Block 5
Block 5
LINEAR ALGEBRA
Indira Gandhi National Open University
School of Sciences
Block
5
INNER PRODUCTS AND QUADRATIC FORMS
UNIT 14 5
Inner Product Spaces
UNIT 15 23
Hermitian and Unitary Operators
UNIT 16 47
Real Quadratic Forms
UNIT 17 74
Conics
Miscellaneous Exercises 97
Curriculum Design Committee
Prof. J.N. Kapur (Chairmain) Prof. M.S. Narasimhan
Jawaharlal Nehru University Tata Institute of Fundamental Research
New Delhi Bombay
Prof. Izhar Husain Prof. (Mrs.) A.R. Singal
Dept. of Mathematics Dept. of Mathematics
Aligarh Muslim University Meerut University
Aligarh
Prof. D.N. Misra Prof. M.P. Singh
Council of Scientific & Industrial Research I.I.T., Delhi
New Delhi
Disclaimer – Any materials adapted from web-based resources in this course are being used for
educational purposes only and not for commercial purpose.
April, 2022
© Indira Gandhi National Open University, 2021
ISBN-
All right reserved. No part of this work may be reproduced in any form, by mimeograph or any other
means, without permission in writing from the Indira Gandhi National Open University.
Further information on the Indira Gandhi National Open University courses, may be obtained from the
University’s office at Maidan Garhi, New Delhi-110 068 and IGNOU website www.ignou.ac.in.
Printed and published on behalf of the Indira Gandhi National Open University, New Delhi by
Prof. Sujatha Varma, Director, School of Sciences.
2
BLOCK 5 INNER PRODUCTS AND QUADRATIC
FORMS
This is the last block of this course. It deals with the interesting properties of a special
class of vector spaces which are known as inner product spaces. In this block the
only vector spaces we consider will be over R or C.
In the first unit (Unit 14) we introduce the basic notion of the inner product of two
vectors, along with its properties. This product helps us in introducing the well-known
geometrical notions of lengths and angles between vectors. We go on to discuss the
concept of orthogonality and the solution of the basic problem of the existence of an
orthonormal basis in a finite-dimensional inner product space.
The second unit (Unit 15) deals with the problem of characterising linear functionals in
inner product spaces. We show that such functionals are represented as inner
products. This further helps us in proving the existence of a unique adjoint for every
given operator. Some interesting relations between an operator and its adjoint lead us
to define self-adjoint and unitary operators. We also establish some of their properties.
Then we introduce you to Hermitian, unitary and orthogonal matrices and the concept
of orthogonal similarity.
In Units 16 and 17 we deal with real vector spaces only. The purpose of these two units
is to use the methods of linear algebra that you have studied in the course so far, to
2
reduce quadratic forms in and 3 into simpler forms. In Unit 17 you will study
various conics in detail.
What you study in these units will stand you in good stead in various mathematics
courses, particularly in geometry and mechanics.
In case you are interested in knowing more about the material covered in this block,
you may look up the books that we have given in the course introduction.
3
UNIT 14
14.1 Introduction 5
Objective
14.2 Inner Product 6
14.3 Norm of a Vector 10
14.4 Orthogonality 13
14.5 Summary 19
14.6 Solution/Answers 19
14.1 INTRODUCTION
So far you have studied many interesting vector spaces over various field. In
this unit, and the following onces, we will only consider real and complex
vector spaces. In Unit 2 you studied geometrical notions like the length of a
vector, the angle between two vectors and the dot product in ℝ2 or ℝ3 . In this
unit we carry these concepts over to a more general setting. We will define a
certain special class of vector spaces which open up new and interesting vistas
for investigations in mathematics and physics. Hence their study is extremely
fruitful as far as the applications of the theory to problems are concerned. This
fact will become clear in Units 16 and 17.
Before going further we suggest that you refer to Unit 2 for the definitions and
properties of the length and the scalar product of vectors of ℝ2 or ℝ3 .
Objectives
After studying this unit, you should be able to:
• define and give examples of inner product spaces;
(x1 , x2 , x3 ) ⋅ (y1 , y2 , y3 ) = x1 y1 + x2 y2 + x3 y3 .
We also remind you that given any complex number z = a + ib, where a, b ∈ ℝ, its
complex conjugate is z = a − ib.
IP1) ⟨x, x⟩ ≥ 0 ∀ x ∈ V.
IP5) ⟨y, x⟩ = ⟨x, y⟩ for all x, y ∈ V. (Here ⟨x, y⟩ denotes the complex conjugate of
the number ⟨x, y⟩.)
The scalar ⟨x, y⟩ is called inner product (or scalar product) of the vector x
with the vector y.
A vector space V over which an inner product has been defined is called an
inner product space, and is denoted by (V, ⟨ , ⟩).
6 i) IP1 is satisfied because ⟨u, u⟩ = x21 + x22 + x23 , which is always non-negative.
Unit
. . . . .14
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Inner
. . . . . . .Product
. . . . . . . . .Spaces
.......
ii) Now, ⟨u, u⟩ = 0 ⇒ x21 + x22 + x23 = 0 ⇒ x1 = 0, x2 = 0, x3 = 0 since the sum of
positive real number is zero if and only if each of them is zero. ∴, u = 0.
Also, if u = 0, then x1 = 0 = x2 = x3 . ∴⟨u, u⟩ = 0.
So, we have shown that IP2 is satisfied by ⟨ , ⟩.
We suggest that you verify IP4 and IP5. That’s what E1 says!
Example 2: What is the value of the standard inner product of u = (5, −1, 2) and
v = (−1, 0, 1)?
∗∗∗
E1) Check that the inner product on ℝ3 in Example 1 satisfies IP4 and IP5.
The inner product that have been given in Example 1 can be generalised to the
inner product ⟨ , ⟩ on ℝn , defined by ⟨(x1 , … , xn ) , (y1 , … , yn )⟩ = x1 y1 + x2 y2 + ⋯ + xn yn .
This is called the standard inner product on ℝn .
Example 3: Take 𝔽 = ℂ and, for x, y ∈ ℂ, define ⟨x, y⟩ = xy. Show that the map
⟨ , ⟩ ∶ ℂ × ℂ → ℂ is an inner product.
Solution: IP1 and IP2 are satisfied because, for any complex number
x, xx ≥ 0. Also, xx = 0 if and only if x = 0.
∗∗∗
E2) Show that IP4 and IP5 are true for Example 3.
In fact, Example 2 can be generalised to ℂn , for any n > 0. We can define the
inner product of two arbitrary vectors x = (x1 , … , xn ) and y = (y1 , … , yn ) ∈ ℂn by
n
⟨x, y⟩ = ∑ xi yi . This inner product is called the standard inner product on ℂn .
i=1 7
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
The next example deals with a general complex vector space.
where ai , bi , ci ∈ ℂ ∀ i = 1, … , n. Then
n
⟨x, x⟩ = ∑ ai ai ≥ 0. Also, ⟨x, x⟩ = 0 ⇔ ai = 0 ∀ i = 1, 2, … , n ⇔ x = 0
i=1
n n n
⟨ x + y, z⟩ = ∑ (ai + bi ) ci = ∑ ai ci + ∑ bi ci = ⟨x, z⟩ + ⟨y, z⟩
i=1 i=1 i=1
n n
ab = a b ∀ a, b ∈ ℂ. Also, for any 𝛼 ∈ ℂ, ⟨𝛼x, y⟩ = ∑ 𝛼ai bi = 𝛼 ∑ ai bi = 𝛼⟨x, y⟩
i=1 i=1
n n n
Finally, ⟨y, x⟩ = ∑ bi ai = ∑ bi ai = ∑ ai bi = ⟨x, y⟩
i=1 i=1 i=1
E4) Let X = {x1 , … , xn } be a set and V be the set of all functions from X to C.
Then, with respect to pointwise addition and scalar multiplication, V is a
vector space over ℂ. Now, for any f, g ∈ V, define
n
⟨f, g⟩ = ∑ f(xi ) g(xi ).
i=1
Example 5: Show that C[a, b], where ‘C[a, b]’ denotes the set of all continuous
real-valued functions defined on the closed interval [a, b], is an inner product
8 space.
Unit
. . . . .14
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Inner
. . . . . . .Product
. . . . . . . . .Spaces
.......
Solution: We need to define an inner product on c[a, b]. For this, we define a
simple inner product
b
⟨f1 , f2 ⟩ = ∫ f1 (x) f2 (x) dx ∀ f1 , f2 ∈ c[a, b].
a
b
i) IP1 is satisfied because ⟨f1 , f1 ⟩ = ∫ f21 (x) dx, which is always non-negative.
a
b
ii) Now ⟨f1 , f1 ⟩ = 0 ⇒ ∫ f21 (x) dx = 0, which is possible if and only if the
a
function f is identically zero.
So, we shown that IP2 is satisfied by ⟨ , ⟩.
b
⟨ f1 + f2 , f3 ⟩ = ∫ (f1 (x) + f2 (x)) f3 (x) dx
a
b b
= ∫ f1 (x) f3 (x) dx + ∫ f2 (x) f3 (x) dx
a a
= ⟨f1 , f3 ⟩ + ⟨f2 , f3 ⟩.
∗∗∗
We suggest that you verify IP4 and IP5. That’s what E6 says!
We now state some properties of inner products that immediately follow from
IP1-IP5.
Theorem 1: Let (V, ⟨ , ⟩) be an inner product space. Then, for any x, y, z ∈ V and
𝛼, 𝜇 ∈ ℂ,
a) ⟨𝛼x + 𝜇y, z⟩ = 𝛼⟨x, z⟩ + 𝜇⟨y, z⟩
c) ⟨0, x⟩ = ⟨x, 0⟩ = 0
d) ⟨x − y, z⟩ = ⟨x, z⟩ − ⟨y, z⟩
e) ⟨x, z⟩ = ⟨y, z⟩ ∀ z ∈ V ⇒ x = y.
Proof: We will prove (a) and (c), and leave the rest to you.
a) ⟨𝛼x + 𝜇y, z⟩ = ⟨𝛼x, z⟩ + ⟨𝜇y, z⟩ (by IP3)
= 𝛼⟨x, z⟩ + 𝜇⟨y, z⟩ (by IP4)
b) The vector 0 ∈ V can be written as 0 = 0 ⋅ y for some y ∈ V.
Thus, ⟨0, x⟩ = ⟨0 ⋅ y, x⟩ = 0⟨y, x⟩ = 0.
Then, ⟨x, 0⟩ = ⟨0, x⟩ = 0 = 0.
■
The proof of this theorem will be complete once you solve E7.
9
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
E6) Show that IP4 and IP5 are true for Example 5.
Definition 2: If (V, ⟨ , ⟩) is an inner product space and x ∈ V, then the norm (or
length) of the vector x is defined to be √⟨x, x⟩. It is denoted by ‖x‖.
Hence, ‖x‖ = √⟨x, x⟩.
1
1/2
c) ‖f‖ = ⟨f, f⟩ = √∫ x2 x2 dx
0
x5 1 1
= √[ ] =
5 0 √5
10 ∗∗∗
Unit
. . . . .14
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Inner
. . . . . . .Product
. . . . . . . . .Spaces
.......
Observe that in E8, x is any vector of V and V is also any vector space. So, ‖xx‖
is a special form of any non-zero vector. It also implies that we can create a unit
vector by using any non-zero vector. So, E8 leads us to the following definition.
We will now prove some results involving norms. The first one is the
Cauchy-Schwarz inequality. It is very simple, but very important because it
allows us to prove many other useful statements.
Now ‖x − 𝛼z‖2 ≥ 0. This means that ‖x‖2 − |⟨x, z⟩|2 + |⟨x, z⟩ − 𝛼|2 ≥ 0 ∀𝛼 ∈ F.
In particular, if we choose 𝛼 = ⟨ x, z⟩, we get 0 ≤ ‖x‖2 − |⟨x, z⟩|2 . Hence,
|⟨x, z⟩| ≤ ‖x‖, that is,
y 1
|⟨x, ⟩| ≤ ‖x‖ ⇔ |⟨ x, y⟩| ≤ ‖x‖ ⇔ |⟨x, y⟩| ≤ ‖x‖ ‖y‖.
‖y‖ ‖y‖
Example 7: Write the expression for the Cauchy-Schwarz inequality for the
vector space given in E4.
n
2
Solution: For any f ∈ V, ‖f‖2 = ⟨f, f⟩ = ∑|f(xi )| . Thus, Theorem 2 says that
i=1
n n n
2 2
2 vectors x and y are |∑ f(xi )g(xi )| ≤ √∑|f(xi )| √∑ |g(xi )| ∀ f, g ∈ V.
i=1 i=1 i=1
called proportional if
∃ 𝛼 ∈ F, 𝛼 ≠ 0, with x = 𝛼y.
∗∗∗
Do try these exercise now.
E10) Write down the expressions for the Cauchy-Schwarz inequality for the
spaces given in Example 1, 3 and 4.
