Notes On Basic Mathematics
Notes On Basic Mathematics
Mathematics
Common Derivations and Techniques
Author
R AIYAN H AQUE
Contents
1 Exponents and Logarithms 4
1.1 Rules for Indices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2 Rules for Logarithms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4 Coordinate Systems 8
4.1 Cartesian Coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.2 Polar Coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
5 Conic Sections 10
5.1 The General Ellipse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
5.2 Circle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
5.3 Parabola . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
5.4 Hyperbola . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
5.5 Rectangular Hyperbola . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
5.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
6 Matrices 20
6.1 Representing Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
6.2 Transpose Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
7 Transformations 21
7.1 Translation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
7.2 Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
7.3 Enlargement or Dilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
7.4 Reflection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
7.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
8 Trigonometry 29
8.1 Trigonometric Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
8.2 Graphs of Trigonometric Function . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
9 Hyperbolic Functions 34
9.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
9.2 Graphs of Basic Hyperbolic Funtions . . . . . . . . . . . . . . . . . . . . . . . . . . 35
9.3 Formulas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
10 Series 42
10.1 The Sigma Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
10.2 Sum of the First n Positive Integers . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
10.3 Sum of the Squares of the First n Positive Integers . . . . . . . . . . . . . . . . . . 43
10.4 Sum of the Cubes of the First n Positive Integers . . . . . . . . . . . . . . . . . . . 43
10.5 Arithmetic Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
10.6 Geometric Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
10.7 Common Series Formulae [Summary] . . . . . . . . . . . . . . . . . . . . . . . . . 47
1
11 Quadratic Functions, Equations & Inequalities 48
11.1 General Quadratic Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
11.2 Quadratic Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
11.3 Quadratics in Graphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
13 Multivariable Calculus 60
13.1 Partial Derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
14 Complex Numbers 61
14.1 Real and Imaginary Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
14.2 Argand Diagrams and Complex Number Representation . . . . . . . . . . . . . . 63
14.3 Simple Operations with Complex Numbers . . . . . . . . . . . . . . . . . . . . . . 67
14.4 Triangle Inequality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
14.5 de Moivre’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
14.6 Circles on the Complex Plane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
14.7 Complex Trigonometric and Hyperbolic Functions . . . . . . . . . . . . . . . . . . 76
14.8 Logarithms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
16 Vector Analysis 85
16.1 Scalars and Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
16.2 Vector Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
16.3 Vector Equation of Straight Line . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
16.4 Vector Calculus: Differentiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
16.5 Vector Calculus: Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
16.6 Green’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
16.7 Kelvin-Stokes Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
16.8 Gauss’s Divergence Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
2
20.1 Fourier Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
20.2 Complex Fourier Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
20.3 Fourier Transform . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
23 Symbols 124
3
1 Exponents and Logarithms
1.1 Rules for Indices
In the following, p, q ∈ R, a, b ∈ R + and m, n ∈ Z + .
1. a p aq = a p+q
ap
2. aq = a p−q
3. ( a p )q = a pq
4. a0 = 1 { a ̸= 0}
5. a− p = 1
ap
6. ( ab) p = a p b p
√ 1
7. n a = a n
√ m
8. n am = a n
√
n a
9. n ba = √
p
n
b
In a p , p is called the exponent, a is the base and a p is called the pth power of a.
The function y = a x is called an exponential function.
4
1.2 Rules for Logarithms
Any logarithmic function is the inverse of an exponential function. It is given in the form,
f ( x ) = loga b
The function is said “the logarithm of b with base a”, and what this means will be cleared out
in a bit. First, we need to know how the numbers a and b make the function behave.
f ( x ) = loge b = ln b
f ( x ) = log10 b ≡ log b ≡ lg b
Logarithm Formulae
1. If a x = b, x = loga b
∴ If (base) power = product, then power = logbase product
2. loga 1 = 0
3. loga ( MN ) = loga M + loga N
4. loga ( M
N ) = loga M − loga N
5. loga Mr = r loga M
6. loga a = 1
1
7. loga b = logb a
logc b
8. loga b = logc a [Base Conversion]
5
2 Permutation and Combination
Permutation and combination are the ways to represent a group of objects by selecting them in
a set and forming subsets. It defines the various ways to arrange a certain group of data. When
we select the data or objects from a certain group, it is said to be permutations, whereas the
order in which they are represented is called combination. Both concepts are very important
in Mathematics.
2.1 Permutation
Permutation relates to the act of arranging all the members of a set into some sequence or or-
der. In other words, if the set is already ordered, then the rearranging of its elements is called
the process of permuting. Permutations occur, in more or less prominent ways, in almost ev-
ery area of mathematics. They often arise when different orderings on certain finite sets are
considered.
A permutation is the choice of r things from a set of n things without replacement and where
the order matters.
2.2 Combination
The combination is a way of selecting items from a collection, such that (unlike permutations)
the order of selection does not matter. In smaller cases, it is possible to count the number of
combinations. Combination refers to the combination of n things taken r at a time without
repetition.
A combination is the choice of r things from a set of n things without replacement and where
order does not matter.
6
3 The Binomial Theorem
Any expression in the form ( a + b)n is called a binomial of power n (also, an expression
( a + b + c)n is known as a trinomial, and so on).
n!
Here, n ∈ Z and we know that n Cr = r!(n−r )!
. Therefore, we can see that the binomial expan-
sion of ( a + b)n becomes,
( a + b ) n = a n (1 + x ) n
∞
= an ∑ (n Cr xr )
r =0
n(n − 1) 2 n(n − 1)(n − 2) 3
n
=a 1 + nx + x + x + ... + x n
2! 3!
Here, the binomial expansion is represented in ascending powers of x with the binomial coef-
ficients, which is wholely multiplied by a factor of an . This is more managable in most cases.
7
4 Coordinate Systems
4.1 Cartesian Coordinates
8
4.2 Polar Coordinates
Polar Coordinates in Two Dimensions
A point P is described as being a distance x horizontally and y verticallt from the origin, where
the horizontal and vertical directions are marked with x and y axes respectively. An alternative
way of describing the position of P is to use polar coordinates.
We use the origin O as a starting point and measure the distance ⃗r of P from O, soOP = ⃗r. We
call the point O the pole. Of course just knowing how far from O the point P is does not help
us to uniquely identify its position. The second measurement we make is the angle measured
from the positive x-axis in an anti-clockwise direction. We call the positive x-axis the inital line
(or initial beam), the angle is ususally denoted as θ and the polar coordinates are written in as
(r, θ ).
The following are some formulae linking polar coordinates and Cartesian coordinates. From
the figure above, it is clear that:
• r cos θ = x
r sin θ = y
• r 2 = x 2 + y2
y
θ = arctan x
9
5 Conic Sections
In mathematics, a conic section (or simply conic) is a curve obtained as the intersection of the
surface of a cone with a plane. The three types of conic section are the hyperbola, the parabola,
and the ellipse; the circle is a special case of the ellipse, though historically it was sometimes
called a fourth type.
Here, a, b, c, f , g, and h are constants. The nature of the conic is determined by the values
of these constants, and can be interpreted using the determinant delta ∆, which is just the
determinant of a 3 × 3 matrix.
a h g
∆= h b f
g f c
10
5.1 The General Ellipse
A standard ellipse with center at the origin O has the Cartesian equation,
x2 y2
+ =1
a2 b2
x2
2. When y = 0, then a2
= 1, and so,
x = ±a
11
Features of an Ellipse
Consider the ellipse above. The labels are explained in the table below.
The conic section is an ellipse only when the eccentricity is such that 0 ≤ E < 1.
12
If the ellipse is translated such that the center is now at the (Cartesian) coordinates (h, k ), the
standard equation can be modified to a more general form.
2 2
x−h y−k
+ =1
a b
a (1 − E2 ) a (1 − E2 )
r= =
1 ± E cos θ 1 ± E sin θ
Note that the polar equation presumes the center of the ellispe to be at origin.
13
5.2 Circle
2
y−k 2
We already found the general equation for an ellipse to be x−a h + b = 1. Now, con-
sider that a and b are equal. So, we can replace them by the parameter r. So, we can write,
2 2
x−h y−k
+ =1
r r
( x − h )2 + ( y − k )2 = r 2
We can see that this is the equation for a circle of radius r, centered at (h, k ).
The eccentricity of a circle can be calculated as follows.
r
b2
E = 1− 2
a
For a = b, √
E= 1−1 = 0
14
5.3 Parabola
A parabola is any curve with the general Cartesian equation in the form,
y2 = 4ax
Here, a ∈ R + is a constant.
This curve is symmetrical about the x-axis. Note that if it had the equation x2 = 4ay, it would
resemble a quadratic function curve (also a parabola), which is symmetric about the y-axis.
A parabola can be said to be the locus of points where every point P( x, y) on the parabola is
the same distance from a fixed point S, called the focus. and a fixed straight line called the
directrix.
15
Parameterizing the Parabola
The parametric form of the equation of the parabola is given by the following set of equations.
(
x = at2
y = 2at
16
5.4 Hyperbola
17
5.5 Rectangular Hyperbola
18
5.6 Summary
19
6 Matrices
6.1 Representing Matrices
Matrices are arrays of elements in an m × n arrangement. A matrix A can be expressed in the
form,
a11 a12 a13 · · · a1n
a21 a22 a23 · · · a2n
A = a31 a32 a33 · · · a3n
.. .. .. .. ..
