Fis Mat 1
Fis Mat 1
PHYSICS I
− Albert Einstein.
− Eugene Wigner.
Mathematical physics is in the first place physics and it could not exist without
experimental investigations
− Peter Debye.
C ONTENTS
1 F OURIER A NALYSIS 1
1.1 T HE INNER PRODUCT . . . . . . . . . . . . . . . . . . . 5
1.2 T HE F OURIER SERIES . . . . . . . . . . . . . . . . . . . 16
1.3 C ONVERGENCE OF THE EXPANSION . . . . . . . . . . . . . 26
1.4 F OURIER SERIES AND NON - PERIODIC FUNCTIONS . . . . . . . 34
1.5 I NTEGRATION AND DIFFERENTIATION OF THE SERIES . . . . . . 37
1.6 T HE F OURIER TRANSFORM. . . . . . . . . . . . . . . . . 41
1.7 P ROPERTIES OF F OURIER TRANSFORMS . . . . . . . . . . . . 51
1.8 F OURIER TRANSFORM IN MORE THAN ONE DIMENSION . . . . . 54
1.9 OTHER INTEGRAL TRANSFORMS : T HE L APLACE TRANSFORM . . . 56
1.10 P ROPERTIES OF THE L APLACE TRANSFORM . . . . . . . . . . 63
1.11 A PPENDIX : D IRAC DELTA FUNCTION . . . . . . . . . . . . 65
1.12 A PPENDIX : T HE L EBESGUE INTEGRAL . . . . . . . . . . . . 74
1.13 F URTHER READING . . . . . . . . . . . . . . . . . . . 79
i
CONTENTS ii
1
F OURIER A NALYSIS
JOSEPH FOURIER
Until now, you have probably used a Taylor expansion whenever you had
to deal with the problem of representing a certain function f (x) in the neigh-
borhood of some point, say x = x 0 . And, as you already know, the Taylor series
expansion of a function f (x) consists in representing the function as an infi-
nite power series in the polynomials (x − x 0 ), and is given by
¯
X∞ 1 dk f ¯ X∞
f (x) = ¯ (x − x ) k
≡ fk xk . (1.1)
k ¯ 0
k=0 k! d x x=x 0 k=0
Thus, to find the Taylor series expansion of f (x) we simply need to take deriva-
tives of f (x) evaluated only at the point of reference, x = x 0 . But, of course, we
can do this if the function f (x) is infinitely differentiable at x = x 0 . For exam-
ple, let us consider what we might call a complicated function
³ x ´3
f (x) = ln(cos(x 2 ) + 2) + , (1.2)
3
which we plot, together with its Taylor expansion around x 0 = 2, in Fig. 1.1.
As we can see in Fig. 1.1, within few terms of the Taylor series, we can get a
good estimation of the function f (x) in Eq. (1.1) in the neighborhood of x 0 = 2.
However, we cannot get the behavior of f (x) in a larger interval, outside the
region x ∼ 2, let’s say −3 ≤ x ≤ 4. To do this, we need functions with more
1
2
Figure 1.1: Plot of f (x) in Eq. (1.2) (solid line) and its Taylor expansion around
x 0 = 2 including 1, 2, 3 and 4 terms only of the expansion.
Zb
d x ω(x) | f (x)|2 (1.3)
a
is defined, and where ω(x) refers to a weight function (a strictly positive real-
valued function). We see then that all this bears a strong formal resemblance
to the problem of expressing a vector in n-dimensional space as a linear com-
bination of n linearly independent vectors. If we consider a vector | f 〉, we
3
where {|e k 〉}nk=0 is an orthonormal basis and f k some coefficients. When con-
sidering Eq. (1.1), we are doing basically the same: if f is a scalar function of
x,
X
∞
f (x) = f k e k (x), (1.5)
k=1
where {e k (x)} would be the basis. If f (x) is infinitely differentiable in the con-
sidered interval, one solution for Eq. (1.5) is x k−1 , i.e.,{e 0 (x), e 1 (x), . . . , e n (x), . . . } =
{1, x, x 2 , . . . , x n , . . . }. If the function f (x) has a finite discontinuity at a finite
number of points x di , i = 1, 2, . . . , N , in the interval considered, where at each
x di the left-hand and right-hand limits of f , i.e.,
exists1 , it is often possible to divide up the interval into subintervals such that
in each subinterval the function f (x) is continuous and monotonic. Such a
f (x) is called piecewise continuous. Notice that every continuous function
is a piecewise continuous function. In other words, f is piecewise continu-
ous if its graph is a smooth curve except for finitely many jumps (where f is
discontinuous) and corners (where d f /d x is discontinuous), but we do not
allow infinite discontinuities (such as f (x) = 1/x has at x = 0). If in addition
f (x) has a piecewise continuous first derivative, then f (x) is called piecewise
smooth. Some examples of piecewise continuous/piecewise smooth func-
tions are shown in Fig. 1.2. We can then obtain a representation of a piecewise
continuous function of the form (1.5).
The general idea of the Fourier analysis is to study how general functions
can be decomposed into trigonometric or exponential functions with definite
frequencies. There are two types of Fourier expansions:
Figure 1.2: Top panel: (Left) The function f (x) = |x|, with −π ≤ x ≤ π. Note
that f is continuous throughout the interval and its derivative is discontin-
½ 2at x = 0. Then, f (x) is piecewise smooth (Right) The function
uous only
x , −π < x < 0
f (x) = . Both the function and its derivative are continu-
x 2 + 1, 0 ≤ x < π
ous except at x = 0. Thus, the function is piecewise smooth. Bottom panel: the
function f (x) = x 1/3 on any interval that includes x = 0 is continuous, but its
derivative is not piecewise continuous, since d f /d x = 1/(3x 2/3 ) is ∞ at x = 0.
Thus, the function is not piecewise smooth. In other words, any region in-
cluding x = 0 cannot be broken up into pieces such that d f /d x is continuous.
½ 1/2
x , x <0
A function, like for example, f (x) = is piecewise continuous,
x 2 + 1, x ≥ 0
df
but its derivative is not, since lim− d x → ∞, where 0− means that we approach
x→0
0 from the left side. Thus, the function f (x) is not piecewise smooth.
function can be written in terms of sinusoidal functions, we can limit our at-
tention to these functions when solving the differential equations. And then
we can build up any other function from these special ones. This is a very
helpful strategy, because it is invariably easier to deal with sinusoidal func-
tions than general ones. Fourier analysis shows up, for example, in classical
mechanics and the analysis of normal modes, in electromagnetism and the
frequency analysis of waves, in noise considerations and thermal physics, in
quantum theory and the transformation between momentum and coordinate
representations, and in quantum field theory and the creation and annihila-
tion operation formalism.
X
n
〈v|w〉 = v 1 w 1 + v 2 w 2 + · · · + v n w n = vk wk . (1.7)
k=1
Observe that we are using a new symbol | 〉 to denote a generic vector. This
object is called ket and this nomenclature is due to Dirac. Since |v〉 and |w〉
are uniquely specified by their components in a given basis, we may, in this
basis, write them as column vectors. In this way,
v1 w1
v2 w2
|v〉 = .. , |w〉 = .. . (1.8)
. .
vn wn
The inner product 〈v|ω〉 is given by the matrix product of the transpose con-
jugate of the column vector representing |v〉, thus, a row vector, which we
denote as 〈v|, and the column vector representing |w〉, i.e.,
w1
¡ ¢
w2 X
n
〈v|w〉 = v 1∗ v 2∗ ... v n∗ .. = v ∗w . (1.9)
. k=1 k k
wn
The symbol 〈 | is called bra and 〈v| denotes the transpose conjugate of |v〉.
In case of vectors in Rn , {v k∗ }nk=1 = {v k }nk=1 , and Eq. (1.9) reduces to Eq. (1.7).
1.1 T HE INNER PRODUCT 6
〈v|v〉 = |v 1 |2 + |v 2 |2 + · · · + |v n |2 , (1.10)
is the sum of the modulus squares of its elements, and hence equal to the
square of its length. Therefore, the Euclidean norm or length of a vector is
found by taking the square root, i.e.,
p q
||v|| = 〈v|v〉 = |v 1 |2 + |v 2 |2 + · · · + |v n |2 . (1.11)
1. Bilinearity,
2. Symmetry,
3. Positivity,
The positivity axiom implies that ||v|| ≥ 0 is real and non-negative, and equals
0 if and only if |v〉 = |0〉 is the zero vector.
A vector space equipped with an inner product is called an inner product
space, and a given vector space can admit many different inner products (if
you wish, you can easily verify the inner product axioms for the Euclidean dot
product).
〈v|w〉 = 2v 1 w 1 + 5v 2 w 2 , (1.17)
which verifies the first bilinearity condition (the second follows by a very sim-
ilar computation). Moreover,
〈v|v〉 = 2v 12 + 5v 22 ≥ 0, (1.19)
is clearly strictly positive for any |v〉 6= |0〉 and equal to zero when |v〉 = |0〉,
which proves positivity and hence establishes Eq. (1.17) as a legitimate inner
product on R2 . The associated weighted norm is
q
||v|| = 2v 12 + 5v 22 . (1.20)
The numbers c i > 0 are the weights: the larger the weight c i , the more the
i th coordinate of v contributes to the norm. Weighted norms are particularly
important in statistics and data fitting, where one wants to emphasize certain
quantities and de-emphasize others; this is done by assigning suitable weights
to the different components of the data vector |v〉.
1.1 T HE INNER PRODUCT 8
Zb
〈 f |g 〉 = d x f ∗ (x)g (x), (1.22)
a
defines an inner product on the vector space (which can be easily proven).
