Lecturenote
Lecturenote
1 Laplace Transformation 5
1.1 The Laplace Transform . . . . . . . . . . . . . . . . . . . . . . 5
1.1.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.2 Existence . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.1.3 General Properties of the Laplace Transform . . . . . . 8
1.1.4 Table of Laplace Transforms . . . . . . . . . . . . . . . 13
1.2 The Inverse Laplace Transform . . . . . . . . . . . . . . . . . 15
1.2.1 Method of Partial Fractions . . . . . . . . . . . . . . . 16
1.3 Applications of Laplace Transforms . . . . . . . . . . . . . . . 19
1.3.1 Initial Value Problems . . . . . . . . . . . . . . . . . . 19
1.3.2 Integral Equations . . . . . . . . . . . . . . . . . . . . 22
2 Fourier Analysis 25
2.1 Periodic Functions . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2 Fourier Series . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2.2 Complex Representation of Fourier Series . . . . . . . . 30
2.2.3 Derivation of the Euler Formulas . . . . . . . . . . . . 30
2.2.4 Even and Odd Functions . . . . . . . . . . . . . . . . . 32
2.2.5 Bessel Inequality and Riemann-Lebesgue Lemma . . . 35
2.2.6 Convergence of Fourier Series . . . . . . . . . . . . . . 37
2.2.7 Functions of Arbitrary Period . . . . . . . . . . . . . . 39
2.2.8 Sine and Cosine Series . . . . . . . . . . . . . . . . . . 41
3
4 CONTENTS
3.1.3 Classification . . . . . . . . . . . . . . . . . . . . . . . 47
3.1.4 Physical Derivation of the Heat Equation . . . . . . . . 49
3.1.5 Boundary and Initial Conditions . . . . . . . . . . . . . 50
3.2 Separation of Variables . . . . . . . . . . . . . . . . . . . . . . 52
3.2.1 Overview of Method . . . . . . . . . . . . . . . . . . . 52
3.2.2 The Heat Equation . . . . . . . . . . . . . . . . . . . . 52
3.2.3 The Wave Equation . . . . . . . . . . . . . . . . . . . . 58
3.2.4 Inhomogeneous Equations . . . . . . . . . . . . . . . . 60
A Answers to Exercises 63
Chapter 1
Laplace Transformation
converges for any s, then F (s) is said to be the Laplace transform 1 of f (t).
1
Pierre Simon de Laplace (1749-1827) French mathematician.
5
6 CHAPTER 1. LAPLACE TRANSFORMATION
−4 −2 2 4
−1
1
θ(t)f (t)
−4 −2 2 4
−1
∞ ∞
1 1
Z
ct −st −(s−c)t
F (s) = e e dt = − e = . (1.1.4)
0 s−c 0 s−c
Note that, for the above integral to converge, we must assume s > c.
1.1.2 Existence
Not any function f (t) have a Laplace transform L[f (t)]. For example, it is
2
easy to see that L[et ] does not exist, since its associated integral diverges
as t → ∞. As a rule, f (t) must be of exponential order to have a Laplace
transform. By this we mean that there must exist a constant, say a, such
that
lim |f (t)e−at | = 0. (1.1.5)
t→∞
If this indeed is the case, then by choosing s > a, we see that the integrand
f (t)e−st of (1.1.1) goes to zero as t tend to infinity and, hence, the integral
for L[f (t)] converges. Let us formalize this result by stating it as a theorem.
for some constants a and C, then the Laplace transform F (s) of f (t) exists.
Solution. Using the above linearity and (1.1.4) with c = ±1, we have
Equating the real and imaginary parts of these two equations, we get
s
L[cos ωt] = , (1.1.13)
s2
+ ω2
ω
L[sin ωt] = 2 . (1.1.14)
s + ω2
1.1. THE LAPLACE TRANSFORM 9
As the last examples show, the definition (1.1.1) is rarely the starting
point for deriving Laplace transforms. Instead, one usually first consults a
table of standard transforms, and then tries to adapt any of these to the
problem at hand using a set of general properties, such as the linearity, of
the Laplace transform. Below, we derive a number of other such properties
and illustrate their use.