. . . . .
am1 am2 am3 · · · amn
a11 a12 a13 ··· a1n
a21
a22 a23 ··· a2n
A T = a31
a32 a33 ··· a3n
.. .. .. .. ..
. . . . .
am1 am2 am3 · · · amn
20
7 Transformations
7.1 Translation
Translation along x-axis
21
Translation along y-axis
Consider the equation of the parabola from before y = f ( x ), for which f ( x ) = ( x − 5)2 .
When we translate this such that y now becomes equal to f ( x ) + 2, the new equation would
be y = f ( x ) + 2 = ( x − 5)2 + 2. By the rule of translation, we can simply draw the new curve
by shifting the entire original curve by 2 units upwards (in the direction of the positive y-axis).
Of course, if the translation is such that the function becomes f ( x ) − 2, the equation becomes
y = f ( x ) − 2 = ( x − 5)2 − 2, and by the rule of translation, we can simply draw the new curve
by shifting the entire original curve by 2 units downwards (in the direction of the negative
y-axis).
22
Translation along both x and y-axes
We can also do a composite translation by shifting a curve along both the axes. A translation
of the curve y = f ( x + a) + b implies that the entire original curve y = f ( x ) is shifted a units
to the right and b units upwards.
For example, for the parabola y = f ( x ) = ( x − 5)2 , a translation of f ( x + 3) − 2 implies a shift
of the original curve by 3 units to the left and 2 units downwards.
23
7.2 Rotation
Rotation of a Function
24
Rotation Matrix
Individual points (or a set of distinct points) can be rotated about the origin using rotation
matrices. The general matrices for rotation in two dimensions by an angle of θ is given as
follows.
• Counter-clockwise (Anti-clockwise)
cos θ − sin θ
sin θ cos θ
• Clockwise
cos θ sin θ
− sin θ cos θ
25
7.3 Enlargement or Dilation
26
7.4 Reflection
Special Case: Inverse Function
27
7.5 Summary
Translation and Enlargement/Dilation
Rotation
28
8 Trigonometry
8.1 Trigonometric Operators
The basic trigonometric operators are the sine and cosine. The other operators are all derived
from these basic ones. The third operator, the tangent, is defined as the ratio of the sine and
cosine. But more on this later.
29
8.2 Graphs of Trigonometric Function
The Sine Function
30
The Tangent Function
31
The Secant Function
32
Comparison
33
9 Hyperbolic Functions
Hyperbolic functions are similar to trigonometric functions in certain aspects. However, hy-
perbolic functions are defined in terms of exponential functions.
9.1 Definitions
Hyperbolic Functions
The basic hyperbolic functions are the hyperbolic sine and the hyperbolic cosine functions.
The other four hyperbolic functions are related to the hyperbolic sine and cosine functions in
the same way that the corresponding trigonometric functions are related to sine and cosine.
Hyperbolic Tangent { x ∈ R }:
sinh x e2x − 1
tanh x = = 2x
cosh x e +1
Hyperbolic Secant { x ∈ R }:
1 2
sech x = = x
cosh x e + e− x
1 2
csch x = = x
sinh x e − e− x
1 cosh x e2x + 1
coth x = = = 2x
tanh x sinh x e −1
34
Inverse Hyperbolic Functions
35
The Hyperbolic Tangent Function
36
The Hyperbolic Cosecant Function
37
Comparison
38
9.3 Formulas
Relationships Among Hyperbolic Functions
1. cosh2 x − sinh2 x = 1
2. sech2 x + tanh2 x = 1
3. coth2 x + csch2 x = 1
Addition Formulas
1. sinh ( x ± y) = sinh x cosh y ± cosh x sinh y
2. cosh ( x ± y) = cosh x cosh y ± sinh x sinh y
tanh x ±tanh y
3. tanh ( x ± y) = 1±tanh x tanh y
1±coth x coth y
4. coth ( x ± y) = coth x ±coth y
q
2. cosh ( 2 ) = cosh2x+1
x
q
x x −1 sinh x cosh x −1
3. tanh ( 2 ) = ± cosh
cosh x +1 {+ if x > 0, − if x < 0} = cosh x +1 = sinh x
39
Multiple Angle Formulas
1. sinh (3x ) = 3 sinh x + sinh3 x
2. cosh (3x ) = 4 cosh3 x − 3 cosh x
3 tanh x +tanh3 x
3. tanh (3x ) =
1+3 tanh2 x
2. cosh2 x = 1
2 cosh (2x ) + 1
2
3. sinh3 x = 1
4 sinh (3x ) − 34 sinh x
4. cosh3 x = 1
4 cosh (3x ) + 34 cosh x
5. sinh4 x = 3
8 − 12 cosh (2x ) + 18 cosh (4x )
6. cosh4 x = 3
8 + 21 cosh (2x ) + 18 cosh (4x )
40
Relationships Between Inverse Hyperbolic Functions
1. arccsch = arcsinh ( 1x )
41
10 Series
10.1 The Sigma Notation
The sigma notation (Σ) is a very useful and concise way to define a series. It makes further
study of series more manageable.
Here, Ur is a function of r.
The notation implies that the series starts from the r th term, all the way to the nth term, which
needs to be added up to for finding the sum. In this case, r = 1 denotes that the series starts
from the 1st term. If the value of r is any number other than 1, we can use the formula:
n n k
∑ Ur = ∑ Ur − ∑ Ur
r =k r =1 r =1
Here, k is another number such that the series starts from the kth term. The logic that is used
for the formula above is that to find the sum of the terms, from the kth term to the nth term of
a series, we can find the sum of the first k terms and subtract this from the sum of the first n
terms.
Series such as arithmetic series, geometric series and the binomial series can all be written in
Σ notation. This will be discussed later.
For finding the summation of the first n integers starting from 1, we can write it as,
Sn = 1 + 2 + 3 + 4 + ... + n
To find the summation formula, first we consider the binomial expansion of (r − 1)3 .
(r − 1)3 = r3 − 3r2 + 3r − 1
To find the summation formula, first we consider the binomial expansion of (r − 1)4 .
43
!
n n n
3
4 ∑r = n4 + 6 (n + 1)(2n + 1) − 4 ( n + 1) + n
r =1
6 2
n
1 1 1
∑ r 3 = 4 n4 + 2 n3 + 4 n2
r =1
n
n2
∑ r3 = 4
( n + 1)2
r =1
Here, U1 is the first term of the series. d is called the “common difference”, which is just the
difference between any two consecutive terms of the sequence, such that d = rk+1 − rk .
To find the sum of the first n terms, we can express it in sigma notation.
n
∑ U1 + (r − 1)d
r =1
Manually, the sum S(n) of the first n terms can be written as,
S(n) = U1 + (U1 + d) + (U1 + 2d) + (U1 + 3d) + ... + (U1 + (n − 1)d)
We could have also started with the nth term and successively subtracted the common differ-
ence. So,
S(n) = Un + (Un − d) + (Un − 2d) + (Un − 3d) + ... + (Un − (n − 1)d)
Both the approaches lead us to the sum of the series. However, if we looked at the equations,
we can see that if we add the two equations together, the terms add out.
2S(n) = (U1 + Un ) + (U1 + Un ) + (U1 + Un ) + ... + (U1 + Un )
We can also modify the equation S(n) = n2 (U1 + Un ). We know that the nth term is Un =
U1 + (n − 1)d. Putting this value in our formula, we get,
n
S(n) = (U1 + U1 + (n − 1)d)
2
n
S(n) = (2U1 + (n − 1)d)
2
This is another formula for finding the sum.
44
10.6 Geometric Series
Geometric series are those whose terms can be defined by the following formula:
U (r ) = U1 Rr−1
Here, U1 is the first term of the series. R is called the “common ratio”, which is just the ratio
r
between any two consecutive terms of the sequence, such that R = kr+k 1 .
To find the sum of the first n terms, we can express it in sigma notation.
n
∑ U1 Rr−1
r =1
Manually, the sum S(n) of the first n terms can be written as,
Note that if we reverse both subtractions in the fraction above, we will obtain the same func-
tion.
xn − 1 1 − xn
=
x−1 1−x
Applying the formula to the geometric summation (by reading R instead of x), we get,
Both of the formulae can be used to find the sum of the first n terms of geometric sequences.
Special Case
45
If the common ratio is such that −1 < R < 1, then the sum of the terms of the geometric series
is convergent to a finite value as the number of terms tend to infinity.
We know that,
n
1 − Rn
r −1
∑ U1 R = U1
1−R
r =1
We can see that for −1 < R < 1, Rn → 0 as n → ∞. Putting this in the summation formula,
n
1−0
r −1 U1
∑ U1 R = U1
1−R
=
1−R
r =1
So, this formula can be used to the sum to infinity of convergent geometric sequences.
46
10.7 Common Series Formulae [Summary]
Series Sigma Notation Formula Validity
n2
Sum of the cubes of the first n integers ∑rn=1 r3 4 (n + 1)2 r, n ∈ Z +
Sum to infinite r, n ∈ Z +
∑r∞=1 U1 Rr−1
U1
terms of a convergent 1− R −1 < R < 1
geometric series
47
11 Quadratic Functions, Equations & Inequalities
11.1 General Quadratic Function
Expanded Form
f ( x ) = ax2 + bx + c
Here, a, b, c ∈ R and a ̸= 0.
