0% found this document useful (0 votes)
150 views297 pages

Mat5101 Am Lecturenotes

The document provides an overview of topics in linear algebra, including matrices and linear systems. It defines matrices and vectors, describes how to add and multiply matrices, and explains that matrix multiplication is only defined if the inner dimensions are equal. It also introduces the concept of solving systems of linear equations using Gauss elimination.

Uploaded by

Mümin Şen
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
150 views297 pages

Mat5101 Am Lecturenotes

The document provides an overview of topics in linear algebra, including matrices and linear systems. It defines matrices and vectors, describes how to add and multiply matrices, and explains that matrix multiplication is only defined if the inner dimensions are equal. It also introduces the concept of solving systems of linear equations using Gauss elimination.

Uploaded by

Mümin Şen
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 297

MAT5101

Applied Mathematics

Department of Mathematics,
Dokuz Eylül University,
35160 Buca, İzmir, Turkey.

Tuesday 15th May, 2018


14:40

Applied Mathematics 1/297


Contents I

1 Linear Algebra
Matrices and Linear Systems
Linear Systems
Gauss Elimination

Vector Spaces

Determinants
Cramer’s Rule

Real Inner Product Spaces


Linear Transformations

Eigenvalues, Eigenvectors and Eigenspaces


Diagonalization

Applied Mathematics 2/297


Contents II

2 Ordinary Differential Equations


First-Order Linear Differential Equations
Homogeneous Equation
Non-Homogeneous Equation
Second-Order Linear Differential Equations with Constant
Coefficients
Homogeneous Equation
Non-Homogeneous Equation

Initial Value Problems and Boundary Value Problems


Initial Value Problems
Boundary Value Problems

Sturm-Liouville Problems
Fourier Series

Applied Mathematics 3/297


Contents III

3 Partial Differential Equations


Heat Equation

Wave Equation

Laplace Equation

Fourier Integral
The Fourier Sine and Cosine Integrals
Solving PDEs by Fourier Integrals

Applied Mathematics 4/297


Linear Algebra Matrices and Linear Systems

Linear Algebra

Matrices and Linear Systems


• Roughly speaking, a matrix is a rectangular array.
• We shall discuss existence and uniqueness of solutions for systems of
linear equations.
• The method of Gauss elimination will be introduced to solve systems.

Applied Mathematics 5/297


Linear Algebra Matrices and Linear Systems

We will denote by capital bold letters such as A, B, · · · , or by writing the


general entry in the brackets such as A := (aij ), and so on. By an m × n
matrix (read as m by n), we mean a matrix with m rows and n columns.
It should be noted that rows come always first! m × n is called the size
of the matrix. Explicitly, an m × n matrix is of the form
 
a11 a12 · · · a1n
 a21 a22 · · · a2n 
A :=  . ..  . (1)
 
.. ..
 .. . . . 
am1 am2 · · · amn

Applied Mathematics 6/297


Linear Algebra Matrices and Linear Systems

Each entry in (1) has two subscripts. The first one stands for the row
number and the second one stands for the column number. That is, a21
is the entry in the second row and first column.
If m = n, then the matrix A will be called as a square matrix. In this
case, a11 , a22 , · · · , ann are called as the main diagonal entries of the
matrix A.

Applied Mathematics 7/297


Linear Algebra Matrices and Linear Systems

If either m = 1 or n = 1, then a matrix is called as a vector. We will


denote vectors by lowercase bold letters such as a, b, · · · or by their
general terms in brackets such as a = (ai ), and so on. If m = 1, then the
vector is called a row vector, i.e.,

a := a1 a2 · · · an

while if n = 1, then the vector is called a column vector, i.e.,


 
a1
 a2 
a :=  .  .
 
 .. 
am

Applied Mathematics 8/297


Linear Algebra Matrices and Linear Systems

Definition 1 (Equality of Matrices)


Two matrices A = (aij ) and B = (bij ) are said to be equal (and written
as A = B) if and only if they have same size and the corresponding
entries are equal, i.e., a11 = b11 , a12 = b12 , and so on. Matrices which
are not equal are called as different. Thus, matrices of different sizes are
always different.

Applied Mathematics 9/297


Linear Algebra Matrices and Linear Systems

Definition 2 (Addition of Matrices)


The sum of two matrices A = (aij ) and B = (bij ) of the same sizes is
written as A + B and is the matrix A + B = (aij + bij ), i.e., the matrix
obtained by adding the corresponding entries of A and B. Matrices of
different sizes cannot be added.

Example 1
Readily, we compute that
     
0 1 2 6 5 4 0+6 1+5 2+4
+ =
9 8 7 3 4 5 9+3 8+4 7+5
 
6 6 6
= .
12 12 12

Applied Mathematics 10/297


Linear Algebra Matrices and Linear Systems

Definition 3 (Scalar Multiplication)


The product of any m × n matrix A = (aij ) by any scalar α is written as
αA and is the m × n matrix αA = (αaij ), i.e., the matrix whose entries
obtained by multiplying each entry of A by α.

Example 2
Obviously, we compute that
   
4 5 6 3·4 3·5 3·6
3 =
7 8 9 3·7 3·8 3·9
 
12 15 18
= .
21 24 27

Applied Mathematics 11/297


Linear Algebra Matrices and Linear Systems

Property 1 (Rules for Matrix Addition and Scalar Multiplication)


From the familiar laws for the addition of numbers, we obtain the
following similar laws for the addition of matrices of the same size m × n.
1. A + B = B + A.
2. (A + B) + C = A + (B + C ), which is written as A + B + C .
3. A + 0 = A.
4. A + (−A) = 0.
Here, 0 denotes the zero matrix (of the size m × n), i.e., the m × n
matrix whose all entries are zero. Similarly, for scalar multiplication, we
have the following rules.
5. α(A + B) = αA + αB.
6. (α + β)A = αA + βA.
7. α(βA) = (αβ)A, which written as αβA.
8. 1A = A.

Applied Mathematics 12/297


Linear Algebra Matrices and Linear Systems

Definition 4 (Matrix Multiplication)

The product C = AB (in this order) of an m × n matrix A = (aij ) times


and p × q matrix B = (bjk ) is defined if and only if n = p and is then the
m × q matrix C = (cik ) whose entries are obtained by
n
X
cik := aij bjk , i = 1, 2, · · · , m, k = 1, 2, · · · , q. (2)
j=1

Applied Mathematics 13/297


Linear Algebra Matrices and Linear Systems

The condition n = p means that the second factor B must have as many
rows as the first factor A has columns. As a diagram of sizes, we can
give the following.
A B = C .
[m×n][n×q] [m×q]

In (2), cik is obtained by multiplying each entry in the i-th row of A by


the corresponding entry in the k-th column of B, and then adding these
n products. For instance, c21 = a21 b11 + · · · + a2n bn1 . One calls this
briefly as “multiplication of rows into columns”.

Applied Mathematics 14/297


Linear Algebra Matrices and Linear Systems

 
b1k · · · b1q
 
 .. . . . 
. . ..
 
 
× 
 bjk · · · bjq


 
.. .
 
..
. ..
 
× 
 . 

bnk · · · bnq
×
  

  
  
 
 ai1 · · · aij · · · ain cik · · · ciq

 
  
 . .  .. .

 . . . ... . . . ...  ..
. ..

 . 
 . 

am1 · · · amj · · · amn cm1 · · · cmk · · · cmq

Figure 1: An illustration of matrix multiplication.

Applied Mathematics 15/297


Linear Algebra Matrices and Linear Systems

Example 3
Using Definition 4, we compute that
  
3 2 2 (−1) 3
4 (−2) 5 3 2
 
3×2+2×5 3 × (−1) + 2 × 3 3×3+2×2
=
4 × 2 + (−2) × 5 4 × (−1) + (−2) × 3 4 × 3 + (−2) × 2
 
16 3 13
= .
(−2) (−10) 8

Note that multiplication in the reverse order is not defined.

Applied Mathematics 16/297


Linear Algebra Matrices and Linear Systems

It should be noted that matrix multiplication is in general not


commutative, i.e., AB 6= BA.
Example 4
    
1 1 (−1) 1 0 0
=
100 100 1 (−1) 0 0
and
    
(−1) 1 1 1 99 99
= .
1 (−1) 100 100 (−99) (−99)

Applied Mathematics 17/297


Linear Algebra Matrices and Linear Systems

Our examples show that order of factors in matrix multiplication must


always be observed very carefully. Otherwise, matrix multiplication
satisfies the following rules similar to those for numbers.

Property 2 (Rules for Matrix Multiplication)


Let α be a scalar while A, B and C be matrices such that the expressions
on the left are defined.
1. (αA)B = α(AB) = A(αB), which is written as αAB or AαB or
ABα.
2. A(BC ) = (AB)C , which is written as ABC .
3. (A + B)C = AC + BC .
4. A(B + C ) = AB + AC .
Second property is called as associate law, third and fourth properties
are called as distributive laws.

Applied Mathematics 18/297


Linear Algebra Matrices and Linear Systems

Definition 5 (Transposition of a Matrix)


The transpose of an m × n matrix A = (aij ) is the n × m matrix AT
(read as A transpose), which has the first row of A as its first column,
second row of A as its second column, and so on. Thus, the transpose of
A in (1) is AT = (aji ), which is written as
 
a11 a21 · · · am1
 a12 a22 · · · am2 
AT :=  . ..  .
 
.. ..
 .. . . . 
a1n a2n · · · amn

As a particular case, transposition converts row vectors to column vectors


and vice versa.

Applied Mathematics 19/297


Linear Algebra Matrices and Linear Systems

Property 3 (Rules for Matrix Transposition)


Let α be a scalar and A, B be matrices such that the expressions on the
left are defined.
T
1. AT = A.
2. (A + B)T = AT + B T .
3. (αA)T = α AT .


4. (AB)T = B T AT .

Applied Mathematics 20/297


Linear Algebra Matrices and Linear Systems

Transposition gives a rise to two useful classes of matrices as follows.

Definition 6 (Symmetric and Skew-Symmetric Matrices)


Let A be a square matrix. A is said to be symmetric matrix if and only
if AT = A while it is said to be skew-symmetric matrix if and only if
AT = −A.

Note that skew-symmetric matrices have zero as their diagonal entries.


Example 5
   
(−2) 3 2 0 (−5) 1
 3 8 (−5)  is symmetric while  5 0 (−3) 
2 (−5) 1 (−1) 3 0
is skew-symmetric.

Applied Mathematics 21/297


Linear Algebra Matrices and Linear Systems

Any real square matrix A may be written as the sum of a symmetric


matrix and a skew-symmetric matrix. More precisely, we write
1  1
A + AT + A − AT .

A=
|2 {z } |2 {z }
symmetric skew-symmetric

Example 6
Consider that
7
(− 32 )
     
3 2 7 3 2 7 0 0
 5 4 3  =  72 4 9
2
+ 3
2 0 (− 23 )  ,
9 3
7 6 4 7 2 4 0 2 0

where the first and the second matrices on the right-hand side are
symmetric and skew-symmetric, respectively.

Applied Mathematics 22/297


Linear Algebra Matrices and Linear Systems

Definition 7 (Upper and Lower Triangular Matrices)


Let A = (aij ) be a square matrix. If aij = 0 for i > j, then A is called as
upper triangular matrix. If aij = 0 for i < j, then A is called as lower
triangular matrix.

   
a11 a12 · · · a1n a11

 a22 · · · a2n 

 a a
 21 22


   
.. .  . .. . .
. ..  ..
  



  . . 

ann an1 an2 · · · ann

Figure 2: Upper and lower triangular matrices, respectively. All the entries in
the red triangles are zero.

Applied Mathematics 23/297


Linear Algebra Matrices and Linear Systems

Definition 8 (Diagonal and Identity Matrices)


Let A = (aij ) be a square matrix. If aij = 0 for i 6= j, then A is called as
diagonal matrix.
In particular, a scalar matrix whose entries on the main diagonal are all 1
is called as the identity matrix (or the unit matrix), and is denoted by
In (or simply I).

   
a11 1
 

 a22 
 
 1 

 
..
 

.
  .. 



  . 
 
ann 1

Figure 3: Diagonal and the identity matrices, respectively. All the entries in the
red triangles are zero.

The diagonal matrix A whose diagonal entries are ak is denoted by


A := diag(ak ).
Applied Mathematics 24/297
Linear Algebra Matrices and Linear Systems

Definition 9 (Orthogonal Matrices)


Let A be a square matrix. A is said to be orthogonal matrix if and only
if AAT = I.

Example 7
 
  2 (−2) 1
1 1 1 1
Show that √ and 1 2 2  are
2 1 (−1) 3
2 1 (−2)
orthogonal matrices.

Applied Mathematics 25/297


Linear Algebra Matrices and Linear Systems

Definition 10 (Linear Systems)


A linear system of m equations and n unknowns x1 , x2 , · · · , xn is the set
of equations of the form

a11 x1 + a12 x2 + · · · + a1n xn =b1


a21 x1 + a22 x2 + · · · + a2n xn =b2
.. (3)
.
am1 x1 + am2 x2 + · · · + amn xn =bm .

The system is called as linear because x1 , x2 , · · · , xn appear in the first


power only, just as in the equation of a straight line. a11 , a12 , · · · , amn are
given numbers called as coefficients of the system. b1 , b2 , · · · , bm on
the right are also given numbers. If b1 , b2 , · · · , bm = 0, then (3) is called
as homogeneous system, otherwise it is called as non-homogeneous
system.

Applied Mathematics 26/297


Linear Algebra Matrices and Linear Systems

From the definition of matrix multiplication, we see that the m equations


of (3) may be written as a single vector equation of the form

Ax = b,

where A, x and b are defined as


     
a11 a12 ··· a1n x1 b1
 a21 a22 ··· a2n   x2   b2 
A :=  .  , x :=   and b :=  .
     
.. .. .. .. ..
 .. . . .   .   . 
am1 am2 ··· amn xn bm

We assume that aij are not all zero, so that A is not a zero matrix. Note
that x has n components and b has m components.

Applied Mathematics 27/297


Linear Algebra Matrices and Linear Systems

The matrix  
a11 a12 ··· a1n b1
 a21 a22 ··· a2n b2 
A
e :=  
 .. .. .. .. .. 
 . . . . . 
am1 am2 ··· amn bm
is called the augmented matrix of the system (3). The vertical line can
be omitted (as we shall do later). It is merely a reminder that the last
column of A e does not belong to A.

Applied Mathematics 28/297


Linear Algebra Matrices and Linear Systems

Although a system can be solved rather quickly by noticing its particular


form, Gauss elimination is a systematic approach and will work in
general, also for larger systems. Let us illustrate this method on the
following example.
Example 8

Consider the system

x1 − x2 + x3 =0
−x1 + x2 − x3 =0
10x2 + 25x3 =90
20x1 + 10x2 =80.

Applied Mathematics 29/297


Linear Algebra Matrices and Linear Systems

The associated augmented matrix is


 
1 (−1) 1 0
 (−1) 1 (−1) 0 
A
e := 
 0
.
10 25 90 
20 10 0 80

Step 1. Elimination of the first unknown. We mark the first row as the
pivot row and its first entry as the pivot. To eliminate x1 in the other
equations, we add 1 times the pivot row to the second row and add
(−20) times the pivot row to the fourth row. This yields
 
1 (−1) 1 0
 0 0 0 0 
 .
 0 10 25 90 
0 30 (−20) 80

Applied Mathematics 30/297


Linear Algebra Matrices and Linear Systems

Step 2. Elimination of the second unknown. First, push the zero row at
the end of the matrix
 
1 (−1) 1 0
 0 10 25 90 
 .
 0 30 (−20) 80 
0 0 0 0

Then, mark the second row as the pivot row and its first non-zero entry
as the pivot. To eliminate x2 in the other equations, we add (−3) times
the pivot row to the third row. Hence, the result is
 
1 (−1) 1 0
 0 10 25 90 
 .
 0 0 (−95) (−190) 
0 0 0 0

Applied Mathematics 31/297


Linear Algebra Matrices and Linear Systems

Step 3. Back substitution. Now, working backwards from the last to the
first row of this triangular matrix, we can readily find x3 , x2 and x1 as
follows.
(−190)
x3 = = 2,
−95x3 = − 190 (−95)
10x2 + 25x3 =90 =⇒ 1
x2 = (90 − 25 × 2) = 4,
x1 − x2 + x3 =0 10
x1 =4 − 2 = 2.

Therefore, we have a unique solution. 

Applied Mathematics 32/297


Linear Algebra Matrices and Linear Systems

Elementary Row Operations for Matrices


• Interchange of two rows.
• Addition of a constant multiple of one row to another row.
• Multiplication of a row by a non-zero constant.

Elementary Operations for Equations


• Interchange of two equations.
• Addition of a constant multiple of an equation to the others.
• Multiplication of an equation by a non-zero constant.

Applied Mathematics 33/297


Linear Algebra Matrices and Linear Systems

Definition 11
We now call a linear system S1 row-equivalent to another linear system
S2 if S1 can be obtained from S2 only by applying finitely many row
operations.

Thus, we can state the following theorem, which also justifies the Gauss
elimination.
Theorem 1
Row-equivalent linear systems have the same set of solutions.

Applied Mathematics 34/297


Linear Algebra Matrices and Linear Systems

At the end of the Gauss elimination the form of the coefficient matrix,
the augmented matrix or the system itself is called row-reduced form.
In it, rows of zeroes, if present, are the last rows, and in each non-zero
row the leftmost non-zero entry is farther to the right than in the
previous row. For instance, (see Example 8) the following coefficient
matrix and its augmented matrix are in the row-reduced form.
   