The associated norm is, according to the basic definition in Eq. (1.16),
v
u b
q uZ
u
|| f || ≡ 〈 f | f 〉 = t d x | f (x)|2 , (1.23)
a
which is known as the norm of the function f over the interval [a, b] and plays
the same role in infinite-dimensional function space that the Euclidean norm
or length of a vector plays in the finite-dimensional Euclidean vector space
Rn . One can also define weighted inner products. The weights along the in-
terval are specified by a continuous positive scalar function ω(x) > 0. The
corresponding weighted inner product and norm are
v
u b
Zb uZ
u
〈 f |g 〉 = d x ω(x) f ∗ (x)g (x), || f || = t d x ω(x)| f (x)|2 . (1.24)
a a
Functions for which such an integral exists and is finite are said to be square-
integrable over the interval [a, b] and the space of square-integrable functions
over the interval [a, b] is denoted by L ω2 (a, b). In this notation, L represents
the name Lebesgue, who generalized the notion of the ordinary Riemann in-
tegral to cases for which the integrand could be highly discontinuous; the su-
perscript 2 indicates the integrability of the square of the modulus of each
function; the values a and b denote the limits of integration and ω refers to the
weight function. When ω(x) = 1, we use the notation L 2 (a, b). Every piece-
wise continuous function defined in a bounded interval belongs to L ω2 (a, b),
but some functions with singularities are also members of the square inte-
grable functions. This includes such non-piecewise continuous functions as
sin(1/x) and x −1/3 , as well as the strange function
½
1, if x is a rational number,
r (x) = (1.25)
0, if x is irrational
1.1 T HE INNER PRODUCT 9
£ ¤
E XAMPLE 1.2. If we take [a, b] = 0, π2 , then the L 2 inner product between
f (x) = sin(x) and g (x) = cos(x) is equal to
π
Z2 ¯x=π/2
1 ¯ 1
〈 f |g 〉 = d x sin(x)cos(x) = − cos (x)¯¯
2
= . (1.26)
2 x=0 2
0
and not 1, as you might have expected. It is also important to note that the
value of the norm depends upon which interval the integral is take over. For
instance, on the longer interval [0, π],
v
uZπ
u p
u
|| f || = t d x 12 = π. (1.29)
0
Thus, when dealing with the L 2 inner product or norm, we must always be
careful to specify the function space, or, equivalently, the interval on which it
is being evaluated.
for every |v〉, |ω〉 ∈ V . Equality holds if and only if |v〉 and |ω〉 are parallel
vectors. In case of the L ω2 (a, b) inner product on function space,
v v v
u b u b u b
uZ uZ uZ
u u u
t d x ω(x)| f (x) + g (x)| ≤ t d x ω(x)| f (x)| + t d x ω(x)|g (x)|2 .
2 2
a a a
(1.33)
Given any inner product on a vector space, we can use the quotient
〈v|w〉
cosθ = , (1.34)
||v|| ||w||
to define the “angle” between the elements |v〉, |w〉 ∈ V . The Cauchy-Schwarz
inequality tells us that the ratio lies between −1 and +1, and hence the angle θ
is well-defined, and in fact, unique of we restrict it to lie in the range 0 ≤ θ ≤ π.
For example, using the standard dot product on R3 , the angle between the
vectors |v〉 = 1|i 〉 + 1|k〉 and w = 1| j 〉 + 1|k〉 is given by
1 1 π
cosθ = p p = =⇒ θ= , (1.35)
2 2 2 3
i.e., 60◦ . Similarly, the “angle” between the functions f (x) = x and g (x) = x 2
defined on the interval [0, 1] is given by
R1
2
d x x3 1 r
〈x|x 〉 0 4 15
cosθ = =s s =q q = , (1.36)
||x|| ||x 2 || R1 R1 1 1 16
d x x2 d x x4 3 5
0 0
so that θ ' 0.25268 radians. Of course, one should not try to give this no-
tion of angle between functions more significance than the formal definition
warrants. It does not correspond to any “angular” properties of their graphs.
Also, the value depends on the choice of inner product and the interval upon
which it is being computed. But even in Euclidean space Rn , the measure-
ment of angle and length depends upon the choice of an underlying inner
product. Different inner products lead to different angle measurements; only
1.1 T HE INNER PRODUCT 11
for the standard Euclidean dot product does angle correspond to our every-
day experience. But the important point is that using the Schwarz inequality,
which holds for any inner product, one can show that the integral in Eq. (1.24)
is defined.
As you already know, two elements |v〉, |w〉 ∈ V of an inner product space
V are called orthogonal if their inner product 〈v|w〉 = 0. For example, the
vectors |v〉 = 1|i 〉 + 2| j 〉 and |w〉 = 6|i 〉 − 3| j 〉 are orthogonal with respect to the
Euclidean dor product in R2 , since 〈v|w〉 = 1·6+2·(−3) = 0. But, interestingly,
the functions, for example, f (x) = x and q(x) = x 2 − 21 are orthogonal with
R1
respect to the inner product 〈 f |g 〉 = d x f ∗ (x)g (x) on the interval [0, 1], since
0
¿ ¯ À Z1 µ ¶ Z1 ³
¯ 2 1 2 1 x´
x ¯x − = dx x x − = d x x3 − = 0. (1.37)
2 2 2
0 0
¿ ¯ À Z2 µ ¶ Z2 ³
¯ 2 1 2 1 x´
x ¯x − = dx x x − = d x x3 − = 3. (1.38)
2 2 2
0 0
Z1 ¯
x 2 ¯¯1 1
〈1|x〉 = dx x = ¯ = . (1.39)
2 x=0 2
0
There are two important theorems for the space L ω2 (a, b) (which we are
not going to proof). One of them is the Stone-Weierstrass theorem: the se-
quence of monomials {x k }∞k=0
are linearly independent and they form a basis
of L ω (a, b). Then, a piecewise continuous function f (x) ∈ L ω2 (a, b) can be
2
written as
X
∞
f (x) = fk xk . (1.40)
k=0
This is very interesting because we can associate vectors with the func-
tion f (x) and the monomials {x k }∞ k=0
in the space L ω2 (a, b). Indeed, if we take
different values of x (let’s call them x 1 , x 2 , etc.) in the interval [a, b], we can
consider the entire set of values of the function f (x) to be represented by a
1.1 T HE INNER PRODUCT 12
vector | f 〉, where
f (x 1 )
¡ ¢
| f 〉 = f (x 2 ) , 〈f | = f ∗ (x 1 ) f ∗ (x 2 ) · · · . (1.41)
..
.
with
Zb Zb
∗
〈f |f 〉 = d x ω(x) f (x) f (x) = d x ω(x)| f (x)|2 . (1.42)
a a
| f + g 〉 = | f 〉 + |g 〉, c| f 〉 (1.43)
f (x 1 ) = f 0 + f 1 x 1 + f 2 x 12 + . . . ,
f (x 2 ) = f 0 + f 1 x 2 + f 2 x 22 + . . . ,
..
. (1.45)
X
∞
| f 〉 = f 0 |0〉 + f 1 |1〉 + f 2 |2〉 + · · · = f k |k〉. (1.47)
k=0
1.1 T HE INNER PRODUCT 13
Zb
〈e k |e l 〉 = d x ω(x)e k∗ (x)e l (x) = δkl , (1.49)
a
〈e k |e l 〉 = δkl , (1.51)
we can determine the coefficients v k in Eq. (1.50) by taking the inner product,
in this case, the scalar product, of |v〉 with each 〈e k | in succession, i.e.,
v k = 〈e k |v〉. (1.52)
X
∞
|f 〉 = f k |e k 〉, (1.53)
k=1
and the coefficients f k can be calculated by taking the inner product [see
Eq. (1.24)] of | f 〉 with each |e k 〉 in succession as
Zb
f k = 〈e k | f 〉 = d x ω(x) e k∗ (x) f (x). (1.54)
a
1.1 T HE INNER PRODUCT 14
These numbers are called the Fourier coefficients of | f 〉 with respect to the
basis {|e k 〉}∞
k=1
and they can be thought of as values of a function f : N → C,
where N is the (infinite) set of natural numbers. The vectors |e k 〉 are com-
plete, i.e., the only vector in L ω2 (a, b) that is orthogonal to all |e k 〉 is the zero
vector, for which all the Fourier coefficients f 1 = f 2 = · · · = 0. For instance,
using Eq. (1.53)
X
∞ X
∞ X
∞ X
∞
〈f |f 〉 = f k∗ f l 〈e k |e l 〉 = f k∗ f l δkl
k=1 l =1 k=1 l =1
X∞
= | f k |2 , (1.55)
k=1
X
∞ Zb
2
〈f |f 〉 = | fk | = d x ω(x)| f (x)|2 , (1.56)
k=1 a
P
∞
and since the integral is finite for f (x) ∈ L ω2 (a, b), we get that | f k |2 con-
k=1
verges. Even more, think now in a vector |v〉 ∈ R3 : if we consider a coordinate
frame and draw a point there which is related to |v〉, the norm ||v|| represents
the distance from the point drawn to the origin of the coordinate frame, mea-
sured following an imaginary line connecting the considered point with the