The quadratic function is also called a second order function because the highest power of x is
2.
Completing Squares
Recall the formula ( p + q)2 = p2 + 2pq + q2 . We can apply this in the bracket as,
2 2 !
b b b
f (x) = a x2 + 2 × ×x+ − +c
2a 2a 2a
2 ! 2
2 b b b
= a x +2× ×x+ +c−a
2a 2a 2a
2
b2
b
= a x+ +c−a× 2
2a 4a
2 2
b b
= a x+ + c−
2a 4a
b b2
Let A = a, B = 2a and C = c − 4a . So, we can write this function as,
f ( x ) = A ( x + B )2 + C
48
11.2 Quadratic Equations
A quadratic equation in x is any equation second order in x in terms of exponent. Obviously,
this is in the form,
ax2 + bx + c = 0
As seen before, we can complete squares of the quadratic function, and thus, the quadrating
function can also be written in the form,
A ( x + B )2 + C = 0
We solve quadratic equations of the form ax2 + bx + c = 0 using the following method.
ax2 + bx + c = 0
b c
x2 + x + = 0
a a
b c
x2 + x = −
a a
2 2
2 b b c b
x + x+ =− +
a 2a a 2a
2
b2
b c
x+ =− + 2
2a a 4a
2
b2 − 4ac
b
x+ =
2a 4a2
r
b b2 − 4ac
x+ =±
2a 4a2
√
b b2 − 4ac
x=− ±
2a 2a
√
−b ± b2 − 4ac
x=
2a
We found that if we solve a quadratic equation, we get two "solutions". These two solutions
are are just as follows. √ √
−b + b2 − 4ac −b − b2 − 4ac
x= ,
2a 2a
The two roots are often denoted as α and β. So, the solution of the quadratic becomes x = α, β.
49
Discriminant
Here, the term inside the square root is known as the determinant. So, the determinant is
D = b2 − 4ac. √
−b ± D
∴x=
2a
The way the discriminant describes the roots is shown in the table below.
Condition Interpretation
50
α and β form of the Quadratic Equation
So, we got α + β = − ba and αβ = ac . Let us return to the general quadratic equation, ax2 + bx +
c = 0. It can be written as,
ax2 + bx + c = 0
b c
x2 + x + = 0
a a
2 b c
x − − x+ =0
a a
∴ x2 − (α + β) x + αβ = 0
Therefore, we can see that any general quadratic equations can be solved without using the
quadratic formula also, by solving two equations in α and β simultaneously.
(
α + β = − ba
αβ = ac
51
Factorized Form
Therefore, the equation can also be written as factors of the roots, that is,
( x − α)( x − β) = 0
Of course, this is not only inherent about the equation, but also for the quadratic function. A
quadratic function can be factorized to f ( x ) = ( x − α)( x − β) from the form f ( x ) = ax2 + bx +
c using the quadratic formula, assuming that f ( x ) = 0.
52
11.3 Quadratics in Graphs
The most basic quadratic function is f ( x ) = x2 . Graphically, it is as follows. For better under-
standing of how the shape of the curve y = ax2 changes as the value of a increases or decreases,
curves with different values of a are provided.
53
Meaning of the Expanded form
54
Meaning of the Completed Square Form
Consider a quadratic equation of whose the squares are completed, y = −2( x − 2)2 + 2. Even
without drawing the curve, we can infer that the curve will have a maximum point at the
coordinates (2, 2) Similarly, the curve of the equation y = 2( x + 2)2 − 2 will have a minimum
point at the coordinates (−2, −2).
55
Meaning of the Factorized Form
The factorized form of quadratic equation is straight-forward in its meaning. If the equation is
in the form f ( x ) = ( x − α)( x − β), then we can infer the roots of the equation directly, x = α, β.
Consider an equation y = −( x − 2)( x − 4). Then, the roots are x = 2, 4 and thus, the x-
intercepts are the coordinates (2, 0) and (4, 0). Also, if the equation is y = x ( x + 3), the roots
are x = 0, 3 and thus, the x-intercepts are the coordinates (0, 0) and (−3, 0). Both the curves
are shown in the figure below.
56
12 Single Variable Calculus
Calculus is the measure of change.
12.1 Differentiation
First Principles
As the point B moves closer to the point A, the gradient of the chord AB gets closer to the
gradient of the tangent to the curve at A.
We can "formalise" this approach by letting the x-coordinate of A be x0 and the x-coordinate
of B be x0 + h. Consider what happens to the gradient of AB as h gets smaller.
57
The vertical distance from A to B is f ( x0 + h) − f ( x0 ), and the horizontal distance between the
points is x0 + h − x0 = h.
As the gradient is simply the ration between the vertical and horizontal distances, we can write
f ( x + h)− f ( x0 )
that the gradient of AB is 0 h .
As h gets smaller, the gradient of AB gets closer to the gradient of the tangent to the curve at
A. This means that the gradient of the curve at A is the limit of this expression as the value of
h tends to zero.
dy
The gradient function, or the derivative, of the curve y = f ( x ) is written as f ′ ( x ) or dx .
f ( x0 + h ) − f ( x0 )
f ′ ( x ) = lim
h →0 h
The gradient function can be used to find the gradient of the curve for any value of x.
Using this rule to find the derivative is called differentiating from the first principles.
58
12.2 Integration
59
13 Multivariable Calculus
13.1 Partial Derivatives
60
14 Complex Numbers
14.1 Real and Imaginary Numbers
Assume a quadratic equation,
ax2 + bx + c = 0
The nature of the roots can be determined even by not solving the equation whatsoever. This
is done by finding out the discriminant of the quadratic equation. The discriminant D of a
quadratic equation of the form as shown above is as follows.
D = b2 − 4ac
The formula we got is known as the quadratic formula. Here, we can see the discriminant
inside a square root, which when is less than zero, cannot give a real solution. This is because
no real numbers can be squared to get a negative number. To solve such problem, complex
numbers are used.
To keep things simple, an imaginary unit, i (or sometimes j in electrical engineering), is intro-
duced which is such that, √
i = −1
( i 2 = −1)
z = a + bi
Here, a, b ∈ R. The real part of z is Re[z] = a and the imaginary part of z is Im[z] = b. We can
write down the complex conjugate, z∗ , of a complex number, which is such that,
z∗ = a − bi
The complex numbers z and z∗ are called the complex conjugate pair of each other. All
quadratic equations with imaginary roots have roots that are the complex conjugate of each
other.
NOTE:
• If a complex number z = a + bi is such that a ̸= 0, then it is a partially complex number.
• If a complex number z = a + bi is such that a = 0, then it isa perfectly complex number.
61
Polynomial Equations with Higher Powers of x
Polynomial equations with more than two roots have some trend in terms of nature of roots.
If all the roots are real, they can be any real number. There cannot be only one complex root
of a polynomial, while other roots are real. This is because complex roots come in conjugate
pairs. So, there have to be even numbers of roots which are complex. For example, a quartic
equation has either,
1. All four roots real
2. Two roots real and the other two roots form a complex conjugate pair
3. Two roots form a complex conjugate pair and the other two roots also form a complex
conjugate pair
z + z∗ = ( a + bi ) + ( a − bi ) = 2a
z − z∗ = ( a + bi ) − ( a − bi ) = 2bi
Therefore, the real and imaginary parts of the complex number z can be written as,
(
Re[z] = 21 (z + z∗ )
Im[z] = 2i1 (z − z∗ ) = − 2i (z − z∗ )
The following inequalities hold for the real and imaginary parts of a complex number.
(
| Re[z]| ≤ |z|
| Im[z]| ≤ |z|
62
14.2 Argand Diagrams and Complex Number Representation
We can represent complex numbers on a diagram, called an Argand diagram. Argand dia-
grams are like graphs, but with a real axis and an imaginary axis. We regard the real "line" or
axis as embedded in the complex plane. R ⊂ C.
NOTE:
• If a number is located on the real axis, it is a real number.
• If a number is located on the imaginary axis, it is a perfectly complex.
• If a number is located anywhere in between, it is partially complex.
Cartesian Form
If we consider the x-axis of a graph to be the real axis and the y-axis to be the imaginary axis,
then any complex number with the form z = x + yi can be shown in the graph as a Cartesian
coordinate, ( x, y). So, we can define an Argand diagram as a Cartesian coordinate diagram to
represent complex numbers.
63
Vector Notation
Polar Form
We can also draw an Argand diagram using polar coordinates. For this, the Cartesian coordi-
nates of z must be converted to the polar equivalent. Then, we will get the polar coordinates
of the complex number in the form (r, θ ).
64
r is called the modulus or norm of z. It is also denoted by |z|. We can calculate the norm of z
using the following method. {r ∈ R }.
We can manipulate this equation and derive another very fundamental formula.
| z |2 = x 2 + y2
= x2 − (yi )2
= ( x + yi )( x − yi )
= zz∗
θ is also known as the argument of z, and is also denoted by arg(z). The argument is such that
−π < θ < π (or −180 < θ < 180). This is sometimes referred to as the principle argument.
The value of θ which is put in the polar coordinate (r, θ ) depends on the quadrant in which the
complex number is located.