1 (−1) 1 1 (−1) 1 0
 0 10 25   and  0 10 25 90
 
 .
 0 0 (−95)   0 0 (−95) (−190) 
0 0 0 0 0 0 0

Note that we do not require that the leftmost non-zero entries be 1 since
this has no theoretic or numeric advantage.

Applied Mathematics 35/297


Linear Algebra Matrices and Linear Systems

Example 9
Transform the following matrices into row-reduced forms.
   
(−1) 4 1 1 (−2) 1 3
 0 0 0 0  and  0 1 1 .
0 0 0 1 2 0 1

Applied Mathematics 36/297


Linear Algebra Matrices and Linear Systems

 
 a11 a12 · · · · · · · · · a1n b1 

 a22 · · · · · · · · · a2n b2 

 
 .. .. .. 

 . . . 

a · · · a br
 
 rr rn 
 

 br +1 

..
 
 

 . 

bm

Figure 4: Row reduced form of the augmented matrix at the end of the Gauss
elimination. Here, r ≤ m and a11 , a22 , · · · , arr 6= 0, and all the entries in the red
triangle and the blue rectangle are zero.

From this, we see that with respect to the solutions of the system with
augmented matrix in Figure 4 (and thus, with respect to the originally
given system) there are three possible cases.

Applied Mathematics 37/297


Linear Algebra Matrices and Linear Systems

Exactly One Solution

If r = n and br +1 , br +2 , · · · , bm = 0 (if present), then we have unique


solution. To get the solution, we solve the n-th equation ann xn = bn , and
then use back substitution.
 
 a11 a12 · · · a1n b1 
a · · · a b
 
 22 2n 2 
 
 .. .. .. 

 . . . 
ann bn 
 

 
 
 
 
 

Figure 5: All the entries in the red triangle and the blue rectangles are zero.

Applied Mathematics 38/297


Linear Algebra Matrices and Linear Systems

Infinitely Many Solutions


If r < n and br +1 , br +2 , · · · , bm = 0 (if present), then we have infinitely
many solutions. To get the solution, we pick xr +1 , xr +2 , · · · , xn as
arbitrary and solve xr from the r -th equation
arr xr + ar (r +1) xr +1 + · · · + arn xn = br , and then use back substitution.
 
 a11 a12 · · · · · · · · · a1n b1 
a22 · · · · · · · · · a2n b2 
 

 
 .. .. .. 

 . . . 

arr · · · arn br 
 

 
 
 
 
 

Figure 6: All the entries in the red triangle and the blue rectangles are zero.

Applied Mathematics 39/297


Linear Algebra Matrices and Linear Systems

No Solution

If r < m and at least one br +1 , br +2 , · · · , bm is non-zero, then we have


no solution.
 
a
 11 12 a · · · · · · · · · a1n b 1 

 a22 · · · · · · · · · a2n b 2


 

 . .. .
.. .
..


 
arr · · · arn br 
 

 

 br +1 
..
 
 

 . 

bm

Figure 7: All the entries in the red triangle and the blue rectangle are zero, but
there exists non-zero entries in the orange rectangle.

Applied Mathematics 40/297


Vector Spaces

Vector Spaces

• A quantity such as work, area or energy, which is described in terms of


magnitude only is called a scalar.
• A quantity, which needs both magnitude and direction to be described
is called a vector.
A vector is an element of vector space, which will be introduced below.

Applied Mathematics 41/297


Vector Spaces

Definition 12 (Vector Space)


A vector space V over the field R is a set satisfying the following
properties.
1. Closure: u + v ∈ V for all u, v ∈ V .
2. Commutativity: u + v = v + u for all u, v ∈ V .
3. Associativity: u + (v + z) = (u + v ) + z for all u, v , z ∈ V .
4. Identity Element: There exists 0 ∈ V such that u + 0 = u for all
u ∈ V.
5. Inverse Element: There exists (−u) ∈ V such that u + (−u) = 0 for
all u ∈ V .
6. Closure: αu ∈ V for all u ∈ V and all α ∈ R.
7. Distributivity of scalar multiplication: α(u + v ) = αu + αv .
8. Distributivity of scalar multiplication: (α + β)u = αu + βu.
9. Compatibility of scalar multiplication: α(βu) = (αβ)u.
10. Identity Element: 1u = u.

Applied Mathematics 42/297


Vector Spaces

Example 10
Some examples of vector spaces are listed below.
1. The singleton {0}.
2. Rn (the set of vectors).
3. Rm×n (the set of m × n matrices).
4. Pn [x] (the set of polynomials).
5. F[a, b] (the set of functions on [a, b]).
6. C1 [a, b] (the set of continuously differentiable functions on [a, b]).

Applied Mathematics 43/297


Vector Spaces

Definition 13 (Linear Combination)


Let u 1 , u 2 , · · · , u n ∈ V and α1 , α2 , · · · , αn ∈ R. Then, the vector

α1 u 1 + α2 u 2 + · · · + αn u n

is called as a linear combination of the vectors u 1 , u 2 , · · · , u n .

Definition 14 (Linear Dependence)


Let u 1 , u 2 , · · · , u n ∈ V . If the equation

α1 u 1 + α2 u 2 + · · · + αn u n = 0

implies that α1 , α2 , · · · , αn = 0, then the vectors u 1 , u 2 , · · · , u n are


called as linearly independent. If there exist scalars α1 , α2 , · · · , αn ∈ R
not all zero such that the linear combination satisfies

α1 u 1 + α2 u 2 + · · · + αn u n = 0,

then the vectors u 1 , u 2 , · · · , u n are called as linearly dependent.

Applied Mathematics 44/297


Vector Spaces

Example 11
1. Show that (1, 1), (−3, 2) are linearly independent.
2. Show that (1, 1), (−3, 2), (2, 4) are linearly dependent.
10 20 2×
(− −→
3,
2)

16

(2, →
4)

1


0

5
(−
−→ )
,1
(1
×
16

-10 0
-10 -3 0 1 2 10 0 10 16 20

Figure 8: The figures show that 16(1, 1) + 2(−3, 2) + (−5)(2, 4) = (0, 0), i.e.,
(1, 1), (−3, 2), (2, 4) are linearly dependent.

3. Show that 1, x, x 2 , (x − 4)2 are linearly dependent.


4. Show that e−t , et , cosh(t) are linearly dependent.

Applied Mathematics 45/297


Vector Spaces

Definition 15 (Rank of a Matrix)


The rank of a matrix A is the maximum number of linearly independent
row vectors of A, and it is denoted by rank(A).

Example 12
Find the rank of the matrix
 
3 0 2 2
A :=  (−6) 42 24 54  .
21 (−21) 0 (−15)

Clearly, we have

12(3, 0, 2, 2) + (−1)(−6, 42, 24, 54) + (−2)(21, −21, 0, −15) = (0, 0, 0, 0)

shows that rank(A) ≤ 2. On the other hand, we obtain

α(3, 0, 2, 2) + β(21, −21, 0, −15) = (0, 0, 0, 0) =⇒ α = 0 and β = 0.

Thus, rank(A) = 2.

Applied Mathematics 46/297


Vector Spaces

Theorem 2 (Rank of Row-equivalent Matrices)


Row-equivalent matrices have the same rank.

In other words, this theorem allows us to use Gauss elimination to find


the rank. In that case, the number of non-zero rows of the row-reduced
matrix is the rank.
Example 13
   
(−2) 2 1 3 (−2) 2 1 3
 0 0 1 1   0 0 1 1 
  and   are row-equivalent.
 2 (−2) 0 1   0 0 0 3 
2 (−2) 4 8 0 0 0 0
The number of non-zero rows of the last matrix is 3, i.e., its rank is 3.
Thus, the first matrix has the rank 3 too.

Applied Mathematics 47/297


Vector Spaces

Theorem 3 (Linear Dependence of a Set of Vectors)


A set of p vectors with n components each are linearly independent if the
matrix with these vectors as row vectors has rank p and linearly
dependent if the rank is less than p.

Example 14
The set{(1, 2, 1), (−2, −3, 1),
 (3, 5, 0)} is linearly dependent since the
1 2 1
matrix  (−2) (−3) 1  has the rank 2, which is less than 3.
3 5 0

Applied Mathematics 48/297


Vector Spaces

Theorem 4 (Rank in Terms of Column Vectors)


The rank of a matrix A is equal to the maximum number of linearly
independent column vectors of A.

Thus, A and AT have the same rank.

Applied Mathematics 49/297


Vector Spaces

Definition 16 (Spanning of a Set)


Let V be a vector space and S be a subset of V . Span(S) is the set of
all linear combinations of the vectors in S, i.e.,

Span(S) := {α1 u 1 +· · ·+αn u n : u 1 , · · · , u n ∈ S, α1 , · · · , αn ∈ R, n ∈ N}.

Example 15
Show that Span{(1, 1), (−3, 2), (2, 4)} = R2 .

Applied Mathematics 50/297


Vector Spaces

Definition 17 (Basis of Vector Spaces)


Let V be a vector space and B be a linearly independent subset of V . B
is a basis of V provided that every element in V can be written as a
linear combination of vectors in B, i.e., for every u ∈ V , there exist
n ∈ N, u 1 , u 2 , · · · , u n ∈ B and α1 , α2 , · · · , αn ∈ R such that
u = α1 u 1 + α2 u 2 + · · · + αn u n . In this case, we write V = Span(B).

Definition 18 (Dimension of Vector Spaces)


Let V be a vector space and B be a basis of V . In this case, dimension
of V is the number of vectors in B and is denoted by dim(V ).

Example 16
Show that {(1, 1), (−3, 2)} is a basis for R2 . Hence, dim(R2 ) = 2.

Applied Mathematics 51/297


Vector Spaces

Definition 19 (Dimension of Row-Space and Column-Space of Matrices)


Let A be a matrix. The row-space (the spanning set of the row vectors)
and the column space (the spanning set of the column vectors) of A have
the same dimension, which is equal to the rank of A.

Finally for a given matrix A, the solution set of the homogeneous system
Ax = 0 is a vector space, which is called as the null space of A, and its
dimension is called as the nullity of A.

Theorem 5 (Dimension Theorem)

Let A be a matrix of the size m × n. Then,

rank(A) + nullity(A) = n.

Applied Mathematics 52/297


Vector Spaces

Consider the non-homogeneous system

a11 x1 + a12 x2 + · · · + a1n xn =b1


a21 x1 + a22 x2 + · · · + a2n xn =b2
.. (4)
.
am1 x1 + am2 x2 + · · · + amn xn =bm ,

where a11 , a12 , · · · , amn , b1 , b2 , · · · , bm are scalars and x1 , x2 , · · · , xn are


unknowns.

Applied Mathematics 53/297


Vector Spaces

Theorem 6 (Fundamental Theorem for Linear Systems)

1. (4) is consistent (i.e., it has solutions) if and only if the coefficient


matrix A and its augmented matrix A e have the same rank.
2. (4) has unique solution if and only if the common rank r of A and A e
is equal to n.
3. (4) has infinitely many solutions if the common rank r is less than n.
In this case, a total of r unknowns can be obtained in terms of (n − r )
unknowns (or parameters).

When (4) admits solutions, one can apply Gauss elimination to find those
solutions.

Applied Mathematics 54/297


Vector Spaces

Consider the homogeneous system

a11 x1 + a12 x2 + · · · + a1n xn =0


a21 x1 + a22 x2 + · · · + a2n xn =0
.. (5)
.
am1 x1 + am2 x2 + · · · + amn xn =0,

where a11 , a12 , · · · , amn are scalars and x1 , x2 , · · · , xn are unknowns. Note
that x1 , x2 , · · · , xn = 0 is always a solution, which is called as the trivial
solution.
According to Theorem 6, (5) admits non-trivial solutions if r < n, where
r := rank(A). If r < n, then these solutions, together with x = 0, form a
vector space of dimension (n − r ), which is called as the solution space
of (5).

Applied Mathematics 55/297


Vector Spaces

In particular, if x 1 and x 2 are two solutions of (5), then the vector


x := c1 x 1 + c2 x 2 , where c1 and c2 are scalars, is also a solution of (5).
This property does not hold for non-homogeneous systems. It should also
be mentioned that the term “solution space” is only used for
homogeneous systems.

Applied Mathematics 56/297


Vector Spaces

The solution space of (5) is also called as the null space of A because of
Ax = 0 for every x in the solution space of (5). Its dimension is called as
the nullity of A. Hence, by Theorem 5, we have

rank(A) + nullity(A) = n,

where n is the number of unknowns (i.e., the number of columns of A).

Applied Mathematics 57/297


Vector Spaces

Example 17
Show that the solution space (null space) of
 
  x  
1 2 3 4   y 
= 0
2 4 7 8  z  0
w

is   

 x 

 y  : x = −2y − 4w and z = 0
  
 z  
 
w
 

or equivalently
     

 (−2) (−4) 

1   0 
  
y
  +w
  : y, w ∈ R .

 0  0  

0 1
 

Hence, the dimension of the solution space (nullity) is 2.


Applied Mathematics 58/297
Vector Spaces

Theorem 7
A homogeneous linear system with fewer equations than the number of
unkowns has always non-trivial solutions.

Applied Mathematics 59/297


Vector Spaces

Theorem 8
If a non-homogeneous linear system is consistent, then all of its solutions
are obtained as
x = xh + xp,
where x h runs through all solutions of the associated homogeneous
system and x p is any (fixed) solution of the non-homogeneous system.

Applied Mathematics 60/297


Vector Spaces

Example 18
Show that 

  x  
3 2 3 (−2)  y  1
 1 1 1 0  = 3 
 z 
1 2 1 (−1) 2
w
has    
1 1
 0   2 
x h := c  
 (−1)  and x p := 
 0 ,

0 3
where c is an arbitrary constant. So, any solution of the system is of the
form    
1 1
 0   2 
x =c  (−1)  +  0  ,
  

0 3
where c is an arbitrary constant.

Applied Mathematics 61/297


Determinants

Determinants

Determinant is a function from square matrices to scalars.


Our efficient computational procedure will be cofactor expansion.
However, for square matrices of the size 2 or 3, we will give simpler rules.

Applied Mathematics 62/297


Determinants

Definition 20 (Determinant of 2 × 2 Matrices)


 
a11 a12
Let A := , then its determinant is the number defined by
a21 a22

a11 a12
a21 a22 := a11 a22 − a12 a21 .

Determinat of A is also denoted by |A| or det(A).

Example 19
We compute that

3 2

(−4) = 3 · 1 − 2 · (−4) = 11.
1

Applied Mathematics 63/297


Determinants

Definition 21 (Determinant of 3 × 3 Matrices)


 
a11 a12 a13
Let A :=  a21 a22 a23 , then det(A) is defined by
a31 a32 a33

a11 a12 a13

a21 a22 a23 :=[a11 a22 a33 + a21 a32 a13 + a31 a12 a23 ]

a31 a32 a33
− [a13 a22 a31 + a23 a32 a11 + a33 a12 a21 ].

Applied Mathematics 64/297


Determinants



a11 a12 a13


a21 a22 a23

a31 a32 a33


(−) a11 a12 a13 (+)
(−) a21 a22 a23 (+)
(−) (+)

Figure 9: A diagram for the Sarrus rule.

Applied Mathematics 65/297


Determinants

Example 20
We compute by Sarrus rule that

1 5 (−3)

2
3 (−4) =[1 · 3 · 2 + 2 · 6 · (−3) + (−1) · 5 · (−4)]
(−1) 6 2
− [(−3) · 3 · (−1) + (−4) · 6 · 1 + 2 · 5 · 2]
=[6 + (−36) + 20] − [9 + (−24) + 20]
=(−10) − 5 = −15.

Applied Mathematics 66/297


Determinants

Before proceeding for the general definition of a square matrix


determinant, we need to introduce two concepts.

Definition 22 (Minors and Cofactors of a Matrix)


Let A := (aij ) be square matrix of size n. For an entry aij , its minor,
which is denoted by Mij , is defined to be the determinant of the matrix
formed by deleting the row and the column where aij is located. For an
entry aij , its cofactor is defined to be Cij := (−1)i+j Mij .

Applied Mathematics 67/297


Determinants

 
a11 ··· a1(j−1) a1j a1(j+1) ··· a1n
 
 .. .. .. .. .. .. .. 

 . . . . . . . 

 a
 (i−1)1 · · · a(i−1)(j−1) a(i−1)j a(i−1)(j+1) · · · a(i−1)n 

ai1 · · · ai(j−1) aij ai(j+1) · · · ain
 
 
 
 a(i+1)1
 · · · a(i+1)(j−1) a(i+1)j a(i+1)(j+1) · · · a(i+1)n 

 
 .. .. .. .. .. .. .. 

 . . . . . . . 

an1 ··· an(j−1) anj an(j+1) ··· ann

Figure 10: A diagram for minors and cofactors.

Applied Mathematics 68/297


Determinants

Definition 23 (Determinant of n × n Matrices by Cofactor Expansion)


Let A := (aij ) be an n × n matrix. Determinant of A through the k-th
row is defined by

det(A) := ak1 Ck1 + ak2 Ck2 + · · · + akn Ckn .

Similarly, cofactor expansion through the k-th column is defined by

det(A) := a1k C1k + a2k C2k + · · · + ank Cnk .

While computing determinant of a matrix by cofactor expansion, it would


be easier to choose the row or the column, where zeroes appear the most.
If A = diag(ak ) is a diagonal matrix of size n, then det(A) = a1 a2 · · · an .