origin of the frame. So if we could measure such distance, let’s say with a
ruler, we have the value for ||v||. If we consider now an orthonormal basis in
R3 and write |v〉 in that basis, i.e., |v〉 = v 1 |i 〉 + v 2 | j 〉 + v 3 |k〉, we can calculate
||v|| as ||v||2 = |v 1 |2 + |v 2 |2 + |v 3 |2 , and, of course, ||v|| will coincide with the
above mentioned distance. But this happens only because we have consid-
ered all vectors which form a basis in R3 , in this case, three, i.e., our basis is
complete. If we would have forgotten to incorporate in the basis one, or more,
of the vectors |i 〉, | j 〉 or |k〉, we, of course, do not have a basis (our basis is not
complete), which means that the value obtained for ||v||2 from measuring the
distance from the corresponding point to the origin of our coordinate frame
P
k<n P
k<n
and the result obtained from |v k |2 would differ, with ||v||2 > |v k |2 . Only
k=1 k=1
if our basis is complete, i.e., if we have all the vectors needed to form a basis of
R3 , we will get that the value obtained for ||v||2 from measuring the distance
P
3
between the point and the origin and that from |v k |2 would coincide. The
k=1
notion of completeness normally is not emphasized when discussing an n-
dimensional vector space as Rn since if you take away some of the vectors of
1.1 T HE INNER PRODUCT 15
the basis, as we stated, you do not have a basis, because you have less than
n vectors. The situation is different in infinite dimension, as it is the case of
L ω2 (a, b). If you start with a basis and take away some of the vectors, you
still have an infinite number of orthonormal vectors. Equation (1.56), analo-
gously to our discussion of |v〉 ∈ R3 , is telling us that if we calculate 〈 f | f 〉 using
Rb P
∞
the integral d x ω(x)| f (x)|2 and the summation | f k |2 and both results do
a k=1
not match, we have forgotten some vector in our basis! The notion of com-
pleteness ensures that no orthonormal vector is taken out of the basis and the
vectors |e k 〉 satisfy the completeness relation
X
∞
|e k 〉〈e k | = 1, (1.57)
k=1
with 1 the identity matrix. Indeed, note that, from Eqs. (1.53) and (1.54)
X
∞ X
∞ X
∞ X
∞
|f 〉 = f k |e k 〉 = 〈e k | f 〉|e k 〉 = |e k 〉〈e k | f 〉 =⇒ |e k 〉〈e k | = 1, (1.58)
k=1 k=1 k=1 k=1
〈 f − g | f − g 〉 = 〈 f | f 〉 − 〈 f |g 〉 − 〈g | f 〉 + 〈g |g 〉
X∞ X
∞ X
∞ X
∞
= | f k |2 − f k∗ g k − f k g k∗ + |g k |2
k=1 k=1 k=1 k=1
X∞ X
∞
= ( f k − g k )∗ ( f k − g k ) = | f k − g k |2 . (1.59)
k=1 k=1
Zb
d x ω(x)| f (x)|2 (1.60)
a
1.2 T HE F OURIER SERIES 16
with
Zb
fk = d x ω(x) e k∗ (x) f (x). (1.62)
a
with m = 1, 2, . . . , which also have the orthonormality property and which are
represented by the trigonometric functions
1 1 1 ³ π ´
e 0+ (x) = e 0 (x) = p , e m +
(x) = p [e m (x) + e −m (x)] = p cos m x ,
2L 2 L L
i 1 ³ π ´
−
em (x) = − p [e m (x) − e −m (x)] = p sin m x . (1.70)
2 L L
In this way,
X
∞ X
∞
f (x) = f m e m (x) = f 0 e 0 (x) + [ f −m e −m (x) + f m e m (x)]. (1.71)
m=−∞ m=1
ZL
1
f 0+ = 〈e 0+ | f 〉= p d x f (x),
2L
−L
ZL ³ π ´
1
f m+ = 〈e m
+
|f 〉 = p d x cos m x f (x), (1.76)
L L
−L
ZL ³ π ´
1
f m− −
= 〈e m |f 〉= p d x sin m x f (x). (1.77)
L L
−L
1.2 T HE F OURIER SERIES 18
a0 X∞ h ³ π ´ ³ π ´i
f (x) = + a m cos m x + b m sin m x , (1.78)
2 m=1 L L
where
ZL ³ π ´
1
am = d x cos m x f (x), m ≥ 0, (1.79)
L L
−L
and
ZL ³ π ´
1
bm = d x sin m x f (x), m > 0. (1.80)
L L
−L
the Fourier series can be simplified. It is easy to verify than if f (x) is even all
b m = 0, while if f (x) is odd, all a m = 0.
It is important to note that since f (x) is a periodic function with period
2L, the Fourier series expansion extends the domain of definition of f (x) to
all the intervals 2kL − L ≤ x ≤ 2kL + L, k = 0, 1, 2, . . . , since f (x) with −L ≤ x ≤ L
is equivalent to f (x − 2kL) with 2kL − L ≤ x ≤ 2kL + L; both will give the same
Fourier series expansion. For this reason we just needed the periodic function
f (x) in [−L, L].
Periodic functions are not always defined on [−L, L]. Let us consider a
periodic function f (u) that is defined on [a, b] with period T = b − a. The
transformation
2L ³ T´
x≡ u−a− , (1.82)
T 2
brings the interval [−L, L] into [a, b], therefore, using Eq. (2.2), we have the
Fourier expansion
1 X
∞
f (u) = p f m e i m(π/L)(2L/T )(u−a−T /2)
2L m=−∞
1 X
∞
=p f m e i m(2π/T )u e −i m(2π/T )(a+T /2) , (1.83)
2L m=−∞
1.2 T HE F OURIER SERIES 19
with
Zb
1 2L
fm = p d u e −i m(π/L)(2L/T )(u−a−T /2) f (u)
2L T
a
p Zb
2L
= d u e −i m(2π/T )u e i m(2π/T )(a+T /2) f (u). (1.84)
T
a
where
Zb
1
Fm = p d u e −i m(2π/T )u f (u). (1.86)
T
a
The functions p1 e i m(2π/T )u are an orthonormal base of L 2 (a, b), thus, Eq. (1.85)
T
is the corresponding Fourier expansion of the function. We can redefine the
variable u as x, and write the Fourier expansion for a function f (x) in the in-
terval [a, b] with period T = b − a as
1 X
∞
f (x) = p f m e i m(2π/T )x , (1.87)
T m=−∞
where now,
Zb
1
fm = p d x e −i m(2π/T )x f (x). (1.88)
T
a
Zb
d x | f (x)| (1.89)
a
Thus, Eq. (1.88) makes also sense for a piecewise continuous function f ∈
L (a, b) and one can also perform a Fourier expansion in this case too.
Equation (1.88) can also be written in terms of sine and cosine functions.
By analogy to Eq. (1.78),
a0 X∞ h ³ 2π ´ ³ 2π ´i
f (x) = + a m cos m x + b m sin m x , (1.91)
2 m=1 T T
where now
Zb ³ 2π ´
2
am = d x cos m x f (x), m ≥ 0, (1.92)
T T
a
and
Zb ³ 2π ´
2
bm = d x sin m x f (x), m > 0. (1.93)
T T
a
V(t)
where
Z2τ
1
Vm = p d t V (t )e −i 2πmt /(2τ) . (1.98)
2τ
0
1.2 T HE F OURIER SERIES 22
E XAMPLE 1.4. Another frequently used voltage is the sawtooth voltage (see
Fig. 1.5). The equation for V (t ), with 0 ≤ t < τ, is
V (t ) = V0 t /τ. (1.106)
In this case, using Eqs. (1.87) and (1.88), we have that the corresponding
Fourier expansion is given by
1 X ∞
V (t ) = p Vm e i 2πmt /τ , (1.107)
τ m=−∞
with
Zτ
1
Vm = p d t V (t )e −i 2πmt /τ . (1.108)
τ
0
1.2 T HE F OURIER SERIES 24
V(t)
In this way,
µ ¶ p
−3/2 τ2 V0 τ
Vm = V0 τ =− , where m 6= 0. (1.110)
−i 2mπ i 2πm
In case of m = 0, from Eq. (1.108),
Zτ Zτ
1 1 t 1 p
V0 = p d t V (t ) = p d t V0 = V0 τ. (1.111)
τ τ τ 2
0 0
Thus, using Eqs. (1.110) and (1.111), we can write Eq. (1.107) as
· p µ −1 ¶¸
1 1 p V0 τ X 1 i 2πmt /τ X ∞ 1
i 2πmt /τ
V (t ) = p V0 τ − e + e
τ 2 i 2π m=−∞ m m=1 m
· p ∞ µ ¶¸
1 1 p V0 τ X 1 i 2πmt /τ −i 2πmt /τ
=p V0 τ − e −e
τ 2 i 2π m=1 m
· µ ¶¸
1 1 X∞ 1 2πmt
= V0 − sin . (1.112)
2 π m=1 m τ
Alternatively, if we could have used Eqs. (1.91), (1.92) and (1.93). In this
case, you can check that all the coefficients a m = 0 with m > 0 are zero. Using
Eq. (1.92) for the case m = 0, we have
Zτ Zτ
2 V0
a0 = d t V (t ) = 2 d t t = V0 . (1.113)
τ τ
0 0
Figure 1.6: Various approximations to the Fourier series in Eq. (1.112). The
dashed line corresponds to the first term of the series, the thick grey line is the
result keeping the first 3 terms, and the solid line represents the first 15 terms
of the expansion.
X
n
| fn 〉 = f k |e k 〉, (1.116)
k=1
X
∞
|f 〉 = f k |e k 〉 = lim | f n 〉. (1.117)
n→∞
k=1
Zb
En ≡ 〈 f − fn | f − fn 〉 = d x| f (x) − f n (x)|2 = 〈 f | f 〉 − 〈 f | f n 〉 − 〈 f n | f 〉 + 〈 f n | f n 〉.
a
(1.118)
Zb
〈f |f 〉 = d x | f |2 , (1.119)
a
lim E n = 0, (1.121)
n→∞
| f 〉 = lim | f n 〉, (1.122)
n→∞
which is read | f 〉 equals the limit in the mean of the sequence | f n 〉 as n ap-
proaches infinity. What we are going to show is that Eq. (1.121) is satisfied as
far as f k = 〈e k | f 〉.
