
Thus in the general solution set exactly one solution exists that satisfies the initial value
condition, viz. the one that emerge when we in (16-51) insert the found value of c .
Exercise 16.25
Again let us consider the linear map f : C1 ( I ) → C0 ( I ) that represents the left-hand side of
a first order linear differential equation:
We know that ker( f ) is one dimensional and has the basis vector e− P(t) . But what is the
image space (the range) for f ?
We end this section by an example that shows how it is possible to “go backwards” from
a given general solution to the differential equation it solves.
First we consider the corresponding homogeneous differential equation. With the structure
theorem in mind we immediately see that
c + p(t)ct = 0 , (16-55)
1
p(t) = − . (16-56)
t
Since we now know p(t) , it only remains to determine the right-hand side q(t) . We find this
by insertion of the particular solution x (t) = te−5t into the left-hand side of the equation.
1
e−5t − 5te−5t − · te−5t = −5te−5t = q(t) . (16-57)
t
Now since both p(t) and q(t) are determined, the whole differential equation is determined
as:
1
x 0 (t) − x (t) = −5te−5t , t > 0 . (16-58)
t
In some cases we know in advance what type of solutions to the differential equation are
of interest. Therefore one can choose to restrict the domain C1 (R) . We end this eNote
with an example where the domain is a finite dimensional subset of C1 (R) which leads
to the introduction of matrix methods.
In this example we are only interested in solutions that belong to the polynomial space
P2 (R) , i.e. the subset of C1 (R) that has the monomial base (1, t, t2 ) .
To find the range f ( P2 (R)) of the linear map f that represents the left-hand side of the
differential equation, we first determine the images of the basis vectors:
Since P3 (R) has the monomial base (1, t, t2 , t3 ) , and the found images lie in their span, we
see that the range f ( P2 (R)) is a subspace of P3 (R) .
F x = b,
where F is the mapping matrix for f with respect to the monomial bases in P2 (R) and
P3 (R) , x is the coordinate matrix for the unknown polynomial with respect to the monomial
basis in P2 (R) , and b is the coordinate matrix for the right-hand side of the differential
equation with respect to the monomial basis in P3 (R) .
Thus, when restricted to P2 (R), the differential equation becomes an inhomogeneous system
of linear equations. The first three columns of the augmented matrix T of the system are
given by F , while the fourth column is b :
1 1 0 0 1 0 0 −1
−2 1 2 7 0 1 0 1
T =
0 −2 → rref(T) = .
1 0 0 0 1 2
0 0 −2 −4 0 0 0 0
Since the rank of T is seen to be 3, the differential equation has only one solution. Since
the fourth column in rref(T) states the coordinate vector of the solution with respect to the
monomial basis in P2 (R) , the solution can immediately be stated as:
x0 (t) = −1 + t + 2t2 .
Exercise 16.28
1. Solve the differential equation in Example 16.27 by the guess method or the general
solution formula.
2. How does the general solution differ from the one found in the example?
eNote 16 16.7 FINITE DIMENSIONAL DOMAIN 433
Exercise 16.29
Replace the right-hand side in the differential equation in Example 16.27 by the function
q(t) = 1 .
1. Show, using matrix computation, that the differential equation does not have a solution
in the subspace P2 (R) given in the example.
2. Using Maple (or other software), find the solution x0 (t) to the differential equation
that satisfies the initial value condition x0 (t) = 0 and draw its graph.
eNote 17 434
eNote 17
This eNote describes systems of linear first-order differential equations with constant
coefficients and shows how these can be solved. The eNote is based on eNote 16, which describes
linear differential equations in general. Thus it is a good idea to read that eNote first. Moreover
eigenvalues and eigenvectors are used in the solution procedure, see eNotes 13 and 14.
(Updated: 9.11.21 by David Brander).
A is called the system matrix. It is now the aim to solve such a system of differential
equations, that is, we wish to determine x(t) = ( x1 (t), x2 (t), . . . , xn (t)).
In order to be able to find the full solution to such an equation one should have as
many equations as one has unknown equations (with corresponding derivatives).
Thus the second equation in the system might be:
We now have as many equations (two), as we have unknown functions (two), and it
is now possible to determine both x1 (t) and x2 (t).
For greater clarity we write the system of differential equations in matrix form. The
system above looks like this:
0
x1 ( t ) 4 −1 x1 ( t ) 0 4 −1
= ⇔ x (t) = x(t) = Ax(t) (17-6)
x20 (t) −6 2 x2 ( t ) −6 2
Disregarding that are operating with vectors and matrices the system of equations
looks like something we have seen before: x 0 (t) = A · x (t), something we were
able to solve in high school. The solution to this differential equation is trivial:
x (t) = ce At , where c is an arbitrary constant. Below we find that the solution to the
corresponding system of differential equations is similar in structure to x (t) = ce At .
eNote 17 436
We now solve the system of differential equations in the following Theorem 17.2. The
theorem contains requirements that are not always satisfied. The special cases where
the theorem is not valid are investigated later. The proof uses a well-known method,
the so-called diagonalization method.