Applied Mathematics 69/297


Determinants

Example 21
We compute that

6 0 (−3) 5

4 13 6 (−8)

(−1) 0 7 4

8 6 0 2

4 6 (−8) 6 (−3) 5
=0(−1)1+2 (−1) 7 4 + 13(−1)2+2 (−1)

7 4
8 0 2 8 0 2

6 (−3) 5 6 (−3) 5
3+2 4+2

+ 0(−1) 4 6 (−8)
+ 6(−1) 4 6 (−8)

8 0 2 (−1) 7 4
=0 · (−1) · 708 + 13 · 1 · (−298) + 0 · (−1) · 48 + 6 · 1 · 674
=(−3874) + 4044 = 170.

Applied Mathematics 70/297


Determinants

Theorem 9
Let A and B be two square matrices of the same size. Then,

det(AB) = det(BA) = det(A) det(B).

Applied Mathematics 71/297


Determinants

Property 4 (Properties of Determinant)


Determinant of a matrix satisfy the following properties.
1. Interchange of two rows or two columns multiplies the value of the
determinant by (−1).
2. Addition of a multiple of a row or a column to another does not
change the value of the determinant.
3. Multiplication of a row or a column by a non-zero constant c
multiplies the value of the determinant by c.

Applied Mathematics 72/297


Determinants

Property 5 (Further Properties of Determinant)

Determinant of a matrix also has the following properties.


1. Transposition of a matrix does not change the value of the
determinant.
2. A zero row or a column renders the value of the determinant zero.
3. Proportional rows or columns render the value of the determinant
zero. In particular, a determinant with two identical rows or columns has
the value zero.

Theorem 10 (Determinant of Orthogonal Matrices)


The determinant of an orthogonal matrix is either 1 or (−1).

Applied Mathematics 73/297


Cramer’s Rule

Cramer’s Rule
Theorem 11 (Cramer’s Rule)

Consider the square system

a11 x1 + a12 x2 + · · · + a1n xn =b1


a21 x1 + a22 x2 + · · · + a2n xn =b2
.. (6)
.
an1 x1 + an2 x2 + · · · + ann xn =bn .

Denote by A and b the coefficient matrix and the right-hand side vector,
respectively. For k = 1, 2, · · · , n, denote by Ak the matrix formed by
replacing k-th column of A by b. Then, we have the following.
det(Ak )
1. If det(A) 6= 0, then (6) has a unique solution given by xk = det(A) .
2. If det(A) = 0 and det(Ak ) = 0 for all k, then (6) has infinitely many
solutions.
3. If det(A) = 0 and det(Ak ) 6= 0 for some k, then (6) has no solutions.

Applied Mathematics 74/297


Cramer’s Rule

As an immediate consequence of Theorem 11, we can give the following


corollary.
Corollary 1

Consider the square homogeneous system

a11 x1 + a12 x2 + · · · + a1n xn =0


a21 x1 + a22 x2 + · · · + a2n xn =0
.. (7)
.
an1 x1 + an2 x2 + · · · + ann xn =0.

Then, (7) has the trivial solution only if and only if det(A) 6= 0.

Note that in this case, det(Ak ) = 0 for all k by the second case of
Corollary 5, and thus last case of Theorem 11 does not appear.

Applied Mathematics 75/297


Cramer’s Rule

Example 22
The system
x + 2y =−3
2x + 6y + 4z = − 6
−x + 2z =1
has a unique solution since the determinant of the coefficient matrix is
(−4). Thus, the solution is given by

(−3) 2 0 1 (−3) 0

(−6) 6 4 2 (−6) 4

1 0 2 (−4) (−1) 1 2
x= = = 1, y = = −2,
(−4) (−4) (−4)

and similarly z = 1.

Applied Mathematics 76/297


Cramer’s Rule

Definition 24 (Singular Matrix and Non-Singular Matrix)


Let A be a square matrix of size n. A is said to be singular if A−1 does
not exist. Otherwise, A is said to be non-singular.

Theorem 12 (Existence of the Inverse Matrix)


Let A be a square matrix of size n. A is non-signular if and only if
det(A) 6= 0. Further, A is non-singular if and only if rank(A) = n.

The inverse matrix A−1 of A can be obtained by transforming the


augmented matrix (A | I) into (I | B) by using elementary row operations.
This procedure is known as Gauss-Jordan elimination process and the
matrix B is the inverse of A, i.e., B := A−1 .

Applied Mathematics 77/297


Cramer’s Rule

Example 23
Find the inverse matrix of the matrix A given below.
 
1 2 3
A :=  0 1 4 .
5 6 0

By using Gauss-Jordan process for the matrix (A | I), we have


   
1 2 3 1 0 0 1 2 3 1 0 0
 0 1 4 0 1 0 ∼ 0 1 4 0 1 0 
5 6 0 0 0 1 0 (−4) (−15) (−5) 0 1
   
1 2 3 1 0 0 1 2 3 1 0 0
∼ 0 1 4 0 1 0 ∼ 0 1 0 20 (−15) (−4) 
0 0 1 (−5) 4 1 0 0 1 (−5) 4 1

Applied Mathematics 78/297


Cramer’s Rule

 
1 2 0 16 (−12) (−3)
∼ 0 1 0 20 (−15) (−4) 
0 0 1 (−5) 4 1
 
1 0 0 (−24) 18 5
∼ 0 1 0 20 (−15) (−4)  .
0 0 1 (−5) 4 1
Hence,  
(−24) 18 5
A−1 :=  20 (−15) (−4)  .
(−5) 4 1


Applied Mathematics 79/297


Cramer’s Rule

Before we introduce a formula for the inverse matrix, we need to


introduce a new matrix.
Definition 25 (Adjoint Matrix)
Let A = (aij ) be a square matrix. The adjoint matrix of A is defined by
adj(A) := (Cij )T , where Cij denotes the cofactor of aij .

Theorem 13 (Inverse of a Matrix)


Let A = (aij ) be a square matrix. Then,

1
A−1 := adj(A)
det(A)

provided that det(A) 6= 0.

It should be noted that if A = diag(ak ) is a diagonal


 matrix such that no
entry on the diagonal is zero, then A−1 = diag a1k .

Applied Mathematics 80/297


Cramer’s Rule

Example 24
Clearly, for  
3 1
A :=
2 4
we have det(A) = 10 and
 T  
4 (−2) 4 (−1)
adj(A) = = .
(−1) 3 (−2) 3

Therefore, we have
 
−1 1 4 (−1)
A = .
10 (−2) 3

Applied Mathematics 81/297


Cramer’s Rule

Example 25
For the matrix  
1 (−1) (−2)
A :=  3 (−1) 1 ,
1 (−3) (−4)
we compute that det(A) = 10. Also, we have
 
7 2 (−3)
adj(A) =  13 (−2) (−7)  .
(−8) 2 2

Thus, the inverse matrix is


 
7 2 (−3)
1 
A−1 = 13 (−2) (−7)  .
10
(−8) 2 2

Applied Mathematics 82/297


Cramer’s Rule

Theorem 14
Let A and B be two square matrices of the same size. Then,

(AB)−1 = B −1 A−1 .

Applied Mathematics 83/297


Cramer’s Rule

The inverse matrix is useful for solving systems of the form

Ax = b.

Once A−1 is known, we can multiply both sides of the above system by
A−1 and get
x = A−1 b.

Example 26
Show that
x − y − 2z = − 10
3x − y + z =5
x − 3y − 4z =20
has the solution

x = −12, y = −28 and z = 13.

Applied Mathematics 84/297


Cramer’s Rule

Remark 1
Some remarks on matrix multiplication are given below.
1. Matrix multiplication is not commutative, i.e., we have in general that

AB 6= BA.

2. AB = 0 does not imply that A = 0 or B = 0 (or BA = 0). For


instance,     
1 1 (−1) 1 0 0
= .
2 2 1 (−1) 0 0
3. AB = AC does not generally imply B = C even when A 6= 0.

Applied Mathematics 85/297


Cramer’s Rule

Theorem 15 (Cancellation Laws)


Let A, B, C be square matrices of the same size. Then, the following
properties hold.
1. If det(A) 6= 0 and AB = AC , then B = C .
2. If det(A) 6= 0 and AB = 0, then B = 0. Hence, if AB = 0 with
A 6= 0 and B 6= 0, then det(A) = 0 or det(B) = 0.
3. If A is singular, then so are AB and BA.

Applied Mathematics 86/297


Real Inner Product Spaces

Real Inner Product Spaces

Definition 26 (Real Inner Product Space)


Let V be a vector space over the field R. V is said to be an inner
product space over R, if there exists a mapping h·, ·i : V × V → R
satisfying the following properties.
1. Positive-Definiteness:
(a) hu, ui ≥ 0 for all u ∈ V .
(b) hu, ui = 0 if and only if u = 0.
2. Symmetry: hu, v i = hv , ui for all u, v ∈ V .
3. Linearity:
(a) hu + v , zi = hu, zi + hv , zi for all u, v , z ∈ V .
(b) hαu, v i = αhu, v i for all u, v ∈ V and all α ∈ R.

Applied Mathematics 87/297


Real Inner Product Spaces

Example 27
Let (u1 , u2 ), (v1 , v2 ) ∈ R2 , then

h(u1 , u2 ), (v1 , v2 )i = u1 v1 + u2 v2

satisfies the properties of vector spaces. Indeed, we have

h(u1 , u2 ), (u1 , u2 )i = u12 + u22 ≥ 0,

and it is zero if and only if u1 = 0 and u2 = 0, i.e., (u1 , u2 ) = (0, 0).


Next, we see that

h(u1 , u2 ), (v1 , v2 )i = u1 v1 + u2 v2 = v1 u1 + v2 u2 = h(v1 , v2 ), (u1 , u2 )i. I

Applied Mathematics 88/297


Real Inner Product Spaces

Finally, we have

h(u1 , u2 ) + (v1 , v2 ), (z1 , z2 )i = h(u1 + v1 , u2 + v2 ), (z1 , z2 )i


=(u1 + v1 )z1 + (u2 + v2 )z2 = (u1 z1 + u2 z2 ) + (v1 z1 + v2 z2 )
=h(u1 , u2 ), (z1 , z2 )i + h(v1 , v2 ), (z1 , z2 )i

and
hα(u1 ,u2 ), (v1 , v2 )i = h(αu1 , αu2 ), (v1 , v2 )i
=(αu1 )v1 + (αu2 )v2 = α(u1 v1 + u2 v2 )
=αh(u1 , u2 ), (v1 , v2 )i.
Thus, R2 is an inner product space on R. 

Applied Mathematics 89/297


Real Inner Product Spaces

Example 28
Two examples of inner product spaces are listed below.
1. Rn is an inner product space with
   
* u1 v1 +
 u2   v2 
 ..  ,  ..  := u1 v1 + · · · + un vn .
   
 .   . 
un vn

2. C[a, b] is an inner product space with


Z b
hf , g i := f (ξ)g (ξ)dξ.
a

Applied Mathematics 90/297


Real Inner Product Spaces

Definition 27 (Orthogonality)
Let V be an inner product space on R, and u, v ∈ V . The vectors u and
v are called orthogonal provided that hu, v i = 0.

Example 29
1. In R3 , u := (2, 1, −1) and v := (1, −1, 1) are orthogonal.
a+b a+b 2

2. In C[a, b], f (x) := x − 2 and g (x) := x 2 − (a + b)x + 2 are
orthogonal.

Applied Mathematics 91/297


Real Inner Product Spaces

Definition 28 (Norm)
Let V be an inner product space on R, and u ∈ V . The norm of u is
defined by p
kuk := hu, ui.
Further, u is said to be a unit vector provided that kuk = 1.

Example 30
1. On Rn , a norm is given by
 
u1

 u2  q
 ..  := u12 + u22 + · · · + un2 .
 
 . 

un

2. On C[a, b], a norm is given by


s
Z b 2
kf k := f (ξ) dξ.
a

Applied Mathematics 92/297


Real Inner Product Spaces

Property 6
Inner products and norms satisfy the following properties.
1. Cauchy-Schwarz Inequality: |hu, v i| ≤ kukkv k for all u, v ∈ V .
2. Triangle Inequality: ku + v k ≤ kuk + kv k for all u, v ∈ V .

3. Parallelogram Equality: ku + v k2 + ku − v k2 = 2 kuk2 + kv k2 for
all u, v ∈ V .

Applied Mathematics 93/297


Linear Transformations

Linear Transformations

Definition 29 (Linear Transformation)


Let V and W be vector spaces over the field R. The mapping
f : V → W is called as a linear transform provided that the following
properties are satisfied.
1. f (u + v ) = f (u) + f (v ) for all u, v ∈ V .
2. f (αu) = αf (u) for all u ∈ V and all α ∈ R.

Applied Mathematics 94/297


Linear Transformations

Example 31
Some examples of linear transforms are listed below.
1. Zero transform, i.e., f (u) = 0.
2. Identity transform, i.e., f (u) = u.
3. Reflection transform, i.e., f (u) = −u.
4. Scaling transform, i.e., f (u) = λu, where λ ∈ R
5. Projection transform.
6. Rotation transform.
7. Differential transform.
8. Integral transform.

Applied Mathematics 95/297


Linear Transformations

Representation Matrix

Let {e 1 , e 2 , · · · , e n } be a basis for Rn . Then for every u ∈ Rn , there


exist scalars α1 , α2 , · · · , αn ∈ R such that

u = α1 e 1 + α2 e 2 + · · · + αn e n .

If f is a linear mapping, then we have

f (u) = α1 f (e 1 ) + α2 f (e 2 ) + · · · + αn f (e n ).

Hence, if we know the values f (e 1 ), f (e 2 ), · · · , f (e n ), then we can


uniquely determine f (u). This idea can be used to write the transform f
as
f (u) = Au,
where A is an m × n matrix provided that f : Rn → Rm .

Applied Mathematics 96/297


Linear Transformations

Example 32
   
1 1
Consider the basis , of R2 . Let us find the rule of
1 (−1)
the linear transform f : R2 → R3 satisfying
   
  5   3
1 1
f =  (−1)  and f =  (−5)  .
1 (−1)
11 1

Applied Mathematics 97/297


Linear Transformations

This implies that the transform f is of the form


      
x x +y 1 x −y 1
f =f +
y 2 1 2 (−1)
   
x +y 1 x −y 1
= f + f
2 1 2 (−1)
   
5 3
x +y  x −y 
= (−1)  + (−5) 
2 2
11 1
 
4x + y
=  −3x + 2y  .
6x + 5y

Applied Mathematics 98/297


Linear Transformations

Thus, f has the matrix representation


 
  4 1  
x x
f =  (−3) 2  .
y y
6 5

Applied Mathematics 99/297


Linear Transformations

Definition 30 (Null Space and Range)


Let f : V → W be a linear transform. The null space of f is defined by

Null(f ) := {u : f (u) = 0}

and the range of f is defined by

Range(f ) := {v : f (u) = v }.

Further,
 
nullity(f ) := dim Null(f ) and rank(f ) := dim Range(f ) .

Applied Mathematics 100/297


Linear Transformations

Recalling Theorem 5 and the definition of the representation matrix, we


can give the following result.
Theorem 16
Let f : V → W be a linear transform. Then,

rank(f ) + nullity(f ) = dim(V ).

Applied Mathematics 101/297


Linear Transformations

Example 33
Find the rank and the nullity of the transform f : R3 → R2 defined by
 
x  
16x + y − 3z
f  y = .
8x − 3y + z
z

We can rewrite f in the form


   
x   x
16 1 (−3)
f  y  =  y ,
8 (−3) 1
z z

Applied Mathematics 102/297


Linear Transformations

We see that    
 1 
Null(f ) = x  5  : x ∈ R
7
 
 
 1 
= Span  5  ,
7
 

and thus nullity(f ) = 1. Then, by Theorem 16, we have rank(f ) = 2. 

Applied Mathematics 103/297


Eigenvalues, Eigenvectors and Eigenspaces

Eigenvalues, Eigenvectors and Eigenspaces

Let A := (aij ) be an n × n matrix and consider the vector equation

Ax = λx. (8)

Here, x is an unknown vector and λ is an unknown scalar. Our task is to


determine the x vectors and λ values for which (8) holds. Geometrically,
we are looking for vectors x for which multiplication by A has the same
effect as multiplication by a scalar λ, i.e., Ax is proportional to x.

Applied Mathematics 104/297


Eigenvalues, Eigenvectors and Eigenspaces

Clearly, x = 0 is a solution of (8) for any value of λ since A0 = 0 = λ0.


This is obvious and is of no interest. A value of λ for which (8) has a
non-trivial solution (x 6= 0) is called as an “eigenvalue” (or a
“characteristic value”) of the matrix A.
The non-trivial solutions x of (8) are called as “eigenvectors” (or
“characteristic vectors”) of the matrix A corresponding to the
eigenvalue λ.
The set of all eigenvalues of A is called as the “spectrum” of A. We will
see in examples that the spectrum of A consists of at least one
eigenvalue and at most of n numerically different eigenvalues.

Applied Mathematics 105/297


Eigenvalues, Eigenvectors and Eigenspaces

This example illustrates the general case as follows. Expanding (8) in its
components, we get

a11 x1 + a12 x2 + · · · + a1n xn =λx1


a21 x1 + a22 x2 + · · · + a2n xn =λx2
..
.
an1 x1 + an2 x2 + · · · + ann xn =λxn .

Applied Mathematics 106/297


Eigenvalues, Eigenvectors and Eigenspaces

Moving the terms on the right-hand side to the left-hand side gives us

(a11 − λ)x1 + a12 x2 + · · · + a1n xn =0


a21 x1 + (a22 − λ)x2 + · · · + a2n xn =0
..
.
an1 x1 + an2 x2 + · · · + (ann − λ)xn =0.