If,
y
tan (φ) =
x
y
φ = arctan
x
Quadrant Value of θ
First quadrant φ
Second quadrant π−φ
Third quadrant −(π − φ)
Fourth quadrant −φ
65
Modulus-Argument Form
z = r (cos θ + i sin θ )
The way to determine the values of r and θ is the same as before, and this is true for a complex
number in any of Argand diagram quadrants.
Exponential Form
We can find the series expansions of the functions cos θ and sin θ. They are,
θ2 θ4 θ6 (−1)r θ 2r
cos θ = 1 − + − + ... + + ...
2! 4! 6! (2r )!
θ3 θ5 θ7 (−1)r θ 2r+1
sin θ = θ − + − + ... + + ...
3! 5! 7! (2r + 1)!
x x2 x3 xr
ex = 1 + + + + ... + + ...
1! 2! 3! r!
It can be proved that the series expansion for e x is also true if x is replaced by a complex
number. If we replace x in e x by iθ, the series expansion becomes,
66
14.3 Simple Operations with Complex Numbers
Let two numbers be,
(
z1 = a + bi
z2 = c + di
z1 + z2 = ( a + bi ) + (c + di )
= a + c + b − +di
= ( a + b ) + ( b + d )i
Multiplication
cz1 = c( a + bi )
= ac + bci
z1 z2 = ( a + bi )(c + di )
= ac + bci + adi + bdi2
= ac + bci + adi − bd
= ( ac − bd) + (bc + ad)i
The result is also in modulus-argument form and has a modulus r1 r2 and argument (θ1 + θ2 ).
67
We can see that this result gives the formula,
Therefore, the complex number z1 z2 = r1 r2 ei(θ1 +θ2 ) is in exponential form and has a modulus
r1 r2 and argument (θ1 + θ2 ).
Division
z1
The value of z2 will be such that,
z1 a + bi
=
z2 c + di
a + bi c − di
= ×
c + di c − di
( a + bi )(c − di )
=
(c + di )(c − di )
( a + bi )(c − di )
=
c2 + d2
( ac + bd) + (bc − ad)i
=
c2+ d2
ac + bd bc − ad
= + i
c2 + d2 c2 + d2
68
z1
Then, z2 is,
z1 r (cos θ1 + i sin θ1 )
= 1
z2 r2 (cos θ2 + i sin θ2 )
r1 (cos θ1 + i sin θ1 ) (cos θ2 − i sin θ2 )
= ×
r2 (cos θ2 + i sin θ2 ) (cos θ2 − i sin θ2 )
r (cos θ1 cos θ2 − i cos θ1 sin θ2 + i sin θ1 cos θ2 − i2 sin θ1 sin θ2 )
= 1
r2 (cos θ2 cos θ2 − i cos θ2 sin θ2 + i sin θ2 cos θ2 + i2 sin θ2 sin θ2 )
r [(cos θ1 cos θ2 + sin θ1 sin θ2 ) − i (cos θ1 sin θ2 − sin θ1 cos θ2 )]
= 1
r2 (cos2 θ2 + sin2 θ2 )
r
= 1 [cos (θ1 + θ2 ) + i sin (θ1 + θ2 )]
r2
Therefore the complex number zz12 = rr21 [cos (θ1 + θ2 ) + i sin (θ1 + θ2 )] is in modulusargument
form and has a modulus rr12 and argument (θ1 − θ2 ).
z1
Then, z2 is,
z1 r eiθ1
= 1 iθ
z2 r2 e 2
r
= 1 eiθ1 e−iθ2
r2
r
= 1 eiθ1 −iθ2
r2
r
= 1 e i ( θ1 − θ2 )
r2
z1 r1 i ( θ1 − θ2 )
Therefore the complex number z2 = r2 e is in modulus-argument form and has a modu-
lus rr12 and argument (θ1 − θ2 ).
Reciprocal
This is similar to the division method. The reciprocal of the complex number z is found as
follows.
1 1 z∗
= × ∗
z z z
z∗
= ∗
zz
z∗
= 2
|z|
69
Summary
• |z1 z2 | = |z1 ||z2 |
• arg (z1 z2 ) = arg (z1 ) + arg (z2 )
z1 | z1 |
• z2 = | z2 |
70
14.4 Triangle Inequality
Consider two complex numbers z and w. Therefore,
|z + w|2 = (z + w)(z∗ + w∗ )
= zz∗ + zw∗ + z∗ w + ww∗
= |z|2 + 2Re[zw∗ ] + |w|2
≤ |z|2 + 2|zw∗ | + |w|2
We know that |zw| = |z||w| and |z∗ | = |z| (geometrically, this is simple and trivial). So, we can
say that |zw∗ | = |z||w∗ | = |z||w|. Now, we can write,
∴ |z + w| ≤ |z| + |w|
71
14.5 de Moivre’s Theorem
de Moivre’s theorem states that,
zn = [r (cos θ + i sin θ )]n = r n (cos (nθ ) + i sin (nθ ))
The theorem states [r (cos θ + i sin θ )]n = r n (cos (nθ ) + i sin (nθ )).
When n = 1,
If de Moivre’s theorem is true for n = k, then it has been shown to be true for n = k + 1. As de
Moivre’s theorem is true for n = 1, it is now also true for all n ≥ 1 and n ∈ Z + by mathemati-
cal induction.
The theorem states [r (cos θ + i sin θ )]n = r n (cos (nθ ) + i sin (nθ )).
If n is a negative integer, it can then be written in the form n = −m, where m is a positive
integer.
Therefore, we have proved that de Moivre’s theorem is true when n is a negative integer.
The theorem states [r (cos θ + i sin θ )]n = r n (cos (nθ ) + i sin (nθ )).
When n = 0,
Extras
73
If z = r (cos θ + i sin θ ) = reiθ , then,
Therefore,
[reiθ ]n = r n einθ
74
14.6 Circles on the Complex Plane
75
14.7 Complex Trigonometric and Hyperbolic Functions
76
14.8 Logarithms
Given that z ̸= 0, we can find w ∈ C, such that ew = z. Writing the expressions of z and w, we
get, (
z = reiθ {r > 0}
w = u + iv
Therefore,
ew = z
eu+iv = reiθ
eu eiv = reiθ
ln z := ln |z| + i arg z
Ln z := ln |z| + iθ
77
15 Calculus of Complex Functions
15.1 Complex Derivatives and Analytics Functions
Definition of Derivative
f (z) − f ( a)
lim =α
z→ a,z̸= a z−a
f ( a + h) − f ( a)
lim =α
h→0,h̸=0 h
Proofs
For (1) ⇔ (2), we can just say z = a + h. Then, we can turn (1) to (2) and vice versa.
f ( a+ h)− f ( a)
(
h − α { h ̸ = 0}
0 { h = 0}
78
Analytic Functions
f (z) − f ( a)
lim =α
z→ a,z̸= a z−a
f (z) − f ( a)
lim = f ′ ( a)
z→ a,z̸= a z−a
The definition of complex differentiation may seem very similar to the defintion of real differ-
entiation, but there are some things to keep in mind. The parameters here are in the complex
plane. Also, the division in the limit is a complex division, and the limit itself is in complex
sense - we can approach a from any directions.
[image]
79
15.2 Rules and Remarks
If the two functions f and g are analytic, ∀z ∈ C and λ ∈ C, then in pointwise notation,
1. ( f + g)(z) := f (z) + g(z)
2. (λ f )(z) := λ f (z)
3. ( f g)(z) := f (z) g(z)
′
f f ′ g− f g′
4. g = g2
Proofs
We know that z = a + h.
Remark (1):
Remark (2):
(λ f )( a + h) − (λ f )( a) − (λ f ′ ( a))h
= λ( f ( a + h) − f ( a) − f ′ ( a)h)
= (λR f ( a; h))h
We can see that λR f ( a; h) = Rλ f ( a; h). When Rλ f ( a; h) → 0, the formula is proved. So, we can
also say that,
(λ f )′ = λ f ′
Remark (3):
( f g)′ = f ′ g + f g′
80
′ ′
It suffices to prove that 1
g = − gg2 , considering f ≡ 1 is a constant function.
1 1 g′ ( a)
− + h
g ( a + h ) g ( a ) g ( a )2
g( a) − g( a + h) g′ ( a)
= + h
g( a + h) g( a) g ( a )2
g( a + h) − g( a) − g′ ( a) h g′ ( a)
1 1
=− − − h
g( a + h) g( a) g( a) g( a + h) g( a)
R g ( a; h) g′ ( a)( g( a + h) − g( a))
= − + h
g( a + h) g( a) g ( a )2 g ( a + h )
R ( a;h) g′ ( a)( g( a+ h)− g( a))
We can see that − g(a+g h) g(a) + g ( a )2 g ( a + h )
= R( 1 ) ( a; h). When R( 1 ) ( a; h) → 0, the formula
g g
is proved.
81
15.3 Cauchy-Riemann Equations
Consider an analytic function f : C → C. The function can be written in the form f (z) = u + iv,
where u = Re[ f ] and v = Im[ f ] are both functions of x and y. If the function is differentiable
at the point z = x + iy, then at z, the first-order partial derivatives of u and v must exist and
satisfy the Cauchy-Riemann (CM) equations below.