Theorem 17.2
Let A ∈n×n . A system of linear differential equations consisting of n equations with
a total of n unknown functions is given by
In the theorem we use the general complex solution for the system of differ-
ential equations. Therefore the general real solution can be found as the real
subset of the complex solution.
Proof
We guess that a solution to the system of differential equations x0 (t) = Ax(t) is a vector v
multiplied by eλt , λ being a constant, such that x(t) = eλt v. We then have the derivative
x0 (t) = λeλt v . (17-9)
If this expression for x0 (t) is substituted into (17-7) together with the expression for x(t) we
get:
λeλt v = Aeλt v ⇔ Av − λv = 0 ⇔ (A − λE)v = 0 (17-10)
eλt is non-zero for every t ∈ R, and can thus be eliminated. The resulting equation is an
eigenvalue problem. λ is an eigenvalue of A and v is the corresponding eigenvector. They
eNote 17 437
can both be determined. We have now succeeded in finding that eλt v is one solution to the
system of differential equations, when λ is an eigenvalue and v the corresponding eigenvec-
tor of A.
In order to find the general solution we use the so-called diagonalization method:
Furthermore we introduce the function y with y(t) = (y1 (t), y2 (t), . . . , yn (t)) such that
We then get x0 (t) = Vy0 (t). If these expressions for x(t) og x0 (t) are substituted into Equation
(17-7) we get
Vy0 (t) = AVy(t) ⇔ y0 (t) = V−1 AVy(t) = Λy(t), (17-13)
where Λ = V−1 AV = diag(λ1 , λ2 , . . . λn ) is a diagonal matrix with the eigenvalues of A .
We now get the equation y0 (t) = Λy(t), which can be written in the following way:
since Λ only has non-zero elements in the diagonal. In this system the single equations are
uncoupled: each of the equations only contains one function and its derivative. Therefore
we can solve them independently and the general solution for every equation is y(t) = ceλt
for all c ∈ C. In total this yields the general solution consisting of the functions below for all
c1 , c2 , . . . , cn ∈ C:
c 1 eλ1 t
y1 ( t )
y 2 ( t ) c 2 eλ2 t
y( t ) = . = . (17-15)
.. ..
yn (t) c n eλ n t
Since we now have the solution y(t) we can also find the solution x(t) = Vy(t):
c 1 eλ1 t
λ2 t
c2 e
x ( t ) = v1 v2 . . . v n .
.. (17-16)
cn eλ n t
λ1 t λ2 t
= c1 e v1 + c2 e v2 + . . . + c n e λ n t v n .
eNote 17 17.1 TWO COUPLED DIFFERENTIAL EQUATIONS 438
Now we have found the general complex solution to the system of equations in Equation
(17-7) consisting of the functions x(t) for all c1 , c2 , . . . , cn ∈ C.
Example 17.3
The solution is found using Theorem 17.2. Another way of writing the solution is to separate
the system of equations so that
Given a linear homogeneous first order system of differential equations with constant
coefficients with n equations and n unknown functions
x0 (t) = Ax(t) . t∈R (17-21)
If the system matrix A has n linearly independent eigenvectors, the real solution can be
found using Theorem 17.2. If the eigenvalues are real then the real solution can be writ-
eNote 17 17.1 TWO COUPLED DIFFERENTIAL EQUATIONS 439
ten directly following formula (17-8) in the theorem, where the n corresponding linearly
independent eigenvectors are real and the arbitrary constants are stated as being real. If
the system matrix has eigenvalues that are not real then the real solution can be found
by extracting the real subset of the complex solution. Also in this case the solution can
be written as a linear combination of n linearly independent real solutions to the system
of differential equations.
We are left with the special case in which the system matrix does not have n linearly in-
dependent eigenvectors. Also in this case the real solution will be a linear combination
of n linearly independent real solutions to the system of differential equations. Here
the method of diagonalization obviously cannot be used and one has to resort to other
methods.
In this section we show the three cases above for systems consisting of n = 2 coupled
differential equations with 2 unknown functions.
eNote 17 17.1 TWO COUPLED DIFFERENTIAL EQUATIONS 440
Method 17.4
The general real solution to the system of differential equations
First determine the eigenvalues of A. For the roots of the characteristic polynomial
A there are three possibilities:
• Two real single roots. In this case both of the eigenvalues λ1 and λ2 have the
algebraic multiplicity 1 and geometric multiplicity 1 and we can put
• Two complex roots. The two eigenvalues λ and λ̄ are then conjugate complex
numbers. We then determine u1 and u2 using Method 17.5.
• One double root. Here the eigenvalue λ has the algebraic multiplicity 2. If the
geometric multiplicity of λ is 1, u1 and u2 are determined using method 17.7.
In the first case in Method 17.4 with two different real eigenvalues, Theorem 17.2 can be
used directly with the arbitrary constants chosen as real, see Example 17.3.