If z = a + ib ∈ ℂ, then
b) ‖x + y‖2 + ‖x − y‖2 = 2 (‖x‖2 + ‖y‖2 ) (Parallelogram law)
a) the real part of z is a,
and is denoted by Re (z), Proof: a) Now
b) z + z = 2Re(z)
c) Re(z) ≤ |z| ‖x + y‖2 = ⟨x + y, x + y⟩ = ‖x‖2 + ⟨x, y⟩ + ⟨y, z⟩ + ‖y‖2 .
= ‖x‖2 + ⟨x, y⟩ + ⟨x, y⟩ + ‖y‖2 .
= ‖x‖2 + 2Re⟨x, y⟩ + ‖y‖2 .
≤ ‖x‖2 + 2|⟨x, y⟩| + ‖y‖2 , since Re ⟨x, y⟩ ≤ |⟨x, y⟩|.
≤ ‖x‖2 + 2‖x‖ ‖y‖ + ‖y‖2 (by Theorem 2)
= (‖x‖ + ‖y‖)2
Hence, ‖x + y‖2 ≤ (‖x‖ + ‖y‖)2 . Taking square roots of both sides we obtain
‖x + y‖ ≤ ‖x‖ + ‖y‖.
⟨x, x⟩ + ⟨x, y⟩ + ⟨y, x⟩ + ⟨y, y⟩ + ⟨x, x⟩ − ⟨x, y⟩ − ⟨y, x⟩ + ⟨y, y⟩ = 2 (‖x‖2 + ‖y‖2 )
Fig. 1: ‖x + y‖ ≤ ‖x‖ + ‖y‖
The reason (a) is called the triangle inequality is that for any triangle the sum of
12 the lengths of any sides is greater than or equal to the length of the third side.
Unit
. . . . .14
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Inner
. . . . . . .Product
. . . . . . . . .Spaces
.......
So, if we consider a triangle in any Euclidean space, two of whose sides are
the vectors x and y, then the third side is x + y (see Fig. 1,) and hence,
‖x‖ + ‖y‖ ≥ ‖x + y‖.
Similarly, (b) is called the parallelogram law because it generalises the fact that
the sum of the squares of the length of the diagonals of a parallelogram in
Euclidean space is always equal to the double of the sum of the squares of its
sides (Fig 2).
E12) Show that |‖x‖ − ‖y‖| ≤ ‖x − y‖ for x, y ∈ (V, ⟨ , ⟩) . Fig. 2: ‖x + y‖2 + ‖x − y‖2 =
(Hint: Use the triangle inequality for y and (x − y) .) 2 (‖x‖2 + ‖y‖2 )
14.4 ORTHOGONALITY
|⟨x,y⟩|
In Theorem 2 we showed that ‖x‖ ‖y‖
≤ 1 for any x, y ∈ V. We knew that, for
2 3 |⟨x,y⟩|
non-zero vectors x and y (in ℝ or ℝ ), ‖x‖ ‖y‖ is equal to the magnitude of the
cosine of the angle between them. We generalise this concept now.
|⟨x,y⟩|
For any inner product space V and for any non-zero x, y ∈ V, we take ‖x‖ ‖y‖ to
be the magnitude of the cosine of the angle between the two vectors x
and y.
E13) Using the definitions of inner product and orthogonality, prove the
following results for an inner product space V.
a) 0 ⟂ x ∀ x ∈ V.
b) ‘ x ⟂ x iff x = 0, where x ∈ V.
c) x ⟂ y ⇒ y ⟂ x, for x ∈ V.
d) x ⟂ y ⇒ 𝛼x ⟂ y for any 𝛼 ∈ F, where x, y ∈ V.
⟨e1 , e1 ⟩ = 1.1 + 0 + ⋯ + 0 = 1.
⟨e2 , e2 ⟩ = 0 + 1 + 0 + ⋯ + 0 = 1.
Similarly, ⟨ei , ei ⟩ = 1 ∀ i = 1, … , n.
Thus, ‖ei ‖ = 1 ∀ i = 1, … , n.
On the lines of Example 8, we can also show that the elements of the standard
basis of ℂn are mutually orthogonal and of unit length with respect to the
standard inner product.
∗∗∗
By definition, every orthonormal set is orthogonal. But the converse is not true,
as the following example tells us.
14 ∗∗∗
Unit
. . . . .14
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Inner
. . . . . . .Product
. . . . . . . . .Spaces
.......
Try the following exercise now.
E15) Let Pn be the real vector space of all polynomials of degree ≤ n. We define
an inner product on Pn by
n n n
⟨∑ ai xi , ∑ bi xi ⟩ = ∑ ai bi .
i=0 i=0 i=0
n
Proof: Let y = ∑ ai yi , where ai ∈ F ∀ i = 1, … , n.
i=1
n n
Then, y ∈ V and ⟨x, y⟩ = ⟨x, ∑ ai yi ⟩ = ∑ ai ⟨x, yi ⟩ = 0, because ⟨x, yi ⟩ = 0 ∀ i. This
i=1 i=1
shows that x ⟂ y. ■
n
Proof: Our hypothesis says that ⟨xi , xj ⟩ = 0 if i ≠ j. Consider y = ∑ ai xi .
i=1
n n n n n n
‖y‖2 = ⟨y, y⟩ = ⟨∑ ai xi , ∑ aj xj ⟩ = ∑ ∑⟨ai xi , aj xj ⟩ = ∑ ∑ ai aj ⟨xi , xj ⟩
i=1 j=1 i=1 j=1 i=1 j=1
n
= ∑ ai ai ⟨xi , xi ⟩ since ⟨xi , xj ⟩ = 0 for i ≠ j
i=1
n
2
= ∑|ai | ‖xi ‖2 . This proves the result.
i=1
Then
n
2 2
‖y‖2 = 0 ⇒ ∑|ai | ‖xi ‖2 = 0 ⇒ |ai | ‖xi ‖2 = 0 ∀ i.
i=1
2
⇒ |ai | = 0 for i = 1, … , n, since ‖xi ‖2 ≠ 0 for any i.
⇒ ai = 0 for i = 1, … , n.
Thus, {x1 , … , xn } is linearly independent. Hence, A is linearly independent. ■
We have just proved that any orthogonal set is linearly independent. Therefore,
any orthogonal set in a vector space V of dimension n must have a maximum
of n elements. So, for example, any orthogonal subset of ℝ3 can have 3
elements, at the most.
We shall use Theorem 6 as a stepping stone towards showing that any inner
product space has an orthonormal set as a basis. But first, some definitions
and remarks.
E16) Let {e1 , … , en } be an orthonormal basis for a real inner product space V.
n n n
Let x = ∑ xi ei and y = ∑ yi ei be elements of V. Show that ⟨x, y⟩ = ∑ xi yi .
i=1 i=1 i=1
v w 1 1 1 −1
{ , } = {( , )( , )}
‖v‖ ‖w‖ √2 √2 √2 √2
Proof: We shall first show that it has an orthogonal basis, and then obtain an
orthonormal basis.
Let {v1 , … , vn } be a basis of V. From this basis we shall obtain an orthogonal
basis {w1 , w2 , … , wn } of V in the following way.
⟨v2 ,w1 ⟩ ⟨v2 ,v1 ⟩
Take w1 = v1 . Define w2 = v2 − ⟨w1 ,w1 ⟩
w1 . Then w2 = v2 − v ,
⟨v1 ,v1 ⟩ 1
and
⟨v2 , v1 ⟩
⟨w2 , v1 ⟩ = ⟨v2 , v1 ⟩ − ⟨v1 , v1 ⟩ = 0.
⟨v1 , v1 ⟩
⟨v2 ,w1 ⟩
That is, ⟨w2 , w1 ⟩ = 0. Further, v2 = c1 v1 + w2 , where c1 = ⟨w1 ,w1 ⟩
∈ F.
This way we obtain an orthogonal set vectors {w1 , w2 , … , wn } , such that the v′i s
are a linear combination of the w′i s. By Theorem 6 this set is linearly
independent, and hence form a basis of V. ■
Example 10: Obtain an orthonormal basis for P2 , the space of all real
polynomials of degree at most 2, the inner product being defined by
1
⟨p1 , p2 ⟩ = ∫ p1 (t)p2 (t)dt.
0
1
⟨t,w1 ⟩ 1 t2 1 1
w2 = t − ⟨w1 ,w1
w . Now ⟨t, w1 ⟩ = ∫ t dt =
⟩ 1
| = . Therefore, w2 = t − 2
0 2 0 2
1 1 2 1
∴, ⟨w2 , w2 ⟩ = ∫ (t − ) dt = .
0 2 12 17
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
⟨t2 , w2 ⟩ ⟨t2 , w1 ⟩
w3 = t2 − w2 − w1
⟨w2 , w2 ⟩ ⟨w1 , w1 ⟩
1 1 1 1
= t2 − 12 { (t − )} − = t2 − t + .
2 12 3 6
1 1 2 1
Also ⟨w3 , w3 ⟩ = ∫ (t2 − t + ) dt = .
0 6 180
Thus, the orthonormal basis is
w1 w2 w3 1 1
{ , , } = {1, √12 (t − ) , √180 (t2 − t + )} .
‖w1 ‖ ‖w2 ‖ ‖w3 ‖ 2 6
∗∗∗
Here’s an exercise.
E17) Obtain an orthonormal basis with respect to the standard inner product,
for
a) the subspace of ℝ3 generated by (1, 0, 3) and (2, 1, 1)
b) the subspace of ℝ4 generated by (1, 0, 2, 0) and (1, 2, 3, 1).
n
Proof: Let x = ∑ ai xi (ai ∈ F) be any linear combination of the elements of A.
i=1
Then
Corollary 1: Let A = {x1 , … , xn } be any orthonormal set (V, ⟨ , ⟩) . Then, for any
y ∈ V,
n
2 2
∑|⟨y, xi ⟩ | ≤ ‖y‖ .
i=1
14.5 SUMMARY
In this unit we have discussed the following points. We have
14.6 SOLUTIONS/ANSWERS
∴ IP5 is satisfied.
E3) a) ⟨u, v⟩ = 2 − 2 = 0
b) ⟨u, v⟩ = 2 + 2 + 9 = 13
E4) Let f, g, h ∈ V and 𝛼 ∈ ℂ. Then
n n
2
⟨f, f⟩ = ∑ f(xi )f(xi ) = ∑|f(xi )| ≥ 0.
i=1 i=1
⟨f, f⟩ = 0 ⇔ f(xi ) = 0 ∀ i = 1, … , n
⇔ f is the zero function.
n
⟨f + g, h⟩ = ∑ (f + g) (xi )h(xi )
i=1
n
= ∑ {f(xi ) + g(xi )} h(xi )
i=1
n n
= ∑ f(xi )h(xi ) + ∑ g(xi )h(xi )
i=1 i=1
= ⟨f, h⟩ + ⟨g, h⟩.
n n
⟨𝛼f, g⟩ = ∑(𝛼f)(xi )g(xi ) = ∑ 𝛼f(xi )g(xi )
i=1 i=1
n
= 𝛼 ∑ f(xi )g(xi ) = 𝛼⟨f, g⟩
i=1
n n
⟨f, g⟩ = ∑ f(xi )g(xi ) = ∑ f(xi )g(xi )
i=1 i=1
n
= ∑ g(xi )f(xi ) = ⟨g, f⟩.
i=1
20 ∴ IP5 satisfied.
Unit
. . . . .14
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Inner
. . . . . . .Product
. . . . . . . . .Spaces
.......
e) ⟨x, z⟩ = ⟨y, z⟩ ∀ z ∈ V
⇒ ⟨x − y, z⟩ = 0 ∀ z ∈ V, by (d) above.
⇒ ⟨x − y, x − y⟩ = 0, taking z = x − y, in particular.
⇒ x − y = 0, by IP2.
⇒ x = y.
x
E8) Let u = ‖x‖
. Then ⟨u, u⟩ = ⟨ ‖xx‖ , ‖xx‖ ⟩ = 1
‖x‖2
⟨x, x⟩ = 1
‖x‖2
‖x‖2 = 1.
∴, ‖u‖ = √⟨u, u⟩ = 1.
= √1 + 1 + 12 − i2 = 2.
E10) In the situation of Example 1 we get
|u ⋅ v| ≤ ‖u‖ ‖v‖ for u, v ∈ ℝ3 . Thus,
2 2 2 2 2 2
|x1 y1 + x2 y2 + x3 y3 | ≤ √x1 + x2 + x3 √y1 + y2 + y3 ∀ (x1 , x2 , x3 ) , (y1 , y2 , y3 ) ∈ ℝ3 .
⇒‖x‖ ≤ ‖y‖ + ‖x − y‖
⇒‖x‖ − ‖y‖ ≤ ‖x − y‖
d) x ⟂ y ⇒⟨x, y⟩ = 0 ⇒ 𝛼⟨x, y⟩ = 0 ∀𝛼 ∈ F
⇒ ⟨𝛼x, y⟩ = 0 ∀𝛼 ∈ F ⇒ 𝛼x ⟂ y ∀ 𝛼 ∈ F.