This is written in the matrix form as

(A − λ I)x = 0.

Applied Mathematics 107/297


Eigenvalues, Eigenvectors and Eigenspaces

By Cramer’s rule, this linear system of equations has a non-trivial


solution if and only if the corresponding determinant of the coefficient
matrix is zero, i.e.,
D(λ) := det(A − λ I) = 0.
Here, (A − λ I) is called as the “characteristic matrix” and D(λ) is
called as the “characteristic determinant” of A. Clearly, D(λ) gives us
a polynomial of n-th degree in λ, which is called as the “characteristic
polynomial” of A.

Applied Mathematics 108/297


Eigenvalues, Eigenvectors and Eigenspaces

Theorem 17 (Eigenvalues)
The eigenvalues of a square matrix A are the roots of the characteristic
equation det(A − λ I) = 0. Hence, an n × n matrix has at least 1
eigenvalue and at most n numerically different eigenvalues.

Applied Mathematics 109/297


Eigenvalues, Eigenvectors and Eigenspaces

For obtaining the eigensystem (eigenvalues and eigenvectors) of a square


matrix, we must first find the eigenvalues from the characteristic
equation det(A − λ I) = 0. Then, we should use them in the system
(A − λ I)x = 0 to find the eigenvectors.

Applied Mathematics 110/297


Eigenvalues, Eigenvectors and Eigenspaces

Example 34
 
2 2
The characteristic polynomial of the matrix A := is
1 3

2−λ 2
D(λ) := = λ2 − 5λ + 4 = (λ − 1)(λ − 4)
1 3−λ

whose roots are λ1 := 1 and λ2 := 4. Thus, the eigenvalues of A are 1


and 4, i.e., the spectrum of A is {1, 4}. I

Applied Mathematics 111/297


Eigenvalues, Eigenvectors and Eigenspaces

It follows from the system (A − λ I)x = 0 with λ1 = 1 that


    
1 2 x1 0
= .
1 2 x2 0

Then, we get x1 = −2x2 , which yields


 
(−2)
x =c ,
1

2
where c is an arbitrary constant. So, we can take as the
(−1)
eigenvector
  corresponding
 to the eigenvalue 1. We could also take
(−1) 2
1 instead of . However, both are linearly dependent
2
(−1)
and the latter one looks more simple. I

Applied Mathematics 112/297


Eigenvalues, Eigenvectors and Eigenspaces

The system (A − λ I)x = 0 with λ2 = 4 becomes


    
(−2) 2 x1 0
= .
1 (−1) x2 0

Solving this, we get  


1
x =c ,
1
 
1
where c is an arbitrary constant. Hence, is an eigenvector, which
1
corresponds to the second eigenvalue 4. 

Applied Mathematics 113/297


Eigenvalues, Eigenvectors and Eigenspaces

Theorem 18 (Eigenvectors and Eigenspaces)


Let x and y be eigenvectors of a matrix A corresponding to the same
eigenvalue λ. Then, (x + y ) (with x + y 6= 0) and αx (with αx 6= 0) are
also eigenvectors, which correspond to λ.
Hence, the eigenvectors corresponding to one and the same eigenvalue of
A, together with 0, form a vector space, which is called as the
“eigenspace” of A corresponding to that eigenvalue.

Applied Mathematics 114/297


Eigenvalues, Eigenvectors and Eigenspaces

Example 35
 
1 2 1
Let us consider the matrix A :=  0 3 2 . The characteristic
(−1) 1 1
equation of A is D(λ) := −λ3 + 5λ2 − 6λ = −λ(λ − 2)(λ − 3). Thus, A
has the eigenvalues
 λ
1 := 0, λ2 := 2and λ3 :=
 3. Further, we get
 the
1 3 1
eigenvector  (−2)  for λ1 = 0,  2  for λ2 = 2 and  1 
3 (−1) 0
for λ3 = 3.

Applied Mathematics 115/297


Eigenvalues, Eigenvectors and Eigenspaces

Example 36
 
1 0 0
Consider the matrix A :=  2 1 0 . The characteristic equation
1 (−2) 3
of A is D(λ) := −λ3 + 5λ2 − 7λ + 3 = −(λ − 1)2 (λ − 3). Thus, A has
the
 eigenvalues
 λ1,2 := 1 and λ3 :=
 3. 
Further, we get the eigenvector
0 0
 1  for the eigenvalue 1 and  0  for the eigenvalue 3. In this
1 1
example, we can find only two eigenvectors.

Applied Mathematics 116/297


Eigenvalues, Eigenvectors and Eigenspaces

Theorem 19 (Eigenvalues of a Transpose)


Let A be a square matrix. The matrix A and its transpose AT have the
same eigenvalues.

Applied Mathematics 117/297


Eigenvalues, Eigenvectors and Eigenspaces

Theorem 20 (Eigenvalues of Symmetric and Skew-Symmetric Matrices)


Symmetric matrices have real eigenvalues while skew-symmetric matrices
have purely imaginary or zero eigenvalues.

Applied Mathematics 118/297


Eigenvalues, Eigenvectors and Eigenspaces

Theorem 21 (Eigenvalues of Orthogonal Matrices)


Let A be an orthogonal square matrix. Then, its eigenvalues are real or
complex conjugate in pairs and have modulus 1.

Applied Mathematics 119/297


Diagonalization

Definition 31 (Similar Matrices)


Let A and B be n × n matrices. B is said to be “similar” to A if there
exists some non-singular matrix P such that

B = P −1 AP.

Applied Mathematics 120/297


Diagonalization

Theorem 22 (Eigenvalues and Eigenvectors of Similar Matrices)


If A and B are similar matrices, then both have the same eigenvalues.
Further, if x is an eigenvector of A, then y := P −1 x is an eigenvector of
B, which corresponds to the same eigenvalue.

Applied Mathematics 121/297


Diagonalization

Theorem 23 (Diagonalization of a Matrix)


Let A be an n × n matrix with n linearly independent eigenvalues. Then,

D := X −1 AX

is a diagonal matrix whose entries on the main diagonal are the


eigenvalues of A. Here, X is the matrix with the eigenvectors as column
vectors.
Further, for m ∈ N, we have

D m = X −1 Am X

or equivalently
Am = X D m X −1 .

Applied Mathematics 122/297


Diagonalization

Example 37
 
4 1
Consider the matrix A := . We see that the eigenvalues
(−8) (−5)
 
1
of A are (−4) and 3, which yield the eigenvectors and
(−8)
   
1 1 1
, respectively. Let us define X := , which
(−1) (−8) (−1)
 
(−1) (−1)
yields X −1 = 17 . Then, we get the diagonal matrix
8 1
 
(−4) 0
D := X −1 AX = .
0 3

Further, note that


 
7 7 −1 4840 2653
A = XD X = .
(−21224) (−19037)

Applied Mathematics 123/297


Ordinary Differential Equations

Ordinary Differential Equations

Definition 32 (Ordinary Differential Equations)

An “ordinary differential equation” (in short, ODE) is an equation


containing one unknown function (dependent variable) of one
independent variable together with its derivatives.

Definition 33 (Order of the ODE)

The highest-order derivative appearing in the equation is called the


“order” of a differential equation.

Applied Mathematics 124/297


Ordinary Differential Equations

Example 38

Let us present some ordinary differential equations.


1. 5y 0 (t) + 3y (t) = 6.
2. y 00 (t) + sin(t)y 0 (t) + 2ty (t) = et .
3. ty 000 (t) + t 2 − 1 y 00 (t) + 4y (t) = 0.


4. y (iv) (t) + 8y 00 (t) + 9y (t) = 0.


Each equation has a dependent variable y (unknown function) and a
single independent variable t. Equation 1 and Equation 4 are said to
have constant coefficients, while Equation 2 and Equation 3 are said to
have variable coefficients. The equations above are of first order, second
order, third order and fourth order, respectively.

Applied Mathematics 125/297


Ordinary Differential Equations First-Order Linear Differential Equations

First-Order Linear Differential Equations

In this section, we will consider ordinary differential equations of the form

y 0 (t) + p(t)y (t) = f (t), t ∈ I, (9)

where p and f are continuous functions on some interval I . If f (t) ≡ 0,


then (9) is said to be “homogeneous”, otherwise it is said to be
“non-homogeneous”.

Applied Mathematics 126/297


Ordinary Differential Equations First-Order Linear Differential Equations

Suppose that yc (complementary solution) denotes all solutions of the


associated homogeneous equation

y 0 (t) + p(t)y (t) = 0, (10)

and yp (particular solution) is a solution of the non-homogeneous


equation (9). Then,
y (t) = yc (t) + yp (t)
denotes all solutions (general solution) of the non-homogeneous
equation (9).

Applied Mathematics 127/297


Ordinary Differential Equations First-Order Linear Differential Equations

Homogeneous Equation

Let us first solve homogeneous equation (10). Suppose that y 6= 0 and


write
y 0 (t)
= −p(t),
y (t)
which by integration yields that

Z t
ln y (t) = − p(ξ)dξ + k,

where k is an arbitrary constant. This can be written as


Rt
y (t) = ce− p(ξ)dξ
, (11)

where c := ±ek is a non-zero arbitrary constant. As y ≡ 0 is also a


solution of (10), we can let c in (11) be any arbitrary constant.

Applied Mathematics 128/297


Ordinary Differential Equations First-Order Linear Differential Equations

The general solution of the homogeneous equation (10) is


Rt
y (t) = ce− p(ξ)dξ
,

where c is an arbitrary constant. This includes all solutions of (10).

Applied Mathematics 129/297


Ordinary Differential Equations First-Order Linear Differential Equations

Non-Homogeneous Equation
Next, we proceed
Rt
to find a solution of (9). To this end, we multiply (9)
by µ(t) := e p(ξ)dξ and get

y 0 (t)µ(t) + y (t)µ(t)p(t) = µ(t)f (t)


=⇒ y 0 (t)µ(t) + y (t)µ0 (t) = µ(t)f (t)
=⇒ (y µ)0 (t) = µ(t)f (t),
which by integration yields that
Z t
y (t)µ(t) = µ(ξ)f (ξ)dξ
Z t
1
=⇒ y (t) = µ(ξ)f (ξ)dξ
µ(t)
or equivalently, we have
Rt
Z t Rξ
− p(ξ)dξ p(ζ)dζ
y (t) = e e f (ξ)dξ, (12)

which is a solution of (9).


Applied Mathematics 130/297
Ordinary Differential Equations First-Order Linear Differential Equations

From (11) and (12), the general solution of (9) (all solutions) are given
by Z t R
− t p(ξ)dξ − t p(ξ)dξ ξ
R R
y (t) = ce| {z } + e e p(ζ)dζ f (ξ)dξ ,
complementary solution | {z }
particular solution

where c is an arbitrary constant.

Applied Mathematics 131/297


Ordinary Differential Equations First-Order Linear Differential Equations

Example 39
Show that the general solution of the equation

y 0 + 3y = et

is
1
yc := ce−3t + et ,
4
where c is an arbitrary constant. Further, if the initial condition y (0) = 1
is given, then the desired solution becomes
3 −3t 1 t
y := e + e.
4 4

Figure 11: The graphic of the solution y over the interval [0, 1].

Applied Mathematics 132/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Second-Order Linear Differential Equations with Constant


Coefficients

In this section, we will be obtaining all solutions of the ODE

ay 00 (t) + by 0 (t) + cy (t) = f (t), t ∈ R,

where a, b, c ∈ R with a 6= 0 and f is a continuous function on R. As in


the first-order case, we will proceed with the associated homogeneous
equation first.

Applied Mathematics 133/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Homogeneous Equation

Consider the
ay 00 (t) + by 0 (t) + cy (t) = 0, (13)
rt
where a, b and c are as mentioned above. Substituting y = e into the
left-hand side of (13), we see that

ay 00 + by 0 + cy =ar 2 er t + br er t + cer t
=y (ar 2 + br + c)
=yP(r ),

where the characteristic parabola is defined by

P(r ) := ar 2 + br + c. (14)

Applied Mathematics 134/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Hence, if P(r0 ) = 0, then

ay 00 + by 0 + cy = 0

and thus
y (t) := er0 t
is a solution of (13).

Applied Mathematics 135/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

2
Distinct Roots: ∆ := b − 4ac > 0

In this case, (14) has two distinct roots given by


√ √
−b − ∆ −b + ∆
r1 := and r2 := .
2a 2a
Thus,
y1 (t) := er1 t and y2 (t) := er2 t (15)
are two solutions of (13).

Applied Mathematics 136/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Further, for the Wronskian of the solutions, we have



y1 (t) y2 (t) er1 t er2 t (r1 +r2 )t 1
1
W (t) = 0
0
= r1 t r2 t = e
y1 (t) y2 (t) r1 e r2 e r1 r2

∆ −bt
= − (r1 − r2 )e(r1 +r2 )t = e a 6= 0.
a
This shows that the solutions in (15) are linearly independent solutions of
(13).
Thus, the general solution of the equation is

yc (t) := c1 er1 t + c2 er2 t ,

where c1 and c2 are arbitrary constants.

Applied Mathematics 137/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Alternatively, using the hyperbolic trigonometric functions, i.e.,

ex + e−x ex − e−x
cosh(x) := and sinh(x) := for x ∈ R,
2 2
we can show that
√  √ 
b ∆ b ∆
y1 (t) := e− 2a t cosh t and y2 (t) := e− 2a t sinh t
2a 2a

are also linearly independent solutions of (13).

Applied Mathematics 138/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Example 40
Let us solve
y 00 + y 0 − 2y = 0.
Clearly, the characteristic equation is r 2 + r − 2 = 0 or equivalently
(r + 2)(r − 1) = 0, which implies r1 := −2 and r2 := 1. Since the roots
are distinct, the general solution of the equation is

yc (t) := c1 e−2t + c2 et ,

where c1 and c2 are arbitrary constants.

Applied Mathematics 139/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Example 41
The differential equation

y 00 (t) − y (t) = 0

has the general solution

yc (t) := c1 cosh(t) + c2 sinh(t),

where c1 and c2 are arbitrary constants.

Applied Mathematics 140/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Repeated Roots: ∆ = 0

Then, (14) has a repeated root given by

b
r1 := − .
2a
Hence,
y1 (t) := er1 t (16)
is a solution of (13). Now, define

y2 (t) := er1 t u(t), (17)

where u is a differentiable function to be determined later.

Applied Mathematics 141/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Recall that y10 = r1 y1 , and thus y100 = r12 y1 . Computing the derivatives of
y2 , we get
y20 =er1 t u 0 + r1 er1 t u
=er1 t [u 0 + r1 u],
y200 =er1 t u 00 + 2r1 er1 t u 0 + r12 er1 t u
=er1 t u 00 + 2r1 u 0 + r12 u .
 

Substituting these into the equation gives us

ay200 + by20 + cy2 =aer1 t u 00 + 2r1 u 0 + r12 u + ber1 t [u 0 + r1 u] + cer1 t u


 

=er1 t au 00 + (2ar1 + b) u 0 + (ar12 + br1 + c) u


 
| {z } | {z }
0 0
=aer1 t u 00 .

Applied Mathematics 142/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

If we let u(t) := t, then y2 in (17) takes the form

y2 (t) := er1 t t (18)

and becomes a solution of (13). Computing the Wronskian of y1 and y2 ,


we obtain

y (t) y2 (t) er1 t er1 t t
W (t) = 10

0
= r1 t r1 t

r1 t
y1 (t) y2 (t) r1 e e + r1 e t
rt
e1 0 b
= = e− a t 6= 0,
r1 er1 t er1 t

i.e.,
y1 (t) = er1 t and y2 (t) = er1 t t
are linearly independent.
Thus, the general solution of the equation is

yc (t) := c1 er1 t + c2 er1 t t,

where c1 and c2 are arbitrary constants.

Applied Mathematics 143/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Example 42
Let us solve
y 00 − 4y 0 + 4y = 0.
Clearly, the characteristic equation is r 2 − 4r + 4 = 0 or equivalently
(r − 2)2 = 0, which implies r1,2 := 2. Since the roots are repeated, the
general solution of the equation is

yc (t) := c1 e2t + c2 e2t t,

where c1 and c2 are arbitrary constants.

Applied Mathematics 144/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Roots in Complex Pairs: ∆ < 0

In this case, complex conjugate roots of (14) are given by


√ √
−b − i −∆ −b + i −∆
r1 := and r2 := .
2a 2a
For simplicity, we let

b −∆
α := − and β := , (19)
2a 2a
and define
z1 (t) := e(α−iβ)t and z2 (t) := e(α+iβ)t .

Applied Mathematics 145/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

As in the first case, we can show that z1 and z2 are solutions of (13).
However, they are complex functions. We will now show that their real
and imaginary parts also form solutions. Using the well-known Euler’s
formula eiθ = cos(θ) + i sin(θ), we get

z1 (t) := eαt [cos(βt) − i sin(βt)]

and
z2 (t) := eαt [cos(βt) + i sin(βt)].

Applied Mathematics 146/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Let us define

y1 (t) := eαt cos(βt) and y2 (t) := eαt sin(βt). (20)

Computing the derivatives of y1 , we get

y10 (t) =αeαt cos(βt) − βeαt sin(βt)


=eαt [α cos(βt) − β sin(βt)]

and
y100 (t) =(α2 − β 2 )eαt − 2αβeαt sin(βt)
=eαt (α2 − β 2 ) cos(βt) − 2αβ sin(βt) .
 