Now follows the method that covers the case with two complex eigenvalues.
eNote 17 17.1 TWO COUPLED DIFFERENTIAL EQUATIONS 441
Method 17.5
Two linearly independent real solutions to the system of equations
Example 17.6
Finally we describe the method that can be used if the system matrix has the eigenvalue
λ with am(λ) = 2 and gm(λ) = 1, that is when diagonalization is not possible.
Method 17.7
If the system matrix A to the system of differential equations
has one eigenvalue λ with algebraic multiplicity 2, but the corresponding eigen-
vector space only has geometric multiplicity 1, there are two linearly independent
solutions to the system of differential equations of the form:
u1 (t) = eλt v
(17-33)
u2 (t) = teλt v + eλt b,
Proof
It is evident that one solution to the system of differential equations is u1 (t) = eλt v. The
difficulty is to find another solution.
The first equation can easily be transformed into Av = λv, which is seen to be true, since v is
eNote 17 17.1 TWO COUPLED DIFFERENTIAL EQUATIONS 443
v + λb − Ab = 0 ⇔
Ab − λb = v ⇔ (17-38)
(A − λE)b = v
If b satisfies the given system of equations, u2 (t) will also be a solution to the system of
differential equations. We now have found two solutions and we have to find out whether
these are linearly independent. This is done by a normal linearity criterion: If the equation
k1 u1 + k2 u2 = 0 only has the solution k1 = k2 = 0 then u1 and u2 are linearly independent.
k1 u1 + k2 u2 = 0 ⇒
k1 eλt v + k2 (teλt v + beλt ) = 0 ⇔
(17-39)
t ( k 2 v) + ( k 1 v + k 2 b) = 0 ⇔
k2 v = 0 ∧ k1 v + k2 b = 0
Since v is an eigenvector, it is not the zero-vector, and hence k2 = 0 according to the first
equation. Thus the other equation is reduced to k1 v = 0, and with the same argument we get
k1 = 0. Therefore the two solutions are linearly independent, and thus the method has been
proved.
Example 17.8
16 − λ −1
det(A − λE) = = (16 − λ)(12 − λ) + 4
4 12 − λ (17-41)
2 2
= λ − 28λ + 196 = (λ − 14) = 0
There is only one eigenvalue, viz. λ = 14, even though it is a 2 × 2-system. The eigenvectors
are determined:
1 − 12
16 − 14 −1 2 −1
A − 14E = = → (17-42)
4 12 − 14 4 −2 0 0
We then obtain the eigenvector ( 12 , 1) or v = (1, 2). We can then conclude that the eigen-
value λ has the algebraic multiplicity 2, but that the corresponding eigenvector space has the
eNote 17 17.2 N-DIMENSIONAL SOLUTION SPACE 444
1 − 12 12
2 −1 1
→ (17-44)
4 −2 2 0 0 0
This yields b = (1, 1), if the free parameter is put at1. The two solutions then are
14t 1
u1 (t) = e
2
(17-45)
14t 1 14t 1
u2 (t) = te +e .
2 1
By use of Method 17.4 the general solution can be determined to the following functions for
all c1 , c2 ∈ R:
14t 1 14t 1 1
x(t) = c1 u1 (t) + c2 u2 (t) = c1 e + c2 e t + . (17-46)
2 2 1
In the preceding section we have considered coupled systems consisting of two linear
equations with two unknown functions. The solution space is two-dimensional, since
it can be written as a linear combination of two linearly independent solutions. This
can be generalized to arbitrary systems with n ≥ 2 coupled linear differential equations
with n unknown functions: The solution is a linear combination of exactly n linearly
independent solutions. This is formulated in a general form in the following theorem.
eNote 17 17.2 N-DIMENSIONAL SOLUTION SPACE 445
Theorem 17.9
Given the linear homogeneous first order system of differential equations with con-
stant real coefficients
x0 (t) = Ax(t), t ∈ R, (17-47)
consisting of n equations and with n unknown functions. The general real solution
to the system is n-dimensional and can be written as
where u1 (t), u2 (t), . . . , un (t) are linearly independent real solutions to the system of
differential equations and c1 , c2 , . . . , cn ∈ R.
Below is an example with a coupled system of three differential equations that exempli-
fies Theorem 17.9.
−9 10 0
We wish to determine the general real solution to the system of differential equations. Eigen-
values and eigenvectors can be determined and are as follows:
λ1 = −4 : v1 = (10, 5, 1)
λ2 = 1 : v2 = (5, 5, 3)
Moreover λ2 has the algebraic multiplicity 2, but the corresponding eigenvector space has
the geometric multiplicity 1. Because n = 3 we need 3 linearly independent solutions to
construct the general solution, as seen in 17.9. The eigenvalues are considered separately:
1) The first eigenvalue, λ1 = −4, has both geometric and algebraic multiplicity equal to 1.