Using Eqs. (1.53) and (1.116) in Eq. (1.118), we get
³X
n X
n ´
En = 〈 f | f 〉 − f k∗ 〈e k | f 〉 + f k 〈 f |e k 〉
k=1 k=1
X
n X
n
+ f k∗ f l 〈e k |e l 〉
k=1 l =1
n ³
X ´ Xn
= 〈f |f 〉− f k∗ 〈e k | f 〉 + f k 〈 f |e k 〉 + f k∗ f k , (1.123)
k=1 k=1
2
the average value of a function g (x) on an interval a ≤ x ≤ b is defined as g =
1
R
b
b−a d x g (x).
a
1.3 C ONVERGENCE OF THE EXPANSION 28
〈e k |e l 〉 = δkl . (1.124)
( f k − 〈e k | f 〉)∗ ( f k − 〈e k | f 〉) = f k∗ f k − ( f k∗ 〈e k | f 〉 + f k 〈e k | f 〉∗ ) + 〈e k | f 〉∗ 〈e k | f 〉,
(1.125)
i.e., when
f k = 〈e k | f 〉. (1.128)
Thus, choosing the expansion coefficient f k in Eq. (1.116) to be the finite in-
tegral transform of f (x) exactly minimizes the mean-square error in the ap-
proximation of f (x) by f n (x). We have then,
X
n
min(E n ) = 〈 f | f 〉 − f k∗ f k . (1.129)
k=1
Zb
En = d x | f (x) − f n (x)|2 , (1.130)
a
or
X
n
〈 f |e k 〉〈e k | f 〉 ≤ 〈 f | f 〉. (1.133)
k=1
X
n
| f k |2 (1.134)
k=1
Zb
0 < 〈f |f 〉 = d x | f (x)|2 < M , (1.135)
a
X
n
| f k |2 < M , (1.136)
k=1
X
n
lim | f k |2 (1.137)
n→∞
k=1
where
Zb
f k = 〈e k | f 〉 = d x e k∗ (x) f (x). (1.139)
a
Next, we shall derive a necessary and sufficient condition for the mean-
square error (1.118) to approach zero as n → ∞. Since both the error E n and
the minimum error min(E n ) are non-negative, we have
0 ≤ min(E n ) ≤ E n . (1.141)
Therefore,
lim E n = 0 (1.142)
n→∞
only if
X
∞ Zb
| f k |2 = 〈 f | f 〉 = d x | f (x)|2 . (1.144)
k=1 a
X
∞
|f 〉 = 〈e k | f 〉|e k 〉. (1.146)
k=1
lim 〈 f n − f | f n − f 〉 = 0 (1.147)
n→∞
lim 〈 f n − f m | f n − f m 〉 = 0. (1.148)
n,m→∞
vector in L 2 (a, b), i.e., there exists a vector | f 〉 ∈ L 2 (a, b) such | f 〉 can be
interpreted as the limit when n → ∞ of the sequence {| f n 〉}. Vector spaces
for which Eq. (1.147) is true are called complete. Thus, L 2 (a, b) is an inner
product space which is complete, and inner product spaces which are com-
plete are also known as Hilbert spaces. Hilbert space is the natural way to let
the number of dimensions become infinite, and at the same time to keep the
geometry of ordinary Euclidean space. Physicists on the 1920’s realized that
Hilbert space was the correct setting to establish Quantum Mechanics!
Instead of considering the mean-square error as in Eq. (1.118), we can ex-
amine instead the actual error in approximating f (x) by
X
n
f n (x) = f k e k (x), (1.149)
k=1
Notice that E n (x) depends upon the particular value of x in the interval a ≤
x ≤ b. If we could show that
for each a ≤ x ≤ b, then we could say that f n (x) converges point-wise to f (x)
in the interval [a, b]. The exact mathematical conditions under which an ex-
pansion in an arbitrary complete set of orthogonal functions is pointwise-
convergent lie outside the scope of our discussion.
We have then established that the orthogonal expansion of a square in-
tegrable piecewise continuous function converges in the mean-square sense.
Nevertheless, a piecewise continuous function f (x) can have a finite number
of finite discontinuities in the interval a ≤ x ≤ b. Near such finite discontinu-
ities the approximating function
X
n
f n (x) = f k e k (x) (1.152)
k=1
fails to match the jump in f (x). However, the series itself does not produce
a discontinuous function, as we saw in Example 1.3. In fact, it is possible to
proof that at a point of finite discontinuity x d , The Fourier series converges to
1
lim[ f (x d + ϵ) + f (x d − ϵ)]. (1.153)
2 ϵ→0
1.3 C ONVERGENCE OF THE EXPANSION 32
This behavior (see Fig. 1.7), which is called Gibbs’s phenomenon, can be qual-
itatively understood on the following basis: At the point x d , the slope of f (x),
i.e., d f /d x, becomes infinite. However f n (x) consists of the sum of a finite
number of smoothly varying functions forming the first n terms of a conver-
gent series. Therefore, d f n /d x must be a smoothly varying bounded function
for a ≤ x ≤ b. Hence, d f n /d x cannot match the infinite slope d f /d x of f (x) at
a point of discontinuity x = x d in [a, b]. The finite series f n (x) tries to achieve
the infinite slope of f (x) at x = x d and thereby overshoots the discontinuity
by a certain amount. As more terms of the series are included, the overshoot
δ moves in position arbitrarily close to the discontinuity, producing spikes of
zero thickness, but it never disappears even in the limit of an infinite num-
ber of terms. Since these additional spikes have zero thickness, they do not
effect the mean-square convergence of the infinite series for f (x), but they do
indicate the limitations of the process of representing f (x) by an orthogonal
expansion. The amount by which lim f n (x) overshoots f (x) at the discon-
n→∞
tinuity x = x d depends on the precise forms of both f (x) and the functions
e k (x). In general it is of the order of 9 percent of the jump in f (x) at x = x d .
A final comment is here in order: if we consider a periodic function f ∈
L (a, b), i.e., a function for which
Zb Zb
∗ 1/2
〈f |f 〉 = d x [ f (x) f (x)] = d x | f (x)| (1.154)
a a
exists and is finite, and calculate the Fourier coefficients for such a function,
we do not necessary have that
X
∞
| f k |2 < ∞, (1.155)
k=−∞
since that happens only if f ∈ L 2 (a, b). However, there is an important Lemma,
which is called the Riemann-Lebesgue lemma, for f ∈ L (a, b), which says
that the coefficients
Zb
1
fm = p d x e −i m(2π/T )x f (x), (1.156)
T
a
of any function f ∈ L (a, b) tends to zero as |m| → ∞. Thus the partial sums
1 X n
f n (x) = p f m e i m(2π/T )x (1.157)
T m=−n
1.3 C ONVERGENCE OF THE EXPANSION 33
Zb µ ¶1/2
∗
lim 〈 f n − f | f n − f 〉 = lim d x [ f n (x) − f (x)] [ f n (x) − f (x)] = 0. (1.158)
n→∞ n→∞
a
1 X
∞
f (x) = p f m e i m(2π/T )x , (1.159)
T m=−∞
1.4 F OURIER SERIES AND NON - PERIODIC FUNCTIONS 34
with
Zb
1
fm = p d x e −i m(2π/T )x f (x), (1.160)
T
a
What happens if the function f (x) is non-periodic in the fixed range given?
In such a case, we may propose another function, let us called it g (x), which
continues the original one outside the range so as to make it periodic. The
Fourier series of this periodic function g (x) would then correctly represent the
non-periodic function f (x) in the desired range. Since we are often at liberty
to extend the function f (x) in a number of ways, we can sometimes make
g (x) odd or even in a symmetric interval about the origin and, then, reduce
the calculation of the Fourier coefficients. In view of the Gibbs’s phenomenon
explained earlier, the choices for g (x) may be reduced, since g (x) must not
be discontinuous at the end-points of the interval of interest, otherwise the
Fourier series will not converge to the required value there.
E XAMPLE 1.5. Let us consider, for example, the function f (x) = x 2 , with 0 ≤
x ≤ 2, which is clearly a non-periodic function. To determine a Fourier ex-
pansion which can be related to f (x) we must first make the function peri-
odic. We do this by extending the range of interest to −2 ≤ x ≤ 2 in such a way
that the new function g (x) is an even function, i.e., g (x) = g (−x) and letting
g (x + T ) = g (x), where T is the period (in this case, T = 4) (see Fig. 1.8(top)).
In this way, all the coefficients b m will be zero. We could also extend the range
so as to make the function g (x) odd, i.e., g (x) = −g (−x) and then make g (x)
periodic in such a way that g (x + T ) = g (x) (see Fig. 1.8(bottom)). In this case
all a m = 0, m ≥ 0, will be zero. Note, however, that within the latter choice,
due to the Gibbs’s phenomenon, the Fourier expansion of g (x) will converge
to zero at x = ±2, while the original function f (x = ±2) = 4. The even exten-
sion is then better, because the Fourier expansion will converge to the original
values of f (x) at x = ±2. Let us consider g (x) to be the even extension of f (x).
Since the interval considered is symmetric about the origin and g (x) is even,
1.4 F OURIER SERIES AND NON - PERIODIC FUNCTIONS 35
g(x)
4
x
-6 -4 -2 2 4 6
g(x)
4
x
-6 -4 -2 2 4 6
-2
-4
Figure 1.8: Periodic extensions of the function f (x) = x 2 . (Top) Even extension
of f (x) plotted in the range −6 ≤ x ≤ 6. (Bottom) Odd extension of f (x) plotted
in the range in the range −6 ≤ x ≤ 6.
1.4 F OURIER SERIES AND NON - PERIODIC FUNCTIONS 36
Z2 µ ¶ Z2 µ ¶
2 2 2πmx 4 2 2πmx
am = d x x cos = d x x cos , (1.161)
4 4 4 4
−2 0
where in the last step we have made used of the fact that g (x) is even in x.