(
∂u ∂v
∂x = ∂y
∂u ∂v
∂y = − ∂x
82
15.4 Differentiations of Complex Logarithms
Theorem: The natural logarithm of a complex function is analytic on C \{0} (away from the
origin). Its derivative is given by,
d 1
ln z = {z ∈ C \{0}}
dz z
83
15.5 Complex Integration
Consider a function F (z) such that F ′ (z) = f (z). This means that F (z) is the (complex indefi-
nite) integral of f (z). F (z) is also called the primitive of f (z). As usual, we use integral sign to
represent the relationship between F (z) and f (z).
Z
F (z) = f (z)dz
Definite Integrals
In the real case, we define definite integral in the interval [ a, b] as,
Z b
f (x) = f ( x )dx
a
We want to find a similar form of definite integrals for complex functions in the interval [α, β],
where α, β ∈ C. However, there are infinite paths connecting the points α and β.
84
16 Vector Analysis
16.1 Scalars and Vectors
A quantity that only has magnitude is called a scalar quantity. Examples of scalar quantities
are mass, time, energy, length, density, and volume. Sacalr quantities can be added to each
other according to the simple laws of arithmetic.
A quantity that possesses both magnitude and direction is called a vector quantity. Exam-
ples of vector quantities are displacement, velocity, acceleration, force and momentum. unlike
scalar quantities, vector additions and subtractions do not obey arithmetic laws. Instead, two
vectors are added or combined according to the laws of vector algebra. This, and other vector
operations, will be discussed here.
Vector Notations
To define the absolution position of objects, let us use the Cartesian coordinate system. The R3
space is considered to make the matter more general, but the dimensions can easily be reduced
as per requirement.
In the R3 Cartesian coordinate system, the three dimensional axes are labelled to be the x, y,
and z axes. They are orthogonal to each other (as a necessity). The origin O of the coordinate
system, ( x, y, z) = (0, 0, 0) is the reference point for all other points in the space.
The arrow over OA indicates that the position is marked at A from the origin O by a vector
arrow, whose "tail" is at O and "head" is at A. In other words, the vector notation implies an
arrow pointing towards the point A starting from the pont O.
⃗ can also be denoted in other forms. For example, we can use the column vector
The vector OA
representation to denote the vector.
X
⃗ = Y
OA
Z
Here, the column vector is a 3 × 1 matrix. The unit vectors need not be written, as the top part
of the matrix represents the x̂ unit vector, the middle part of the matrix represents the ŷ unit
vector and the the bottom part of the matrix represents the ẑ unit vector. This method is very
simple to use, especially for the applications.
85
We usually denote the Cartesian coordinates of a point in the form ( x, y, z). We know that the
⃗ can also be denoted using angle
position of A is at the coordinates ( X, Y, Z ). So, the vector OA
brackets as,
⃗ = ⟨ X, Y, Z ⟩
OA
To add two vectors, we add the like components together. The general rule for the vectors
⃗ = ⟨ A x , Ay , Az ⟩ and ⃗B = ⟨ Bx , By , Bz ⟩ is,
A
Ax Bx A x + Bx
⃗ + ⃗B = Ay + By = Ay + By
A
Az Bz Az + Bz
86
Vector Multiplications
A vector can be multiplied with both scalars, or another vector. However, there are two kinds
of vector multiplications in the case of when two vectors are being multiplied. These are the
dot product and the cross product. The result of the dot product between two vectors give a
scalar result, and hence it is also called the scalar product of two vectors. On the other hand,
the cross product gives a vector-valued result, and hence is called the vector product.
⃗ = ⟨ A x , Ay , Az ⟩ and ⃗B =
The general format for finding the dot product of two vectors A
⟨ Bx , By , Bz ⟩ is,
Ax Bx
⃗ ⃗
A · B = Ay · By = A x Bx + Ay By + Az Bz
Az Bz
For n dimensions,
n
⃗ · ⃗B =
A ∑ Ai Bi
i =1
The dot product of two vectors is the product of the length projection of one vector on the other
with the modulus of the other.
⃗ · ⃗B = | A
A ⃗ ||⃗B| cos θ
87
The Cross Product:
The cross product of two vectors is found by arranging their values in a 3 × 3 matrix, and
⃗ = ⟨ A x , Ay , Az ⟩ and ⃗B = ⟨ Bx , By , Bz ⟩, the cross product
finding its modulus. For the vectors A
is,
Ax Bx x̂ ŷ ẑ Ay Bz − Az By
⃗ × ⃗B = Ay × By = A x Ay Az = Az Bx − A x Bz
A
Az Bz Bx By Bz A x By − Ay Bx
⃗ × ⃗B = | A
A ⃗ ||⃗B| sin θ n̂
88
16.3 Vector Equation of Straight Line
In three dimensions, the equation of a straight line can be written in the form, ⃗r = ⃗p + λ⃗q.
px qx p x + λq x
⃗r = py + λ qy = py + λqy
pz qz pz + λqz
Here, ⃗p is a position vector in R3 lying on the straight line. ⃗q is a direction vector which points
to any direction along the straight line. Of course, it must be parallel to the line. The scalar
λ scales the direction vector without ever changing its direction. Hence, this is what draws
the straight line, starting from the position vector ⃗p to either sides of the line directed by the
direction vector ⃗q.
The lines can be parallel if and only if their direction vectors are the same, or if they are just
the negative of each other. So, for the two lines above, the condition that must be met is,
⃗b = ±d⃗
This is because, the negative can be taken out and included with the scalar.
For two lines to be perpendicular, the dot product of their direction vectors must be zero.
Consider the lines,
ax bx
l1 : ⃗r = ⃗a + λ⃗b = ay + λ by
az bz
cx dx
l2 : ⃗r = ⃗c + µd⃗ = cy + µ dy
cz dz
89
To understand this, recall the definition of the dot product. Dot product of two vectors ⃗b and d⃗
is given by,
⃗b · d⃗ = |⃗b||d⃗| cos θ
If the vectors are perpendicular, i.e., the angle between the vectors is to be θ = 90◦ , then the
dot product becomes,
⃗b · d⃗ = |⃗b||d⃗| cos 90 = 0
90
16.4 Vector Calculus: Differentiation
The Nabla Operator
Gradient (grad)
Gradient of a scalar function defines the slope of the function in each direction. The gradient
operator works only on scalar fields, but it is included in vector calculus because the result is
a vector-valued solution.
We can see that it is takes the derivative with respect to each coordinates, and thus, when it acts
on a scalar field ψ( x, y, z), it outputs the gradient of the scalar field in each respective direction,
giving us ∇ψ, which is a vector-valued function.
∂ψ
∂x
∇ψ( x, y, z) = ∂ψ
∂y
∂ψ
∂z
For this reason, the action of the nabla on a scalar function is also known as finding the "direc-
tional derivative" of the scalar function.
Divergence (div)
The divergence of a vector field measures the density of change in the strength of the vector
field. In other words, the divergence measures the instantaneous rate of change in the strength
of the vector field along the direction of flow.
Ax
In mathematical terms, the divergence of a vector field A⃗ ( x, y, z) = Ay is its dot product
Az
with the nabla operator. ∂A
∂ x
∂x Ax ∂x
⃗ = ∂ ∂Ay
∇·A ∂y · Ay = ∂y
∂ Az ∂Az
∂z ∂z
NOTE: The divergence of a vector field is a scalar field because of the nature of the dot product.
The curl is a vector operator that describes the infinitesimal circulation of a vector field in
three-dimensional Euclidean space. The curl at a point in the field is represented by a vector
whose length and direction denote the magnitude and axis of the maximum circulation. The
91
curl of a field is formally defined as the circulation density at each point of the field.
Ax
⃗ ( x, y, z) = Ay is its cross product
In mathematical terms, the divergence of a vector field A
Az
with the nabla operator.
∂A ∂A
∂
x̂ ŷ ẑ z
− ∂zy
Ax ∂y
⃗ = ∂x
∂ ∂ ∂ ∂ ∂Ax
∇×A × Ay = ∂x = − ∂A z
∂y ∂y ∂z ∂z ∂x
∂ Az ∂Ay
∂z
Ax Ay Az
∂x − ∂A
∂y
x
NOTE: The curl of a vector field is also a vector field because of the nature of the cross product.
The nabla operator was a first order directional derivative operator. For the second order
directional derivative, we use the Laplacian operator. It is defined as the square of the nabla
operator as follows.
∂2
∂x2
2 ∂2
∆ = ∇ = ∂y 2
∂2
∂z2
92
16.5 Vector Calculus: Integration
93
① ∂M
16.6 Green’s Theorem
∂L
I
− dS = ( Ldx + Mdy)
∂x ∂y C
D
④⃗ ④ ②
16.8 Gauss’s Divergence Theorem
A · d⃗S = ⃗ · n̂)dS =
(A ⃗ )dV
(∇ · A
S S V
94
17 Ordinary Differential Equations
Differential equations arise whenever it is easier to describe change than the absolute amounts.
For example, it is easier to say why population size grow or shrink than it is to describe why
they have the particular values that they do at a given point in time.