Theorem 3 (1st Shifting Rule). If f (t) has the Laplace transform F (s)
then for any constant c, we have
Proof. Inserting ect f (t) directly into the definition (1.1.1) gives, with
s > c,
Z ∞ Z ∞
ct ct −st
L[e f (t)] = e f (t)e dt = f (t)e−(s−c)t dt = F (s − c). (1.1.16)
0 0
Proof. Let g(t) = 1t f (t), i.e., f (t) = tg(t). The previous theorem then
gives F (s) = −G′ (s). By the fundamental theorem of calculus, and the fact
that G(s) → 0 as s → ∞, we have
Z ∞
G(s) = F (ω) dω. (1.1.28)
s
1.1. THE LAPLACE TRANSFORM 11
sin t
Example 7. Find the Laplace transform of .
t
Solution. Recall that L[sin t] = (s2 + 1)−1 . Since
sin t
lim = 1, (1.1.29)
t→0 t
the assumptions of the last theorem are satisfied and thus we have
Z ∞
dω π
1
L[ t sin t] = 2
= [arctan ω]∞
s = − arctan s. (1.1.30)
s ω +1 2
A fundamental property of the Laplace transform is the fact that, roughly
speaking, taking the derivative of the original function f (t) corresponds to
multiplying its transform F (s) by s.
Theorem 7. Suppose f (t) and f ′ (t) are two continuous, piecewise smooth
functions satisfying the inequality (1.1.6) for the same values of M and a.
Then, it holds
L[f ′ (t)] = sF (s) − f (0). (1.1.31)
Proof. Integrating by parts, we have
Z ∞
′
L[f (t)] = f ′ (t)e−st dt
0
Z ∞
−st ∞
f (t)e−st dt = −f (0) + sF (s).
= f (t)e 0
+s (1.1.32)
0
Similarly,
L[f (n) (t)] = sn F (s) − sn−1 f (0) − sn−2 f ′ (0) − . . . − f (n−1) (0). (1.1.35)
12 CHAPTER 1. LAPLACE TRANSFORMATION
is given by 1s F (s).
Rt
Proof. Let h(t) = 0 f (τ ) dτ . By construction we then have h′ (t) = f (t)
and h(0) = 0. Applying then the result (1.1.31) to h(t) we immediately get
F (s) = sH(s) − h(0). Hence, H(s) = 1s F (s).
a. t b. t2 c. t3 d. tn
1
e. t + 1 f. (t − 1)2 g. (1 + t)4 h. t
t2
i. e−t j. e3t+4 k. tet l. e
m. cosh t n. cos t o. sin 2t p. sinh2 t
g. sin(ωt + α) h. t sin 2t i. ln t
1
j. t
(1 − cos t) k. cosh t cos t l. cos2 t
1.1. THE LAPLACE TRANSFORM 13
f (t) F (s)
tf (t) −F ′ (s)
e−at f (t) F (s + a)
1
θ(t)
s
tn 1
n! sn+1
1
e−at
s+a
s
cosh at
s2 − a2
a
sinh at
s2 − a2
s
cos bt
s2 + b2
b
sin bt
s2 + b2
t s
sin bt
2b (s2 + b2 )2
1 1
(sin bt − bt cos bt)
2b3 (s2 + b2 )2
a 2 /4t √
√ e−a e−a s
4πt3
e−s e−2t
F (s) = − 4 . (1.2.3)
s2 s
Solution. From a table of Laplace transforms, we have
A + B = 0, C − B = 1, A − C = 2, (1.2.28)
1 1 s+1 1
a. b. c. d.
s+1 s2 +4 s2 + 1 s2 −1
s + 12 s s+1 e−s
e. f. g. h.
s2 + 4s (s + 2)2 (s − 3)4 s
Problem 4. Find the inverse Laplace transform of the following functions.
s s+2 1
a. b. c.
s2 − 2s − 3 s2 + 4s + 5 (s − 2)2 + 9
s+1 3s 1
d. e. f.
s3 + s2 − 6s s2 + 2s − 8 s(s + 1)(s + 2)
1.3. APPLICATIONS OF LAPLACE TRANSFORMS 19
3s + 3 A B (A + B)s − 3A + 2B
= + = , (1.3.9)
(s + 2)(s − 3) s+2 s−3 (s + 2)(s − 3)
3 12
Solving this, we find A = 5
and B = 5
. Hence,
3 12
5 5
Y (s) = + . (1.3.11)
s+2 s−3
Consulting a table of standard Laplace transforms, we finally have
1. Take the Laplace transform of both sides of the given hard problem
for y(t). As a result a simple algebraic equation for Y (s) = L[y(t)] is
obtained.