Thus, integrating by parts twice, we get
· µ ¶¯ Z2 µ ¶¸
2 2 πmx ¯¯2 4 πmx
am = x sin − d x xsin
πm 2 ¯0 πm 2
0
· µ ¶¯ Z2 µ ¶
8 πmx ¯¯2 8 πmx
= 2 2 xcos − d x cos
π m 2 ¯0 π 2 m 2 2
0
16 16
= 2 2 cos(πm) = 2 2 (−1)m , m > 0. (1.162)
π m π m
In case of m = 0, from Eq. (1.92)
Z2 Z2 Z2
2 4 8
a0 = d x g (x) = d xg (x) = x 2d x = . (1.163)
4 4 3
−2 0 0
In this way, using Eq. (1.91), we can write the Fourier expansion of g (x) in the
range −2 ≤ x ≤ 2 as
µ ¶
4 16 X ∞ (−1)m πmx
g (x) = + 2 cos . (1.164)
3 π m=1 m 2 2
Zx X∞ h ³ 2π ´¯x
a0 am ¯
g (x) = d y f (y) = C + x+ sin m y ¯
2 m=1 (2πm/T ) T 0
0
³ 2π ´¯x i
bm ¯
− cos m y ¯ , (1.167)
(2πm/T ) T 0
Zx X
∞ bm a0
g (x) = d y f (y) = C + + x
m=1 (2πm/T ) 2
0
X∞ h bm ³ 2π ´ am ³ 2π ´i
+ − cos m x + sin m x (1.168)
m=1 (2πm/T ) T (2πm/T ) T
P
∞
bm
Since m(2π/T )
produces a finite number, we can reabsorbed this term in
m=1
the constant of integration C , which we need to determine, i.e.,
X
∞ bm
C+ →C (1.169)
m=1 (2πm/T )
1.5 I NTEGRATION AND DIFFERENTIATION OF THE SERIES 38
and write
Zx X∞ h ³ 2π ´
a0 bm
g (x) = d y f (y) = C + x+ − cos m x
2 m=1 (2πm/T ) T
0
am ³ 2π ´i
+ sin m x . (1.170)
(2πm/T ) T
The right hand side of Eq. (1.170) is not, strictly speaking, a Fourier series,
since we do not have a periodic function due to the term a 0 x/2. There are two
ways to interpret this formula within the Fourier framework. Either we can
write
a0 X∞ h bm ³ 2π ´
g (x) − x =C + − cos m x
2 m=1 (2πm/T ) T
am ³ 2π ´i
+ sin m x (1.171)
(2πm/T ) T
and interpret the right hand side as the Fourier series of the function on the
left hand side, or, alternatively, we could replace the function x by its Fourier
series. Then, we can Rconsider the right hand side of Eq. (1.170) as the Fourier
x
series of the integral 0 d y f (y).
Considering the properties of the sine and cosine functions, it is easy to
show that
Zb ³ 2π ´ 1 Zb ³ 2π ´
1
d x cos m x = d x sin m x = 0. (1.172)
T T T T
a a
Zb ³
1 a0 ´
C= d x g (x) − x . (1.173)
T 2
a
Zb Zb
a0 1 1
= d x f (x) = d x f (x) = f (1.174)
2 T b−a
a a
is the mean or average of the function f (x) on the interval [a, b]. If the func-
tion f (x)Rhas zero mean, i.e., a 0 = 0, then g (x) can be considered as the Fourier
x
series of 0 d y f (y).
1.5 I NTEGRATION AND DIFFERENTIATION OF THE SERIES 39
Zb Zb
a0 1 1
= d x f (x) = d x f (x) = f (1.175)
2 T b−a
a a
is the mean or average of the function f (x) on the interval [a, b]. Then, a func-
tion has no a 0 term in its Fourier series if and only if it has zero mean. It can
be easily shown that the mean zero functions are precisely the ones which
remain periodic upon integration.
In any case, we can see from Eq. (1.171) that the effect of integration if to
place and additional power of m in the denominator of each coefficient. Thus,
we have a faster convergence than before.
Term-by-term differentiation, however, is a much more precarious mat-
ter. Differentiation of the series produces m factors in the numerator, thus,
we can destroy the convergence of the series. Therefore, to justify taking the
derivative of a Fourier series, we need to know that the differentiated function
remains reasonable nice. If f (x) is a continuous function of x for all x (thus,
f (x) has not finite jumps) and f (x) is also periodic, then the Fourier series
that results from differentiating term by term converges to d f /d x, provided
that d f /d x itself is piecewise continuous.
These properties of the Fourier series may be useful in calculating com-
plicated Fourier series, since simple Fourier series may easily be evaluated (or
found from standard tables) and often the more complicated series can then
be build up by integration and/or differentiation.
E XAMPLE 1.6. Let’s determine the Fourier series of x 3 for 0 < x ≤ 2 by using
the result from example 1.5, in which we determined the Fourier series for x 2
in the range 0 < x ≤ 2 by extending the function x 2 to make it periodic in the
range −2 ≤ x ≤ 2. If
Zx Zx
x3
g (x) = d y f (y) = d y y2 = , (1.176)
3
0 0
we can get the Fourier series for g (x) in the range 0 < x ≤ 2 by using Eqs. (1.171),
(1.162), (1.163). In this way, we can write
µ ¶
4 X
∞ 1 16 m 2πm
g (x) − x = C + (−1) sin
m=1 2πm/4 π m
3 2 2 4
µ ¶
32 X
∞ (−1) m
πm
=C + 3 sin . (1.177)
π m=1 m 3 2
1.5 I NTEGRATION AND DIFFERENTIATION OF THE SERIES 40
Be careful with differentiating term by term! We could think in using Eq. (1.181),
and differentiate it term by term to get d f /d x with f (x) = x. We find
µ ¶ µ ¶
X
∞
m πmx X
∞
m+1 πmx
1 = −2 (−1) cos =2 (−1) cos , (1.183)
m=1 2 m=1 2
which makes no sense, since the left-hand side of the equation is a constant,
while the right-hand side depends on x! Why is this happening? This is be-
cause Eq. (1.181) is obtained from the periodic extension of x 2 , thus, the se-
ries in Eq. (1.181) does not converge to x, but rather to its periodic exten-
sion. In fact, since there are only sine term in Eq. (1.181), the Fourier series
of Eq. (1.181) corresponds to the one obtained for an odd periodic extension
of x (see Fig. 1.9).
Such periodic extension of x has a jump discontinuity at odd multiples of
2, thus, the function is not continuous for all values of x and term by term dif-
ferentiation of the corresponding Fourier series does not converge to d (x)/d x =
1.
1.6 T HE F OURIER TRANSFORM 41
Next, we can extend the range of f (x) by introducing the function f Λ (x)
defined in the interval [a −Λ/2, b +Λ/2], where Λ is an arbitrary positive num-
1.6 T HE F OURIER TRANSFORM 42
In this way, we have managed to separate various copies of the original func-
tion by Λ (see Fig. 1.11). It should be clear that if Λ → ∞, we can completely
isolate the function and stop the repetition. In other words,
We can now obtain the Fourier expansion of f Λ (x). Using Eqs. (1.87) and
(1.88), we have
1 X
∞
f Λ (x) = p f mΛ e i 2πmx/(L+Λ) , (1.186)
L + Λ m=−∞
where
Z
b+Λ/2
1
f mΛ =p f Λ (x)e −i 2πmx/(L+Λ) . (1.187)
L +Λ
a−Λ/2
1 X∞
f Λ (x) = p f˜Λ (k)e i kx ∆k, (1.189)
2π k=−∞
with
Z
b+Λ/2
˜Λ 1
f (k) = p d x f Λ (x)e −i kx . (1.190)
2π
a−Λ/2
1 1 1 1 2π 1 1
p p ∆k → p p =p p , (1.191)
2π 2π 2π 2π L + Λ L +Λ L +Λ
which is the same factor that we obtain when substituting Eq. (1.187) into
(1.186).
If we consider now the limit Λ → ∞, ∆k becomes vanishingly small, i.e.,
∆k → d k, and k becomes a continuum. In other words, as m changes by one
unit, k changes only slightly. Thus, the infinite sum of terms in the Fourier
series of Eq. (1.189) becomes an integral in the limit Λ → ∞. We would then
have
Z∞
1
f (x) = p d k f˜(k)e i kx , (1.192)
2π
−∞
with
Z∞
1
f˜(k) = p d x f (x)e −i kx . (1.193)
2π
−∞
Equations (1.194) and (1.195) are called the Fourier integral transforms of
f˜(k) and f (x), respectively. The function f˜(k) is called the Fourier transform
while f (x) is the inverse Fourier transform of f˜(k). Note that the
of f (x), p
factor 1/ 2π appearing in Eqs. (1.194) and (1.195) is clearly arbitrary, with the
only requirement being that their product should be equal to 1/(2π). It would
have been possible, for example, to define
Z∞
f (x) = d k f˜(k)e i kx , (1.194)
−∞
1.6 T HE F OURIER TRANSFORM 44
with
Z∞
1
f˜(k) = d x f (x)e −i kx . (1.195)
2π
−∞
and you might find different conventions in different books. We state with the
convention of Eqs. (1.194) and (1.195), which is more symmetric.
In general, the Fourier transform f˜(k) of a function f (x) is a complex-
valued function of k. Thus, we can write f˜(k) in polar form as
Z∞
1
f (x) = p d kr (k)e i [kx−ϕ(k)] , (1.198)
2π
−∞
Z∞
0≤ d x| f (x)| ≤ M < ∞, (1.199)
−∞
then
Z∞
f˜(k) = d x f (x)e −i kx (1.200)
−∞
What happens if the function f (x) is an even or odd function? Can we get
a simplification of the Fourier transform as in case of the Fourier series? Note
that Eq. (1.195) can be written as
Z∞
1
f˜(k) = p d x f (x)[cos(kx) − i sin(kx)]
2π
−∞
= f˜c (k) − i f˜s (k), (1.204)
where
Z∞
1
f˜c (k) = p d x f (x)cos(kx),
2π
−∞
Z∞
1
f˜s (k) = p d x f (x)sin(kx), (1.205)
2π
−∞
are called the Fourier cosine and sine transform, respectively, of a function
f (x).
If the function f (x) is an even function of x, i.e., f (x) = f (−x),
Z∞
1
p d x f (x)sin(kx) = 0, (1.206)
2π
−∞
since the integral of an odd function over a symmetric interval about the ori-
gin is zero. Then
Z∞ r Z∞
1 2
f˜(k) = p d x f (x)cos(kx) = d x f (x)cos(kx), (1.207)
2π π
−∞ 0
1.6 T HE F OURIER TRANSFORM 46
because the integral of an even function over a symmetric interval about the
origin is twice the integral taken over one-half the interval. This shows that if
f (x) is an even function of x, its Fourier transform is purely real. Similarly, if
f (x) is an odd function of x, i.e., f (x) = − f (−x),
Z∞ r Z∞
i 2
f˜(k) = − p d x f (x)sin(kx) = −i d x f (x)sin(kx), (1.208)
2π π
−∞ 0
because the integrand in the above equation is again even, i.e., it is the prod-
uct of two odd functions of x.