Differential equations come in two flavors: Ordinary Differential Equations (ODE) and Partial
Differential Equations (PDE). Ordinary differential equations deal with functions with a single
input, often thought of as time, while partial differential equations deal with multiple inputs,
and can be thought of as a whole continuum of values changing with time, for example, the
temperature at every point in a solid body, or the velocity of a fluid at every point in space.
dy
= f ( x ) g(y)
dx
Then, its general function can be found by the following integration.
1
Z Z
dy = f ( x )dx + C
g(y)
Here, C is the integration constant to be found in order to determine the particular solution.
Non-Separable ODEs
The method for finding the general solution for these differential equations is based on the
product rule of differentiation.
dv du d
u +v = (uv)
dx dx dx
where, u and v are functions of x.
dy
u + u̇y = f ( x )
dx
du
where u̇ = dx , then the general solution is derived by,
d
(uy) = f ( x )
dx
Z
uy = f ( x )dx + C
Z
1
y= × f ( x )dx + C
u
95
Assume that a differential equation is such that,
dy
a + by = c
dx
dy b c
+ =
dx a a
where a, b and c are functions of x.
b
Assuming that P = a and Q = ac , then the equation becomes,
dy
+ Py = Q
dx
If a differential equation has this format, we need to determine an integrating factor to solve
for the general solution.
R
Integrating factor: e Pdx
Pdx dy
R R R
Pdx Pdx
e +e Py = e Q
dx
Now, the differential equation becomes such that the first derivative of the coefficient of y
dy
becomes the first derivative of the coefficient of dx .
R R
d
i.e. dx e Pdx = Pe Pdx
d R Pdx
R
Pdx
e y = Qe
dx
R Z R
Pdx Pdx
e y= Qe dx + C
R
Z R
− Pdx Pdx
y=e × Qe dx + C
Bernoulli ODEs
dy
+ P( x )y = Q( x )yn
dx
Here, n > 1, and nR. Note, that when n = 0, 1, the ODE is simply separable/non-separable as
discussed before.
96
Now, let us return to the Bernoulli ODE. If we multiply y−n on both sides, we get,
dy
y−n + P ( x ) y 1− n = Q ( x )
dx
It would be much more convenient to transform this differential equation in terms of a new
variable v, such that v( x ) = y1−n . But for that, we need to manipulate this equation a bit
dy
more. The term y−n dx can be put in the form of the chain rule of differentiation, by simply
multiplying (1 − n) on both sides of the equation.
dy
(1 − n ) y − n + ( 1 − n ) P ( x ) y 1− n = ( 1 − n ) Q ( x )
dx
d 1− n
( y ) + (1 − n ) P ( x ) y 1− n = ( 1 − n ) Q ( x )
dx
dv
+ (1 − n ) P ( x ) v = (1 − n ) Q ( x )
dx
This is a linear first order non-homogeneous differential equation, which we can easily now
solve using the technique we solved non-separable ODEs. But still, the general formula for the
solution can be found as follows.
R
The integrating factor of the found ODE is e(1−n) P( x )dx . Multiplying this on both sides of the
ODE, we get,
P( x )dx dv
R R R
e (1− n ) + ( 1 − n ) e (1− n ) P( x )dx
P ( x ) v = ( 1 − n ) e (1− n ) P( x )dx
Q( x )
dx
d (1− n ) R P( x )dx
R
e v = ( 1 − n ) e (1− n ) P( x )dx
Q( x )
dx
R Z R
(1− n ) P( x )dx
e v = (1 − n ) e (1− n ) P( x )dx
Q( x ) dx
R Z R
v = ( 1 − n ) e ( n −1) P( x )dx
e (1− n ) P( x )dx
Q( x ) dx
R Z R 1−1 n
( n −1) P( x )dx (1− n ) P( x )dx
y = (1 − n ) e e Q( x ) dx
R
Z R 1−1 n
( n −1) P( x )dx (1− n ) P( x )dx
∴ y = (1 − n ) e e Q( x ) dx + C
97
Exact ODEs
dy M( x, y)
=−
dx N ( x, y)
→ M( x, y)dx + N ( x, y)dy = 0
∂M ∂N
If this ODE is such that ∂y = ∂x , then the ODE is said to be an exact differential equation.
We can either integrate M with respect to x or integrate N with respect to y. Both the ap-
proaches are shown below.
h( x ) is the integration constant (constant in y, but can very well be a function of x). Now, you
can see that if we partially differential the function Ψ with respect to x, we get M.
Z
∂
Ψ x ( x, y) = N ( x, y)dy + h( x ) = M( x, y)
∂x
Z
∂
N ( x, y)dy + h′ ( x ) = M( x, y)
∂x
Finally, the implicit solution to the differential equation becomes Ψ( x, y) = C, which is,
Z Z
∂
Z
N ( x, y)dy + M( x, y) − N ( x, y)dy dx = C
∂x
98
Here, C is a constant.
h(y) is the integration constant (constant in x, but can very well be a function of y). Now, you
can see that if we partially differential the function Ψ with respect to y, we get N.
Z
∂
Ψy ( x, y) = M ( x, y)dx + h(y) = N ( x, y)
∂y
Z
∂
M ( x, y)dx + h′ (y) = N ( x, y)
∂y
Finally, the implicit solution to the differential equation becomes Ψ( x, y) = C, which is,
Z Z
∂
Z
M( x, y)dx + N ( x, y) − M( x, y)dx dy = C
∂y
Here, C is a constant.
Although the notations differ in the two answers in appearnace, they are the same. Both ap-
proaches will lead to the same general solution.
99
17.2 Second Order Ordinary Differential Equations
The general solution of the linear second order differential equation,
d2 y dy
a 2
+ b + cy = f ( x )
dx dx
where a, b and c are constants, is found using the equation,
Complementary Function
To solve a differential equation of the format shown above, we need to break it down. To find
the complementary function, we need to solve the equation assuming f ( x ) = 0.
d2 y dy
a 2
+ b + cy = 0
dx dx
As the equation above resembles a quadratic equation, we can write it in a simplified manner.
am2 + bm + c = 0
This equation is known as the auxiliary equation. The roots of this equation when we solve for
m determines the complementary function of the differential equation. The table below shows
the format of the complementary functions for the nature of roots.
100
Particular Integral
d2 y dy
a + b + cy = 0
dx2 dx
To find the particular integral of a differential equation, a suitable form of particular integral is
taken. Let us assume the equation to be y = g( x ), where g( x ) is a form of particular integral.
ẏ = g′ ( x )
ÿ = g′′ ( x )
d2 y dy
a 2
+ b + cy = f ( x )
dx dx
where a, b and c are known constants.
The left hand side of the equation is rearranged as such that it is of the same form as f ( x ). After
that, the unknown constants (λ, µ, γ and/or p) can be found out by comparing the coefficients
of both sides of the equation. The now known constants are put in the assumed equation,
y = g( x )
This equation, hence, becomes the particular integral of the differential equation.
101
17.3 Clairaut Differential Equation
Clairaut’s differential equation is an ODE in the form,
y = xy′ + f (y′ )
From this, we can see that y′′ = 0. So, y′ = C is a constant. Therefore, we can now write
Clairaut’s ODE as,
y = Cx + f (C )
102
18 Partial Differential Equations
103
19 Approximation Using Infinite Series
Under certain conditions, a wide range of functions of x can be expressed as an infinite se-
ries in ascending powers of x; these are often referred to as power series. Integrals of many
2
common functions, e.g. e−kx , familiar to students of statistics, cannot be expressed in terms of
elementary functions, but approximations to any required degree of accuracy can be found.
f : x → f ( x ), { x ∈ R }
The providing that f (0), f ′ (0), f ′′ (0), f ′′′ (0), ... , f (r) (0) all have finite values,
The plot above shows the Maclaurin approximations of the standard cosine function f ( x ) =
cos x. The true cosine function can be compared with the different approximation curves plot-
ted by taking different numbers of terms (r ) into the calculation.
104
19.2 Taylor Series
We can find an approximation to a function of x close to x = a, where a ̸= 0, using Taylor’s
expansion of the function.
The construction of the Maclaurin expansion focuses on x = 0 and for a value of x very close
to 0, a few terms of the series may well give a good approximation of the function. For values
of x further away from 0, even if they are in the interval of validity, more and more terms of
series are required to give a good degree of accuracy. To overcome these problems, a series
expansion focusing on x = a can be derived.
f ( x + a) and f ( a) are the two forms of Taylor’s expansion or series. When a = 0, they both be-
come the Maclaurin expansion. So, we can say that Maclaurin series is a special case of Taylor
series.
Therefore, we can say that a = −φ if we want to compare it with our derived formula.
f ′′ (−φ) f ′′′ (−φ) f (n) (−φ)
f ( x ) = f (−φ) + f ′ (−φ)( x + φ) + ( x + φ )2 + ( x + φ)3 + ... + ( x + φ)n
2! 3! n!
The Taylor series formula can be written as the sigma notation as follows.
!
∞
f (r ) ( x0 ) r
f (x) = ∑ ( x − x0 )
r →0 r!
Note that the Taylor series formula becomes the Maclaurin series when x0 = 0. The Maclaurin
series is a special case of the Taylor series expansion.
105
19.3 Laurent’s Series
∞
∑ ( ar ( z − z0 )r )
r =−∞
Here, ar is,
1 f (z)
Z
ar = dz
2πi c ( z − z 0 ) n +1
106
19.4 Common Series Expansions of Functions
Function Series When is Valid/True
1 + x + x2 + x3 + x4 + ...