Example 10. Solve the following initial value problem for t > 0
Solution. We have
or
(s + 3)(s + 1)Y (s) = 3s + 13. (1.3.17)
1.3. APPLICATIONS OF LAPLACE TRANSFORMS 21
Solving for Y (s) and using a decomposition into partial fractions, we get
3s + 13 A B
Y (s) = = +
(s + 3)(s + 1) s+3 s+1
A(s + 1) + B(s + 3) (A + B)s + A + 3B
= = , (1.3.18)
(s + 3)(s + 1) (s + 3)(s + 1)
from which we obtain A = −2, and B = 5. Thus,
2 5
Y (s) = − + . (1.3.19)
s+3 s+1
Recalling (1.1.4) it is obvious that
A simple way to check whether the correct solution has been obtained
is to see if the initial condition is satisfied by the found function y(t). Here
we have y(0) = −2 + 5 = 3 and, since y ′(t) = 6e−3t − 5e−t , we also have
y ′ (0) = 6 − 5 = 1. Hence, the given initial conditions are indeed satisfied by
(1.3.21).
Problem 5. Solve the following differential equations for t > 0
a. y ′ + 2y = e−3t , y(0) = 4.
h. y ′′ + 2y ′ + 2y = 2, y(0) = 0, y ′(0) = 0.
22 CHAPTER 1. LAPLACE TRANSFORMATION
v(t) C
i(t)
or, since R = C = 1,
I(s) 1
I(s) + = (e−s − e−2s ). (1.3.26)
s s
Solving for I(s) we obtain, after elementary manipulations,
1
I(s) = (e−s + e−2s ). (1.3.27)
s+1
1
Noting that L−1 [ s+1 ] = e−t , we then use the 2nd Shifting Rule, to obtain
Hence,
0,
t < 1,
1 −t
i(t) = e e , (1.3.29)
1 < t < 2,
1
(e − e2 )e−t , t > 2.
i(t)
0.5
0 t
1 2 3
−0.5
−1.0
Fourier Analysis
25
26 CHAPTER 2. FOURIER ANALYSIS
f (x + p) = f (x), (2.1.1)
for all x. Any positive number p with this property is called a period of f (x).
For example, f (x) = cos x has periods 2π, 4π, etc. However, the smallest
number p > 0 with the property (2.1.1) is called the prime period, and it
is generally this value that is meant when a function is referred to as being
p-periodic, or, of period p.
0 p 2p
Proof: Let
Z a+p Z a+p Z a
g(a) = f (x) dx = f (x) − f (x) dx. (2.1.3)
a 0 0
1, cos x, sin x, cos 2x, sin 2x, cos 3x, sin 3x, . . . (2.2.2)
Here the term 12 a0 is due to the constant function cos 0 = 1, the factor 12
being included for reasons of later convenience. Further, b0 does not exist,
since sin 0 = 0.
Fourier managed to solve several problems of heat flow using such series
representations, and, as a result, (2.2.3) is today called the Fourier series of
f (x). Similarly, the corresponding numbers an and bn are called the Fourier
coefficients of f (x).
1
Jean Baptiste Fourier (1768-1830) French physicist and mathematician.
2
Recall that if {φ1 , φ2 , φ3 , . . .} is a collections of things (e.g., numbers, vectors, or
functions) that can be multiplied by scalars and added together, then a linear combination
is any expression of the form
c1 φ1 + c2 φ2 + c3 φ3 + . . . , (2.2.1)
Example 11. Find the Fourier series of the 2π-periodic ramp function
6.28
3.14
0
−12.56 −6.28 0.00 6.28 12.56
Figure 2.2: Graph of the ramp function.
and for n = 1, 2, 3, . . .
