Ema 481 Lecture Note
Ema 481 Lecture Note
LECTURE NOTE
1
Chapter One
Numerical methods for first-order differential equations
1.1 Initial Value Problem
Most physical systems can be described in mathematical terms through differential equations.
Analytical methods have been used to solve specific types of differential equation (i.e. the
separable-variable type, the homogeneous type and the linear type). However, differential
equations such as those used to solve real-life problems may not necessarily be directly
solvable, i.e. do not have closed form solutions. Instead, solutions can be approximated using
numerical methods and in science and engineering, a numeric approximation to the solution is
often good enough to solve a problem. There are a number of numerical methods available and
the simplest of these is called Euler’s method.
Fig. 1
Hence at some point f (h) in Fig. 1:
If the y-axis and origin are moved a units to the left, as shown in Fig. 1.1, the equation of the
same curve relative to the new axis becomes y = f (a + x) and the function value at P is f (a).
At point Q in Fig. 2:
2
Fig.2
which is a statement called Taylor’s series. If h is the interval between two new ordinates y0
and y1, as shown in Fig. 3, and if f (a)= y0 and y1 = f (a + h), then Euler’s method states
(2)
Fig. 3
The approximation used with Euler’s method is to take only the first two terms of Taylor’s
series shown in Equation (1).
Hence if y0, h and (y)0 are known, y1, which is an approximate value for the function at Q in
Fig. 1.3, can be calculated. Euler’s method is demonstrated in the worked problem as follows.
Problem 1. Obtain a numerical solution of the differential equation
given the initial conditions that x =1 when y =4, for the range x =1.0 to x =2.0 with intervals of
0.2.
Draw the graph of the solution
3
Solution
Fig.4
Thus at point R,
y1 = y0 + h(y)0 from Equation (2)
= 4.4 + (0.2)(2.2) = 4.84
When x1 =1.4 and y1 =4.84,
(y)1 =3(1 + 1.4) - 4.84=2.36
This step-by-step Euler’s method can be continued and it is easiest to list the results in a table,
as shown in Table 1. The results for lines 1 to 3 have been produced above
Table 1
4
For line 4, where x0 = 1.6:
y1 = y0 + h(y)0 = 4.84 + (0.2)(2.36) = 5.312 and ( y)0 = 3(1 + 1.6) - 5.312 = 2.488
For line 5, where x0 =1.8:
y1 = y0 + h(y)0 = 5.312 + (0.2)(2.488) = 5.8096 and ( y)0 = 3(1 + 1.8) - 5.8096 = 2.5904
For line 6, where x0 = 2.0:
y1 = y0 + h(y)0 = 5.8096 + (0.2)(2.5904) = 6.32768
(As the range is 1.0 to 2.0 there is no need to calculate (y)0 in line 6.) The particular solution is
given by the value of y against x
conditions x =1 and y =4 is shown in Fig. 1.5.
In practice it is probably best to plot the graph as each calculation is made, which checks that
there is a smooth progression and that no calculation errors have occurred.
Fig.5
The percentage error is given by the formular below, where the actual value is obtained by
analytical method and estimated value is obtained from the numerical method.
5
Fig. 6
to obtain an approximate value of y1 at point Q. QR in Fig. 6 is the resulting error in the result.
In an improved Euler method, called the Euler–Cauchy method, the gradient at P(x0, y0) across
half the interval is used and then continues with a line whose gradient approximates to the
gradient of the curve at x1.
Let yP1 be the predicted value at point R using Euler’s method, i.e. length RZ, where
(3)
Fig 7
The error shown as QT in Fig. 7 is now less than the error QR used in the basic Euler method
and the calculated results will be of greater accuracy. The corrected value, yC1 in the improved
Euler method is given by:
(4)
Problem 2.
Apply the Euler–Cauchy method to solve the differential equation
6
in the range 0(0.1)0.5, given the initial conditions that at x =0, y =2
solution
Since the initial conditions are x0 =0 and y0=2 then (y)0 =2-0=2. Interval h =0.1, hence
x1 = x0 + h =0 + 0.1=0.1
From Equation (3)
7
For line 5, x1 =0.4
yP1 = y0 + h(y)0 = 2.649232625 + (0.1)(2.349232625) = 2.884155887
yC1 = y0 + 12 h[(y)0 + f (x1, yP1)] = 2.649232625 + 1 2 (0.1)[2.349232625 + (2.884155887 - 0.4)]
= 2.89090205
(y)0 = yC1 - x1= 2.89090205 - 0.4 = 2.49090205
For line 6, x1 =0.5
yP1 = y0 + h(y)0 = 2.89090205 + (0.1)(2.49090205) = 3.139992255
yC1 = y0 + 12 h[(y)0 + f (x1, yP1)] = 2.89090205 + 12 (0.1)[2.49090205+ (3.139992255 - 0.5)]
= 3.14744676
The Runge–Kutta method
(c) Runge-Kutta 4th Order Method
The Runge–Kutta method for solving first-order differential equations is widely used and
provides a high degree of accuracy. Again, as with the two previous methods, the Runge–
Kutta method is a step-by-step process where results are tabulated for a range of values of x.
Although several intermediate calculations are needed at each stage, the method is fairly
straightforward.
The seven-step procedure for the Runge–Kutta method, without proof, is as follows:
To solve the differential equation
dy
dx
= f (x, y) given the initial condition y = y0 at x = x0 for a range of values of
x = x0(h)xn:
1. Identify x0, y0 and h, and values of x1, x2, x3,....
2. Evaluate k1 = f(xn, yn) starting with n =0
3. Evaluate k2 = f xn + h2 , yn + h2 k1
4. Evaluate k3 = f xn + h2 , yn + h2 k2
5. Evaluate k4 = f (xn + h, yn + hk3)
6. Use the values determined from steps 2 to 5 to evaluate:
yn+1 = yn +h/6{k1 + 2k2 + 2k3 + k4}
7. Repeat steps 2 to 6 for n =1, 2, 3, ..
Problem 3.
Use the Runge–Kutta method to solve the differential equation:
dy
dx
=y-x
In the range 0(0.1)0.5, given the initial conditions that at x =0, y =2
Solution
Using the above procedure:
1. x0 =0, y0 =2 and since h =0.1, and the range is from x =0 to x =0.5, then x1 =0.1, x2 =0.2, x3 =
0.3, x4 = 0.4, and x5 = 0.5
Let n=0 to determine y1:
2. k1= f (x0, y0)= f (0,2);
8
since
dy
dx
= y - x, f (0,2)=2 - 0=2
Table 3
9
dy
=y - x may be solved analytically using the integrating factor method with the solution:
dx
y=x+1+ex
Substituting values of x of 0, 0.1, 0.2,..., 0.5 will give the exact values. A comparison of the
results obtained by Euler’s method, the Euler–Cauchy method and the Runga–Kutta method,
together with the exact values is shown in Table 4. It is seen from Table 4 that the Runge–
Kutta method is exact, correct to 5 decimal places
Table 4
10
Chapter Two
Introduction to Partial Differential Equations (PDE’s)
2u 2u 2u
A 2 +B +C 2 + D = 0 (2.1)
x xy y
u u
Where A, B, and C are functions of x and y and D is a function of x, y, u and , .
x y
Depending on the value of B 2 − 4 AC , a 2nd order linear PDE can be classified into three categories.