Given an arbitrary function f (x) we can either take the complex Fourier
transform of f (x) directly, or the sine and cosine Fourier transforms.
E XAMPLE 1.7. Let us evaluate the Fourier transform of the Gaussian function
2
f (x) = ae −bx , a,b > 0. Using Eq. (1.195),
Z∞ 2 Z∞
a −b(x 2 +i kx/b) ae −k /(4b) 2
f˜(k) = p dx e = p d x e −b(x+i k/2b) . (1.209)
2π 2π
−∞ −∞
Then, we have
a 2
f˜(k) = p e −k /(4b) , (1.211)
2b
which is also a Gaussian function.
f (x) = 〈e x | f 〉, (1.212)
e k (x) = 〈e x |e k 〉. (1.213)
1.6 T HE F OURIER TRANSFORM 47
We physicist prefer more the notation |x〉 and |k〉 for the vectors |e x 〉 and |e k 〉.
Then, we write
and
Note that the inner product of two functions f (x) and g (x) belonging to
L ω2 (a, b) can now be written as
Zb Zb
∗
〈g | f 〉 = d x ω(x) g (x) f (x) = d x ω(x) 〈g |x〉〈x| f 〉
a a
³ Zb ´
= 〈g | d x ω(x) |x〉〈x| | f 〉, (1.216)
a
Zb
d x ω(x)|x〉〈x| = 1. (1.217)
a
X Zb
→ d x ω(x). (1.218)
i
a
³ Zb ´ Zb
|f 〉 = d x ω(x)|x〉〈x| | f 〉 = d x ω(x) f (x)|x〉, (1.219)
a a
1.6 T HE F OURIER TRANSFORM 48
which shows how to expand a vector | f 〉 in terms of the |x〉’s. If we take now
the inner product of Eq. (1.219) with 〈x 0 |, we obtain
Zb
0 0
〈x | f 〉 = f (x ) = d x ω(x) f (x)〈x 0 |x〉, (1.220)
a
which is the Dirac delta function and remember that for a function f defined
on the interval [a, b] has the following property (if this is the first time you
have encountered with the Dirac delta function, you should take a look at the
Appendix 1.11).
Zb ½
0 f (x 0 ) if x 0 ∈ (a, b),
d x f (x)δ(x − x ) = (1.222)
0 otherwise
a
Let us now particularize all this for Eqs. (1.194) and (1.195). In our case,
ω(x) = 1 and the range of x is [−∞, ∞]. Using Eqs. (1.214) and (1.215), and
interpreting f˜(k) as the component with “index” k of | f˜〉, i.e.,
Z∞ µ Z∞ ¶
〈x| f 〉 = d k〈k| f˜〉〈x|k〉 = 〈x| d k|k〉〈k| | f˜〉, (1.224)
−∞ −∞
where
1
〈x|k〉 = p e i kx . (1.225)
2π
1.6 T HE F OURIER TRANSFORM 49
which is the same as Eq. (1.217) for our particular case, i.e., ω(x) = 1 and the
interval [−∞, ∞]. Then, Eq. (1.221) yields,
and use this equation when substituting Eq. (1.195) into (1.194), getting the
identity f (x) = f (x).
1.6 T HE F OURIER TRANSFORM 50
In this way, {|x〉}x∈R and {|k〉}k∈R form two bases of L 2 (−∞, ∞) and we
can express a vector | f 〉 ∈ L 2 (−∞, ∞) in terms of these two bases by using
the inner product of L 2 (−∞, ∞): 〈x| f 〉 corresponds to the components of | f 〉
in the basis {|x〉}x∈R , i.e., f (x), while 〈k| f 〉 represents the component of | f 〉
in the basis {|k〉}k∈R . How are these two bases connected? They are related
through Eq. (1.225). The Fourier transform, and its inverse, establishes the
way of obtaining f˜(k) = 〈k| f 〉 given f (x) = 〈x| f 〉 and vice versa.
Z∞ Z a µ ¶
1 −i kx b −i kx 2ab sin(ka)
f˜(k) = p d x f (x)e =p dx e =p . (1.232)
2π 2π −a 2π ka
−∞
Simple, right? Let us discuss this result in detail. First, note that if a → ∞, the
function f (x) becomes a constant function over the entire real line, and we
get from Eq. (1.232)
2b sin(ka) 2b
f˜(k) = p lim = p πδ(k) (1.233)
2π a→∞ k 2π
To get the last equation one has to note that Eq. (1.228) can be written as
Za
1 0 1 sin(a(k − k 0 ))
δ(k − k 0 ) = lim d x e i (k −k)x = lim . (1.234)
a→∞ 2π a→∞ π k − k0
−a
1.7 P ROPERTIES OF F OURIER TRANSFORMS 51
Next, let b → ∞ and a → 0 in such a way that 2ab, which is the area under
f (x), is 1. Then f (x) will approach the δ-function, and f˜(k) becomes
So, since f (x) approaches to the δ-function under the limit considered, we
p obtained that the Fourier transform of the δ-function is the constant
have
1/ 2π.
Finally, we note that the width of f (x) is ∆x = 2a, and the width of f˜(k) is
roughly the distance, on the k-axis, between its first two roots, k + and k − , on
either side of k = 0: ∆k = k + − k − = 2π/a. Thus, increasing the width of f (x)
results in a decrease in the width of f˜(k). In other words, when the function
is wide, its Fourier transform is narrow. In the limit of infinite width (a con-
stant function), we get infinite sharpness (the δ-function). The last two state-
ments are very general.p For instance, in Example 1.7 the width of f (x), which
is proportional top 1/ b, is the inverse relation to the width of f˜(k), which is
proportional to b. In fact, it can be shown that ∆x∆k ≥ 1 for any function
f (x). When both sides of this inequality are multiplied by the reduced Planck
constant ħ = h/(2π), the result is the celebrated Heisenberg uncertainty rela-
tion of quantum mechanics ∆x∆p ≥ ħ, where p = ħk is the momentum of the
particle. In this context, the width of the function, which corresponds to the
so-called wave packet, measures the uncertainty in the position x of a quan-
tum mechanical particle. Similarly, the width of the Fourier transform mea-
sures the uncertainty in k, which is related to the momentum p of the particle
via p = ħk.
1. Differentiation:
· ¸
dn f
F = (i k)n f˜(k). (1.236)
d xn
1.7 P ROPERTIES OF F OURIER TRANSFORMS 52
2. Integration:
·Z x ¸
1 ˜
F f (y)d y = f (k) + 2πC δ(k), (1.237)
ik
where the last term in the above equation corresponds to the Fourier
transform of the constant of integration C associated with the indefinite
integral on the left side of Eq. (1.237).
3. Scaling:
µ ¶
1 ˜ k
F [ f (ax)] = f . (1.238)
a a
4. Translation:
F [ f (x + a)] = e i ak f˜(k). (1.239)
5. Exponential multiplication:
· ¸
αx
F e f (x) = f˜(k + i α), (1.240)
as well. Let
Z∞
1
h̃(k) = p d x h(x)e −i kx ,
2π
−∞
Z∞
1
f˜(k) = p d x f (x)e −i kx ,
2π
−∞
Z∞
1
g̃ (k) = p d x g (x)e −i kx (1.243)
2π
−∞
1.7 P ROPERTIES OF F OURIER TRANSFORMS 53
be the Fourier transforms of h(x), f (x), and g (x), respectively. The con-
volution theorem says that if
Z∞
h(x) = d y f (x − y)g (y), (1.244)
−∞
then
p
h̃(k) = 2π f˜(k)g̃ (k). (1.245)
Z∞ Z∞ Z∞
1 −i kx 1 −i kx
h̃(k) = p d x h(x)e =p dx e d y f (x − y)g (y).
2π 2π
−∞ −∞ ∞
(1.246)
Z∞ Z∞ Z∞ Z∞
1 −i k(z+y) 1 −i kz
h̃(k) = p d z d ye f (z)g (y) = p d z f (z)e d y g (y)e −i k y
2π 2π
−∞ ∞ −∞ −∞
1 hp ˜ ihp i p
=p 2π f (k) 2πg̃ (k) = 2π f˜(k)g̃ (k). (1.247)
2π
1
F [ f (x)g (x)] = p f˜(k) ∗ g̃ (k). (1.248)
2π
7. Parseval’s theorem: If
Z∞
1
f˜(k) = p d x f (x)e −i kx , (1.249)
2π
−∞
1.8 F OURIER TRANSFORM IN MORE THAN ONE DIMENSION 54
and
Z∞
1
f (x) = p d k f˜(k)e i kx , (1.250)
2π
−∞
then
Z∞ Z∞
2
d x | f (x)| = d k | f˜(k)|2 . (1.251)
−∞ −∞
Z∞ Z∞ Z∞ · Z∞ ¸∗ · Z∞ ¸
∗ 1 1
2
d x | f (x)| = d x [ f (x)] f (x) = dx p d l f˜(l )e ilx
p d l f˜(k)e i kx
2π 2π
−∞ −∞ −∞ −∞ −∞
Z∞ · Z∞ ¸· Z∞ ¸
1 1
= dx p d l [ f˜(l )]∗ e −i l x p d l f˜(k)e i kx
2π 2π
−∞ −∞ −∞
Z∞ Z∞
1
= dl d k [ f˜(l )]∗ f˜(k)(2π)δ(k − l )
2π
−∞ −∞
Z∞
= d k| f˜(k)|2 , (1.252)
−∞
We use can easily generalized Eqs. (1.194) and (1.195) if more than one di-
mension is involved. By noticing that in three dimensions, kx corresponds to
the projection of r = xiˆ + y jˆ + zk̂ in the direction of k = k iˆ, we can write for
x = (x 1 , x 2 , . . . , x n ) and k = (k 1 , k 2 , . . . , k n )
Z
1
f (xx ) = d n k e ikk ·xx f˜(k
k ),
(2π)n/2
Z
1
f˜(k
k) = d n x e −ikk ·xx f (xx ). (1.253)
(2π)n/2
1.8 F OURIER TRANSFORM IN MORE THAN ONE DIMENSION 55
and
Z
0 1 0
δ(k
k −k k )= n
d n x e i (kk −kk )·xx ,
(2π)
Z
0 1 0
δ(xx − x ) = n
d n k e i (xx −xx )·kk , (1.254)
(2π)
1 1
〈xx |k
k〉 = n/2
e ikk ·xx , 〈k
k |xx 〉 = e −ikk ·xx . (1.255)
(2π) (2π)n/2
Equations (1.254) and (1.255) and the indentification | f˜〉 ≡ | f 〉 exhibit a strik-
ing resemblance between |xx 〉 and |k k 〉. In fact, any given abstract vector | f 〉
can be expressed either in terms of its x-representation, 〈xx | f 〉 = f (xx ), or in
terms of its k representation, 〈kk | f 〉 ≡ f˜(k
k ). These two representations are
completely equivalent, and there is one-to-one correspondence between the
two, which is given by Eq. (1.253). The representation that is used in practice
is dictated by the physical application. In quantum mechanics, for instance,
most of the time the x-representation, corresponding to the position, is used,
because then the operator equations turn into differential equations that are
(in many cases) linear and easier to solve than the corresponding equations
in the k-representation, which is related to the momentum.