1 π
Z
an = f (x) cos nx dx
π −π
1 2π
Z
= x cos nx dx
π 0
2π Z 2π
x sin nx 1
= − sin nx dx
π n 0 nπ 0
2π
1
= cos nx = 0. (2.2.8)
πn2 0
Similarly,
1 π
Z
bn = f (x) sin nx dx
π −π
1 2π
Z
= x sin nx dx
π 0
h x cos nx i2π Z 2π
1 2
= − + cos nx dx = − . (2.2.9)
π n 0 nπ 0 n
2.2. FOURIER SERIES 29
3.14
0
−12.56 −6.28 0.00 6.28 12.56
6.28
3.14
0
−12.56 −6.28 0.00 6.28 12.56
Figure 2.3: Partial sums of the Fourier series for the ramp function.
whilst Z π Z π
i(n−k)x
e dx = dx = 2π, for n = k. (2.2.16)
−π −π
Finally, using the fact that an = cn + c−n and that bn = i(cn − c−n ) we find
for n = 0,
1 π
Z
a0 = 2c0 = f (x) dx, (2.2.20)
π 0
and for n = 1, 2, 3 . . .,
an = cn + c−n
Z π π
1 1
Z
= f (x)(einx + e−inx ) dx = f (x) cos nx dx, (2.2.21)
2π −π π −π
bn = i(cn − c−n )
Z π π
i 1
Z
= f (x)(e−inx − einx ) dx = f (x) sin nx dx. (2.2.22)
2π −π π −π
32 CHAPTER 2. FOURIER ANALYSIS
2 g(x) h(x)
1
−6 −4 −2 2 4 6
−1
−2
Example 12. Find the Fourier series of the square wave, defined by
(
+1, 0 < x < π,
f (x) = (2.2.27)
−1, −π < x < 0,
f (x) = f (x + 2π). (2.2.28)
2
1
0
−1
−2
−9.42 −6.28 −3.14 0.00 3.14 6.28 9.42
Figure 2.5: Graph of the square wave.
2 π 2 h cos nx iπ 2 (−1)n − 1
Z
bn = sin nx dx = − =− , (2.2.29)
π 0 π n 0 π n
2 4
(sin x + 13 sin 3x + 51 sin 5x)
π
1
2 4
(sin x + 31 sin 3x + . . . + 1
sin 17x)
π 17
1
Figure 2.6: Partial sums of the Fourier series for the square wave.
∞ π
1
X Z
2
|cn | ≤ |f (x)|2 dx. (2.2.33)
−∞
2π −π
Proof. Let
N
X
f
SN (x) = cn e−inx , (2.2.34)
n=−N
denote the N-th partial sum of the Fourier series for f (x). We then have
N
X
f
|f (x) − SN (x)|2 = |f (x) − cn einx |2
n=−N
N
! N
!
f¯(x) −
X X
= f (x) − cn einx c̄n e−inx
n=−N n=−N
N
cn f¯(x)einx + c̄n f (x)e−inx
X
= |f (x)|2 −
n=−N
N
X N
X
+ cm c̄n ei(m−n)x , (2.2.35)
n=−N m=−N
π N
1
Z X
0≤ |f (x) − cn einx |2 dx
2π −π n=−N
π N Z π Z π
1 1 X
Z
= |f (x)| dx − 2
cn f¯(x)e dx + c̄n
inx ¯
f(x)einx
dx
2π −π 2π n=−N −π −π
N N Z π
1 X X
+ cm c̄n ei(m−n)x dx. (2.2.36)
2π n=−N m=−N −π
36 CHAPTER 2. FOURIER ANALYSIS
Now, using the orthogonality (2.2.14) and the formula (2.2.19) for cn , this
simplifies, viz.,
π N N
1
Z X X
2
0≤ |f (x)| dx − (cn c̄n + c̄n cn ) + |cn |2
2π −π n=−N n=−N
π N
1
Z X
= |f (x)|2 dx − |cn |2 . (2.2.37)
2π −π n=−N
lim cn = 0, (2.2.38)
|n|→∞
converge to f (x).
The question of pointwise convergence, for example, concerns whether
f
lim SN (x0 ) = f (x0 ), (2.2.43)
N →∞
f f (x0− ) + f (x0+ )
lim SN (x0 ) = , (2.2.44)
N →∞ 2
f
for all points x0 . In particular, limN →∞ SN (x) = f (x0 ) for every point x0 at
which f (x) is continuous.
i.e., the average error between f (x) and its Fourier series tends to zero as N
tend to infinity. However, this means that the Bessel inequality (2.2.33) is
actually a genuine equality. Usually, it is then called the Parseval Indentity.
