Intelligent Routines II - Solving Linear Algebra and Differential Geometry With Sage
Intelligent Routines II - Solving Linear Algebra and Differential Geometry With Sage
George A. Anastassiou
Iuliana F. Iatan
Intelligent
Routines II
Solving Linear Algebra and Differential
Geometry with Sage
Intelligent Systems Reference Library
Volume 58
Series editors
Janusz Kacprzyk, Polish Academy of Sciences, Warsaw, Poland
e-mail: [email protected]
The aim of this series is to publish a Reference Library, including novel advances
and developments in all aspects of Intelligent Systems in an easily accessible and
well structured form. The series includes reference works, handbooks, compendia,
textbooks, well-structured monographs, dictionaries, and encyclopedias. It con-
tains well integrated knowledge and current information in the field of Intelligent
Systems. The series covers the theory, applications, and design methods of
Intelligent Systems. Virtually all disciplines such as engineering, computer sci-
ence, avionics, business, e-commerce, environment, healthcare, physics and life
science are included.
George A. Anastassiou Iuliana F. Iatan
•
Intelligent Routines II
Solving Linear Algebra and Differential
Geometry with Sage
123
George A. Anastassiou Iuliana F. Iatan
Department of Mathematical Sciences Department of Mathematics and Computer
University of Memphis Science
Memphis Technical University of Civil Engineering
USA Bucharest
Romania
G. A. Anastassiou
G. A. Anastassiou
Virgil, Georgics
Molière, Le Misanthrope, I, 1
Boileau, Satires
German proverb
Preface
Linear algebra can be regarded as a theory of the vector spaces, because a vector
space is a set of objects or elements that can be added together and multiplied by
numbers (the result remaining an element of the set), so that the ordinary rules of
calculation are valid. An example of a vector space is the geometric vector space
(the free vector space), presented in the first chapter of the book, which plays a
central role in physics and technology and illustrates the importance of the vector
spaces and linear algebra for all practical applications.
Besides the notions which operates mathematics, created by abstraction from
environmental observation (for example, the geometric concepts) or quantitative
and qualitative research of the natural phenomena (for example, the notion of
number) in mathematics there are elements from other sciences. The notion of
vector from physics has been studied and developed creating vector calculus,
which became a useful tool for both mathematics and physics. All physical
quantities are represented by vectors (for example, the force and velocity).
A vector indicates a translation in the three-dimensional space; therefore we
study the basics of the three-dimensional Euclidean geometry: the points, the
straight lines and the planes, were in the second chapter.
The linear transformations are studied in the third chapter, because they are
compatible with the operations defined in a vector space and allow us to transfer
algebraic situations and related problems in three-dimensional space.
Matrix operations clearly reflect their similarity to the operations with linear
transformations; so the matrices can be used for the numerical representation of
the linear transformations. The matrix representation of linear transformations is
analogous to the representation of the vectors through n coordinates relative to a
basis.
The eigenvalue problems (also treated in the third chapter) are of great
importance in many branches of physics. They make it possible to find some
coordinate systems in which changes take the simplest forms. For example, in
mechanics the main moments of a solid body are found with the eigenvalues of a
symmetric matrix representing the vector tensor. The situation is similar in con-
tinuous mechanics, where the body rotations and deformations in the main
directions are found using the eigenvalues of a symmetric matrix. Eigenvalues
have a central importance in quantum mechanics, where the measured values of
the observable physical quantities appear as eigenvalues of operators. Also, the
vii
viii Preface
1 Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Geometric Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Free Vectors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Operations with Free Vectors . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.1 Addition of the Free Vectors . . . . . . . . . . . . . . . . . . . . 4
1.2.2 Scalar Multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.3 Vector Decomposition . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.2.4 Defining of the Products in the Set
of the Free Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3 Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.4 Vector Subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
1.5 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
ix
x Contents
2.4 Distances in E3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
2.4.1 Distance from a Point to a Straight Line . . . . . . . . . . . . 74
2.4.2 Distance from a Point to a Plane . . . . . . . . . . . . . . . . . 75
2.4.3 Distance Between Two Straight Lines . . . . . . . . . . . . . . 79
2.5 Angles in E3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
2.5.1 Angle Between Two Straight Lines . . . . . . . . . . . . . . . . 83
2.5.2 Angle Between Two Planes . . . . . . . . . . . . . . . . . . . . . 83
2.5.3 Angle Between a Straight Line and a Plane . . . . . . . . . . 84
2.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
3 Linear Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.1 Linear Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
3.2 Matrix as a Linear Mapping . . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.3 Changing the Associated Matrix to the Change of Basis . . . . . . 107
3.4 Eigenvalues and Eigenvectors . . . . . . . . . . . . . . . . . . . . . . . . . 110
3.4.1 Characteristic Polynomial of an Endomorphism . . . . . . . 111
3.4.2 Determining the Eigenvalues and the Eigenvectors
for an Endomorphism . . . . . . . . . . . . . . . . . . . . . . . . . 116
3.4.3 Diagonalization Algorithm of an Endomorphism . . . . . . 117
3.4.4 Jordan Canonical Form . . . . . . . . . . . . . . . . . . . . . . . . 122
3.5 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
Symbols
E3, 1
!
AB , 1
AB, 3
V3, 3
d (A, B), 13
a
b, 14
a b, 17
(Rn,?,•), 25
(Rn [X],?,•), 29
(Mm, n (R),?,•), 25, 26
xB, 36
MðB1 ;B2 Þ , 39
Msn (R), 46, 47
Man (R), 48, 49
, 52
n, 62, 63
L(U, V), 92
End(V), 92
Ker T, 92
Im T, 92
eðB1 ;B2 Þ , 97
T
P (k), 112
gk0 , 117
ak0 , 117
Jp (k), 123
Bp (k), 123
\x; y[, 135
kxk, 138
ru (x), 152
sO (x), 153
f (x), 166
b (x, y), 166
s, 199
xiii
xiv Symbols
b, 199
m, 199
pN, 200
pO, 200
pr, 200
U1, 223
dr, 229
Chapter 1
Vector Spaces
AB
A
−
→ −→
Definition 1.4 (see [1], p. 108). Two oriented segments AB and CD, A = B C = D
one call equipollent if they have the same direction, the same sense and the same
−
→ −→ −→ −→
norm; if AB is equipollent with CD we shall write AB ∼ CD.
1.1 Geometric Vector Spaces 3
Theorem 1.5 (see [1], p. 108). The equipollent relation defined on the set of the
oriented segments is an equivalence relation.
Proof
The equipollent relation is an equivalence relation since it is:
−→ − →
1. reflexive: AB ∼ AB;
−
→ −→ −→ − →
2. symmetric: AB ∼ CD involves CD ∼ AB;
−
→ −→ −→ − → −
→ − →
3. transitive: AB ∼ CD and CD ∼ EF involves AB ∼ EF.
We can classify the vectors in the following way:
(a) free vectors, which have the arbitrary origin at any point in space, but whose
direction, sense and length on space are prescribed;
(b) bound vectors, whose origin is prescribed;
(c) sliding vectors moving along the same support line, and their origin can be
anywhere on the line.
Definition 1.6 (see [4], p. 86). We call a free vector (geometric vector) characterized
−
→ −
→
by an oriented segment AB, the set of the oriented segments, equipollent with AB:
−→ −→ − →
AB = CD|CD ∼ AB .
Any oriented segment of this set is called the representative of the free vector
−→
AB; therefore CD ∈ AB.
A free vector of the length:
• 1 is called versor (unit vector); generally one denotes by e;
• 0 is called null vector; one denotes by 0.
Definition 1.7 (see [4], p. 86). The length, direction and sense of a free nonzero
vector means the length, direction and sense corresponding to the oriented segment
that it represents.
The set of the geometric vectors from the space E3 will be denote by V3 :
V3 = AB| A, B ∈ E3 ,
−
→
namely V3 means the set of equivalence
classes of the oriented segment AB.
We can use the notations: a AB or d (A, B) in order to designate the length
of a free vector a or AB.
Definition 1.8 (see [1], p. 109). We say that two free vectors are equal and we write
a = b if their representatives are equipollent.
Definition 1.9 (see [1], p. 109). Two free non-null vector a and b are collinear if
they have the same direction (see Fig. 1.3).
4 1 Vector Spaces
Definition 1.10 (see [1], p. 109). Two collinear vectors which have the same length
but they have opposite directions are called opposite vectors. The opposite of a free
vector a is −a (see Fig. 1.4).
Definition 1.11 (see [1], p. 109). Three free vectors a , b , c are called coplanar if
their support lines lie in the same plane (see Fig. 1.5).
Therefore, the sum of two or more vectors is also a vector, which can be obtained
through the following methods:
(A) if vectors are parallel or collinear and
(a) they have the same sense, then the sum vector has the direction and the
sense of the component vectors, and their length is equal to the sum of the
lengths corresponding to the component vectors;
(b) they have opposite sense, then the sum vector has the common direction,
the sense of the larger vector, and its magnitude is given by the differences
of the two vector magnitudes.
(B) if the vectors have only a common origin, then their sum is determined using
the parallelogram rule.
Definition 1.12 (the parallelogram rule, see [1], p. 110). Let a, b ∈ V3 be two
free vectors, which have a common origin and A ∈ E3 be an arbitrary fixed point.
−
→ −
→ −→
If OA ∈ a (OA means the representative of the free vector a ) and OC ∈ b then the
−
→
free vector c represented by the oriented segment OB is called the sum of the free
−→ − → −→
vectors a and b; one writes c = a + b or OB = OA + OC (Fig. 1.6).
(C) if the vectors are arranged so as one extremity to be the origin of the other, to
achieve their sum one applies the triangle rule.
Definition 1.13 (the triangle rule, see [1], p. 110). Let a, b ∈ V3 be two free vectors
−→ −
→
and A ∈ E3 an arbitrary fixed point. If AB ∈ a (AB is the representative of the free
−
→ −
→
vector a) and BC ∈ b then the free vector c represented by the oriented segment AC
−→ − → − →
is called the sum of the free vectors a and b; one writes c = a + b or AC = AB + BC
(Fig. 1.7).
Remark 1.14 (see [11], p. 118 and [10], p. 8). For many vectors, sitting in the same
way, one applies the polygon rule, which is a generalization of the triangle rule; the
sum vector is that which closes the polygon, joining the origin of the first component,
with the last extremity.
The addition of the vectors is based on some experimental facts (composition of
forces, velocities).
Example 1.15 (see [5], p. 58). We suppose a segment AB and the points M1 and M2
that divide the segment into three equal parts. If M is an arbitrary point outside the
segment, express the vectors MM1 and MM2 depending on the vectors MA = a and
MB = b.
Solution
Using the triangle rule we have:
MM1 = MA + AM1 .
Hence
AB = AM + MB = −a + b.
We deduce
1 b−a
AM1 = AB =
3 3
and
b−a 2a + b
MM1 = a + = .
3 3
Similarly,
2 b−a a + 2b
MM2 = MA + AM2 = MA + AB = a + 2 = .
3 3 3
1.2 Operations with Free Vectors 7
Theorem 1.16 (see [6], p. 7). The addition of the free vectors determines an abelian
group structure (V3 , +) on the set of the free vectors.
Proof
We note that the addition of the free vectors is an internal well-defined algebrical
operation, i.e. the free vector c = a + b doesn’t depend on the choice point A since
from AB = A B and BC = B C it results AC = A C .
We check the properties of (Fig. 1.8):
8 1 Vector Spaces
1. associativity:
a + b + c = a + b + c, (∀) a, b, c ∈ V3 .
⎫
a + b + c = OA + AB + BC = OB + BC = OC ⎬
=⇒ a + b + c = a + b + c.
a + b + c = OA + AB + BC = OA + AC = OC ⎭
Let OA = a, OO = 0, AA = 0.
We have:
OO + OA = OA ⇐⇒ 0 + a = a,
OA + AA = OA ⇐⇒ a + 0 = a.
3. simetrizable element:
a + (−a) = AB + BA = AA = 0,
(−a) + a = BA + AB = BB = 0.
Remark 1.17 (see [4], p. 88). The existence of an opposite for a free vector allows
the substraction definition of the free vectors a, b ∈ V3 (Fig. 1.9):
a − b = a + −b .
1.2 Operations with Free Vectors 9
4. commutativity:
a + b = b + a, (∀) a, b ∈ V3 .
OA + AB = OB,
OC + CB = OB.
• : R × V3 → V3 , (t, a) → ta,
Theorem 1.18 (see [6], p. 7). The multiplication of the free vectors with scalar has
the following properties:
1. distributivity of scalar multiplication with respect to the vector addition:
t a + b = ta + tb, (∀) t ∈ R, (∀) a, b ∈ V3 ;
4. 1 · a = a, (∀) a ∈ V3 .
Proposition 1.19
(the decomposition of a vector in a direction, see [7], p. 8). Let
be a, b ∈ V3 \ 0 . The vectors a and b are collinear if and only if (∃) t ∈ R unique
such that b = ta.
Theorem 1.20 (see [7], p. 8). Let be a, b ∈ V3 \ 0 . The vectors a and b are collinear
if and only if (∃) α, β ∈ R nonsimultaneous equals to zero (i.e. α2 + β 2 = 0) such
that αa + βb = 0.
The decomposition of a vector after two directions is the reverse operation to the
addition of the two vectors.
Proposition 1.21 (the decompositionofa vector after two noncollinear directions,
see [7], p. 9). Let be a, b, c ∈ V3 \ 0 . If a, b, c are coplanar then (∃) α, β ∈ R
uniquely determined such that c = αa + βb.
Theorem 1.22 (see [7], p. 10). Let be a, b, c ∈ V3 \ 0 . The vectors a, b, c are
coplanar if and only if (∃) α, β, γ ∈ R nonsimultaneous equal to zero (namely
α2 + β 2 + γ 2 = 0) such that αa + βb + γc = 0.
Proposition 1.23 (the decomposition of a vector after three noncoplanar direc-
tions, see [7], p. 10). Let be a, b, c, d ∈ V3 \ 0 . If a, b, c are noncoplanar then
(∃) α, β, γ ∈ R uniquely determined such that d = αa + βb + γc.
We suppose a point O in E3 called origin and three non-coplanar versors i j k whose
we attach the coordinate axes Ox, Oy, Oz that have the same sense as the sense of these
versors (see Fig. 1.10). The ensemble O, i, j, k is called the Cartesian reference
in E3 .
Whereas the versors i, j, k are non-coplanar, then under the Proposition 1.23, for
any vector v ∈ V3 (∃) r, s, t ∈ R uniquely determined such that v is expressed as
v = ri + sj + tk, called analytical expression of the vector v. The numbers (r, s, t)
are called the Euclidean coordinates (the components) of v relative to the reference
O, i, j, k .
Definition 1.24 (see [8], p. 449 and [10], p. 10) Let M ∈ E3 be a fixed point. The
vector OM is called the position vector ofthe pointM. The coordinates of the
position vector OM relative to the reference O, i, j, k are called the coordinates
of the point M. If OM = xi + yj + zk then one writes M (x, y, z).
Example 1.25 (see [9], p. 50 ). Find λ ∈ R such that the vectors
⎧
⎨ v1 = 2i + (λ + 2) j + 3k
v2 = i + λj − k
⎩
v3 = 4j + 2k
to be coplanar.
With this values of λ decompose the vector v1 after the directions of the vectors
v2 and v3 .
Solution
Using the Theorem 1.22, v1 v2 , v3 are coplanar if and only if (∃) α, β, γ ∈ R, α2 +
β + γ 2 = 0 such that αv1 + βv2 + γv3 = 0.
2
We obtain:
i.e. ⎧
⎨ 2α + β = 0
(λ + 2) α + λβ + 4γ = 0
⎩
3α − β + 2γ = 0;
v 1 = α v 2 + β v 3 ;
it results
12 1 Vector Spaces
2i − 6j + 3k = α i + −8α + 4β j + −α + 2β k;
therefore ⎧
⎨ α = 2
−8α + 4β = −6
⎩
−α + 2β = 3 ⇒ 2 + 2β − 3 ⇒ β = 25 .
We achieve:
5
v1 = 2v2 + v3 .
2
Using Sage we shall have:
Definition 1.26 (see [10], p. 12 ). If A (x1 , y1 , z1 ) , B (x2 , y2 , z2 ) are two given points
from E3 (see Fig. 1.11)
1.2 Operations with Free Vectors 13
then we have
and the distance from the points A and B denoted d (A, B) is calculated using the
formula:
d (A, B) = AB = (x2 − x1 )2 + (y2 − y1 )2 + (z2 − z1 )2 . (1.2)
−
→ −
→
Definition 1.27 (see [1], p. 115). Let be a, b ∈ V3 0 , O ∈ E3 and OA ∈ a, OB ∈ b.
−
→ −
→
The angle ϕ ∈ [0, π] determined by the oriented segments OA and OB is called the
angle between the free vectors a and b (see Fig. 1.12).
The free vectors a and b are called orthogonal if the angle between them is π2 .
Definition 1.28 (see [1], p. 116). The scalar (dot) product of the free vectors a
and b is the scalar a · b given by
a b cos ϕ, a = 0, b = 0
a·b= (1.3)
0, a = 0, b = 0.
One important physical application of the scalar product is the calculation of work.
−
→
The scalar product represents (see [11], p. 126) the work done by the force F
−
→
required to move a mobile to a straight line, with the director vector d, that makes
−
→
the angle ϕ with the direction of the force F .
Proposition 1.29 (see [7], p. 15). The scalar product of the free vectors has the
following properties:
1. commutativiy:
a · b = b · a, (∀) a, b ∈ V3 ;
2. t a · b = ta · b = a · tb, (∀) a, b ∈ V3 , (∀) t ∈ R;
3. distributivity of scalar multiplication with respect to the vector addition:
a · b + c = a · b + a · c, (∀) a, b, c ∈ V3 ;
a + b · c = a · b + b · c, (∀) a, b, c ∈ V3 ;
a · a > 0, (∀) a ∈ V3 \ 0 ;
4.
a · a = 0 ⇔ a = 0;
5. a · b = 0 ⇔ a and b are orthogonal, (∀) a, b ∈ V3 \ 0 ;
6. if a, b ∈ V3 ,
a = a1 i + a2 j + a3 k
b = b1 i + b2 j + b3 k
a · b = a1 b1 + a2 b2 + a3 b3 . (1.4)
Particularly,
a · a = a12 + a22 + a32 = a2 . (1.5)
7. the angle between the vectors a, b ∈ V3 \ 0 is given by the formula:
a·b a1 b1 + a2 b2 + a3 b3
cos ϕ = = , ϕ ∈ [0, π] ; (1.6)
a b a12 + a22 + a32 · b12 + b22 + b32
1.2 Operations with Free Vectors 15
one notices that the vectors and are orthogonal if and only if
a1 b1 + a2 b2 + a3 b3 = 0.
Example 1.30 (see [12], p. 161). Prove that the heights of a triangle are concurrent.
Solution
Let H be the intersection of the heights from A and B. We must show that CH is
perpendicular on AB, namely CH · AB = 0.
Using the triangle rule we have:
AB = AC − BC.
We shall obtain
CH · AB = CH · AC − BC = CH · AC − CH · BC
= − HB + BC · AC + HA + AC · BC,
i.e.
CH · AB = −HB · AC − BC · AC + HA · BC + AC · BC = −HB · AC + HA · BC = 0.
We have used the commutative property of the scalar product and the fact that BH
is perpendicular on AC and AH is perpendicular on BC.
We need the following Sage code to solve this problem:
Example 1.31 (see [12], p. 161). Prove that th diagonals in a rhombus are
perpendicular.
16 1 Vector Spaces
Solution
We shall deduce
(u − v) · (u + v) = u · (u + v) − v · (u + v) = u · u + u · v − v · u − v · v = u2 − v2 .
(u − v) · (u + v) = 0,
• Its sense is given by the right-hand rule: if the vector a × b is grasped in the right
hand and the fingers curl around from a to b through the angle ϕ, the thumb points
in the direction of a × b (Fig. 1.13).
Proposition 1.33 (see [2], p. 17). The algebraic properties of the cross product of
the free vectors are:
1. Anticommutativiy:
a × b = − b × a , (∀) a, b ∈ V3 ;
a×
4. a × a = 0, (∀) a ∈ V3 ;
5. a × 0 = 0 × a = 0, (∀) a ∈ V3 ;
6. if a, b ∈ V3 ,
a = a1 i + a2 j + a3 k
b = b1 i + b2 j + b3 k
Proposition 1.34 (see [12], p. 170 and [2], p. 17).The geometric properties of the
cross product of the free vectors are:
⎫
a · a × b = 0⎬
1. namely a × b is orthogonal both on a and b.
b · a × b = 0⎭
2. Lagrange identity:
2 2 2
a × b = a2 b − a · b , (∀) a, b ∈ V3 ;
3. a × b is the positive area of the parallelogram determined by the vectors a and
b, having the same origin (Fig. 1.14).
− → − → →
→ − → − −
→ −
AOBCA = OA OB sin ≺ OA, OB = OA × OB .
But
AOBCA = 2AπOAB .
It results that − →
→ −
OA × OB
AπOAB = .
2
i j k
AB × BC = 3 −4 0 = 19k
1 5 0
AB × BC AB CD AB × BC 19
AπABC = =
=⇒ CD = = .
2 2 AB 5
Example 1.36 (see [5], p. 65). Let the vectors a = 3m − n b = m + 3n be such that
m = 3 and n = 2 and ≺ (m, n) = π2 . Determine the area of the triangle formed
by the vectors a and b.
Solution
According to Proposition 1.34, the area of the triangle formed by the vectors a
and b is:
1
Aπ = a × b .
2
We have
We shall obtain
1 π
Aπ = · 10 · m × n = 5 m · n · sin = 30.
2 2
20 1 Vector Spaces
Definition 1.37 (see [2], p. 19). Let a, b, c ∈ V3 \ 0 be three free vectors. The
mixed product (also called
the scalar triple product or box product) of these free
vectors is the scalar a · b × c .
Remark 1.38 (see [1], p. 121). If the free vectors a, b, c ∈ V3 \ 0 are noncoplanar,
then the volume
of the parallelepiped
determined by these three vectors (see Fig. 1.15)
is given by a · b × c .
If we denote by:
• θ the angle between the vectors b and c,
• ϕ the angle between the vectors a and d = b × c, then
a· b × c = a·d = a d ·cos ϕ = b × c a cos ϕ = ±Ah = ±Vparalelipiped ,
±2h
A
i.e.
a · b × c = Vparalelipiped .
1.2 Operations with Free Vectors 21
Proposition 1.39 (see [2], p. 19 and [1], p. 121). The mixed product of the free
vectors has the following properties:
1. a · b × c = c · a × b = b · (c × a)
2. a · b × c = −a · c × b
3. ta · b × c = a · tb × c = a · b × tc , (∀) t ∈ R;
4. a + b · c × d = a · c × d + b · c × d
5. Lagrange identity:
a · c a · d
a × b · c × d =
b·c b·d
6. a · b × c = 0 if and only if:
⎧
⎨ a = a1 i + a2 j + a3 k
b = b1 i + b2 j + b3 k
⎩
c = c1 i + c2 j + c3 k
Example 1.40 (see [9], p. 52). Find λ ∈ R such that the volume of the parallelepiped
determined by the vectors a = 2i − 3j + k, b = i + j − 2k, c = λi + 2j be equal to
5.
Solution
The volume of the parallelepiped determined by these vectors is
2 −3 1
V = ± 1 1 −2 = ± (10 + 5λ) .
λ 2 0
Solution
(a)+(b) The points A, B, C, D are coplanar if the volume of the tetrahedron ABCD
is equal to 0.
We have
As
BC = i + j + 5k, BD = −5i + j − k
we shall have
i j k
BC × BD = 1 1 5 = −6i − 24j + 6k
−5 1 −1
and √ √
BC × BD = 36 + 576 + 36 = 18 2;
therefore
1 √ √
AπBCD = · 18 2 = 9 2.
2
It will result
30 10
d (A, (BCD)) = √ = √ .
9 2 3 2
(d) We have:
⎫
cos ≺ BAC = cos ≺ AB, AC = AB·AC
= √ 4√ ⎪
⎬
AB·AC 6· 29
=⇒
cos ≺ BAD = cos ≺ AB, AD = AB·AD
= √ 4√ ⎪
⎭
AB·AD 6· 29
Linear algebra can be regarded as the theory of the vector spaces, as a vector space
is a set of some objects or elements, that can be added together and multiplied by
the numbers (the result remaining an element of the set), so that the ordinary rules
of calculation to be valid.
An example of a vector space is the geometric vector space (the free vector
space), which plays a central role in physics and technology and illustrates the
importance of the vector spaces and linear algebra for all practical applications.
Let K be a commutative field and V be a non-empty set. The elements of K are
called scalars and we shall denote them by Greek letters, and the elements of V are
called vectors and we shall denote them by Latin letters, with bar above.
Definition 1.42 (see [14], p. 1). The set V is called a vector space over the field K
if the following are defined:
1. an internal algebraic operation, denoted additive “+”, + : V × V → V , called
addition, in respect to which V is a commutative group;
2. an external algebraic operation, denoted multiplicativ “ • ”, •: K × V → V ,
called multiplication by a scalar, that satisfies the axioms:
(a) (α+ β) a = αa + βa, (∀) α, β ∈ K and (∀) a ∈ V3
(b) α a + b = αa + αb, (∀) α ∈ K and (∀) a, b ∈ V3
(c) α (βa) = (αβ) a, (∀) α, β ∈ K and (∀) a ∈ V3
(d) 1 · a = a, (∀) a ∈ V3 .
1.3 Vector Spaces 25
If K is the field of the real numbers, V is called the real vector space. In the case
when K is the field of the complex numbers, V is called the complex vector space.
Examples of vector spaces
(1) The real arithmetic vector space with n - dimensions (Rn , +, ·)
Rn = R × R ×
. . . × R
n times
is the set of the ordered systems formed with n reale numbers, namely
Rn = x = x (1) , x (2) , . . . , x (n) |x (i) ∈ R, i = 1, n .
def
· : R × Rn → Rn , αx = αx (1) , αx (2) , . . . , αx (n) .
(2) The vector space of the polynomials in the indeterminate X with real coef-
ficients, of degree ≤ n, i.e. (Rn [X], +, ·)
Rn [X] means the set of the polynomials in the indeterminate X, with real coeffi-
cients, of degree ≤ n.
Let α ∈ R, P, Q ∈ Rn [X],
P (X) = a0 + a1 X + . . . + an X n , Q (X) = b0 + b1 X + . . . + bn X n ;
then
def
+ : Rn × Rn → Rn , P (X) + Q (X) = a0 + b0 + (a1 + b1 ) X + . . . + (an + bn ) X n
(1.10)
def
· : R × Rn → Rn , αP (X) = αa0 + αa1 X + . . . + αan X n . (1.11)
(3) The vector space (Mm,n (R), +, ·) of the m×n matrices, with real coefficients.
Example 1.43 (see [13]). Show that the set of the matrices of real numbers with m
lines and n columns forms a vector space on R, toward the addition of the matrices
and the scalar multiplication from R.
Solution
Stage I. One proves that (Mm,n (R), +, ·) is a commutative group (abelian group).
If A, B ∈Mm,n (R) then A + B ∈Mm,n (R), i.e. Mm,n (R) is a stable part in relation
to the addition of the matrices (the addition is well defined). Since are easily to check
the axioms on:
26 1 Vector Spaces
• Associativity:
def
· : R × F → F, (αf ) (x) = αf (x) .
Theorem 1.44 (see [6], p. 9). If V is a real vector space then the following statements
occur:
(i) 0 · a = 0 (∀) a ∈ V
(ii) α · 0 = 0 (∀) α ∈ R
(iii) (−1) · a = −a (∀) a ∈ V
(iv) if (∀) α ∈ K, (∀) a ∈ V such that α · a = 0 then α = 0 or a = 0.
1.3 Vector Spaces 27
Definition 1.45 (see [14], p. 6). A vector system {x 1 , . . . , x m } from the space vector V
over K is linearly dependent if there are the scalars α(i) ∈ K, α(i) = 0, (∀) i = 1, m,
such that
α(1) x 1 + . . . + α(m) x m = 0.
α(1) = . . . = α(m) = 0
⎦
m
v= α(i) x i , x i ∈ S, α(i) ∈ K, i = 1, m.
i=1
Definition 1.47 (see [14], p. 6). Let V be a vector space over the field K. The finite
system S of vectors from V is called system of generators for V if any vector from
V is a linear combination of vectors from S.
Definition 1.48 (see [6], p. 12). Let V be a space vector over the field K. The finite
system B of vectors from V is called basis of V if:
(a) B is linearly independent;
(b) B is a system of generators for V .
Example 1.49 (see [13]). B = Eij , i = 1, m, j = 1, n is a basis in Mm,n (R).
Solution
We consider ⎛ ⎞
a11 · · · a1n
⎜ .. .. ⎟ .
A ∈ Mm,n (R), A = ⎝ . . ⎠
am1 · · · amn
We can write
⎛ ⎞ ⎛ ⎞
a11 0 · · · 0 0 0 a12 · · · 0 0
⎜ 0 0 ··· 0 0 ⎟ ⎜ 0 0 ··· 0 0 ⎟
A=⎜ ⎟ ⎜ ⎟
⎝ ··· ··· ··· ··· ···⎠ + ⎝··· ··· ··· ··· ···⎠ + ··· +
0 0 ··· 0 0 0 0 ··· 0 0
⎛ ⎞ ⎛ ⎞
0 0 · · · 0 a1n 0 0 ··· 0 0
⎜ 0 0 ··· 0 0 ⎟ ⎜ ⎟
⎜ ⎟ + ··· + ⎜ 0 0 ··· 0 0 ⎟
⎝··· ··· ··· ··· ··· ⎠ ⎝··· ··· ··· ··· ··· ⎠
0 0 ··· 0 0 0 0 · · · 0 amn
28 1 Vector Spaces
⎛ ⎞ ⎛ ⎞
1 0 ··· 0 0 0 1 ··· 0 0
⎜ 0 0 ··· 0 0 ⎟ ⎜ 0 0 ··· 0 0 ⎟
= a11 ⎜
⎝···
⎟ + a12 ⎜ ⎟
··· ··· ··· ···⎠ ⎝··· ··· ··· ··· ···⎠
0 0 ··· 0 0 0 0 ··· 0 0
⎛ ⎞
0 0 ··· 0 0
⎜ 0 0 ··· 0 0 ⎟
+ · · · + amn ⎜ ⎟
⎝··· ··· ··· ··· ···⎠.
0 0 ··· 0 1
We denote
⎛ j ⎞
0 ··· 0
⎜ .. .. ⎟.
Eij = i⎝ . 1 .⎠
0 ··· 0
namely
B = Eij , i = 1, m, j = 1, n
is a system of generators.