This yields exactly one solution
10
u1 (t) = eλ1 t v1 = e−4t 5 (17-50)
1
Therefore we can use method 17.7 in order to find two solutions. First b is determined:
−10 10 0
5
( A − λ 2 E ) b = v2 ⇒ − 3 0 5 b = 5 (17-51)
1 −4 5 3
A particular solution to this system of equations is b = (0, 12 , 1). With this knowledge we
have two additional linearly independent solutions to the system of differential equations:
5
λ2 t t
u2 (t) = e v2 = e 5
3
(17-52)
5 0
u3 (t) = teλ2 t v2 + eλ2 t b = tet 5 + et 21
3 1
We leave it to the reader to show that all three solutions are linearly independent.
According to Method 17.9 the general real solution consists of the following linear combina-
tion for all c1 , c2 , c3 ∈ R:
x(t) = c1 u1 (t) + c2 u2 (t) + c3 u3 (t) (17-53)
Thus this yields
10 5 5 0
x(t) = c1 e−4t 5 + c2 et 5 + c3 tet 5 + et 12 (17-54)
1 3 3 1
According to the Structural Theorem 17.9 the general solution to a system of differential
equations with n equations contains n arbitrary constants. If we have n initial condi-
tions, then the constants can be determined, and we then get a unique solution. This is
formulated in the following existence and uniqueness theorem.
eNote 17 17.3 EXISTENCE AND UNIQUENESS OF SOLUTIONS 447
Theorem 17.11
A first order system of differential equations consisting of n equations in n unknown
functions with constant coefficients is given by
For every t0 ∈ I and every number set y0 = (y1 , y2 , . . . , yn ) exactly one solution
exists x(t) = ( x1 (t), x2 (t) . . . , xn (t) ) satisfying the initial conditions
x(t0 ) = y0 , (17-56)
that is
x1 ( t0 ) = y1 , x2 ( t0 ) = y2 , . . . , x n ( t0 ) = y n . (17-57)
Example 17.12
In Example 17.3 we found the general solution to the system of differential equations
0 1 2
x (t) = x(t), t ∈ R, (17-58)
3 0
viz.
x1 ( t ) 3t 1 −2t 2
x( t ) = = c1 e + c2 e , t∈R (17-59)
x2 ( t ) 1 −3
Now we wish to determine the unique solution x(t) = ( x1 (t), x2 (t)) that satisfies the initial
condition x(0) = ( x1 (0), x2 (0)) = (6, 6). This yields the system of equations
6 0 1 0 2 1 2 c1
= c1 e + c2 e = (17-60)
6 1 −3 1 −3 c2
By ordinary Gauss-Jordan elimination we get
1 2 6 1 2 6 1 0 6
→ → (17-61)
1 −3 6 0 −5 0 0 1 0
Thus we obtain the solution (c1 , c2 ) = (6, 0), and the unique conditional solution is therefore
3t 1
x(t) = 6e , t ∈ R, (17-62)
1
which is equivalent to
x1 (t) = 6e3t x2 (t) = 6e3t . (17-63)
In this particular case the two functions are identical.
eNote 17 17.4 TRANSFORMATION OF NTH ORDER DIFFERENTIAL EQUATIONS 448
Method 17.13
An nth order linear differential equation
for t ∈ R, can be transformed into a first order system of differential equations and
the system will look like this:
0
x1 ( t ) 0 1 0 ··· 0 x1 ( t )
x 0 (t) 0 1 ···
0 0 x2 (t)
2
.. . .. .. .. .. ..
= .. . (17-65)
0 . . . .
.
x n −1 ( t ) 0 0 0 ··· 1 xn−1 (t)
0
xn (t) − a 0 − a 1 − a 2 · · · − a n −1 xn (t)
The proof of this rewriting is simple but gives a good understanding of the transforma-
tion.
Proof
These new expressions are substituted into the differential equation (17-64):
xn0 (t) + an−1 xn (t) + an−2 xn−1 (t) + . . . + a1 x2 (t) + a0 x1 (t) = 0 (17-67)
Now this equation can together with equations (17-66) be written in matrix form.
Example 17.14
We wish to determine the general solution. Therefore the following functions are introduced
x1 ( t ) = x ( t )
x2 (t) = x10 (t) = x 0 (t) (17-70)
x3 (t) = x20 (t) = x 00 (t)
And we can then gather the last three equations in a system of equations.
x10 (t) = x2 ( t )
x20 (t) = x3 ( t ) (17-72)
0
x3 (t) = −10x1 (t) + 7x2 (t) + 4x3 (t)
But we need only the solution of x1 (t) = x (t), and we isolate this from the general solution
to the system. Furthermore we introduce three new arbitrary constants k1 , k2 , k3 ∈ R, that are
equal to the product of the c’s and the first coordinates of the eigenvectors. The result is
This constitutes the general solution to the differential equation (17-69). If the first coordinate
in v1 is 0 , we put k1 = 0 ; otherwise k1 can be an arbitrary real number. Similarly for k2 and
k3 .
eNote 18 451
eNote 18
Following eNotes 16 and 17 about differential equations, we now present this eNote about
second-order differential equations. Parts of the proofs closely follow the preceding notes and a
knowledge of these notes is therefore a prerequisite. In addition, complex numbers are used.