E XAMPLE 1.9. In this example we are going to evaluate the Fourier transform
of the Yukawa potential
qe −αr
Vα (r ) = , α>0 (1.256)
r
Z∞ Z1 Z2π −αr
q 2 −i kr cosθ e
k) =
Ṽα (k d r r d cosθ d ϕe . (1.258)
(2π)3/2 r
0 −1 0
1.9 OTHER INTEGRAL TRANSFORMS : T HE L APLACE TRANSFORM 56
The ϕ integration simply gives a factor 2π, while in case of the θ integration
we get
Z1
1
d cosθe −i kr cosθ = (e i kr − e −i kr ). (1.259)
i kr
−1
Then,
Z∞ −αr
q(2π) 2e 1
k) =
Ṽα (k 3/2
d r r (e i kr − e −i kr )
(2π) r i kr
0
Z∞ · ¸
q 1 (−α+i k)r −(α+i k)r
= 1/2
dr e −e
(2π) i k
0
· ¯ ¯ ¸
q 1 e (−α+i k)r ¯¯∞ e (−α+i k)r ¯¯∞
= + . (1.260)
(2π)1/2 i k −α + i k ¯0 α + i k ¯0
2q 1
k) = p
Ṽα (k . (1.261)
2π k + α2
2
The parameter α is a measure of the range of the potential. It is clear that the
larger α is, the smaller the range. In fact, it was in response to the short range
of nucleon forces that Yukawa introduced α, which turns out to be related to
the mass of a pion.
The Fourier transform exists only for a function f (x) which satisfies the con-
dition
Z∞
0< d x | f (x)| ≤ M < ∞, (1.262)
−∞
with M being some positive real number. However, even for simple functions
like f (x) = e i kx , the Fourier transform fails to converge. What to do then with
such functions? Furthermore, we might be interested in a given function only
for x > 0 (for instance, consider x to represent the time variable t ). This leads
1.9 OTHER INTEGRAL TRANSFORMS : T HE L APLACE TRANSFORM 57
provided that the integral exists. We assume here that s is real and positive, but
complex values with Re(s) > 0 would have to be considered in a more detailed
study. Through Eq. (1.263) we define a linear transformation L that converts
functions of the variable x to functions of a new variable s:
L[a f 1 (x) + b f 2 (x)] = a L[ f 1 (x)] + b L[ f 2 (x)] = a f˜1 (s) + b f˜2 (s). (1.264)
A few comments on the existence of the integral are in order. The infinite
integral of f (x), i.e.,
Z∞
d x f (x), (1.265)
0
need not exist. For instance, f (x) may diverge exponentially for large x. How-
ever, if there are some constants s 0 , M and x 0 ≤ 0 such that for all x > x 0
the Laplace transform will exist for s > s 0 ; f (x) is then said to have exponential
2
growth of order s 0 . As a counterexample, f (x) = e x does not satisfy the con-
2
dition given by Eq. (1.266) and is not of exponential order. Thus, L[e x ] does
not exist.
The Laplace transform may also fail to exist because of a sufficiently strong
singularity in the function f (x) as x → 0. For example,
Zx
d x x n e −sx (1.267)
0
diverges at the origin for n ≤ −1. The Laplace transform L[x n ] does not exist.
Before continuing with further discussions, I guess you might wonder about
the origin of the definition of the Laplace transformation of a function, and
you might be surprised to know that such a definition can be considered sim-
ply as the continuous analog of the Taylor series of a function! Indeed, if we
have a function F (x) which admits a Taylor expansion, we can write
X
∞
F (x) = fn xn . (1.268)
n=0
1.9 OTHER INTEGRAL TRANSFORMS : T HE L APLACE TRANSFORM 58
Z∞
F (x) = d t f (t )x t . (1.270)
0
For the series in Eq. (1.269) to converge, we consider 0 < x < 1, since negative
values of x or x > 1 can mess up the convergence. This means that lnx < 0 and
we can introduce a change of variable,
−s = ln(x), (1.271)
or simply,
Z∞
F (s) = d t f (t )e −st . (1.274)
0
then
Z∞
1
f˜(s) = d xe −sx = , for s > 0. (1.276)
s
0
1.9 OTHER INTEGRAL TRANSFORMS : T HE L APLACE TRANSFORM 59
Next, let
Z∞ · (α−s)x ¸¯∞
e ¯
f˜(s) = αx −sx
dx e e = ¯ = 1 , for s > α. (1.278)
α − s ¯0 s −α
0
Using this relation, we can determine the Laplace transform of certain other
functions. For example, since
1 1
cosh(αx) = (e αx + e −αx ), sinh(αx) = (e αx − e −αx ) (1.279)
2 2
we have
µ ¶
1 1 1 s
L[cosh(αx)] = + = 2 ,
2 s −α s +α s − α2
µ ¶
1 1 1 α
L[sinh(αx)] = − = 2 , (1.280)
2 s −α s +α s − α2
it is evident that we can obtain transforms of the sine and cosine if α is re-
placed by i α in Eq. (1.280):
s α
L[cos(αx)] = , L[sin(αx)] = , (1.282)
s 2 + α2 s 2 + α2
both valid for s > 0. It is a curious fact that lim L[sin(αx)] = 1/α despite the
s→0
R∞
fact that d x sin(αx) does not exists.
0
Last case: f (x) = x n . We have then
Z
Γ(n + 1)
L[x n ] = d x x n e −sx = , s > 0, n > −1, (1.283)
s n+1
Table 1.1: Some common Laplace transforms. The transforms are valid for
s > s0 .
f (x) f˜(s) s0
c c/s 0
cx n cn!/s n+1 0
sin(αx) α/(s 2 + α2 ) 0
cos(αx) s/(s 2 + α2 ) 0
e αx 1/(s − α) α
x n e αx n!/(s − α)n+1 α
sinh(αx) α/(s 2 − α2 ) |α|
cosh(αx) s/(s 2 − α2 ) |α|
e αx sin(βx) β/[(s − α)2 + β2 ] α
e αx cos(βx) (s − α)/[(s − α)2 + β2 ] α
1
x 1/2 3 1/2
2 (π/s ) 0
x −1/2 (π/s)1/2 0
δ(x − x 0 ) e −sx0 0
½
1 for x ≥ x 0
θ(x − x 0 ) = e −sx0 /s 0
0 for x < x 0
Zs0
d x [ f 1 (x) − f 2 (x)] = 0. (1.284)
0
This result is known as Lerch’s theorem, and is not quite equivalent to f 1 (x) =
f 2 (x), because it permits f 1 (x) and f 2 (x) to differ at isolated points. How-
ever, in most problems studied in physics this ambiguity is not important and
for all practical purposes when finding inverse Laplace transforms using Ta-
ble 1.1, the inverse Laplace transform would be considered unique. The in-
verse Laplace transformation is linear, thus
E XAMPLE 1.11. Using Table 1.1, let us calculate the function f (x) whose Laplace
transform is given by
α2
f˜(s) = . (1.286)
s(s 2 + α2 )
First, this function is not listed in Table 1.1. However, we can rewrite Eq. (1.286)
as
1 s
f˜(s) = − 2 . (1.287)
s s + α2
In this way,
· ¸ · ¸
−1
˜ −1 1 −1 s
f (x) = L [ f (s)] = L −L . (1.288)
s s 2 + α2
E XAMPLE 1.12. In this example we are going to use the Laplace transform and
Table 1.1, to evaluate a definite integral. In particular, we are going to calculate
Z∞
sin(y x)
f (x) = d y . (1.290)
y
0
1.9 OTHER INTEGRAL TRANSFORMS : T HE L APLACE TRANSFORM 62
Using Table 1.1, the factor in square brackets is just the Laplace transform of
sin(x y). Then, we have
Z∞ µ ¶¯∞
1 1 ¯
−1 y ¯ π
f˜(s) = d y = tan ¯ = . (1.293)
s2 + y 2 s s y=0 2s
0
1. Differentiation:
· ¸
d f (x)
L = s L[ f (x)] − f (+0), s > 0. (1.298)
dx
This property can be proven starting from the definition of the Laplace
transform for d f /d x and integrating by parts. Naturally, both f (x) and
its derivative must be such that the integrals do not diverge. Since f (x)
and/or d f /d x can be piecewise continuous, strictly speaking, the zero
of x needs to be approached from the positive side of x. For this reason
we write f (+0) instead of f (0). An extension of Eq. (1.298) to higher
derivatives is also possible:
· n ¸ ¯
d f (x) n n−1 d n−1 f ¯¯
L = s L[ f (x)] − s f (0) − · · · − , s > 0. (1.299)
d xn d x n−1 ¯x=+0
2. Change of scale:
µ ¶
1 ˜ s
L[ f (ax)] = f . (1.300)
a a
3. Substitution:
· ¸
˜ ax
f (s − a) = L e f (x) . (1.301)
4. Translation:
where b > 0 and remember that we are considering that f (x) = 0 if x < 0,
thus, f (x − b) = 0 for 0 ≤ x < b.