We consider the linear null combination of the matrices from B:
⎛ ⎞
a11 · · · a1n
⎦
m ⎦
n
⎜ .. ⎟ = O ⇐⇒ a = 0,
αij Eij = O ⇐⇒ ⎝ ... . ⎠ ij (∀) i = 1, m, j = 1, n;
i=1 j=1 am1 · · · amn
therefore
B = Eij , i = 1, m, j = 1, n
is linearly independent.
It results that
B = Eij , i = 1, m, j = 1, n
A solution in Sage will be given to prove that, in the case of M2,3 (R):
Solution
(i) Let be α ∈ R, P, Q ∈ Rn [X], P (X) = a0 + a1 X + . . . + an X n , Q (X) =
b0 + b1 X + . . . + bn X n .
Stage I. It proves that (Rn [X], +, ·) is an abelian group, where “+” and “·” are
defined in (1.10) and respectively (1.11).
As
deg ree (P + Q) ≤ max (deg ree (P) , deg ree (Q)) ≤ n
it results
P + Q ∈ Rn [X].
(ii) We denote ⎧
⎪
⎪ P0 (X) = 1
⎪
⎨ P1 (X) = X
..
⎪
⎪ .
⎪
⎩
Pn (X) = X n .
As
α(0) + α(1) X + . . . + α(n) X n = ORn [X]
Let be f , g ∈ V , f (x) = α sin (x + β) , g (x) = α sin x + β . We shall check
if (V , +) is a commutative group.
We show only that f + g ∈ V , namely (∃) α0 , β0 ∈ R such that
i.e.
We denote:
A = α cos β + α cos β
B = α sin β + α sin β .
We deduce
A = α0 cos β0
B = α0 sin β0 .
It results that
sin β0 B B B
= ⇒ tg β0 = ⇒ β0 = arctg + kπ, k ∈ Z,
cos β0 A A A
A
α0 = .
cos β0
We have
It is known that
π π
cos x = sin − x = − sin x − .
2 2
We denote π
f1 (x) = sin x, f2 (x) = − sin x − .
2
Hence
f = αf1 cos β + αf2 sin β.
We prove that f1 , f2 ∈ V .
As
sin x
= 1 · sin (x + 0) ⇒
f1 ∈ V , for α = 1, β = 0;
f1 (x) =
f2 (x) = − sin x − π2 = (−1) sin x − π2 ⇒ f1 ∈ V , for α = −1, β = − π2 .
it results that
α(1) sin x + α(2) cos x = 0, (∀) x ∈ R.
Solution
(i) Check the axioms from the definition of the vector space.
(ii) Let be B = {a1 , a2 , . . . , an } a basis for V . We intend to show that B =
{ia1 , ia2 , . . . , ian } a basis for CV .
Let be (u1 , u2 ) ∈ CV , u1 , u2 ∈ V .
As B is a basis for V it results that
• (∀) u1 ∈ V , (∃) x (i) ∈ K, i = 1, n unique, such that
u1 = x (1) a1 + . . . + x (n) an ;
u2 = y(1) a1 + . . . + y(n) an .
1.3 Vector Spaces 35
We compute
(u1 , u2 ) = x (1) a1 + . . . + x (n) an , y(1) a1 + . . . + y(n) an
= x (1) a1 , y(1) a1 + . . . + x (n) an , y(n) an
It results that B = 0, a1 , . . . , 0, an is a system of generators for C V .
Let be a null combination of elements from B :
namely
− β (1) a1 − . . . − β (n) an = 0 (1.13)
and
α(1) a1 + . . . + α(n) an = 0. (1.14)
Proposition 1.56 (see [1], p. 12). Let V be a space vector of n dimension and
B = {a1 , . . . , an } ⊂ V . Then:
(a) if B is liniarly independent it results that B is a basis,
(b) if B is system of generators it results that B is a basis.
Theorem 1.57 (see [14], p. 7). Let V be a space vector over K and B = {a1 , . . . , an } ⊂
V . Then B is a basis of V if and only if any vectors from V can be written in an unique
way as a linear combination of the vectors from B.
Definition 1.58 (see [14], p. 7). The unique scalars x (i) ∈ K, i = 1, m that appear as
coefficients in the writing of the vector x ∈ V as a linear combination of the vectors
from the basis B are called the coordinates of the vector x relative to the basis B.
We shall denote by x B the column matrix formed with the coordinates of the vector
x relative to the basis B; therefore
⎛ ⎞
x (1)
⎜ x (2) ⎟
⎜ ⎟
xB = ⎜ . ⎟ .
⎝ .. ⎠
x (m)
Remark 1.59 (see [6], p. 14). The writing of a vector in a basis is unique.
Example 1.60 (see [9], p. 31). In R4 the following vectors are given:
Solution
(i) dim R4 = 4 ⇒ it suffices to show that the vectors are linearly independent.
Let be the null linear combination
It results that
α(1) , α(1) , 2α(1) , α(1) + α(2) , −α(2) , 0, α(2) + 0, 0, −α(3) , α(3) +
α(4) , 2α(4) , 2α(4) , 0 = (0, 0, 0, 0) .
As
1 1 0 1
1 −1 0 2
d = = −4 = 0
2 0 −1 2
1 1 1 0
We have
(1, 1, 1, 1) = α(1) , α(1) , 2α(1) , α(1) + α(2) , −α(2) , 0, α(2)
+ 0, 0, −α(3) , α(3) + α(4) , 2α(4) , 2α(4) , 0 .
38 1 Vector Spaces
whose solution is
1 (2) 1 1 1
α(1) = , α = , α(3) = , α(4) = .
4 4 2 2
Solving this problem with Sage, we achieve:
Hence ⎛ ⎞
1/4
⎜ 1/4 ⎟
vB = ⎜ ⎟
⎝ 1/2 ⎠ .
1/2
Let B1 = {e1 , . . . , en } , B1 = f 1 , . . . , f n be two bases of V x ∈ V and
We suppose that the vectors from B2 can be written as a linear combination of the
vectors from the basis B1 :
(1) (n)
f 1 = α1 e1 + . . . + α1 en (1.18)
..
.
f n = αn(1) e1 + . . . + αn(n) en .
1.3 Vector Spaces 39
Writing the column coefficients of these linear combinations we get (see [14],
p. 8) the matrix ⎛ ⎞
(1) (1)
α1 . . . αn
⎜ (2) ⎟
⎜ α . . . αn(2) ⎟
M(B1 ,B2 ) = ⎜ 1 ⎟, (1.19)
⎝ ... ... ... ⎠
(n) (n)
α1 . . . αn
which represents the transition matrix from the basis B1 to the basis B2 .
Remark 1.61 (see [6], p. 18). The matrix M(B1 ,B2 ) is always a nonsingular matrix
due to the linear independence of the basis vectors.
Example 1.62 (see [13]). Find for the space of polynomials byat most four degree,
the transition matrix from the basis B1 = 1, X, X 2 , X 3 , X 4 to the basis B2 =
1, (X + 1) , (X + 1)2 , (X + 1)3 , (X + 1)4 .
Solution
The vectors from B2 can be written as a linear combination of the vectors from
B1 thus:
1=1
X +1 = 1+X
(X + 1)2 = 1 + 2X + X 2
(X + 1)3 = 1 + 3X + 3X 2 + X 3
(X + 1)4 = 1 + 4X + 6X 2 + 4X 3 + X 4 .
It results that ⎛ ⎞
1 1 1 1 1
⎜0 1 2 3 4⎟
⎜ ⎟
M(B1 ,B2 ) =⎜
⎜0 0 1 3 6⎟⎟.
⎝0 0 0 1 4⎠
0 0 0 0 1
Due to the uniqueness of writing a vector into a basis, from (1.16) and (1.20) it
results (see [6], p. 18)
⎧ (1) (1) (1) (n) (1)
⎨ x = y α1 + . . . + y αn
⎪
.. (1.21)
⎪ .
⎩ (n) (n) (n)
x = y α1 + . . . + y(n) αn .
(1)
The formulas (1.21) are (see [6], p. 18) the formulas of changing a vector
coordinates when the basis of a vector space one changes.
The relations (1.21) can be written in the matrix form as:
or
x B2 = M−1
(B1 ,B2 ) · x B1 . (1.23)
Example 1.63 (see [13]). In the arithmetic vector space R3 the following vectors are
considered:
Solution
(a) dim R3 = 3 ⇒ it is enough to prove that the vectors are linearly independent.
Consider the null linear combination
1.3 Vector Spaces 41
As
2 1 0
d = −1 −1 3 = −8 = 0
2 2 2
We have
(−1, 2, 3) = 2α(1) , −α(1) , 2α(1) + α(2) , −α(2) , 2α(2) + 0, 3α(3) , 2α(3) .
whose solution is
13 (2) 9 7
α(1) = − , α = , α(3) = .
8 4 8
Hence ⎛ ⎞
−13/8
x B = ⎝ 9/4 ⎠ .
7/8
We know that ⎛ ⎞
(1) (1) (1)
α1 α2 α3
⎜ ⎟
M(B1 ,B2 ) = ⎝ α1(2) α2(2) α3(2) ⎠ .
(3) (3) (3)
α1 α2 α3
We can write
(1) (2) (3)
b1 = α1 a1 + α1 a2 + α1 a3 (1.24)
One obtains ⎛ ⎞
0.625 1.875 −0.875
M(B1 ,B2 ) = ⎝ −1.25 −1.75 0.75 ⎠ .
0.125 0.375 0.625
(d) Let
vB1 = x (1) a1 + x (2) a2 + x (3) a3
We shall have
⎛ (1) ⎞ ⎛ (1) ⎞ ⎧ (1)
x y ⎨ x = 0.625y(1) + 1.875y(2) − 0.875y(3)
⎝ x (2) ⎠ = M(B1 ,B2 ) · ⎝ y(2) ⎠ ⇒ x (2) = −1.25y(1) − 1.75y(2) + 0.75y(3)
(3) (3) ⎩ (3)
x y x = 0.125y(1) + 0.375y(2) + 0.625y(3) .
Example 1.64. In the vector space M2,2 (R) the following matrices are considered:
!
" ! " ! " ! "
10 1 −1 1 −1 1 −1
C1 = , C2 = , C3 = , C4 = ,
00 0 0 1 0 1 −1
! " ! " ! " ! "
00 0 0 0 1 −1 −1
A1 = , A2 = , A3 = , A4 = .
01 −1 1 −1 1 −1 1
One requires:
(a) Prove that both B1 = {C1 , C2 , C3 , C4 } and B2 = {A1 , A2 , A3 , A4 } are bases for
M2,2 (R).
(b) Find the transition matrix from the basis B1 to the basis B2 .
(c) Write the formulas of changing a vector coordinates when one passes from the
basis B1 to the basis B2 .
Solution
(a) Let
We have
! " ! " ! " ! " ! "
00 0 0 0 1 −1 −1 00
α(1) + α(2) + α(3) + α(4) = ⇔
01 −1 1 −1 1 −1 1 00
! " ! " ! " ! " ! "
0 0 0 0 0 α(3) −α(4) −α(4) 00
+ + + = .
0 α(1) −α(2) α(2) −α(3) α(3) −α(4) α(4) 00
(c) Let
Definition 1.65 (see [14], p. 8). Let V be a vector space over the n -dimensional
K field. The nonempty subset W of V it’s called a vector subspace of V if the
following conditions are satisfied:
(1) x + y ∈ W (∀) x, y ∈ W
(2) αx ∈ W (∀) α ∈ K, (∀) x ∈ W .
Proposition 1.66 (see [14], p. 8). Let V be a vector space over the n-dimensional K
field. The nonempty subset W of V it’s called a vector subspace of V if and only if
αx + βy ∈ W , (∀)α, β ∈ K, (∀)x, y ∈ W .
We can write
⎛ ⎞ ⎛ ⎞
1 0 ... 0 0 0 1 ... 0 0
⎜ 0 0 ... 0 0 ⎟ ⎜ 1 0 ... 0 0 ⎟
A = a11 ⎜
⎝... ...
⎟ + a12 ⎜ ⎟
... ... ...⎠ ⎝... ... ... ... ...⎠
0 0 ... 0 0 0 0 ... 0 0
=F1 =F2
⎛ ⎞
0 0 ... 0 1 ⎛ ⎞
0 0 ... 0 0
⎜ 0 0 ... 0 0 ⎟ ... 0 0 ⎠
+ . . . + a1n ⎜
⎝... ...
⎟ + a22 ⎝ 0 1
... ... ...⎠ ... ... ... ... ...
1 0 ... 0 0 0 0 ... 0 0
=Fn+1
=Fn
⎛ ⎞
0 0 ... 0 0
+ . . . + ann ⎝ . 0. . . 0. . . . . 0 0 ⎠.
... ... ...
0 0 ... 0 1
=Fn(n+1)/2
n (n + 1)
dim Mns (R) = .
2
We can check that in Sage, too:
(3) Mna (R) = A ∈ Mn (R)|At = −A the vector subspace of the antisymmetric
matrices is a vector subspace of Mn (R).
We shall prove that the subset Mna (R) is a vector subspace of Mn (R) and we shall
determine its dimension, highlighting a basis of its.
Let α, β ∈ R and A, B ∈ Mna (R). It results that At = −A and Bt = −B.
We shall have
We can write
⎛ ⎞ ⎛ ⎞
0 1 ... 0 0 0 0 ... 0 1
⎜ −1 0 ... 0 0 ⎟ ⎜ ... 0 ⎟
⎜
A = a12 ⎝ ⎟ + . . . + a1n ⎜ 0 0 0 ⎟
... ... ... ... ...⎠ ⎝... ... ... ... ...⎠
0 0 ... 0 0 1 0 ... 0 0
=G1 =Gn−1
⎛ ⎞
0 0 ... 0 0
⎜ 0 0 ... 0 0 ⎟
+ . . . + ann ⎜
⎝...
⎟.
... ... ... ...⎠
0 0 ... 0 1
=Gn(n−1)/2
n (n − 1)
dim Mna (R) = .
2
Using Sage we shall have:
50 1 Vector Spaces
Proposition 1.67 (see [6], p. 20). The set of the solutions corresponding to a linear
and homogeneous system with m equations and n unknowns is a vector subspace of
Rn , having the dimension n − r , r being the rank of the associated matrix system A.
Example 1.68 (see [13]). In the vector space R5 we consider
W = x = (x1 , x2 , x3 , x4 , x5 ) ∈ R5
and
⎛ ⎞
111
π = ⎝ 0 1 2 ⎠ = −3 = 0.
120
1.4 Vector Subspaces 51
We denote
x3 = t, x4 = u, t, u ∈ R;
we obtain
−x2 + x5 = −t + u
x2 + 2x5 = t − u,
so we deduce that x5 = 0.
Finally, we get the solution of the system
⎧
⎪ x1 = −2t + 2u
⎪
⎪
⎪
⎨ x2 = t − u
x3 = t , (∀) t, u ∈ R.
⎪
⎪
⎪
⎪ x4 = u
⎩
x5 = 0
For
• x3 = 0 and x4 = 1 we obtain a1 = (2, −1, 0, 1, 0) ;
• x3 = 1 and x4 = 0 we obtain a2 = (−2, 1, 1, 0, 0)
and B1 = {a1 , a2 } is a basis of W .
We shall give the solution in Sage, too:
Proposition 1.69 (see [14], p. 8 and [6], p. 22). Let V be a vector space over the field
K and W1 W2 be two of its vector subspaces. Then W1 ∩ W2 is a vector subspace
of V .
52 1 Vector Spaces
Definition 1.70 (see [14], p. 8 and [6], p. 22). Let V be a vector space over the field
K and W1 W2 be two of its vector subspaces. The sum of two vector subspaces W1
and W2 is defined as:
W1 + W2 = {x 1 + x 2 |x 1 ∈ W1 and x 2 ∈ W2 } .
Proposition 1.71 (see [14], p. 8 and [6], p. 22). Let V be a vector space over the field
K and W1 W2 be two of its vector subspaces. Then W1 + W2 is a vector subspace
of V .
Theorem 1.72 (Grassmann- the dimension formula, see [14], p. 8 and [6],
p. 22). Let V be a finite n dimensional vector space over the field K, and W1 W2
be two of its vector subspaces. Then there is the relation
Definition 1.73 (see [14], p. 8 and [6], p. 23). Let V be a n dimensional finite vector
space over the field K, and W1 W2 be two of its vector subspaces. The sum of the
vector subspaces W1 and W2 is called the direct sum, denoted by W1 ⊕ W2 if any
vector from W1 + W2 can be uniquely written in the form:
x = x 1 + x 2 , x 1 ∈ W1 , x 2 ∈ W2 .
Proposition 1.74 (see [14], p. 9 and [6], p. 23). Let V be a n dimensional finite vector
space over the field K and W1 W2 be two of its vector subspaces. Then V = W1 ⊕ W2
if and only if the following conditions are satisfied:
(1) W1 ∩ W2 = 0
(2) dimW1 + dimW2 = dimV .
Proposition 1.75 (see [6], p. 21). Let V be a n dimensional finite vector space over
the field K. If W is a vector space of V , then the dimension of W is finite and
dimW ≤ dimV .
Proposition 1.76 (see [6], p. 22). Let V be a n dimensional finite vector space over
the field K. If W is a vector space of V , then dimW = dimV if and only if W = V .
Example 1.77 (see [13]). In the vector space R3 [X] we consider
Solution
(a) Let be α, β ∈ K and P, Q ∈ W1 . It results that P(1) = 0 and Q(1) = 0.
1.4 Vector Subspaces 53
We shall have
⇒ P = a0 (X − 1) + a1 X (X − 1) + a2 X 2 (X − 1) = a0 P1 + a1 P2 + a2 P3 ;
=P1 =P2 =P3
α1 P1 + α2 P2 + α3 P3 = 0 ⇒ α1 (X − 1) + α2 X (X − 1) + α3 X 2 (X − 1) = 0.
1 = t · P1
P = t ·
=P1
X − 1|P ⇒ P = (X − 1) Q, degree Q ≤ 2
• P ∈ W2 ; it results degree P = 0.
We deduce
W1 ∩ W2 = {P0 } .
It is noticed that
1.5 Problems
Solution
We shall present the solution in Sage:
2. Show that the points A(3, −1, 1) B(4, 1, 4) and C(6, 0, 4) are the vertices of a
right triangle.
Solution
Using Sage we shall have:
v1 = 2i + j − k, v2 = 3i + 2j + k, v3 = −j + 2k.
Solution
The solution in Sage will be given:
Solution
The answer of this question will be find in Sage:
1.5 Problems 57
8. Establish the transformation formulas of the coordinates when passing from the
basis B to the basis B , if
References
OM = xi + y j + zk
then M (x, y, z) .
We denote
r 0 = OM 0 = x0 i + y0 j + z 0 k.
r = OM = xi + y j + zk.
M0 M × v = 0. (2.1)
However
M0 M = r − r 0 . (2.2)
From (2.1) and (2.2) it results (see [6]) the vector equation of the straight line
d = (M0 , v) :
(r − r 0 ) × v = 0;
M0 M = tv.
Taking account the relation (2.2) it results that r − r 0 = tv; we obtain (see [6])
the vector parametric equation of the straight line d:
r − r 0 = tv, t ∈ R. (2.3)
xi + y j + zk = x0 i + y0 j + z 0 k + tai + tb j + tck;
we deduce the parametric equations (see [4], p. 49) of the straight line d:
2.1 Equations of a Straight Line in E3 61
x = x0 + ta
y = y0 + tb , t ∈ R. (2.4)
z = z 0 + tc
If in the relation (2.4) we eliminate the parameter t we obtain (see [4], p. 49) the
Cartesian equations of the straight line d:
x − x0 y − y0 z − z0
= = . (2.5)
a b c
In the relation (2.5), there is the following convention: if one of the denominators
is equal to 0, then we shall also cancel that numerator.
Let be:
• M1 ∈E3 , M1 (x1 , y1 , z 1 ) , r 1 = OM 1 and
• M2 ∈E3 , M2 (x2 , y2 , z 2 ) , r 2 = OM 2 .
We want to determine the equation of straight line determined by the points M1
and M2 , denoted by d = (M1 , M2 ) and represented in Fig. 2.2.
M1 M × M1 M 2 = 0
it results that vector equation (see [6]) of the straight line d is:
(r − r 1 ) × (r 2 − r 1 ) = 0.
Since (within the Proposition 1.1 from the Chap. 1), (∃) t ∈ R unique, such that
M1 M = t M1 M 2
62 2 Plane and Straight Line in E3
it results that
r − r 1 = t (r 2 − r 1 ) , (∀) t ∈ R,
r = r 1 + t (r 2 − r 1 ) , (∀) t ∈ R; (2.6)
x − x1 y − y1 z − z1
= = . (2.8)
x2 − x1 y2 − y1 z2 − z1
2.2 Plane in E3
A plane can be determined (see [6]) in E3 as follows:
(1) a point and a nonnull vector normal to plane,
(2) a point and two noncollinear vectors,
(3) three non collinear points,
(4) a straight line and a point that does not belong to the straight line,
(5) two concurrent straight lines,
(6) two parallel straight lines.
Definition 2.2 (see [6]). The straight line d which passes through M0 and has the
direction of the vector n is called the normal to the plane through M0 ; the vector n
is the normal vector of the plane.
We propose to obtain the plane equation determined by the point M0 and by the
vector n, denoted with π = (M0 , n).
A point M (x, y, z) ∈ π → M0 M and n are orthogonal.
We denote
r = O M = xi + y j + zk.
As n is perpendicular on M0 M it results
M0 M · n = 0,
namely
(r − r 0 ) · n = 0;
from here we deduce the normal equation (see [6]) of the plane π :
r · n − r 0 · n = 0. (2.9)
ax + by + cz − ax0 − by0 − cz 0 = 0
a (x − x0 ) + b (y − y0 ) + c (z − z 0 ) = 0. (2.10)
If we denote
ax0 − by0 − cz 0 = −d
then from the equation (2.10) one deduces the general Carthesian equation (see [6])
of the plane π :
ax + by + cz + d = 0. (2.11)
u = l1 i + m 1 j + n 1 k, v = l2 i + m 2 j + n 2 k
We want to find the equation of the plane determined by the point M0 and by the
free vectors u and v, denoted by π = (M0 , u, v).
Let
−−−∀
• M0 M 1 be a representative for the free vector u,
−−−∀
• M0 M 2 be a representative for the free vector v.
A point M (x, y, z) ∈ π → M0 M, M0 M 1 , M0 M 2 are coplanar. The coplanarity
of these vectors can be expressed as:
(a) using the Proposition 1.21 from the Chap. 1, (∃) t1 , t2 ∈ R uniquely determined,
such that
M 0 M = t1 M 0 M 1 + t2 M 0 M 2 ; (2.12)
M0 M · (u × v) = 0. (2.13)
r − r 0 = t1 u + t2 v
r = r 0 + t1 u + t2 v, (∀) t1 , t2 ∈ R (2.14)
From the relation (2.13) we obtain the vector equation (see [6]) of the plane π :
(r − r 0 ) · (u × v) = 0. (2.16)
2.2 Plane in E3 65
As
M0 M = (x − x0 ) i + (y − y0 ) j + (z − z 0 ) k (2.17)
we have
x − x0 y − y0 z − z 0
M0 M · (u × v) = l1 m1 n 1 ;
l2 m2 n2
so from the relation (2.13) we deduce the Carthesian equations (see [6]) of the
plane π :
x − x0 y − y0 z − z 0
l1 m1 n 1 = 0. (2.18)
l2 m2 n2
We note that π = (M0 , M1 , M2 ) coincides with π1 = M0 , M0 M 1 , M0 M 2 ,
namely we are in the case presented in the previous paragraph. We have
M0 M 1 = (x1 − x0 ) i + (y1 − y0 ) j + (z 1 − z 0 ) k,
M0 M 2 = (x2 − x0 ) i + (y2 − y0 ) j + (z 2 − z 0 ) k.
Using M0 M from (2.17) we obtain the following Cartesian equation (see [6]) of
the plane π :
66 2 Plane and Straight Line in E3
x − x0 y − y0 z − z 0
x1 − x0 y1 − y0 z 1 − z 0 = 0. (2.19)
x2 − x0 y2 − y0 z 2 − z 0
where a = a1 i + a2 j + a3 k and A (x A , y A , z A ) .
We want to find the equation of the plane determined by the straight lines d1
and d2 .
Noting that π = (d1 , d2 ) coincides with π = (P, a 1 , a 2 ), i.e. with the plane
which passes through P and has the direction vectors a 1 and a 2 . If M (x, y, z) ∈ π
we deduce that the vector equation (see [6]) of the plane is:
(r − r P ) · (a 1 × a 2 ) = 0; (2.22)
where: a 1 = l1 i + m 1 j + n 1 k, a 2 = l2 i + m 2 j + n 2 k, P (x P , y P , z P ) .
Example 2.3 Check if the following straight lines are concurrent:
y−7
d1 : x−1
2 = 1 = z−5
4
y+1
d2 : x−6
3 = −2 = z
1
and then write the equation of the plane which they determine.
Solution
We note that the direction vectors of the two straight lines are: a 1 = 2i + j + 4k
and respectively a 2 = 3i − 2 j + k. As
i j k
a 1 × a 2 = 2 1 4 = 9i + 10 j − 7k = 0
3 −2 1
π : 9 (x + 3) + 10 (y − 5) − 7 (z + 3) = 0,
namely
π : 9x + 10y − 7z − 44 = 0.
• d2 is the straight line which passes through A2 and has the direction vector a,
d2 = (A2 , a) .
The plane determined by d1 and d2 is the plane determined by A1 and the two
non collinear vectors a and A1 A2 .
If M (x, y, z) ∈ π then the vector equation (see [6]) of the plane π is:
r − r A1 · a × A1 A2 = 0. (2.24)
where: a = a1 i + a2 j + a3 k, A1 x A1 , y A1 , z A1 , A2 x A2 , y A2 , z A2 .
The intersection of the planes π1 and π2 is the set of solutions of the system of
equations determined by the equations of π1 and π2 .
We denote
⎬ ⎭
a1 b1 c1
A= .
a2 b2 c2
We denote
u = n 1 × n 2 , u = li + m j + nk.
We have
i j k
n 1 × n 2 = a1 b1 c1 = (b1 c2 − b2 c1 ) i + (a2 c1 − a1 c2 ) j + (a1 b2 − a2 b1 ) k.
a2 b2 c2
We deduce
⎨ b1 c1
⎨
⎨ l =
⎨
⎨ b2 c2
⎨
⎨
⎨
⎨
c1 a1
m=
⎨
⎨ c2 a2
⎨
⎨
⎨
⎨
⎨
⎨ a1 b1
⎨
n= .
a2 b2
x − x0 y − y0 z − z0
= = , (2.26)
l m n
Solution
(a) The vector equation of the plane is
π : (r − r M ) · (v 1 × v 2 ) = 0.
Where as
i j k
• v 1 × v 2 = 1 −2 1 = −10i − j + 8k,
3 2 4
• r − r M = (x + 2) i + (y − 3) j + (z − 4) k
we obtain
π : −10 (x + 2) − (y − 3) + 8 (z − 4) = 0
or
π : −10x − y + 8z − 49 = 0.
π : (−2) (x − 1) + (y + 1) + 3 (z − 1) = 0 → π : −2x + y + 3z = 0.
Let be d = π1 ∃ π2 ⎫
π1 : a1 x + b1 y + c1 z + d1 = 0
π2 : a2 x + b2 y + c2 z + d2 = 0.
Definition 2.5 (see [1], p. 62 and [2], p. 681). The set of the planes which contain the
straight line d is called a plane fascicle of axis d (see Fig. 2.10). The straight line d
is called the fascicle axis and π1 , π2 are called the base planes of the fascicle.
π : a1 x + b1 y + c1 z + d1 + λ (a2 x + b2 y + c2 z + d2 ) = 0, λ ∈ R√ .
74 2 Plane and Straight Line in E3
Example 2.6 Determine a plane which passes through the intersection of the planes
π1 : x + y + 5z = 0 and π2 : x − z + 4 = 0 and which forms with the plane
π : x − 4y − 8z + 12 = 0 an angle ϕ = π2 .
Solution
Let be d = d1 ∃ d2 . A plane of the plane fascicle of axis d has the equation
π ≡ : x + 5y + z + λ (x − z + 4) = 0,
namely
π ≡ : (1 + λ) x + 5y + (1 − λ) z + 4λ = 0.
It results that
n ≡ = (1 + λ) i + 5 j + (1 − λ) k.
As n = i − 4 j − 8k we shall deduce
π n · n≡ −27 + 9λ
cos ≤ π, π ≡ = cos = ⎩ ⎩= ∞ → λ = 3.
2 ⎩ ≡⎩ 27 + 2λ2
∗n∗ · ⎩n ⎩ 9
π ≡ : (1 + 3) x + 5y + (1 + 3) z + 4 · 3 = 0
or
π ≡ : 4x + 5y − 2z + 12 = 0.
2.4 Distances in E3
−−∀
Let A A≡ be a representative for a. The equation of the straight line is:
x − xA y − yA z − zA
= = .
a1 a2 a3
We know that ⎩ ⎩
⎩ ⎩
A A A≡ P M = ⎩A A≡ × M A⎩ . (2.27)
However ⎩ ⎩
⎩ ⎩
A A A≡ P M = ⎩ A A≡ ⎩ · ρ (M, d) . (2.28)
From (2.27) and (2.28) it results that the distance formula from a point to a straight
line is [6] ⎩ ⎩
⎩ ≡ ⎩ ⎩ ⎩
⎩A A × M A⎩ ⎩a × M A⎩
ρ (M, d) = ⎩ ⎩ = . (2.29)
⎩ ≡⎩ ∗a∗
⎩A A ⎩
Let d = (M0 , n) be the normal line to the plane which passes through M0 ,
n = ai + b j + ck . The equation of this straight line is
x − x0 y − y0 z − z0
d: = = =t
a b c
or
x = x0 + ta
d : y = y0 + tb, t ∈ R.
z = z 0 + tc
As M1 ∈ d we deduce
x1 − x0 y1 − y0 z1 − z0
d: = = =t ⇔
a b c
x1 = x0 + ta
y1 = y0 + tb, t ∈ R. (2.30)
z 1 = z 0 + tc
Multiplying the first equation of (2.31) with a, the second with b and the third
with c we have:
ax1 = ax0 + ta 2
by = by0 + tb2 , t ∈ R. (2.32)
1
cz 1 = cz 0 + tc2
namely
ax0 + by0 + cz 0 + d
t =− . (2.34)
a 2 + b2 + c2
We have
2.4 Distances in E3 77
M0 M1 = (x1 − x0 ) i + (y1 − y0 ) j + (z 1 − z 0 ) k;
therefore
⎩ ⎩
⎩ M0 M1 ⎩ = (x1 − x0 )2 + (y1 − y0 )2 + (z 1 − z 0 )2 (2.30)
= t 2 a 2 + b2 + c2 ⇔
⎩ ⎩
⎩ M0 M1 ⎩ = |t| a 2 + b2 + c2 . (2.35)
Substituting (2.34) into (2.35) we can deduce [6] the distance formula from a
point to a plane:
|ax0 + by0 + cz 0 + d|
ρ (M0 , π ) = ∞ . (2.36)
a 2 + b2 + c2
Solution
(a) Using the formul (2.36) we achieve:
∞
2 2 3
ρ (A, π ) = ∞ = .