E14) If x ⟂ y, then ⟨x, y⟩ = 0. Hence,
∴ ‖w2 ‖ = √ 94 + 1 + 1
4
= √ 72
∴ { 1
(1, 0, 3) , √ 27 ( 32 , 1, −21 )} is the required orthonormal basis.
√10
b) w1 = (1, 0, 2, 0)
7 2 1
w2 = (1, 2, 3, 1) − (1, 0, 2, 0) = (− , 2, , 1) ,
5 5 5
‖w1 ‖ = √5, ‖w2 ‖ = √ 26
5
w w
Then { ‖w1‖ , ‖w2‖ } is the required basis.
1 2
22
UNIT 15
15.1 INTRODUCTION
In the preceding unit we discussed general properties of linear product spaces.
In this unit we will show that we can precisely determine the nature of linear
functionals defined over inner product spaces.
We, then, discuss the adjoint of an operator. The behaviour of this adjoint leads
us to the concepts of self-adjoint operators and unitary operators. As usual, we
will discuss their matrix analogues also. This will entail studying the definitions
and properties of Hermitian, unitary and orthogonal matrices.
Regarding the notation in this unit, F will always denote R or C. And, unless
otherwise mentioned, the inner product of Rn or Cn will be the standard inner
product. Also, if T is a function acting on x, then we will often write Tx for T(x),
for our convenience.
Being reading this unit we advise you to look and recall the definitions of a
linear functional and a dual space. 23
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
Objectives
After studying this unit, you should be able to:
• prove and use the fact that a matrix is unitary iff its rows (or columns) form an
orthogonal set of vectors;
• use the fact that any real symmetric matrix is orthogonally similar to a
diagonal matrix.
f(v) = ⟨v, x⟩ ∀ v ∈ V.
Before going into the detailed study of such functions let us consider an
example.
x = (x1 , x2 ) ∈ ℝ2 , f ∶ ℝ2 → ℝ
Solution: Firstly,
Let us now consider any inner product space (V, ⟨ , ⟩) . We choose a vector
z ∈ V and fix it. With the help of this vector we can obtain a linear functional
f ∈ V∗ = L (V, F) in the following way:
is a linear functional on V.
Proof: As dim V = n, there exists a finite orthogonal basis for V. Let this basis
be B = {e1 , e2 , … , en } . Then
0, i ≠ j
⟨ei , ej ⟩ = {
1, i = j
Then
n n n
f(x) = f (∑ bi ei ) = ∑ bi f(ei ) = ∑ bi ai …(1)
i=1 i=1 i=1
n
Now consider the vector z ∈ V such that z = ∑ ai ei .
i=1 25
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
As each ai is known to us, z is a known vector of V. Also,
n n
⟨x, z⟩ = ⟨∑ bi ei , ∑ aj ej ⟩
i=1 j=1
n n
= ∑ ∑⟨bi ei , aj ej ⟩
i=1 j=1
n n
= ∑ ∑ bi aj ⟨ei , ej ⟩
i=1 j=1
n
= ∑ bi ai , since B is an orthonormal set.
i=1
= f(x), from (1) above.
Then,
f(x) = ⟨x, z⟩ ∀ x ∈ V.
Let us now use linear functionals to define the adjoint of a linear transformation
from V to V.
f(x) = ⟨Tx, y⟩ ∀ x ∈ V.
By E3 and Theorem 2, ∃ a unique element z ∈ V such that f = ⟨•, z⟩, that is,
f(x) = ⟨x, z⟩ ∀ x ∈ V, that is, ⟨Tx, y⟩ = ⟨x, z⟩ ∀ x ∈ V.
Note that the choice of this vector z depends upon the fixed vector y. This is
because if the fixed vector y is replaced by another vector y1 , we shall get
another linear functional f1 , and f1 will be represented as an inner product with
some other vector z1 . Of course, you can see that f depends on T also!
T∗ ∶ V → V ∶ T∗ (y) = z.
⟨Tx, y⟩ = ⟨x, T∗ y⟩ for all x, y ∈ V (since both are equal to ⟨x, z⟩).
Theorem 3: If (V, ⟨ , ⟩) is an inner product space over the field F and T ∈ A(V),
then T∗ is a linear transformation, i.e. T∗ ∈ A(V).
So, we have shown that given T ∈ A(V) ∃ T∗ ∈ A(V), such that ⟨Tx, y⟩ = ⟨x, T∗ y⟩ for
x, y ∈ V. Now, we will show that T∗ is unique. ■
Definition 1: If (V, ⟨ , ⟩) is an inner product space over the field F and T ∈ A(V),
then the unique operator T∗ ∈ A(V) for which ⟨Tx, y⟩ = ⟨x, T∗ y⟩ holds for all
x, y, ∈ V, is called the adjoint of the operator T. (We also call T∗ the adjoint
operator.)
Example 2: Let Pn (ℂ) denote the vector space of all polynomials of degree ≤ n
with complex coefficients. Show that we can define an inner product on
Pn (ℂ) = Pn as follows:
n
⟨f, g⟩ = ∑ ai bi where f = a0 + a1 t + ⋯ + an tn and g = b0 + b1 t + ⋯ + bn tn . Find T∗ for the
i=0
operator T defined by (Tf)(t) = af(t), a ∈ ℂ.
Solution: Take B = {1, t, t2 , … , tn } in Example 4 of Unit 14. Then you can see
that ⟨ , ⟩, defined above, is an inner product. Now for f, g ∈ Pn .
∗∗∗
Theorem 5: Let (V, ⟨ , ⟩) be an inner product space over F. Then, for S, T ∈ A(V),
28 the following relations hold.
Unit
. . . . .15
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Hermitian
. . . . . . . . . . .and
. . . . Unitary
. . . . . . . . .Operators
..........
a) I∗ = I, I being the identity operator.
b) (S + T)∗ = S∗ + T∗
f) T∗ T = 0 iff T = 0.
g) (T ∘ S)∗ = S∗ ∘ T∗ .
Proof: We will prove (e), (f) and (g) here, assuming (a) to (d). We leave the
proof of (a)-(d) to you (see E5.)
e) Choose any two vectors x, y ∈ V. Then,
This is true for any y ∈ V. Therefore, T∗∗ (x) = T(x) ∀ x ∈ V. Hence, T∗∗ = T.
T(x) = 0 ∀ x ∈ v
⇒ T∗ T(x) = 0 ∀ x ∈ V
⇒ T∗ T = 0.
g) For any x, y ∈ V,
Now, look closely at (e) and (f) of Theorem 5. They tell us that for any T ∈ A(V)
E7) Show that the map 𝜙 ∶ A(V) → A(V) ∶ 𝜙(T) = T∗ is sesquilinear, that is ,
𝜙(S + T) = 𝜙(S) + 𝜙(T), and 𝜙(𝛼S) = 𝛼𝜙(S), ∀ S, T ∈ A(V) and 𝛼 ∈ F.
E8) Using Theorem 5, prove that if T ∈ A(V) and T−1 exists then (T−1 )∗ = (T∗ )−1 .
Now that you are familiar with the adjoint operator, let us look at some
operators whose adjoints have special properties.
for any x, y ∈ V.
If V is a real inner product space and T is self-adjoint, then the above condition
is reduced to ⟨Tx, y⟩ = ⟨Ty, x⟩ (since z = z ∀ z ∈ ℝ).
In this case T is said to be symmetric.
In Unit 12, you have already studied about the eigenvalues and eigenvectors of
operators. Let us see what they look like in the case of self-adjoint operators.
E11) Let V be a complex inner product space and T ∈ A(V) such that T∗ = −T.
Show that
a) iT is self-adjoint, where i = √−1.
b) the eigenvalues of T are purely imaginary number or 0.
T = 0 ⇒ Tx = 0 ∀ x ∈ V ⇒ ⟨Tx, x⟩ = 0 ∀ x ∈ V.
⟨Tx, y⟩ − ⟨Ty, x⟩ = 0 ∀ x, y ∈ V.
This theorem will come in useful in the next sub section, where we look at
another type of linear transformation.
TT∗ = I = T∗ T.
Can you think of an example of a unitary operator? Does the identity operator
satisfy the equation II∗ = I = I∗ I? Yes.
Another example is f ∶ ℝ2 → ℝ2 ∶ f(x, y) = (y, x).
From E9 you know that f = f∗ . Also
Similarly f∗ f = I. ∴ f is unitary.
In both these examples you may have noticed that the operators are also
self-adjoint. The following exercise will give you an example of a unitary
operator which is not self-adjoint.
T ∶ ℝ3 → ℝ3 ∶ T (x1 , x2 , x3 ) = (x3 , x1 , x2 )
We will now prove a theorem that shows the utility of a unitary (orthogonal)
operator.
Theorem 8: If (V, ⟨ , ⟩) is an inner product space over F and T ∈ A(V), then the
following conditions are equivalent.
a) T∗ T = I.
Proof: We shall prove (a)⇒(b)⇒(c)⇒(a). This will show that all three
statements are equivalent.
(a) ⇒(b): Assume (a). Then, for any x, y ∈ V, ⟨x, y⟩ = ⟨Ix, y⟩.
(b) ⇒(c): If (b) holds for all x, y ∈ V, then it also holds when x = y. This means
that, ∀ x ∈ V.
You will learn about some properties of unitary operators from the following
exercises.
E13) If V is a given inner product space over ℂ and S, T ∈ A(V) are unitary
operators, show that
a) S ∘ T is a unitary operator.
b) 𝛼T is a unitary operator for 𝛼 ∈ ℂ iff |𝛼| = 1.
E14) Show that the characteristic roots of a unitary operator have absolute
value 1.
Let us now talk about the action of a unitary operator on an orthonormal basis.
From Unit 14 Theorem 7 you know that (V⟨ , ⟩) has an orthonormal basis. The
following theorem characterises unitary operators in terms of their action on an
orthonormal basis.
We will first show that if T is unitary then {Te1 , … , Ten } is an orthonormal basis of
V. Now, since T preserves inner products, we get ⟨Tei , Tej ⟩ = 0 for i ≠ j, and
⟨Tei , Tej ⟩ = 1 ∀ i, j = 1, … , n. Also, since T is invertible (in fact, T−1 = T), you know
that T maps a basis to a basis. Hence {Te1 , … , Ten } is an orthonormal basis.
= ∑ ∑ 𝛼i 𝛽j ⟨Tei , Tej ⟩
i j
∗∗∗
Note that this example shows us that Theorem 7 is false if T is not self-adjoint.
So far we have been discussing various kinds of operators. You may have
wondered about their matrix analogues. That is what we will discuss in the next
section.
and
n n
⇒⟨∑ aki ek , ej ⟩ = ⟨ei , ∑ bkj ek ⟩
k=1 k=1
n n
⇒ ∑ aki ⟨ek , ej ⟩ = ∑ bkj ⟨ei , ek ⟩
k=1 k=1
0, if i ≠ j
⇒aji = bij , since ⟨ei , ej ⟩ = {
1, if i = j.
Recall that given a matrix A = [aij ], its conjugate transpose is the matrix A∗ = [a∗ij ],
where a∗ij = aji , i.e., A∗ = At .
0 1 0 0 0 0
∗
[0 0 2] ∴ [D ]B = [1 0 0]
0 0 0 0 2 0
Theorem 11: Let V be an inner product space over F and T ∈ A(V). Let the
matrix representation of T with respect to an orthonormal basis B = {e1 , … , en }
be A. Then T is self-adjoint iff A is Hermitian.
Proof: Let [T]B = A = [aij ]. Then, by Theorem 10, [T∗ ]B = [bij ], where bij = aij . That
is, [T∗ ]B = A∗ .
n n
⟨Tei , ek ⟩ = ⟨∑ aji ej , ek ⟩ = ∑ aji ⟨ej , ek ⟩ = aki = aik ∀ i, k = 1, … , n.
j=1 j=1
n
Also T∗ ei = ∑ a∗ji ej ∀ i = 1, 2, … , n.
j=1
∗
∴ ⟨T∗ ei , ek ⟩ = aki = aik ∀ i, k = 1, … , n.
∴ ⟨Tei , ek ⟩ = ⟨T∗ ei , ek ⟩ ∀ i, k = 1, … , n.
So, by Theorem 11 we know the matrix of the operator in E9, with respect to
0 1
the standard basis, is a Hermitian matrix. That is, [ ] is Hermitian.
1 0
k a + ib c + id
2 1+i
[3], [ ] and [a − ib m e + if ], where a, b, c, d, e, f, k, m, n ∈ ℝ.
1−i 0
c − id e − if n
E17) How any characteristic roots of a Hermitian matrix are purely imaginary?
(Hint: Use Theorem 6.)