Linear second-order differential equations with constant coefficients look like this:
a0 , a1 ∈ R are constant coefficients of x (t) and x 0 (t), respectively. The right hand side
q(t) is a continuous real function, with the domain being an interval I (which could be
all of R ). The equation is called homogeneous if q(t) = 0 for all t ∈ I and otherwise
inhomogeneous.
The left hand side is linear in x, i.e., the map f : C ∞ (R) → C ∞ (R) given by
satisfies the linearity requirements L1 and L2 . The method used in this eNote for solving
the inhomogeneous equation exploits this linearity.
eNote 18 18.1 THE HOMOGENEOUS EQUATION 452
2. The general solution set Linhom for an inhomogeneous linear second-order dif-
ferential equation
(a) First the general solution Lhom to the corresponding homogeneous equation is
determined. This is produced by setting q(t) = 0 in (18-4).
(b) Then a particular solution x0 (t) to (18-4) is determined e.g. by guessing.
Concerning this see section 18.2.
where a0 and a1 are real constants. We wish to determine the general solution. This can
be accomplished using exact formulas that depend on the appearance of the equation.
eNote 18 18.1 THE HOMOGENEOUS EQUATION 453
λ2 + a1 λ + a0 = 0. (18-8)
The type of roots to this equation determines how the general solution Lhom to the
homogeneous differential equation will appear.
• Two complex roots λ = α ± βi, with Im(λ) = ± β 6= 0, yield the real solution
In all three cases the respective functions for all c1 , c2 ∈ R constitute the general
solution Lhom .
In Section 17.4 you find the theory for rewriting this type of differential equa-
tion as a system of first-order differential [Link] method works here.
The system will then look like this:
0
x1 ( t ) 0 1 x1 ( t )
= (18-12)
x20 (t) − a0 − a1 x2 ( t )
where x1 (t) = x (t) and x2 (t) = x10 (t) = x 0 (t). The problem can now be solved
using the theory and methods outlined in that section.
eNote 18 18.1 THE HOMOGENEOUS EQUATION 454
Proof
where x1 (t) = x (t) is the wanted solution that constitutes the general solution. The proof
begins with the theorems and methods in Section 17.1. For the proof we need the eigenvalues
of the system matrix A:
−λ 1
det(A − λE) = = λ2 + a1 λ + a0 = 0, (18-14)
− a0 − a1 − λ
which is also the characteristic equation for the differential equation. The type of roots of this
equation determines the solution x (t) = x1 (t), which gives the following three parts of the
proof:
First part
The characteristic equation has two different real roots: λ1 and λ2 . By using Method 17.4 we
obtain two linearly independent solutions u1 (t) = v1 eλ1 t and u2 (t) = v2 eλ2 t , where v1 and v2
are eigenvectors corresponding to the two eigenvalues , respectively. The general solution is
then spanned by:
x(t) = k1 u1 (t) + k2 u2 (t) = k1 eλ1 t v1 + k2 eλ2 t v2 , (18-15)
for all k1 , k2 ∈ R. The first coordinate x1 (t) = x (t) is the solution wanted:
which for all the arbitrary constants c1 , c2 ∈ R constitutes the general solution. c1 and c2
are two new arbitrary constants and they are the products of the k-constants and the first
coordinates of the eigenvectors: c1 = k1 v11 and c2 = k2 v21 .
Second part
The characteristic equation has the complex pair of roots λ = α + βi and λ̄ = α − βi. It is
possible to find the general solution using Method 17.5.
v is an eigenvector corresponding to λ and k1 and k2 are arbitrary constants. The first coordi-
nate x1 (t) = x (t) is the wanted solution, and is according to the above given by
For all c1 , c2 ∈ R, x (t) constitutes the general solution. c1 and c2 are two new arbitrary
constants given by c1 = k1 Re(v1 ) + k2 Im(v1 ) and c2 = −k1 Im(v1 ) + k2 Re(v1 ). v1 is the first
coordinate of v.
Third part
The characteristic equation has the double root λ. Because of the appearance of the system
matrix (the matrix is equivalent to an upper triangular matrix) it is possible to see that the
geometric multiplicity of the corresponding eigenvector space is 1, and it is then possible to
use Method 17.7 to find the general solution.
x(t) = k1 u1 (t) + k2 u2 (t) = k1 eλt v + k2 (teλt v + eλt b) = eλt (k1 v + k2 b) + k2 teλt v, (18-19)
which for all c1 , c2 ∈ R constitutes the general solution. c1 d and c2 are two new arbitrary
constants, given by c1 = k1 v1 + k2 b1 and c2 = k2 v1 , in which v1 is the first coordinate in v, as
b1 is the first coordinate in b.
All the three different cases of roots of the characteristic equation have now been treated thus
proving the theorem.
Dividing this equation by eλt , which is non-zero for all values of t, yields the charac-
teristic equation.