38 CHAPTER 2. FOURIER ANALYSIS
or, equivalently,
π ∞
1
Z X
2
|f (x)| dx = 1
|a |2
2 0
+ |an |2 + |bn |2 . (2.2.47)
π −π n=1
2 π 2 π
Z Z
an = |x| cos nx dx = x cos nx dx. (2.2.49)
π 0 π 0
Thus, for n = 0, π
π
2 2 x2
Z
a0 = x dx = = π, (2.2.50)
π 0 π 2 0
and for n > 0,
π
2 π 2 π sin nx
2 x sin nx
Z Z
an = x cos nx dx = − dx
π 0 π n 0 π 0 n
2 h cos nx iπ 2 cos nπ − 1 2 (−1)n − 1
= = = . (2.2.51)
π n2 0 π n2 π n2
since sin nπ = 0 and cos nπ = (−1)n . Consequently,
∞
π 2 X (−1)n − 1
f (x) = + cos nx. (2.2.52)
2 π n=1 n2
Furthermore,
π π
2 3
Z Z
2
|f (x)| dx = 2 x2 dx = π . (2.2.53)
−π 0 3
2.2. FOURIER SERIES 39
i.e.,
2 2 π 2 16
1 1 1 1
π = + 2 + + + + ... . (2.2.55)
3 2 π 14 34 54 74
Hence,
1 1 1 4 π4
+ + + + . . . = . (2.2.56)
14 34 54 74 96
Problem 12. Compute the sum
∞
X 1
2
, (2.2.57)
n=1
n
can be written as
∞
1
X nπx nπx
a
2 0
+ an cos + bn sin . (2.2.60)
n=1
L L
40 CHAPTER 2. FOURIER ANALYSIS
Example 14. Find the Fourier series of the rectified sine wave,
Hence, we have
∞
2 4X 1
f (x) = − 2
cos 2nπx. (2.2.65)
π π n=1 (4n − 1)
2.2. FOURIER SERIES 41
Problem 13. Let the function f (x) be periodic of period 4 and let
f (x) = |x|, |x| < 2.
a. Graph f (x) for |x| < 6.
b. Find the Fourier series of f (x).
c. Use the series obtained to compute the sum
1 1 1
1+ 2
+ 2 + 2 + ...
3 5 7
d. Show, using the Parseval identity, that
1 1 1 π4
1+ + + + . . . = .
34 54 74 96
2.2.8 Sine and Cosine Series
In various physical and engineering problems there is a practical need to
apply trigonometric series expansions to functions f (x) only defined for a
finite interval, say 0 ≤ x ≤ L. This can be accomplished by first defining
f (x) for −L ≤ x ≤ 0 and then extending it to a periodic function of period
2L. In doing so, we can use the freedom of definition to make the extension
either even or odd as is convenient.
Example 15. Develop a sine series for the function f (x) = cos x, 0 ≤ x ≤ π.
1
2 π
Z
bn = cos x sin nx dx
π 0
1 π 1 π
Z Z
= sin(n − 1)x dx + sin(n + 1)x dx
π 0 π 0
π π
1 cos(n − 1)x 1 cos(n + 1)x
= − + −
π n−1 0 π n−1 0
n n
1 (−1) − 1 1 (−1) − 1 2n (−1)n − 1
= + = . (2.2.73)
π n−1 π n+1 π n2 − 1
Hence, the sine series of cos x is given by
∞
X n (−1)n − 1
2 sin nx. (2.2.74)
n=1
π n2 − 1
2.2. FOURIER SERIES 43
Problem 14. Classify the following functions as even, odd, or neither even
nor odd.
a. ex b. x sin x c. ln x
2
d. sin2 x e. e−x f. 1 + x + x5 + x7
Problem 15. Develop a cosine series for the function
f (x) = x, 0 ≤ x ≤ 1.
f (x) = x2 , 0 ≤ x ≤ π.
Problem 17. Let the function f (x) be odd and periodic of period 2π, where
45
46 CHAPTER 3. PARTIAL DIFFERENTIAL EQUATIONS
is linear in u, even though its so-called coefficients esin y and log(1 + x3 ) are
highly non-linear functions of x and y. However, the much simpler-looking
minimal surface equation
u = c1 u 1 + c2 u 2 , (3.1.6)
3.1.3 Classification
Let it suffice to classify linear second-order equations of the type
Here, A = 1, B = u, and C = 1 + u2 , so
c. uxy + x1 uy = 0.
b. uxx + x3 uyy = 0.