Theorem 12: let V be an inner product space over F with dim V = n. Let
U ∈ A(V) have a matrix representation A = [aij ], with respect to an orthonormal
basis B of V. If U is unitary, then
n
a) ∑ aik ajk = 𝛿ij
k=1
n
b) ∑ aki akj = 𝛿ij , ∀ i, j, … , n.
k=1
1 −1
Example 5: Show that the matrix A = [ ] is not orthogonal.
2 3
1 −1 1 2 2 −1
AA∗ = [ ] [ ]=[ ]≠I
2 3 −1 3 −1 13 37
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
This means that A is not orthogonal.
∗∗∗
The following exercises will give you some examples of unitary matrices.
0 1 i 0 i 1 1+i
[ ], [ ], [ ].
−1 i 2 i 0 1−i 0
cos 𝜃 sin 𝜃
E20) Is [ ] orthogonal?
− sin 𝜃 cos 𝜃
Theorem 13: For a square matrix A over ℂ the following are equivalent.
a) A is unitary.
R1
(a)⇔(b): Let A = [aij ] = [ ⋮ ] , where Ri is the ith row of A. Then Rti will be the ith
Rn
∗
column of A .
R1
∗ t t
∴ AA = I ⇔ [ ⋮ ] [R1 , … , Rn ] = I.
Rn
aj1
⇔ [ai1 , … , ain ] [ ⋮ ] = 𝛿ij
ajn
n
⇔ ∑ aik ajk = 𝛿ij .
k=1
⇔ the set of vectors { (ai1 , … , ain )| i = 1, … , n} are orthonormal.
⇔ the rows of A are orthonormal.
Similarly, using the fact that A∗ A = I, we can prove that (a)⇔(c). Hence, we
38 have proved the theorem. ■
Unit
. . . . .15
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Hermitian
. . . . . . . . . . .and
. . . . Unitary
. . . . . . . . .Operators
..........
Note: Just as we have proved Theorem 13, we can prove that a real square
matrix is orthogonal iff rows (or columns) form an orthonormal set of vectors.
You can apply what we have just said to solve the following exercise.
Now let us look at real matrices only for the rest of the section.
Recall that a matrix A is symmetric if A = At . In Unit 12, you also came across
the concept of similar matrices. We now define an allied concept.
Definition 5: Two square matrices A and B, of the same order, are said to be
orthogonally similar if A = P−1 BP, for some orthogonal matrix P.
1 2 −2 −1
Example 6: Show that [ ] and [ ] are orthogonally similar.
−1 −2 2 1
a b
Solution: Suppose P = [ ] is an orthogonal matrix satisfying
c d
1 2 −2 −1
[ ] = Pt [ ] P.
−1 −2 2 1
Then we have
1 2 a c −2 −1 a b
[ ]=[ ] [ ] [ ]
−1 −2 b d 2 1 c d
−2a + 2c −a + c a b (2a + c)(c − a) (2b + d)(c − a)
=[ ] [ ]=[ ]
−2b + 2d −b + d c d (2a + c)(d − b) (2b + d)(d − b)
1 = (2a + c)(c − a)
−1 = (2a + c)(d − b)
2 = (2b + d)(c − a)
−2 = (2b + d)(d − b),
we get
0 1 0 −1
P=[ ] or [ ].
1 0 −1 0 39
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
t
1 2 0 1 −2 −1 0 1
∴ [ ]=[ ] [ ] [ ]
−1 −2 1 0 2 1 1 0
t
1 2 0 −1 −2 −1 0 −1
or [ ]=[ ] [ ] [ ].
−1 −2 −1 0 2 1 −1 0
Check that these equalities do hold, by multiplying the right hand side.
∗∗∗
This example shows that there can be several orthogonal matrices P such that
A = Pt BP.
a) P is orthogonal.
n x1 y1
t
X ⋅ Y = ∑ xi yi = X Y ∀ X = [ ⋮ ] , Y = [ ⋮ ] in Vn (ℝ).
i=1
xn yn
Now,
Since 𝛼1 ≠ 𝛼2 , we get X1 ⋅ X2 = 0.
Similarly Xi ⋅ Xj = 0 ∀ i ≠ j.
Also ‖Xi ‖ = 1 ∀ i = 1, … , n, since the X‵i s are normalised vectors.
Therefore, (X1 , X2 , … , Xn ) is an orthonormal set.
Therefore, by Theorem 13, P is orthogonal.
b) From Unit 12 (Theorem 5) you know that P−1 AP = diag (𝛼1 , … , 𝛼n ) . That is,
Pt AP = diag (𝛼1 , … , 𝛼n ) .
40 ■
Unit
. . . . .15
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Hermitian
. . . . . . . . . . .and
. . . . Unitary
. . . . . . . . .Operators
..........
What Theorem 14 says is that any real symmetric n × n matrix with n
distinct eigenvalues is orthogonally similar to a diagonal matrix.
Note: Though we have proved Theorem 14 for real symmetric matrices with
distinct eigenvalues, it is true for any real symmetric matrix. That is, any real
symmetric matrix is orthogonally similar to a diagonal matrix. The proof of
this result is beyond the scope of this course.
1 1 1
Example 7: Reduce [1 1 −1] to diagonal form.
1 −1 −1
1 1 1
Solution: The matrix A = [1 1 −1] is a real symmetric matrix. Its
1 −1 −1
𝜆−1 −1 −1
characteristic equation is | −1 𝜆−1 1 | = 0. This shows us that the
−1 1 𝜆+1
eigenvalues of A are 1, 2, −2.
Eigenvectors corresponding to them are (1, −1, 1), (1, 1, 0) and (−1, 1, 2),
respectively. Therefore, the normalised eigenvectors are
(1/√3) (1, −1, 1), (1/√2) (1, 1, 0), (1/√6) (−1, 1, 2).
1 0 0
t
Then, we get P AP = [0 2 0 ] .
0 0 −2
∗∗∗
7 −1 −10
E22) Reduce [ −1 7 10 ] to diagonal form.
−10 10 −2
(Its eigenvalues are 6, −12 and 18.)
15.6 SUMMARY
As in the previous unit, the vector spaces considered in this unit are all defined
over the fields ℂ or ℝ. We made the following points in this unit. 41
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
1. Any linear functional on an inner product space is represented by the
inner product with a fixed vector.
15.7 SOLUTIONS/ANSWERS
E1) For any x1 , x2 ∈ ℝ2 , we have
∴ fy is a linear functional on ℝ2 .
42 ∴ f ∈ V∗ .
Unit
. . . . .15
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Hermitian
. . . . . . . . . . .and
. . . . Unitary
. . . . . . . . .Operators
..........
E4) ⟨T(x1 , … , xn ), (y1 , … , yn )⟩ = ⟨(x1 , 0, … , 0), (y1 , … , yn )⟩
= x1 y1 = ⟨(x1 , … , xn ), (y1 , 0, … , 0)⟩.
∴ T∗ (y1 , … , yn ) = (y1 , 0, … , 0)
∴ T∗ = T in this case.
⇒⟨x, I∗ y⟩ = ⟨x, y⟩
⇒I∗ (y) = y ∀ y ∈ V ⇒ I∗ = I.
b) For any x, y ∈ V,
∴ (S + T)∗ = S∗ + T∗ .
c) For any x, y ∈ V,
∴ (𝛼T)∗ = 𝛼T∗ .
d) ⟨T∗ y, x⟩ = ⟨x, T∗ y⟩ = ⟨Tx, y⟩ = ⟨y, Tx⟩.
E6)
T = 0 ⇒ Tx = 0 ∀ x ∈ V ⇒ ⟨Tx, y⟩ = 0 ∀ x, y, inV.
⇒ ⟨x, T∗ y⟩ = 0 ∀ x, y ∈ V.
In particular, for x = T∗ (y) ∈ V, we get
⟨T∗ y, T∗ y⟩ = 0 ∀ y ∈ V.
⇒ T∗ y = 0 ∀ y ∈ V ⇒ T∗ = 0.
⇒ I∗ = (T−1 )∗ ⋅ (T∗ )
⇒ I = (T−1 )∗ ⋅ (T∗ )
∴ 𝛼 = −𝛼 ⇒ 𝛼 = 0 or 𝛼 is purely imaginary.
∴ 𝛼 = 𝛼−1 ⇒ 𝛼𝛼 = 1 ⇒ |𝛼| = 1.
E16) a) Let A = [aij ], B = [bij ] and AB = C = [cij ]. Then the (i, j)th element of
n
C∗ = conjugate of the (j,i)th element of C = ∑ ajk bki …(3)
k=1
Now, if B∗ = [dij ] and A∗ = [eij ], then dij = bji and eij = aji . Also,the (i, j)th
element of
n n
B∗ A∗ = ∑ dik ekj = ∑ bki ajk
k=1 k=1
n
= ∑ ajk bki …(4)
k=1
0 i 0 −i
Now, if A = [ ] , then A∗ = [ ].
i 0 −i 0
0 i 0 −i 1 0
∴ AA∗ = [ ] [ ]=[ ].
i 0 −i 0 0 1
Similarly, A∗ A = I.
∴ A is unitary.
1 1+i 1 1+i
If A = [ ] , then A∗ = [ ].
1−i 0 1−i 0
1 1+i 1 1+i 3 1+i
∴ AA∗ = [ ] [ ]=[ ] ≠ I.
1−i 0 1−i 0 1−i 2
∴ A is not unitary.
cos 𝜃 sin 𝜃 cos 𝜃 − sin 𝜃
E20) Let A = [ ] . Then A∗ = [ ].
− sin 𝜃 cos 𝜃 sin 𝜃 cos 𝜃
1 0
∴ AA∗ = [ ] . Also A∗ A = I.
0 1
∴ A is orthogonal.
cos 𝜃 − sin 𝜃 0
E21) The matrix is A = [ sin 𝜃 cos 𝜃 0]
0 0 1
cos 𝜃 sin 𝜃 0
∗
Then A = [− sin 𝜃 cos 𝜃 0]
0 0 1
∗ ∗
∴ AA = A A = I.
∴ A is unitary. ∴, its columns form an orthonormal set of vectors.
1 1 1
E22) The eigenvectors corresponding to 6, −12 and 18 are [1] , [−1] and [−1] ,
0 2 −1
respectively.
∴, the normalised eigenvectors are
1 1 1
1/√2 [1] , 1/√6 [−1] , 1/√3 [−1] .
0 2 −1 45
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
1/√2 1/√6 1/√3
∴ P = [1/√2 −1/√6 −1/√3] is an orthogonal matrix such that
0 2/√6 −1/√3
6 0 0
t
P AP = [0 −12 0 ] .
0 0 18
46
UNIT 16
16.1 INTRODUCTION
So far you have studied various kinds of matrices and inner products. In this
unit, we shall discuss a particular kind of inner product, which is closely
connected to symmetric matrices. This is called a quadratic form. It can also
be thought of as a particular kind of second degree polynomial, which is the
way we shall first define it. We will discuss the geometric aspect of a particular
case of quadratic forms in the next unit.
Before going further make suer that you are familiar with Units 14 and 15.
Objectives
After studying this unit, you should be able to:
• identify a real quadratic form;
ax2 + bx + c = 0, a, b, c ∈ R, a ≠ 0, …(1)
which are called quadratic equations. The left hand side of Eqn. (1) is a
quadratic function in one variable over R. We call the second degree term in
Eqn. (1), i.e., ax2 , a quadratic form of order one. It is called of order one,
since it involves only one variable.
The most general quadratic equation over R involving two variable x and y is
is called a quadratic form of order two, since it involves two variables x and y.
By now you can see how we can generalise this concept. We call the non-zero
form
n
∑ aij xi xj
i,j=1
a quadratic form R of order n, where the a′ij s are real constants and x1 , x2 , … , xn
are real variables.
Remark 1: These expressions are called quadratic, since they are of second
degree. They are called forms, since every term in them has the same degree.
A quadratic form is real if its variables can only take real values and the
coefficients are real number. We have already stated, in the unit introduction,
that all spaces considered in this unit shall be over R. Therefore, by a
quadratic form we shall always mean a real quadratic form.
From the definition of a quadratic form it is clear that a real valued function will
be a quadratic form if and only it it satisfies each of the following conditions:
a) it is a polynomial,
b) it is homogeneous, and
c) it is of degree two.
a) x2 + x + 1
b) 2x2 + y2 + z2
c) x2 − √2y2 = 0
d) 3x21 + x1 x2 − √3x22
e) x31 − x22 + x2 x3
f) x3 + x2 y − y3
g) x2 + log x.
Solution: (c) is an equation, and not a polynomial. (a) and (e) are
polynomials, but they are not homogeneous. (f) is a polynomial which is
homogeneous, but its degree is three and not two. (g) is not a polynomial. Only
(b) and (d) represent quadratic forms. (b) involves three variables, and hence,
its order is three. (d) involves two variables, and thus, has order two.