1
1− x x ∈ (−1, 1)
∑∞
n =0 xn
x2 x3 x4
1+x+ 2! + 3! + 4! + ...
ex x∈R
xn
∑∞
n=0 n!
x2 x4 x6 x8
1− 2! + 4! − 6! + 8! − ...
cos x x∈R
n x2n
∑∞
n=0 (−1) (2n)!
x3 x5 x7 x9
x− 3! + 5! − 7! + 9! − ...
(2n−1)
sin x ∑∞
n=1 (−1)
( n −1) x
(2n−1)!
x∈R
(2n+1)
∑∞ n x
n=0 (−1) (2n+1)!
x2 x3 x4 x5
x− 2 + 3! − 4! + 5! − ...
n
ln 1 + x ∑∞
n=1 (−1)
( n −1) x
n x ∈ (−1, 1]
n
∑∞
n=1 (−1)
( n +1) x
n
x3 x5 x7 x9
x− 3 + 5! − 7! + 9! − ...
(2n−1)
arctan x ∑∞
n=1 (−1)
( n −1) x
(2n−1)
x ∈ [−1, 1]
(2n+1)
∑∞ nx
n=1 (−1) (2n+1)
107
20 Fourier Analysis
20.1 Fourier Series
Almost any periodic function can be represented by the sum of a series of sine and cosine
functions. The formula for the series is,
∞
1 2πk 2πk
f ( x ) = a0 + ∑ ak cos x + bk sin x
2 k =1
L L
Here, a0 , ak and bk are Fourier coefficients. These are found using the integrals,
RL
a0 = − L f ( x )dx
RL
ak = − L cos 2πk x dx
L
L
2πk
b =
R
k − L sin L x dx
108
20.2 Complex Fourier Series
Consider a Fourier series in the domain [−π, π ] that has the general form,
∞
1
f (x) = a0 + ∑ ( ak cos (kx ) + bk sin (kx ))
2 k =1
We can use the following formulae to represent the sine and cosine terms as exponential func-
tions. (
cos (kx ) = 21 (eikx + e−ikx )
sin (kx ) = 2i1 (eikx − e−ikx ) = − 2i (eikx − e−ikx )
If these are inserted into the original Fourier series equation, we get,
∞
1
f (x) = a0 + ∑ ( ak cos (kx ) + bk sin (kx ))
2 k =1
∞
1 1 ikx −ikx i ikx −ikx
= a0 + ∑ a k (e + e ) + bk − ( e − e )
2 k =1
2 2
∞
ak − ibk ak + ibk
1 ikx −ikx
= a0 + ∑ e + e
2 k =1
2 2
∞ ∞
ak − ibk ak + ibk
1 ikx −ikx
= a0 + ∑ e +∑ e
2 k =1
2 k =1
2
Now, let c0 = 12 a0 and ck = ak −2ibk . Also, for the last second summation term, change k to −k.
So, we can write, the last summation term as,
−1
a−k + ib−k
ikx
∑ 2
e
k =−∞
a−k +ib−k
Here, we can see that the term 2 is equivalent to ck . So, we can write the Fourier series
as,
∞ −1
f ( x ) = c0 + ∑ ck eikx + ∑ ck eikx
k =1 k=−∞
∞
∴ f (x) = ∑ ck eikx
k =−∞
Now, we need to find the constants ck . For this, let us multiply both sides by e−imx , where
m ∈ Z is a constant.
∞
f ( x )e−imx = ∑ ck eikx e−imx
k=−∞
We integrate both side with respect to x in the entire domain of [−π, π ], we get,
Z π ∞ Z π
f ( x )e−imx dx = ∑ ck eikx e−imx dx
−π k =−∞ −π
109
(
Rπ 0 {k ̸= m}
We can see that the integral −π eikx e−imx dx = . For our purposes, n = m. So,
2π {k = m}
we get,
Z π ∞ Z π
f ( x )e−imx dx = ∑ ck eikx e−imx dx = 2πcn = 2πcm
−π k =−∞ −π
Z π
2πcm = f ( x )e−imx dx
−π
Z π
1
cm = f ( x )e−imx dx
2π −π
As k = m, we can write, Z π
1
ck = f ( x )e−ikx dx
2π −π
where, Z π
1
ck = f ( x )e−ikx dx
2π −π
110
20.3 Fourier Transform
Fourier Transform and the Inverse Fourier Transform
Relation between the Modulus of a Function and the Modulus of its Fourier Transform
R∞
Now, consider the integral −∞ | f ( x )|2 dx. We know that | f ( x )|2 = f ( x ) f ∗ ( x ). So, we can write,
Z ∞ Z ∞
2
| f ( x )| dx = f ( x ) f ∗ ( x )dx
−∞ −∞
R∞ ′
From the inverse Fourier relation, f ∗ ( x ) = √12π −∞ fˆ(k′ )eik x dk′ . Now, the equation above
becomes,
Z ∞ Z ∞ Z ∞
1 1 ′x
2
| f ( x )| dx = √ fˆ(k )e dk
ikx
√ fˆ(k )e dk
′ ik ′
−∞ 2π −∞ 2π −∞
Z ∞ Z ∞ Z ∞
ˆ∗ ˆ ′ 1 − x (k′ −k)
= f (k) f (k ) e dx dk′ dk
−∞ −∞ 2π −∞
R ∞ −ixk′ −k
We can see from the definition of the delta function that δ(k′ − k ) = 2π 1
−∞ e dx. So, we
can write, Z ∞ Z ∞ Z ∞
2
| f ( x )| dx = ˆ∗
f (k) ˆ ′ ′ ′
f (k )δ(k − k )dk dk
−∞ −∞ −∞
111
21 The Dirac Delta Function
The delta function is a not a "true function". We call it a function out of convenience, but to
mathematicians, this is a "generalized function" or a "distribution".
Consider a Gaussian distribution with a mean of x0 and a variance of σ. Then, we can write
the probability density function as,
1 x − x0 2
e− 2 ( )
1
G ( x − x0 ) = √ σ
2πσ
The Dirac delta function can be said to be a Gaussian distribution such that the width tends to
zero (σ → 0).
1 x−x
− 12 ( σ 0 )
2
δ( x − x0 ) = lim √ e
σ →0 2πσ
In other words, the Dirac delta function is generally written as the non-continuos function as
follows. (
∞ { x = x0 }
δ ( x − x0 ) =
0 { x ̸ = x0 }
The plots above shows the Gaussian distribution for different σ values. It can be seen that as
σ → 0, the graph resembles the delta function.
112
Dirac Delta Function as the Derivative of the Heaviside Step Function
The Heaviside step function (also known as a "jump" fuction) H ( x ) is a function defined as,
(
0 { x < 0}
H (x) =
1 { x ≥ 0}
We can see here that the graph of H ( x ) resembles a single step of a stair, from which it gets its
name.
Now, let us see what the derivative of the step function. For finding the derivative, we can
divide the function’s domain into three parts, and then find the derivatives at the respective
parts.
1. When t < x0 , the expression for the step function is a constant (0), and so the derivative
in this interval is zero.
2. When t > x0 , the expression for the step function is a constant (1), and so the derivative
in this interval is zero.
3. When t = x0 , the function jumps from 0 to 1 instantaneously, which implies an infinite
gradient.
113
From these, we find that the derivative of the Heaviside step function is 0 everywhere, but at
the point x0 . So, we can write,
0
{ x < x0 }
′
H ( x − x0 ) = ∞ { x = x0 }
0 { x > x0 }
114
21.2 Integrals Involving the Dirac Delta
For the integration of the Dirac delta, first recall what an integral implies. A definite integral
of a function is the area under the curve within a certain domain. Also recall that the delta
function is just a standard Gaussian distribution whose variance tends to zero. So, let us begin
by integrating the standard Gaussian distribution first.
Let Q be the integral of the standard Gaussian distribution function in an infinite domain. So,
Z ∞
1 − 21 (
x − x0
)
2
Q= √ e σ dx
−∞ 2πσ
We solve this integral using substitution to simplify the problem. Let u = x−σx0 . So, the differ-
ential is du = σ1 dx, which implies dx = σdu. The domain in u is still (−∞, ∞) according to the
definition of u. With this substitution, the integral simplifies into,
Z ∞
1 − 12 u2
Q= √ e σdu
−∞ 2πσ
Z ∞
1 1 2
=√ e− 2 u du
2π −∞
It turns out that it is easier to find the integral of Q2 instead of Q. So, that is what we are going
to do.
1
Z ∞ 2
2 − 12 u2
Q = √ e du
2π −∞
1
Z ∞ 2
− 21 u2
= e du
2π −∞
Remember that a2 = a × b if a = b. Using this property, we can write the integral as,
Z ∞ Z ∞ 1 2
1 − 21 u2
2
Q = e du e− 2 v dv
2π −∞ −∞
Here, u and v are serving the same purpose. We can further simplify this by writing this as a
double integral as follows.