1. if B 2 − 4 AC 0 , it is called elliptic
2. if B 2 − 4 AC = 0 , it is called parabolic
3. if B 2 − 4 AC 0 , it is called hyperbolic
2T 2T
+ =0 (2.2)
x 2 y 2
Using the general form of second order linear PDEs with one dependent variable and two independent
variables,
2u 2u 2u
A + B + C +D=0
x 2 xy y 2
A = 1, B = 0, C = 1, D = 0 ,
11
gives B 2 − 4 AC = 0 − 4(1)(1)
= −4
= −4 0
This classifies Equation (2.2) as elliptic.
T 2T
=k 2 (2.3)
t x
Using the general form of second order linear PDEs with one dependent variable and two independent
variables,
2u 2u 2u
A 2 +B +C 2 + D = 0
x xy y
A = k , B = 0, C = 0, D = −1 ,
gives B 2 − 4 AC = 0 − 4(0)(k )
=0
This classifies Equation (2.3) as parabolic.
2 y 1 2 y
= (2.4)
x 2 c 2 t 2
Using the general form of second order linear PDEs with one dependent variable and two independent
variables,
2u 2u 2u
A + B + C +D=0
x 2 xy y 2
1
A = 1, B = 0, C = − ,D = 0
c2
12
−1
gives B 2 − 4 AC = 0 − 4(1)( )
c2
4
=
c2
4
= 0
c2
This classifies Equation (2.4) as hyperbolic.
There are three basic numerical methods in solving boundary value problems of ordinary differential
equations (ODEs) and partial differential equations (PDEs). These are:
This lecture note provides an introduction to the finite difference method (FDM) for solving partial
differential equations (PDEs). In addition to specific FDM details, general concepts such as stability,
boundary conditions and verification.
𝑑𝑦 𝑑𝑦
| = 𝐶1 , | = 𝐶2
𝑑𝑥 𝑥=0 𝑑𝑥 𝑥=1
𝑑𝑦 𝑑𝑦
𝑦+ | = 𝐶1 , 𝑦 + | = 𝐶2
𝑑𝑥 𝑥=0 𝑑𝑥 𝑥=1
13
2.5 Parabolic Partial Differential Equations
One of the simple examples of a parabolic PDE is the heat-conduction equation for a metal rod (Figure
2.1)
2T T
= (2.5)
x 2 t
where
k
=
C
where
14
i = node number along the x − direction, i = 0,1,...., n ,
x
x x
i −1 i i +1
Figure 2.2: Schematic diagram showing the node representation in the model
L
x = (2.7)
n
The time is similarly broken into time steps of t . Hence Ti j corresponds to the temperature at node i
, that is,
x = (i )(x )
and time,
t = ( j )(t ) ,
where
t = time step.
The time derivative of the right hand side of Equation (2.3) is approximated by the forward divided
difference approximation
T Ti j +1 − Ti j
(2.8)
t i, j t
Substituting the finite difference approximations given by Equations (2.6) and (2.8) in Equation (2.3)
gives
15
Solving for the temperature at the time node j + 1 , gives
t
Ti j +1 = Ti j +
(x) 2
(Ti +j1 − 2Ti j + Ti −j1 )
Choosing
t
= (2.9)
(x) 2
(
Ti j +1 = Ti j + Ti +j1 − 2Ti j + Ti −j1 ) (2.10)
Equation (2.10) can be solved explicitly because it can be written for each internal location node of the
rod for time node j + 1 in terms of the temperature at time node j . In other words, if we know the
temperature at node j = 0 , and knowing the boundary temperatures, which is the temperature at the
external nodes, we can find the temperature at the next time step. We continue the process by first
finding the temperature at all nodes j = 1 , and using these to find the temperature at the next time
node, j = 2 . This process continues till we reach the time at which we are interested in finding the
temperature.
Example 1
A rod of steel is subjected to a temperature of 100C on the left end and 25C on the right end. If the
rod is of length 0.05m , use the explicit method to find the temperature distribution in the rod from
t = 0 and t = 9 seconds. Use x = 0.01m , t = 3s . Given:
W kg J
k = 54 , = 7800 3 , C = 490
m−K m kg − K
Solution
k
=
C
54
= = 1.4129 10 −5 m 2 / s
7800 490
Then
t
=
(x )2
3
= 1.4129 10−5 = 0.4239
(0.01)2
16
t final − t initial 9−0
Number of time steps= = =3
t 3
i=0 1 2 3 4 5
T =100 C T = 25 C
0.01m
Figure 2.3: Schematic diagram showing the node distribution in the rod
T0 j = 100C
for all j = 0,1,2,3 (2.11)
T5 j = 25C
The initial temperature of the rod is 20C , that is, all the temperatures of the nodes inside the rod are
at 20C when time, t = 0 sec except for the boundary nodes as given by Equation (2.11). This could be
represented as
Initial temperature at the nodes inside the rod (when t=0 sec)
17
T11 = T10 + (T20 − 2T10 + T00 )
= 20 + 0.4239(20 − 2(20) + 100)
= 20 + 0.4239(80)
= 20 + 33.912
= 53.912C
18
T12 = T11 + (T21 − 2T11 + T01 )
= 53.912 + 0.4239(20 − 2(53.912) + 100)
= 53.912 + 0.4239(12.176)
= 53.912 + 5.1614
= 59.073C
(
T42 = T41 + T51 − 2T41 + T31 )
= 22.120 + 0.4239(25 − 2(22.120) + 20)
= 22.120 + 0.4239(0.76)
= 22.120 + 0.032220
= 22.442C
𝑇52 = 250 𝐶
19
(
T13 = T12 + T22 − 2T12 + T02 )
= 59.073 + 0.4239(34.375 − 2(59.073) + 100)
= 59.073 + 0.4239(16.229)
= 59.073 + 6.8795
= 65.953C
(
T23 = T22 + T32 − 2T22 + T12 )
= 34.375 + 0.4239(20.899 − 2(34.375) + 59.073)
= 34.375 + 0.4239(11.222)
= 34.375 + 4.7570
= 39.132C
(
T33 = T32 + T42 − 2T32 + T22 )
= 20.899 + 0.4239(22.442 − 2(20.899) + 34.375)
= 20.899 + 0.4239(15.019)
= 20.899 + 6.367
= 27.266C
(
T43 = T42 + T52 − 2T42 + T32 )
= 22.442 + 0.4239(25 − 2(22.442) + 20.899)
= 22.442 + 0.4239(1.0150)
= 22.442 + 0.4303
= 22.872C
To better visualize the temperature variation at different locations at different times, temperature
distribution along the length of the rod at different times is plotted in the Figure2.4.