3 3
(b) The distance from the point A to the straight line d (Fig. 2.13) is computed using
the formula
⎩ ⎩
⎩ ⎩ ⎩ ⎩
⎩ M A≡ × AM ⎩ ⎩a × AM ⎩
ρ (A, d) = ⎩ ⎩ = .
⎩ ⎩ ∗a∗
⎩ M A≡ ⎩
78 2 Plane and Straight Line in E3
Denoting z M = u ∈ R we deduce
1
3y M = 3 − u → y M = 1 − u
3
and
1
x M = y M + 1 = 2 − u.
3
xM = 2
We can suppose that u = 0; we obtain: y M = 1 =⇔ M (2, 1, 0) . We have:
zM = 0
AM = i − 2k and
i j k
a × AM = −1 −1 3 = 2i + j + k;
1 0 −2
⎩ ⎩ ∞
therefore ⎩a × AM ⎩ = 6. We shall obtain
∞
6
ρ (A, d) = ∞ = 0.739.
11
2.4 Distances in E3 79
where:
• π is the plane which passes through d1 and it is parallel with d2 ,
• ρ (A2 , π ) is the height which corresponds to the vertex A2 of the oblique paral-
lelepiped built on the vectors a, b, A1 A2 .
Therefore, the distance formula between two straight lines is [6]:
V parallelepi ped a · b × A1 A2
ρ (d1 , d2 ) = = ⎩ ⎩ . (2.37)
Abase ⎩a × b⎩
Definition 2.8 (see [6]). The support straight line corresponding to the segment which
represents the distance between two straight lines is called the common perpendic-
ular of the two straight lines.
Let δ be the straight line which represents the common perpendicular. To deter-
mine the equations of the straight line δ:
1. find the direction of the the common perpendicular n = a × b, a and b being the
direction vectors of the two straight lines;
2. write the equation of a plane π1 , which passes through d1 and contains n;
3. write the equation of a plane π2 , which passes through d2 and contains n;
4. δ = π1 ∃ π2 is the common perpendicular searched by us.
If
• d1 = (A1 , a) , A1 (x1 , y1 , z 1 ) ,
• d2 = (A2 , a) , A2 (x2 , y2 , z 2 ) ,
2.4 Distances in E3 81
• n = n 1 i + n 2 j + n 3 k,
• a = a1 i + a2 j + a3 k,
• b = b1 i + b2 j + b3 k,
then the equations of the common perpendicular δ are [6]:
⎨ x − x1 y − y1 z − z 1
⎨
⎨
⎨
⎨ a1 a2 a3 = 0
⎨
n1 n2 n 3
δ : (2.38)
⎨
⎨ x − x2 y − y2 z − z 2
⎨
⎨
⎨
⎨ b1 b2 b3 = 0.
n1 n2 n3
Solution
(a) We have
a·b
cos ≤ (d1 , d2 ) = cos ≤ a, b = ⎩ ⎩,
∗a∗ ⎩b⎩
where
• d1 = (A,
a) , A (1, 3, −2) , a = 2i + j + k,
• d2 = B, b , B (1, −2, 9) , b = i − 4 j + 2k.
We obtain π
cos ≤ (d1 , d2 ) = 0 ⇔≤ (d1 , d2 ) = .
2
(b) The direction of the common perpendicular is
n = a × b,
namely
i j k
n = 2 1 1 = 6i − 3 j − 9k.
1 −4 2
As
• the equation of the plane which passes through d1 and contains n is
82 2 Plane and Straight Line in E3
x − 1 y − 3 z + 2
π1 : 2 1 1 = 0 → π1 : x − 4y + 2z + 15 = 0;
6 −3 −9
where
• AB = (x B − x A ) i + (y B −
y A ) j + (z B − z A ) k = −5 j + 11k,
2 1 1
• a · b × AB = 1 −4 2 = −84,
0 −5 11
⎩ ⎩ ∞
• ⎩a × b⎩ = 62 + (−3)2 + (−9)2 = 126;
therefore
84 ∞
ρ (d1 , d2 ) = ∞ = 2 14.
126
2.5 Angles in E3
Definition 2.11 (see [3], p. 112). Let d1 , d2 be two straight lines, which have the
direction vectors a = a1 i + a2 j + a3 k and respectively b = b1 i + b2 j + b3 k. The
angle between the straight lines d1 and d2 is the angle between the vectors a and
b (see Fig. 2.15).
Hence
a·b a1 b1 + a2 b2 + a3 b3
cos ϕ = ⎩ ⎩= , ϕ ∈ [0, π ] . (2.39)
∗a∗ ⎩b⎩ a12 + a22 + a32 b12 + b22 + b32
a · b = 0 → a1 b1 + a2 b2 + a3 b3 = 0.
Let be
• π1 : a1 x + b1 y + c1 z + d1 = 0 and the normal vector n 1 = a1 i + b1 j + c1 k,
84 2 Plane and Straight Line in E3
Definition 2.13 (see [3], p. 113). The angle ϕ between the planes π1 and π2 is the
angle between the vectors n 1 and n 2 (see Fig. 2.16).
Hence
n1 · n2 a1 a2 + b1 b2 + c1 c2
cos ϕ = = , ϕ ∈ [0, π ] . (2.41)
∗n 1 ∗ ∗n 2 ∗ a12 + b12 + c12 a22 + b22 + c22
(2) π1 ≺π2 ⊂⇔ n 1 · n 2 = 0 ⊂⇔ a1 a2 + b1 b2 + c1 c2 = 0.
Let d be the straight line with the direction vector a = a1 i + a2 j + a3 k and the plane
π having the normal vector n = n 1 i + n 2 j + n 3 k.
Definition 2.15 (see [3], p. 113). The angle ϕ between the straight line d and the
plane π is the angle between the straight line d and the projection of this straight
line on the plane π (see Fig. 2.17).
2.5 Angles in E3 85
The angle between the straight line d and the plane π is related to the angle θ , the
angle of the vectors a and n, through the relations: θ = π2 ± ϕ as the vectors are on
the same side of the or in different parts. Hence:
π π
cos θ = cos ± ϕ = ± sin ϕ, θ ∈ [0, π ] ⇔ ϕ ∈ 0, .
2 2
As
n·a
cos θ = , θ ∈ [0, π ]
∗n∗ ∗a∗
it results that
|a1 n 1 + a2 n 2 + a3 n 3 | π
sin ϕ = , ϕ ∈ 0, . (2.42)
n 21 + n 22 + n 23 a12 + a22 + a32 2
Compute:
(a) the angle of these planes;
(b) the angle between the straight line and the plane π1 .
Solution
(a) We have n 1 = (2, −1, 0) , n 2 = (1, −5, 3) ; hence
∞
(2.41) n1 · n2 7
cos ≤ (π1 , π2 ) = cos ϕ = = = 0.529.
∗n 1 ∗ ∗n 2 ∗ 5
86 2 Plane and Straight Line in E3
(b) As the direction vector of the straight line d is a = (2, −1, 5) we obtain
(2.42) |a1 n 1 + a2 n 2 + a3 n 3 | 1
sin ≤ (d, π1 ) = sin ϕ = =∞ .
n 21 + n 2 + n 3 a1 + a2 + a3
2 2 2 2 2 6
2.6 Problems
1. Check if the points M1 (3, 0, 1), M2 (0, 2, 4) , M3 1, 43 , 3 are collineare.
Solution
Using Sage we shall have:
2.6 Problems 87
2. Write the equation of the plane determined by the points: M1 (3, 1, 0),
M2 (0, 7, 2), M3 (4, 1, 5) .
Solution
We shall give a solution in Sage:
5. Let d be the straight line determined by the point P0 (2, 0, −1) and the direction
vector v = i − j. Compute the distance from the point P (1, 3, −2) to the straight
line d.
6. Write the ecuation of the perpendicular from the point M (−2, 0, 3) on the plane
π : 7x − 5y + z − 11 = 0.
Hint. The perpendicular from a point to a plane is the straight line which passes
through that point and has the normal vector of the plane as a direction vector.
Solution
Solving this problem with Sage, we shall have:
2.6 Problems 89
7. It gives a tetrahedron ABC D defined by the points A (3, 0, 0), B (2, 4, 0),
C (−3, −1, 0), D (0, 0, 5). Write the equations of its faces, the edge equations
and the equations corresponding to the heights of the tetrahedron ABC D.
8. Write the equation of the plane which passes through Oz and is perpendicular
on the plane π : 8x + y + 2z − 1 = 0.
Hint. The equation of the plane which passes through Oz is: ax + by = 0.
Solution
We shall present the solution in Sage:
9. Determine the projection equation of the straight line having the equations
⎫
x − 3z + 1 = 0
d:
y − 2z − 3 = 0
on the plane π : x − y + 2z − 1 = 0.
10. Let be the straight lines:
⎫
x − 2y + z + 1 = 0
d1 :
y−z =0
90 2 Plane and Straight Line in E3
and
x −1 y+3 z
d2 : = = .
2 1 8
(a) Find the equation of the common perpendicular.
(b) Compute the distance between the two straight lines.
References
1. S. Chiriţă (ed.), Probleme de matematici superioare (Didactică şi Pedagogică, Bucureşti, 1989)
2. V. Postelnicu, S. Coatu (eds.), Mică enciclopedie matematică (Tehnică, Bucureşti, 1980)
3. C. Udrişte, Algebră liniară, geometrie analitică (Geometry Balkan Press, Bucureşti, 2005)
4. I. Vladimirescu, M. Popescu (eds.), Algebră liniară şi geometrie n- dimensională (Radical,
Craiova, 1996)
5. I. Vladimirescu, M. Popescu, Algebră liniară şi geometrie analitică, ed. Universitaria,
Craiova,1993
6. I. Vladimirescu, M. Popescu, M. Sterpu, Algebră liniară şi geometrie analitică, Note de curs şi
aplica ţii, Universitatea din Craiova, 1993
Chapter 3
Linear Transformations
The linear transformations should be studied [1] because they are compatible with
the operations defined in a vector space and allow the transfer of some algebraic
situations or problems from a space to another. Matrix operations clearly reflect
their similarity with the operations with linear transformations; hence the matrices
can be used for numerical representation of the linear transformations. The matrix
representation of linear transformations is [1] analogous to the representation of the
vectors through n coordinates, relative to a basis.
Definition 3.1 (see [2], p. 41). Let U and V two vector spaces over the field K . The
mapping T : U ∈ V is called a Linear transformation (or a linear mapping) if
the following conditions are satisfied:
(1) T (x + y) = T (x) + T (y) , (→) x, y ∈ U, namely T is additive;
(2) T (πx) = πT (x) , (→) x ∈ U, namely T is homogeneous.
The two properties of the linear maps can be formulated in a single.
Proposition 3.2 (see [2], p. 41). The mapping T : U ∈ V is linear if and only if
Theorem 3.9 (see [5], p. 45). Let T : Un ∈ Vn be a linear mapping between two
vector spaces with the same dimension. Then the following statements are equivalent:
(i) T is injective;
(ii) T is surjective;
(iii) T is bijective.
Definition 3.10 (see [2], p. 42). A linear mapping T : U ∈ V which is injective
and surjective is called isomorphism between the vector spaces U and V .
Theorem 3.11 (see [3], p. 29). Let U and V be two finite dimensional vector spaces
over the field K . Then they are isomorphic if and only if dim U = dim V .
Definition 3.12 (see [4], p. 26). Let S, T ∈ L(U, V ). The sum of the two linear
transformations is the linear transformation R ∈ L (U, V ) ,
Definition 3.13 (see [4], p. 26). Let T ∈ L(U, V ). The scalar multiples of the linear
transformation T, denoted πT ∈ L (U, V ) is defined as:
3.1 Linear Transformations 93
Definition 3.14 (see [5], p. 41). The composition of the two linear transformations
is called the product (or the multiplication) and is defined as in the case of the
functions.
Remark 3.15 (see [5], p. 41). The composition isn’t commutative but it is associative.
Proposition 3.16 (see [4], p. 27). If U, V, W are some vector spaces over K and
T : U ∈ V , S : V ∈ W are two linear transformations, then the mapping
S ⇒ T : U ∈ V,
(S ⇒ T ) (x) = S (T (x)) , (→) x ∈ U (3.6)
is linear.
Example 3.17 (see [6], p. 36). Let be the linear transformations
Solution
(a) f is a linear mapping∃
(b) We obtain
Using the Proposition 3.16 it results that the transformations f ⇒ g and g ⇒ f are
linear.
We can check this result in Sage, too:
Definition 3.19 (see [5], p. 42). Let T ∈ L(U, V ) be a bijective linear mapping
(hence it is invertible). Its inverse, T −1 ∈ L(U, V ) is a linear mapping.
Example 3.20 (see [6], p. 36). Let be the linear mapping
Solution
(a) We know (see the Definition 3.6) is injective if and only if Ker T = 0R3 . Let
be x ∈ R3 .
f (x) = 0R3 ∃ (x1 − 2x2 + x3 , 2x1 + x2 − x3 , x2 − 3x3 ) = (0, 0, 0)
x1 − 2x2 + x3 = 0
∃ 2x1 + x2 − x3 = 0
x2 − 3x3 = 0.
Since the determinant of the matrix A, associated to the previous system is:
⎫ ⎫
⎫ 1 −2 1 ⎫
⎫ ⎫
ρ = ⎫⎫ 2 1 −1 ⎫⎫ ⇐= 0 ⇔ rank (A) = 3 ⇔
⎫ 0 1 −3 ⎫
We obtain f −1 : R3 ∈ R3 ,
⎭ ⎧
1 1 1
f −1 (x) = (2x1 + 5x2 − x3 ) , (−6x1 + 3x2 − 3x3 ) , (−2x1 + x2 − 5x3 ) .
12 12 12
(b) Using the Definition 3.19 it results that the mapping f −1 is linear.
Theorem 3.21 (see [5], p. 41). The set L(U, V ) is a vector space over the field K
relative to the addition of the linear transformations and the scalar multiplication of
a linear transformation.
Theorem 3.22 (see [5], p. 43). Let T ∈ L(U, V ). We have the following properties:
(i) Ker T is a vector subspace of U ;
(ii) Im T is a vector subspace of V .
Proposition 3.23 (see [5], p. 40 and [3], p. 27). If T ∈ L (U, V ). Then:
(1) T 0U = 0V , i.e. a linear transformation maps the null vector, in the null vector;
(2) T (−x) = −T (x) , (→) x ∈ U ;
(3) If W is a vector subspace of U , then T (W ) is a vector subspace of V ;
(4) If the vectors x 1 , . . . , x n ∈ U are linearly dependent then the vectors T (x 1 ) ,
. . . , T (x n ) ∈ V are linearly dependent, too;
(5) Being given the vectors x 1 , . . . , x n ∈ U , if the vectors T (x 1 ) , . . . , T (x n ) ∈ V
are linearly independent then the vectors x 1 , . . . , x n are linearly independent,
too.
Proposition 3.24 (see [3], p. 28). Let T ∈ L (U, V ) be an injective linear mapping.
If the system of vectors {x 1 , . . . , x n } ≺ U is linearly independent, the system of
vectors T (x 1 ) , . . . , T (x n ) is linearly independent.
Proposition 3.25 (see [3], p. 28). Let U, V be two vector spaces over the field K and
T ∈ L (U, V ). The following statements hold:
(a) If {x 1 , . . . , x n } ≺ U is a system of generators for U , then T (x 1 ) , . . . , T (x n )
is a system of generators for Im T .
(b) If T is surjective and {x 1 , . . . , x n } ≺ U is a system of generators for U , then
T (x 1 ) , . . . , T (x n ) is a system of generators for V .
Theorem 3.26 (see [3], p. 29 and [2], p. 44). Let U, V be two vector spaces over the
field K . If dim U < √ and T ∈ L(U, V ), then
We shall denote by
⎩ (1) (1)
π1 . . . πn
⎨(B1 ,B2 ) .. ..
T = . . (3.7)
(m) (m)
π1 ... πn
the m × m matrix, where the column by index i contains the coordinates of the vector
T (ei ).
⎨(B1 ,B2 ) is called the associated matrix
Definition 3.28 (see [2], p. 43). The matrix T
of the linear transformation T relative to the bases B1 and B2 , fixed in the vector
spaces U and V .
Example 3.29 (see [9]). Let be
We shall obtain ⎭ ⎧
⎨ 2 1 −1
f (B1 ,B2 ) = .
01 7
Let be x ∈ R3 .
dim Ker f = n − r = 3 − 2 = 1,
we obtain
dim Im f = 2.
(d) We know that Im f ≡ R2 i.e. Im f is a vector space of R2 (see the Theorem 3.22).
3.1 Linear Transformations 99
As
Im f ≡ R2 ⎪
Definition 3.6
dim Im f = 2 ⇔ Im f = R2 =⇔ f is surjective.
⎛
dim R2 = 2
hence
f (x) = π1 f (e1 ) + π2 f (e2 ) + π3 f (e3 ) .
We obtain
Ker f = x ∈ V | f (x) = 0V = {x ∈ V |π1 + π2 + π3 = 0} ;
therefore
dim Ker f = n − r = 3 − 1 = 2.
Whereas
100 3 Linear Transformations
π1 = −π2 − π3
we deduce
Im f = { f (x) |x ∈ V } .
We shall obtain
Example 3.31 (see [9]). Let T : M2 (R) ∈ M2 (R) be a linear mapping, defined by
⎭ ⎧ ⎭ ⎧
1 −2 10
T (A) = A , (→) A ∈ M2 (R) .
0 1 11
Build the matrix of the linear mapping T in the canonical basis of that space.
3.1 Linear Transformations 101
Solution
We know that
⎭ ⎧ ⎭ ⎧ ⎭ ⎧ ⎭ ⎧⎠
10 01 00 00
B = {E 11 , E 12 , E 21 , E 22 } = , , ,
00 00 10 01
⎭ ⎧ ⎭ ⎧ ⎭ ⎧ ⎭ ⎧
1 −2 1 0 1 0 10
T (E 11 ) = · · = = E 11
0 1 0 0 1 1 00
⎭ ⎧ ⎭ ⎧ ⎭ ⎧ ⎭ ⎧
1 −2 0 1 1 0 11
T (E 12 ) = · · = = E 11 + E 12
0 1 0 0 1 1 00
⎭ ⎧ ⎭ ⎧ ⎭ ⎧ ⎭ ⎧
1 −2 0 0 1 0 −2 0
T (E 21 ) = · · = = −2E 11 + E 21
0 1 1 0 1 1 1 0
⎭ ⎧ ⎭ ⎧ ⎭ ⎧ ⎭ ⎧
1 −2 0 0 1 0 −2 −2
T (E 22 ) = · · =
0 1 0 1 1 1 1 1
= − 2E 11 − 2E 12 + E 21 + E 22 .
It results ⎩
1 1 −2 −2
0 1 0 −2
⎨(B)
T =
0
.
0 1 1
0 0 0 1
f (P) = P ≤ , (→) P ∈ R4 [X ],
102 3 Linear Transformations
(b) We compute
⎬ f (1) = 1≤ = 0 = 0 · 1 + 0 · X + 0 · X 2 + 0 · X 3
⎬
⎬
⎬
f (X ) = X≤ = 1 = 1 · 1 + 0 · X + 0 · X2 + 0 · X3
2 2 ≤
f X = X = 2X = 0 · 1 + 2 · X + 0 · X 2 + 0 · X 3
⎬
⎬ ≤
⎬
⎬ f X 3 = X 3 = 3X 2 = 0 · 1 + 0 · X + 3 · X 2 + 0 · X 3
4 4 ≤
f X = X = 4X 3 = 0 · 1 + 0 · X + 0 · X 2 + 4 · X 3 .
We deduce that ⎩
0 1 0 0 0
0 0 2 0 0
⎨
f (B1 ,B2 ) =
0
.
0 0 3 0
0 0 0 0 4
(c) We have
Ker f = P ∈ R4 [X ]| f (P) = O R3 [X ] .
P = a0 + a1 X + a2 X 2 + a3 X 3 + a4 X 4 ;
then
Ker f = {P ∈ R4 [X ]|P = a0 = a0 · 1} ,
However
Im f ≡ R3 [X ] ⎪
dim Im f = 4 ⇔ Im f = R3 [X ].
⎛
dim R3 [X ] = 4
We propose to define the matrix operations, starting from the corresponding opera-
tions of linear maps.
104 3 Linear Transformations
Let be S,T ∈ L(U, V ) and B1 = {e1 , . . . , en }, B2 = f 1 , . . . , f m be two bases
in U and respectively in V. Let A, B be the associated matrices of S and T relative
to the two bases: A = ⎨
S(B1 ,B2 ) , A = ai j 1∗i∗m and B = T ⎨(B1 ,B2 ) , B = bi j 1∗i∗m.
1∗ j∗n 1∗ j∗n
We have:
⎬
⎬ S (e1 ) = a11 f 1 + · · · + am1 f m
⎬
⎬ ..
⎬
⎬
.
S e j = a1 j f 1 + · · · + am j f m
⎬
⎬ ..
⎬
⎬
⎬
⎬ .
S (en ) = a1n f 1 + · · · + amn f m ,
⎬
⎬ T (e1 ) = b11 f 1 + · · · + bm1 f m
⎬
⎬ ..
⎬
⎬
.
T e j = b1 j f 1 + · · · + bm j f m
⎬
⎬ ..
⎬
⎬
⎬
⎬ .
T (en ) = b1n f 1 + · · · + bmn f m .
a1 j f 1 + · · · ai j f i + · · · + am j f m = b1 j f 1 + · · · + bi j f i + · · · + bm j f m (3.9)
As f 1 , . . . , f m is a basis and we know that the writing of a vector into a basis
is unique, from (3.9) it results that ai j = bi j , (→) i = 1, m, (→) j = 1, n.
Definition 3.33 (see [8], p. 60). The matrices A = ai j 1∗i∗m and B = bi j 1∗i∗m
1∗ j∗n 1∗ j∗n
are equal if and only if
ai j = bi j , (→) i = 1, m, (→) j = 1, n.
namely
(S + T ) e j = a1 j + b1 j f 1 + · · · ai j + bi j f i + · · · + am j + bm j f m .
(3.10)
3.2 Matrix as a Linear Mapping 105
However
(S + T ) e j = c1 j f 1 + · · · + ci j f i + · · · + cm j f m . (3.11)
ci j = ai j + bi j , (→) i = 1, m, (→) j = 1, n.
Definition 3.34 (see [8], p. 60). The sum of matrices A = ai j 1∗i∗m and B =
1∗ j∗n
bi j 1∗i∗m is the matrix C = ci j 1∗i∗m , ci j = ai j +bi j , (→) i = 1, m, (→) j = 1, n.
1∗ j∗n 1∗ j∗n
We denote C = A + B.
However
(πS) e j = c1 j f 1 + · · · + ci j f i + · · · + cm j f m . (3.13)
We have:
(T ⇒ S) e j = T S e j = T a1 j f 1 + · · · ak j f k + · · · + am j f m
= a1 j T f 1 + · · · + ak j T f k + · · · + am j T f m ,
i.e.
106 3 Linear Transformations
(T ⇒ S) e j = a1 j b11 g 1 + · · · + bi1 g i + · · · + b p1 g p + · · · + (3.14)
ak j b1k g 1 + · · · + bik g i + · · · + b pk g p + · · · +
am j b1m g 1 + · · · + bim g i + · · · + b pm g p .
We denote by C = ci j 1∗i∗ p the associated matrix of the linear mapping T ⇒ S ∈
1∗ j∗n
L (U, W ). It results
(T ⇒ S) e j = c1 j g 1 + · · · + ci j g i + · · · + c pj g p . (3.15)
We write (3.14) as
(T ⇒ S) e j = b11 a1 j + · · · + b1k ak j + · · · + b1m am j g 1 (3.16)
+ · · · + bi1 a1 j + · · · + bik ak j + · · · + bim am j g i
+ · · · + b p1 a1 j + · · · + b pk ak j + · · · + b pm am j g p .
AB = B A = In .
Theorem 3.41 (see [8], p. 61). Let U, V be two vector spaces over the field K and
T ∈ L(U, V ). If B1 = {e1 , · · · , en } is a basis in U and B2 = f 1 , · · · , f n is a
basis in V and A is the associated matrix of the linear mapping T relative to the bases
B1 and B2 , then
rank T = rank A.
Proposition 3.42 (see [8], p. 62). A square matrix is invertible if and only if it is
nonsingular.
x = x (1) e1 + · · · + x (n) en .
We shall have
namely
⎦ ! ⎦ !
(1) (m)
T (x) = x (1) π1 + · · · + x (n) πn(1) f 1 + · · · + x (1) π1 + · · · + x (n) πn(m) f m .
(3.18)
As T (x) ∈ V it results
therefore
⎨(B1 ,B2 ) · x B1 .
(T (x)) B2 = T (3.20)
108 3 Linear Transformations
Let B1≤ be another basis of U and B2≤ be another basis of V . Let C be the transition
matrix from the basis B1 to the basis B1≤ and D the transition matrix from the basis
B2 to the basis B2≤ .
Within the relation (3.20) we have
⎨ B ≤ ,B ≤ · x B ≤ .
(T (x)) B2≤ = T (3.21)
( 1 2) 1
We know that
x B1 = C( B1 ,B ≤ ) · x B1≤ (3.22)
1
and
(T (x)) B2 = D( B2 ,B ≤ ) · (T (x)) B2≤ . (3.23)
2
⎨(B1 ,B2 ) · x B1 .
D( B2 ,B ≤ ) · (T (x)) B2≤ = T (3.24)
2
⎨(B1 ,B2 ) · C B ,B ≤ · x B ≤ .
D( B2 ,B ≤ ) · (T (x)) B2≤ = T (3.25)
2 ( 1 ) 1 1
If in the relation (3.25) we multiply at the left, the both members with the inverse
of D, we obtain
⎨ B ≤ ,B ≤ · x B ≤ = D −1 ≤ · T
T ⎨ ·C ≤ · x ≤,
( 1 2) 1 ( B2 ,B2 ) (B1 ,B2 ) ( B1 ,B1 ) B1
i.e.
⎨ B ≤ ,B ≤ = D −1 ≤ · T
T ⎨ ·C ≤ . (3.27)
( 1 2) ( B2 ,B2 ) (B1 ,B2 ) ( B1 ,B1 )
The formula (3.27) constitutes [9] the changing formula of the associated matrix
of a linear mapping when one changes the bases in the two vector spaces U and V .
Example 3.43 (see [5], p. 48). Let T1 , T2 ∈ End R3 , be defined as
⎦ !
T1 (x) = 5x (1) − x (2) − 5x (3) , 20x (1) − 15x (2) + 8x (3) , 3x (1) − 2x (2) + x (3)
⎦ !
T2 (x) = 10x (1) − 10x (2) + 10x (3) , 0, 5x (1) − 5x (2) + 5x (3) ,
⎦ !
(→) x = x (1) , x (2) , x (3) ∈ R3 .
3.3 Changing the Associated Matrix to the Change of Basis 109
Find the sum of the two endomorfisme matrix T = T1 + T2 relative to the basis
B ≤ = {v 1 = (2, 3, 1) , v 2 = (3, 4, 1) , v 3 = (1, 2, 2)} ≺ R3 .
Solution
We have:
we shall obtain ⎩
15 −11 5
⎨(B)
T = 20 −15 8 .
8 −7 6
we have ⎩
231
C = 3 4 2.
112
Hence ⎩
100
⎨(B ≤ )
T ⎨(B) · C = 0 2 0 .
= C −1 · T
003
The eigenvalue problems are of great importance in many branches of physics. They
make it possible to find some coordinate systems in which the linear transformations
take the simplest forms.
For example, in mechanics the main moments of a solid body one finds with the
eigenvalues of a symmetric matrix representing the vector tensor. The situation is
similar in continuous mechanics, where a body rotations and deformations in the
main directions are found using the eigenvalues of a symmetric matrix.
Eigenvalues have [1] a central importance in quantum mechanics, where the mea-
sured values of the observable physical quantities appear as eigenvalues of operators.
Also, the eigenvalues are useful in the study of differential equations and contin-
uous dynamical systems that arise in areas such as physics and chemistry.
Definition 3.44 (see [2], p. 45). Let V be a vector space over the field K and
T ∈End(V ). The subspace vector W of V is called an invariant subspace rela-
tive to T if from x ∈ W it results T (x) ∈ W or T (W ) ≡ W .
Definition 3.45 (see [2], p. 45). Let V be a vector space over the field K and
T ∈End(V ). We say that the scalar ϕ ∈ K is an eigenval for T if there is x ∈ V \ 0
such that
T (a i ) = ϕi a i , i = 1, p,
⎨(B) be the associated matrix of the linear mapping T relative to the basis B,
Let T
⎩ (1) (1)
π1 . . . πn
⎨(B) .. .. .
T = . . (3.31)
(n) (n)
π1 ... πn
x = x (1) e1 + · · · + x (n) en .
i.e. a linear and homogeneous system in the unknowns x (1) , x (2) , . . . , x (n) .
The scalar ϕ is an eigenvalue of T if and only if the system (3.35) admits the
nonbanal solutions, i.e. there is a scalar system x (1) , x (2) , . . . , x (n) , not all null, that
verifies the system (3.35).
This is achieved if and only if
⎫ ⎫
⎫ π (1) − ϕ π (1) · · · π (1) ⎫
⎫ 1 2 n ⎫
⎫ ⎫
⎫ π1(2) π2(2) − ϕ · · · πn(2) ⎫
ρ=⎫ ⎫ = 0,
⎫ ··· ··· ··· ··· ⎫
⎫ ⎫
⎫ π (n) π2
(n) (n)
· · · πn − ϕ ⎫
1
i.e.
det T⎨(B) − ϕIn = 0, (3.36)
Remark 3.51 (see [3], p. 66). The polynomial P (ϕ) is a polynomial in ϕ, of degree n:
where
3.4 Eigenvalues and Eigenvectors 113
⎬
⎬ n
(i)
⎬ δ = trace ⎨(B) =
T πi
⎬
⎬ 1
⎬
⎬ ⎫ i=1 ⎫
⎬
⎬
⎫⎫ π (i) π (i) ⎫⎫
δ2 = ⎫ (i j) (j j) ⎫
⎬
⎬ ⎫ πi π j ⎫
⎬
⎬ 1∗i< j∗n
⎬
⎬ ..