Applied Mathematics 147/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Substituting these into (13) and using (19) yield that

ay100 (t) + by10 (t) + cy1 (t) =aeαt (α2 − β 2 ) cos(βt) − 2αβ sin(βt)
 

+ beαt [α cos(βt) − β sin(βt)] + ceαt cos(βt)


=eαt a(α2 − β 2 ) + bα + c cos(βt)
 
| {z }
0

− (2aα + b) sin(βt) = 0,
| {z }
0

which shows that y1 is a solution of (13). Similarly, we can show that y2


is also a solution of (13).

Applied Mathematics 148/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Finally, we show that they are linearly independent, i.e.,



eαt cos(βt) eαt sin(βt)
W (t) =
αeαt cos(βt) − βeαt cos(βt) αeαt sin(βt) + βeαt sin(βt)

eαt cos(βt) eαt sin(βt) 2αt cos(βt)
sin(βt)
= αt αt = βe
−βe cos(βt) βe sin(βt) − cos(βt) sin(βt)
=βe2αt 6= 0.

Thus, y1 and y2 defined in (20) are linearly independent solutions of (13).


Therefore, the general solution of the equation is

yc (t) := c1 eαt cos(βt) + c2 eαt sin(βt),

where c1 and c2 are arbitrary constants.

Applied Mathematics 149/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Example 43
Let us solve
y 00 − 4y 0 + 13y = 0.
Clearly, the characteristic equation is r 2 − 4r + 13 = 0 or equivalently
(r − 2)2 + 32 = 0, which implies r1,2 := 2 ± 3i. Since the roots are in
complex conjugates, the general solution of the equation is

yc (t) := c1 e2t cos(3t) + c2 e2t sin(3t),

where c1 and c2 are arbitrary constants.

Applied Mathematics 150/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Non-Homogeneous Equation

Now, we return to the equation

ay 00 (t) + by 0 (t) + cy (t) = f (t), t ∈ R, (21)

where a, b, c ∈ R with a 6= 0 and f is a continuous function on R.


Suppose that y1 and y2 are two linearly independent solutions of the
associated homogeneous differential equation (14). Suppose that a
solution of (21) is of the form

yp (t) := y1 (t)u1 (t) + y2 (t)u2 (t), (22)

where u1 and u2 are functions to be determined later.

Applied Mathematics 151/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Taking the derivatives, we get

yp0 =[y10 u1 + y20 u2 ] + [y1 u10 + y2 u20 ],


yp00 =[y100 u1 + y200 u2 ] + [y10 u10 + y20 u20 ] + [y1 u10 + y2 u20 ]0 ,

which yields upon substituting into (21) that

f =ayp00 + byp0 + cyp


=a [y100 u1 + y200 u2 ] + [y10 u10 + y20 u20 ] + [y1 u10 + y2 u20 ]0
 

+ b [y10 u1 + y20 u2 ] + [y1 u10 + y2 u20 ] + c[y1 u1 + y2 u2 ]


 

=u1 [ay100 + by10 + cy1 ] +u2 [ay200 + by20 + cy2 ]


| {z } | {z }
0 0
+ a [y10 u10 + y20 u20 ] + [y1 u10 + y2 u20 ]0 + b[y1 u10 + y2 u20 ].
 

Applied Mathematics 152/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

If we assume that  0 0
y1 u1 + y2 u2 =0
y10 u10 + y20 u20 = f ,
a
then by Cramer’s rule, we have
Z t
f (ξ)y2 (ξ)

fy2

0
u1 = − u1 (t) = − dξ
 

aW =⇒ aW (ξ)
Z t (23)
u 0 = fy1
 
u2 (t) = f (ξ)y1 (ξ)
dξ,

2
aW aW (ξ)

where W is the Wronskian of y1 and y2 , i.e., W := y1 y20 − y10 y2 .

Applied Mathematics 153/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Using (23) in (22), we get


Z t Z t
f (ξ)y2 (ξ) f (ξ)y1 (ξ)
yp (t) = −y1 (t) dξ + y2 (t) dξ. (24)
aW (ξ) aW (ξ)

Since the complementary solution of (22) is

yc (t) := c1 y1 (t) + c2 y2 (t), (25)

where c1 and c2 are arbitrary constants. Finally, by (24) and (25), the
general solution of (21) is
Z t Z t
f (ξ)y2 (ξ) f (ξ)y1 (ξ)
y (t) = c1 y1 (t) + c2 y2 (t) −y1 (t) dξ + y2 (t) dξ ,
| {z } aW (ξ) aW (ξ)
complementary solution | {z }
particular solution

where c1 and c2 are arbitrary constants.

Applied Mathematics 154/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Example 44
Let us solve
y 00 + y = 3.
Clearly, the characteristic equation is r 2 + 1 = 0, which implies
r1,2 := ±i. Since the roots are complex conjugates, the complementary
solution of the equation is

yc := c1 cos(t) +c2 sin(t),


| {z } | {z }
y1 y2

where c1 and c2 are arbitrary constants. I

Applied Mathematics 155/297


Ordinary Differential Equations Second-Order Linear Differential Equations with Constant Coefficients

Note that the Wronskian of cos and sin is 1. Using the variation of
parameters formula, we have
Z t Z t
yp (t) := − cos(t) 3 sin(ξ)dξ + sin(t) 3 cos(ξ)dξ

=3 cos2 (t) + 3 sin2 (t) = 3.

Therefore, the general solution of the equation is

y (t) := c1 cos(t) + c2 sin(t) + 3. 

Applied Mathematics 156/297


Initial Value Problems and Boundary Value Problems

Initial Value Problems and Boundary Value Problems

Two main contents of this section are the initial-value problems (IVP)
and the boundary-value problems (BVP).

Applied Mathematics 157/297


Initial Value Problems and Boundary Value Problems

Initial Value Problems

Initial value problems (IVP) for second-order linear differential equations


are of the form
(
ay 00 (t) + by 0 (t) + cy (t) = f (t) for t > 0
(26)
y (0) = α and y 0 (0) = β,

where a, b, c ∈ R with a 6= 0, f is a continuous function on [0, ∞) and


α, β are real numbers.

Applied Mathematics 158/297


Initial Value Problems and Boundary Value Problems

To find the unique solution of (26), we first write the general solution of
the ODE in the first line of (26) and then apply the initial conditions in
the second line to determine for which arbitrary constants these
conditions hold. Say

yc (t) := c1 y1 (t) + c2 y2 (t) + yp (t) for t ≥ 0

is the general solution of (26). Applying the initial conditions, we get


( (
yc (0) =α c1 y1 (0) + c2 y2 (0) =α − yp (0)
=⇒
0
yc (0) =β c1 y10 (0) + c2 y20 (0) =β − yp0 (0)

from which we can solve c1 and c2 .

Applied Mathematics 159/297


Initial Value Problems and Boundary Value Problems

Example 45
Show that the solution of the IVP
(
y 00 + y 0 − 2y = 0 for t > 0
y (0) = 1 and y 0 (0) = −5

is
y (t) := 2e−2t − et for t ≥ 0.

Figure 12: The graphic of the solution y over the interval [0, 1].

Applied Mathematics 160/297


Initial Value Problems and Boundary Value Problems

Example 46
Show that the solution of the IVP
(
y 00 − 4y 0 + 4y = 0 for t > 0
0
y (0) = 1 and y (0) = 5

is
y (t) := e2t + 3e2t t for t ≥ 0.

1
1

Figure 13: The graphic of the solution y over the interval [0, 1].

Applied Mathematics 161/297


Initial Value Problems and Boundary Value Problems

Example 47
Show that the solution of the IVP
(
y 00 − 4y 0 + 13y = 0 for t > 0
0
y (0) = 2 and y (0) = −5

is
y (t) := e2t 2 cos(3t) − 3 sin(3t) for t ≥ 0.
 

2
1

Figure 14: The graphic of the solution y over the interval [0, 1].

Applied Mathematics 162/297


Initial Value Problems and Boundary Value Problems

Example 48
Show that the solution of the IVP
(
y 00 + y = 3 for t > 0
0
y (0) = 5 and y (0) = −1

is
y (t) := 2 cos(t) − sin(t) + 3 for t ≥ 0.

Figure 15: The graphic of the solution y over the interval [0, 1].

Applied Mathematics 163/297


Initial Value Problems and Boundary Value Problems

Example 49
Show that the solution of the IVP
(
y 00 − y = 3 for t > 0
y (0) = −1 and y 0 (0) = −3

is
y (t) := 2 cosh(t) − 3 sinh(t) − 3 for t ≥ 0.

1
−1

Figure 16: The graphic of the solution y over the interval [0, 1].

Applied Mathematics 164/297


Initial Value Problems and Boundary Value Problems

Boundary Value Problems

Before we start off this section, we need to make it very clear that we are
only going to scratch the surface of the topic of BVPs. There is enough
material in the topic of BVPs that we could devote a whole class to it.
The intent of this section is to give a brief (and we mean very brief) look
at the idea of BVPs and to give enough information to allow us to do
some basic partial differential equations in the next chapter.

Applied Mathematics 165/297


Initial Value Problems and Boundary Value Problems

Now, with that out of the way, the first thing that we need to do is to
define just what we mean by a BVP. With IVPs, we had a differential
equation and we specified the value of the solution and an appropriate
number of derivatives at the same point (collectively called “initial
conditions”). For instance, for a second-order differential equation the
initial conditions are as follows.

y (0) = α and y 0 (0) = β.

Applied Mathematics 166/297


Initial Value Problems and Boundary Value Problems

With BVPs, we will have a differential equation and we will specify the
function and/or derivatives at different points, which we’ll call boundary
values. For second-order differential equations, which will be looking at
pretty much exclusively here, any of the following can (and will) be used
for boundary conditions.

y (t1 ) = α and y (t2 ) = β,


0
y (t1 ) = α and y 0 (t2 ) = β,
0
y (t1 ) = α and y (t2 ) = β,
y (t1 ) = α and y 0 (t2 ) = β.

Applied Mathematics 167/297


Initial Value Problems and Boundary Value Problems

As we’ll soon see much of what we know about initial value problems will
not hold here. We can, of course, solve the ODE provided the
coefficients are constant. None of that will change. The changes (and
perhaps the problems) arise when we move from initial conditions to
boundary conditions.

Applied Mathematics 168/297


Initial Value Problems and Boundary Value Problems

Example 50
Show that the solution of the BVP
(
y 00 + 4y = 0 for 0 < t < π
4
y (0) = −2 and y ( π4 ) = 10

is
π
y (t) := −2 cos(2t) + 10 sin(2t) for 0 ≤ t ≤ 4.

π
−2 4

Figure 17: The graphic of the solution y over the interval [0, π4 ].

Applied Mathematics 169/297


Initial Value Problems and Boundary Value Problems

Example 51
Show that solutions of the BVP
(
y 00 + 4y = 0 for 0 < t < π
y (0) = −2 and y (π) = −2

are
y (t) := −2 cos(2t) + c sin(2t) for 0 ≤ t ≤ π,
where c is an arbitrary constant.

π
−2

Figure 18: The graphic of the solutions y over the interval [0, π] for
c = −5, −4, · · · , 5.

Applied Mathematics 170/297


Initial Value Problems and Boundary Value Problems

Example 52
Show that the BVP
(
y 00 + 4y = 0 for 0 < t < π
y (0) = −2 and y (π) = 3

has no solutions.

Applied Mathematics 171/297


Initial Value Problems and Boundary Value Problems

Example 53
Show that the solution of the BVP
(
y 00 + 3y = 0 for 0 < t < π
0
y (0) = 7 and y (π) = 0

is √  √  √ 
y (t) := 7 cos 3t + 7 tan 3π sin 3t for 0 ≤ t ≤ π.

Figure 19: The graphic of the solution y over the interval [0, π].

Applied Mathematics 172/297


Initial Value Problems and Boundary Value Problems

Example 54
Show that the BVP
(
y 00 + 25y = 0 for 0 < t < π
0 0
y (0) = 5 and y (π) = 5

has no solutions.

Applied Mathematics 173/297


Initial Value Problems and Boundary Value Problems

Example 55
Show that solutions of the BVP
(
y 00 + 9y = cos(t) for 0 < t < π
2
y 0 (0) = 3 and y ( π2 ) = −1

are
1 π
y (t) := c cos(3t) + sin(3t) + cos(t) for 0 ≤ t ≤ 2,
8
where c is an arbitrary constant.

−1 π
2

Figure 20: The graphic of the solutions y over the interval [0, π] for
c = −5, −4, · · · , 5.

Applied Mathematics 174/297


Sturm-Liouville Problems

Sturm-Liouville Problems

Legendre’s, Bessel’s and other ODEs of importance in engineering can be


written as a Sturm-Liouville equation
0
p(t)y 0 + [q(t) + λr (t)]y = 0,

(27)

where λ is a real parameter. The boundary value problem consisting of


an ODE (27) and given Sturm-Liouville boundary conditions
(
α1 y (a) + α2 y 0 (a) =0
(28)
β1 y (b) + β2 y 0 (b) =0

is called a “Sturm-Liouville problem”.

Applied Mathematics 175/297


Sturm-Liouville Problems

We shall see further that these problems lead to useful series of


developments in terms of particular solutions of (27) and (28). Crucial in
this section is the notion of orthogonality, which will be discussed later
in this section.
In (27), we suppose that p, q, r and p 0 are continuous on the interval I
with r (t) > 0 for all t ∈ I . Further, we suppose in (28) that α1 , α2 are
given constants, not both zero, and so are β1 , β2 .

Applied Mathematics 176/297


Sturm-Liouville Problems

Clearly, y ≡ 0 is a solution (called as the “trivial solution”) for any λ as


(27) is homogeneous and (28) has zeroes on the right-hand side. This is
of no interest. We want to find eigenfunctions y , that is, solutions of
(27) satisfying (28) without being identically zero. We call a number λ
for which an eigenfunction exists an eigenvalue of the Sturm-Liouville
problem (27) and (28).

Applied Mathematics 177/297


Sturm-Liouville Problems

Example 56
Consider the Sturm-Liouville problem
(
y 00 (t) + λy (t) = 0
y (0) = 0 and y (π) = 0.

Let us find its eigenvalues and eigenfunctions. I

Applied Mathematics 178/297


Sturm-Liouville Problems

To obtain the non-trivial solutions of the EVP, we consider the following


there distinct cases.
Case 1. Let λ > 0. Say λ := µ2 , where µ > 0. In this case, the general
solution of the DE is

yc (t) := c1 cos(µt) + c2 sin(µt),

where c1 and c2 are arbitrary constants. Using the boundary conditions,


we get

 c1 :=0
( 
c1 = 0
=⇒ c2 :=arbitrary
c1 cos(µπ) + c2 sin(µπ) = 0 
µn :=n, n = 1, 2, · · · .

Hence, for n = 1, 2, · · · , we get the eigenvalues λn := n2 and the


eigenfunctions ϕn (t) := sin(nt), where we let c2 := 1 for simplicity. I

Applied Mathematics 179/297


Sturm-Liouville Problems

Case 2. Let λ = 0. In this case, the general solution of the DE is

yc (t) := c1 + c2 t,

where c1 and c2 are arbitrary constants. Applying the boundary


conditions, we get
( (
c1 = 0 c1 :=0
=⇒
c1 + πc2 = 0 c2 :=0,

which leads us to the trivial solution only. I

Applied Mathematics 180/297


Sturm-Liouville Problems

Case 3. Let λ < 0. Say λ := −µ2 , where µ > 0. The general solution of
the DE is
yc (t) := c1 cosh(µt) + c2 sinh(µt),
where c1 and c2 are arbitrary constants. Using the boundary conditions,
we get ( (
c1 = 0 c1 :=0
=⇒
c1 cosh(µπ) + c2 sinh(µπ) = 0 c2 :=0.
Hence, we get the trivial solution.
Thus, the Sturm-Liouville problem has the eigenvalues

λn := n2 for n = 1, 2, · · ·

and the eigenfunctions

ϕn (t) := sin(nt) for n = 1, 2, · · · . 

!!! Insert Graphic Here!!!

Applied Mathematics 181/297


Sturm-Liouville Problems

Example 57
Consider the Sturm-Liouville problem
(
y 00 (t) + λy (t) = 0
y (0) = 0 and y 0 (π) = 0.

Show that its eigenvalues are


 1 2
λn := n − for n = 1, 2, · · ·
2
and eigenfunctions are
 
1
ϕn (t) := sin n − t for n = 1, 2, · · · .
2

!!! Insert Graphic Here!!!

Applied Mathematics 182/297


Sturm-Liouville Problems

Example 58
Consider the Sturm-Liouville problem
(
y 00 (t) + λy (t) = 0
y 0 (0) = 0 and y 0 (π) = 0.

Show that its eigenvalues are

λn := n2 for n = 0, 1, · · ·

and eigenfunctions are

ϕn (t) := cos(nt) for n = 0, 1, · · · .

!!! Insert Graphic Here!!!

Applied Mathematics 183/297


Sturm-Liouville Problems

Example 59
Consider the periodic Sturm-Liouville problem
(
y 00 (t) + λy (t) = 0
y (−π) = y (π) and y 0 (−π) = y 0 (π).

Show that its eigenvalues are

λn := n2 for n = 0, 1, · · ·

and eigenfunctions are

ϕ2n (t) := cos(nt) and ϕ2n+1 (t) := sin(nt) for n = 0, 1, · · · .

!!! Insert Graphic Here!!!