20
Temperature distribution along the length of the rod
Temperature, T (oC)
t=3 secs
t= 6 secs
t= 9 secs
Also, the explicit method does not guarantee stability which depends on the value of the time step,
location step and the parameters of the elliptic equation. For the PDE
2T T
= ,
x 2 t
the explicit method is convergent and stable for
t 1
(2.13)
(x) 2
2
These issues are addressed by using the implicit method. Instead of the temperature being found one
node at a time, the implicit method results in simultaneous linear equations for the temperature at all
interior nodes for a particular time.
The second derivative on the left hand side of the equation is approximated by the central divided
difference scheme at time level j + 1 at node i as
21
2T Ti +j1+1 − 2Ti j +1 + Ti −j1+1
(2.14)
x 2 i , j +1
(x )2
The first derivative on the right hand side of the equation is approximated by backward divided
difference approximation at time level j + 1 and node i as
T Ti j +1 − Ti j
(2.15)
t i , j +1 t
giving
where
t
=
(x )2
Now Equation (2.16) can be written for all nodes (except the external nodes), at a particular time level.
This results in simultaneous linear equations which can be solved to find the nodal temperature at a
particular time.
Example 2
A rod of steel is subjected to a temperature of 100C on the left end and 25C on the right end. If the
rod is of length 0.05m , use the implicit method to find the temperature distribution in the rod from
t = 0 to t = 9 seconds. Use x = 0.01m and t = 3s .
Given
W kg J
k = 54 , = 7800 3 , C = 490 .
m−K m kg − K
Solution
k 54
= =
C 7800 490
= 1.4129 10−5 m 2 / s
Then
22
t 3
= = 1.412 10−5 = 0.4239
(x )2
(0.01)2
i=0 1 2 3 4 5
T = 100 C T = 25 C
0.01m
Figure 2.5: Schematic diagram showing the node representation in the model
T0 j = 100C
for j = 0,1,2,3 (2.17)
T5 j = 25C
The initial temperature of the rod is 20C , that is, the temperatures of all the nodes inside the rod are
at 20C when time, t = 0 except for the boundary nodes where the temperatures are given by
satisfying the Equation (2.17). This could be represented as
Initial temperature at the nodes inside the rod (when t=0 sec)
𝑇00 = 1000 𝐶
𝑇10 = 200 𝐶
𝑇20 = 200 𝐶
𝑇30 = 200 𝐶
𝑇40 = 200 𝐶
𝑇50 = 250 𝐶
23
For all the interior nodes, putting j = 0 and i =1, 2, 3, 4 in Equation (2.16) gives the following equations
i=1
i=2
i=3
i=4
The simultaneous linear equations (2.19) – (2.22) can be written in matrix form as
The above coefficient matrix is tri-diagonal. Special algorithms such as Thomas’ algorithm can be used to
solve simultaneous linear equation with tri-diagonal coefficient matrices. The solution is given by
24
T11 39.451
1
T2 = 24.792
T31 21.438
1
T4 21.477
Hence, the temperature at all the nodes at time, t = 3 sec is
T01 100
1
T1 39.451
T21 24.792
1 =
T3 21.438
T 1 21.477
41
T5 25
For all the interior nodes, putting j = 1 and i =1, 2, 3, 4 in Equation (2.16) gives the following equations
i=1
i=2
i=3
25
i=4
The simultaneous linear equations (2.23) – (2.26) can be written in matrix form as
T12 51.326
2
T2 = 30.669
T32 23.876
2
T4 22.836
Hence, the temperature at all the nodes at time, t = 6 sec is
T02 100
2
T1 51.326
T22 30.669
2 =
T3 23.876
T 2 22.836
42
T5 25
For all the interior nodes, setting j = 2 and i =1, 2, 3, 4 in Equation (2.16) gives the following equations
i=1
26
− T03 + (1 + 2 )T13 − T23 = T12
(−0.4239 100) + (1 + 2 0.4239)T13 − (0.4239T23 ) = 51.326
− 42.39 + 1.8478T13 − 0.4239T23 = 51.326
i=2
i=3
i=4
T13 59.043
3
T2 = 36.292
T33 26.809
3
T4 24.243
27
T03 100
3
T1 59.043
T23 36.292
3 =
T3 26.809
T 3 24.243
43
T5 25
To better visualize the temperature variation at different locations at different times, the temperature
distribution along the length of the rod at different times is plotted in Figure 6.
t=3 secs
t= 6 secs
t= 9 secs
28
2T 1 Ti +j1 − 2Ti j + Ti −j1 Ti +j1+1 − 2Ti j +1 + Ti −j1+1
+ (2.30)
x 2 i, j
2 (x )2 (x )2
The first derivative on the right side of Equation (2) is approximated using forward divided difference
approximation at time level j + 1 and node i as
T Ti j +1 − Ti j
(2.31)
t i, j t
giving
− Ti −j1+1 + 2(1 + )Ti j +1 − Ti +j1+1 = Ti −j1 + 2(1 − )Ti j + Ti +j1 (2.33)
where
t
=
(x )2
Now Equation (2.33) is written for all nodes (except the external nodes). This will result in simultaneous
linear equations that can be solved to find the temperature at a particular time.
Example 3
A rod of steel is subjected to a temperature of 100C on the left end and 25C on the right end. If the
rod is of length 0.05m , use Crank-Nicolson method to find the temperature distribution in the rod from
t = 0 to t = 9 seconds. Use x = 0.01m , t = 3s .