3.1. BASIC CONCEPTS 49
However, since the length of the rod L is arbitrary, we must conclude that
et + jx = 0. (3.1.14)
Experimental studies show that the relation between the total thermal
energy H of a body of mass m and its temperature u is roughly linear and
given by
H = cmu, (3.1.15)
where c is known as the specific heat capacity of the material. Because our
rod is long and thin, it is reasonable to believe that its temperature u only
varies with longitudinal position x and time t, i.e., u = u(x, t). Moreover,
3
It originates back to Fourier and his thesis Theorie Analytique de la Chaleur (1822).
50 CHAPTER 3. PARTIAL DIFFERENTIAL EQUATIONS
RL
since we have defined the total energy of the rod by H = 0
e dx, it follows
that
e = ̺cu, (3.1.16)
where ̺ is the mass density of the rod.
Heat always flow from warm regions to cooler ones, but at rate which is
dependent upon the conducting medium. For example, if a copper rod and
an iron rod are joined together end to end, and the ends are subsequently
heated, the heat will conduct through the copper end more quickly than the
iron end because copper has a higher thermal conductivity.
Fourier considered the above properties and summarized them elegantly,
viz.,
j = −κux , (3.1.17)
a relation which determines the heat flux j for a given temperature profile u
and thermal conductivity κ.
Substituting (3.1.16) and (3.1.17) into (3.1.14) and assuming κ constant,
we finally arrive at
̺cut = κuxx , (3.1.18)
which is generally know as the heat equation.
Boundary and Initial Conditions for the Heat Equation. Since the
heat equation contains a total of three derivatives, two spatial ones and
one with respect to time, we expect its solution to contain three degrees of
freedom, that is, constants of integration. Consequently, it is necessary to
supply three auxiliary constraints along with this equation to get a unique
solution.
Regarding our heat conducting rod, we must thus know how heat flows
through the ends of the rod, and we must know the temperature distribution
within the rod at some initial time.
If the ends of the rod are perfectly insulated, so that the heat flux across
these at x = 0 and x = L are zero, then by (3.1.17) we have the boundary
conditions
ux (0, t) = 0, ux (L, t) = 0, t > 0. (3.1.20)
On the other hand, if the ends of the rod are held at the fixed temperature
zero, e.g., by submerging them into ice water, then we have the boundary
conditions
u(0, t) = 0, u(L, t) = 0, t > 0. (3.1.21)
Either of these boundary conditions can be inhomogeneous, i.e., have a non-
zero right hand side, and we could of course also have mixed conditions,
where one end of the rod is insulated and the other end is held at a constant
temperature.
A boundary condition that specifies the value of the solution is called a
Dirichlet condition, while a condition that specifies the value of a derivative
is called a Neumann condition. If the boundary condition is of mixed type,
then it is called a Robin condition.
To completely determine the temperature as a function of time t and
space x, we must also specify the temperature profile at some initial time t0
via a so-called initial condition
u(x, t0 ) = f (x). (3.1.22)
52 CHAPTER 3. PARTIAL DIFFERENTIAL EQUATIONS
ut = kuxx , 0 ≤ x ≤ L, t ≥ 0, (3.2.2)
u(0, t) = u(L, 0) = 0, (3.2.3)
u(x, 0) = f (x). (3.2.4)
impossible, unless both sides are equal to a constant, λ say. Hence, we arrive
at
1 T ′ (t) X ′′ (x)
= = λ, (3.2.7)
k T (t) X(x)
where λ is called the separation constant. Rearranging, we have