∗∗∗
a) x2 − xy
b) x1 + x2
c) x31
d) x3 − xy2
e) sin (x2 + 2y2 )
f) x21 − √2x22 = 0
E3) Find the values of the integer k for which the following will represent
quadratic forms.
a) x2 − 2y2 − kxy2
b) xk + 2y2
c) x41 + 2x1 x2 − xk1
E4) Let Q1 and Q2 be two quadratic forms, both of order n, in the n variables
x1 , x2 , … , xn . Which of the following will be a quadratic form?
x 2 1
Putting X = [ ]and A = [ ] , we find that
y 1 3
2 1 x
Q = Xt AX = [xy] [ ] [ ] …(2)
1 3 y
The question now is whether we can replace the matrix A by another matrix
without changing the quadratic form Q. In fact, you can check that
2 2 2 −1
Q = Xt BX, where B = [ ] , and Q = Xt CX, where C = [ ].
0 3 3 3
Note that we can also write Q = ⟨AX, X⟩, where ⟨Y, Z⟩ = Zt Y for any Y, Z ∈ V2 (R).
So, as you go along, remember that we are simultaneously discussing the
50 representation of Q as a matrix product, as well as an inner product.
Unit
. . . . .16
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Real
. . . . . .Quadratic
. . . . . . . . . . .Forms
......
Look carefully at the matrices A, B and C, given above. Do they have a
common feature? You must have noticed that the diagonal elements of all
these matrices are the same, i.e., A, B and C have the same diagonal. Now,
what about the off-diagonal (i.e., non-diagonal) entries? Have you noticed that
the sum of the off-diagonal entries in all these matrices is 2? Note that the
coefficient of the term xy, of the given quadratic form, is also 2.
E5) Change one of the diagonal entries of A and verify that this will change
the quadratic form.
2 a
In fact, any matrix P = [ ] , with a + b = 2, can replace A without changing the
b 3
quadratic form Q. This is because the coefficient of xy in the quadratic form
Xt PX is (a + b). However, if we insist that the matrix P should be symmetric, then
2 1
we must have a = b; and hence, the choice is unique, namely, [ ].
1 3
We, therefore, conclude that A is the only symmetric matrix for which Q = Xt AX.
This symmetric matrix A is called the matrix of the quadratic form Q, or the
matrix associated to the quadratic form Q. Observe that
Given a quadratic form Q of order 2, there are infinitely many square matrices
B for which Q = Xt BX. However, there will be unique symmetric matrix A for
which Q = Xt AX. This matrix A, which is called the matrix of the quadratic form
Q, is given by the rule
1 −1
Example 2: What is the quadratic form generated by A = [ ]?
−1 1
1 −1 x
Solution: The quadratic form generated by A is [xy] [ ] [ ].
−1 1 y 51
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
On expanding this we get x2 − 2xy + y2 .
Observe that you could have obtained the quadratic form simply by applying
the rule (3) as follows:
∗∗∗
𝛼1 0
Example 3: A general diagonal matrix of order 2 is A = [ ].
0 𝛼2
What is the corresponding quadratic form?
𝛼1 0 x
Xt AX = [xy] [ ] [ ] = 𝛼 1 x2 + 𝛼 2 y2 ,
0 𝛼2 y
∗∗∗
a) x2
b) −y2 − 4xy
Solution: Rule (3) is very handy for writing the symmetric matrix of a given
quadratic form. It is easy to see that the corresponding matrices will be
1 0 0 −2
a) [ ], b) [ ].
0 0 −2 −1
∗∗∗
The above discussion involved matrices and quadratic forms of order two. It
can be extended to matrices and quadratic forms of higher orders. Let us look
52 at the case of quadratic forms of order 3.
Unit
. . . . .16
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Real
. . . . . .Quadratic
. . . . . . . . . . .Forms
......
Let us consider a general 3 × 3 matrix
Q = Xt AX, …(4)
where Xt = [x1 x2 x3 ] .
Q = a11 x21 + a22 x22 + a33 x23 +(a12 + a21 ) x1 x2 +(a23 + a32 ) x2 x3 +(a13 + a31 ) x1 x3 . …(5)
Observe that the diagonal elements of A, i.e., a11 , a22 and a33 , are the
coefficients of x21 , x22 , and x23 , respectively, in Q given by Eqn. (5).
Also note that the sum of the two entries a12 and a21 determines the
coefficients of x1 x2 , while these two entries do not occur elsewhere in Eqn. (5).
So, if we replace a12 and a21 by two different numbers a′12 and a′21 such that
a′12 + a′21 = a12 + a21 , while keeping other entries of A unchanged, the new matrix
A′ , thus obtained, will not be equal to A. But the quadratic forms generated by A
and A′ will be the same, i.e.,
Q = Xt AX = Xt A′ X.
Similar changes can be made for the entries contributing to the coefficients of
x1 x3 , to obtain matrices different from A which can replace A without changing
the quadratic form. However, if the matrix A′ is restricted to being symmetric
then the choice is unique, i.e.,
a′12 = a′21 = 1
2
(a12 + a21 ) = 1
2
(coef. of x1 x2 ),
a′13 = a′31 = 1
2
(a13 + a31 ) = 1
2
(coef. of x1 x3 ),
Therefore, the unique symmetric matrix corresponding to the quadratic form (5)
will be
coef. of x21 1
2
coef. of x1 x2 1
2
coef. of x1 x3
A′ = [ 12 coef. of x1 x2 coef. of x22 1
2
coef. of x2 x3 ] …(6)
1
2
coef. of x1 x3 1
2
coef. of x2 x3 coef. of x23
Given a quadratic form of order 3, there are infinitely many matrices of order 3
which will generate it. However, a symmetric matrix that will generate a
quadrate form of order three is unique. This symmetric matrix is called the
matrix associated to the quadratic form, or simply, the matrix of the quadratic
form.
But, a quicker way is to use the rule (6). Comparing the entries of A′ in (6) with
those of A above we can obtain all the coefficients of the quadratic form as
follows:
Coefficients of x21 , x22 , x23 will be the elements of the diagonal in A, i.e. 1, 4 and 2,
respectively.
∗∗∗
2 1 −3
Solution: Using the rule (5), we can write the matrix as [ 1 −1 0 ]
−3 0 1
∗∗∗
Example 7: Find the quadratic form associated with the zero matrix of order
three.
Solution: All the entries of a zero matrix are zero. Therefore, using (6), we
get all the coefficients to be zero. The associated quadratic form is, then,
∗∗∗
𝜆1 0 0
Example 8: Consider the general diagonal matrix of order three, [ 0 𝜆2 0 ].
0 0 𝜆3
54 What is the associated quadratic form?
Unit
. . . . .16
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Real
. . . . . .Quadratic
. . . . . . . . . . .Forms
......
Solution: The associated quadratic form is the diagonal form 𝜆1 x22 + 𝜆2 x22 + 𝜆3 x23 .
∗∗∗
Can we extend the comments about quadratic forms of order two and three to
a quadratic form of any finite order n? Yes. You know that a general quadratic
form of order n is given by
n
Q = ∑ aij xi xj , where aij = aji ∀ i, j = 1, … , n.
i,j=1
Before going further, we would like to remind you that the quadratic form of
order n, Xt AX, is simply the inner product ⟨AX, , X⟩ in Vn (R).
Let us now see what happens to the matrix of a quadratic form if we change
the basis of the underlying vector space.
Let P be the matrix of the change of basis from B to B′ . Then P = [aij ], where
n
e′j = ∑ aij ei .
i=1
Previously, we have seen that P is invertible. Note that the columns of P are
the components of the vectors of the new basis B′ , expressed in terms of the
original basis B.
a11 … a1n
x1 y1
… … …
[ ⋮ ]=[ ][ ⋮ ]
… … …
xn yn
an1 … ann
i.e., X = PY.
The above discussion shows that, under a change of basis given by the
invertible matrix P, the coordinate transformation is given by X = PY, and the
quadratic form Xt AX gets transformed into another quadratic form Yt CY, where
C = Pt AP. This leads us to the following definitions.
In particular, if the matrices A and B are orthogonally similar (see Unit 15) then
the corresponding quadratic forms, Xt AX and Yt BY are called orthogonally
equivalent.
b) Let Q(X) = x21 − 2x1 x2 + 4x22 . Find the expression of Q in terms of y1 and y2 .
Solution:
x1 1 1 y1
[ ]=[ ] [ ] , or X = PY, say. …(7)
x2 0 2 y2
x1 y + y2
[ ]=[ 1 ]
x2 2y2
i.e., x1 = y1 + y2
x2 = 2y2
1 −1 x1
b) Now Q(X) = [x1 , x2 ] [ ] [ ]
−1 4 x2
where
1 0 1 −1 1 1
Pt AP = [ ] [ ] [ ]
1 2 −1 4 0 2
1 −1
=[ ]
−1 13
Thus, under the change of basis given by X = PY, the given quadratic form
transforms into (9).
∗∗∗
The following exercises will give you some more practice in dealing with
quadratic forms under a change of basis.
E10) Verify that the matrix P in Example 9 is not orthogonal. (Therefore, (7) is
not a orthogonal transformation. Therefore, (8) and (10) are equivalent,
but not orthogonally equivalent.)
with respect to the standard basis B1 = {(1, 0), (0, 1)} of ℝ2 . Find its
expression with respect to the basis B2 = {(2, 1), (1, −2)} .
Recall that multiplication by a non-singular matrix does not change the rank of
a matrix. Therefore,
Definition 3: The rank of a quadratic form is the rank of its associated matrix.
You may think that this definition is not meaningful, because the associated
matrix depends on the basis of the vector space. But Theorem 2 will assure us
that the definition is meaningful. So, basically if we find the rank of any
associated matrix, that rank will be the rank of the given quadratic form. (See
Example 10(b)).
Theorem 2: The rank of a quadratic form does not change under a change of
basis.
E13) Verify that the rank of a diagonal form is the number of non-zero terms in
its expression.
where [x1 , x2 , x3 ] are the coordinates of X with respect to the standard basis of
ℝ3 .
1 1 −2 1 −1 1 1 1
B = {( , , ),( , , 0) , ( , , )}
√6 √6 √6 √2 √2 √3 √3 √3
a) Let Yt = [y1 , y2 , y3 ] denote the coordinates with respect to the new basis B.
Then, the change of coordinates is given by
1 1 1
√6 √2 √3
−1
X = [ √1 √2
1
√3
] Y = PY (say)
6
−2 1
0
√6 √3
2 1 −3
A = [ 1 2 −3]
−3 −3 6
Using this, we get Q(Y) = 9y21 + y22 , which is the required quadratic form.
Note that P is an orthogonal matrix. ∴ Q(X) and Q(Y) are orthogonally
equivalent.
b) Now, let us obtain rank (Q) directly. We know that rank (A) = 2.
∴ rank (Xt AX) = 2, i.e., the rank of Q is 2.
Another way of showing that rank Q(X) = 2 is as follows: Q(X) and Q(Y)
are equivalent, and the rank of the diagonal quadratic form Q(Y) is two. ∴,
rank of Q(X) is also two.
∗∗∗
The following exercise will give you some practice in obtaining the rank of a
quadratic form.
We shall now use the relation (11) to transform any quadratic form to a
diagonal form.
X = RY, …(12)
Yt = [y1 , y2 , … , yn ] begin the coordinates with respect to the new basis. R being
orthogonal, (12) is an orthogonal transformation which will convert Xt AX into
because of (11).
We say that the orthogonal transformation (12) has reduced the quadratic
form Xt AX into its orthogonal canonical form, given by (13). The form in
(13) is orthogonal since the transformation used to convert Xt AX into it is
orthogonal. It is called canonical as the reduced form is the Simplest
orthogonal reduction of Xt AX. The elements of the basis which diagonalise the
quadratic form (in this case they are U1 , … , Un ) are called the principal axes of
the quadratic form.
So far we have spoken about the orthogonal canonical form in an abstract way.
Let us now look at a practical method of reducing a quadratic form to its
orthogonal canonical form.
2. Form the characteristic equation det (A − 𝜆I) = 0 and find the eigenvalues of
A. Let 𝜆1 , … , 𝜆r be the non-zero eigenvalues arranged in decreasing order,
i.e., 𝜆1 ≥ 𝜆2 ≥ … ≥ 𝜆r .
7. The new basis {U1 , U2 , … , Un } is called the canonical basis and its
elements are the principal axes of the given quadratic form.
In step 2 you are required to find the eigenvalues, i.e., the roots of the
characteristic equation. In a realistic situation the roots can be irrational
numbers and we may have to use numerical methods determine such roots.
We have avoided irrational numbers by carefully selecting the quadratic forms
in our examples and exercises so that the roots of characteristic equations are
rational numbers.
To clarify the procedure given above we present some examples and exercises.
Example 11: Obtain the unique orthogonal canonical form of the quadratic
form 5x21 − 6x1 x2 + 5x22 .
5 −3
Solution: The matrix of this quadratic form is A = [ ].