Given
W kg J
k = 54 , = 7800 3 , C = 490
m−K m kg − K
Solution
k 54
= = = 1.412910−5 m 2 / s
C 7800 490
Then
t
=
(x )2
29
3
= 1.412 10−5 = 0.4239
(0.01)2
i=0 1 2 3 4 5
0.01m
Figure 2.7: Schematic diagram showing the node representation in the model
T0 j = 100C
for j = 0,1,2,3 (2.34)
T5 j = 25C
The initial temperature of the rod is 20C , that is, all the temperatures of the nodes inside the rod are
at 20C at, t = 0 except for the boundary nodes given by Equation (2.34). This could be represented as
Initial temperature at the nodes inside the rod (when t=0 sec)
𝑇00 = 1000 𝐶
𝑇10 = 200 𝐶
𝑇20 = 200 𝐶
𝑇30 = 200 𝐶
𝑇40 = 200 𝐶
𝑇50 = 250 𝐶
30
For all the interior nodes, setting j = 0 and i =1, 2, 3, 4 in Equation (2.33) gives the following equations
i=1
i=2
i=3
i=4
The coefficient matrix in the above set of equations is tridiagonal. Special algorithms such as Thomas’
algorithm are used to solve equation with tridiagonal coefficient matrices
31
T11 44.372
1
T2 = 23.746
T31 20.797
1
T4 21.607
Hence, the temperature at all the nodes at time, t = 3 sec is
T01 100
1
T1 44.372
T21 23.746
1 =
T3 20.797
T 1 21.607
41
T5 25
For all the interior nodes, putting j = 1 and i =1, 2, 3, 4 in Equation (2.33) gives the following equations
i=1
i=2
i=3
32
− T22 + 2(1 + )T32 − T42 = T21 + 2(1 − )T31 + T41
− 0.4239T22 + 2(1 + 0.4239)T32 − 0.4239T42 =
(0.4239)23.746 + 2(1 − 0.4239)20.797 + (0.4239)21.607
− 0.4239T22 + 2.8478T32 − 0.4239T42 = 10.066 + 23.962 + 9.1592
− 0.4239T22 + 2.8478T32 − 0.4239T42 = 43.187 (2.42)
i=4
T12 55.883
2
T2 = 31.075
T32 23.174
2
T4 22.730
T02 100
2
T1 55.883
T22 31.075
2 =
T3 23.174
T 2 22.730
42
T5 25
33
For all the interior nodes, setting j = 2 and i =1, 2, 3, 4 in Equation (2.33) gives the following equations
i=1
i=2
i=3
i=4
34
T13 62.604
3
T2 = 37.613
T33 26.562
3
T4 24.042
Hence, the temperature at all the nodes at time, t = 9 secis
T03 100
3
T1 62.604
T23 37.613
3 =
T3 26.562
T 3 24.042
43
T5 25
To better visualize the temperature variation at different locations at different times, the temperature
distribution along the length of the rod at different times is plotted in Figure 2.8.
t=3 secs
t= 6 secs
t= 9 secs
35
2.9 Analytical Method
The parabolic heat conduction equation is formulated as
2T T
= 0 x 0.05, t 0
x 2 t
with boundary conditions
T = 100C at x = 0, t 0 (2.48)
We split the problem into a steady state problem and a transient (homogeneous) problem. The
solutions of the steady state problem and transient problem are found separately and by applying the
principle of superposition, the final solution would be obtained. This formulation can be represented as
T ( x, t ) = Ts ( x) + Th ( x, t ) (2.51)
where
d 2Ts
= 0, 0 x 0.05 (2.52)
dx 2
with boundary conditions
Ts = 100C at x = 0 (2.53)
dTs
=A
dx
where A is a constant of integration and by integrating again to give
Ts = Ax + B (2.55)
36
where B is another constant of integration. By substituting the boundary condition (2.53), we obtain
A(0) + B = 100
B = 100
By substituting the boundary condition (2.54), we obtain
A(0.05) + 100 = 25
− 75
A=
0.05
= −1500
Plugging back the values of A and B in Equation (2.55), we get the steady state solution as
2Th Th
= , 0 x 0.05 (2.57)
x 2 t
with boundary conditions
Th = 0C at x = 0 (2.58)
T ( x, t ) = Ts ( x) + Th ( x, t )
and by substituting Equations (2.53) and (2.54), the boundary conditions of Th are obtained.
Th = 20 − Ts , t = 0, 0 x 0.05
= 20 − (−1500x + 100)
= 20 + 1500x − 100
To obtain solution for the transient problem, let us assume Th ( x, t ) is function of the product of a
spatial function and a temperature function. That is
Th ( x, t ) = X ( x). (t ) (2.61)
37
Substituting Equation (2.61) in Equation (2.57), we get
d2X d
2
=X
dx dt
1 d2X 1 d
= (2.62)
X dx 2
dt
The left hand side of Equation (2.62) represents the spatial term and the right hand side represents the
temporal (time) term. We will attempt to find the solutions of the spatial and temporal term
independently. To do so, let us assume that both the left hand side and the right hand side of the
Equation (2.62) is equal to a constant − 2 (say)
1 d2X 1 d
= = − 2 (2.63)
X dx 2
dt
1 d2X
2
= − 2
X dx
d2X
+ 2X = 0 (2.64)
dx 2
The Equation (2.64) is a homogeneous second order ordinary differential equation. These type of
equations have the solution of the form X ( x) = e mx . Substituting X ( x) = e mx in Equation (2.64) we
get,
m 2 e mx + 2 e mx = 0
e mx (m 2 + 2 ) = 0
m2 + 2 = 0
m1 , m2 = i , − i
Temporal solution
38
1 d
= − 2
dt
d
+ 2 = 0 (2.66)
dt
The above equation is a homogeneous first order ordinary differential equation. These type of equations
have the solution of the form (t ) = e mt . Substituting (t ) = e mt in Equation (2.66) we get
me mt + 2 e mt = 0
e mt (m + 2 ) = 0
m + 2 = 0
m = − 2
(t ) = Ee −
2
t
(2.66)
e − t F .1 + G.0 = 0
2
e − t [ F ] = 0
2
Th ( x, t ) = Ge − t sin( x)
2
(2.68)
Ge − t sin(0.4 ) = 0
2
sin(0.05 ) = 0
0.05 = n
39
n
= = 20n
0.05
Substituting the value of in Equation (2.68) gives
Th ( x, t ) = Ge − ( 20 n ) t sin(20nx)
2
G
n =1
n sin(20nx) = 1500x − 80
0.05 0.05
n =1 0
Gn sin(20nx) sin(20mx)dx = (1500x − 80) sin(20mx)dx
0
0.05 0.05
Gn
n =1 2
2 sin(20nx) sin(20mx)dx = (1500x − 80) sin(20mx)dx
0 0
Gn 0.05 0.05
cos(20(m − n)x)dx −
n =1 2 0
cos(20(m + n)x)dx =
0
0.05 0.05
1500 x sin(20mx)dx − 80 sin(20mx)dx
0 0
cos(20(m − n)x)dx = 0 ,
0
mn
0.05
0.05
cos(20(m + n)x)dx = 0,
0
for any m
we get
40
0.05 0.05
Gm
0.05 = 1500 x sin(20mx)dx − 80 sin(20mx)dx
2 0 0
− x cos(20m ) 0.05 1
0.05
0.05
= 1500
20m
0
+
20m 0 cos(20m ) dx
− 80 0 sin(20m )dx
− x cos(20m ) 0.05 1
0.05
0.05
= 1500
20m
0
+
20m 0 cos(20m ) dx
− 80 0 sin(20m )dx
Gm =
− 150(−1) m 160
m
+
m
(−1) m − 1
10(−1) m − 160
= (2.70)
m
Substituting Equation (2.70) in Equation (2.69), we get
10(−1) m − 160 − 2.5 m 2 t
Th ( x, t ) = e sin(2.5mx) (2.71)
m =1 m
Substituting Equations (2.71) and (2.56) in Equation (2.51) we have
10(−1) m − 160 − 20 m 2 t
T ( x, t ) = −1500x + 100 + e sin(20mx) (2.72)
m =1 m
Now
k
=
C
54
=
7800 490
= 1.4129 10−5 m 2 / s
41
Equation (2.73) is the analytical solution of the problem. Substituting the values of x and t gives the
temperature inside the rod at a particular location and time. For example using the analytical solution,
we will find the temperature of the rod at the first node, that is, x = 0.01m when t = 9 secs.