Applied Mathematics 184/297


Sturm-Liouville Problems

Fourier Series

Definition 34 (Orthogonality)
Functions f1 , f2 , · · · defined on some interval I : a < t < b are called as
“orthogonal” on this interval with respect to the weight function µ
(with µ(t) > 0 for all t ∈ I ) if the inner product satisfies
Z b
hfm , fn i := fm (ξ)fn (ξ)µ(ξ)dξ = 0 for all distinct m and n.
a

The norm of a function f is defined by


s
Z b
2
kf k := f (ξ) µ(ξ)dξ.
a

If the weight function is 1, then we briefly say that f1 , f2 , · · · are


orthogonal instead of orthogonal with respect to 1. Further, if f1 , f2 , · · ·
are orthogonal and all have norm 1, then they are called “orthonormal”.

Applied Mathematics 185/297


Sturm-Liouville Problems

Let f0 , f1 , · · · be an orthogonal set of functions with respect to the weight


function µ on the interval I : a < t < b. Suppose that the function f
can be represented by a convergent series

X
f (t) = am fm (t) = a0 f0 (t) + a1 f1 (t) + · · · , (29)
m=0

where a0 , a1 , · · · are scalars. This is called as the “orthogonal


expansion” or the “generalized Fourier series”. If f0 , f1 , · · · are
eigenfunctions of a Sturm-Liouville problem, then we call (29) as an
“eigenfunction expansion”.

Applied Mathematics 186/297


Sturm-Liouville Problems

Note that we have


Z b
hf , fn i = f (ξ)fn (ξ)µ(ξ)dξ
a
Z ∞
b X
= am fm (ξ)fn (ξ)µ(ξ)dξ
a m=0

X Z b
= am fm (ξ)fn (ξ)µ(ξ)dξ
m=0 a

X∞
= am hfm , fn i.
m=0

Applied Mathematics 187/297


Sturm-Liouville Problems

Because of the orthogonality, all the terms inside the sum are zero except
when m = n. Hence, the infinite series reduces to

hf , fn i = an hfn , fn i = an kfn k2 ,

which yields
hf , fn i
an = , n = 0, 1, · · · (30)
kfn k2
provided that kfn k =
6 0.

Applied Mathematics 188/297


Sturm-Liouville Problems

Recalling (29), we explicitly have

f (t) = a0 f0 (t) + a1 f1 (t) + · · · ,

where Rb
f (ξ)fn (ξ)µ(ξ)dξ
an = Ra b 2 , n = 0, 1, · · ·
a
fn (ξ) µ(ξ)dξ
provided that kfn k =
6 0.

Applied Mathematics 189/297


Sturm-Liouville Problems

Definition 35 (Completeness)
Let f1 , f2 , · · · be a sequence of orthogonal functions on an interval
I : a < t < b and S be a set of functions defined on I . If every function
f ∈ S can be approximated arbitrarily closely by a linear combination of
the functions f1 , f2 , · · · , then f1 , f2 , · · · is said to be “complete” in the
set S. More precisely, let f ∈ S, if for every ε > 0, there exist a positive
integer m, scalars a1 , a2 , · · · , am and functions f1 , f2 , · · · , fm such that
m
X

f − ak k < ε.
f

k=0

Applied Mathematics 190/297


Sturm-Liouville Problems

Lemma 1 (Completeness of the Trigonometric System)


The trigonometric system 1, cos(mt), sin(mt) for m = 1, 2, · · · is
orthogonal on the interval I : −π < t < π (hence, on any interval of
length 2π because of the periodicity), i.e., the inner product of any two
of these functions over the period is 0. Explicitly, we have the following.

1. −π cos(mξ)dξ = 0 for any m.

2. −π sin(mξ)dξ = 0 for any m.
(
Rπ π, m = n
3. −π cos(mξ) cos(nξ)dξ =
0, m 6= n.
(
Rπ π, m = n
4. −π sin(mξ) sin(nξ)dξ =
0, m 6= n.

5. −π cos(mξ) sin(nξ)dξ = 0 for any m and n.
Further, the trigonometric system is complete in the set of piecewise
smooth functions (i.e., functions, which are piecewise continuous
together with their derivatives) defined on the closed interval [−π, π].

Applied Mathematics 191/297


Sturm-Liouville Problems

Theorem 24 (Fourier Series)

Let f be piecewise smooth and be periodic of period 2π. Then, f can be


represented by its Fourier series

X  
a0 + am cos(mt) + bm sin(mt) ,
m=1

where Z π
1
a0 = f (ξ)dξ,
2π −π
1 π
Z
am = f (ξ) cos(mξ)dξ, m = 1, 2, · · · ,
π −π
1 π
Z
bm = f (ξ) sin(mξ)dξ, m = 1, 2, · · · .
π −π
Further, the series converges to f (t) if f is continuous at t and to
f (t + )+f (t − )
2 if f has a jump type discontinuity at t.

Applied Mathematics 192/297


Sturm-Liouville Problems

Example 60

Consider the following 2π-periodic function


(
1, 0<t<π
f (t) :=
(−1), −π < t < 0

whose graphic is given below.


1

−2π −π π 2π

−1

Figure 21: The graphic of f over the interval [−2π, 2π]. I

Applied Mathematics 193/297


Sturm-Liouville Problems

Computing the Fourier coefficients, we see that


Z 0 Z π 
1
a0 = (−1)dξ + 1dξ = 0,
2π −π 0
Z 0 Z π 
1
am = (−1) cos(mξ)dξ + cos(mξ)dξ = 0,
π −π 0
Z 0 Z π 
1
bm = (−1) sin(mξ)dξ + sin(mξ)dξ
π −π 0
 0 π 
1 1 1
= cos(mξ) − cos(mξ)

π m −π m 0
2  2 
1 − (−1)m
 
= 1 − cos(mπ) =
mπ mπ
for m = 1, 2, · · · . I

Applied Mathematics 194/297


Sturm-Liouville Problems

Substituting these back into the series, we get



2 X 1
1 − (−1)m sin(mt)

f (t) ∼
π m=1 m

or equivalently

4X 1 
f (t) ∼ sin (2k − 1)t ,
π 2k − 1
k=1

which is the desired series. I

Applied Mathematics 195/297


Sturm-Liouville Problems

This Fourier series converges to the 2π-periodic function




 0, t=π

1, 0<t<π
g (t) :=
0,
 t=0

(−1), −π < t < 0

whose graphic is given below.


1

−2π −π π 2π

−1

Figure 22: The graphic of g over the interval [−2π, 2π]. 

Applied Mathematics 196/297


Sturm-Liouville Problems

Theorem 25 (Fourier Series for 2L-Periodic Functions)

The Fourier series of a function of period 2L is of the form


∞     
X π π
f (t) ∼ a0 + am cos mt + bm sin mt
m=1
L L

with coefficients
Z L
1
a0 := f (ξ)dξ,
2L −L
1 L
Z  
π
am := f (ξ) cos mξ dξ, m = 1, 2, · · · ,
L −L L
Z L  
1 π
bm := f (ξ) sin mξ dξ, m = 1, 2, · · · .
L −L L

Applied Mathematics 197/297


Sturm-Liouville Problems

Example 61
Consider the following 2-periodic function
(
t, 0 < t < 1
f (t) :=
2, −1 < t < 0

whose graphic is given below.


2

−2 −1 1 2

Figure 23: The graphic of f over the interval [−2, 2]. I

Applied Mathematics 198/297


Sturm-Liouville Problems

The series
∞ 
1 − (−1)m 2 − (−1)m
   
5 X
− cos(πmt) + sin(πmt)
4 m=1 π 2 m2 πm

is the Fourier series for f . I

Applied Mathematics 199/297


Sturm-Liouville Problems

This series converges to the 2-periodic function


3
2, t = 1


t, 0 < t < 1
g (t) :=
1, t = 0


2, −1 < t < 0

whose graphic is given below.


2

−2 −1 1 2

Figure 24: The graphic of g over the interval [−2, 2]. 

Applied Mathematics 200/297


Sturm-Liouville Problems

Theorem 26 (Fourier Series for Even Functions)

The Fourier series of an even function of period 2L is a Fourier Cosine


series of the form
∞  
X π
f (t) ∼ a0 + am cos mt
m=1
L

with coefficients
Z L
1
a0 := f (ξ)dξ,
L 0
Z L  
2 π
am := f (ξ) cos mξ dξ, m = 1, 2, · · · .
L 0 L

Applied Mathematics 201/297


Sturm-Liouville Problems

Theorem 27 (Fourier Series for Odd Functions)

The Fourier series of an odd function of period 2L is a Fourier Sine


series of the form
∞  
X π
f (t) ∼ bm sin mt
m=1
L
with coefficients
Z L  
2 π
bm := f (ξ) sin mξ dξ, m = 1, 2, · · · .
L 0 L

Applied Mathematics 202/297


Sturm-Liouville Problems

Example 62
Consider the following 2-periodic function

f (t) := t for |t| < 1

whose graphic is given below.


1

−4 −3 −2 −1 1 2 3 4

−1

Figure 25: The graphic of f over the interval [−4, 4]. I

Applied Mathematics 203/297


Sturm-Liouville Problems

The series

2 X (−1)m
− sin(πmt)
π m=1 m
is the Fourier Sine series for f . I

Applied Mathematics 204/297


Sturm-Liouville Problems

This series converges to the 2-periodic function



0, t = 1

g (t) := t, −1 < t < 1

0, t = 0

whose graphic is given below.


1

−4 −3 −2 −1 1 2 3 4

−1

Figure 26: The graphic of g over the interval [−4, 4]. 

Applied Mathematics 205/297


Sturm-Liouville Problems

Example 63
Consider the 2L-periodic function defined as
(
t(L − t), 0≤t≤L
f (t) :=
t(L + t), −L ≤ t ≤ 0

whose graphic is given below.

−3L −L L 3L

Figure 27: The graphic of f over the interval [−3L, 3L]. I

Applied Mathematics 206/297


Sturm-Liouville Problems

Show that

4L2 X 1 − (−1)m
 
π
sin mt for − L ≤ t ≤ L
π 3 m=1 m3 L

or equivalently

8L2 X
 
1 π
sin (2k − 1)t for − L ≤ t ≤ L
π3 (2k − 1)3 L
k=1

is the Fourier series of f . 

Applied Mathematics 207/297


Partial Differential Equations

Partial Differential Equations

Definition 36 (Partial Differential Equations)

A “partial differential equation” (in short, PDE) is an equation


containing one unknown function (dependent variable) of at least two
independent variables together with its derivatives.

Definition 37 (Order of the PDE)

The highest-order derivative appearing in the equation is called the


“order” of a differential equation.

Applied Mathematics 208/297


Partial Differential Equations

Example 64

Let us present some important partial differential equations.


2
∂u
1. One-Dimensional Heat Equation: ∂t = c 2 ∂∂xu2 .
∂2u 2
2. One-Dimensional Wave Equation: ∂t 2 = c 2 ∂∂xu2 .
∂2u ∂2u
3. Two-Dimensional Laplace Equation: ∂x 2 + ∂y 2 = 0.
2 2
∂ u ∂ u
4. Two-Dimensional Poisson Equation: ∂x 2 + ∂y 2 = f (x, y ).
2
 2

∂ u ∂ u ∂2u
5. Two-Dimensional Wave Equation: ∂t 2 = c2 ∂x 2 + ∂y 2 .
∂2u ∂2u ∂2u
6. Three-Dimensional Laplace Equation: ∂x 2 + ∂y 2 + ∂z 2 = 0.
Here, c is a positive constant, t is time while x, y , z are Cartesian
coordinates, and the dimension is the number of these Cartesian
variables in the equation.

In this section, we will consider Equation 1, Equation 2 and Equation 3.

Applied Mathematics 209/297


Partial Differential Equations

Consider the second-order quasilinear PDE

Auxx + Buxy + Cuyy = F (x, y , u, ux , uy ), (x, y ) ∈ R, (31)

where A, B and C are only functions of x and y , and R is a open


connected set in R2 . Such equations are classified by checking
∆ := B 2 − 4AC as follows.
1. Hyperbolic: ∆ > 0.
2. Parabolic: ∆ = 0.
3. Elliptic: ∆ < 0.
So, we can say that the heat equation is parabolic, the wave equation
is hyperbolic while the Laplace equation is elliptic.

Applied Mathematics 210/297


Partial Differential Equations

For a PDE of the particular form (31) to be meaningful, we need


boundary conditions such as the followings.
1. Dirichlet Boundary Condition: u = f (x, y ) for (x, y ) ∈ ∂R, where
∂R denotes the boundary of R.
∂u
2. Neumann Boundary Condition: ∂n = f (x, y ) for (x, y ) ∈ ∂R,
where n denotes the normal vector to the boundary of R.
∂u
3. Robin Boundary Condition: αu + β ∂n = f (x, y ) for (x, y ) ∈ ∂R.
4. Mixed Boundary Condition: This case includes at least two of the
conditions above. For instance, u = f (x, y ) for (x, y ) ∈ J1 and
∂u
∂n = g (x, y ) for (x, y ) ∈ J2 , where J1 and J2 nonintersecting curves with
J1 ∪ J2 = ∂R.

Applied Mathematics 211/297


Partial Differential Equations

Theorem 28 (Superposition Principle)


Let u1 and u2 be solutions of the second-order linear homogeneous
PDE
Auxx + Buxy + Cuyy + Dux + Euy + Fu = 0,
where A, B, C , D, E and F are only functions of x and y , in some region
R. Then,
u := c1 u1 + c2 u2 ,
where c1 and c2 are arbitrary constants, is also a solution.

Applied Mathematics 212/297


Partial Differential Equations Heat Equation

Heat Equation

In this section, we will study finding the solution of the one-dimensional


heat equation
∂u ∂2u
= c2 2 , (32)
∂t ∂x
K
where c 2 = σρ , which gives us the temperature u(x, t) in a body of
homogeneous rod. Here, c is the thermal diffusivity, K is the thermal
conductivity, σ is the specific heat and ρ is the density of the metal rod.
This equation is also known as “diffusion equation”.

Applied Mathematics 213/297


Partial Differential Equations Heat Equation

We consider the temperature in a long thin metal rod of constant cross


section and homogeneous material, which is oriented along the
hotizontal-axis and is perfectly insulated laterally. So that the heat can
only flow in the hotizontal-direction. See Figure 28.

Figure 28: The rod under consideration.

Applied Mathematics 214/297


Partial Differential Equations Heat Equation

We shall solve (32) for some important types of boundary and initial
conditions. We begin with the case in which the ends x = 0 and x = L of
the bar are kept at temperature zero. So that we have the boundary
conditions

u(0, t) = 0 and u(L, t) = 0 for all t ≥ 0. (33)

Furthermore, the initial temperature in the bar at time t = 0 is given as


f (x). Hence, we have the initial condition

u(x, 0) = f (x) for 0 ≤ x ≤ L. (34)

Here, we must have f (0) = 0 and f (L) = 0 for consistency with (33).

Applied Mathematics 215/297


Partial Differential Equations Heat Equation

We shall determine a solution u(x, t) of (32) satisfying (33) and (34).


Our method will use the technique of separation of variables, which will
be the main idea for solving the wave equation and the Laplace equation.
The steps of the method are outlined below.
Step 1. We let u(x, t) := ϕ(x)ψ(t) to obtain from the PDE (32) two
ODEs, one for ϕ and one for ψ.
Step 2. We determine solutions of these ODEs satisfying the boundary
conditions (33) and the initial condition (34).
Step 3. Finally, using Fourier series, we compose the solutions in Step 2
to obtain the solution of (32) satisfying (33) and (34).

Applied Mathematics 216/297


Partial Differential Equations Heat Equation

Step 1: Two ODEs from the Heat Equation

In the method of separating variables (or the product method), we


determine solution of the heat equation (32) of the form

u(x, t) := ϕ(x)ψ(t), (35)

which is a product of two functions, each depending only on one of the


variables x and t. This is a powerful general method that has various
applications in engineering mathematics. Differentiating (35), we obtain

∂u ∂2u
= ϕψ̇ and = ϕ00 ψ,
∂t ∂x 2
where dots denote the derivatives with respect to t and the primes the
derivatives with respect to x.

Applied Mathematics 217/297


Partial Differential Equations Heat Equation

By substituting this into (32), we have

ϕψ̇ = c 2 ϕ00 ψ.

Dividing by c 2 ϕψ and simplifying gives

ϕ00 1 ψ̇
= 2 .
ϕ c ψ
The variables are separated now, the left-hand side depends only on x
and the right-hand side only on t. Hence, both sides must be constant
because if they were variables, then changing one would affect only one
side, leaving the other side unaltered. Thus,

ϕ00 1 ψ̇
= 2 = −λ,
ϕ c ψ
where λ is called as the “separation constant”.

Applied Mathematics 218/297


Partial Differential Equations Heat Equation

Multiplying by the denominators gives immediately two ODEs

ϕ00 + λϕ = 0 (36)

and
ψ̇ + c 2 λψ = 0. (37)

Applied Mathematics 219/297


Partial Differential Equations Heat Equation

Step 2: Fulfilling the Boundary Conditions

We now determine solutions ϕ and ψ of (36) and (37), respectively, so


that u = ϕψ satisfies the boundary conditions (33), that is,

u(0, t) = ϕ(0)ψ(t) = 0 and for all t ≥ 0.


u(L, t) = ϕ(L)ψ(t) = 0
(38)
we first solve the spatial problem (36). If ψ ≡ 0, then u = ϕψ ≡ 0, which
is of no interest. Hence, ψ 6≡ 0 and then by (38), we get

ϕ(0) = 0 and ϕ(L) = 0. (39)

Applied Mathematics 220/297


Partial Differential Equations Heat Equation

We will show that λ must be positive.