Z ∞ Z ∞
1 1 2 1 2
Q2 = e− 2 u e− 2 v du dv
2π −∞ −∞
Z ∞ Z ∞
1 1 2 2
= e− 2 (u +v ) du dv
2π −∞ −∞
115
Here, u and v are the bases for the Cartesian coordinate system, which we have been working
on till now. To solve the integral, it is easier to do it in polar coordinates. So, the orthogonal u
and v coordinates become the corresponding r and θ polar coordinates using the relations,
(
u = r cos θ
v = r sin θ
Here, the extent of r is [0, ∞) and the extent of θ is [0, 2π ]. With this, the integral in polar
coordinate system becomes,
Z 2π Z ∞
1 1 2
Q2 = e− 2 r r dr dθ
2π 0 0
The extra r is just the Jacobian of the transformation. We now perform another substituion
using ω = r2 . So, the differential dω = 2rdr and rdr = 21 dω. the extent of omega is also [0, ∞).
Now, the integral becomes,
Z 2π Z ∞
1 1
1
Q2 = e− 2 ω dω dθ
2π 0 0 2
1 2π
Z Z∞ 1
2
Q = e− 2 ω dω dθ
4π 0 0
1 ∞ − 1 ω
Z
= e 2 dω
2 0
1h 1
i∞
= −2e− 2 ω
2 0
= −(−1)
=1
∴Q=1
This shows that the integral of the standard Gaussian distribution function over in infinite
domain is 1. Z ∞
1 1 x − x0 2
√ e− 2 ( σ ) dx = 1
−∞ 2πσ
116
Integral of the Dirac Delta in Infinite Domain
This just implies the area of bounded by the delta function. Now, instead of the delta function
inside the integral, we can write the expression of the Gaussian distribution.
Z ∞ ( x − x0 )2
1 −
lim √ e 2σ2 dx
σ →0 − ∞ 2πσ
The standard Gaussian distribution is a probability distribution, so the area under the curve
between two points tells you the probability of variables taking on a range of values. The total
area under the curve is 1 or 100% for any σ value. This was derived previously. Therefore, the
integral of the Dirac delta across infinite domain has a value of,
Z ∞
δ( x )dx = 1
−∞
We saw that the delta function is just a standard Gaussian distribution whose σ → 0. As the
Gaussian distribution is a probability distribution, the area integral of it has a value of 1. But
that is for an infinite domain, of course. The Dirac delta, however, has discrete values as given
by, (
∞ { x = x0 }
δ ( x − x0 ) =
0 { x ̸ = x0 }
Now, if the domain is finite such that the domain is [ x1 , x2 ], we can propose an integral,
Z x2
δ( x − x0 )dx
x1
This integral implies the area under the distribution curve within that domain. Howver, since
the distribution is a delta function, we can use its defnition to deduce the solution.
The delta function δ( x − x0 ) is such that it has a non-zero value only when x = x0 . Therefore,
the integral of the delta function in a domain which does not include x0 must be zero. With the
same reason, if the integral domain includes the point x0 , the integral will have value of 1.
Z x2
(
1 { x0 ∈ [ x1 , x2 ]}
δ( x − x0 )dx =
x1 0 { x0 ∈
/ [ x1 , x2 ]}
117
21.3 Properties of the Dirac Delta Function
Translation Property
Here, the function f ( x ) is multiplied by a Dirac delta, which is at a point x0 . Therefore, the
function f ( x )δ( x − x0 ) is zero everywhere, but when x = x0 , the function has a value of f ( x0 ).
So, we can write, Z ∞
f ( x0 ) δ( x − x0 )dx
−∞
R∞
By the definition of the integral of the Dirac delta in infinite domain, we know −∞ δ( x − x0 )dx =
1. So, the integral of a function multiplied by the Dirac delta is,
Z ∞
f ( x )δ( x − x0 )dx = f ( x0 )
−∞
Recall the translation property, which states that for any continuous function f ( x ), it holds
that, Z ∞
f (x) = f ( x0 )δ( x − x0 )dx0
−∞
f ( x0 ) δ ( x − x0 ) = f ( x0 ) δ ( x0 − x )
∴ δ ( x − x0 ) = δ ( x0 − x )
118
Scaling Property
Let the parameter y be such that y = a( x − x0 ) = ax − ax0 , where a > 0 is constant. So, the
y
parameter x can be written as x = a + x0 and the differential of it becomes dx = 1a dy.
R∞
Recall the integral, −∞ f ( x )δ( x − x0 )dx. For our case, this becomes,
Z ∞ Z ∞ y
1
f ( x )δ( a( x − x0 ))dx = f + x0 δ ( y ) dy
−∞ −∞ a a
Z ∞ y
1
= f + x0 δ(y)dy
a −∞ a
R∞
According to the formula −∞ f ( x )δ( x − x0 )dx = f ( x0 ), we get,
1
= f ( x0 )
a
Now, let the parameter y be such that y = − a( x − x0 ) = − ax + ax0 , where a > 0 is constant. So,
y
the parameter x can be written as x = − a + x0 and the differential of it becomes dx = − 1a dy.
The integral thus becomes,
Z ∞ Z −∞ y
1
f ( x )δ(− a( x − x0 ))dx = f − + x0 δ(y) − dy
−∞ ∞ a a
Z −∞
1 y
= − f − + x0 δ(y)dy
a ∞ a
1 ∞ y
Z
= f − + x0 δ(y)dy
a −∞ a
1
= f ( x0 )
a
1
∴ f ( x0 )
| a|
(
1
Z ∞
| a|
f ( x0 ) if a < 0
f ( x )δ( a( x − x0 ))dx = 1
a f ( x0 ) if a > 0
−∞
119
Thus, in general, Z ∞
1
f ( x )δ( a( x − x0 ))dx = f ( x0 )
−∞ | a|
for all a.
R∞
Since f ( x0 ) = −∞ f ( x )δ( x − x0 )dx, based on the translation property, we have,
Z ∞ Z ∞
1
f ( x )δ( a( x − x0 ))dx = f ( x )δ( x − x0 )dx =
−∞ | a| −∞
1
f ( x )δ( a( x − x0 )) = f ( x ) δ ( x − x0 )
| a|
1
∴ δ( a( x − x0 )) = δ ( x − x0 )
| a|
Proof of xδ( x ) = 0
There are two approaches to prove this, and both is shown below.
(
∞ if x = 0
The first is this. The Dirac delta is such that δ( x ) = . Therefore, it has a non-
0 if x ̸= 0
zero value only at x = 0. However, multiplying the δ( x ) with x nullifies any possibilities of a
non-zero result, because when x = 0, the x term becomes zero and when x ̸= 0, the δ( x ) term
becomes zero. So, ∀ x ∈ R,
xδ( x ) = 0
This was an intuitive explanation. For the second approach, we use the integral formula below.
Z ∞ Z ∞
f ( x )( xδ( x ))dx = ( x f ( x ))δ( x )dx
−∞ −∞
Z ∞
g( x )δ( x )dx = g(0)
−∞
= f (0) × 0
=0
120
21.4 The Dirac Comb: Fourier Series Representation of the Dirac Delta Function
For any function f ( x ), its Fourier series is given by,
∞ nπ nπ
1
f (x) = a0 + ∑ an cos x + bn sin x
2 n =1
L L
Here, RL
1
a0 =
L R−LL
f ( x )dx
1 nπ
an = f ( x ) cos L x dx
L
1
R −LL nπ
bn = − L f ( x ) sin L x dx
L
When the function f ( x ) is a delta function δ( x − x0 ) which is in the interval [− L, L] (such that
{− L < x < L}), and that the function is periodic, the function can be expressed as a Fourier
series. For this series, the expressions if a0 , an and bn are calculated below.
For a0 ,
Z L
1
a0 = δ( x − x0 )dx
L −L
1
=
L
For an ,
1 L nπ
Z
an = δ( x − x0 ) cos x dx
L −L L
1 nπ
= cos x0
L L
For bn ,
1 L nπ
Z
bn = δ( x − x0 ) sin x dx
L −L L
1 nπ
= sin x0
L L
Putting these in the general Fourier series formula, we get the Fourier series of the periodic
delta function to be,
121
∞ nπ nπ 1 nπ nπ
1 1
δ ( x − x0 ) = +∑ cos x0 cos x + sin x0 sin x
2L n=1 L L L L L L
1 1 ∞ nπ nπ nπ nπ
= + ∑ cos x0 cos x + sin x0 sin x
2L L n=1 L L L L
Using the trigonometric formula cos A cos B − sin A sin B = cos ( A − B), we can write,
1 1 ∞ nπ nπ
2L L n∑
δ ( x − x0 ) = + cos x0 − x
=1 L L
1 1 ∞ nπ
= + ∑ cos ( x0 − x )
2L L n=1 L
This is the general Fourier series expression for a Dirac delta function δ( x − x0 ). The following
graph shows the plot when L = 1 and x0 = 0 for different numbers of terms n taken into
consideration.
Due to the characteristic shape of the plot, the periodic Dirac delta function is known as the
Dirac comb.
122
22 The Kronecker Delta
The Kronecker delta is a function of two variables, usually just non-negative integers. The
function is 1 if the variables are equal, and 0 otherwise.
(
1 {i = j }
δij =
0 {i ̸ = j }
Mathematicians use the Kronecker delta function to convey in a single equation what might
otherwise take several lines of text.
1 0 0 ··· 0
0 1 0 · · · 0
δij = 0 0 1 · · · 0
.. .. .. . . ..
. . . . .
0 0 0 ··· 1
123
23 Symbols
124