⎬
⎬ .
⎬
δn = det T ⎨(B) .
Definition 3.52 (see [2], p. 45). The polynomial P (ϕ) defined in (3.37), in the
indeterminate ϕ, of degree n is called the characteristic polynomial associated of
the endomorphism T .
Definition 3.53 (see [3], p. 66). The equation P (ϕ) = 0 is called the characteristic
equation associated of the endomorphism T .
Remark 3.54 (see [3], p. 66). The scalar ϕ is an eigenvalue of T if and only if ϕ is
the root of the characteristic equation.
Example 3.55 (see [9]). Let T be be an endomorphism of R3 such that T has
the eigenvalues: ϕ1 = 1, ϕ2 = −1, ϕ3 = 2 with the eigenvectors x 1 = (1, 0, 1),
x 2 = (−1, 2, 1), x 3 = (2, 1, −1). Write the associated matrix of T in the canonical
basis from R3 .
Solution
We note that
x 1 = e1 + e3
x 2 = −e1 + 2e2 + e3
x 3 = 2e1 + e2 − e3 .
As
T (x 1 ) = ϕ1 x 1
we deduce
T (e1 + e3 ) = ϕ1 (e1 + e3 ) . (3.38)
Similarly, because
T (x 2 ) = ϕ2 x 2
we deduce
T (−e1 + 2e2 + e3 ) = ϕ2 (−e1 + 2e2 + e3 ) . (3.40)
Similarly, as
T (x 3 ) = ϕ3 x 3
we shall have
T (2e1 + e2 − e3 ) = ϕ3 (2e1 + e2 − e3 ) . (3.42)
We shall solve the system of equations resulting from the relations (3.39), (3.41)
and (3.43) to determine the expression of T (ei ) , i = 1, 3 as a linear combination of
the elements of the basis, i.e.:
T (e1 ) + T (e3 ) = e1 + e3
−T (e1 ) + 2T (e2 ) + T (e3 ) = e1 − 2e2 − e3 (3.44)
2T (e1 ) + T (e2 ) − T (e3 ) = 4e1 + 2e2 − 2e3 .
i.e.
T (e2 ) + T (e3 ) = e1 − e2 . (3.45)
Substituting (3.47) and (3.48) into the first equation of the system (3.44) we obtain
i.e.
5e1 − e2 − 4e3 = 4T (e2 ) ;
3.4 Eigenvalues and Eigenvectors 115
hence
5 1
T (e2 ) = e1 − e2 − e3 .
4 4
We shall have
⎭ ⎧
5 1 5 3
T (e1 ) = 5e1 − 3e3 − 3 e1 − e2 − e3 = e1 + e2
4 4 4 4
and ⎭ ⎧
5 1 1 3
T (e3 ) = e1 − e2 − e1 − e2 − e3 = − e1 − e2 + e3 .
4 4 4 4
We obtain ⎩
5/4 5/4 −1/4
⎨(B)
T = 3/4 −1/4 −3/4 .
0 −1 1
Definition 3.56 (see [5], p. 64). The set of the roots of the characteristic equation, that
is associated to the endomorphism T is called the spectrum of the endomorphism
T . If all the roots are simple on says that T is an endomorphism with a simple
spectrum. We denote by θ (T ) the spectrum of T .
Theorem 3.56 (Hamilton-Cayley, see [2], p. 49). Let V be a n dimensional vector
space over K , n ∞ 1 and be an endomorphism T ∈End(V ). If P (ϕ) is the character-
istic polynomial of A = T⎨(B) (the matrix of the endomorphism T relative to a basis
B of V , then P (A) = 0End(V ) .
Example 3.57 (see [7], p. 29). Compute
Solution
The characteristic polynomial, that is associated to the matrix A is
⎫ ⎫
⎫2 − ϕ 0 0 0 ⎫⎫
⎫
⎫ 1 3−ϕ 1 1 ⎫⎫
P (ϕ) = ⎫⎫ ⎫ = ϕ − 8ϕ + 24ϕ − 32ϕ + 16.
4 3 2
⎫ 0 0 1 − ϕ −1 ⎫
⎫ −1 −1 0 2 − ϕ⎫
Definition 3.58 (see [3], p. 66). The dimension of the eigensubspace Wϕ0 associated
of the eigenvalue ϕ0 is called the geometric multiplicity of ϕ0 and it is denoted
by gϕ0 .
Definition 3.59 (see [3], p. 66). The algebraic multiplicity of the eigenvalue ϕ0 ,
denoted by aϕ0 means the multiplicity of ϕ0 as a root of the characteristic polynomial
P (ϕ), associated to the endomorphism T .
Proposition 3.60 (see [3], p. 66). The characteristic polynomial P (ϕ) is invariant
relative to the basis changing in the vector space V .
Theorem 3.61 (see [3], p. 66). Let V be a vector space over the field K , dimV =
n < √, T ∈End(V ) and ϕ0 an eigenvalue of T . Then the geometric multiplicity of
ϕ0 is not greater than the algebraic multiplicity of ϕ0 , i.e. gϕ0 ∗ aϕ0 .
Proposition 3.62 (see [5], p. 64). Let V be a vector space over the field K and
T ∈End(V ) . Then each eigenvector of T corresponds to a single eigenvalue ϕ ∈
θ (T ).
The matrix ⎩
ϕ1 0 ··· 0
0 ϕ2 0
⎨(B)
T =
0 .
0 .. 0
0 0 ϕn
is a diagonal matrix.
Definition 3.63 (see [5], p. 69). Let V be a vector space over the finite dimensional
field K , n ∞ 1. We say that the endomorphism T ∈End(V ) is diagonalizable if
there is a basis of V relative to which its matrix is a diagonal matrix.
Theorem 3.64 (see [3], p. 68). Let V be a vector space over the n finite dimen-
sional field K , n ∞ 1. The necessary and sufficient condition that the endomorphism
T ∈End(V ) to be diagonalizable is that the characteristic polynomial P (ϕ) to have
all the roots in K and the geometric multiplicity of each eigenvalue to be equal to its
algebraic multiplicity.
118 3 Linear Transformations
B ≤ = B1 ⊂ B2 ⊂ . . . ⊂ B p .
7. build the transition matrix from the basis B to the basis B ≤ , i.e. M(B,B ≤ ) ;
8. test the correctness of the calculations using the relation
⎨(B ≤ ) = M−1 ≤ · T
T ⎨(B) · M(B,B ≤ ) .
(B,B )
Example 3.65 (see [9]). On the vector space of matrices of second order one considers
the mapping:
T : M2 (R) ∈ M2 (R) , T (A) = At .
(a) Write the associated matrix of T relative to the canonical basis of the space
M2 (R).
(b) Determine the eigenvalues and the corresponding eigenspaces.
3.4 Eigenvalues and Eigenvectors 119
(c) Determine a basis B ≤ of the vector space M2 (R) relative to which the associated
matrix of T has a diagonal form.
Solution
⎭ ⎧ ⎭ ⎧ ⎭ ⎧ ⎭ ⎧⎠
10 01 00 00
B = {E 11 , E 12 , E 21 , E 22 } = , , ,
00 00 10 01
The associated matrix of T relative to the canonical basis of the space M2 (R)
will be ⎩
1000
⎨(B) = 0 0 1 0 .
T 0 1 0 0
0001
(b) We determine
⎫ ⎫
⎫1 − ϕ 0 0 0 ⎫⎫
⎫
⎫ 0 ⎫⎫
P (ϕ) = det T⎨(B) − ϕI4 = ⎫ 0 −ϕ 1
⎫ 0 1 −ϕ 0 ⎫⎫
⎫
⎫ 0 0 0 1 − ϕ⎫
⎦ !
= (1 − ϕ)2 ϕ2 − 1 = (ϕ − 1)3 (ϕ + 1) .
Hence ⎭ ⎧⎠
a11 a12
Wϕ1 = A ∈ M2 (R) |A = .
a12 a22
We can write
⎭ ⎧ ⎭ ⎧ ⎭ ⎧
10 01 00
A = a11 + a12 + a22
00 10 01
⎭ ⎧
01
= a11 E 11 + a12 + a22 E 22 ;
10
it results that
⎭ ⎧⎠
01
B1 = E 11 , E 22 , (3.49)
10
it results ⎭ ⎧ ⎭ ⎧
a11 a12 00
= ,
a12 a22 00
i.e.
3.4 Eigenvalues and Eigenvectors 121
Wϕ2 = A ∈ M2 (R) |At = −A .
Therefore ⎭ ⎧⎠
0 a12
Wϕ2 = A ∈ M2 (R) |A = .
−a12 0
We can write ⎭ ⎧
0 1
A = a12 .
−1 0
Similarly, we obtain ⎭ ⎧⎠
0 1
B2 = (3.50)
−1 0
aϕ1 = gϕ1 ⎪
aϕ2 = gϕ2 ⇔ T is diagonalizable.
⎛
the characteristic equation has some real roots
The basis of the vector space M2 (R) relative to which the associated matrix of T
has the canonical diagonal form is
⎭ ⎧ ⎭ ⎧⎠
≤ (3.49)+(3.50) 01 0 1
B = B1 ⊂ B2 = E 11 , E 22 , , = {F1 , F2 , F3 , F4 }
10 −1 0
and
⎩
1 0 0 0
0 1 0 0
⎨(B ≤ )
T =
0
.
0 1 0
0 0 0 −1
We shall have:
F1 = E 11 = 1 · E 11 + 0 · E 12 + 0 · E 21 + 0 · E 22
F2 = E 22 = 0 · E 11 + 0 · E 12 + 0 · E 21 + 1 · E 22
122 3 Linear Transformations
F3 = E 12 + E 21 = 0 · E 11 + 1 · E 12 + 1 · E 21 + 0 · E 22
F4 = E 12 − E 21 = 0 · E 11 + 1 · E 12 − 1 · E 21 + 0 · E 22 .
Therefore
⎩
1 0 0 0
0 0 1 1
M(B,B ≤ ) =
0
.
0 1 −1
0 1 0 0
It results that
⎩
1 0 0 0
0 1 0 0
M−1 ⎨
(B,B ≤ ) · T(B) · M(B,B )
≤ =
0
=T
⎨(B ≤ ) .
0 1 0
0 0 0 −1
⎩
ϕ1 0 00
⎭ ⎧ 0 ϕ 1 0 0
J2 (ϕ) O
B5 (ϕ) = =
0 0 ϕ 1 0
,
O J3 (ϕ) 0 0 0 ϕ 1
00 0 0ϕ
124 3 Linear Transformations
⎩
ϕ 10 00 0 00
0 ϕ1 00 0 0 0
0ϕ 0 0
⎭ ⎧ 0 10 0
J4 (ϕ) O 0 00 ϕ1 0 0 0
B8 (ϕ) = =
0
O J4 (ϕ) 00 0ϕ 1 0 0
0 00 00 ϕ 1 0
0 00 00 0 ϕ 1
0 00 00 0 0ϕ
Definition 3.70 (see [2], p. 47). A square matrix of order n, which has Jordan blocks
on the main diagonal, i.e. of the form
⎩
Bn 1 (ϕ1 ) O
Bn 2 (ϕ2 )
J= . ∈ Mn (K ) , (3.53)
..
O Bnr (ϕr )
Let V be a finite n dimensional vector space over the field K , n ∞ 1 and let
T ∈End(V ) be an endomorphism.
The jordanization algorithm of an endomorphism T consists [2] of the following
steps:
1. Choose a basis B of V and write the associated matrix of T relative to this basis,
⎨(B) .
i.e. the matrix T
2. Determine the characteristic polynomial P (ϕ) using (3.37); there are two cases:
126 3 Linear Transformations
6. Determine the number of the Jordan cells of order h ∈ {1, 2, . . . , si } within the
formula
dh = rank (T − ϕi V )h+1 + rank (T − ϕi V )h−1 − 2 · rank (T − ϕi V )h , (3.57)
where
⎬
⎬ rank (T − ϕi V )0 = n
⎬
rank (T − ϕi V )si +1 = rank (T − ϕi V )si
⎬ si (3.58)
⎬
⎬ h · dh = m i .
h=1
(a) Show that T is jordanizable, write its matrix to the Jordan canonical form J
and find the basis of R3 relative to which T has the matrix J.
(b) Compute An , n ∈ N∩ .
Solution
P (ϕ) = (2 − ϕ)3
dim Ker (T − ϕi V ) = 3 − r,
where
r = rank (A − 2I3 ) = 1;
therefore
n 1 = dim Ker (T − ϕi V ) = 3 − 1 = 2.
We note that
hence
i.e. s1 = 2.
We have
d1 = 0 + 3 − 2 = 1,
128 3 Linear Transformations
d2 = 0 + 1 − 0 = 1.
Therefore, there will be two Jordan cells in the achieved matrix in the Jordan
canonical form, namely a first order cell J1 (2) and a second-order cell J2 (2). The
matrix J of T will be the following in the Jordan canonical form:
⎩
⎭ ⎧ 200
J1 (2) O
J= = 0 2 1
O J2 (2)
002
and the basis of R3 relative to which T has this form is B ≤ = f 1 , f 2 , f 3 .
Using the definition of the associated matrix to an endomorphism we shall have:
T f 1 = 2 f 1
T f = 2f2
2
T f 3 = f 2 + 2 f 3.
As the number of Jordan cells for the eigenvalue ϕ1 is equal to the maximum
number of the corresponding linearly independent eigenvectors, it results that the
vectors f 1 , f 2 are some eigenvectors for T , i.e. their coordinates are the solution of
system: ⎩ ⎩ ⎩
3 −2 1 x x
2 −2 2 y = 2 y ∃
3 −6 5 z z
3x − 2y + z = 2x x − 2y + z = 0
2x − 2y + 2z = 2y ∃ 2x − 4y + 2z = 0
3x − 6y + 5z = 2z 3x − 6y + 3z = 0.
y = t, z = u, (→) t, u ∈ R;
it results x = 2t − u.
We shall have
Vϕ1 = f ∈ R3 | f = (2t − u, t, u) , t, u ∈ R .
T f 3 = f 2 + 2 f 3,
y = v, z = w, (→) v, w ∈ R;
it results x = 2t − u + 2v − w.
For t = 2, u = 3 we consider f 2 = (1, 2, 3). In the case when t = 2, u = 3 for
v = w = 0 we obtain f 3 = (1, 0, 0) .
(b) The transition matrix from the canonical basis B to the basis B ≤ will be
⎩
211
C = M(B,B ≤ ) = 1 2 0.
030
We shall obtain
⎩ ⎩ ⎩
211 200 0 1 −2/3
A = CJC −1 = 1 2 0 · 0 2 1 · 0 0 1/3
030 002 −1 2 1
and
⎩ ⎩ n ⎩
211 2 0 0 0 1 −2/3
An = CJn C −1 = 1 2 0 · 0 2n n · 2n−1 · 0 0 1/3
030 0 0 2n −1 2 1
⎩ n
2 +n·2 n−1 −n · 2 n n·2 n−1
⇔ An = n · 2n (1 − 2n) 2n n · 2n .
3n · 2 n−1 −3n · 2 (2 + 3n) 2n−1
n
3.5 Problems
Solution
We shall use Sage to find the matrix A :
T : R2 [X ] ∈ R2 [X ] , T (P) = (4X + 1) P ≤ .
(a) Find the eigenvalues and the eigenspace of the corresponding eigenvectors.
(b) Decide if T is diagonalizable or not and if so write the diagonal form of the
matrix and specify the form relative of which the matrix is diagonal.
Solution
The solution of the problem in Sage is:
132 3 Linear Transformations
⎩
1 −1 2
A = 1 0 1
1 0 −1
relative to the canonical basis of R3 . Determine the matrix of T relative to the basis
B1 = f 1 = (1, 2, 3) , f 2 = (3, 1, 2) , f 3 = (2, 3, 1) .
Solution
This matrix can be determined in Sage:
(a) Determine the eigenvalues and the eigenvectors associated to the endomor-
phism T .
(b) Check if T is diagonalizable and then determine a space basis relative to
which the associated matrix of T has the canonical diagonal form.
8. Let the mapping T :R3 [X ] ∈ R4 , T (P) = P (−3) P (−1) P (1) P (3) .
3.5 Problems 133
2 "1
T : R3 [X ] ∈ R3 [X ] , T (P) = X k
t k P (t) dt.
k=1 −1
(a) Write the associated matrix of T relative to the canonical basis of the vector
space R3 [X ].
(b) Determine KerT and ImT .
Solution
Using Sage to solve the problem we achieve:
Hence:
Ker f = {P ∈ R3 [X ] |P = a2 Q 1 + a3 Q 2 }
and
⎭ ⎧ ⎭ ⎧ ⎠
2 2 2 2
Im f = { f (P) |P ∈ R3 [X ]} = a1 + a3 X + a0 + a2 X 2 .
3 3 3 5
134 3 Linear Transformations
(a) Show that T is jordanizable, write its matrix in the Jordan canonical form J
and determine the basis of R3 relative to which J is the matrix of T .
(b) Compute An , n ∈ N∩ .
References
The study of the Euclidean vector space is required to obtain the orthonormal bases,
whereas relative to these bases, the calculations are considerably simplified. In a
Euclidean vector space, scalar product can be used to define the length of vectors
and the angle between them.
Let E be a real vector space.
Definition 4.1 (see [6], p. 101). The mapping <, >: E∈E is called a scalar product
(or an Euclidean structure) on E if the following conditions are satisfied:
(a) < x, y >=< y, x >, (→) x, y ∈E
(b) < x + y, z >=< x, z > + < y, z >, (→) x, y, z ∈E
(c) α < x, y >=< αx, y >, (→) x, y ∈E, (→) α ∈ R
(d) < x, x > ≥ 0, (→) x ∈E; < x, x > = 0 ⇔ x = 0.
The scalar < x, y >∈ R is called the scalar product of vectors x, y ∈E.
Definition 4.2 (see [6], p. 101). A real vector space on which a scalar product
is defined, is called an Euclidean real vector space and it should be denoted by
(E, <, >).
Proposition 4.3 (see [6], p. 101). A scalar product on E has the following properties:
(i) < 0, x >=< x, 0 >= 0, (→) x ∈E
(ii) < x, y + z >=< x, y > + < x, z >, (→) x, y, z ∈E
(iii) < x, α y >= α < x, y >, (→) x, y ∈E, (→) α ∈ R
n
m
n
m
(iv) < α (i) x i , β ( j) y j >= α (i) β ( j) < x i , y j >, (→) x i , y j ∈ E,
i=1 j=1 i=1 j=1
(i) (
(→) α , β ∈ R, i
j) = 1, n, j = 1, m.
n
< x, y >= x (i) y (i) , (→) x, y ∈ Rn , x = x (1) , x (2) , . . . , x (n) , y = y (1) , y (2) , . . . , y (n)
i=1
is a scalar product on Rn .
2. Mn,n (R) , <, > the real Euclidean space of the square matrices, with the scalar
product
< A, B >= trace At B , (→) A, B ∈ Mn,n (R) .
is a scalar product on V3 .
This concret scalar product was the model from which, by abstraction has reached
the concept of the scalar product.
Example 4.4 (see [8]). Let V3 be the vector space of geometric vectors of the usual
physical space. Indicate if the next transformation is a scalar product on V3 :
Solution
If the transformation would be a scalar product, then according to the Definition
4.1 it should be satisfied the condition:
i.e.
∀αx∀ ∀y∀ = α ∀x∀ ∀y∀ , (→) x, y ∈ V3 , (→) α ∈ R.
Since the previous relationship can not be held for α < 0 and x, y ∃= 0 it results
that the given transformation is not a scalar product.
This fact can also be checked using Sage:
4.1 Euclidean Vector Spaces 137
< x, y > B = x1 y1 + x2 y2 + x3 y3 ,
it results ⎫ ⎭
0
(1 + X ) B = ⎬ 1 ⎧ ,
0
< 1 + X, 1 + X > B = 0 · 1 + 1 · 1 + 0 · 0 = 1.
and
138 4 Euclidean Vector Spaces
2 + 4X + 6X 2 = −2 − 2 − 2X + 6 + 6X + 6X 2
= (−2) · 1 + (−2) · (1 + X ) + 6 · 1 + X + X 2
it results ⎫ ⎭
3
3 − X2 = ⎬ 1 ⎧ ,
B
−1
⎫ ⎭
−2
2 + 4X + 6X 2 = ⎬ −2 ⎧ ,
B
6
Definition 4.6 (see [7], p. 155). Let (E, <, >) be a real Euclidean vector space. It’s
called a norm on E, a mapping ∀∀ :E∈ R+ which satisfies the properties:
(i) ∀x∀ > 0, (→) x ∈E and ∀x∀ = 0 ⇔ x = 0 (positivity)
(ii) ∀αx∀ = |α| ∀x∀ , (→) x ∈E, (→) α ∈ R (homogeneity)
(iii) ∀x + y∀ ⇐ ∀x∀ + ∀y∀ , (→) x, y ∈E (Minkowski’s inequality or the triangle
inequality).
4.1 Euclidean Vector Spaces 139
Theorem 4.7 (see [7], p. 154). Let (E, <, >) be a real Euclidean vector space. The
function ∀∀ :E∈ R+ , defined by
⎨
∀x∀ = < x, x >, (→) x ∈ E (4.1)
is a norm on E. The norm defined in the Theorem 4.7 is called the Euclidean norm.
Remark 4.8 (see [7], p. 154). If x ∈ V3 , then its norm (length), in the sense of
Theorem 4.7 coincides with the geometric meaning of its length.
Proposition 4.9 (see [7], p. 154). Let (E, <, >) be a real Euclidean vector space.
(a) For all x, y ∈E, the Cauchy- Schwartz- Buniakowski inequality occurs
Definition 4.10 (see [7], p. 155). Let (E, <, >) be a real Euclidean vector space.
It’s called the angle of the non-zero vectors x, y ∈E, the unique number ϕ ∈ [0, π ]
for which
< x, y >
cos ϕ = . (4.4)
∀x∀ ∀y∀
Definition 4.11 (see [7], p. 156). Let (E, <, >) be a real Euclidean vector space.
We shall say that the vectors x, y ∈E are orthogonal and we shall denote x⇔y if
< x, y >= 0.
Remark 4.12 (see [7], p. 156). The null vector is orthogonal on any vector x ∈E.
Example 4.13 (see [8]). In the vector space R2 one considers B = {e1 , e2 } , ∀e1 ∀ = 2,
∀e2 ∀ = 4, ⇒ (e1 , e2 ) = π3 . The vectors a = 2e1 − 3e2 , b = −e1 + e2 are given.
Compute ⇒ a, b .
Solution
We have:
hence, we obtain:
1
< e1 , e2 >= ∀e1 ∀ ∀e2 ∀ cos ⇒ (e1 , e2 ) = 2 · 4 · = 4.
2
140 4 Euclidean Vector Spaces
It results that
< a, a >=< 2e1 − 3e2 , 2e1 − 3e2 >= 4 < e1 , e1 > −12 < e1 , e2 > +9 < e2 , e2 >= 112
< a, b >=< 2e1 − 3e2 , −e1 + e2 >= −2 < e1 , e1 > +5 < e1 , e2 > −3 < e2 , e2 >= −36
< b, b >=< −e1 + e2 , −e1 + e2 >=< e1 , e1 > −2 < e1 , e2 > + < e2 , e2 >= 12.
⎨ ≺
∀a∀ = < a, a > = 4 7
⎩ ⎩ ≺
⎩b⎩ = < b, b > = 2 3.
Therefore,
(4.4) < a, b > 36
cos ⇒ a, b = ⎩ ⎩ =− ≺ ≺ .
⎩
∀a∀ b ⎩ 4 7·2 3
Proposition 4.14 (see [7], p. 156). Let (E, <, >) be a real Euclidean vector space and
a 1 , a 2 , . . . , a p ∈E be some nonnull vectors which are pairwise orthogonal. Then,
the vectors a 1 , a 2 , . . . , a p are linearly independent.
Definition 4.15 (see [7], p. 156). If x is a nonnull vector of the real Euclidean vector
space (E, <, >) then the vector
x
x0 =
∀x∀
Definition 4.18 (see [7], p. 156). Let (E, <, >) be a real Euclidean vector space and
S √ E.
(a) The system S is orthogonal if its vectors are nonnull and pairwise orthogonal.
(b) The system S is orthonormal (or orthonormat) if it is orthogonal and each of
its vectors has the length equal to 1.
(c) If dim E = n < ≡ then the basis B = {e1 , e2 , . . . , en } of E is called orthonor-
mal if < ei , e j >= δi j , (→) i, j = 1, n,
where
1, i = j
δi j = , i, j = 1, n
0, i ∃= j
We have:
n
n
n
n
(i) ( j)
< x, y >=< x ei , y e j >= x (i) y ( j) < ei , e j > . (4.5)
i=1 j=1 i=1 j=1
We denote
the matrix G signifies the matrix of the scalar product <, > relative to the
basis B.
Remark 4.20 (see [7], p. 157). The
matrix
G is symmetric (G t = G ) and positive
definite ( x Gx > 0, (→) x ∈ R \ 0 ).
t n
n
n
< x, y >= x (i) y ( j) gi j , (4.6)
i=1 j=1
relation that constitutes the analytical expression of the scalar product <, >
relative to the basis B.
Definition 4.22 (see [7], p. 157). The equality (4.6) is equivalent to the equality
called the matrix representation of the scalar product <, > relative to the
basis B.
Example 4.23. One considers the mapping <, >: R3 × R3 ∈ R which, relative to
the canonical basis B = {e1 , e2 , e3 } of R3 has the analytical expression:
< x, y >= (x1 − 2x2 ) (y1 − 2y2 ) + x2 y2 + (x2 + x3 ) (y2 + y3 ) =< y, x >, (→) x, y ∈ R3
< x + z, y >= (x1 − 2x2 ) (y1 − 2y2 ) + (z 1 − 2z 2 ) (y1 − 2y2 ) + x2 y2 + z 2 y2 +
(x2 + x3 ) (y2 + y3 ) + (z 2 + z 3 ) (y2 + y3 ) =< x, y > + < z, y >, (→) x, y, z ∈ R3
< αx, y >= α [(x1 − 2x2 ) (y1 − 2y2 ) + x2 y2 + (x2 + x3 ) (y2 + y3 )]
= α < x, y >, (→) x ∈ R3 , (→) α ∈ R
(b) As
• a = e1 − e2 + 2e3 ≤ a = (1, −1, 2) ,
• b = −e1 + e2 + 9e3 ≤ b = (−1, 1, 9)
we shall obtain < a, b >= 0; hence a and b are orthogonal.
(c) As
(e) We have
⎫ ⎭
1 −2 0
G = ⎬ −2 6 1 ⎧ ,
0 1 1
n
n
< x, y >= x (i) y ( j) (4.8)
i=1 j=1
The equalities (4.8) and (4.9) justifies the importance of considering the ortho-
normal basis which consists in the fact that relative to these bases, the computations
are more simplified.
Definition 4.24 (see [7], p. 159). The matrix A ∈ Mn (R) is orthogonal if At A =
A At = In , In being unit matrix of order n.
Theorem 4.25 (theorem of change the orthonormal bases, see [7], p. 160).
Let (E, <, >) be a real Euclidean vector space of finite dimension n and B1 =
{e1 , e2 , . . . , en }, B2 = {u 1 , u 2 , . . . , u n } are two orthonormal bases of E. Then the
transition matrix from the basis B1 to the basis B2 is orthogonal.
Example 4.26 (see [6], p. 112). In the real Euclidean vector space E one considers
the bases B1 = {e1 , e2 , e3 }, B2 = {u 1 , u 2 , u 3 } . If x is an arbitrary vector from E,
x (i) , y (i) , i = 1, 3 are the coordinates of the vector relative to the bases B1 , B2 , B1
is an orthonormal basis and
(1)
x = 27 y (1) + 37 y (2) + 67 y (3)
x (2) = 67 y (1) + 27 y (2) + αy (3)
(3)
x = −αy (1) − 67 y (2) + 27 y (3)
Solution
In order that B2 be an orthonormal basis it is necessary (within the Theorem 4.25)
that the transition matrix from the basis B1 to the basis B2 , i.e.
⎫ 2 3 6⎭
7 7 7
M(B1 ,B2 ) =
⎬
6
7
2
7 α
⎧
−α − 67 2
7
to be an orthogonal one.
From the condition
it results that α = − 37 .
The solution is Sage is:
Theorem 4.27 (Gram-Schmidt orthogonalization, see [1], p. 26). If (E, <, >) is
a real Euclidean vector space of finite dimension n and B = {a 1 , a 2 , . . . , a n } is a
basis of E then there is a basis B ∗ = {e1 , e2 , . . . , en } of E which has the following
properties:
(i) the basis B ∗ is orthonormal;
(ii) the sets {a 1 , . . . , a k } and {e1 , . . . , ek } generate the same vector subspace Wk √
E, for each k = 1, n.
The Gram- Schmidt orthogonalization procedure (described in detail in [2],
p. 150) can be summarized as:
b1 = a 1
(1)
b 2 = α2 b 1 + a 2
(1) (2)
b 3 = α3 b 1 + α3 b 2 + a 3
.
.
. (4.10)
(1) (i−1)
bi = αi b1 + . . . + αi bi−1 + a i
.
.
.
(1) (n−1)
b n = αn b 1 + . . . + αn bn−1 + a n ,
( j)
where the scalars αi ∈ R, i = 2, n, j = 1, i − 1 are determined from the condition
that bi ⇔ b j , i, j = 1, n, i ∃= j.
146 4 Euclidean Vector Spaces
bi
ei = ⎩ ⎩ , i = 1, n. (4.11)
⎩b i ⎩
1
< P, Q >= P (t) Q (t) dt.
−1
As
1
< f 1 , X >=< 1, X >= 1 · tdt = 0
−1
We obtain
1
2
< f 1 , X >=< 1, X >=
2 2
1 · t 2 dt =
3
−1
4.1 Euclidean Vector Spaces 147
and
1
< f 1 , f 1 >=< 1, 1 >= 1 · 1 dt = 2;
−1
hence α1 = −1/3.
From the orthogonality condition f 2 and f 3 we deduce:
As
1
< f 2 , X 2 >=< X, X 2 >= t · t 2 dt = 0
−1
it results α2 = 0. Therefore
1
f 3 = X2 − .
3
∗∗
(4.11) f
g i = ⎩ i ⎩ , (→) i = 1, 3.
⎩ f i⎩
Whereas
⎩ ⎩ ≺
⎩ f 1⎩ = < f 1, f 1 > = 2
it results that
f 1
g1 = ⎩ 1 ⎩ = ≺ .