Case 1. Let λ < 0, then we may take λ := −µ2 , where µ > 0. Hence,
the general solution of (36) is

ϕ(x) := A cosh(µx) + B sinh(µx),

where A and B are arbitrary constants. Applying the boundary conditions


in (39) yields A := 0 and B := 0, which leads us to the trivial solution
ϕ ≡ 0.

Applied Mathematics 221/297


Partial Differential Equations Heat Equation

Case 2. Let λ = 0, then the general solution of (36) is

ϕ(x) := A + Bx,

where A and B are arbitrary constants. An application of the boundary


conditions in (39) yields A := 0 and B := 0, which again leads us to the
trivial solution ϕ ≡ 0.

Applied Mathematics 222/297


Partial Differential Equations Heat Equation

Case 3. Let λ > 0, then we may take λ := µ2 , where µ > 0. In this


case, the general solution of (36) is

ϕ(x) := A cos(µx) + B sin(µx),

where A and B are arbitrary constants. Using (39), we get

A=0 and B sin(µL) = 0.

We must take B 6= 0 since otherwise ϕ ≡ 0. Hence, sin(µL) = 0, which


yields µn := πL n for n = 1, 2, · · · . Hence, the eigenvalues are
 2
π
λn := n , n = 1, 2, · · ·
L

and the eigenfunctions are



π
ϕn (x) := sin nx , n = 1, 2, · · · .
L

Applied Mathematics 223/297


Partial Differential Equations Heat Equation

π
2
Now, we turn to the solution of the problem (37) with λn := Ln , i.e.,
 2
π
ψ̇ + c n ψ=0
L

whose general solution is


π 2
ψn (t) := Cn e−(c L n) t ,

where C1 , C2 , · · · are arbitrary constants to be determined later.

Applied Mathematics 224/297


Partial Differential Equations Heat Equation

Hence, the functions


 
π π 2
un (x, t) := ϕn (x)ψn (t) = Cn sin nx e−(c L n) t , n = 1, 2, · · · (40)
L

are solutions of the heat equation (32) satisfying (33). These are the
eigenfunctions of the problem corresponding to the eigenvalues c πL n
for n = 1, 2, · · · .

Applied Mathematics 225/297


Partial Differential Equations Heat Equation

Step 3: Complete Solution by Fourier Series


So far, we have the solutions in (40) satisfying the boundary conditions
(33). To obtain a solution that also satisfies the initial condition (34), we
consider the series of eigenfunctions
∞ ∞  
X X π π 2
u(x, t) = un (x, t) = Cn sin nx e−(c L n) t . (41)
n=1 n=1
L

Letting t = 0 in (41) and using (34), we have


∞  
X π
u(x, 0) = Cn sin nx = f (x).
n=1
L

Hence, for (41) to satisfy (34), Cn ’s must be the coefficients of the


Fourier Sine series of f by Theorem 27, i.e.,

2 L
Z  
π
Cn := f (ξ) sin nξ dξ, n = 1, 2, · · · .
L 0 L

Applied Mathematics 226/297


Partial Differential Equations Heat Equation

Example 65
Show that the solution of the heat equation

1
ut = 64 uxx for 0 < x < L and t > 0

u(0, t) = 0 and u(L, t) = 0 for t > 0

u(x, 0) = x(L − x) for 0 ≤ x ≤ L

is

4L2 X 1 − (−1)n
 
π π 2
u(x, t) := 3 3
sin nx e−( 8L n) t .
π n=1 n L

!!! Insert Graphic Here !!!

Applied Mathematics 227/297


Partial Differential Equations Heat Equation

Example 66
Show that the solution of the heat equation

1
ut = 64 uxx for 0 < x < L and t > 0

u(0, t) = 0 and u(L, t) = 0 for t > 0
u(x, 0) = sin πL x + 3 sin 5 πL x − 8 sin 17 πL x
   
for 0 ≤ x ≤ L

is    
π π 2 π π 2
u(x, t) := sin x e−( 8L ) t + 3 sin 5 x e−(5 8L ) t
L L
 
π π 2
− 8 sin 17 x e−(17 8L ) t .
L

!!! Insert Graphic Here !!!

Applied Mathematics 228/297


Partial Differential Equations Heat Equation

Example 67
Show that the solution of the heat equation

1
ut = 64 uxx for 0 < x < L and t > 0

ux (0, t) = 0 and ux (L, t) = 0 for t > 0
u(x, 0) = 1 − cos 3 πL x
 
for 0 ≤ x ≤ L

is  
π π 2
u(x, t) := 1 − cos 3 x e−(3 8L ) t .
L

!!! Insert Graphic Here !!!

Applied Mathematics 229/297


Wave Equation

Wave Equation

The model of a vibrating elastic string (a violin string, for instance)


consists of the one-dimensional wave equation

∂2u 2
2∂ u
= c , (42)
∂t 2 ∂x 2
where c 2 := Tρ . Here, c is the velocity of propagation with a tension
of T and a mass density of ρ.

Applied Mathematics 230/297


Wave Equation

Since the string is fastened at the ends x = 0 and x = L, we have the


two boundary conditions

u(0, t) = 0 and u(L, t) = 0 for all t ≥ 0. (43)

Furthermore, the form of the motion of the string will depend on its
initial deflection (the deflection at time t = 0) f (x), and on its initial
velocity (the velocity at time t = 0) g (x). Hence, we have the two
initial condition

u(x, 0) = f (x) and ut (x, 0) = g (x) for 0 ≤ x ≤ L. (44)

We will find a solution of the PDE (42) satisfying (43) and (44). The
technique applied here is almost same as the one used for the heat
equation.

Applied Mathematics 231/297


Wave Equation

Step 1: Two ODEs from the Wave Equation

Substitution of a product u(x, t) = ϕ(x)ψ(t) into (42) gives

ϕψ̈ = c 2 ϕ00 ψ,
2
d2 ψ
where ϕ00 := ddxϕ2 and ψ̈ := dt 2 . To separate the variables, we divide by
c 2 ϕψ, which yields
ϕ00 1 ψ̈
= 2 .
ϕ c ψ
The left-hand side depends only on x and the right-hand side only on t,
so both sides must be equal to the so-called separation constant λ, i.e.,

ϕ00 1 ψ̈
= 2 = −λ. (45)
ϕ c ψ

Applied Mathematics 232/297


Wave Equation

We can show that for λ < 0 and λ = 0, the only solution u = ϕψ


satisfying (43) is the trivial solution, i.e., u ≡ 0. For λ = µ2 > 0, where
µ > 0, we have from (45) that

ϕ00 1 ψ̈
= 2 = −µ2 .
ϕ c ψ
Multiplication by the denominators gives immediately the two ODEs

ϕ00 + µ2 ϕ = 0 (46)

and
ψ̈ + c 2 µ2 ψ = 0. (47)

Applied Mathematics 233/297


Wave Equation

Step 2: Fulfilling the Boundary Conditions

We first solve (46). A general solution is

ϕ(x) := A cos(µx) + B sin(µx), (48)

where A and B are constants. From the boundary conditions in (43), it


follows that

u(0, t) = ϕ(0)ψ(t) = 0 and u(L, t) = ϕ(L)ψ(t) = 0.

We only require ϕ(0) = 0 and ϕ(L) = 0 since ψ ≡ 0 would yield u ≡ 0.


Hence, ϕ(0) = A = 0 by (48) and then ϕ(L) = B sin(µL) = 0 with B 6= 0
(to avoid ϕ ≡ 0), we get sin(µL) = 0, which implies µn := πL n for
n = 1, 2, · · · . Letting B := 1, we obtain the solutions ϕn (x) := sin πL nx


of (46) satisfying (43)

Applied Mathematics 234/297


Wave Equation

π
Now, we solve (47). For µn := L n, as just obtained, (47) becomes
 2
π
ψ̈ + c n ψ=0
L

whose general solution is


   
π π
ψn (t) := Cn cos c nt + Dn sin c nt ,
L L

where C1 , C2 , · · · and D1 , D2 , · · · are constants.

Applied Mathematics 235/297


Wave Equation

Therefore, the functions


     
π π π
un (x, t) := sin nx Cn cos c nt +Dn sin c nt , n = 1, 2, · · ·
L L L
(49)
are solutions of the wave equation (42) satisfying (43). The functions in
(49) are the eigenfunctions of the problem corresponding to the
eigenvalues c πL n for n = 1, 2, · · · .

Applied Mathematics 236/297


Wave Equation

Step 3: Complete Solution by Fourier Series

So far, we have the solutions in (49) satisfying the boundary conditions


(43). To obtain a solution that also satisfies the initial condition (44), we
consider the series of eigenfunctions
∞      
X π π π
u(x, t) := sin nx Cn cos c nt + Dn sin c nt . (50)
n=1
L L L

Applied Mathematics 237/297


Wave Equation

Letting t = 0 in (50) and using (44), we have


∞  
X π
u(x, 0) = Cn sin nx = f (x).
n=1
L

Hence, for (50) to satisfy the first condition in (44), Cn ’s must be the
coefficients of the Fourier Sine series of f by Theorem 27, i.e.,

2 L
Z  
π
Cn := f (ξ) sin nξ dξ, n = 1, 2, · · · .
L 0 L

Applied Mathematics 238/297


Wave Equation

Letting t = 0 after differentiating (50) with respect to t and using (44),


we have
∞  
X π π
ut (x, 0) = Dn c n sin nx = g (x).
n=1
L L

Hence, for (50) to satisfy the second condition in (44), Dn ’s must be the
coefficients of the Fourier Sine series of g by Theorem 27, i.e.,
Z L  
2 π
Dn := g (ξ) sin nξ dξ, n = 1, 2, · · · .
cπn 0 L

Applied Mathematics 239/297


Wave Equation

Example 68
Show that the solution of the wave equation

utt = 4uxx for 0 < x < L and t > 0

u(0, t) = 0 and u(L, t) = 0 for t > 0

u(x, 0) = x(L − x) and ut (x, 0) = x for 0 ≤ x ≤ L

is

L2 X 4 1 − (−1)n
   
π π
u(x, t) := 2 sin n x cos 2n t
π n=1 L π n3 L
(−1)n
 
π
− 2
sin 2n t .
n L

!!! Insert Graphic Here !!!

Applied Mathematics 240/297


Wave Equation

Example 69
Show that the solution of the wave equation

utt = 4uxx for 0 < x < L and t > 0

u(0, t) = 0 and u(L, t) = 0 for t > 0
u(x, 0) = 4 sin 5 πL x and ut (x, 0) = 3 sin 14 πL x
  
for 0 ≤ x ≤ L

is
       
π π 3L π π
u(x, t) := 4 sin 5 x cos 10 t + sin 14 x sin 28 t .
L L 28π L L

!!! Insert Graphic Here !!!

Applied Mathematics 241/297


Wave Equation

Example 70
Show that the solution of the wave equation

utt = 4uxx for 0 < x < L and t > 0

ux (0, t) = 0 and u(L, t) = 0 for t > 0
π π
  
u(x, 0) = 2 cos 9 2L x and ut (x, 0) = 3 cos 15 2L x for 0 ≤ x ≤ L

is
       
π π L π π
u(x, t) := 2 cos 9 x cos 9 t + cos 15 x sin 15 t .
2L L 5π 2L L

!!! Insert Graphic Here !!!

Applied Mathematics 242/297


Laplace Equation

Laplace Equation

In this section, we will consider the Laplace equation

∂2u ∂2u
2
+ 2 = 0, (51)
∂x ∂y
which can be obtained from two-dimensional heat equation
 2
∂2u

∂u ∂ u
= c2 +
∂t ∂x 2 ∂y 2
∂u
by taking the time as constant. So that ∂t ≡ 0, which leads us to (51).

Applied Mathematics 243/297


Laplace Equation

We will consider here (51) under Dirichlet boundary condition

u(x, 0) = 0 and u(x, H) = f (x) for 0 ≤ x ≤ L (52)

and
u(0, y ) = 0 and u(L, y ) = 0 for 0 ≤ y ≤ H. (53)
See Figure 29.

f (x)

H
0 0

Figure 29: The plate under consideration.

Applied Mathematics 244/297


Laplace Equation

Step 1: Two ODEs from the Laplace Equation

We will solve this problem again by using the method of separation of


variables. Substituting u(x, y ) = ϕ(x)ψ(y ) into (51), dividing by ϕψ,
and then
ϕ00 ψ̈
= − = −λ, (54)
ϕ ψ
where λ is the so-called separation constant. We can show that for
λ < 0 and λ = 0, the only solution u = ϕψ satisfying (53) is the trivial
solution, i.e., u ≡ 0. For λ = µ2 > 0, where µ > 0, we have from (54)
that
ϕ00 ψ̈
= − = −µ2 .
ϕ ψ

Applied Mathematics 245/297


Laplace Equation

Step 2: Fulfilling the Boundary Conditions

From this and (53), we get the ODE

ϕ00 + µ2 ϕ = 0

with the boundary conditions

ϕ(0) = 0 and ϕ(L) = 0.

This gives us µn := πL n and ϕn (x) := sin πL nx for n = 1, 2, · · · .




Applied Mathematics 246/297


Laplace Equation

π
Next, using µn := Ln in the second ODE, we get
 2
π
ψ̈ − n ψ=0
L

with the initial condition


ψ(0) = 0
whose solution is  
π
ψn (y ) := Cn sinh ny ,
L
where C1 , C2 , · · · are constants to be determined later.

Applied Mathematics 247/297


Laplace Equation

Thus, for n = 1, 2, · · · , we see that


   
π π
un (x, y ) := Cn sin nx sinh ny
L L

π
2
are the eigenfunctions for the eigenvalues λn := Ln .

Applied Mathematics 248/297


Laplace Equation

Step 3: Complete Solution by Fourier Series

Defining
∞    
X π π
u(x, y ) := Cn sin nx sinh ny
n=1
L L

and using (52), we see that


∞    
X π π
u(x, H) = Cn sin nx sinh nH = f (x).
n=1
L L

By Theorem 27, the coefficients of the solution u are given by


Z L  
2 π
Cn := f (ξ) sin nξ dξ, n = 1, 2, · · · .
L sinh πL nH 0

L

Applied Mathematics 249/297


Laplace Equation

Example 71
Show that the solution of the Laplace equation
 2
∂ u ∂2u
 ∂x 2 + ∂y 2 = 0 for 0 < x, y < 2 

u(x, 0) = 0 and u(x, 2) = sin 3 π2 x for 0 ≤ x ≤ 2

u(0, y ) = 0 and u(2, y ) = 0 for 0 ≤ y ≤ 2

is    
1 π π
u(x, y ) := sin 3 x sinh 3 y .
sinh(3π) 2 2

!!! Insert Graphic Here!!!

Applied Mathematics 250/297


Laplace Equation

Example 72
Show that the solution of the Laplace equation
 2
∂ u ∂2u
 ∂x 2 + ∂y 2 = 0 for 0 < x, y < 2

u(x, 0) = 0 and u(x, 2) = 0 for 0 ≤ x ≤ 2

u(0, y ) = − sin π2 y and u(2, y ) = 0 for 0 ≤ y ≤ 2
 

is    
1 π π
u(x, y ) := − sinh π − x sin y .
sinh(π) 2 2

!!! Insert Graphic Here!!!

Applied Mathematics 251/297


Laplace Equation

Example 73
Show that the solution of the Laplace equation
 2
∂ u ∂2u
 ∂x 2 + ∂y 2 = 0 for 0 < x, y < 2

u(x, 0) = 3 sin 7 π2 x and u(x, 2) = sin 3 π2 x

for 0 ≤ x ≤ 2

u(0, y ) = − sin π2 y and u(2, y ) = 0 for 0 ≤ y ≤ 2
 

is    
1 π π
u(x, y ) := sin 3 x sinh 3 y
sinh(3π) 2 2
   
3 π π
+ sin 7 x sinh 7π − 7 y
sinh(7π) 2 2
   
1 π π
− sinh π − x sin y .
sinh(π) 2 2

!!! Insert Graphic Here!!!

Applied Mathematics 252/297


Fourier Integral

Fourier Integral

We develop the Fourier integral with an intuitive approach. However, at


the conclusion of our approach, we state the Fourier integral theorem.
The Fourier integral theorem states all the sufficient conditions necessary
for the Fourier integral representation of a function. We now start with
our intuitive approach.

Applied Mathematics 253/297


Fourier Integral

Suppose we have an arbitrary smooth function fL on the interval [−L, L].


Then, we may represent fL by the Fourier series
∞     
X π π
fL (t) = a0 + am cos mt + bm sin mt ,
m=1
L L

where Z L
1
a0 := f (ξ)dξ,
2L −L
1 L
Z  
π
am := f (ξ) cos mξ dξ, m = 1, 2, · · · ,
L −L L
1 L
Z  
π
bm := f (ξ) sin mξ dξ, m = 1, 2, · · · .
L −L L
We will try to figure out what happens when L → ∞.

Applied Mathematics 254/297


Fourier Integral

π π
Let ω := L m, then ∆ω := L. Thus, we have
Z L ∞  Z L 
1 X 1
fL (t) = f (ξ)dξ + f (ξ) c(ωξ)dξ c(ωt)
2L −L m=1
L −L
 Z L  
1
+ f (ξ) s(ωξ)dξ s(ωt)
L −L
Z L ∞ Z L 
1 1X
= f (ξ)dξ + f (ξ) c(ωξ)dξ c(ωt)
2L −L π m=1 −L
Z L   (55)
+ f (ξ) s(ωξ)dξ s(ωt) ∆ω,
−L

where we denote cos and sin with their first letters above.