Similarly using Equation (2.73), the temperature of the rod at any location at any time can be found by
substituting the corresponding values of x and t .
Table 1: Comparison of temperature obtained at interior nodes using different methods discussed
in this chapter (absolute true error is given in parenthesis)
42
Chapter Three
Elliptic Partial Differential Equations
Tt
W Tl Tr
x
Tb
L
Figure 3.1: Schematic diagram of the plate with the temperature boundary conditions
The partial differential equation that governs the temperature T ( x, y ) is given by
2T 2T
+ =0 (3.1)
x 2 y 2
To find the temperature within the plate, we divide the plate area by a grid as shown in Figure 3.2.
43
y
Tt
(0, n)
x
y Tr
Tl (i, j )
x
(0,0)
Tb (m,0)
(i, j + 1)
y x
(i − 1, j ) (i, j ) (i + 1, j )
(i, j − 1)
L
x = (3.2)
m
W
y = (3.3)
n
Now we will apply the finite difference approximation of the partial derivatives at a general interior node
( i, j ).
44
Equations (3.4) and (3.5) are central divided difference approximations of the second derivatives.
Substituting Equations (3.4) and (3.5) in Equation (3.1), we get
x = y
Now we can write this equation at all the interior nodes of the plate, that is (m − 1) (n − 1) nodes. This
will result in an equal number of equations and unknowns. The unknowns are the temperatures at the
interior (m − 1) (n − 1) nodes. Solving these equations will give us the two-dimensional profile of the
temperature inside the plate.
Example 1
A plate 2.4 m 3.0 m is subjected to temperatures as shown in Figure 3.3. Use a square grid length of
0.6 m . Using the direct method, find the temperature at the interior nodes.
y
300 C
75 C
3 .0 m 100 C
x
50 C
2 .4 m
Solution
x = y = 0 . 6 m
45
L 2 .4
m= = =4
x 0 .6
W
n=
y
3
=
0 .6
=5
T0, 4
T1, 4 T2, 4 T3, 4 T4, 4
T0,3
T1,3 T2,3 T3,3 T4,3
T0, 2
T1, 2 T2, 2 T3, 2 T4, 2
All the nodes on the left and right boundary have an i value of zero and m , respectively. While all the
nodes on the top and bottom boundary have a j value of zero and n , respectively.
The corner nodal temperature of T0,5 , T4,5 , T4,0 and T0,0 are not needed. Now to get the temperature at
the interior nodes we have to write Equation (3.7) for all the combinations of i and j,
i = 1,...., m − 1; j = 1,...., n − 1 .
46
i=1 and j=1
T2,1 + T0,1 + T1, 2 + T1,0 − 4T1,1 = 0
47
i=2 and j=4
T3, 4 + T1, 4 + T2,5 + T2,3 − 4T2, 4 = 0
Equations (3.8) to (3.19) represent a set of twelve simultaneous linear equations and solving them gives
the temperature at the twelve interior nodes. The solution is
48
T1,1 73.8924
T
1, 2 93.0252
T1,3 119.907
T1, 4 173.355
T2,1 77.5443
T2, 2 103.302
T = 138.248 C
2,3
T2, 4 198.512
T 82.9833
3,1
T3, 2 104.389
T3,3 131.271
T3, 4 182.446
75 100
173 199 182
75 100
120 138 131
75 100
93 103 104
75 74 78 83 100
x
50 50 50
Figure 3.5: Temperatures at the interior nodes of the plate
49
Ti +1, j + Ti −1, j + Ti , j +1 + Ti , j −1
Ti , j = , i = 1,2,3,4; j = 1,2,3,4,5 (3.20)
4
Now Equation (3.20) is solved iteratively for all interior nodes until all the temperatures at the interior
nodes are within a pre-specified tolerance.
Example 2
A plate 2.4 m 3.0 m is subjected to the temperatures as shown in Fig. 3.6. Use a square grid length of
0.6 m . Using the Gauss-Seidel method, find the temperature at the interior nodes. Conduct two iterations
at all interior nodes. Find the maximum absolute relative error at the end of the second iteration. Assume
the initial temperature at all interior nodes to be 0 C .
y
300 C
75 C
3 .0 m 100 C
x
50 C
2 .4 m
Figure 3.6: A rectangular plate with the dimensions and boundary temperatures
Solution
x = y = 0 . 6 m
L 2 .4
m= = =4
x 0 .6
W
n=
y
50
3
=
0 .6
=5
T0, 4
T1, 4 T2, 4 T3, 4 T4, 4
T0,3
T1,3 T2,3 T3,3 T4,3
T0, 2
T1, 2 T2, 2 T3, 2 T4, 2
All the nodes on the left and right boundary have an i value of zero and m , respectively. All of the nodes
on the top or bottom boundary have a j value of either zero or n , respectively.
The corner nodal temperature of T0,5 , T4,5 , T4,0 and T0,0 are not needed. Now to get the temperature at
the interior nodes we have to write Equation (3.20) for all of the combinations of i and j ,
i = 1,...., m − 1; j = 1,..., n − 1 .
Iteration 1
For iteration 1, we start with all of the interior nodes having a temperature of 0C .
51
T2,1 + T0,1 + T1, 2 + T1, 0
T1,1 =
4
0 + 75 + 0 + 50
=
4
= 31.2500C
52
T3, 2 + T1, 2 + T2,3 + T2,1
T2, 2 =
4
0 + 26.5625 + 0 + 20.3125
=
4
= 11.7188C
53
T4,3 + T2,3 + T3, 4 + T3, 2
T3,3 =
4
100 + 9.27735+ 0 + 38.5742
=
4
= 36.9629C
Iteration 2
For iteration 2, we use the temperatures from iteration 1.