−3 5
𝜆−5 3
The eigenvalues of A are given by | | = 0, i.e.,
3 𝜆−5
𝜆2 − 10𝜆 + 16 = 0 ⇒ 𝜆 = 8, 2.
Thus, the new orthonormal basis is {U1 , U2 } , which is the canonical basis. U1
and U2 are the principal axes of the given form.
x1 −1/√2 1/√2 y1
[ ]=[ ] [ ]
x2 1/√2 1/√2 y2
63
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
Find its orthogonal canonical reduction and the corresponding new basis.
1 1 1
Solution: The matrix of (14) is A = [1 1 1] .
1 1 1
3x21 , …(15)
1/√3
A normalised eigenvector corresponding to the eigenvalue 3 is [1/√3] .
1/√3
1 1 1 x 0
[1 1 1] [y] = [0] ,
1 1 1 z 0
i.e., x + y + z = 0 …(16)
Here we can choose any two mutually orthogonal normalised vectors satisfying
1/√2 1/√6
(16). Let us choose [−1/√2] and [ 1/√6 ] .
0 −2/√6
which is the canonical basis. Its elements are the principal axes of (14). The
change of basis needed to convert (14) into (15) is given by
∗∗∗
The next few exercises will give you some practice in applying the procedure of
reduction.
E16) Find the orthogonal canonical forms to which the following quadratic
forms can be reduced by means of an orthogonal change of basis. Also
obtain a set of principal axes for them.
a) x2 + 4xy + y2
b) 8x2 − 4xy + 5y2
c) 3x22 + 3x23 + 4x1 x2 + 4x1 x3 − 2x2 x3
E18) Show that the quadratic forms x2 − 2y2 + z2 and z21 − 2x21 + y21 are
orthogonally equivalent. Find the orthogonal transformation which will
transform the first of these into the second.
We will now try to reduce the matrix of a quadratic form to a diagonal form
whose diagonal elements are only 1, −1 or 0.
Let us now look at some examples that will help you in understanding the
procedure.
z1 = √8y1 , z2 = √2y2
√8 0 z y
i.e., z = [ ] Y, where Z = [ 1 ] and Y = [ 1 ] .
0 √2 z2 y2
This transformation, which is non-singular but not orthogonal, will convert (22)
into z21 + z22 , which is the required normal canonical form.
66 ∗∗∗
Unit
. . . . .16
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Real
. . . . . .Quadratic
. . . . . . . . . . .Forms
......
Example 14: Reduce the diagonal form 2x21 − 3x22 − 7x23 into its normal
canonical form.
y1 = √2x1
y2 = √3x2
y3 = √7x3
√2 0 0
i.e., Y = [ 0 √3 0 ] X.
0 0 √7
This will convert the given diagonal form into y21 − y22 − y23 .
∗∗∗
E20) Reduce the following quadratic forms to their normal canonical forms.
E21) Show that the rank of a normal canonical form is the number of non-zero
terms in its expression.
E22) Show that a quadratic form and its normal canonical reduction have the
same rank.
Thus, (23) and (24) are both normal canonical reduction of Q, in which the
number of positive terms are p are p′ , respectively. To prove the theorem we
have to prove that p = p′ . Let U and V be the subspaces of ℝn generated by
{u1 , … , up } and {vp′+1 , … , vn } , respectively.
Suppose U ∩ V ≠ 0. Let 0 ≠ u ∈ U ∩ V.
⇒ p + n − p′ ≤ n
⇒ p ≤ p′ …(27)
p′ ≤ p …(28)
Thus, s = p − (r − p) = 2p − r.
For example, for the form Example 13, we have p = 2, r = 2 and s = 2. For the
form in Example 14, p = 1, r = 3, s = −1.
E23) Find the rank and signature of the quadratic forms given in E 20.
The rank and the signature completely determine the normal canonical
reduction. Also, any two quadratic forms having the same canonical reduction
will be equivalent. We can, therefore, state the following result.
Theorem 7: Two quadratic forms are equivalent if and only if they have the
same rank and signature.
Ip 0 0
[ 0 −Ir−p 0 ]
0 0 0n−r×n−r
And now we end the unit by briefly recalling what we have done in it.
16.8 SUMMARY
In this unit all the spaces considered are over the field R. In it we have covered
the following points.
16.9 SOLUTIONS/ANSWERS
E4) The first three will be quadratic forms, if they are non-zero. Q1 Q2 will be
degree 4. Q1 /Q2 will also not be quadratic; in fact, it may not even be a
polynomial.
1 1
E5) For example, the matrix [ ] gives us the quadratic form
1 3
1 1
Xt [ ] X = x2 + 2xy + 3y2 .
1 3
0 0 2 0 0 1 p q/2
E6) a) [ ], b) [ ], c) [ ], d) [ ].
0 −1 0 1 1 0 q/2 r
7 −1 −10 x
E7) a) [x y z] [ −1 7 10 ] [y]
−10 10 −2 z
−1
1 2
x1
b) [x1 x2 ] [ −1 ] [ ]
2
1 x2
1 −1 0 x1
c) [x1 x2 x3 ] [−1 0 0] [x2 ]
70 0 0 0 x3
Unit
. . . . .16
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Real
. . . . . .Quadratic
. . . . . . . . . . .Forms
......
0 0 1 x
d) [x y z] [0 0 1] [y]
1 1 0 z
E8) a) ax2 + by2 + cz2 + 2hxy + 2gxz + 2fyz
b) x2 − y2
c) 4x2 − √2y2 − z2
E9) Xt AX = 4x21 + x22 + 4x24 + 2x1 x2 + 3x1 x3 + 2x1 x4 + 6x2 x3 + x3 x4 .
4 1 3/2 1
1 1 3 0
A′ = [ ]
3/2 3 0 1/2
1 0 1/2 4
1 0 1 −1 1 1 1 0
Pt AP = [ ] [ ][ ]=[ ]
1 1 −1 4 0 1 0 3
2 2
∴ Q(Y) = y1 + 3y2 .
7 26
E12) A = [ ]
26 −32
The coordinate transformation corresponding to the change from B1 to B2
2 1
is given by the matrix P = [ ] . ∴, the matrix of the form will now be
1 −2
2 1 7 26 2 1 100 0
Pt AP = [ ] [ ] [ ]=[ ]
1 −2 26 −32 1 −2 0 −225
5 −2 0
E14) a) The rank of the form = rank of [−2 6 −2] = 3, since its determinant
0 −2 7
rank is 3.
1 1 1
b) rank (Q) =rank of [1 1 1] = 1, since its row-reduced echelon form is
1 1 1
1 1 1
[0 0 0] .
0 0 0 71
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
2 1 1
c) rank (Q) =rank of [1 2 1] = 3, since its determinant is non-zero.
1 1 2
1 0 0
d) rank (Q) =rank of [0 1 0] = 2, using the determinant rank method.
0 0 0
E15) The required transformation is X = PY, where P = [−U1 − U2 ] , i.e.,
1
x1 = (y1 − y2 )
√2
−1
x2 = (y1 + y2 )
√2
1 2
E16) a) The matrix of the form is A = [ ] . Its eigenvalues are 3 and −1. ∴,
2 1
the given form is equivalent to 3x21 − y21 . Normalised eigenvectors
1
1/√2
corresponding to 3 and −1 are [ ] and [ √−12 ] , respectively. ∴, they
1/√2
√2
form a set of principal axes of the form. Remember that the principal
axes are not unique.
b) Its orthogonal canonical form is 9x21 + 4y21 .
−2/√5 1/√5
A set of principal axes is {[ ],[ ]} .
1/√5 2/√5
c) Its orthogonal canonical reduction is 4y21 + 4y22 − 2y23 .
Eigenvectors corresponding to the eigenvalues 4 are given by
0 2 2 x x
[2 3 −1] [y] = 4 [y] ⇒ 2x − y − z = 0.
2 −1 3 z z
0 1/√5
[ 1/√2 ] , [ 0 ]
−1/√2 2/√5
E17) Any two forms are orthogonally equivalent iff they the same orthogonal
canonical forms as given in Theorem 4. ∴, their matrices should have the
72 same eigenvalues (including repetitions).
Unit
. . . . .16
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Real
. . . . . .Quadratic
. . . . . . . . . . .Forms
......
Now, the eigenvalues of the matrices in (a) and (c) are 12, 12 and −6. ∴,
the forms in (a) and (c) are orthogonally equivalent. The matrix of the
form in (b) has eigenvalues 9, 9, −9. ∴, it is not orthogonally equivalent to
the others.
E18) Both the forms have the same diagonal forms, as given in Theorem 4,
x x′ x1 x′
namely x′2 + y′2 − 2z′2 . If [y] = P [y′ ] and [y1 ] = Q [y′ ] , then PQ−1 will
z z′ z1 z′
transform the first to the second, and
1 1
P = diag ( ,⋯, , 1, … , 1) .
√|𝜆1 | √|𝜆r |
73
UNIT 17
CONICS
Structure
Page Nos.
17.1 Introduction 74
Objective
17.2 Definitions and Equations 75
What is a Conic?
Standard Equations of Conics
17.3 Ellipse 79
Description
Geometrical Properties
17.4 Hyperbola 83
Description
Geometrical Properties
17.5 Parabola 86
Description
Geometrical Properties
17.6 The General Theory of Second Order Curves in ℝ2 88
17.7 Summary 69
17.8 Solution/Answers 70
17.1 INTRODUCTION
In Unit 16 you have studied about real quadratic forms of any order n. This unit
is only a geometric extension of the previous one. In it we shall confine
ourselves to the two dimensional case.
Circles, parabolas, hyperbolas and ellipses are curves which we come across
quite often. The ancients Greeks studied these curves and named them coins
sections, since they could be obtained by taking a plane section of a right
circular double cone (Fig. 1). However, from the analytic viewpoint, the Greek
definition of conics, as sections of a cone, is not particularly useful. We shall
Fig. 1: Right circular
consider a conic to be a curve which can be represented by an equation of
double cone
second degree.
After defining conics, we shall list the different types of standard conics. Then
74 we shall study the ellipse, the hyperbola, and the parabola in detail. In the last
Unit
. . . . .17
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Conics
.......
section we will look at one of the basic problems of plane analytic geometry
that deals with conics–how to obtain a rectangular coordinate system in which
the equation of a given conic takes the standard form.
Objectives
After studying this unit, you should be able to:
• recognise different types of conics and their standard equations;
Now, (1) represent a curve in ℝ2 . We call this curve a conic. Let us make some
formal definitions now.
x2 + y2 = a, a ∈ R. …(2)
A conic consisting of only Case 2: If a = 0, then the only real solution of (2) is x = 0 and y = 0. Hence, the
one point is called a point conic represented by (2) will consist of just one point, i.e., (0, 0).
conic. 2
Case 3: If a > 0, then √a ∈ R and a = (√a) . ∴, a point (x, y) will satisfy (2) if and
only if the distance of (x, y) from the origin is √a. Hence, the conic represented
by (2) will be a circle of radius √a and centre (0, 0).
∗∗∗
2x2 − xy − 3x = 0. …(3)
The examples above show that a circle, a point and a pair of straight lines are
conics.
E2) Find the nature of the conics represented by the following equations.
a) x2 − 2xy + y2 = 0
b) 4x2 − 9x + 2 = 0
c) x2 = 0
d) xy = 0
In the examples and exercises that you have done so far, you have dealt with
simple second degree equations. These and other simple forms are what we
will discuss now.
x2 + 5xy + y2 + 2x − 6y + 10 = 0
x2 y2
Ellipse a2
+ b2
= 1, a, b > 0
Circle x2 + y2 = a2 , a ≠ 0
x2 y2
Hyperbola a2
− b2
= 1, a, b > 0
x2 y2
Pair of intersecting lines a2
− b2
= 0 a, b ≠ 0
x2 y2
Point conic a2
+ b2
= 0, a, b ≠ 0
x=Y
} …(4)
y=X
to the conic.
There are several types of standard conics to which a general quadratic
equation can be reduced. The classification is made on the basis of the
coefficients of the various terms and the constant term appearing in the
equation. In Table 1 we list different types of real conics along with their
standard equations.
From the standard equations of conics that we have listed Table 1, we can
obtain other equality simple equations by the following two methods.
ii) Reversing the direction of an axis: For example, the direction of the
x-axis can be reversed by applying the orthogonal transformation
x = −X
} …(5)
y=Y
to the conic.
Similarly, we can reverse the direction of the y-axis by applying the
orthogonal transformation x = X, y = −Y.
All three equations represent the same parabola with respect to different
coordinate systems.
∗∗∗
E3) What are the different forms of the equation of the circle x2 + y2 = a2 that
we get on applying the transformations (4) and (5) given above?
Let us now study some of these conics in detail. In the following sections we
will describe ellipses, hyperbolas, parabolas and other conics. As we go along
we will also pictorially show you how conics occur as planar sections of a right
78 circular double cone.
Unit
. . . . .17
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Conics
.......