⎩ f 1⎩ 2
We compute
1
2
< f 2 , f 2 >=< X, X >= t · t dt = ;
3
−1
148 4 Euclidean Vector Spaces
we shall have ≺
f2 3
g 2 = ⎩ ⎩ = ≺ X.
⎩ f 2⎩ 2
We achieve that:
1 ⎞ ⎟
1 1 1 2 8
< f 3 , f 3 >=< X − , X 2 − >=
2
t −
2
dt =
3 3 3 45
−1
and
≺ ⎞ ⎟
f 3 5 1
g3 = ⎩ 3 ⎩ = ≺ X2 − .
⎩ f 3⎩ 2 2 3
Example 4.29 We consider the real vector space of the symmetric matrices, of the
order n, with real elements, Mns (R) and
<, >: Mns (R) × Mns (R) ∈ R, < A, B >= trace At B .
4.1 Euclidean Vector Spaces 149
Solution
We consider the orthogonal system {B1 , B2 , B3 }, trace Bit B j = 0, (→) i ∃= j, as
follows:
B1 = A1
B2 = A2 + α B1
B3 = A3 + α1 B1 + α2 B2 .
i.e.
therefore
trace (A2 B1 ) 2
α=− =− .
trace (B1 B1 ) 3
We obtain
⎞ ⎟ ⎞ ⎟ ⎞ ⎟
01 2 11 −2/3 1/3
B2 = − = .
12 3 10 1/3 2
We shall obtain
1 4
B3 = A3 + B1 − B2 .
3 7
The orthonotormal system {C1 , C2 , C3 } will be:
Bi
Ci = , i = 1, 3.
∀Bi ∀
In investigating the Euclidean vector spaces are very useful the linear transformations
compatible with the scalar product, i.e. the orthogonal transformations.
Definition 4.30 (see [7], p. 199). Let (E, <, >) be a finite dimensional real Euclid-
ean space. The endomorphism T ∈ End(V ) is called orthogonal operator or
orthogonal transformation if T transforms the orthonormal basis into some ortho-
normal basis, i.e. if B = {e1 , e2 , . . . , en } is an orthonormal basis of E then
B ∗ = {T (e1 ) , T (e2 ) , . . . , T (en )} is an orthonormal basis of E, too.
Theorem 4.31 (see [7], p. 199). For an operator T ∈ End(V ) the following statements
are equivalent:
4.2 Linear Operators in Euclidean Vector Spaces 151
1. T is orthogonal,
2. T is bijective and T −1 is orthogonal,
3. T preserves the scalar product, i.e., < T (x) , T (y) >=< x, y >, (→) x, y ∈ E,
4. T stores the length of vectors, i.e., ∀T (x)∀ = ∀x∀, (→) x ∈ E,
5. the operator matrix T relative to an orthonormal basis of E is orthogonal.
Corollary 4.32 (see [7], p. 200). If T ∈ End(V ) is orthogonal then T preserves the
vector angles, i.e.
Proposition 4.33 (see [7], p. 200). Let T, S ∈ End(V ) be two orthogonal operators
and be α ∈ R. Then:
1. T ∞ S is an orthogonal operator,
2. αT is orthogonal⇔ α = ±1.
We denote by
O (E) = {T ∈ End (V ) | T orthogonal} .
Proposition 4.34 (see [7], p. 200). If T ∈ O (E) and A is the associated matrix of T
relative to an orthonormal basis B of E then det A = ±1.
Definition 4.35 (see [7], p. 200). It’s called an orthogonal operator of the first
kind or the rotation operator, an orthogonal operator for which the determinant of
the associated matrix in an orthonormal basis of E is equal to −1.
Definition 4.36 (see [7], p. 200). It’s called an orthogonal operator of the second
kind, an orthogonal operator for which the determinant of the associated matrix in
an orthonormal basis of E is equal to −1.
We denote by:
• O+ (E) = the set of the orthogonal operators of the first kind,
• O− (E) = the set of the orthogonal operators of the second kind.
Proposition 4.37 (see [7], p. 200). The roots of the characteristic equation of an
orthogonal operator have their absolute values equal to 1. In particular, the eigenval-
ues of an orthogonal operator are equal to ±1.
Proposition 4.38 (see [7], p. 202). For an orthogonal operator, the eigenvectors that
correspond to different eigenvalues are orthogonal.
Theorem 4.39 (see [3], p. 95). The orthogonal matrices of M2 (R) are of the form:
⎞ ⎟ ⎞ ⎟
cos ϕ sin ϕ − cos ϕ sin ϕ
, , (4.13)
sin ϕ − cos ϕ − sin ϕ − cos ϕ
⎞ ⎟ ⎞ ⎟ ⎞ ⎟
− cos ϕ sin ϕ − cos ϕ − sin ϕ cos ϕ − sin ϕ
, , ,
sin ϕ cos ϕ − sin ϕ cos ϕ sin ϕ cos ϕ
⎞ ⎟ ⎞ ⎟
cos ϕ sin ϕ − cos ϕ − sin ϕ
, ,
− sin ϕ cos ϕ sin ϕ − cos ϕ
152 4 Euclidean Vector Spaces
(→) ϕ ∈ [0, 2π ].
Definition 4.40 (see [7], p. 203 and [3], p. 95). An orthogonal matrix with det A = 1
is called a rotation matrix in Rn .
Theorem 4.41 (see [8]). The orthogonal transformations in the Euclidean plane are:
the rotations, the reflections or the compositions of rotations with reflections.
Proposition 4.42 (see [8]). The rotation of the plane vectors around the origin, in
the counterclockwise, with the angle ϕ, rϕ : R2 ∈ R2 ,
rϕ (x) = rϕ x (1) , x (2) = x (1) cos ϕ − x (2) sin ϕ, x (1) sin ϕ + x (2) cos ϕ
(4.14)
is an orthogonal transformation.
Remark 4.43 (see [8]). If O is the center of rotation then each point M has associated
the point M ∗ , such that (see Fig. 4.1):
⎦ ⎩ ⎩ ⎩ ⎩
⎩O M ⎩ = ⎩ ⎩
⎩O M ∗ ⎩ = a
⇒ M O M ∗ = the rotation angle in the counterclockwise.
We have:
T (x) = rϕ (x) = Ax,
where
⎞ ⎟
cos ϕ − sin ϕ
A=
sin ϕ cos ϕ
Proposition 4.44 (see [8]). The rotation through the angle π around the origin,
T (x) = sO (x) = −x (1) , −x (2) (4.15)
coincides with the reflection with respect to the origin (see Fig. 4.2).
Fig. 4.2 Rotation through
angle π around the origin
Proposition 4.45 (see [8]). The reflection across the Ox axis (see Fig. 4.3),
T (x) = sd (x) = x (1) , −x (2) (4.16)
is an orthogonal transformation.
Fig. 4.3 Reflection across the
axis Ox
Proposition 4.46 (see [8]). The reflection across the Oy axis (see Fig. 4.4),
T (x) = sd∗ (x) = −x (1) , x (2) (4.17)
is an orthogonal transformation.
Proposition 4.47 (see [8]). The composition of the rotation rϕ with the reflection sd
is an orthogonal transformation.
Proof
We shall have
(4.15)
T (x) = rϕ ∞ sd (x) = rϕ (sd (x)) = rϕ x (1) , −x (2)
(4.13)
= x (1) cos ϕ + x (2) sin ϕ, x (1) sin ϕ − x (2) cos ϕ .
is an orthogonal transformation.
Proposition 4.48 (see [8]). The composition of the rotation rϕ with the reflection sO
is an orthogonal transformation.
Proof
We shall achieve
(4.14)
T (x) = rϕ ∞ sO (x) = rϕ (sO (x)) = rϕ −x (1) , −x (2)
(4.13)
= −x (1) cos ϕ + x (2) sin ϕ, − x (1) sin ϕ − x (2) cos ϕ .
We can note ⎞ ⎟
− cos ϕ sin ϕ
A=
− sin ϕ − cos ϕ
We deduce
(4.16)
T (x) = x ∗ , y ∗ = rϕ ∞ sd∗ (x) = rϕ sd∗ (x) = rϕ (−x, y)
(4.13)
= (−x cos ϕ − y sin ϕ, − x sin ϕ + y cos ϕ) .
Case 2. We have a rotation, followed by a reflection across the Ox ∗ axis (Fig. 4.6).
We obtain
156 4 Euclidean Vector Spaces
(4.15)
T (x) = x ∗ , y ∗ = rϕ ∞ sd (x) = rϕ (sd (x)) = rϕ (x, −y)
(4.13)
= (x cos ϕ + y sin ϕ, x sin ϕ − y cos ϕ) .
y = 3 + 23 = 3.232.
Proposition 4.50 (see [4], p. 463). The rotation of a rectangular coordinate system
around the origin, in the counterclockwise, through the angle ϕ, rϕ : R2 ∈ R2 is an
orthogonal transformation.
Proof
By rotating the rectangular coordinate system xOy around the origin, in the coun-
terclockwise, through the angle ϕ one gets the system x ∗ Oy ∗ . A point
M which has
the coordinates (x, y) in the old system will have the coordinates x ∗ , y ∗ in the new
system.
We choose in the plane an orthonormal reference with the origin in the center of
rotation (Fig. 4.7).
We note that:
OC1 = x ∗ cos ϕ
AC1 = y ∗ sin ϕ
OC2 = x ∗ sin ϕ
C2 B = y ∗ cos ϕ
and
158 4 Euclidean Vector Spaces
O A = OC1 − AC1 = x ∗ cos ϕ − y ∗ sin ϕ
O B = OC2 + C2 B = x ∗ sin ϕ + y ∗ cos ϕ.
It turns out that the equations corresponding to the transformation of the coordinate
system xOy by rotating it in the counterclockwise, through the angle ϕ will be:
x = x ∗ cos ϕ − y ∗ sin ϕ
y = x ∗ sin ϕ + y ∗ cos ϕ
i.e.
x ∗ = x cos ϕ + y sin ϕ
(4.18)
y ∗ = −x sin ϕ + y cos ϕ.
We obtain:
Rϕ (x) = Rϕ (x, y) = x ∗ cos ϕ − y ∗ sin ϕ, x ∗ sin ϕ + y ∗ cos ϕ ,
where ⎞ ⎟
cos ϕ − sin ϕ
A= ;
sin ϕ cos ϕ
where:
• x ∗ , y ∗ are the coordinates of the point M in the plane reported to the rectangular
axes Ox ∗ , Oy ∗ ;
• x, y, are the coordinates of the point M in the plane reported to the rectangular
axes Ox, Oy;
• θ is the angle to be rotated the axes.
Multiplying the first equation with cos ϕ and the second with − sin ϕ and adding
the obtained equations, we deduce:
x ∗ cos θ − y ∗ sin θ = x,
while multiplying the first equation with sin θ and the second with cos θ and adding
the obtained equations, we deduce:
4.2 Linear Operators in Euclidean Vector Spaces 159
x ∗ sin θ + y ∗ cos θ = y.
By emphasizing the condition that the point M belongs to the Ox ∗ axis (i.e. y ∗ = 0)
we have
∗ ∗
x cos θ = x x cos θ = 1 ≺
⇔ ⇔ x ∗2 = 2 ⊂≤ x ∗ = ± 2.
x ∗ sin θ = y x ∗ sin θ = 1
≺
In the case when x ∗ = 2 it results
≺
cos θ = ≺1 = 2
2 ≺2
sin θ = ≺1 = 2
2 2
i.e. θ = π/4. ≺
In the case when x ∗ = − 2 it results
≺
cos θ = − ≺1 = − 2
2 2 ≺
sin θ = ≺1 = 2
2 2
i.e.
π 5π
θ =π+ = .
4 4
So, the new coordinates of M if:
≺
π
• the axes one rotate with the angle θ = 4 are M 2, 0 ,
≺
• axes one rotate with the angle θ = 5π
4 are M − 2, 0 .
4.3 Problems
is a scalar product.
2. In the real Euclidean vector space (E, <, >) having a basis B = {e1 , e2 } such
that: ∀e1 ∀ = 1, ∀e2 ∀ = 4, ⇒ (e1 , e2 ) = π/4 one considers the vector x =
e1 + 5e2 . Compute ∀x∀.
3. In the real Euclidean vector space R3 one assumes the vectors: a 1 = (1, 0, 3) ,
a 2 = (1, 1, 0) , a 3 = (1, 1, 1).
(a) Determine if B1 = {a 1 , a 2 , a 3 } constitutes a basis of R3 .
(b) Map the system of vectors B1 into one orthonormal.
Solution
Using Sage we shall have:
,
4.3 Problems 161
4. Let R3 be the arithmetic vector space and the scalar products <, >1 , <, >2 :
R3 × R3 ∈ R3 ,defined by:
Solution
Using Sage, we shall have:
Solution
With Sage, it will result:
162 4 Euclidean Vector Spaces
Solution
We shall use Sage:
8. One gives the point M (1, 1) in the plane reported to the rectangular axes
Ox, Oy. Determine the angle that the axes should be rotated so that the
point M belongs to the Oy ∗ axis. Find the new coordinates of M in these
conditions.
9. Let be a triangle, having the vertices A (3, 1) , B (7, 1) , C (7, 4) . Find its image
through the rotation with the center O and the angle π3 .
Solution
Solving this problem in Sage, we achieve:
4.3 Problems 163
10. One considers the rotation through the angle ϕ in the counterclockwise, rϕ :
R2 ∈ R2 .
(a) Justify the linearity of the transformation rϕ .
(b) Build the associated matrix in the canonical basis from R2 and in the basis
B1 = {e1 + e2 , e1 − e2 }.
164 4 Euclidean Vector Spaces
References
The theory of bilinear form and quadratic form is used [5] in the analytic geometry
for getting the classification of the conics and of the quadrics.
It is also used in physics, in particular to describe physical systems subject to small
vibrations. The coefficients of a bilinear form one behave to certain transformations
like the tensors coordinates. Tensors are useful in theory of elasticity (the deformation
of an elastic medium is described through the deformation tensor).
Definition 5.1 (see [1], p. 150). A mapping b : V × V ∈ K is called a bilinear
form on V if it satisfies the conditions:
1. b (πx + λ y, z) = πb (x, z) + λb (y, z) , (→) π, λ ∈ K , (→) x, y, z ∈ V,
2. b (x, π y + λz) = πb (x, y) + λb (x, z) , (→) π, λ ∈ K , (→) x, y, z ∈ V.
Definition 5.2 (see [1], p. 150). We say that the bilinear form b : V × V ∈ K is sym-
metric (antisymmetric) if b (x, y) = b (y, x) (respectively, b (x, y) = −b (y, x).
Consequences 5.3 (see [2], p. 116). If the mapping b : V × V ∈ K is a bilinear
form then:
(1) b 0, x = b x, 0 = 0, (→) x∈V
n n
(2) (a) b π (i) x i , y = π (i) b (x i , y) ,
i=1 i=1
1
b (x, y) = [ f (x + y) − f (x) − f (y)] , (→) x, y ∈ V (5.1)
2
Definition 5.6 (see [1], p. 150). The symmetric bilinear form b associated to the
quadratic form f is called the polar form of the quadratic form f .
Example 5.7 (see [3], p. 93). The quadratic form corresponding to the real scalar
product (which is a symmetric bilinear form) is the square of the Euclidean norm:
n
n
x= x (i) a i , y = y ( j) a j ;
i=1 j=1
therefore
n
n
b (x, y) = ai j x (i) y ( j) , (5.2)
i=1 j=1
where
ai j = b a i , a j , (→) i, j = 1, n.
The expression (5.2) constitutes [1] the analytic expression of the bilinear form b
relative to the basis B, and A ∈Mn (K ), A = ai j 1∀i, j∀n represents the associated
matrix of the bilinear form b relative to the basis B.
From (5.2) one obtains [1] the analytic expression of the bilinear form f : V ∈ K
relative to the basis B of V :
n n
n
f (x) = ai j x (i) y ( j) , (→) x = x (i) a i ∈ V. (5.3)
i=1 j=1 i=1
Definition 5.8 (see [1], p. 151) We call the associated matrix of a quadratic
form f : V ∈ K relative to a basis of V , the matrix of the bilinear mapping
b : V × V ∈ K from which derives f relative to the considered basis.
Example 5.9 (see [4]) Let be b : R4 × R4 ∈ R,
and relative to the canonical basis and highlight the link between them.
(c) Determine the expression of the quadratic form f associated to b.
Solution
(a) According to the Definition 5.1, b is a bilinear functional if those two conditions
are accomplished. We shall check the first condition as for the others one proceeds
similarly. Let be π, λ ∈ K and x, y, z ∈ V ; we have:
We achieve: ⎭ ⎩
5 2 2 3
⎧ 4 2 2 2
A⇒ = ⎧
⎨5
.
2 1 2
2 0 −1 2
1 0 0 −1
168 5 Bilinear and Quadratic Forms
We have:
b (x, y) = x tB Ay B
and
A⇒ = Mt(B,B ⇒ ) AM(B,B ⇒ ) .
Definition 5.10 (see [1], p. 152) The rank of the quadratic form f is the rank of its
matrix relative to a basis of V and one denotes with rank f .
Remark 5.11 (see [2], p. 122) Because of the symmetry of the associated matrix of
a quadratic form, relative to a basis B of V , the relation (5.3) is written
n 2
n
f (x) = aii x (i) + 2 ai j x (i) x ( j) . (5.4)
i=1 i, j=1
i< j
Definition 5.12 (see [2], p. 122) If the associated matrix of the quadratic form f :
V ∈ K relative to the basis B = {e1 , e2 , . . . , en } of V is diagonal, i.e. A =
diag (π1 , . . . , πn ) ; we shall say that:
• the basis B is a canonical basis for f,
• the analytical expression of f relative to the basis B, i.e.
n 2
n
f (x) = πi x (i) , (→) x = x (i) ei ∈ V (5.5)
i=1 i=1
If f is the quadratic null form, then f has the canonical expression in any basis
of V . Hence, we can assume that isn’t null.
We can also assume that (∃) i = 1, n such that aii ⇐= 0. Otherwise, if ar p ⇐= 0,
for r ⇐= p then we make the change of coordinates:
170 5 Bilinear and Quadratic Forms
⎫ x (r ) = t (r ) + t ( p)
x ( p) = t (r ) − t ( p) (5.6)
⎬
x (i) (i)
= t , i ∈ {1, . . . , n} \ {r, p}
2
n
n
f (x) = a11 x (1) + 2 a1k x (1) x (k) + ai j x (i) x ( j) . (5.7)
k=2 i, j⇐=1
We shall add and subtract the necessary terms in (5.7) to write it in the form:
1 2
n
f (x) = (1) (2)
a11 x + a12 x + . . . + a1n x (n)
+ ai⇒ j x (i) x ( j) , (5.8)
a11
i, j=2
n
where ai⇒ j x (i) x ( j) doesn’t contain x (1) .
i, j=2
We make the change of coordinates:
(1)
z = a11 x (1) + a12 x (2) + . . . + a1n x (n)
⎫ z (2) = x (2)
..
.
⎬
z (n) = x (n)
..
.
⎬
x (n) = z (n) .
The transition to the new coordinates z (1) , z (2) , . . . , z (n) is achieved through the
relation:
x B = M(B,B1 ) · x B1 , (5.9)
The form Q has the following analytical expression relative to the basis B1 :
1 (1) 2
n
f (x) = z + ai⇒ j x (i) x ( j) . (5.10)
a11
i, j=2
The sum
n
Q1 = ai⇒ j x (i) x ( j)
i, j=2
from the right member of the relation (5.10) is a quadratic form in n − 1 variables,
therefore can be treated by the process described above, as well
as the form Q.
Finally, after at most n − 1 steps we obtain a basis B ⇒ = e⇒1 , e⇒2 , . . . , e⇒n of V ,
relative to which the quadratic form Q is reduced to the canonical expression.
Example 5.14 (see [4]). Let be the quadratic form
2 2
Q : R4 ∈ R, Q (x) = x (1) + 2x (1) x (2) + 2 x (2) − 4x (2) x (3)
2 2
+ x (3) + x (1) x (4) − x (4) .
is ⎭ ⎩
1 1 0 1/2
⎧ 1 2 −2 0
A=⎧
⎨ 0
.
−2 1 0
1/2 0 0 −1
2 2
x (4)
Q (x) = x (1) + x (2) + + x (2) − 4x (2) x (3)
2
5 (4) 2 2
− x − x (2) x (4) + x (3) .
4
By making the change of coordinates:
x (4)
y (1) = x (1) + x (2) +
⎫ (2) 2
y = x (2)
y (3) = x (3)
⎬ (4)
y = x (4)
it results y (4)
x (1) = y (1) − y (2) −
⎫ (2) 2
x = y (2)
(3) = y (3)
⎬ x (4)
x = y (4) .
The transition matrix associated with this change of coordinates will be:
⎭ ⎩
1 −1 0 −1/2
⎧0 1 0 0
M(B,B1 ) =⎧
⎨0
,
0 1 0
0 0 0 1
2 2 2 3 2
(1) (2) (3) 1
Q= y + y − 2y − y (4) − 3 y (3) − y (4) − 2y (3) y (4) .
2 2
hence (1)
y = z (1)
⎫ (2)
y = z (2) + 2z (3) + 21 z (4)
y (3) = z (3)
⎬ (4)
y = z (4) .
The transition to the new coordinates z (1) , z (2) , . . . , z (n) is achieved through the
relation
x B1 = M(B1 ,B2 ) · x B2 ,
2 2 1 7 2
Q = z (1) + z (2) − 3z (3) + z (4) − z (4) .
3 6
We shall make the change of coordinates:
(1)
t = z (1)
⎫ t (2) = z (2)
t (3) = 3z (3) + z (4)
⎬ (4)
t = z (4) ;
we have ⎭ ⎩
1 0 0 0
⎧0 1 0 0
M(B2 ,B3 ) =⎧
⎨0
.
0 1/3 −1
0 0 0 1
is 2 2 1 2 7 2
Q = t (1) + t (2) − t (3) − t (4) ,
3 6
so we have obtained the canonical expression of Q.
We shall get:
It follows that transition matrix from the initial basis B of the space R4 to the
basis B3 , relative to which Q has the canonical expression is:
Theorem 5.15 (Jacobi, see [3], p. 100). Let V be ann finite dimensional vector space
over K , f : V ∈ K , a quadratic form and A = ai j 1∀i, j∀n its relative matrix to
the basis B = {e1 , e2 , . . . , en } of V .
If all the principal minors
ϕ1 = a⎪ 11 ⎪
⎪ ⎪
⎫ ϕ2 = ⎪ a11 a12 ⎪
⎪ a21 a22 ⎪
(5.11)
..
.
⎬
ϕn = det A
are all non- null, then there is a basis B ⇒ = e⇒1 , e⇒2 , . . . , e⇒n of V , relative to which
the quadratic form Q has the canonical expression
ϕi−1 2
n
f (x) = y (i) , (5.12)
ϕi
i=1
where
• y (i) , i = 1, n are the coordinates of x in the basis B ⇒ ,
• ϕ0 = 1.
Proof
We are looking for the vectors e⇒1 , e⇒2 , . . . , e⇒n by the form
⇒
e1 = c11 e1
e⇒2 = c21 e1 + c22 e2
..
⎫
.
⇒ (5.13)
e = ci1 e 1 + ci2 e2 + . . . + cii ei
i
..
.
⎬ ⇒
en = cn1 e1 + cn2 e2 + . . . + cnn en ,
We obtain:
j = 1 : b ei⇒ , e1 = ci1 a11 + ci2 a12 + . . . + cii a1i = 0
⎫ j = 2 : b ei⇒ , e2 = ci1 a21 + ci2 a22 + . . . + cii a2i = 0
.. (5.16)
.
⇒
j = i − 1 : b e , e
i i−1 = ci1 i−1,1 + ci2 ai−1,2 + . . . + cii ai−1,i = 0
a
⎬
j = i : b ei⇒ , ei = ci1 ai1 + ci2 ai2 + . . . + cii aii = 1
We deduce that the quadratic form has the following canonical expression in the
basis B ⇒ :
ϕi−1 (i) 2
n
f (x) = ai⇒ j y (i) y ( j) = y
ϕi
i, j=1
Example 5.16 (see [3], p. 101). Using the Jacobi method find the canonical expression
and the basis in which to do this for the quadratic form
Solution
The matrix of the quadratic form relative to the canonical basis of the space R3 is
⎭ ⎩
1 −4 −8
A = ⎨ −4 7 −4 .
−8 −4 1
n
ϕi−1 ϕ 0 2 ϕ 1 2 ϕ2 2
Q (x) = yi2 = y + y + y
ϕi ϕ 1 1 ϕ2 2 ϕ 3 3
i=1
1 2 1 2
= y12 −
y2 + y .
9 81 3
We shall determine the new basis B ⇒ = e⇒1 , e⇒2 , . . . , e⇒n , relative to which Q has
the canonical expression:
⇒
⎫ e1 = c11 e1
e⇒ = c21 e1 + c22 e2
⎬ 2⇒
e3 = c31 e1 + c32 e2 + c33 e3 ,
1 1
b (x, y) = x1 y1 − x2 y2 + x3 y3 .
9 81
We have:
b ei⇒ , e1 = b (c11 e1 , e1) = c11
b (e1 , e1 ) = c11 a11 = c11
⇔ c11 = 1;
b e⇒1 , e1 = 1
therefore: e⇒1 = e1 .
We shall compute:
b e⇒2 , e1 = b (c21 e1 + c22 e2 , e1 ) = c21 b (e1 , e1 ) + c22 b (e2 , e1 )
= c21 a11 + c22 a21 = c21 − 4c22
⇒
b e2 , e2 = b (c21 e1 + c22 e2 , e2 ) = c21 b (e1 , e2 ) + c22 b (e2 , e2 )
= c21 a12 + c22 a22 = −4c21 + 7c22 .
i.e.
4 1
e⇒2 = − e1 − e2 .
9 9
5.3 Reducing a Quadratic Form to a Canonical Expression by Jacobi Method 179
⎫ c31 − 4c32 − 8c33 = 0
8 4 1
−4c31 + 7c32 − 4c33 = 0 ⇔ c31 = − , c32 = − , c33 = ;
⎬ 81 81 81
−8c31 − 4c32 + c33 = 1
it results:
8 4 1
e⇒3 = − e1 − e2 + e3 .
81 81 81
The solution in Sage will be given, too:
180 5 Bilinear and Quadratic Forms
Theorem 5.17 (Eigenvalue method, see [1], p. 153). Let V be an Euclidean real
vector space and let f : V ∈ R be a real quadratic form. Then there is an orthonormal
basis B ⇒ = e⇒1 , e⇒2 , . . . , e⇒n of the vector space V relative to which the canonical
expression of the form is
n 2
f (x) = ρi y (i) , (5.19)
i=1
where:
• ρ1 , . . . , ρn are the eigenvalues of the associated matrix of the quadratic form,
relative to an orthonormal basis B (each eigenvalue being included in sum such
many times as its multiplicity),
• y (1) , . . . , y (n) are the coordinates of the vector x relative to the basis B ⇒ .
To apply the eigenvalue method for reducing a quadratic form to canonical expres-
sion one determines as follows:
1. choose an orthonormal basis B = {e1 , e2 , . . . , en } of V and write the matrix A,
associated to f relative to the basis B;
2. determine the eigenvalues: ρ1 , . . . , ρr ∈ R of the matrix A, with the correspond-
ing algebraic multiplicities aρ1 , . . . , aρr , with aρ1 + . . . + aρr = n ;
3. for the eigensubspaces Wρ1 , . . . , Wρr associated to the eigenvalues ρ1 , . . . , ρr
determine the orthonormal bases B1 , . . . , Br , using the Gram-Schmidt orthogo-
nalization procedure;
4. one considers the orthonormal basis: B ⇒ = B1 ≺ . . . ≺ Br of V and one writes
the canonical expression of f relative to the basis B ⇒ with (5.19), where x B ⇒ =
(1) t
y , . . . , y (n) .
Example 5.18 (see [4]). Use the eigenvalue method to determine the canonical
expression and the basis relative to which can be made this, for the quadratic form:
5.4 Eigenvalue Method for Reducing a Quadratic Form into Canonical Expression 181
Solution
The associated matrix of f relative to the canonical basis of the space R3 is
⎭ ⎩
1 1/2 1/2
A = ⎨ 1/2 1 1/2 .
1/2 1/2 1
We have:
⎪ ⎪
⎪ 1 − ρ 1/2 1/2 ⎪ 2
⎪ ⎪ 1
P (ρ) = ⎪⎪ 1/2 1 − ρ 1/2 ⎪⎪ = (2 − ρ) −ρ ,
⎪ 1/2 1/2 1 − ρ ⎪ 2
We deduce:
⎫ x1 + 21 x2 + 21 x3 = 2x1 ⎫ −x1 + 21 x2 + 21 x3 = 0
1
x + x + x = 2x2 √
1 1
x − x + 1x = 0
⎬ 21 1 12 2 3 ⎬ 12 1 1 2 2 3
2 x 1 + 2 x 2 + x 3 = 2x 3 2 x 1 + 2 x 2 − x 3 = 0.
Denoting x3 = t, t ∈ R we achieve:
−2x1 + x2 = −t
⇔ x1 = t, x2 = t.
x1 − 2x2 = −t
Therefore
⎫
Wρ1 = x ∈ R3 | x = (t, t, t) = t · (1, 1, 1) = tc1 .
⎬ ⎞ ⎟⎠ ⎦
c1
The orthonormal basis B1 will be B1 = f 1 , where
c1 1 1 1 1
f1 = = ≡ c1 = ≡ ,≡ ,≡ .
c1 3 3 3 3
182 5 Bilinear and Quadratic Forms
We achieve:
1 1 1
x1 + x2 + x3 = x1
⎫
2 2 2
1 1 1
x1 + x2 + x3 = x2 √
2
2 2
⎬ 1x + 1x + x = 1x
1 2 3 3
2 2 2
1 1 1
x1 + x2 + x3 = 0 √ x1 + x2 + x3 = 0.
2 2 2
We denote x1 = t1 , x2 = t2 , t1 , t2 ∈ R; hence x3 = −t1 − t2 .
Therefore:
⎫
Wρ2 = x ∈ R3 | x = (t1 , t2 , −t1 − t2 ) = t1 (1, 0, −1)
⎬ ⎞ ⎟⎠ ⎦
c2
+ t2 (0, 1, −1) = t1 c2 + t2 c3 . (5.20)
⎞ ⎟⎠ ⎦
c3
⎜ ⇒ ⇒
⎝
We consider the orthogonal system f 2 , f 3 , where
⎛ ⇒
f 2 = c2
⇒ ⇒
f 3 = c3 + π f 2 ,
⇒ ⇒
where π is obtained from the condition that f 3 and f 2 to be orthogonal, i.e.