T1,present − T1,previous
a 1,1 = 1
present
1
100
T1,1
42.9688 − 31.2500
= 100
42.9688
= 27.27%
54
T1,present − T1,previous
a 1, 2 = 2
present
2
100
T1, 2
38.7696 − 26.5625
= 100
38.7696
= 31.49%
T1,present − T1,previous
a 1,3 = 3
present
3
100
T1,3
55.7862 − 25.3906
= 100
55.7862
= 54.49%
T1,present − T1,previous
a 1, 4 = 4 4
100
T1,present
4
133.283 − 100.098
= 100
133.283
= 24.90%
55
42.5781+ 42.9688 + 11.7188 + 50
=
4
= 36.8164C
T2present − T2previous
a 2,1 = ,1 ,1
100
T2present
,1
36.8164 − 20.3125
= 100
36.8164
= 44.83%
T2present − T2previous
a 2, 2 = ,2
present
,2
100
T2, 2
30.8594 − 11.7188
= 100
30.8594
= 62.03%
T2present − T2previous
a 2,3 = ,3
present
,3
100
T2,3
56.4881− 9.27735
= 100
56.4881
= 83.58%
T2present − T2previous
a 2, 4 = ,4
present
,4
100
T2, 4
156.150 − 102.344
= 100
156.150
= 34.46%
T3present − T3previous
a 3,1 = ,1 ,1
present
100
T3,1
56.3477 − 42.5781
= 100
56.3477
= 24.44%
57
T3present − T3previous
a 3, 2 = ,2 ,2
present
100
T3, 2
56.0425 − 38.5742
= 100
56.0425
= 31.70%
T3present − T3previous
a 3, 3 = ,3 ,3
present
100
T3,3
86.8394 − 36.9629
= 100
86.8394
= 57.44%
T3present − T3previous
a 3, 4 = ,4 ,4
present
100
T3, 4
160.747 − 134.827
= 100
160.747
= 16.12%
58
y
75
133 156 161 100
75
56 56 87 100
75
39 31 56 100
75
43 37 56 100
x
50 50 50
It took ten iterations to get all of the temperature values within 1% error. The table below lists the
temperature values at the interior nodes at the end of each iteration:
1 2 3 4 5
59
T3, 2 38.5742 56.0425 75.7847 87.6890 94.6990
6 7 8 9 10
Ti ,relaxed
j = Ti ,new
j + (1 − )Ti , j
old
(3.22)
60
where
Ti ,new
j = value of temperature from current iteration,
= weighting factor, 1 2 .
Again, these iterations are continued till the pre-specified tolerance is met for all nodal temperatures. This
method is also called the Lieberman method.
Example 3
A plate 2.4 m 3.0 m is subjected to the temperatures as shown in Fig. 3.9. Use a square grid length of
0.6 m . Use the Gauss-Seidel with successive over relaxation method with a weighting factor of 1.4 to find
the temperature at the interior nodes. Conduct two iterations at all interior nodes. Find the maximum
absolute relative error at the end of the second iteration. Assume the initial temperature at all interior
nodes to be 0 C .
300 C
75 C
3 .0 m 100 C
x
50 C
2 .4 m
Figure 3.9: A rectangular plate with the dimensions and boundary temperatures
Solution
x = y = 0 . 6 m
61
L
m=
x
2 .4
=
0 .6
=4
W
n=
y
3
=
0 .6
=5
T0, 4
T1, 4 T2, 4 T3, 4 T4, 4
T0,3
T1,3 T2,3 T3,3 T4,3
T0, 2
T1, 2 T2, 2 T3, 2 T4, 2
All of the nodes on the left and right boundary have an i value of zero and m , respectively. All of the
nodes on the top or bottom boundary have a j value of either zero or n , respectively.
62
The corner nodal temperature of T0,5 , T4,5 , T4,0 and T0,0 are not needed. Now to get the temperature at
the interior nodes, we have to write Equation (3.20) for all of the combinations of i and j , i = 1 to m − 1
, j = 1 to n − 1 . After getting the temperature from Equation (3.20), we have to use Equation (3.22) to
apply the over relaxation method.
Iteration 1
For iteration 1, we start with all of the interior nodes having a temperature of 0C .
= 1.4(31.2500) + (1 − 1.4)0
= 43.7500C
= 1.4(29.6875) + (1 − 1.4)0
= 41.5625C
63
= 1.4(29.1406) + (1 − 1.4)0
= 40.7969C
= 1.4(103.949) + (1 − 1.4)0
= 145.529C
= 1.4(23.4375) + (1 − 1.4)0
= 32.8215C
= 1.4(18.5938) + (1 − 1.4)0
= 26.0313C
64
T3,3 + T1,3 + T2, 4 + T2, 2
T2 , 3 =
4
0 + 40.7969 + 0 + 26.0313
=
4
= 16.7071C
relaxed
T2,3 = T2,3
new
+ (1 − )T2,3
old
= 1.4(16.7071) + (1 − 1.4)0
= 23.3899C
= 1.4(117.230) + (1 − 1.4)0
= 164.122C
= 1.4(45.7031) + (1 − 1.4)0
= 63.9844C
65
100 + 26.0313+ 0 + 63.9844
=
4
= 47.5039C
relaxed
T3,2 = T3,2
new
+ (1 − )T3,2
old
= 1.4(47.5039) + (1 − 1.4)0
= 66.5055C
= 1.4(47.4739) + (1 − 1.4)0
= 66.4634C
= 1.4(157.646) + (1 − 1.4)0
= 220.704C
Iteration 2
For iteration 2, we take the temperatures from iteration 1.