Before starting these sections you may like to recall what you studied about
curve tracing in Block 2 of the Calculus course.
17.3 ELLIPSE
In your school, you have already studied that any planet orbits the sun in an
elliptical path. The sun is at a focus of these ellipses. In this sections, you will
see what exactly an ellipse is and study some of its geometrical properties. In
Fig. 2 you can see why an ellipse is called a conic.
17.3.1 Description
Fig. 2: Ellipse as a
From Sec. 17.2 you know that the standard equation of an ellipse is
section of a double
2 2 2 2
x /a + y /b = 1, a, b > 0 …(6) cone
We may assume a > b. (If b > a, then we can interchange the x and y axes to
arrive at the assumed case.) We want to trace the ellipse (1). For this purpose
we start gathering information.
b) (6) is a central conic: If we replace both x and y by (−x) and (−y) in (6), it
remains unchanged. Thus, the ellipse is symmetric with respect to the
origin. Hence, (0, 0) is the centre of the ellipse.
(a) and (b) tell us that it is enough to sketch the graph in the first quadrant
only, i.e., for x, y 0.
a 2 2
or x = √b − y , 0 ≤ y ≤ b …(8)
b 79
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
x2 y2
Fig. 3: The ellipse a2
+ b2
=1
E4) Prove that the tangents at (a, 0) and (0, b) of the ellipse (6) are x = a and
y = b, respectively.
From the above information the ellipse (6) will be represented by the curve in
Fig. 3.
ii) A′ A and B′ B are called the major and minor axes of the ellipse,
respectively. Their lengths are 2a and 2b, respectively. These axes are
the principal axes of the ellipse. Can you see why? It is because they are
1 0
given by the normalised eigenvectors [ ] and [ ] , of the form
0 1
Fig. 4: Circle as a
x2 /a2 + y2 /b2 .
section of a cone.
iii) The positive real number e defined by a2 e2 = a2 − b2 , is called the
eccentricity of the ellipse. Note that 0 < e < 1.
iv) The points (ae, 0) and (−ae, 0) are called the foci (plural of focus).
∗∗∗
We have seen what happens if a = b in (6). But what happens if b > a in (6)?
The role of the major and minor axes will be interchanged and the terminology
given for an ellipse will have to be suitably modified as follows:
ii) B′ B and A′ A will be the major and minor axes, and their lengths will be 2b
and 2a, respectively.
iv) the points (0, ±be) will be the foci. They will lie on the y-axis. Therefore,
the major axis will lie along the y-axis.
v) The lines y = b/e and y = −b/e will be the directrices corresponding to the
foci (0, be) and (0, −be), respectively.
By now you must be ready to describe an ellipse yourself. Try the following
exercise.
E5) Find the vertices, eccentricity, foci and directrices of the ellipse
Fig. 5: The ellipse
a) 9x2 + 4y2 = 36 (see Fig. 5.) x2 y2
4
+ 9
= 1.
b) 16x2 + 25y2 = 400
Proof: Let P (x1 , y1 ) be a point on the ellipse x2 /a2 + y2 /b2 = 1, a > b (see Fig.
6). Let F1 (ae, 0) be the focus under consideration. The directrix corresponding 81
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
to F1 is x = a/e. Let D be the foot of the perpendicular from P to the directrix
x = a/e. Since P lies on the ellipse we have
which is equivalent to PF21 = e2 PD2 , i.e., PF1 = e(PD), which proves the
Fig. 6: The ellipse
x2 y2 statement for the focus F1 . For completing the proof, try E6. ■
a2
+ b2
= 1.
String Property: For each point P of the ellipse the sum of the distances of P
from the two foci of the ellipse is the same, and is equal to the length of the
major axis.
Proof: let P be a point on the ellipse whose foci are F1 and F2 (see Fig. 6). Let
D1 and D2 be the feet of the perpendiculars from P to the two directrices. Using
the focus-directrix property, we get
You may wonder why this property is called the string property. It provides a
Fig. 7: Sketching an
mechanical method to construct an ellipse by using a string. Let us see what
ellipse using staring
the method is.
E7) Use the method we have just given to draw an ellipse whose eccentricity
is 0 and minor axis is 3 inches in length, on a piece of paper.
An ellipse has another important property which we shall state, but not prove in
this course.
82
Unit
. . . . .17
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Conics
.......
Reflected Wave Property: A ray of light (or sound, or any other type of wave)
emitted from one focus of an ellipse is reflected back from its reflecting interior
to the other focus (see Fig. 8). An interesting consequence of this property is
that rooms with an ellipsoidal ceiling have whispering galleries. A person
standing at one focus of the ellipse can whisper so as to be heard by a person Fig. 8: Reflected wave
at the other focus, while the people in between cannot hear what is said. property
17.4 HYPERBOLA
In this section we shall present the description and some geometrical
properties of a hyperbola. See Fig. 9 for a representation of a hyperbola as a
planner section of a double cone.
17.4.1 Description
From Table 1 you know that the standard equation of a hyperbola is
You can check that this is symmetric about both the axes, and hence about the
origin. The origin is, therefore, the centre of the hyperbola. Thus, the hyperbola
is a central conic. The x-axis meets the hyperbola in (± a, 0) while the y-axis
does not meet it at all.
Due to symmetry about both the axes, it is enough to sketch the hyperbola in
the first quadrant only, i.e., for x, y ≥ 0. In this quadrant it is given by
Fig. 9: Hyperbola as a
x2 y2
y = b √ 2 − 1 (or x = a√ 2 + 1) . section of a double
a b
cone
This provides the following information.
b) y = 0 for x = a.
Can you see that the hyperbola consists of two branches? Of all the conics,
this property is typical of hyperbolas only.
x2 y2
Fig. 10: The hyperbola a2
− b2
=1
ii) The line segment joining the vertices is called the principal (or
transversal) axis, while the line segment joining B and B′ is called the
conjugate axis. The length of the principal axis is 2a, while the length of
the conjugate axis is 2b.
As in the case of an ellipse, these axes are in the direction of the
1 0
normalised eigenvectors [ ] and [ ] , of the matrix of the form
0 1
2 2 2 2
x /a − y /b .
E8) Find the vertices, eccentricity, foci and directrices of the hyperbola
a) 9x2 − 16y2 = 144
b) 25x2 − 9y2 = 225.
Proof: We will start the proof and you can complete it! Let P (x1 , x2 ) be any
point of the hyperbola x2 /a2 − y2 /b2 = 1, a, b > 0. Consider the foci F1 (ae, 0) and
F2 (−ae, 0). Now do E9. ■
E9) Prove that PF1 = ePD, where D =distance of P from the directrix x = a/e.
Also show that PF2 = ePD′ , where D′ =distance of P from the line x = −a/e.
String Property: For each point of a hyperbola the absolute value of the
difference of its distances from the two foci is the same, and is equal to the
length of the principal axis.
Proof: Let P be a point of the hyperbola whose foci are F1 and F2 . Let D1 and
D2 be the feet of the perpendicular from P on the two directrices. Fig. 11 shows
the two cases, when P is on the branch or the other. ■
PF1 = ePD1
PF2 = ePD2 .
85
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
Hence,
You must have noticed the similarly in the properties of an ellipse and a
hyperbola. Sometimes an ellipse or a hyperbola is defined by the
focus-directrix property, an ellipse being when e < 1, and a hyperbola when
e > 1. What happens when e = 1? In other words, what is the locus of a point
whose distance from a fixed point (a focus) is equal to its distance from a fixed
line (a directrix)? We shall answer this question in the next section.
17.5 PARABOLA
Have you ever noticed the path of a projectile when it is acted upon by the
force of gravity only? It is a parabola. In this section we will discuss parabolas
in some detail. In Fig. 12 we show how it can be represented by a planner
section of a cone.
17.5.1 Description
Fig. 12: Parabola as a Table 1 tells you that the standard equation of a parabola is y2 = 4px, p > 0.
section of a double
You can verify the following information about it, as you have done for an
cone
ellipse or a hyperbola.
b) For x < 0 there are no real value of y, and hence, this parabola does not
exists in the second and third quadrants.
In view of (a) and (b), it is enough to sketch the parabola in the first quadrant
only. The part of the parabola in the first quadrant is given by
E10) Find the coordinates of the focus, and the equation of the directrix, of the
parabola
a) y2 = 3x, b) x2 = 4ay, c) y2 = −4ax. Fig. 14: PF = PD
Proof: Let the parabola have standard equation y2 = 4px. Then F(p, 0) is its
focus. Let P (x1 , y1 ) be any point on the parabola (see Fig. 14). Then
y21 = 4px1 . 87
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
Now
2 2 2
PF2 = (x1 − p) + y21 = (x1 − p) + 4px1 = (x1 + p)
= PD2
= (distance of P from the directrix x = −p)2 .
A paraboloid is a surface As a consequence of this property paraboloid surface is used in the headlight
generated by revolving a of cars, optical and radio telescopes, radars, etc. The focus-directrix property is
parabola about its axis. common to an ellipse, a hyperbola and a parabola. Each of them can be
considered as a locus of a point whose distance from a fixed point (a focus) is
a constant, e, times its distance from a fixed line (a directrix). The locus is an
The ellipse, hyperbola ellipse, parabola or hyperbola accordingly as e < 1, e = 1, e > 1. The
and parabola are called focus-directrix property, therefore, unifies all these conics. What about the rest
non-degenerate conics. of the conics given in Table 1? They are all limiting cases of an ellipse, a
hyperbola or a parabola.
Similarly, the ellipse x2 /a2 + y2 /b2 = 1 degenerates into the pair of parallel lines
given by y2 = b2 , as a → ∞.
So far you have studied quite a few conics. But you must be wondering about
curves that are represented by the general equation of second degree.
We will now look at any conic and see how to reduce it to one of the standard
forms given in Sec. 17.2.
We will see how to reduce this equation to standard form, that is, one of the
forms listed in Table 1. You will see that the whole of this section will be
devoted to using the following theorem.
We will give a rough outline of the proof of this theorem. The idea is to first
reduce the quadratic forms ax2 + 2hxy + by2 to the orthogonal canonical form
2 2
𝜆1 x1 + 𝜆2 y1 , with 𝜆1 ≥ 𝜆2 (ref. Sec. 16.6). Let this transformation be given by
x x1
[ ] = P[ ].
y y1
x1 = X + 𝛼, y1 = Y + 𝛽, 𝛼, 𝛽 ∈ R.
We will choose 𝛼 and 𝛽 in such a manner that the linear terms are reduced to
zero. Then our conic (10) will finally be transformed to one of the standard
conics.
Our proof may seem vague to you. To understand the method of reduction
consider the following examples.
7 −4
Solution: The matrix of the quadratic form 7x2 − 8xy + y2 is [ ].
−4 1
∗∗∗
Its eigenvalues are 9 and −1. ∴, form Unit 16 (Theorem 4) you know that we
can find an orthogonal transformation which will reduce 7x2 − 8xy + y2 into
9X2 − Y2 . This transformation will reduce the given conic to 9X2 − Y2 = a.
Solution: The second degree terms in the given equation are the same as in
the quadratic form considered in Example 11 of Unit 16. The orthogonal 89
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
coordinate transformation
x = (1/√2) (−y1 + y2 )
y = (1/√2) (y1 + y2 )
will convert 5x2 − 6xy + 5y2 into 8y21 + 2y22 , and hence will transform the given
equation into
will transform the above equation into 8X2 + 2Y2 = a + 1/2, which is in standard
form.
∗∗∗
The nature of this conic will depend on the value of a. We have the following
three cases:
Case 1: a + 1/2 < 0. In this case no real values of X and Y satisfy the conic, and
hence the conic is imaginary.
X2 Y2
+ = 1,
(2a + 1)/16 (2a + 1)/4
In the following example you can see what a conic looks like before and after
reduction to standard form.
Fig. 16: The ellipse 5x2 − 6xy + 5y2 + √2(x + y) = 4 (a) before reduction. (b) after
reduction.
X2 Y2
+ = 1.
9/16 9/4
We give the sketch of the original equation in Fig. 16(a) and the sketch of the
reduced equation in Fig. 16(b).
So, you see, the shape and size of the conic remains unchanged under the
transformations that we apply to reduced it to standard form.
∗∗∗
x2 + 2xy + y2 − 6x − 2y + 4 = 0.
1 1
Solution: The matrix of the quadratic form x2 + 2xy + y2 is [ ] , whose
1 1
eigenvalues are 2, 0. Normalised eigenvectors corresponding to the 91
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
eigenvalues 2, 0 are ( 1 1 ) and (−1/√2, 1/√2) , respectively. Hence, the
√2 √2
coordinate transformation
x 1/√2 −1/√2 y1
[ ]=[ ] [ ]
y 1/√2 1/√2 y2
i.e., x = (y1 − y2 ) /√2, y = (y1 + y2 ) /√2, will convert x2 + 2xy + y2 into 2y21 , and the
given equation into
y1 − √2 = X, y2 = Y,
Step 1: Use the method of Section 16.6 to reduce ax2 + 2hxy + by2 to 𝜆1 y21 + 𝜆2 y22
using an orthogonal transformation. This transformation will reduce (11) to
2 2
𝜆1 y1 + 𝜆2 y2 + Ay1 + By2 + C = 0 …(12)
Step 2: Now use a suitable translation of axes (y1 , y2 ) ↦ (X, Y) to eliminate the
linear terms and reduce (12) into one of the standard forms. This will give the
reduction of (11).