The characteristic equation has the roots λ1 = −5 and λ = 4, since −5 · 4 = −20 and −(−5 +
4) = 1 are the coefficients of the characteristic equation. Therefore the general solution to the
homogeneous equation is
We wish to determine Lhom , the general solution to the homogeneous equation. The charac-
teristic equation is
λ2 − 8λ + 16 = 0 ⇔ (λ − 4)2 = 0 (18-26)
Thus we have the double root λ = 4, and the general solutions set is composed of the follow-
ing function for all c1 , c2 ∈ R:
As can be seen from the two preceding examples it is relatively simple to determine
the solution to the homogeneous equation. In addition it is possible to determine the
differential equation from the solution, that is "go backwards". This is illustrated in the
following example.
Since the solution only includes terms with arbitrary constants, the equation must be homo-
geneous. Furthermore it is seen that the solution structure is similar to the solution structure
eNote 18 18.2 THE INHOMOGENEOUS EQUATION 457
in equation (18-10) in Theorem 18.2. This means that the characteristic equation of the second-
order differential equation has two complex roots: λ = 2 ± 7i. The characteristic equation
given these roots reads:
Directly from coefficients of the characteristic equation we can write the differential equation
as:
x 00 (t) − 4x 0 (t) + 53x (t) = 0, t ∈ R. (18-30)
This can also be seen from Theorem 18.2.
We wish to find a particular solution, because it is part of the general solution Linhom
together with the general solution Lhom to the corresponding homogeneous equation cf.
Method 18.1.
In this eNote we do not use a specific solution formula. Instead we use different meth-
ods depending on the form of q(t). In general one might guess that a particular solu-
tion x0 (t) has a form that somewhat resembles q(t), as will appear from the following
methods. Notice that these methods cover some frequently occurring forms of q(t), but
certainly not all.
Finally we will introduce the complex guess method. The complex guess method can be
eNote 18 18.2 THE INHOMOGENEOUS EQUATION 458
used if the right hand side q(t) of the equation is the real part of a simple complex ex-
pression, e.g. q(t) = et sin(3t) that is the real part of −ie(1+3i)t . Solving an equation with
a simple right hand side is easier, and therefore the corresponding complex equation is
solved instead. The solutions to the real equation and to the corresponding complex
equation are closely related.
We wish to determine a particular solution x0 (t) to the inhomogeneous equation. Since the
right hand side is a second degree polynomial we insert an unknown polynomial of second
degree in the left-hand side of the equation and equate this with the right-hand side:
x0 (t) = b2 t2 + b1 t + b0 , t ∈ R. (18-34)
The coefficients are determined by substituting the expression into the differential equation
eNote 18 18.2 THE INHOMOGENEOUS EQUATION 459
From the first equation it is evident that b2 = 2, and by substituting this in the second equa-
tion we get b1 = −4. Finally the last equation yields b0 = 9. Therefore a particular solution
to Equation (18-33) is given by
Given the following differential equation where the right-hand side is a first degree polyno-
mial:
x 00 (t) = t + 1 , t ∈ R. (18-37)
Show that you have to go to the third degree in order to find a polynomial that is a particular
solution to the equation.
where A and B are determined by substitution of the expression for x0 (t) as a solu-
tion into the inhomogeneous equation.
We wish to determine a particular solution x0 (t). By the use of Method 18.9 a particular
solution is
x0 (t) = A sin(ωt) + B cos(ωt) = A sin(3t) + B cos(3t). (18-41)
In addition we have
x00 (t) = 3A cos(3t) − 3B sin(3t)
(18-42)
x000 (t) = −9A sin(3t) − 9B cos(3t)
This is substituted into the equation
From this we get that A = 2, and a particular solution to the differential equation is then
Note that the number ω = 3 is the same in the arguments of both cosine
and sine in Example 18.10, and this is the only case that Method 18.9 facili-
tates. If two different numbers are present Method 18.9 does not apply, e.g.
q(t) = 3 sin(t) + cos(10t). But either superposition or the complex guess method
can be applied, and they will be described in section 18.2.2 and section 18.2.3,
respectively.
eNote 18 18.2 THE INHOMOGENEOUS EQUATION 461
where γ is determined by substituting the expression for x0 (t) as a solution into the
inhomogeneous equation. We emphasize that α must not be a root of the character-
istic equation for the differential equation.
As commented by the end of Method 18.11 the exponent α must not be a root of
the characteristic equation. If this is the case the guess will be a solution to the
corresponding homogeneous equation c.f. Theorem 18.2. This is a “problem”
for all orders of differential equations.
Thus we have succeeded in finding γ, and therefore we have a particular solution to the
differential equation:
x0 (t) = 4e−t , t ∈ R. (18-50)
eNote 18 18.2 THE INHOMOGENEOUS EQUATION 462
where q(t) = βeλt , β ∈ R and λ is a root in the characteristic equation of the differ-
ential equation, has the following form:
Consequently we use Method 18.13, and we guess a solution of the form x0 (t) = γteλt =
γte2t . We then have
x00 (t) = γe2t + 2γte2t
(18-56)
x000 (t) = 2γe2t + 2γe2t + 4γte2t = 4γe2t + 4γte2t
This is substituted into the equation in order to determine γ.