66
32.8125 + 75 + 41.5625 + 50
=
4
= 49.8438C
= 1.4(49.8438) + (1 − 1.4)43.75
= 52.2813C
T1,present − T1,previous
a 1,1 = 1
present
1
100
T1,1
52.2813 − 43.7500
= 100
52.2813
= 16.32%
= 1.4(48.5274) + (1 − 1.4)41.5625
= 51.3133C
T1,present − T1,previous
a 1, 2 = 2
present
2
100
T1, 2
51.3133− 41.5625
= 100
51.3133
= 19.00%
67
T1,3relaxed = T1,3new + (1 − )T1,3old
= 1.4(73.8103) + (1 − 1.4)40.7969
= 87.0157C
T1,present − T1,previous
a 1,3 = 3
present
3
100
T1,3
87.0157 − 40.7969
= 100
87.0157
= 53.12%
= 1.4(156.534) + (1 − 1.4)145.529
= 160.936C
T1,present − T1,previous
a 1, 4 = 4 4
100
T1,present
4
160.936 − 145.529
= 100
160.936
= 9.57%
68
= 1.4(48.0743) + (1 − 1.4)32.8125
= 54.1790C
T2present − T2previous
a 2,1 = ,1
present
,1
100
T2,1
54.1790 − 32.8125
= 100
54.1790
= 39.44%
= 1.4(48.8469) + (1 − 1.4)26.0313
= 57.9732C
T2,2present − T2,2previous
a 2,2 = 100
T2,2present
57.9732 − 26.0313
= 100
57.9732
= 55.10%
= 1.4(93.8936) + (1 − 1.4)23.3899
= 122.095C
69
T2present − T2previous
a 2,3 = ,3
present
,3
100
T2,3
122.095 − 23.3899
= 100
122.095
= 80.84%
= 1.4(200.934) + (1 − 1.4)164.122
= 215.659C
T2present − T2previous
a 2, 4 = ,4 ,4
present
100
T2, 4
215.659 − 164.122
= 100
215.659
= 23.90%
= 1.4(67.6711) + (1 − 1.4)63.9844
= 69.1458C
70
T3present − T3previous
a 3,1 = ,1 ,1
present
100
T3,1
69.1458 − 63.9844
= 100
69.1458
= 7.46%
= 1.4(73.3956) + (1 − 1.4)66.5055
= 76.1516C
T3present − T3previous
a 3, 2 = ,2 ,2
present
100
T3, 2
76.1516 − 66.5055
= 100
76.1516
= 12.67%
= 1.4(129.738) + (1 − 1.4)66.4634
= 155.048C
71
T3present − T3previous
a 3, 3 = ,3 ,3
present
100
T3,3
155.048 − 66.4634
= 100
155.048
= 57.13%
= 1.4(192.677) + (1 − 1.4)220.704
= 181.466C
T3present − T3previous
a 3, 4 = ,4 ,4
present
100
T3, 4
181.466 − 220.704
= 100
181.466
= 21.62%
75
161 216 181 100
75
87 122 155 100
75
51 58 76 100
75
52 54 69 100
x
50 50 50
72
Figure 3.11: Temperature distribution after two iterations
It took nine iterations to get all of the temperature values within 1% error. The table below lists the
temperature values at all nodes after each iteration.
1 2 3 4 5
73
Node Number of Iterations
6 7 8 9
However, the node (m + 1, j ) is not inside the plate. The derivative boundary condition needs to be used
to account for these additional unknown nodal temperatures on the right edge. This is done by
approximating the derivative at the edge node (m, j ) as
74
T Tm+1, j − Tm−1, j
(3.25)
x m, j 2(x)
giving
T
Tm+1, j = Tm−1, j + 2(x) (3.26)
x m, j
T
2Tm−1, j + 2(x) + Tm, j −1 + Tm, j +1 − 4Tm, j = 0 (3.27)
x m, j
T
=0 (3.28)
x m, j
substituting Equation (3.28) in Equation (3.27), gives an equation to use at the Neuman Boundary
condition
Example 4
A plate 2.4 m 3.0 m is subjected to the temperatures and insulated boundary conditions as shown in
Fig. 3.12. Use a square grid length of 0.6 m . Assume the initial temperatures at all of the interior nodes
to be 0 C . Find the temperatures at the interior nodes using the direct method.
300 C
75 C
3 .0 m Insulated
x
50 C
2 .4 m
75
Solution
x = y = 0 . 6 m
L
m=
x
2 .4
=
0 .6
=4
W
n=
y
3
=
0 .6
=5
T0,4
T1,4 T2,4 T3,4 T4,4
T0,3
T1,3 T2,3 T3,3 T4,3
T0,2
T1,2 T2,2 T3,2 T4,2
T0,1
T1,1 T2,1 T3,1 T4,1
T0,0 x
T1,0 T2,0 T3,0 T4,0
All of the nodes on the boundary have an i value of either zero or m . All of the nodes on the boundary
have a j value of either zero or n .
76
T0, j = 75; j = 1,2,3,4
Ti ,0 = 50; i = 1,2,3,4
Ti ,5 = 300; i = 1,2,3,4 (3.30)
T
= 0; j = 1,2,3,4
x 4, j
Now in order to find the temperatures at the interior nodes, we have to write Equation (3.7) for all of the
combinations of i and j . We express this using i from 1 to m − 1 and j from 1 to n − 1 . For the right
side boundary nodes, where i = m = 4 , we have to write Equation (18) for j = 1,2,3,4 . This would give
m n − 1 simultaneous linear equations with m n − 1 unknowns.
i=1 and j=1
T2,1 + T0,1 + T1, 2 + T1,0 − 4T1,1 = 0
77
T3,1 + T1,1 + T2, 2 + 50 − 4T2,1 = 0
78
Now for i = 4 (for this problem m = 4 ), all of these nodes are on the right hand side boundary which is
insulated, so we use Equation (3.29) for j = 1,2,3 and 4 .Substituting i for m variables gives
Equations (3.31) to (3.46) represent a set of sixteen simultaneous linear equations, and solving them gives
the temperature at sixteen interior nodes. The solution is
79
T1,1 76.8254
T 99.4444
1, 2
T1, 3 128.617
T
1, 4 180.410
T2 ,1 82.8571
T2 , 2 117.335
T 159.614
2,3
T2 , 4 218 . 021
T = C
87. 2678
3 ,1
T3, 2 127.426
T3, 3 174.483
T3, 4 232.060
T4 ,1 88. 7882
T 130.617
4, 2
T4 , 3 178.830
232.738
T4 , 4
80
Chapter Four
First Order Non-Linear Equations
Non-linear differential equations are formed by the products of the unknown function and its
derivatives are allowed and its degree is > 1. There are very few methods of solving nonlinear
differential equations exactly; those that are known typically depend on the equation having
particular symmetries. Nonlinear differential equations can exhibit very complicated behavior over
extended time intervals, characteristic of chaos. Even the fundamental questions of existence,
uniqueness, and extendability of solutions for nonlinear differential equations, and well-posedness
of initial and boundary value problems for nonlinear PDEs are hard problems and their resolution
in special cases is considered to be a significant advance in the mathematical theory.
Recall that a first order ODE has the form y' = f (t,y), and is linear if f is linear in y, and nonlinear if f
is nonlinear in y. Examples: y' = ty - et, y' = ty2.
where A is a function of y only and B a function of x only, and dx and dy are differentials. Thus, we could
also write A(y)y′ = B(x). An alternate form is written as follows:
𝑑𝑦
B(x) + A(y) 𝑑𝑥 = 0 (4.2)
(Here, there is obviously a difference of sign in the functions A(y) in the two forms)
When an equation is separated like this, it is then possible to integrate directly on both sides of the
equality to find a solution, which may be given as an implicit function.