By now you must be wanting to try and reduce equations on your own. Try this
exercise.
E11) Reduce the following second degree equations to standard form. (Here
a ∈ ℝ.) What is the type of conic they represent?
a) x2 + 4xy + y2 = a
b) 8x2 − 4xy + 5y2 = a
c) 3x2 − 4xy = a
d) 4x2 − 4xy + y2 = 1
e) 16x2 − 24xy + 9y2 − 104x − 172y + 44 = 0
f) 4x2 − 4xy + y2 − 12x + 6y + 9 = 0
We end this unit with briefly mentioning what has been done in it.
17.7 SUMMARY
92 In this unit we have covered the following points.
Unit
. . . . .17
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Conics
.......
1. A conic is defined to be the set of points in ℝ2 that satisfy an equation of
second degree. Conics can be real or imaginary.
3. All these conic, except for a pair of parallel lines, can be obtained by
taking a plane section of a right circular double cone.
6. An ellipse (a hyperbola) satisfies the string property, i.e., for each point P
on the ellipse (hyperbola), the sum (absolute value of the difference) of
the distances of P from the two foci is constant, and is equal to the length
of the major (principal) axes.
17.8 SOLUTIONS/ANSWERS
∴, the tangents to (a, 0) and (0, b) are the lines x = a and y = b, respectively.
x2 y2
E5) The given ellipse is 4
+ 9
= 1. ∴ a = 2, b = 3.
√5
∴, the vertices are (0, ±3), e = √ 99−4 = 3
, the foci are (0, ±√5) and the
corresponding directrices are y = ±9.
√5
a) Here a = 4, b = 3. ∴ e = √ 1616+9 = 54 .
The vertices are (± 4, 0).
The foci are (± 5, 0).
16
The corresponding directrices are x = ± 5
.
x2 y2
b) The hyperbola is 9
− 25
= 1.
√9+25
Here, a = 3, b = 5. ∴ e = 9
= √ 34
9
. The vertices are (±3, 0). The foci
9
are (± √34, 0). The corresponding directrices are x = ± .
√34
E11) a) The second degree terms give the quadratic form x2 + 4xy + y2 . This
reduces to 3x21 − x22 . ∴, the given coins reduces to 3x21 − x22 = a.
If a = 0, this is a pair of straight lines.
If a ≠ 0, this is a hyperbola.
b) 8x2 − 4xy + 5y2 = a reduces to 9x21 + 4x22 = a.
If a = 0, this is a point conic.
If a < 0, this is imaginary.
94 If a > 0, this is an ellipse.
Unit
. . . . .17
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Conics
.......
x 4/5 3/5 x1
[ ]=[ ] [ ],
y −3/5 4/5 y1
2
⇒25x1 + 20x1 − 200y1 + 44 = 0.
2
⇒ (5x1 + 2) − 40 (5y1 − 1) = 0.
−2/√5 1/√5
[ ],[ ]
1/√5 2/√5
∴, the transformation
x −2/√5 1/√5 x1
[ ]=[ ] [ ]
y 1/√5 2/√5 y1
transforms the conic to
12 6
5x21 − (−2x1 + y1 ) + (x1 + 2y1 ) + 9 = 0.
√5 √5 95
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
2
⇒ (√5x1 + 3) = 0.
Now we apply the translation X = √5x1 + 3, Y = y. We get X2 = 0. This
represents a pair of coincident lines.
96
Miscellaneous
. . . . . . . . . . . . . . . .Exercises
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Conics
.......
MISCELLANEOUS EXERCISES
The few exercises, given below cover the concepts and processes you have
studied in this block. Solving the exercises, will give a better understanding of
the underlying concepts concerned. Trying exercises on your own will also give
you more practice and confidence in solving such problem.
1. Calculate the norm of f(x) = x + 1, using the simple inner product defined
in C[1, 2].
3. Let V be a complex inner product space and T ∈ A(V) such that T∗ = −T.
Show that eigenvectors of T corresponding to distinct eigenvalues are
mutually orthogonal.
5. For which pairs of vectors does the equality hold in the Cauchy-Schwartz
inequality?
8. Let P2 , the space of all real polynomials of degree at most 2, have the
inner product
1
⟨p, q⟩ = ∫ p(x) q(x) dx
−1
16. Express the quadratic form in the matrix notation XT AX, where A is a
symmetric matrix.
a) 3x21 + x22 − 6x1 x2 b) 2x2 + 3y2 − z2 + 2xy − 4xz + yz
to the basis
21. Reduce the following equations to standard form and identify the conic
section represented by the equation.
a) −x2 − y2 − 6xy = 8
b) 5x2 − 6xy + 5y2 = 16
SOLUTIONS/ANSWERS TO
MISCELLANEOUS EXERCISES
1
E1) ‖f‖ = ⟨f, f⟩1/2 = √∫ (x + 1)2 dx
0
1
= √∫ (x2 + 2x + 1) dx
0
1
x3
= √[ + x3 + x]
3 0
1
=√ +1+1
3
7
=√
98 3
Miscellaneous
. . . . . . . . . . . . . . . .Exercises
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Conics
.......
E2) v ⟂ (1, 0, 0) ⇒ x ⋅ 1 + y ⋅ 0 + z ⋅ 0 = 0 ⇒ x = 0
v ⟂ (−1, 2, 0) ⇒ x ⋅ (−1) + y ⋅ 2 + z ⋅ 0 = 0 ⇒ −x + 2y = 0.
So, we get x = 0, y = 0. Thus, v is of the form (0, 0, z) for z ∈ ℝ.
E3) Let 𝛼, 𝛽 ∈ C be distinct eigenvalues of T. Let v, w ∈ V be eigenvectors
corresponding to 𝛼 and 𝛽, respectively. Then Tv = 𝛼v and Tw = 𝛽w.
Now
E4) Let A be an upper triangular Hermitian matrix. Then aij = 0 for i > j. Also
A = A∗ . ∴ aij = aji .
∴, for i < j, aij = aji = 0 = 0, since j > i.
∴ ∀ i ≠ j, aij = 0. ∴ A is a diagonal matrix.
Similarly, if A is Hermitian and lower triangular, it must be a diagonal
matrix.
E5) Let (V, ⟨ , ⟩) be an inner product space and x, y ∈ V. Now if x and y are
linearly dependent then one of them is a scalar multiple of the other (in
that case without any loss of generality you can assume that y = 𝛼 x,
where 𝛼 is a scalar).
Then
So, in this case, we can see that the equality holds in the
Cauchy-Schwartz inequality.
E6) Yes, See Example 2 in unit 14.
E7) No. By definition, it should be positive.
1 1
E8) ⟨p, q⟩ = ∫ x. x2 dn = ∫ x3 dn = 0.
−1 −1
So, p ⟂ q.
E9) Let v1 = (1, i, 0, 1) , v2 = (1, 0, i, 0) , v3 = (−i, 0, 1, −1) .
u u u ⟨v2 ,u1 ⟩
We want the set { ‖u1‖ , ‖u2‖ , ‖u3‖ } , where u1 = v1 , u2 = v2 − ⟨u1 ,u1 ⟩
u1 ,
1 2 3
⟨v3 ,u1 ⟩ ⟨v3 ,u2 ⟩
and u3 = v3 − ⟨u1 ,u1 ⟩
u1 − ⟨u2 ,u2 ⟩
u2
Now, ⟨v2 , u1 , ⟩ = ⟨v2 , v1 , ⟩ = 1 + 0 + 0 + 0 = 1.
Also ⟨u1 , u1 ⟩ = ⟨v1 , v1 ⟩ = 3, so that ‖u1 ‖ = √3.
1
∴ u2 = (1, 0, i, 0) − 3
(1, i, 0, 1) = ( 32 , − 3i , i, − 13 ) 99
Block
. . . . . . .5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .Inner
. . . . . . Products
. . . . . . . . . . .and
. . . . .Quadratic
. . . . . . . . . . .Forms
......
√15
∴ ‖u2 ‖ = 3
.
−1−i 1−5i 2
∴ u3 = (−i, 0, 1, −1) − 3
(1, i, 0, 1) − 5
( 3 , − 3i , i, − 13 )
= ( 51 , 2i
5
, − 5i , − 35 ).
√15
∴ ‖u3 ‖ = 5
.
1 3
∴ { (1, i, 0, 1) , ( 23 , − 3i , i, − 13 ) , 5
( 15 , 2i
5
, − 5i , − 35 )} is the required
√3 √15 √15
orthonormal basis.
E10) Let v1 = (1, i, 0) , v2 = (−i, 0, 2) , v3 = (0, −i, 2) .
u u u ⟨v2 ,u1 ⟩
We want the set { ‖u1‖ , ‖u2‖ , ‖u3‖ } , where u1 = v1 , u2 = v2 − ⟨u1 ,u1 ⟩
u1 ,
1 2 3
⟨v3 ,u1 ⟩ ⟨v3 ,u2 ⟩
and u3 = v3 − ⟨u1 ,u1 ⟩
u1 − ⟨u2 ,u2 ⟩
u2
Now, ⟨v2 , u1 , ⟩ = ⟨v2 , v1 , ⟩ = −i + 0 + 0 = −i.
Also ⟨u1 , u1 ⟩ = ⟨v1 , v1 ⟩ = 2, so that ‖u1 ‖ = √2.
−i
∴ u2 = (−i, 0, 2) − 2
(1, i, 0) = (− 2i , − 21 , 2)
3
∴ ‖u2 ‖ = .
√2
⟨v3 ,u1 ⟩ ⟨v3 ,u2 ⟩
∴ u3 = v3 − ⟨u1 ,u1 ⟩
u1 − ⟨u2 ,u2 ⟩
u2
= ( 49 + 4i 4
,
9 9
− 4i 2
,
9 9
− 2i
9
).
2√2
∴ ‖u3 ‖ = 3
.
1 √2
∴ { (1, i, 0) , 3
(− 2i , − 12 , 2) , 3
(4 + 4i 4
, − 4i 2
, − 2i
)} is the required
√2 2√2 9 9 9 9 9 9
orthonormal basis.
E11) We have for u, v ∈ ℂn , ⟨T𝛼 u, v⟩ = ⟨𝛼u, v⟩ = ⟨u, 𝛼v⟩ = ⟨u, T𝛼 v⟩. Since
⟨T𝛼 u, v⟩ = ⟨u, T∗𝛼 v⟩ it follows that T∗𝛼 v = 𝛼v. If T𝛼 is self adjoint, we have
T𝛼 v = T𝛼 v, i.e 𝛼v = 𝛼v for all v ∈ ℂn . If v ≠ 0, (𝛼 − 𝛼)v = 0 ⇔ 𝛼 − 𝛼 = 0. So,
T𝛼 = T𝛼 ⇔ 𝛼v = 𝛼v ⇔ 𝛼 − 𝛼 = 0 ⇔ 𝛼 is a real number
T𝛼 is unitary ⇔ T∗𝛼 ∘ T𝛼 = I
⇔ (T∗𝛼 ∘ T𝛼 )(v) = v
⇔ |𝛼|2 v = v ∀v ∈ ℂn
⇔ |𝛼| = 1
TT∗ = I, so T is unitary.
2 1 −2
b) A = [ 1 3 21 ]
−2 12 −1
1 1
x − x1
[ ] = [ √12 1
√2
][ ].
y y1
√2 √2
2 2 0
b) The matrix of the quadratic form is [2 2 0].
0 0 2
The eigenvalues are 4, 2, 0.
The required orthogonal canonical form is 4x21 + 2y21 .
1 1
0 −
√2 √2
The eigenvectors corresponding to 4, 2, 0 are [ 1 ] , [0] and [ 1 ].
√2 √2
0 1 0
∴, The orthogonal coordinate transformation is given by
1 1
x 0 − x1
√2 √2
[y] = [ 1 0 1
] [y1 ]
√2 √2
z 0 1 0 z1
E20) a) hyperbola
b) ellipse
c) parabola
d) circle
E21) a) The second degree terms give the quadratic form −x2 − y2 − 6xy. This
reduces to −4x21 + 2y21 . (After using the transformation as we have
done in E9).
2 2
∴, The given conic reduces to −4x1 + 2y1 = 8.
∴ The standard form is
y21 x21
− =1
22 (√2)2
x21 y21
+ =1
2 2
(√2) (√8)
Which is an ellipse with y-axis as major axis and x-axis as minor axis.
102