4γe2t + 4γte2t − 7(γe2t + 2γte2t ) + 10γte2t = −3e2t ⇔
(4γ − 14γ + 10γ)t + (4γ − 7γ + 3) = 0 ⇔ (18-57)
γ=1
eNote 18 18.2 THE INHOMOGENEOUS EQUATION 463
We have now succeeded in finding γ, and therefore a particular solution to the equation is
18.2.2 Superposition
Within all types of linear equations the concept of superposition exists. We present the
concept here for second-order linear differential equations with constant coefficients.
Superposition is here used in order to determine a particular solution to the inhomo-
geneous equation, when the right hand side (q(t)) is a combination (addition) of more
types of functions, e.g. a sine function added to a polynomial.
is a particular solution to
Proof
Superposition is a consequence of the differential equation being linear. We will here give a
general proof for all types of linear differential equations.
The left hand side of a differential equation is called f ( x (t)). We now posit n differential
eNote 18 18.2 THE INHOMOGENEOUS EQUATION 464
equations:
f ( x01 (t)) = q1 (t), f ( x02 (t)) = q2 (t), ..., f ( x0n (t)) = qn (t) (18-62)
where x01 , x02 , . . . , x0n are particular solutions to the respective inhomogeneous differential
equations. Define x0 = x01 + x02 + . . . + x0n and substitute this into the left hand side:
On the right hand side we get the sum of the functions q1 , q2 , . . . , qn , which sum we call q.
The Theorem is thus proven.
Now we treat Equation (18-66), where a particular solution is a polynomial of at the most
first degree, cf. Method 18.6, thus x02 (t) = b1 t + b0 . Hence x00 2 (t) = b1 and x0002 (t) = 0. This is
substituted into the differential equation.
0 − b1 − 3(b1 t + b0 ) = 3t − 14 ⇔ (−3b1 − 3)t + (−b1 − 3b0 + 14) = 0 (18-68)
Thus we have two equations in two unknowns, and we find that b1 = −1, and therefore that
b0 = 5. Thus a particular solution is x02 (t) = −t + 5. The general solution to (18-64) is then
found as the sum of the already found particular solutions to the two split equations:
x0 (t) = x01 (t) + x02 (t) = e4t − t + 5, t ∈ R. (18-69)
eNote 18 18.2 THE INHOMOGENEOUS EQUATION 465
The complex guess method is used when it is easy to rewrite the right hand side of the
differential equation as a complex expression, such that the given real right hand side is
the real part of the complex.
If e.g. the original right hand side is 2e2t cos(3t) , adding i (−2e2t sin(3t)) , we get
Here it is evident that Re(2e(2−3i)t ) = 2e2t cos(3t) . One then finds a complex particular
solution with complex right hand side. The wanted real particular solution to the origi-
nal equation is then the real part of the found complex solution.
Note that this method can be used because the equation is linear. It is exactly the lin-
earity that secures that the real part of the complex solution found is the wanted real
solution. This is shown by interpreting the left hand side of the equation as linear map
f (z(t)) in the set of complex functions of one real variable and using the following gen-
eral theorem:
Theorem 18.17
Given a linear map f : (C ∞ (R), C) → (C ∞ (R), C) and the equation
If we state z(t) and s(t) in rectangular form as z(t) = x (t) + i · y(t) and s(t) =
q(t) + i · r (t) , then (18-71) is true and if and only if
Proof
Given the function z(t) and letting the linear map f and the functions z(t) and s(t) be given
as in Theorem 18.17. As a consequence of the qualities of a linear map, cf. Definition ??, the
following applies:
f (z(t)) = s(t) ⇔
f ( x (t) + i · y(t)) = q(t) + i · r (t) ⇔
(18-73)
f ( x (t)) + i · f (y(t)) = q(t) + i · r (t) ⇔
f ( x (t)) = q(t) and f (y(t)) = r (t) .
The complex particular solution has the form z0 (t) = (c + di )e(α+ωi)t , where c and d
are determined by substitution of z0 (t) into Equation (18-76).
A decisive reason for using the complex guess method is that it is so easy to
determine the derivative of the exponential function, even when it is complex.
We wish to determine a particular solution. It is evident that we can use the complex guess
method in Method 18.18. Initially the following is true for the right hand side:
q(t) = 19e4t cos(t) − 35e4t sin(t) = Re (19 + 35i )e(4+i)t . (18-79)
We shall now instead of the original problem find a complex particular solution to the differ-
ential equation
z00 (t) − 2z0 (t) − 2z(t) = (19 + 35i )e(4+i)t , t ∈ R. (18-80)
by guessing that z0 (t) = (c + di )e(4+i)t is a solution. We also have
These expressions are substituted into the complex equation in order to determine c and d:
These are two equations in two unknowns. The augmented matrix of the system of equations
is written:
1 − 65 19
5 −6 19 5 1 0 5
→ 61 → . (18-83)
6 5 35 0 61 5 5 0 1 1
thus we have that c = 5 and d = 1, which yields z0 (t) = (5 + i )e(4+i)t . Therefore a particular
solution to the equation (18-78) is
x0 (t) = Re(z0 (t)) = Re (5 + i )e(4+i)t = 5e4t cos(t) − e4t sin(t), t ∈ R. (18-84)
eNote 18 18.3 EXISTENCE AND UNIQUENESS 468
Here we formulate a theorem about existence and uniqueness for differential equations
of the second order with constant coefficients. We need two initial value conditions: The
value of the function and its first derivative at the chosen initial point.
such that
x ( t0 ) = x0 and x 0 ( t0 ) = v0 , (18-86)
where t0 ∈ I, x0 ∈ R and v0 ∈ R.