which is two simple ordinary differential equations for X and T that can be
solved by any elementary method. Solving the ODE for T by inspection, we
have
T (t) = Cekλt , (3.2.9)
where C is arbitrary. Contemplating this result, we see that T (t) increases
exponentially with time, if λ ≥ 0. However, without any additional heating
mechanism, such a behavior is obviously unphysical for the temperature u
of the rod. Therefore, we choose the separation constant negative. To force
this we let
λ = −µ2 . (3.2.10)
As a result, the ODE for X is then
Thus,
X(x) = A cos µx + B sin µx, (3.2.12)
where A and B are constants of integration. Multiplying X and T , we get
2
u(x, t) = X(x)T (t) = e−µ t (A cos µx + B sin µx). (3.2.13)
i.e., A = 0, and
2
u(L, t) = Be−µ t sin µL = 0, (3.2.15)
54 CHAPTER 3. PARTIAL DIFFERENTIAL EQUATIONS
or, equivalently
n2 π 2
λ=− , n = 1, 2, 3, . . . (3.2.17)
L2
2 n2 t/L2 nπx
un (x, t) = Bn e−kπ sin , n = 1, 2, 3 . . . , (3.2.18)
L
but, since (3.2.2) obeys the superposition principle, its general solution is
given by a linear combination of these individual solutions un , that is to say,
∞
X 2 n2 t/L2 nπx
u(x, t) = Bn e−kπ sin . (3.2.19)
n=1
L
∞
X nπx
u(x, 0) = Bn sin = f (x), (3.2.20)
n=1
L
and the problem is to choose the numbers Bn , so that the series for u(x, 0)
equals f (x). However, recognizing (3.2.20) as a Fourier sine series, we know
that this can be done by choosing Bn as
L
2 nπx
Z
Bn = f (x) sin dx. (3.2.21)
L 0 L
1
As a particular example, let us consider the case L = 1, k = 10
, and
3.2. SEPARATION OF VARIABLES 55
Solution. Put u(x, t) = X(x)T (t) and plug it into (3.2.24). We get
T ′ (t) X ′′ (x)
= = λ, (3.2.27)
T (t) X(x)
i.e., B = 0, and
2
ux (1, t) = −Ae−tµ sin µ = 0, (3.2.30)
3.2. SEPARATION OF VARIABLES 57
This last result suggests that we should expand the initial condition f (x)
using a cosine series. In doing so, it is better to write the solution u(x, t),
viz.,
∞
X 2 n2 t
u(x, t) = 12 A0 + An e−π cos nπx.
n=1
and for n = 1, 2, 3, . . .,
Z 1 Z 1
An = 2 f (x) cos nπx dx = 2 cos nπx dx
0 1/2
1 −1, if n = 1, 5, 9, . . .
sin nπx 2
=2 = 0, if n = 2, 4, 6, . . . (3.2.32)
nπ 1/2 nπ
1, if n = 3, 7, 11, . . .
so if we let
q
s′′ (x) = − , (3.2.53)
k
then v(x, t) satisfies the homogeneous equation vt = kvxx . However, we also
desire homogeneous boundary conditions for v(x, t). To accomplish this, we
note that
which suggest that we put s(0) = α and s(L) = β, to get v(0, t) = 0 and
v(L, t) = 0, as desired.
As a last step, the ansatz (3.2.51) is inserted into the initial condition,
i.e.,
u(x, 0) = v(x, 0) + s(x) = 0. (3.2.56)
yielding the initial condition for the v(x, t) problem as
v(0, t) = −s(x). (3.2.57)
Hence, the original inhomogeneous problem has been converted into two
problems, each of which is straightforward to solve. First, s(x) is found by
solving
q
s′′ (x) = − , s(0) = α, s(L) = β. (3.2.58)
k
Next, v(x, t) is determined by applying separation of variables to
vt = vxx , 0 < x < L, t > 0, (3.2.59)
v(0, t) = 0, v(L, t) = 0, (3.2.60)
v(x, 0) = −s(x). (3.2.61)
Finally, the solutions to the two separate problems given by (3.2.58) and
(3.2.59) are added to give the desired space-time temperature distribution,
u(x, t).
Problem 28. Solve the heat equation
ut = uxx + sin 3πx, 0 < x < 1, t > 0,
u(0, t) = 0, u(1, t) = 0,
u(x, 0) = sin πx.
Problem 29. Solve the problem
ut = uxx + sin πx, 0 < x < 1, t > 0,
u(0, t) = 0, u(1, t) = 0,
u(x, 0) = 1.
Problem 30. Solve the wave equation
uxx = utt − 1, 0 < x < 1, t > 0,
u(0, t) = 0, u(1, t) = 1,
u(x, 0) = x, ut (x, 0) = 0.