⇒ ⇒
< f 3 , f 2 >= 0.
From
⇒
⇒ ⇒ < c3 , f 2 > 1
< c3 + π f 2 , f 2 >= 0 ⇔ π = − ⇒ ⇒ =− .
< f 2, f2 > 2
It results
⇒ 1 ⇒ 1 1 1
f 3 = c3 − f 2 = (0, 1, −1) − (1, 0, −1) = − , 1, − .
2 2 2 2
5.4 Eigenvalue Method for Reducing a Quadratic Form into Canonical Expression 183
The basis B = f 2 , f 3 is orthonormal, where
⇒
f2 1 ⇒ 1 1
$ $
f 2 = ⇒ = ≡ f 2 = ≡ , 0, − ≡ ,
$ $ 2 2 2
$ f 2$
⇒ ≡ ≡
f3 2 ⇒ 1 2 1
f 3 =$ ⇒ $ = ≡ f 3 = −≡ , ≡ , −≡ .
$ $ 3 6 3 6
$ f 3$
We achieve:
B ⇒ = B1 ≺ B2 = { f 1 , f 2 , f 3 }.
1 2 1 2
f (x) = 2y12 + y + y .
2 2 2 3
We need the following Sage code to implement this method:
184 5 Bilinear and Quadratic Forms
(ii) The number q of the negative coefficients from a canonical expression of the
quadratic form f is called the negative index of f .
(iii) The pair ( p, q, d) is called the signature of the quadratic form, where d =
n − ( p + q) is the number of the null coefficients.
The following theorem allows us to decide if a quadratic form is positive or negative
definite, without being obliged to determine one of its canonical expression.
Theorem 5.24 (Sylvester’s criterion, inertia theorem, see [3], p. 104). Let V be
an n finite dimensional real vector space and A = ai j 1∀i, j∀n , A ∈Mn (R) be a
symmetric matrix associated of the quadratic form f : V ∈ R relative to the basis
B = {e1 , e2 , . . . , en } of V . Then
1. f is positive definite if and only if all the principal minors ϕ1 , ϕ2 , . . . , ϕn of
the matrix A are strictly positive,
2. f is negative definite if and only if (−1)k ϕk > 0, (→) k = 1, n.
Remark 5.25 (see [3], p. 104).
(i) The quadratic form f is positive (negative) definite if and only if rank f =
n = p (respectively rank f = n = q).
(ii) The law of inertia states that following any of the three methods to obtain the
canonical expression of a quadratic form, the signature of the quadratic form
(inferred from obtained canonical the expression) is always the same.
(iii) Given a quadratic form f : V ∈ R, its associated matrix relative to a basis of
the space V , f is positive definite if and only if any of the following conditions
are satisfied:
Example 5.26 (see [1], p. 156). Let f : R4 ∈ R be a quadratic form whose analytical
expression form relative to the canonical basis of R4 is
(a) Write the matrix of f relative to the canonical basis of R4 and the analytical
expression of the polar of f , relative to the same basis.
(b) Use the Gauss method to determine a canonical expression for f and a basis of
R4 , relative to which f has this canonical expression.
(c) Indicate the signature of f .
Solution
(a) The matrix of the quadratic form relative to the canonical basis of R4 is
186 5 Bilinear and Quadratic Forms
y1 y2 y3 y4
⎭ ⎩
x1 0 1/2 0 1/2
x ⎧ 1/2 0 −1/2 0
A = 2⎧ .
x3 ⎨ 0 −1/2 0 1/2
x4 1/2 0 1/2 0
Remark 5.27 (see [4]). In the writing of the matrix A occurs both x1 , x2 , x3 , x4 and
y1 , y2 , y3 , y4 to obtain the analytical expression of the polar of f : multiply each
element of the matrix A with the index cooresponding to the line denoted by xi
respectively of the column, denoted by y j to intersection which is this element.
The analytical expression of the polar of f relative to the canonical basis of R4
will be
1 1 1 1 1 1
b (x, y) = x1 y2 + x1 y4 + x2 y1 − x2 y3 − x3 y2 + x3 y4
2 2 2 2 2 2
1 1
+ x4 y1 + x4 y3 , (→) x, y ∈ R4 . (5.21)
2 2
(b) As a12 ⇐= 0 we make the change of coordinates:
x1 = y1 + y2
y1 = 1
2 x1 + 21 x2
⎫ ⎫
x2 = y1 − y2 y2 = 1
2 x1 − 21 x2
⇔
x3
= y3 y3
= x3
⎬ ⎬
x4 = y4 y4 = x4 .
i.e.
f (x) = y12 − y22 − y1 y3 + y2 y3 + y3 y4 + y1 y4 + y2 y4 .
5.5 Characterization Criteria for Positive (Negative) Definite Matrices 187
it results
y = z 1 + 21 z 3 − 21 z 4
⎫ 1
y2 = z2
y = z3
⎬ 3
y4 = z4.
The transition matrix associated with this change of coordinates, through the
relation
x B1 = M(B1 ,B2 ) x B2
will be: ⎭ ⎩
1 0 1/2 −1/2
⎧0 1 0 0
M(B1 ,B2 ) =⎧
⎨0
,
0 1 0
0 0 0 1
1 1 3
f (x) = z 12 − z 22 − z 32 − z 42 + z 3 z 4 + z 2 z 3 + z 2 z 4 .
4 4 2
188 5 Bilinear and Quadratic Forms
it results
z1 = t1
⎫
z2 = −t2 + 21 t3 + 21 t4
z = t3
⎬ 3
z4 = t4 .
The transition matrix associated with this change of coordinates, through the
relation
x B2 = M(B2 ,B3 ) x B3
will be: ⎭ ⎩
1 0 0 0
⎧0 −1 1/2 1/2
M(B2 ,B3 ) =⎧
⎨0
,
0 1 0
0 0 0 1
⇒⇒⇒ = 0, a ⇒⇒⇒ ⇐ = 0.
We note that a33 34
Making the change of coordinates
t1 = u1
⎫
t2 = u2
t3 = u3 + u4
⎬
t4 = u3 − u4;
it results
u 1 = t1
⎫u = t
2 2
u 3 = 2 (t3 + t4 )
1
⎬
u 4 = 21 (t3 − t4 ) .
The associated transition matrix of this change of coordinates, through the relation
x B3 = M(B3 ,B4 ) x B4
will be ⎭ ⎩
1 0 0 0
⎧0 1 0 0
M(B3 ,B4 ) =⎧
⎨0
,
0 1 1
0 0 1 −1
f (x) = u 21 − u 22 + 2u 23 − 2u 24 .
(c) We have
rank f = rank A = 4, p = 2, q = 2.
We obtain that the quadratic form f has the signature (2, 2, 0).
We can also solve this problem in Sage, too:
5.6 Problems
f (x) = x1 x2 + x2 x3 .
(a) Write the matrix of f relative to the canonical basis of R3 and the analytical
expression corresponding to the polar of f , relative to the same basis.
5.6 Problems 191
(b) Use the Gauss method to determine a canonical expression for f and a basis
of R3 , relative to which f has this canonical expression.
(c) Indicate the signature of f .
Solution
Solving this problem in Sage, we obtain:
f (x) = x12 + 5x22 + 4x32 − x42 + 6x1 x2 − 4x1 x3 − 12x2 x3 − 4x2 x4 − 8x3 x4 .
Use the Gauss method to determine a canonical expression for f and a basis of
R4 , relative to which f has this canonical expression.
192 5 Bilinear and Quadratic Forms
b ( x, y) = x1 y2 − x2 y1 + x1 y3 − x3 y1 + x1 y4 − x4 y1 + x2 y3 − x3 y2
+ x2 y4 − x4 y2 + x3 y4 − x4 y3 , (→) x, y ∈ R4 (5.22)
Solution
Using Sage, we shall have:
5. Let B = {e1 , e2 , e3 } be the canonical basis of the arithmetic vector space R3 and
let b : R3 × R3 ∈ R be the bilinear form for which:
⎫ b (e1 , e1 ) = −1, b (e2 , e2 ) = 3, b (e3 , e3 ) = −6
b (e1 − e2 , e2 ) = 2, b (e2 , e1 + 2e2 ) = 5, b (e3 − e1 , e1 ) = 4
⎬
b (2e1 + e2 , e3 ) = −7, b (e1 + e3 , e2 ) = 4, b (e1 − 2e2 , e3 ) = −1.
(a) Write the matrix corresponding to the bilinear functional b, relative to the
basis B.
(b) Is b a symmetic bilinear functional?
Solution
With Sage, we achieve:
5.6 Problems 193
Solution
We shall solve in Sage this problem:
194 5 Bilinear and Quadratic Forms
Solution
The solution in Sage of this problem is:
5.6 Problems 195
8. Use the eigenvalue method to determine the canonical expression and the basis
in which makes this for the quadratic form:
10. Reduce to the canonical expression through the three method, the following
quadratic form:
4
f : R4 ∈ R, f (x) = xi x j .
i< j
i=1
196 5 Bilinear and Quadratic Forms
References
Definition 6.1 (see [5], p. 544). The regular arc of a curve is defined as the set π
of the points M (x, y, z) from the real three-dimensional Euclidean space R3 , whose
coordinates x, y, z check one of the following systems of equations:
F (x, y, z) = 0
, (∈) (x, y, z) → D ⊆ R3 (the implicit representation) (6.1)
G (x, y, z) = 0
z = f (x, y)
, (∈) (x, y) → D ⊆ R2 (the explicit representation) (6.2)
z = g (x, y)
x = f 1 (t)
y = f 2 (t) , (∈) t → (a, b) (the parametric representation) , (6.3)
z = f 3 (t)
is not equal to 0.
Definition 6.2 (see [5], p. 545). The regular curve is the reunion of the regular curve
arcs.
Definition 6.3 (see [5], p. 545). Let π be a curve given by its parametric equations
x = x (t)
y = y (t) , (∈) t → (a, b) (6.4)
z = z (t)
Definition 6.4 (see [5], p. 552). One considers the regular curve π and let be the
points M, M1 → π. The limit position of the chord M, M1 when M1 tends to M is
called the tangent to the curve at the point M (see Fig. 6.1).
x − x0 y − y0 z − z0
T :
= = , (6.6)
x (t0 ) y (t0 ) z (t0 )
6.2 Tangent and Normal Plane to a Curve in the Space 199
where
x = x (t0 )
y = y (t0 ) , (∈) t0 → (a, b) ;
z = z (t0 )
(b) the curve is given implicitly (see the relation (6.1)) are
x − x0 y − y0 z − x0
T : D(F,G)
= D(F,G)
= D(F,G)
; (6.7)
D(y0 ,z 0 ) D(z 0 ,x0 ) D(x0 ,y0 )
(c) the curve is given explicitly (see the relation (6.2)) are
x − x0 y − y0 z − x0
T : = = . (6.8)
g y0 − f y0 f x0 − gx 0 f x0 g y0 − f y0 gx 0
Definition 6.5 (see [5], p. 553). One considers the regular curve π and let be M → π.
The normal plane to the curve π in the point M is the plane π N perpendicular to
the tangent T to the curve π in the point M.
The equation of the normal plane π N to the curve π in the point M, in the case
when:
(a) the curve is given parametrically (as in the relation (6.4)) is:
Definition 6.6 (see [5], p. 576). Let π be a regular curve and the point M0 → π.
The Frenet Trihedron attached to the curve π in the point M0 is a right trihedron
determined by the versors τ , β, ν (Fig. 6.2).
200 6 Differential Geometry of Curves and Surfaces
x − x (t0 ) y − y (t0 ) z − z (t0 )
π0 :
= 0; (6.12)
x (t0 ) y (t0 ) z (t0 )
x − x0 y − y0 z − z0
D(F,G) D(F,G) D(F,G)
π0 :
= 0; (6.13)
D2 (F,G) D2 (F,G) D2 (F,G)
2
D (y0 ,z 0 ) D (z 0 ,x0 ) D (x0 ,y0 )
2 2
x − x0 y − y0 z − z0
g
π0 :
y0 − f f − g f g − f g
y0 x0 x0 x0 y0 y0 x0
= 0. (6.14)
g − f f − g f g − f g
y0 y0 x0 x0 x0 y0 y0 x0
6.3 Frenet Trihedron. Frenet Formulas 201
If
v = a1 i + a2 j + a3 k, (6.15)
then the equation of the rectified plane to the curve π in the point M0 (x0 , y0 , z 0 ) → π
is:
πr : a1 ( x − x (t0 )) + a2 (y − y (t0 )) + a3 (z − z (t0 )) = 0. (6.16)
1. the tangent in M0 to the curve (the tangent versor is denoted by τ ); the equations
of the tangent to the curve π in a point M0 (x0 , y0 , z 0 ) → π can be expressed with
the relations (6.6), (6.7) or (6.8);
2. the binormal of a curve in M0 (the normal which is perpendicular to the osculator
plane, that passes through the point M0 ; the binormal versor is denoted by β) has
the equations:
x − x0 y − y0 z − z0
β :
=
=
.
y (t0 ) z (t0 )
z (t0 ) x (t0 )
x (t0 ) y (t0 )
y (t0 ) z (t0 )
z (t0 ) x (t0 )
x (t0 ) y (t0 )
3. the principal normal of a curve in M0 (the straight line contained in the normal
plane and in the osculator plane passing through M0 ; the principal normal versor
is denoted by ν) has the equations:
x − x0 y − y0 z − z 0
πr :
t1 t2 t3
= 0, (6.18)
b1 b2 b3
Definition 6.7 (see [5], p. 550). The arc element of a curve π is the differential ds
of the function s = s (t), which signifies the length of the respective arc, from the
curve π.
If the curve π is given by the parametric equations (6.4), then
202 6 Differential Geometry of Curves and Surfaces
⎫
ds = dx 2 + dy 2 + dz 2 . (6.19)
If we dispose by the vector equation from the relation (6.5) of the curve π, then
ds = dr . (6.20)
Find the equations of the edges (the tangent, the principal normal and the binormal)
and of the planes (the normal plane, the rectified plane and the osculator plane)
corresponding to the Frenet trihedron in the point t = 1.
Solution
The tangent equations are:
6.3 Frenet Trihedron. Frenet Formulas 203
x − 2t y − t2 z − ln t
(τ ) : = = 1
.
2 2t t
y (1) z (1)
z (1) x (1)
x (1) y (1)
i.e.
x −2 y−1 z
β :
=
=
2 1
1 2
2 2
2 −1
−1 0
0 2
or
x −2 y−1 z
β : = = .
−4 2 4
π N : 2 (x − 2) + 2 (y − 1) + z = 0
or
π N : 2x + 2y + z − 6 = 0.
x − x (1) y − y (1) z − z (1)
π0 :
= 0;
x (1) y (1) z (1)
i.e
204 6 Differential Geometry of Curves and Surfaces
x − 2 y − 1 z − 1
π0 :
2 2 1
= 0
0 2 −1
or
πo : 2x − y − 2z = 3.
The principal normal is the intersection between the osculator plane and the normal
plane:
2x − y − 2z = 3
(ν) :
2x + 2y + z = 6.
As the rectified plane is the plane that passes through the point t = 1 and contains
the director vectors of the tangent and of the binormal in this point, based on (6.18),
the equation of the rectified plane will be:
x − x (1) y − y (1) z − z (1)
πr :
a1 a2 a3
= 0,
b1 b2 b3
where:
it will result:
x −2 y −1 z
πr :
2 2 1
= 0.
−4 2 4
The Frenet’s formulas establish some relations between the edge versors that bears
its name and their derivatives.
Theorem 6.10 (see [5], p. 582). Let π be a regular curve and the point M → π be a
current point, having the position vector r . Let τ , β, ν be the versors of the tangent,
binormal and the principale normal in M. If ds is the arc element on the curve π,
then the following relations are satisfied, called Frenet’s formulas:
1. the first Frenet formula:
dτ ν
= (6.25)
ds R
2. the second Frenet formula:
dβ ν
=− (6.26)
ds T
3. the third Frenet formula:
⎧ ⎨
dν τ β
=− − . (6.27)
ds R T
206 6 Differential Geometry of Curves and Surfaces
The scalars 1/R and 1/T, that are introduced through the Frenet formulas are called
the curvature and the torsion of the given curve in the point M.
We shall examine the geometric interpretation of these scalars.
Definition 6.11 (see [5], p. 579). The mean curvature is the variation of the tangent
direction, per arc unit; the curvature at a point of a curve is the limit of the mean
curvature, when the considered arc element tends to 0 (see Fig. 6.3), i.e.
λτ
K = lim
. (6.28)
λs∀0 λs
Definition 6.12 (see [5], p. 579). The curvature radius in a point from a curve is
equal to the inverse of curvature at that point and is denoted with R.
Remark 6.13 (see [2], p. 30). The curvature of the curve indicates the speed with
which the curve moves away from the tangent.
Theorem 6.14 (see [5], p. 586). Let π be a regular curve. The necessary and sufficient
condition that this curve to be a straight line is that K = 0.
Definition 6.15 (see [5], p. 581). The torsion in a point of a curve is the variation of
the binormal direction, per unit of arc, when the considered arc element tends to 0.
Definition 6.16 (see [5], p. 581). The torsion radius in a point of a curve is equal to
the inverse of the torsion at that point and is denoted by T .
Remark 6.17 (see [2], p. 30). The torsion of the curve indicates the speed with which
the curve moves away from the osculator plane.
6.4 Curvature and Torsion of the Space Curves 207
Theorem 6.18 (see [5], p. 587). Let π be a regular curve. The necessary and sufficient
condition that this curve to be a plane curve is that T1 = 0.
Theorem 6.19 (computing the curvature, see [5], p. 589). Let π be a regular curve
and M → π be the current point, with position vector r , ds be the arc element on the
curve π and R be the radius of curvature of the the curve π in the point M. Then
⎬ ⎬
⎬ ⎬
1 ⎬r × r "⎬
K = = ⎬ ⎬3 . (6.29)
R ⎬ ⎬
⎬r ⎬
Proof
Using the first Frenet formula we have:
dτ ν d2 r ν
= ⇒ 2 = ∃
ds R ds R
d2 r
ν = R 2. (6.30)
ds
We shall achieve:
⎬ 2 ⎬ ⎬ 2 ⎬
⎬d r ⎬ ⎬ ⎬
⎬ ⎬ = 1 · ν ⇒ ⎬ d r ⎬ = 1 ;
⎬ ds 2 ⎬ R ⎩ ⎬ ds ⎬
2 R
=1
therefore
2 2 2
1 d2 x d2 y d2 z
= + + .
R ds 2 ds 2 ds 2
Using the third formula from (6.24) and the relation (6.30) we deduce:
⎬ ⎬ ⎬ ⎬
⎬ ⎬ ⎬ dr d2 r ⎬
⎬
⎬β ⎬ = ⎬ × R 2 ⎬ =1∃
ds ds ⎬
⎬ ⎬
1 ⎬ dr d2 r ⎬
=⎬
⎬ ds × ⎬. (6.31)
R ds 2 ⎬
As
dr dr dt
= · (6.32)
ds dt ds
208 6 Differential Geometry of Curves and Surfaces
we have:
d2 r d dr d dr dt d dr dt dr d2 t
= = · = · + ·
ds 2 ds ds ds dt ds ds dt ds dt ds 2
d dr dt dt dr d2 t
= · · + · ,
dt dt ds ds dt ds 2
i.e.
2
d2 r d2 r dt dr d2 t
= · + · . (6.33)
ds 2 dt 2 ds dt ds 2
We deduce
⎪ 2 2 ⎛
dr d2 r dr dt d r dt dr d2 t
× 2 = · × · + ·
ds ds dt ds dt 2 ds dt ds 2
3
dt dr d2 r dt d2 t dr dr
= · × 2 + · · × .
ds dt dt ds ds 2 dt dt
⎩
=0
Therefore 3 d2 r
dr d2 r dt dr d2 r
dr
dt × dt 2
× 2 = · × 2 = ds 3 ,
ds ds ds dt dt
dt
i.e. d2 r
dr d2 r
dr
(6.20) dt × dt 2
× 2 = ⎬ ⎬3 . (6.34)
ds ds ⎬ dr ⎬
⎬ dt ⎬
Theorem 6.20 (computing the torsion, see [5], p. 593). Let π be a regular curve
and M → π be the current point, with position vector r , ds be the arc element on the
curve π and R be the radius of curvature of the the curve π in the point M. Then
1 r · r × r
= ⎬ ⎬ . (6.35)
T ⎬r × r ⎬2
6.4 Curvature and Torsion of the Space Curves 209
Proof
Taking into account the relations (6.32) and (6.33), we have
⎧ 2 ⎨ ⎧ 2 ⎨
d3r d d2 r dt dr d2 t d d2 r dt d dr d2 t
= · + · = · + ·
ds 3 ds dt 2 ds dt ds 2 ds dt 2 ds ds dt ds 2
2 ⎧ ⎨ 2
d d2 r dt d2 r d dt 2 d dr d t dr d3 t
= · + · + · + ·
ds dt 2 ds dt 2 ds ds ds dt ds 2 dt ds 3
2 2
d d r dt dt d2 r d dt dt
= · · + · ·
dt dt 2 ds ds dt 2 ds ds ds
2
d dr dt d t dr d3 t
+ · · 2+ ·
dt dt ds ds dt ds 3
3
d3r dt d2 r d2 t dt d2 r dt d2 t d2 r dt d2 t
= 3 · + 2 · 2· + 2 · · 2+ 2 · ·
dt ds dt ds ds dt ds ds dt ds ds 2
dr d3 t
+ · ,
dt ds 3
i.e. 3
d3r d3 r dt d2 r dt d2 t dr d3 t
= · +3 · · + · . (6.36)
ds 3 dt 3 ds dt 2 ds ds 2 dt ds 3
We obtain:
2
d2 r 3 3 d3 r d2 r
1
dr
ds · ds 2
× dds r3 dr
(6.31) ds
· dds r2 × dds r3 ds 3
· dr
ds × ds 2
= = ⎬ ⎬ = ⎬ ⎬2 . (6.39)
T 1 ⎬ dr 2 ⎬2 ⎬ dr d2 r ⎬
R2 ⎬ ds × dds r2 ⎬ ⎬ ds × ds 2
⎬
We have:
d2 r
dr d2 r
dr
(6.34) ds × ds 2
× 2 = ⎬ ⎬3 ;
ds ds ⎬ dr ⎬
⎬ dt ⎬
i.e
3
d3 r dr d2 r 1 d r dr
(6.34) d2 r
· × 2 = ⎬ ⎬3 · ×
ds 3 ds ds ⎬ dr ⎬ ds 3 ds ds 2
⎬ dt ⎬
⎪⎧ 3 ⎨ ⎛
1 d3 r dt d2 r d2 t dt dr d3 t dr d2 r
= ⎬ ⎬3 · +3 2 · 2 · + · · × 2
⎬ dr ⎬ dt 3 ds dt ds ds dt ds 3 dt dt
⎬ dt ⎬
⎜
3 3
1 ⎝ ⎝ dt d r dr d2 r d2 t dt d2 r dr d2 r
= ⎬ ⎬3 ⎝ · 3 · × 2 +3 2 · · · × 2
⎬ dr ⎬ ⎞ ds dt dt dt ds ds dt 2 dt dt
⎬ dt ⎬ ⎩
=0
⎟
⎠
d3 t dr dr d2 r ⎠
+ · · × 2 ⎠.
ds 3 dt dt dt ⎦
⎩
=0
Finally, we get
dt 3
· d3 r
· dr
× d2 r d3 r
· dr
× d2 r
d3 r dr d2 r ds dt 3 dt dt 2 dt 3 dt dt 2
· × 2 = ⎬ ⎬3 = ⎬ ds 3 ,
⎬3
ds 3 ds ds ⎬ dr ⎬ ⎬ dr ⎬
⎬ dt ⎬ ⎬ dt ⎬ · dt
i.e.
d3 r
·
2
× ddt 2rdr d3 r
· ×
dr d2 r
d3r dr d2 r dt 3 dt dt 3 dt dt 2
· × 2 = ⎬ ⎬ ⎬3 =⎬3 ⎬ ⎬6 . (6.40)
ds 3 ds ds ⎬ dr ⎬ ⎬ ⎬ ⎬ dr ⎬
⎬ dt ⎬ · ⎬ dr
dt ⎬ ⎬ dt ⎬
6.4 Curvature and Torsion of the Space Curves 211
Determine:
Solution
(a) Using (6.21), the tangent versor is:
−3 sin t i + 3 cos t j + 4k 1
τ= ⇐ = −3 sin t i + 3 cos t j + 4k .
9 + 16 5
r = 3 cos t i + 3 sin t j + 4t k.
We compute:
dτ dτ dt 1 dτ 1 dτ 1 dτ 1
= · = dt · =⎬ ⎬· = ⎬ ⎬ · = −3 cos t i − 3 sin t j .
ds dt ds dt ⎬ ⎬ dt ⎬r ⎬ dt 25
ds ⎬ dr
dt ⎬
ν = − cos t i − sin t j.
6.4 Curvature and Torsion of the Space Curves 213
The curvature of the curve can be also determined using the formula (6.29).
Solving (b) + (c) + (e) with Sage, we have:
(e) the binormal versor can be determined using one of the two formulas: (6.22)
or the third relation from (6.24). We achieve:
1
β= 4 sin t i − 4 cos t j + 3k .
5
(f) The binormal equations are determined using the formulas (6.16); it results:
x − 3 cos t y − 3 sin t z − 4t
β : = = .
4 sin t −4 cos t 3
(g) As
dβ dβ dt 1 dβ 1 dβ
= · = ds · =⎬ ⎬·
ds dt ds dt ⎬ dr ⎬ dt
dt ⎬ dt ⎬
1 dβ 1
= ⎬ ⎬ · = 4 cos t i + 4 sin t j ,
⎬r ⎬ dt 25
1 4
− − cos t i − sin t j = cos t i + sin t j ;
T 25
214 6 Differential Geometry of Curves and Surfaces
1 4
= .
T 25
The same value can be found using Sage:
(h) The equation of the osculator plane in M can be determined with (6.12):
x − 3 cos t y − 3 sin t z − 4t
π0 :
−3 sin t 3 cos t 4
= 0.
−3 cos t −3 sin t 0
F (x, y, λ) = 0, (6.41)
where λ is a real parameter represents a family of curves in the plane xOy, each
curve from the family being determined by the value of the respective parameter λ.
Definition 6.23 (see [3], p. 121). The envelope of a family of curves is the tangent
curve in every of its points, to a curve from that family (see Fig. 6.4).
As the considered family of curves depends on the parameter λ it results that as
well as the envelope points will depend on the values of λ. Therefore, the envelope
of a family of curves has the parametric representation:
x = x (λ)
I : (6.42)
y = y (λ) .
6.5 Envelope of a Family of Curves in Plane 215
The common points of the envelope and of the curves that belong to the respective
family check the equation:
i.e.
The equation of the tangent in a point M (x, y) to a curve from the family of
curves that has the Eq. (6.41) is
Fx
Y −y=− (X − x) , (6.46)
Fy
yλ
Y−y= (X − x) . (6.47)
xλ
As the envelope and the family of curves are tangent in the common points it
results that the slopes of tangents coincide, i.e. from the relations (6.46) and (6.47)
we deduce
yλ Fx
= − ⇒ Fx xλ + Fy yλ = 0. (6.48)
xλ Fy
Example 6.24 (see [4], p. 331). Determine the envelope of the family of straight
lines :
Fλ (x, y, λ) = x cos λ + y sin λ − a = 0,
Solution
We have
Fλ (x, y, λ) = −x sin λ + y cos λ.
Solving the system, which contains the Eqs. (6.41) and (6.49), i.e.:
Fλ (x, y, λ) = 0
Fλ (x, y, λ) = 0
it will result:
x cos λ + y sin λ − a = 0
(6.50)
−x sin λ + y cos λ = 0.
By multiplying the first equation of the system (6.50) with sin λ and the second
equation with cos λ we achieve
y = a sin λ.
By multiplying the first equation of the system (6.50) with cos λ and the second
equation with − sin λ we achieve
x = a cos λ.
We shall deduce
x 2 + y 2 = a 2 sin2 λ + a 2 cos2 λ = a 2 .
Hence, the elimination of the parameter λ from the system (6.50) gives us the
equation
x 2 + y2 = a2,
6.5 Envelope of a Family of Curves in Plane 217
i.e. the envelope is a circle centered in the origin and having the radius a.
Solving this problem in Sage, we obtain:
Definition 6.25 (see [5], p. 602). A regular portion of a surface is the set ϕ of
the points M (x, y, z) from the three-dimensional Euclidean real space R3 , whose
coordinates x, y, z check one of the following systems of equations:
x = f 1 (u, v)
y = f 2 (u, v) = 0, (u, v) → D ⊆ R2 (the parametric representation) , (6.53)
z = f 3 (u, v)
Theorem 6.30 (see [5], p. 609). Let ϕ be a regular surface. If (u, v) is a curvilinear
coordinate system on the surface ϕ, then any curve π ⇔ ϕ can be analytically
represented by one of the following equation:
u = u (t)
(6.57)
v = v (t)
f (u, v) = 0 (6.58)
u = g (v) . (6.59)
6.7 Tangent Plane and Normal to a Surface 219
Definition 6.31 (see [5], p. 615). The tangent plane in a point of the surface ϕ is
set of the tangents pursued to all of the curves from the surface, passing through that
point (see Fig. 6.5).
x − x0 y − y0 z − z 0
πT :
xu 0 yu 0 z u 0
= 0, (6.60)
xv yv0 z v0
where
⎭
⎭ x0 = x (u 0 , v0 ) , y0 = y (u 0 , v0 ) , z 0 = z (u 0 , v0 )
⎭
⎭
∂x ∂y ∂z
xu 0 = |u=u 0 ,v=v0 , yu 0 = |u=u 0 ,v=v0 , z u 0 = |u=u 0 ,v=v0
⎭ ∂u ∂u ∂u
⎭
⎭
xv = ∂x |u=u ,v=v , yu = ∂ y |u=u ,v=v , z v = ∂z |u=u ,v=v .
⎭
0
∂v 0 0 0
∂v 0 0 0
∂v 0 0
∂F ∂F
πT : (x − x0 ) |x=x0 ,y=y0 ,z=z 0 + (y − y0 ) |x=x0 ,y=y0 ,z=z 0
∂x ∂y
∂F
+ (z − z 0 ) |x=x0 ,y=y0 ,z=z 0 = 0 (6.61)
∂z
∂z ∂z
πT : (X − x0 ) |x=x0 ,y=y0 +(Y − y0 ) |x=x0 ,y=y0 −(Z − z 0 ) = 0. (6.62)
∂x ∂y
220 6 Differential Geometry of Curves and Surfaces
Find the equation of the tangent plane to the surface ϕ in the point M, for u 0 =
2, v0 = 1.