Applied Mathematics 255/297


Fourier Integral

Now, let L → ∞, and assume that the limiting function

f (t) := lim fL (t)


L→∞

is a non-periodic absolutely integrable function, i.e.,


Z ∞
|f (ξ)|dξ < ∞.
−∞

Then, for all L > 0, we have


Z L Z L Z ∞
1 1 1

2L f (ξ)dξ ≤
|f (ξ)|dξ ≤ |f (ξ)|dξ,
−L 2L −L 2L −∞

which implies
Z L
1
lim f (ξ)dξ = 0. (56)
L→∞ 2L −L

Applied Mathematics 256/297


Fourier Integral

Letting L → ∞, then ∆ω → 0, and we get from (55) and (56) that


Z ∞
f (t) = {A(ω) cos(ωt) + B(ω) sin(ωt)}dω, (57)
0

where
Z ∞ Z ∞
1 1
A(z) := f (ξ) cos(zξ)dξ and B(z) := f (ξ) sin(zξ)dξ.
π −∞ π −∞

It is clear that the intuitive approach here merely suggests the


representation in (57) but, by no means, establishes it. In general, the
limit of the series in (55) as ∆ω → 0 is not the definition of the integral
in (57). Sufficient conditions for the validity of (57) are given below.

Applied Mathematics 257/297


Fourier Integral

Theorem 29 (Fourier Integral)

Let f be a piecewise continuous function on every finite interval, and


have a right-hand and left-hand derivative
R∞ at every point. Suppose also
that f is absolutely integrable, i.e., −∞ |f (ξ)|dξ < ∞. Then, f can be
represented by the Fourier integral
Z ∞
f (t) ∼ {A(ω) cos(ωt) + B(ω) sin(ωt)}dω,
0

where
Z ∞ Z ∞
1 1
A(z) := f (ξ) cos(zξ)dξ and B(z) := f (ξ) sin(zξ)dξ.
π −∞ π −∞

Further, the integral converges to f (t) if f is continuous at t and to


f (t + )+f (t − )
2 if f has a jump type discontinuity at t.

Applied Mathematics 258/297


Fourier Integral

Example 74

Show that the Fourier integral of




 0, t ≤ −h

1, −h < t < 0
f (t) :=


 2, 0<t≤h
0, t > h,

where h > 0, is
Z ∞
1 sin(hω)
f (t) ∼ 3 cos(ωt)
π 0 ω

1 − cos(hω)
+ sin(ωt) dω for t ∈ R. I
ω

Applied Mathematics 259/297


Fourier Integral

By Theorem 29, we see that the Fourier integral converges to the function
1


 2 , t = −h
1, −h < t < 0




3, t = 0

g (t) := 2
2, 0 < t < h


1, t = h





0, |t| > h.

Applied Mathematics 260/297


Fourier Integral

Example 75

Show that the Fourier integral of


(
1, |t| ≤ h
f (t) :=
0, |t| > h,

where h > 0, is
Z ∞
2 sin(hω)
f (t) ∼ cos(ωt)dω for t ∈ R. I
π 0 ω

Applied Mathematics 261/297


Fourier Integral

By Theorem 29, we see that the Fourier integral converges to the function

1, |t| < h

g (t) := 12 , |t| = h

0, |t| > h.

Applied Mathematics 262/297


Fourier Integral

Example 76

Show that the Fourier integral of


(
sgn(t), |t| < h
f (t) :=
0, |t| ≥ h,

where h > 0, is

1 − cos(hω)
Z
2
f (t) ∼ sin(ωt)dω. I
π 0 ω

Applied Mathematics 263/297


Fourier Integral

By Theorem 29, we see that the Fourier integral converges to the function
 1

 (− 2 ), t = −h

(−1), −h < t < 0





0, t=0
g (t) :=
1,
 0<t<h

1
, t=h


2



0, |t| > h.

Applied Mathematics 264/297


Fourier Integral

The Fourier Cosine and Sine Integrals

In the previous subsection, we first stated Fourier’s integral theorem.


Then, we gave the preferred form of Fourier’s integral. In Example 75
and Example 76, we see that B(z) ≡ 0 and A(z) ≡ 0, respectively. In
this subsection, we will present that this property holds in general for
even and for odd functions.

Applied Mathematics 265/297


Fourier Integral

Theorem 30 (Fourier Integral Theorem for Even Functions)


Fourier integral of an even function is a Fourier Cosine integral of the
form Z ∞
f (t) ∼ A(ω) cos(ωt)dω,
0
where Z ∞
2
A(z) := f (ξ) cos(zξ)dξ.
π 0

This actually gives us the Fourier transform of f (|t|) for t ∈ R if


f : [0, ∞) → R.

Applied Mathematics 266/297


Fourier Integral

Theorem 31 (Fourier Integral Theorem for Odd Functions)


Fourier integral of an odd function is a Fourier Sine integral of the form
Z ∞
f (t) ∼ B(ω) sin(ωt)dω,
0

where Z ∞
2
B(z) := f (ξ) sin(zξ)dξ.
π 0

This actually gives us the Fourier transform of sgn(t)f (|t|) for t ∈ R if


f : [0, ∞) → R.

Applied Mathematics 267/297


Fourier Integral

Example 77

Show that the Fourier Cosine integral of

f (t) = e−αt for t ≥ 0,

where α > 0, is
Z ∞
2 α
f (|t|) ∼ cos(ωt)dω for t ∈ R.
π 0 α2 + ω2

!!! Insert Graphic Here !!!

Applied Mathematics 268/297


Fourier Integral

Example 78

Show that the Fourier Cosine integral of


2
f (t) = e−αt for t ≥ 0,

where α > 0, is
Z ∞
1 ω2
f (|t|) ∼ √ e− 4α cos(ωt)dω for t ∈ R.
απ 0

!!! Insert Graphic Here !!!

Applied Mathematics 269/297


Fourier Integral

Example 79

Show that the Fourier Sine integral of

f (t) = e−αt for t ≥ 0,

where α > 0, is
Z ∞
2 ω
sgn(t)f (|t|) ∼ sin(ωt)dω for t ∈ R.
π 0 α2 + ω 2

Applied Mathematics 270/297


Fourier Integral

Solving PDEs by Fourier Integrals

Now, we will solve some examples by using the notion of Fourier integral.
Example 80

Consider the heat equation under Dirichlet type boundary conditions in


the upper half-plane

u − uxx = 0, −∞ < x < ∞, t > 0



 t

2
u(x, 0) = e−x , −∞ < x < ∞


u is bounded.

We will use the Fourier integral concept to find the solution. I

Applied Mathematics 271/297


Fourier Integral

Let u(x, y ) := ϕ(x)ψ(y ), then

ϕ00 + λϕ = 0 and ψ̇ + λψ = 0.

Then, we have

−µx
a(λ)e
 + b(λ)eµx , λ<0
ϕλ (x) := a(λ) + b(λ)x, λ=0

a(λ) cos(µx) + b(λ) sin(µx), λ > 0,

where µ > 0 satisfies λ = sgn(λ)µ2 , and

ψλ (t) := c(λ)e−λt . I

Applied Mathematics 272/297


Fourier Integral

Case 1. Let λ < 0. u is not bounded if a(λ) 6= 0 or b(λ) 6= 0.


Case 2. Let λ = 0. u is not bounded if b(λ) 6= 0.
Case 3. Let λ > 0. u is bounded for any a(λ) and b(λ).
This can be written as

ϕλ (x) := a(λ) cos(µx) + b(λ) sin(µx) for λ ≥ 0.

Then, we have
2
uµ (x, t) := A(µ) cos(µx) + B(µ) sin(µx) e−µ t

for µ ≥ 0. I

Applied Mathematics 273/297


Fourier Integral

By the superposition principle, we have


Z ∞
2
A(ω) cos(ωx) + B(ω) sin(ωx) e−ω t dω.

u(x, t) :=
0

Then, by Example 78, we have


Z ∞
2
e−x = u(x, 0) =

A(ω) cos(ωx) + B(ω) sin(ωx) dω,
0

which yields
1 z2
A(z) := √ e− 4 and B(z) :≡ 0. I
π

Applied Mathematics 274/297


Fourier Integral

Therefore,
Z ∞
1 ω2 2
u(x, t) := √ e− 4 cos(ωx)e−ω t dω
π 0
Z ∞
1 1 2
=√ cos(ωx)e−(t+ 4 )ω dω. I
π 0

Applied Mathematics 275/297


Fourier Integral

The last integral above can be computed as


x2
e− 1+4t
u(x, t) := √ .
1 + 4t

x 1 −1
t

2
− x
Figure 30: Graphic of e√ 1+4t
. 
1+4t

Applied Mathematics 276/297


Fourier Integral

Example 81

Consider the Laplace equation under Dirichlet type boundary conditions


in the upper half-plane

uxx + uyy = 0, −∞

 (
< x < ∞, y > 0

 1, |x| < π
u(x, 0) =


 0, |x| ≥ π

u is bounded.

We will use the Fourier integral concept to find the solution. I

Applied Mathematics 277/297


Fourier Integral

Let u(x, y ) := ϕ(x)ψ(y ), then

ϕ00 + λϕ = 0 and ψ̈ − λψ = 0.

Then, we have

−µx
a(λ)e
 + b(λ)eµx , λ<0
ϕλ (x) := a(λ) + b(λ)x, λ=0

a(λ) cos(µx) + b(λ) sin(µx), λ > 0

and 
c(λ) cos(µy ) + d(λ) sin(µy ), λ < 0

ψλ (y ) := c(λ) + d(λ)y , λ=0
c(λ)e−µy + d(λ)eµy ,

λ > 0,

where µ > 0 satisfies λ = sgn(λ)µ2 . I

Applied Mathematics 278/297


Fourier Integral

Case 1. Let λ < 0.


• ϕ is not bounded if a(λ) 6= 0 or b(λ) 6= 0.
• ψλ is bounded for any c(λ) and d(λ).
Case 2. Let λ = 0.
• ϕ is not bounded if b(λ) 6= 0.
• ψλ is not bounded if d(λ) 6= 0.
Case 3. Let λ > 0.
• ϕλ is bounded for any a(λ) and b(λ).
• ψλ (y ) is not bounded if d(λ) 6= 0. I

Applied Mathematics 279/297


Fourier Integral

Thus, 
0,
 λ<0
ϕλ (x) := a(λ), λ=0

a(λ) cos(µx) + b(λ) sin(µx), λ > 0

and 
c(λ) cos(µy ) + d(λ) sin(µy ), λ < 0

ψλ (y ) := c(λ), λ=0
c(λ)e−µy ,

λ > 0.

Applied Mathematics 280/297


Fourier Integral

Therefore, we have

uµ (x, y ) := {A(µ) cos(µx) + B(µ) sin(µx)}e−µy for µ ≥ 0,

which yields
Z ∞
u(x, y ) := {A(ω) cos(ωx) + B(ω) sin(ωx)}e−ωy dω. I
0

Applied Mathematics 281/297


Fourier Integral

Using the initial condition, we get


Z ∞
f (x) = u(x, 0) = {A(ω) cos(ωx) + B(ω) sin(ωx)}e−ωy dω.
0

By Example 75, we find that

2 sin(πz)
A(z) := and B(z) :≡ 0.
π z
Therefore, we have
Z ∞
2 sin(πω)
u(x, y ) := cos(ωx)e−ωy dω. I
π 0 ω

Applied Mathematics 282/297


Fourier Integral

The last integral above can be computed as


    
1 π+x π−x
u(x, y ) := arctan + arctan .
π y y

x 1 −1
y

1 π+x π−x 
  
Figure 31: Graphic of π
arctan y
+ arctan y
.

Applied Mathematics 283/297


Fourier Transform

Fourier Transform

Using the definition of Fourier integral, we will introduce the so-called


Fourier transform. Let f be a continuous function and consider the
Fourier integral

1 ∞
Z Z ∞ Z ∞ 
f (t) = f (ξ) c(ωξ) c(ωt) + f (ξ) s(ωξ) s(ωt) dξdω
π 0 −∞ −∞
Z ∞Z ∞
1 
= f (ξ) cos ω(ξ − t) dξdω
π 0 −∞
Z ∞Z ∞
1 
= f (ξ) cos ω(ξ − t) dξdω
2π −∞ −∞
Z ∞Z ∞
1 h i
= f (ξ) e−iω(ξ−t) + eiω(ξ−t) dξdω
4π −∞ −∞
"Z #
∞ Z ∞ Z ∞Z ∞
1
= f (ξ)e−iω(ξ−t) dξdω + f (ξ)eiω(ξ−t) dξdω .
4π −∞ −∞ −∞ −∞

Applied Mathematics 284/297


Fourier Transform

Substituting ω 7→ −ω in the second integral, we get


"Z #
∞ Z ∞ Z ∞Z ∞
1 −iω(ξ−t) iω(ξ−t)
f (t) = f (ξ)e dξdω + f (ξ)e dξdω
4π −∞ −∞ −∞ −∞
Z ∞Z ∞
1
= f (ξ)e−iω(ξ−t) dξdω
2π −∞ −∞
Z ∞ Z ∞ 
1 1 −iωξ
=√ √ f (ξ)e dξ eiωt dω.
2π −∞ 2π −∞

Applied Mathematics 285/297


Fourier Transform

Define the Fourier transform by


Z ∞
1
F{f }(z) := √ f (ξ)e−izξ dξ,
2π −∞

then Z ∞
1
√ F{f }(ω)eiωt dω = f (t)
2π −∞

or equivalently 
F F{f } (−t) = f (t).
That is, if F is a Fourier transform of a function f , then

F −1 {F }(t) = F{F }(−t) = f (t).

Applied Mathematics 286/297


Fourier Transform

Definition 38 (Fourier Transform)

Let f : R → R be a piecewise continuous function such that


Z ∞
|f (ξ)|dξ < ∞.
−∞

Then, the Fourier transform F of f is defined by


Z ∞
1
F{f }(z) := √ f (ξ)e−izξ dξ.
2π −∞

Applied Mathematics 287/297


Fourier Transform

Example 82

Show that the Fourier transform of


(
1, |t| < h
f (t) :=
0, |t| ≥ h,

where h > 0, is r
21
F{f }(z) = sin(hz).
πz

!!! Insert Graphic Here!!!

Applied Mathematics 288/297


Fourier Transform

Example 83

Show that r
 −α|·|
2 α
F e (z) = ,
π α2 + z 2
where α > 0.
!!! Insert Graphic Here!!!

Applied Mathematics 289/297


Fourier Transform

Example 84

Show that
2 1 z2
F e−α· (z) = √ e− 4α ,


where α > 0.
!!! Insert Graphic Here!!!

Applied Mathematics 290/297


Fourier Transform

Theorem 32
The Fourier transform is linear, i.e.,

F{αf + βg }(z) = αF{f }(z) + βF{g }(z),

where α, β are constants and f , g are piecewise continuous functions.

Theorem 33
Let f be a piecewise differentiable function such that limt→±∞ f (t) = 0.
Then,
F{f 0 }(z) = izF{f }(z).

Applied Mathematics 291/297


Fourier Transform

Definition 39 (Convolution)

Let f and g be suitable functions. The convolution f ∗ g is defined by


Z ∞
(f ∗ g )(t) := f (t − τ )g (τ )dτ.
−∞

Theorem 34

Let F and G denote the Fourier transforms of the functions f and g ,


respectively. Then,
F(f ∗ g )(z) = F (z)G (z).

Applied Mathematics 292/297


Fourier Transform

Definition 40 (Inverse Fourier Transform)

The integral Z ∞
1
F −1 {F }(t) := √ F (ω)eiωt dω.
2π −∞

Theorem 35

If F is the Fourier transform of f , then


Z ∞
1
√ F (ω)eiωt dω = f (t).
2π −∞

Applied Mathematics 293/297


Fourier Transform

Example 85

Consider the heat equation under Dirichlet type boundary conditions in


the upper half-plane

ut − uxx = 0, −∞ < x < ∞, t > 0

u(x, 0) = e−|x| , −∞ < x < ∞

u is bounded.

We will use the Fourier transform to find the solution. I

Applied Mathematics 294/297


Fourier Transform

Let us denote by U(z, t) the Fourier transform of u(x, t) in the variable


x, i.e., Z ∞
1
U(z, t) := √ u(ξ, t)e−izξ dξ.
2π −∞
Taking the Fourier transform of the PDE, we get

Ut + z 2 U = 0,

which is a first-order linear ordinary differential equation in the dependent


variable U and the independent variable t. Solving this, we get
2
U(z, t) := C (z)e−z t ,

where C is a function of z, which will be determined from the initial


condition. I

Applied Mathematics 295/297


Fourier Transform

This implies that


2
U(z, t) = U(z, 0)e−z t .
It follows from the initial condition and Example 83 that
r
 −|·| 2 1
U(z, 0) = F e (z) = .
π 1 + z2
By Example 84 and Theorem 34, we have
r  
2 1 −z 2 t 1  −|·| 2
− ·4t
U(z, t) = e =√ F e (z)F e
π 1 + z2 2t
 
1 ·2
= √ F e−|·| ∗ e− 4t . I
2t

Applied Mathematics 296/297


Fourier Transform

Taking the inverse Fourier transform, we get


 
1 −|·|
2
− ·4t
u(x, t) := √ e ∗e (x)
2t
or explicitly Z ∞
1 τ2
u(x, t) = √ e−|x−τ | e− 4t dτ,
2t −∞

which is the desired solution. 


!!! Insert Graphic Here!!!

Applied Mathematics 297/297

You might also like