Example 1
9y2dy=2xdx
81
3y3=x+C
(Although both integrals would yield a constant of integration, we can combine the two constants into
one.)
Now the function y is given implicitly. We could solve for y to make the solution explicit:
(Note further that since C is an arbitrary constant, we might also just write y = , changing C
to C/3.)
We can of course, solve for the constant C if we wish. (We can do this with the equation in either the
implicit or explicit form.)
Example 2
Solve the initial value problem
, and y(0) = 2
We separate:
Finally, we solve for y, which is not too difficult here. We proceed by multiplying through by 2 and
completing the square:
= x2 + x + 6
y2 + 4y = 2x2 + 2x + 12 (multiply by 2)
y2 + 4y + 4 = 2x2 + 2x + 16 (completing the square)
(y + 2)2 = 2x2 + 2x + 16 (factoring)
82
(We know that it is the positive square root, because we must have y(0) = 2, which is positive.)
We make a further note about our solution: since 2x2 + 2x + 16 > 0 for all values of x, the solution is valid
for all x. If it were not, we would need to restrict the values of x in the solution. Further, note that our
function never has y = −2, so the function is always a solution to the original differential equation which
has y + 2 in the denominator of the right hand side.
Some first order equations may be both separable and linear, and so may be solved different ways:
Example 3
y′ = x − 1 + xy − y is a linear differential equation, as can be seen by rewriting it as y′ = (x − 1) + (x − ′1)y.
We could use an integrating factor exp (x − x /2) for the equation in the form y + (1 − x)y = x − 1, but the
y′ = (x − 1) + (x − 1)y = (x − 1)(y + 1)
So we get
,or
Then we can integrate on both sides to get ln, or (by solving) the equivalent
1, where A = ±eC.
Example 4
y′ = 3y + ex is not separable. Note that if we rewrite it using differentials, the best we can do is
dy = (3y + ex)dx
which is not separated. However, if we rewrite the equation as y′ −3y = ex, we could solve the equation
by an integrating factor of e−3x.
83
Example 5
As we saw above, sometimes a separable equation gives a result that is an implicit function. If it’s not
too hard to solve for y, go ahead. Otherwise, it might be best to leave it in implicit form
Example 6
Find a solution to the initial value problem y′ = cos2(x)cos2(2y), y(π/2) = π:
Thus, we have
Now we need to know find a value of C which can satisfy the initial condition y(π/2) = π. That is, y = π
when x = π/2, so we see that we have tan(2 C, or 0 = π/2 + C, so C = −π/2.
So the solution to the initial value problem is
.
Is it reasonable to solve for y? If we use the arctangent, we have the problem that the range of the
standard branch of the arctangent has range (−π/2,π/2), so that we cannot have y(π/2) = π! Therefore,
we might be better off leaving this solution as implicit.
Note that we could write down a solution by adjusting the arctangent. In fact, given that tan(x) repeats
every π, an explicit solution is given by the following:
84
4.2 Differences between linear and Nonlinear ode
In this section, we will see that first order linear and nonlinear equations differ in a number of
ways, including:
*The theory describing existence and uniqueness of solutions, and corresponding domains, are
different.
*Solutions to linear equations can be expressed in terms of a general solution, which is not usually
the case for nonlinear equations.
*Linear equations have explicitly defined solutions while nonlinear equations typically do not, and
nonlinear equations may or may not have implicitly defined solutions.
*For both types of equations, numerical and graphical construction of solutions are important.
Theorem 1
(4.2)
If the functions p and g are continuous on an open interval (α, β) containing the point t = t 0, then
there exists a unique solution y = φ(t) that satisfies the IVP for each t in (α, β). Hence by the method
of integrating factor Equation (4.2) has a unique solution as follows:
Theorem 2
(4.3)
Suppose f and ∂f/∂y are continuous on some open rectangle (t, y) ∈ (α, β) x (γ, δ) containing the
point (t0, y0). Then in some interval (t0 h, t0 + h) ⊆ (α, β) there exists a unique solution y = φ(t) that
satisfies the IVP.
Proof discussion: Since there is no general formula for the solution of arbitrary nonlinear first order
IVPs, this proof is difficult, and is beyond the scope of this course.
It turns out that conditions stated in Theorem 2 are sufficient but not necessary to guarantee
existence of a solution, and continuity of f ensures existence but not uniqueness of φ
85
Example 3.7: Linear IVP
We have
The solution to this initial value problem is defined for t > 0, the interval on which p (t) = -2/ t is
continuous. If the initial condition is y (-1) = 2, then the solution is given by same expression as
above, but is defined on t < 0. In either case, Theorem1 guarantees that solution is unique on
corresponding interval.
86
and are continuous except on line y = 1. Thus we can draw an open rectangle about (0, -1) on
which f and ∂f/∂y are continuous, as long as it doesn’t cover y = 1. How wide the rectangle will be
is a function defined by the solution (Using separable variable).
With
which are continuous except on line y = 1. If we change initial condition to y(0) = 1, then Theorem
2 is not satisfied. Solving this new IVP, we obtain
87
Thus, a solution exists but is not unique
Thus f continuous everywhere, but ∂f/∂y doesn’t exist at y = 0, and hence Theorem 2 is not
satisfied. Solutions exist but are not unique. Separating variables and solving, we obtain
88
If initial condition is not on t-axis, then Theorem 2 does guarantee existence and uniqueness.
Thus f and ∂f/∂y are continuous at t = 0, so Theorem2 guarantees that solutions exist and are
unique.
Separating variables and solving, we obtain
The solution y(t) is defined on (-∞, 1). Note that the singularity at t = 1 is not obvious from
original IVP statement.
89
Chapter Five
Complex integration
5.1 Cauchy integral theorem and Cauchy integral formulas
Definite integral of a complex-valued function of a real variable
which is assumed to be a piecewise continuous function defined in the closed interval a ≤ t ≤ b. The
integral of f(t) from t = a to t = b, is defined as
90
91
92
93
94
95
5.2 Cauchy integral theorem
Let f(z) = u(x, y)+iv(x, y) be analytic on and inside a simple closed contour C and let f′(z) be also
continuous on and inside C, then
Proof
The proof of the Cauchy integral theorem requires the Green theorem for a positively oriented closed
contour C: If the two real functions P (x, y) and Q(x, y) have continuous first order partial derivatives on
and inside C, then
One can infer from the continuity of f′(z) that u(x, y) and v(x, y)
have continuous derivatives on and inside C. Using the Green theorem, the two real line integrals can be
transformed into double integrals
Both integrands in the double integrals are equal to zero due to the Cauchy-Riemann relations, hence
the theorem.
96