Solution
We have
xu = 1, yu = 1, z u = v
xv = 1, yv = −1, z v = u.
x − 3 y − 1 z − 2
πT :
1 1 1
= 0,
1 −1 2
i.e πT : 3 (x − 3) − (y − 1) − 2 (z − 2) = 0 ⇒
πT : 3x − y − 2z − 4 = 0.
ϕ : x 2 + 2x y + y 2 + 4x z + z 2 + 2x + 4y − 6z + 8 = 0.
Find the equation of the tangent plane to the surface in the point M (0, 0, 2).
Solution.
We have
F (x, y, z) = x 2 + 2x y + y 2 + 4x z + z 2 + 2x + 4y − 6z + 8.
6.7 Tangent Plane and Normal to a Surface 221
Whereas ∂F
⎭
⎭ = 2x + 2y + 4z + 2
∂x
∂F
⎭ ∂ y = 2x + 2y + 4
⎭
∂F
∂z = 4x + 2z − 6,
with (6.61), the equation of the tangent plane to the surface in the point M (0, 0, 2)
will be
πT : 10x + 4y − 2 (z − 2) = 0,
i.e
πT : 5x + 2y − z + 2 = 0.
Definition 6.34 (see [5], p. 621). A normal in a point of a surface is the straight line
perpendicular to the tangent plane to surface in that point (Fig. 6.6).
x − x0 y − y0 z − z0
λN :
=
=
; (6.63)
yu z u
z u xu
xu yu
0 0
0 0
0 0
yv z v
z v xv
xv yv
0 0 0 0 0 0
x − x0 y − y0 z − z0
λN : ∂F
= ∂F
= ∂F
; (6.64)
∂x |x=x0 ,y=y0 ,z=z 0 ∂ y |x=x0 ,y=y0 ,z=z 0 ∂z |x=x0 ,y=y0 ,z=z 0
X − x0 Y − y0 Z − z0
λN : ∂z
= ∂z
= . (6.65)
∂x |x=x0 ,y=y0 ∂ y |x=x0 ,y=y0
−1
Example 6.35 (see [5], p. 623). Find the equation of the normal to the surface from
the Example 6.32, in the point u 0 = 2, v0 = 1.
Solution
For u 0 = 2, v0 = 1 it results that x0 = 2, y0 = 1, z 0 = 2.
We shall have:
!
∂x ∂y ∂z
∂u = 1, ∂u = 1, ∂u = v
∂x ∂y ∂z
∂v = 1, ∂v = −1, ∂v = u;
therefore, based on (6.63), the equations of the normal to the surface will be:
x −3 y − y0 z − z0
λN :
=
=
1 1
1 1
1 1
−1 2
2 1
1 −1
or
x −3 y−1 z−2
λN : = = .
3 −1 −2
Example 6.36 (see [5], p. 623). Find the equation of the normal λ N to the surface
from the Example 6.33, in the point M0 (0, 0, 2) → ϕ.
Solution
As in the Example 6.33 we computed: ∂∂xF , ∂∂Fy , ∂∂zF , using the relation (6.64) we
achieve
x −3 y z−2 x y z−2
λN : = = ⇒ λN : = = .
10 4 −2 5 2 −1
Remark 6.38 (see [5], p. 637). The first fundamental form ρ1 is also called the
metric of the surface ϕ.
Theorem 6.39 (expression of the first fundamental form, see [5], p. 637). Let ϕ
be a regular surface, given through its parametric equations (6.54).
224 6 Differential Geometry of Curves and Surfaces
Then
where
E = xu2 + yu2 + z u2
F = xu xv + yu yv + z u z v (6.68)
G = xv2 + yv2 + z v2
and !
∂y
xu = ∂x ∂z
∂u , yu = ∂u , z u = ∂u
∂y (6.69)
xv = ∂x ∂z
∂u , yu = ∂u , z u = ∂u .
Definition 6.40 (see [5], p. 643). Let π ⇔ ϕ be the curve given by (6.54). If M1 , M2
are two points corresponding to the values of t = t1 and t = t2 then the length of
the arc M1 M2 is
"
" t 2 2
t2
2 du du dv dv
s =
ds
E + 2F · +G dt
, (6.70)
t1
t1 dt dt dt dt
Solution
The curvilinear coordinates of the points are M1 , M2 , M3 are
1 1
M1 u = a, v = 1 , M2 (u = 0, v = 0) , M3 u = − a, v = 1 .
2 2
We shall have
#
7
⎭
0 1 2
⎭
⎭ s 1 =
a v + 1 dv
= 6 a,
⎭
⎭
1 2
# 1
s2 =
0 a 21 v 2 + 1 dv
= 76 a
⎭
⎭
⎭
⎭
# 21 a
⎭
s3 =
1 du
= a.
−2a
7 7 20 10
P = s1 + s2 + s3 = a+ a+a = a= a.
6 6 6 3
This problem will be also solve in Sage:
6.8 First Fundamental Form of a Surface. Curves on a Surface 227
Theorem 6.43 (see [5], p. 645). Let the surface ϕ given by (6.54 ) and π1 ⇔ ϕ, π2 ⇔
ϕ be two curves traced on the surface ϕ. The angle between the curves π1 , π2 in
the point M → π1 ≺ π2 is given by the relation
where
• E, F, G are the coefficients from the metric of the surface ϕ computed in M,
• du, dv are two differentials along the curve π1 , and ∂u, ∂v are two differentials
along the curve π2 .
Remark 6.44 (see [3], p. 146). If π1 : u = C1 , π2 : v = C2 , C1 , C2 → R (the curves
have constant coordinates); as du = 0, ∂v = 0 it results that cos ϕ = ⇐ F .
EG
Example 6.45 (see [3], p. 147). Compute the angle between the curves π1 : u = v
and π2 : u + v = 2, traced on the surface
ϕ : r = u 2 + v 2 i + u 2 − v 2 j + 2uvk.
Solution
The parametric equations of the surface ϕ will be
x = u 2 + v2
y = u 2 − v2
z = 2uv.
228 6 Differential Geometry of Curves and Surfaces
Using (6.68), we shall compute the coefficients of the first fundamental form:
E = 4u 2 + 4u 2 + 4v 2 = 8u 2 + 4v 2
F = 2u · 2v − 2u · 2v + 2v · 2u = 4uv
G = 4v 2 + 4v 2 + 4u 2 = 4u 2 + 8v 2 .
Theorem 6.46 (see [5], p. 655). Let ϕ be a regular surface. The area of the surface
ϕ is calculated using the following surface integral
Definition 6.47 (see [5], p. 655). Let ϕ be a regular surface and σ from (6.72) be its
area. The surface area element corresponding to ϕ, denoted by dσ is the expression
&
dσ = E G − F 2 dudv. (6.73)
Theorem 6.48 (see [5], p. 655). Let ϕ be a regular surface. The area and the surface
area element have respectively the expressions from bellow; if
(a) the surface is given parametrically (as in the relation (6.54)), then
## ⇐
σ= ⇐ ϕ A√2 + B √2 + C √2 dudv
(6.74)
dσ = A + B √2 + C √2 dudv,
√2
where
√ =
yu z u
⎭
⎭ A
⎭
⎭
yv z v
⎭
⎭
z x
B √ =
u u
(6.75)
⎭
⎭ z v xv
⎭
⎭
⎭
⎭
x y
⎭
C √ =
u u
;
xv yv
(b) the surface is given implicitly (see the relation (6.54)), then
' 2 2 2
⎭
⎭ ## ∂F
+ ∂∂Fy + ∂∂zF
⎭
∂x
σ= ϕ ∂F dxdy
'
∂z
2 2
(6.76)
⎭
⎭
⎭
dσ = 1+ ∂z
+ ∂z
dxdy.
∂x ∂y
230 6 Differential Geometry of Curves and Surfaces
6.9 Problems
(a) Determine the analytical expressions of the versors of the Frenet trihedron
in an arbitrary point of the curve.
(b) Write the equations of the edges and of the planes of the Frenet trihedron in
an arbitrary point of the curve.
(c) Find the curve and the torsion of the curve in an arbitrary point.
2. Find the versors of the Frenet trihedron in the origin for the curve:
π : r = t i + t 2 j + t 3 k.
Solution
Solving in Sage, we get:
6.9 Problems 231
3. Write the tangent equations and the equation of the normal plane to the curve:
x (t) = et cos 3t
π : y (t) = et sin 3t
z (t) = e−2t
in the point t = 0.
4. Find the relation between the curvature and the torsion of the curve:
t2 t3
π:r =ti+ j + k.
2 6
Solution
We shall use Sage to determine the asked relation:
232 6 Differential Geometry of Curves and Surfaces
π : r = t −1 i + t j + 2t 2 − 1 k
Solution
We need the following Sage code to solve this problem:
Solution
The solution in sage of this problem is:
(a) Find the equation of the tangent plane to the surface in the point u = 0, v = 2;
(b) Find the equation of the normal to the surface in that point.
8. Let be the surface
ϕ : z = 5x 2 + 4y − 3.
(a) Find the equation of the tangent plane in the point M (1, 0, 2).
(b)Write the equation of the of normal to the surface ϕ in the point M (1, 0, 2).
Solution
The solution in Sage of this problem is:
ϕ : r = u cos v i + u sin v j + (u + v) k.
ϕ : r = u cos v i + u sin v j + av k.
References
1. G.h. Atanasiu, G.h. Munteanu, M. Postolache, Algebră liniară, geometrie analitică şi
diferenţială, ecuaţii diferenţiale, ed (ALL, Bucureşti) (1998)
2. I. Bârză, Elemente de geometrie diferenţială, ed (MatrixRom, Bucureşti) (2007)
3. T. Didenco, Geometrie analitică şi diferenţială (Academia Militară, Bucureşti) (1977)
4. C. Ionescu- Bujor, O. Sacter, Exerciţii şi probleme de geometrie analitică şi diferenţială, vol II,
ed (Didactică şi pedagogică, Bucureşti) (1963)
5. E. Murgulescu, S. Flexi, O. Kreindler, O. Sacter, M. Tîrnoveanu, Geometrie analitică şi
diferenţială, ed (Didactică şi pedagogică, Bucureşti) (1965)
6. V. Postelnicu, S. Coatu, Mică enciclopedie matematică, ed (Tehnică, Bucureşti) (1980)
Chapter 7
Conics and Quadrics
The analytical geometry replaces the definition and the geometrical study of the
curves and the surfaces with that algebraic: a curve respectively a surface is defined
by an algebraic equation and the study of the curves and of the surfaces is reduced
to the study of the equation corresponding to each of them.
Definition 7.1 (see [1], p. 158). We consider the function
A second order algebraic curve or a conic is the set π of the points M (x, y)
from the plane, whose coordinates relative to an orthonormal Cartesian reference
check thegeneral equation
f (x, y) = 0 (7.2)
where the coefficients a11 , a12 , a22 , b1 , b2 , c are some real constants, with a11
2 +
π = M (x, y) | (x, y) ∈ R2 , f (x, y) = 0 . (7.3)
Definition 7.2 (see [2]). The invariants of a conic are those expressions made with
the coefficients of the conic equation, that keep the same value to the changes of an
orthonormal reference.
Proposition 7.3 (see [2] and [3], p. 56). We can assign three invariants of the conics
from (7.2):
the first invariant is linear, the second is quadratic, while the third is a cubical one.
The invariant λ determines the nature of a conic. Thus, if:
• λ →= 0, we say that π is a non-degenerate conic (the circle, the ellipse, the hyper-
bola and the parabola)
• λ = 0, we say that π is a degenerate conic.
With the help of δ one establishes the type of a conic. Thus, if:
• δ > 0, we say that π has an elliptic type
• δ < 0, we say that π is of a hyperbolic type
• δ = 0, we say that π has a parabolic type.
Definition 7.4 (see [1], p. 168). The center of symmetry of a conic π (in the case
when this exists, the conic is called a conic with center) is a point C from the plane,
which has the property that for any point M ∈ π, the reflection of M with respect to
C satisfies the equation of the conic π, too.
Theorem 7.5 (see [1], p. 168). The conic π from (7.2) admits a unique center of sym-
metry C (x0 , y0 ) if and only if its invariant δ is non-null; in this case, its coordinates
are the solutions of the linear system
∂f
∂x =0 a11 x + a12 y + b1 = 0
∂f ⇔ (7.5)
∂y =0 a12 x + a22 y + b2 = 0.
Theorem 7.6 (see [4], p. 174). Each conic has one of the following canonical forms:
(1) imaginary ellipse:
x2 y2
2
+ 2 + 1 = 0; (7.7)
a b
(2) real ellipse:
x2 y2
2
+ 2 − 1 = 0; (7.8)
a b
(3) hyperbola:
x2 y2
2
− 2 − 1 = 0; (7.9)
a b
(4) two imaginary concurrent lines having a real intersection
x2 y2
2
+ 2 = 0; (7.10)
a b
(5) two real concurrent lines
x2 y2
2
− 2 = 0; (7.11)
a b
(6) parabola
y 2 = 2 px; (7.12)
x2
+ 1 = 0; (7.13)
a2
(8) two real parallel lines
x2
− 1 = 0; (7.14)
a2
238 7 Conics and Quadrics
x 2 = 0. (7.15)
7.2.1 Circle
Definition 7.7 (see [5], p. 117). The circle is the set of the equally spaced points
from the plane by a fixed point called the center, the distance from the center to the
points of the circle, being the radius.
We report the circle plane to an orthogonal Cartesian reference O, i, j . Let
C (a, b) be the center of the circle and M(x, y) be an arbitrary point of it (see
Fig. 7.1).
From the Definiton 7.7, it results that the distance from C and M is constant and
equal to the radius r of the circle, i.e.:
C M = r ;
hence
(x − a)2 + (y − b)2 = r.
If we open the squares from (7.16) we achieve the circle equation in the form:
Denoting by ⎫
⎬ m = −a
n = −b (7.18)
⎭
p = a 2 + b2 − r 2 ,
(x + m)2 + (y + n)2 = m 2 + n 2 − p
it results that:
1. if m 2 + n 2 ⎧
− p > 0 then the circle π will have the center C(−m, −n) and the
radius r = m 2 + n 2 − p;
2. if m 2 + n 2 − p = 0 then the circle π one reduces to the point C(−m, −n);
3. if m 2 + n 2 − p < 0 then π = ϕ (empty set).
For m 2 + n 2 − p > 0, the Eq. (7.19) is called the general Cartesian equation of
the circle π.
If θ is that angle made by the radius with the positive direction of the axis Ox,
then the parametric equations of the circle will be:
x = a + r cos θ
, θ ∈ [0, 2π] .
y = b + r sin θ
Example 7.8 (see [5], p. 121). Find the equation of the circle determined by the
points: M (−1, 1) , N (2, −1) , P (1, 3).
Solution
Using the Eq. (7.19) we deduce
⎫ ⎫
⎬ 2m − 2n − p = 2 ⎬ m = − 10
11
4m − 2n + p = −5 ⇒ n = − 10
9
⎭ ⎭
2m + 6n + p = −10; p=−5.12
22 18 12
x 2 + y2 − x− y− =0
10 10 5
240 7 Conics and Quadrics
or ⎨ ⎩
5 x 2 + y 2 − 11x − 9y − 12 = 0.
7.2.2 Ellipse
Let c > 0 be a positive real number and F, F ∀ be two fixed points of the plane such
that F F ∀ = 2c.
Definition 7.9 (see [5], p. 150). The ellipse is the set of the points M of the plane
that satisfy the relation
M F + M F ∀ = 2a = ct, (7.21)
Therefore F(c, 0) and F ∀ (−c, 0); the points F and F ∀ are called the foci of the
ellipse and the distance F F ∀ constitutes the focus distance of the ellipse. M F, M F ∀
are the focus radiuses (focus radii) of the point M. The ellipse admits a unique center
of symmetry O and two axes of symmetry Ox,Oy. The ellipse is a bounded curve
(there is a rectangle that contains all its points).
To find the ellipse equation we shall transform analytically the Eq. (7.21). If
M (x, y), then the relation (7.21) becomes
or
a (x + c)2 + y 2 = a 2 + cx;
therefore
⎨ ⎩ ⎨ ⎩
a 2 − c2 x 2 + a 2 y 2 − a 2 a 2 − c2 = 0. (7.23)
x 2 + y2 = r 2 (7.25)
242 7 Conics and Quadrics
and it represents a circle centered in origin and having the radius r . Therefore, the
circle is a special case of an ellipse.
Thereby:
2 y2
2 x
π = M (x, y) | (x, y) ∈ R , 2 + 2 = 1 . (7.26)
a b
To find the points of intersection of the curve with the coordinate axes we shall
make by turn y = 0 and x = 0. It results:
• A (a, 0), A∀ (−a, 0) are situated on Ox,
• B (0, b), B ∀ (0, −b) are situated on Oy.
The segment
• A A∀ = 2a is called the major axis of the ellipse;
• B B ∀ = 2b is called the minor axis of the ellipse.
Their halves, i.e. OA = a and OB = b are the semi-axes of the ellipse. The points
A, A∀ , B, B ∀ are called the vertices of the ellipse.
From the Eq. (7.8) one deduces the explicit Cartesian equations of the ellipse:
b⎧ 2
y=± a − x 2 , x ∈ [−a, a] . (7.27)
a
To obtain the parametric equations of the ellipse, one proceeds (see [6], p. 378)
as follows (see Fig. 7.3):
(1) build two concentric circles with the radii a and respectively b, a > b;
(2) trace a semi- line through origin, which intersects the two circles in the points
A and respectively B;
(3) bulid through the points A and B some lines, that are parallel with the axes; the
intersection of these points will be a point M of the ellipse;
(4) denote by θ the angle formed by the radius OA with the axis Ox.
7.2.3 Hyperbola
As in the case of the ellipse, we shall also consider c > 0 a real positive number and
F, F ∀ two fixed points from the plane such that F F ∀ = 2c.
Definition 7.10 (see [5], p. 156). The hyperbola is the set of the points M of the
plane that satisfy the relation
M F − M F ∀ = 2a = ct, (7.29)
i.e. that have constantly the difference of the distances to two fixed points.
−−∈
We choose F F ∀ as Ox axis and the mediator of the segment F F ∀ as Oy axis (see
Fig. 7.4).
M F, M F ∀ are the focus radii of the point M. The hyperbola admits a unique
center of symmetry O and two axes of symmetry Ox, Oy.
From Fig. 7.4 we can note that the hyperbola is an unbounded curve.
The points F (c, 0) and F ∀ (c, 0) are called the foci of the hyperbola and the
distance F F ∀ constitutes the focus distance of the hyperbola.
To find the hyperbola equation we shall transform analytically the Eq. (7.29). If
M (x, y), then the relation (7.29) becomes:
We shall obtain:
By reducing the similar terms and passing in the first part of all those terms that
don’t contain radical signs, we have:
cx − a 2 = ±a (x − c)2 + y 2 ;
hence ⎨ ⎩ ⎨ ⎩
c2 − a 2 x 2 − a 2 y 2 − a 2 c2 − a 2 = 0. (7.31)
c2 − a 2 = b2 . (7.32)
By dividing the relation (7.31) with a 2 b2 it results the implicit Cartesian equation
(7.9) of the hiperbola.
To find the points of intersection of the curve with the coordinate axes we shall
make by turn y = 0 and x = 0. It results that
That is why, the Ox axis is called the transverse axis and the Oy axis is called the
untransverse axis.
The points A, A∀ represent the vertices of the hyperbola.
From the Eq. (7.9) one deduces the explicit Cartesian equations of the hyperbola:
b⎧ 2
y=± x − a 2 , x ∈ (−∃, −a] ⇐ [a, ∃) . (7.33)
a
The hyperbola admits two oblique asymptotes:
b
y = ± x. (7.34)
a
From (7.9) we shall have:
⎨x y⎩⎨x y⎩
+ − = 1.
a b a b
7.2 Conics on the Canonical Equations 245
Example 7.11 (see [5], p. 158). Determine the vertices, the foci and the asymptotes
of the hyperbola
2x 2 − 5y 2 − 8 = 0.
Solution
Writing the equation of the hyperbola in the form
x2 y2
− 8 − 1 = 0,
4 5
we deduce: ⎫
⎬ a2 = 4
b2 = 85
⎭ 2
c = a 2 + b2 = 5 .
28
The hyperbola vertices are: A (2, 0), A∀ (−2, 0) and its foci: F 2 5 , 0 ,
7
∀
F −2 5 , 0 . The equations of the hyperbola asymptotes are:
7
2
y=± x.
5
246 7 Conics and Quadrics
7.2.4 Parabola
Definition 7.12 (see [5], p. 182). The parabola is the set of the points from plane,
that are equidistant from a fixed line and a fixed point.
The fixed line is called the directrix of the parabola and the fixed point is called
the focus of the parabola.
To find the equation of the parabola we choose a Cartesian reference whose axes
are:
• the perpendicular from the focus F on the directrix d as Ox axis,
• the parallel to d going to the midway between the focus and the directrix d as Oy
axis.
We denote A = d⇒Ox. Let M be a point of the parabola and N be its projection
on the directrix (see Fig. 7.5).
7.2 Conics on the Canonical Equations 247
The parabola has no center of symmetry and it has a single axis of symmetry, Ox.
It is an unbounded curve.
We denote AF = p; it results F 2p , 0 and A − 2p , 0 . If M is an arbitrary
point of the parabola, then within the Definition 7.12, the point M has to satisfy the
relation:
M F = M N. (7.35)
As ⎨ p⎩ p
MN = x − − =x+ ,
2 2
the relation (7.35) becomes:
⎨
p ⎩2 p
x− + y2 = x + ;
2 2
y 2 = 2 px. (7.36)
Remark 7.13 (see [1], p. 165). In the case when x ⇔ 0, the implicit Cartesian
equation of the parabola will become:
y 2 = −2 px.
The Ox axis cuts the parabola in the point O(0, 0) called the parabola vertex.
From the Eq. (7.36) one deduces the explicit Cartesian equations of the parabola:
⎧
y = ± 2 px, x ≺ 0, (7.37)
p being a positive number called the parabola parameter, which indicates its form.
As p is smaller, both the focus and the directrix one approach [6] by the Oy axis,
and the parabola one approaches by the Ox axis (when p ∈ 0 then the parabola
degenerates into the Ox axis).
248 7 Conics and Quadrics
As p is greater, both the focus and the directrix one depart [6] by the Oy axis,
and the parabola one approaches by the Oy axis (when p ∈ ∃ then the parabola
degenerates into the Oy axis).
The parametric equations of the parabola are:
2
x = 2t p
, t ∈ R. (7.38)
y=t
Definition 7.14 (see [5], p. 187). The eccentricity denoted by e is an element which
characterizes the circle, the ellipse, the hiperbola and the parabola, representing the
ratio of the distances from an arbitrary point of the respective conic to the focus and
the directrix, namely:
• e = 0 characterizes the circle e = ac = a0 ,
• 0 < e < 1 characterizes the ellipse e = ac ,
• e > 1 characterizes the hyperbola e = ac ,
• e = 1 characterizes the parabola e = M MF
N .
associated to the form Q relative to the canonical basis B of R2 and then build the
characteristic polynomial:
a − λ a12
P (λ) = 11 = λ2 − I λ + δ. (7.40)
a12 a22 − λ
7.3 Reducing to the Canonical Form of a Conic Equation 249
(3) make a change of an orthonormal reference such that the center of symmetry to
be the origin of the new reference.
The transition from the coordinates (x, y) to the coordinates x ∀ , y ∀ in the
new reference is achieved by a translation of the vector OC, characterized by the
equations:
x = x0 + x ∀
(7.41)
y = y0 + y ∀ .
i.e.
a11 x ∀2 + 2a12 x ∀ y ∀ + a22
2 ∀2
y + c∀ = 0, c∀ = f (x0 , y0 ) . (7.42)
As δ →= 0 it results that
λ
= c∀ .
δ
(6) obtain the canonical equation of the conic:
λ
λ1 x ∀∀2 + λ2 y ∀∀2 + = 0. (7.45)
δ
250 7 Conics and Quadrics
Remark 7.15 (see [2]). If λ1 and λ2 have the same sign and λδ has an opposite one,
then from (7.45) one obtains an ellipse. If λ1 and λ2 have different signs then from
(7.45) one obtains an hyperbola.
Example 7.16 (see [4], p. 209). Let be the conic
π : 9x 2 − 4x y + 6y 2 + 16x − 8y − 2 = 0.
Bring it to the canonical form, pointing the necessary reference changes, recognize
the obtained conic and plot its graph.
Solution
Using (7.4), we have:
I = 9 + 6 = 15,
9 −2
δ= = 50 > 0
−2 6
9 −2 8
λ = −2 6 −4 = −500 →= 0.
8 −4 −2
As
• δ > 0, the conic π has an elliptic type
• λ →= 0, the conic π is non-degenerate
• δ →= 0, the conic π admits an unique center of symmetry.
The center of the conic is given by the system:
18x − 4y + 16 = 0
−4x + 12y − 8 = 0,
i.e. it is the point C − 45 , 25 .
We shall make a change of an orthonormal reference so that the center of symmetry
to be the origin of the new reference:
x∀ = x + 4
x = − 45 + x ∀
5 ⇔
y∀ = y − 2
5 y = 25 + y ∀ .
or
π : 9x ∀2 − 4x ∀ y ∀ + 6y ∀2 − 10 = 0. (7.46)
We note that
(7.43) 4 2
c∀ = f − , ,
5 5
where
f (x, y) = 9x 2 − 4x y + 6y 2 + 16x − 8y − 2.
(7.40)
P (λ) = λ2 − 15λ + 50 = (λ − 10) (λ − 5)
We achieve:
9 −2 v1 v1
=5 ;
−2 6 v2 v2
therefore
9v1 − 2v2 = 5v1
⇒ 2v1 = v2 ;
−2v1 + 6v2 = 5v2
it results
⎫ ⎝
⎬ ⎞
Vλ1 = v ∈ R2 | v = v1 (1, 2), (√) v1 ∈ R .
⎭ ⎪⎛ ⎜ ⎟
w
We shall obtain the orthonormal basis B1 = f 1 , where:
w 1 2
f1 = = ≤ ,≤ .
≡w≡ 5 5
252 7 Conics and Quadrics
Vλ2 = u ∈ R2 |Au = λ2 u .
We achieve:
9 −2 u1 u1
= 10 ;
−2 6 u2 u2
therefore
9u 1 − 2u 2 = 10u 1
⇒ u 1 = −2u 2 ;
−2u 1 + 6u 2 = 10u 2
it results
⎫ ⎝
⎬
⎞
Vλ2 = u ∈ R |u = u 2 (−2, 1), (√) u 2 ∈ R .
2
⎭ ⎪⎛ ⎜
⎟
z
We shall obtain the orthonormal basis B2 = f 2 , where:
z 2 1
f2 = = −≤ , ≤ .
≡z≡ 5 5
It will result
B = B1 ⇐ B2 = f 1 , f 2 .
therefore
x∀ = ≤1 x ∀∀ + ≤2 y ∀∀
5 5 (7.47)
y∀ = − ≤2 x ∀∀ + ≤1 y ∀∀ .
5 5
x ∀∀2
π: + y ∀∀2 − 1 = 0
2
and it represents an ellipse.
The ellipse axes have the equations x ∀∀ = 0 and respectively y ∀∀ = 0.
Solving the system from (7.47) we deduce
⎫
⎬ x ∀∀ = x ∀≤
−2y ∀
5 (7.48)
∀ +y ∀
⎭ y ∀∀ = 2x≤
.
5
x + 45 −2 y − 25 = 0
2 x + 45 + y − 25 = 0,
i.e.
x − 2y + 85 = 0
(7.50)
2x + y + 65 = 0.
The ellipse center will be O∀∀ x ∀∀ = 0, y ∀∀ = 0 . We have:
⎫ ∀ ∀
⎬x≤−2y
x ∀∀
= 0 (7.48) = 0 (7.49)
∀∀ ⇔ 5
∀ +y ∀ ⇔
y =0 ⎭ ≤ =0
2x
5
∀
x − 2y ∀ = 0 (7.50) x − 2y + 85 = 0
⇔
2x ∀ + y ∀ = 0 2x + y + 65 = 0
x = − 45
⇔
y = 25 ;
hence the ellipse center will be O∀∀ x = − 45 , y = 25 .
The vertices of the ellipse will be:
254 7 Conics and Quadrics
⎨ ≤ ⎩ ⎨ ≤ ⎩
A x ∀∀ = 2, y ∀∀ = 0 , A∀ x ∀∀ = − 2, y ∀∀ = 0 ,
B x ∀∀ = 0, y ∀∀ = 1 , B ∀ x ∀∀ = 0, y ∀∀ = −1 .
We deduce
⎫ ∀ ∀ ≤
≤ ⎬x≤−2y
x ∀∀ = 2 (7.48) = 2 (7.49)
⇔ ∀
5
∀ ⇔
y ∀∀ = 0 ⎭ 2x≤+y = 0
5
∀ ≤ ≤
x − 2y ∀ = 10 (7.50) x − 2y + 85 = 10
⇔
2x ∀ + y ∀ = 0 2x + y + 65 = 0
⎫ ≤
⎬ x = 10−4
⎨ 5 ⎩ ≤
⇔
⎭ 2 1− 10
y= 5 ;
⎫ ≤
≤ ⎬ x = − 10−4
x ∀∀ =− 2 ⎨ ≤ 5 ⎩
⇔
y ∀∀ = 0 ⎭ 2 1+ 10
y= 5 ;
⎫ ∀ ∀
⎬x≤−2y
x ∀∀
= 0 (7.48) = 0 (7.49)
∀∀ ⇔ 5
∀ +y ∀ ⇔
y =1 ⎭ ≤ =1
2x
5
∀
x − 2y ∀ =≤0 (7.50) x − 2y + 85 =≤0
⇔
2x ∀ + y ∀ = 5 2x + y + 65 = 5
≤
x=2 ≤ 5−4
⇔ 5
y= 5 ;2+ 5
≤
x ∀∀ = 0 x = −2 5≤5−4
⇔
y ∀∀ = −1 y = 2−5 5 .
∀
x x
= M(B,B ∀ ) · . (7.52)
y y∀
λ 0
δ = 1 = λ1 λ2 .
0 λ2
We note that:
λ1 0 0
λ = 0 0 b2∀ = −λ1 b2∀2 .
0 b∀ c∀∀
2
λ1 X 2 + 2b2∀ Y = 0, (7.54)
π : 4x 2 − 4x y + y 2 − 3x + 4y − 7 = 0.
Bring it to the canonical form, pointing the necessary reference changes, recognize
the obtained conic and plot its graph.