62 CHAPTER 3. PARTIAL DIFFERENTIAL EQUATIONS
Appendix A
Answers to Exercises
Problem 1
a. 1/s2 b. 2/s3
c. 6/s4 d. n!/sn+1
e. 1/s2 + 1/s f. 2/s3 − 2/s2 + 1/s
g. 24/s5 + 24/s4 + 12/s3 + 4/s2 + 1/s h. do not exist
i. 1/(s + 1) j. e4 /(s − 3)
k. 1/(s − 1)2 l. do not exist
2
m. s/(s − 1) n. s/(s2 + 1)
o. 2/(s2 + 4) p. 2/(s(s2 − 4))
Problem 2
a. (s − a)/((s − a)2 + b2 ) b. e−s /s
c. e−(s+1) /(s + 1) d. 2(3s2 + 1)/(s2 − 1)3
e. 6/(s − 1)4 f. s(2 + s)/(s2 + 2s + 2)2
g. (ω cos α + s sin α)/(s2 + ω 2) h. s/(s2 + 1/4)2
1
i. do not exist j. 2
ln(1 + 1/s2)
k. s3 /(s4 + 4) l. (2 + s2 )/(s(s2 + 4))
Problem 3
63
64 APPENDIX A. ANSWERS TO EXERCISES
1
a. e−t b. 2
sin 2t c. cos t + sin t d. sinh t
Problem 4
d. − 61 − 15
2 −3t 3 2t
e + 10 e e. 2e−4t + e2t f. 1
2
+ 21 e−2t − e−t
Problem 5
b. y = e2t − 2et
Problem 7
Problem 8
∞
1 X 1 − (−1)n
f (x) = + sin nx
2 n=1 nπ
1 2 1 1
= + sin x + sin 3x + sin 5x + . . .
2 π 3 5
Problem 9
∞ ∞
π X (−1)n − 1 X (−1)n+1
f (x) = + cos nx + sin nx
4 n=1 πn2 n=1
n
π 2 1 1
= − cos x + cos 3x + cos 5x + . . .
4 π 9 25
1 1 1
+ sin x − sin 2x + sin 3x − sin 4x + . . .
2 3 4
Problem 10
∞
π2 X (−1)n
x2 = +4 2
cos nx
3 n=1
n
π2
1 1
= + 4 − cos x + cos 2x − cos 3x + . . .
3 4 9
Let x = 0 to get
∞
X (−1)n+1 1 1 1 π2
=1− + − + ... = .
n=1
n2 4 9 16 12
Problem 13
a. -
b.
8 πx 1 3πx 1 5πx 1 7πx
|x| = 1 − 2 cos + cos + cos + cos + ...
π 2 9 2 25 2 49 2
c. π 2 /8 (put x = 0)
d. -
Problem 14
a. Neither odd nor even b. Even c. Neither odd nor even
Problem 15
∞
1 X (−1)n − 1
f (x) = + 2 cos nπx
2 n=1
n2 π 2
Problem 16
∞
π2 X (−1)n
f (x) = +4 2
cos nx
3 n=1
n
Problem 17
a. -
∞
X sin nx
b. f (x) = −2
n=1
n
c. π/4
Problem 19
67
Problem 20
∞
X 4 −n2 π2 t
u(x, t) = e sin nπx
n=1,3,5,...
nπ
Problem 21
Problem 22
∞
4 X (−1)k − 1 (2k+1)2 π2 t
u(x, t) = e 4 sin(2k + 1)πx
π 2 k=0 (2k + 1)2
4 1
− π2 t 1 − 9 π2 t 1 − 25 π2 t
= 2 e 4 sin πx − e 4 sin 3πx + e 4 sin 5πx − . . .
π 9 25
Problem 23
2 +1) t
u(x, t) = e−(π sin πx
Problem 24
∞
X 4 −(n2 π2 +1)t
u(x, t) = e sin nπx
n=1,3,5,...
nπ
Problem 25
∞
1 4 X 1 −n2 π2 t
u(x, t) = − 2 2
e cos nπx
2 π n=1,3,5,...
n
Problem 26
1
u(x, t) = cos 6t sin 3x + sin 8t sin 4x
4
68 APPENDIX A. ANSWERS TO EXERCISES
Problem 27
∞
16 X 1
u(x, t) = − 3 3
(sin(k + 21 )π(x + t) + sin(k + 21 )π(x − t))
π k=0 (2k + 1)
Problem 28
2
−π 2 t 1 − e−9π t
u(x, t) = e sin πx + sin 3πx
9π 2
Problem 29
2t
!
4 −π2 t 1 − e−π
u(x, t) = e + sin πx
π π2
∞
4X 1 2 2
+ e−(2k+1) π t sin(2k + 1)πx
π 2k + 1
k=0
Problem 30
∞
x2 3x X 4
u(x, t) = − + − 3 3
sin nπx cos nπt.
2 2 n=1,3,5,...
π n