6 − 45
6 = −4c1 + 9(5 − c1 ) = −13c1 + 45 ⇔ c1 = =3 (18-92)
−13
Therefore c2 = 5 − 3 = 2 and the conditional solution is
Note that one can determine a unique and conditional solution to a homogeneous
differential equation, as in this case. The right hand side needs not be different from
zero. The general solution for the equation is Linhom = Lhom , since x0 (t) = 0.
Below is an example going through the whole solution procedure for an inhomogeneous
equation with a double initial value condition. After that an example is presented where
the purpose is to find the differential equation given the general solution. It is analogous
to example 18.5, but now we have a right hand side different from zero.
First we solve the corresponding homogeneous equation, and the characteristic equation
looks like this:
λ2 + 6λ + 5 = 0 (18-95)
This has the roots λ1 = −5 and λ2 = −1, since (λ + 5)(λ + 1) = λ2 + 6λ + 5. Because these
roots are real and different, cf. Theorem 18.2, the general homogeneous solution set is given
by
Lhom = c1 e−5t + c2 e−t , t ∈ R c1 , c2 ∈ R .
(18-96)
Now we determine a particular solution to the inhomogeneous equation. Since the right hand
side is a second degree polynomial we guess that x0 (t) = b2 t2 + b1 t + b0 , using Method 18.6.
We then have that x00 (t) = 2b2 t + b1 and x000 (t) = 2b2 . This is substituted into the differential
equation.
2b2 + 6(2b2 t + b1 ) + 5(b2 t2 + b1 t + b0 ) = 20t2 + 48t + 13 ⇔
(5b2 − 20)t2 + (12b2 + 5b1 − 48)t + (2b2 + 6b1 + 5b0 − 13) = 0 ⇔ (18-97)
5b2 − 20 = 0 og 12b2 + 5b1 − 48 = 0 og 2b2 + 6b1 + 5b0 − 13 = 0.
eNote 18 18.3 EXISTENCE AND UNIQUENESS 470
The first equation easily yields b2 = 4. If this is substituted into the second equation we get
b1 = 0. Finally in the third equation we get b0 = 1. A particular solution to the inhomoge-
neous equation is therefore
x0 (t) = 4t2 + 1, t ∈ R. (18-98)
Following the structural theorem, e.g. Method 18.1, the general solution to the inhomoge-
neous equation is given by
We now determine the solution that satisfies the given initial value conditions. An arbitrary
solution has the form
5= c1 + c2 + 1
(18-102)
−8 = −5c1 − c2
Given the general solution to a linear second-order differential equation with constant coef-
ficients:
Linhom = c1 e−2t + c2 e2t − 21 sin(2t) , t ∈ R c1 , c2 ∈ R
(18-105)
It is now the aim to find the differential equation, which in general looks like this:
First we split the solution into a particular solution and the general homogeneous solution
set:
Now we consider the general homogeneous solution. The looks of this complies with the first
case of in Theorem 18.2. The characteristic equation has two real roots and they are λ1 = −2
and λ2 = 2. Therefore the characteristic equation is
(λ + 2)(λ − 2) = λ2 − 4 = 0 (18-108)
This determines the coefficients on the left hand side of the differential equation: a1 = 0 and
a0 = −4. The differential equation so far looks like this:
Since x0 (t) is a particular solution to the inhomogeneous equation the right hand side q(t)
can be determined by substituting x0 (t). We have that x000 (t) = 2 sin(2t).
18.4 Summary
In this note linear second-order differential equations with constant coefficients are writ-
ten as:
x 00 (t) + a1 x 0 (t) + a0 x (t) = q(t) (18-112)
• This equation is solved by first determining the general solution to the correspond-
ing homogeneous equation and then adding this to a particular solution to the
inhomogeneous equation, see Method 18.1.
• In particular we have the complex guess method for the determination of the partic-
ular solution x0 (t). The complex guess method can be used when the right hand
side has this appearance:
q(t) = Re ( a + bi )e(α+ωi)t = aeαt cos(ωt) − beαt sin(ωt). (18-114)
The solution is then determined by rewriting the differential equation in the cor-
responding exponential form, see Method 18.18.
Index
vector, 251
vector space, 251
vector space over the complex numbers, 251
vector space over the real numbersl, 251
vector space with inner product, 364