Solution
Using (7.4), we have:
I = 4 + 1 = 5,
4 −2
δ= =0
−2 1
4 −2 −3/2
25
λ = −2 1 2 = − →= 0.
−3/2 2 −7 4
As
• λ →= 0, the conic π is non-degenerate
• δ = 0, the conic π doesn’t admit an unique center of symmetry.
The matrix associated to the quadratic form is
4 −2
A= .
−2 1
(7.40)
P (λ) = λ2 − 5 = λ (λ − 5)
We achieve:
4 −2 u1 u1
=0· ;
−2 1 u2 u2
therefore
4u 1 − 2u 2 = 0
⇒ u 2 = 2u 1 ;
−2u 1 + u 2 = 0
it results ⎫ ⎝
⎬ ⎞
Vλ1 = v ∈ R2 |u = u 1 (1, 2), (√) u 1 ∈ R .
⎭ ⎪⎛ ⎜ ⎟
w
We shall obtain the orthonormal basis B1 = f 1 , where:
w 1 2
f1 = = ≤ ,≤ .
≡w≡ 5 5
We achieve:
4 −2 v1 v
=5 1 ;
−2 1 v2 v2
therefore
4v1 − 2v2 = 5v1
⇒ v1 = −2v2 ;
−2v1 + v2 = 5v2
it results
⎫ ⎝
⎬
⎞
Vλ2 = v ∈ R |v = v2 (−2, 1), (√) u 2 ∈ R .
2
⎭ ⎪⎛ ⎜
⎟
z
We shall obtain the orthonormal basis B2 = f 2 , where:
z 2 1
f2 = = −≤ , ≤ .
≡z≡ 5 5
It will result
B = B1 ⇐ B2 = f 1 , f 2 .
260 7 Conics and Quadrics
The transition to the new coordinates (x ∀ , y ∀ ) one achieves through the relation
(7.52), where ⎠ 1 ⎦
2 ≤ ≤
M(B,B ∀ ) = 5 5 ;
− ≤ ≤1
2
5 5
therefore
x= ≤1 x ∀ + ≤2 y ∀
5 5
y= − ≤2 x ∀ + ≤1 y ∀ .
5 5
We will form a perfect square, writing the Eq. (7.55) in the form
2
1∀ 11 176
5 y − ≤ − ≤ x∀ − = 0.
5 5 5 25
We can write:
⎠ ≤ ⎦
∀∀2 11 ∀∀ 176 5
5y −≤ x − · = 0.
5 25 11
11
π : Y2 = ≤ X
5
The obtained
parabola can be represented not only in the (Ox y) plane, but in the
Ox ∀∀ y ∀∀ plane, too:
An algebraic surface of the second order or a quadric is the set ρ of the points
M (x, y, z) from the space, whose coordinates relative to an orthonormal Cartesian
reference verify general the equation
f (x, y, z) = 0, (7.57)
7.4 General Equation of a Quadric 263
where the coefficients a11 , a22 , a33 , a12 , a13 , a23 , b1 , b2 , b3 , c are some real con-
%
3 %3
stants, with ai2j →= 0; therefore
i=1 j=1
ρ = M (x, y, z) | (x, y, z) ∈ R3 , f (x, y, z) = 0 . (7.58)
Definition 7.19 (see [2]). The invariants of a quadric are those expressions formed
with the coefficients of the conic equation, that keep the same value to the changes
of an orthonormal reference.
Proposition 7.20 (see [3], p. 69). We can assign four invariants of the quadric from
(7.58):
⎫∂f ⎫
⎬ ∂x = 0 ⎬ a11 x + a12 y + a13 z + b1 = 0
∂f
∂ = 0 ⇔ a12 x + a22 y + a23 z + b2 = 0 (7.60)
⎭∂f
y ⎭
a13 x + a23 y + a33 z + b3 = 0.
∂z = 0
Theorem 7.21 (see [4], p. 176). Any quadric has one of the following canonical
forms:
(1) imaginary ellipsoid:
x2 y2 z2
+ + + 1 = 0, a ≺ b ≺ c > 0 (7.61)
a2 b2 c2
264 7 Conics and Quadrics
x2 y2 z2
2
+ 2 − 2 − 1 = 0, a ≺ b > 0, c > 0 (7.63)
a b c
(4) hyperboloid of two sheets
x2 y2 z2
+ − + 1 = 0, a ≺ b > 0, c > 0 (7.64)
a2 b2 c2
(5) imaginary second-order cone
x2 y2 z2
+ + = 0, a ≺ b > 0, c > 0 (7.65)
a2 b2 c2
(6) real second-order cone
x2 y2 z2
2
+ 2 − 2 = 0, a ≺ b > 0, c > 0 (7.66)
a b c
(7) elliptic paraboloid
x2 y2
+ − 2z = 0, a ≺ b > 0 (7.67)
a2 b2
(8) hyperbolic paraboloid
x2 y2
− − 2z = 0, a > 0, b > 0 (7.68)
a2 b2
(9) imaginary elliptic cylinder
x2 y2
+ + 1 = 0, a ≺ b > 0 (7.69)
a2 b2
x2 y2
+ − 1 = 0, a ≺ b > 0 (7.70)
a2 b2
(11) hyperbolic cylinder
x2 y2
2
− 2 − 1 = 0, a > 0, b > 0 (7.71)
a b
7.4 General Equation of a Quadric 265
x2 y2
+ = 0, a, b > 0 (7.72)
a2 b2
(13) pair of secant planes
x2 y2
2
− 2 = 0, a, b > 0 (7.73)
a b
(14) parabolic cylinder
x2
− 2z = 0, a > 0 (7.74)
a2
(15) pair of parallel imaginary planes
x2
+ 1 = 0, a > 0 (7.75)
a2
(16) pair of parallel real planes
x2
− 1 = 0, a > 0 (7.76)
a2
(17) pair of confounded planes
x 2 = 0. (7.77)
λ
S1 x 2 + S2 y 2 + S3 z 2 + = 0, (7.78)
δ
where S1 , S2 , S3 are the roots of the secular equation
S 3 − I S 2 + J S − δ = 0. (7.79)
S1 x 2 + S2 y 2 + S3 z 2 = 0. (7.81)
Example 7.22 (see [8]). Determine the nature of the following quadrics:
(a) x 2 + y 2 + z 2 + 7x y + yz − 6z = 0
(b) 36x 2 + y 2 + 4z 2 + 72x + 6y − 40z + 109 = 0.
Solution
(a) We have
f (x, y, z) = x 2 + y 2 + 7x y + yz − 6z.
⎫
⎬ a11 = 1, a22 = 1, a33 = 1, a12 = 27 , a13 = 0, a23 = 1
2
b1 = 0, b2 = 0, b3 = −3
⎭
c = 0.
We shall obtain:
I = 1 + 1 + 1= 2
1 7/2 1 0 1 1/2
J = + + = − 19
7/2 1 0 1 1/2 1 2
1 7/2 0
23
δ = 7/2 1 1/2 = − →= 0
0 1/2 1 2
7.4 General Equation of a Quadric 267
1 7/2 0 0
7/2 1 1/2 0 405
λ = = →= 0.
0 1/2 1 −3 4
0 0 −3 0
(b) We have
I = 36 + 1 + 4 = 41
36 0 36 0 1 0
J = + + = 184
0 1 0 4 0 4
36 0 0
δ = 0 1 1/2 = 144 →= 0
0 0 4
36 0 0 36
0 1 1/2 3
λ= = −5184 →= 0.
0 0 4 −20
36 3 −20 109
x 2 + 4y 2 + 36z 2 − 36 = 0
or
x2 y2
+ + z2 − 1 = 0
36 9
i.e. it is a real ellipsoid.
Using Sage, we shall have:
270 7 Conics and Quadrics
7.5.1 Sphere
Definition 7.23 (see [1], p. 185). The sphere is the set of the points from space
equally distant from a fixed point called the center of the sphere, the distance from
the center to the points of the sphere is called the radius of the sphere.
We shall report the sphere plane at an orthonormal Cartesian reference.
Let C (a, b, c) be center of the sphere and M (x, y, z) an arbitrary point of the
sphere (see Fig. 7.6).
Remark 7.24 (see [2]). The sphere is a quadric of rotation which is obtained by
rotating a circle (semicircle) around one of its diameter.
From the Definition 7.23 it results that the distance between C and M is constant
and equal to the radius R of the sphere:
C M = R,
i.e.
If we open the squares in (7.82) we get the general Cartesian equation of the
sphere
The relations between the Cartesian coordinates (x, y, z) of a point M from space
and its spherical coordinates (ρ, θ, ϕ) are:
272 7 Conics and Quadrics
⎫
⎬ x = ρ sin θ cos ϕ
y = ρ sin θ sin ϕ (7.84)
⎭
z = ρ cos θ,
where:
• ρ ≺ 0 is the distance from the point M to the origin of axes,
• θ, θ ∈ [0, π] is angle made by the position vector of the point M with the Oz axis,
• ϕ, ϕ ∈ [0, 2π] means the angle made by the projection of the position vector of
the point M on the plane (xOy) with the Ox axis.
Remark 7.25 (see [6], p. 667). Each triplet of spherical coordinate corresponds to a
point, but not any point corresponds to a triplet, as is the case when M is on Oz or
in origin.
If we make a change of an orthonormal reference, such that C (a, b, c) constitutes
the origin of thenew reference,
then the transition from the coordinates (x, y, z) to
the coordinates x ∀ , y ∀ , z ∀ in the new reference is achieved by a translation of the
vector OC, characterized by the equations:
⎫
⎬ x = a + x∀
y = b + y∀
⎭
z = c + z∀.
The parametric equations of the sphere with the center C (a, b, c) and the radius
ρ ≺ 0 will be:
⎫
⎬ x = a + ρ sin θ cos ϕ
y = b + ρ sin θ sin ϕ , θ ∈ [0, π] , ϕ ∈ [0, 2π] . (7.85)
⎭
z = c + ρ cos θ.
Example 7.26 (see [9], p. 104). Write the equation of the sphere with the center on
the line:
x y−1 z+2
d: = = ,
1 −1 1
≤
having the radius R = 2 and passing through the point A (0, 2, −1).
Solution
We have
x y−1 z+2
= = = t, t ∈ R.
1 −1 1
As the center of the sphere is situated on the straight line d, it results that the point
C (t, 1 − t, t − 2) is the center of the sphere.
From the condition
2
C A = R 2
we deduce
3t 2 = 0 ⇒ t = 0.
It follows that C (0, −1, 2) is the the center of the sphere and the equation of the
sphere will be
S : x 2 + (y − 1)2 + (z + 2)2 = 2.
7.5.2 Ellipsoid
Definition 7.27 (see [7], p. 351 and [10], p. 131). The ellipsoid is the set of the
points in space, whose coordinates relative to an orthonormal reference check the
Eq. (7.62), where a, b, c are some strictly positive real numbers called the semi-axes
of the ellipsoid.
We consider the orthonormal reference, relative to which is given the ellipsoid
equation.
To plot the ellipsoid we shall determine its intersections with: the coordinate axes,
the coordinate planes, the planes parallel to the coordinate planes.
y=0 x2
Ox : ⇒ 2 − 1 = 0 ⇒ x = ±a;
z=0 a
hence the ellipsoid crosses Ox in two points: A (a, 0, 0) and A∀ (−a, 0, 0).
x =0 y2
Oy : ⇒ 2 − 1 = 0 ⇒ y = ±b;
z=0 b
hence the ellipsoid crosses Oy in two points: B (0, b, 0) and B ∀ (0, −b, 0).
x =0 z2
Oz : ⇒ 2 − 1 = 0 ⇒ z = ±c;
y=0 c
hence the ellipsoid crosses Oz in two points: C (0, 0, c) and C ∀ (0, 0, −c).
The points A, A∀ , B, B ∀ , C, C ∀ are the vertices of the ellipsoid, while the symmetry
axes of the ellipsoid are Ox,Oy,Oz.
2 2
(Ox y) : z = 0 ⇒ ax 2 + by2 − 1 = 0 ⇒ an ellipse of semi-axes a and b.
x2 z2
(Ox z) : y = 0 ⇒ a2
+ c2
− 1 = 0 ⇒ an ellipse of semi-axes a and c.
y2 z2
(Oyz) : x = 0 ⇒ b2
+ c2
− 1 = 0 ⇒ an ellipse of semi-axes b and c.
The intersections with the planes parallel to the plane (Ox y), by equation z = k,
can be determined from:
x2 y2 k2
+ + − 1 = 0.
a2 b2 c2
If k ∈ (−c, c) then the intersections with the planes parallel to the plane (Ox y)
are some ellipses having the equations (Fig. 7.8):
7.5 Quadrics on Canonical Equations 275
x2 y2
⎨ ≤ ⎩2 + ⎨ ≤ ⎩2 − 1 = 0.
a
c 2 − k2 b
c 2 − k2
c c
The ellipsoid has: an unique center of symmetry (the origin), symmetry axes (the
coordinate axes), planes of symmetry (the coordinate planes).
The parametric equations corresponding to an ellipsoid are:
⎫
⎬ x = a sin u cos v
y = b sin u sin v , u ∈ [0, π] , v ∈ [0, 2π] . (7.86)
⎭
z = c cos u
Remark 7.28 (see [10], p. 132). The sphere is a special case of ellipsoid, obtained if
all the semi-axes of the ellipsoid are equal between themselves. If two semi-axes are
equal, then one achieves a rotating ellipsoid, which can be generated by the rotation
of an ellipse around of an axis. For example, if a = b, then the ellipsoid is of rotation
around of Oz.
The next figure shows an ellipsoid built in Sage, using (7.86):
276 7 Conics and Quadrics
7.5.3 Cone
Definition 7.29 (see [10], p. 153). The cone of the second order is the set of the
points in space, whose coordinates relative to an orthonormal reference check the
Eq. (7.66).
We consider the orthonormal reference, relative to which is given the cone equa-
tion. To plot the cone we shall determine its intersections with: the coordinate axes,
the coordinate planes, the planes parallel to the coordinate planes.
y=0 x2
Ox : ⇒ 2 =0⇒x =0
z=0 a
hence the cone crosses Ox in origin. Similarly, the cone crosses Oy and O z in origin,
too. 2
2
(Ox y) : z = 0 ⇒ ax 2 + by2 = 0 ⇒ x = y = 0.
x2 z2
(Ox z) : y = 0 ⇒ a2
− c2
=0⇔
x
− z
=0
a c ⇒ two lines concurrent in the origin.
x
a + z
c =0
y2 z2
(Oyz) : x = 0 ⇒ b2
− c2
=0⇔
y
− z
=0
by c ⇒ two lines concurrent in the origin.
b + z
c =0
7.5 Quadrics on Canonical Equations 277
The intersections with the planes parallel to the plane (Ox y), by equation z = k,
can be determined from:
x2 y2 k2
+ − = 0;
a2 b2 c2
therefore the intersections with the planes parallel to the plane (Ox y) are some
ellipses having the equations (Fig. 7.9):
x2 y2
a 2 + b 2 − 1 = 0.
ck ck
y2 z2
− =0 (7.87)
a2 c2
around the Oz axis.
278 7 Conics and Quadrics
The parametric equations corresponding to the cone of the second order are:
⎫
⎬ x = av cos u
y = bv sin u , u ∈ [0, 2π] , v ∈ R. (7.88)
⎭
z = ±c
We shall plot in Sage a cone of the second order, defined by the Eq. (7.66):
Definition 7.32 (see [10], p. 141). The one-sheeted hyperboloid is the set of the
points from space, whose coordinates relative to an orthonormal reference check the
Eq. (7.63).
We consider the orthonormal reference, relative to which is given the one-sheeted
hyperboloid equation.
y=0 x2
Ox : ⇒ 2 − 1 = 0 ⇒ x = ±a;
z=0 a
x2 y2
⎨ ≤ ⎩2 + ⎨ ≤ ⎩2 − 1 = 0
a
c 2 + k2 b
c 2 + k2
c c
y2 z2
− −1=0 (7.90)
b2 c2
around of the Oz axis.
We shall use parametric equation (7.89) to represent the one-sheeted hyperboloid
in Sage:
Definition 7.33 (see [10], p. 137). The two-sheeted hyperboloid is the set of the
points from space, whose coordinates relative to an orthonormal reference check the
Eq. (7.64).
The number of the sheets is given by the number of the squares that have the same
sign with the free term.
We consider the orthonormal reference, relative to which is given the two-sheeted
hyperboloid equation.
y=0 x2
Ox : ⇒ 2 = −1;
z=0 a
hence the one-sheeted hyperboloid crosses in two points: C (0, 0, c) and C ∀ (0, 0, −c).
2 2 2 2
(Oyz) : x = 0 ⇒ by2 − cz 2 + 1 = 0 ⇒ π1 : cz 2 − by2 = 1 a hyperbola.
2 2 2 2
(Ox z) : y = 0 ⇒ ax 2 − cz 2 + 1 = 0 ⇒ π2 : cz 2 − ax 2 = 1 a hyperbola.
The intersections with the planes parallel to the plane (Ox y), by equation z = k,
can be determined from:
x2 y2 k2
+ = − 1.
a2 b2 c2
If k ∈ (−∃, −c) ⇐ (c, ∃) then the intersections with the planes parallel to the
plane (Ox y) are some ellipses having the equations:
x2 y2
⎨ ≤ ⎩2 + ⎨ ≤ ⎩2 − 1 = 0.
c c −k c c −k
a 2 2 b 2 2
y2 z2
− +1=0 (7.92)
b2 c2
around of the Oz axis.
The following Sage code allows us to plot a two-sheeted hyperboloid:
Definition 7.35 (see [10], p. 150). The elliptic paraboloid is the set of the points in
space, whose coordinates relative to an orthonormal reference check the Eq. (7.67),
where a, b, z are some strictly positive real numbers.
We consider the orthonormal reference, relative to which is given the elliptic
paraboloid.
y=0 x2
Ox : ⇒ 2 = 0;
z=0 a
x2 y2
πλ : ⎨ ≤ ⎩2 + ⎨ ≤ ⎩2 = 1.
a 2k b 2k
⎫ ≤
⎬ x = a ≤2v cos u
y = b 2v sin u , u ∈ [0, 2π] , v > 0. (7.93)
⎭
z=v
Remark 7.36 (see [10], p. 128) If a = b then elliptic paraboloid is of rotation around
of Oz, i.e. it can be generated by the rotation of the parabola
y 2 = 2a 2 z (7.94)
Definition 7.37 (see [10], p. 146). The hyperbolic paraboloid is the set of the
points in space, whose coordinates relative to an orthonormal reference satisfy the
Eq. (7.68).
We consider the orthonormal reference, relative to which is given the hyperbolic
paraboloid.
y=0 x2
Ox : ⇒ 2 = 0;
z=0 a
x2 y2
πλ : ⎨ ≤ ⎩2 − ⎨ ≤ ⎩2 = 1.
a 2k b 2k
Remark 7.38 (see [6], p. 686). There is not a hyperbolic paraboloid of rotation; the
hyperbolic paraboloid is the only surface of second degree, which is not a surface of
rotation (because any section through a hyperbolic parabolid is not an ellipse).
The hyperbolic paraboloid is a surface of translation, this being obtained by the
translation of a parabola (which has the opening in the bottom)
y 2 = −2b2 z (7.96)
x 2 = 2a 2 z. (7.97)
Definition 7.39 (see [1], p. 196). A surface that can be generated by moving a straight
line G, which one relieses on a given curve π from space is called a ruled surface.
The straight line G is called the generatrix (or generator) of the surface.
Definition 7.40 (see [2]). A family of straight lines, which has the property that
every straight line of the family can generate a surface represents the rectilinear
generators of the respective surface.
7.6 Ruled Surfaces. Surface Generation 287
Proposition 7.41 (see [4], p. 174). The straight line d passing through the point
A (x0 , y0 , z 0 ) and which has the director vector v = (l, m, n) constitutes the recti-
linear generators of the quadric from (7.58) if and only if the following conditions
are satisfied:
⎫
⎬ f (x0 , y0 , z 0 ) = 0
a11 l 2 + a22 m 2 + a33 n 2 + 2a12 lm + 2a13ln + 2a23 mn = 0 (7.98)
⎭ ∂f
l ∂x (x0 , y0 , z 0 ) + m ∂∂ yf (x0 , y0 , z 0 ) + n ∂∂zf (x0 , y0 , z 0 ) = 0.
The only ones ruler surfaces are the two non-degenerate quadrics: the one-sheeted
hyperboloid and the hyperbolic paraboloid.
Example 7.42 (see [4], p. 239). Determine the rectilinear generators of the quadric
by equation:
ρ : x 2 + y 2 + z 2 + 2x y − 2x z − yz + 4x + 3y − 5z + 4 = 0
f (x, y, z) = x 2 + y 2 + z 2 + 2x y − 2x z − yz + 4x + 3y − 5z + 4.
As
⎫∂f ⎫∂f
⎬ ∂x (x, y, z) = 2x + 2y − 2z + 4
⎬ ∂x (−1, −1, 1) = −2
∂f ∂f
∂y (x, y, z) = 2y + 2x − z + 3 ⇒ ∂ y (−1, −1, 1) = −2
⎭ ∂f
⎭ ∂f
∂z (x, y, z) = 2z − 2x − y − 5 ∂z (−1, −1, 1) = 0,
therefore, the two rectilinear generators pass through the point A (−1, −1, 1). Their
equations are:
x+1
= y+1
d1 : −1 1
z=1
and, respectively:
x +1 y+1 z−1
d2 : = = .
1 −1 1
Proposition 7.43 (see [4], p. 176). The one-sheeted hyperboloid has two families of
rectilinear generators:
x y
a − c = λ 1 − b
z
(7.81)
λ ax + cz = 1 + by
and
x
−x
z
= λ 1 + by
a c , (√) λ ∈ R. (7.81)
λ a + cz = 1 − by
Proposition 7.44 (see [4], p. 176). The hyperbolic paraboloid has two families of
rectilinear generators:
x y
a− b = 2λz (7.82)
λ ax + by = 1
and y
a − b = λ , (√) λ ∈ R.
x
(7.83)
λ ax + by = 2z
All of the degenerate quadrics have generators. Therefore, the cylindrical surfaces
and the conical surfaces (including the quadrics by the cone type and the cylinder
type) are ruled surfaces.
Let V (a, b, c) be a fixed point and
f (x, y, z) = 0
π: (7.84)
g (x, y, z) = 0
be a given curve.
290 7 Conics and Quadrics
Definition 7.45 (see [11], p. 87). The surface generated by moving a straight line D
called the generatrix, passing through the fixed point V called vertix and one relieses
on the given curve π, called director curve is called conical surface (Fig. 7.14).
We want to find the equation of the conical surface. An arbitrary straight line,
which passes through the point V has the equations:
x −a y−b z−c 2
= = , l + m 2 + n 2 →= 0.
l m n
We denote l
=α
n , n →= 0. (7.85)
m
n =β
The straight line, which generates the conical surface (the generatrix) will be
x −a y−b z−c
= = . (7.86)
α β 1
The condition of supporting the generatrix on the director curve returns to the
algebraic condition of compatibility of the system:
⎫
⎬ f (x, y, z) = 0
g (x, y, z) = 0 (7.87)
⎭ x−a y−b
α = β = 1 .
z−c
Eliminating now α and β from (7.86) and (7.88) it results that the requested
conical surface has the equation:
7.6 Ruled Surfaces. Surface Generation 291
x −a y−b
ϕ , = 0. (7.89)
z−c z−c
Example 7.46. Find the Cartesian equation of the conical surface, which has the
vertix V (2, 1, 3) and the director curve
z=0
π:
x 2 + y 2 = 9.
Solution
The generatrix of the conical surface has the equations:
x −2 y−1 z−3 2
G: = = , l + m 2 + n 2 →= 0.
l m n
Imposing the condition that the straight line G to support on the curve π we
deduce that the system
⎫ x−2
⎬ l = y−1 m = n
z−3
z=0 (7.90)
⎭
x 2 + y2 = 9
is compatible.
Using the notation (7.85) we achieve:
x −2 y−1 z−3
G: = = . (7.91)
α β 1
2 2
x −2 y−1 x −2 y−1
9 +9 − 12 −6 − 4 = 0.
z−3 z−3 z−3 z−3
Definition 7.47 (see [11], p. 89). We call a cylindrical surface, the surface generated
by a straight line which remains parallel to a given direction and which supports on
a given curve, called the director curve (Fig. 7.15).
7.6 Ruled Surfaces. Surface Generation 293
Let v = (l, m, n) be the given direction and π from (7.84) be the director curve.
An arbitrary straight line, by the director vector v has the equations:
nx − lz = λ
(7.94)
ny − mz = μ.
The condition of supporting the generator on the director curve returns to the
algebraic condition of compatibility of the system:
⎫
f (x, y, z) = 0
⎬
g (x, y, z) = 0
nx − lz = λ
⎭
ny − mz = μ.
ϕ (λ, μ) = 0. (7.95)
Eliminating now λ and μ from (7.94) and (7.95) it results that the requested
cylindrical surface has the equation:
Example 7.48 (see [4], p. 215). Write the equation of the cylindrical surface, which
has the director curve
z=0
π:
x 2 + y 2 + 2x − y = 0
x −1 y−2 z
d: = = .
1 −1 2
Solution
We note that v = (1, −1, 2) is the director vector of the straight line d. An arbitrary
straight line, by the director vector v has the equations:
x +y−3=λ
, (√) λ, μ ∈ R. (7.97)
2x − z − 2 = μ
The condition of supporting the generatrix on the director curve returns to the
algebraic condition of compatibility of the system:
⎫ 2
x + y 2 + 2x − y = 0 ⎫ 2
⎬ ⎬ x + y 2 + 2x − y = 0
z=0
⇔ x +y−3=λ
x +y−3=λ ⎭
⎭ 2x − 2 = μ.
2x − z − 2 = μ
μ+2
x= ; (7.98)
2
substituting this expression of x in the second equation of the system we get:
μ+2 2λ − μ + 4
y =λ+3− = . (7.99)
2 2
If in the first equation of the system, we take into account of (7.98) and (7.99) we
achieve:
2 2
μ+2 2λ − μ + 4 μ + 2 2λ − μ + 4
+ +2· − = 0,
2 2 2 2
i.e
Eliminating μ and λ between (7.100) and (7.97) it follows that the cylindrical
surface equation is
7.6 Ruled Surfaces. Surface Generation 295
Definition 7.49 (see [11], p. 91). We call a surface of rotation, a surface generated
by rotating a curve around of a straight line called the axis of rotation (Fig. 7.16).
296 7 Conics and Quadrics
x − x0 y − y0 z − z0 2
= = , l + m 2 + n 2 →= 0
l m n
and the curve π, which one rotates being defined in (7.84).
The surface from the above figure can also result by the displacement of a circle,
parallel with itself, with variable radius, perpendicular to the axis of rotation and
which one relies on the given curve.
In order to satisfy the conditions (7.95), the circle equations will be achieved by
crossing a sphere with a variable radius and having the center on the rotation axis
with a family of planes, perpendicular on the rotation axis; therefore
(x − x0 )2 + (y − y0 )2 + (z − z 0 )2 = α2
G: (7.101)
lx + my + nz = β.
The condition of supporting the generatrix on the director curve returns to the
algebraic condition of compatibility of the system:
⎫
f (x, y, z) = 0
⎬
g (x, y, z) = 0
(x − x0 )2 + (y − y0 )2 + (z − z 0 )2 = α2
⎭
lx + my + nz = β.
Eliminating now α and β from (7.101) and (7.88) it results that the requested
cylindrical surface has the equation:
⎨ ⎩
ϕ (x − x0 )2 + (y − y0 )2 + (z − z 0 )2 , lx + my + nz = 0. (7.102)
7.7 Problems
1. Let be the points: A (−1, 4) , B (3, −2). Write the equation of the circle, which
has AB as a diameter.
Solution
Solving this problem in Sage, we have:
298 7 Conics and Quadrics
2x 2 + 4y 2 − 5 = 0.
π : 4x 2 − 12x y + 9y 2 − 2x + 3y − 2 = 0.
Bring it to the canonical form, indicating the required reference changes and
recognize the achieved conic.
4. Write the equation of the parabola which passes through the points:
Solution
The solution in Sage is:
7.7 Problems 299
5. Let be the points A (−1, 1, 2) , B (1, 3, 3). Write the equation of the sphere with
center in the point and which passes through the point B.
6. Write the equations of the rectilinear generators
x2 y2
− = z,
16 4
that are parallel to the plane π : 3x + 2y − 4z = 0.
7. Find the rectilinear generators of the quadrics :
ρ : x 2 + 3y 2 + 4yz − 6x + 8y + 8 = 0.
Solution
we need the following Sage code to solve this problem:
300 7 Conics and Quadrics
7.7 Problems 301
302 7 Conics and Quadrics
8. Find the equation of the conical surface having as vertex the point V (−3, 0, 0)
and the director curve:
2
3x + 6y 2 − z = 0
π:
x + y + z − 1 = 0.
9. Determine the equation of the cylindrical surface that has the director curve:
x 2 + y2 − 1 = 0
π:
z=0
d : x = y = z.
10. Write the equation of the surface generated by rotating the parabola
y 2 = 2 px
p:
z=0
References
A Dimension formula, 52
Algebraic multiplicity, 117 Direct sum, 52
Angle, 13, 14, 83–85, 139, 227 Distance, 13, 74, 75, 79, 80
Arc element of a curve, 201
Associated matrix, 97
Automorphism, 92 E
Eigensubspace, 110
Eigenval, 110
B Eigenvalue method, 180
Basis, 27 Eigenvector, 110, 111
Bilinear form, 165 Ellipse, 240, 241
Binormal, 201 Ellipsoid, 274
Binormal versor, 202 Elliptic paraboloid, 282, 283
Endomorphism, 92
Envelope, 214
C Euclidean coordinates, 11
Canonical basis, 169 Euclidean real vector space, 135
Cartesian reference, 10
Center of symmetry, 236
F
Change the orthonormal bases, 144
Family of curves, 214
Characteristic equation, 113
Fascicle axis, 73
Characteristic polynomial, 113
First Frenet formula, 205
Circle, 238
First fundamental form, 223
Common perpendicular, 80
Formulas of changing a vector coordinates,
Cone, 276
40
Conic, 235
Free vector, 3
Conical surface, 289, 290
Frenet Trihedron, 199
Coordinates, 36, 37
Cross product, 16, 17
Curvature, 206 G
Curvilinear coordinates, 218 Gauss-Lagrange, 169
Cylindrical surface, 292, 293 Geometric multiplicity, 117
Gram-Schmidt orthogonalization, 145
D
Defect, 97 H
Diagonalizable, 117 Hamilton-Cayley, 115