5. Derivative of a transform:
· ¸
d n f (s) n
= L (−x) f (x) . (1.303)
d sn
In this case, e −sx f (x) needs to converge exponentially for large s, such
that all the integrals obtained on the right-hand side of the equation
when calculating the transform will be uniformly convergent due to the
decreasing exponential behavior of e −sx f (x).
1.10 P ROPERTIES OF THE L APLACE TRANSFORM 64
Z∞
f (x) = d ye −x y f (y) (1.304)
0
Z∞ · ¸
f (x)
d x f (x) = L , (1.305)
x
s
provided limx→0 f (x)/x exists. The lower limit s must be chosen large
enough so that f˜(s) is within the region of uniform convergence.
7. Convolution theorem: If the functions f (x) and g (x) have Laplace trans-
forms f˜(s) and g̃ (s) then
· Zx ¸
L[ f ∗ g ] = L d y f (y)g (x − y) = f˜(s)g̃ (s),
0
Zx
L−1 [ f˜(s)g̃ (s)] = d y f (y)g (x − y) = f ∗ g . (1.306)
0
d2
−α2 sin(αx) = [sin(αx)] (1.308)
d x2
we can determine the Laplace transform for sin(αx). Effectively, if we apply
the Laplace transform to Eq. (1.308) we get
· 2 µ ¶¸
2 d
−α L[sin(αx)] = L sin(αx) . (1.309)
d x2
1.11 A PPENDIX : D IRAC DELTA FUNCTION 65
Then,
α
L[sin(αx)] = , (1.311)
s 2 + α2
which confirms the result obtained in Eq. (1.282).
Figure 1.13: (a) The charged line segment and (b) its linear density function.
x 0 −L/2, q/L for values between x 0 −L/2 and x 0 +L/2, and zero again for values
greater than x 0 + L/2. Let us call this function λ(x). Then, we can write,
0, if x < x 0 − L/2,
λ(x) = q/L, if x 0 − L/2 < x < x 0 + L/2, (1.312)
0, if x > x 0 + L/2.
Now suppose that we squeeze the segment on both sides so that the length
shrinks to L/2 without changing the position of the midpoint and the amount
of charge. The new function describing the linear charge density will now be
0, if x < x 0 − L/4,
λ(x, x 0 ) = q 2/L, if x 0 − L/4 < x < x 0 + L/4, (1.313)
0, if x > x 0 + L/4.
The charge q has been factorized for later convenience. We have also intro-
duced a second argument to emphasize the dependence of the function λ on
the midpoint. Instead of one-half, we can shrink the segment to any fraction,
while still keeping both the amount of charge and the midpoint unchanged.
Shrinking the size L/n and renaming the function as λn (x, x 0 ) to reflect its de-
pendence on n, we have
0, if x < x 0 − L/(2n),
λn (x, x 0 ) = q n/L, if x 0 − L/(2n) < x < x 0 + L/(2n), (1.314)
0, if x > x 0 + L/(2n).
This function is depicted in Fig. 1.14 for n = 10 as well as for some smaller
values of n. As you can see, the height of λn (x, x 0 ) increases at the same time
that its width decreases.
Instead of a charge distribution that abruptly changes from zero to some
finite value and just abruptly drops to zero, let us consider a charge distribu-
tion that smoothly rises to a maximum value and just as smoothly falls to zero.
1.11 A PPENDIX : D IRAC DELTA FUNCTION 67
Figure 1.14: The linear density λn (x, x 0 ) in Eq. (1.314) versus x as n increases.
There are, of course, many functions which could be used to describe such a
charge distribution, but one which is convenient is a Gaussian distribution.
For example,
r
n −n(x−x0 )2
λn (x, x 0 ) = q e , (1.315)
π
p p
has a peak of height q n/π at x = x 0 (why the factor n/π, you are going to
understand it in the next paragraph) and drops to smaller and smaller values
as we get farther and farther away from x 0 in either direction, as shown in
Fig. 1.15. It is clear from the figure that the width of the graph of λn (x, x 0 ) gets
smaller as n → ∞.
Figure 1.15: λn (x, x 0 ) in Eq. (1.315) versus x for several values of n. The Gaus-
sian bell-shaped curve approaches the Dirac delta function as the width of the
curve approaches zero. The value of n is 1 for the dashed curve, 4 for the thick
solid line, and 20 for the thin solid line.
In both cases λn (x, x 0 ) is a true linear charge density in the sense that its
integral gives the total charge. This is evident in the first case, Eq. (1.314),
1.11 A PPENDIX : D IRAC DELTA FUNCTION 68
because of the way the function was defined. In the second case, Eq. (1.315),
once we integrate from −∞ to +∞, we also obtain the total charge q. The
region of integration extends over all real numbers in the second case because
at every point of the real line we have some nonzero charge. Furthermore,
we can extend the interval of integration over all real numbers even for the
first case, because the function vanishes outside the interval [x 0 − L/(2n), x 0 +
L/(2n)] and no other contribution to the integral arises. We thus can write
Z∞
d xλn (x, x 0 ) = q (1.316)
−∞
In other words, the integral of δn (x, x 0 ) over all the real numbers is one, and, in
particular, independent of n. Using δn (x, x 0 ) we define the Dirac delta func-
tion, δ(x, x 0 ) as
and, since the integral in Eq. (1.320) is independent of n, has the following
property:
Z∞
d xδ(x, x 0 ) = 1. (1.322)
−∞
1.11 A PPENDIX : D IRAC DELTA FUNCTION 69
The Dirac delta function has infinite height and zero width at x 0 , but these two
undefined quantities compensate for one another to give a finite area under
the graph of the function. The Dirac delta function is actually not a function
in the usual sense, because at the only point that it is nonzero, it is infinite!
Although we have separated the arguments of the Dirac delta function by
a comma, the function depends only on the difference between the two argu-
ments. This becomes clear if we think of the Dirac delta function as the limit
when n → ∞ of the exponential in Eq. (1.315), because the latter is a function
of x − x 0 . We therefore have the important relation
We can think of the last equality as an identity satisfied by the Dirac delta func-
tion: the Dirac delta function is zero everywhere except at the point which
makes its argument zero, in which case the Dirac delta function is infinite.
Since the Dirac delta function is zero almost everywhere, we can shrink the
region of integration in Eq. (1.322) to a smaller interval and write
Zb
d xδ(x − x 0 ) = 1 (1.325)
a
as long as x 0 lies in the interval [a, b]. If x 0 is outside the interval, then the
integral will be zero because the delta function would always be zero in the
region of integration. We can then summarize these results in the following
way:
Zb ½
1, if a < x 0 < b,
d xδ(x − x 0 ) = (1.326)
0, otherwise.
a
Equation (1.322) is then a special case of this for which −∞ < x 0 < ∞ for any
value of x 0 . Any function, as the Dirac delta function, whose integral over all
real numbers is one is called a linear density function. What kind of distri-
bution the Dirac delta function describes? For example, consider mδ(x − x 0 ),
where m designates mass. This function is zero everywhere except at x 0 and
its integral is the total mass m. Thus, it it is to be a mass distribution, it has
to be a point mass located at x 0 . The linear density of a point mass is infinite
because its length is zero, and this is precisely what mδ(x − x 0 ) describes.
1.11 A PPENDIX : D IRAC DELTA FUNCTION 70
Z∞ xZ0 +ϵ
' f (x 0 ) d xδn (x − x 0 )
x 0 −ϵ
Z+∞
' f (x 0 ) d xδn (x − x 0 ) ' f (x 0 ). (1.327)
−∞
The approximation in the second line follows from the fact that f (x) is almost
constant in the small interval [x 0 −ϵ, x 0 +ϵ]. The third approximation is a result
of the smallness of δn outside the interval, and the equality follows because δn
is a linear density function. In the limit n → ∞, δn becomes the Dirac delta
function and the approximation in Eq. (1.327) becomes an equality, i.e.,
Z∞
d x f (x)δ(x − x 0 ) = f (x 0 ). (1.328)
−∞
This is equivalent to
Zb ½
f (x 0 ), if a < x 0 < b,
d x f (x)δ(x − x 0 ) = (1.329)
0, otherwise.
a
1.11 A PPENDIX : D IRAC DELTA FUNCTION 71
In words, the result of integration is the value of f at the root (i.e., the zero)
of the argument of the delta function, provided this root is inside the range of
integration. In this way, the result of integration is always well defined then
because it is simply the value of a good function f at a point, say x 0 . In fact,
the result of integration is so nice that one can even define the derivative of the
Dirac delta function by differentiating Eq. (1.329) with respect to x 0 , obtaining
Z∞ ¯
d d f ¯¯
d x f (x) [δ(x − x 0 )] = − . (1.330)
dx d x ¯x=x0
−∞
Zb ½
(n) (−1)n f (n) (x 0 ), ifa < x 0 < b,
d x f (x)δ (x − x 0 ) = (1.331)
0, otherwise,
a
can be written as
This fact strongly suggests the identification of the derivative of the θ-function
as the Dirac delta function. Noting that
½
1, if x > x 0 ,
θ(x − x 0 ) = (1.337)
0, if x < x 0 ,
and that d [θ(x − x 0 )]/d x is zero everywhere except at x 0 , for any well-behaved
function f (x) we obtain
Z∞ xZ0 +ϵ xZ0 +ϵ
d [θ(x − x 0 )] d [θ(x − x 0 )] d [θ(x − x 0 )]
d x f (x) = d x f (x) ' f (x 0 )
dx dx dx
−∞ x 0 −ϵ x 0 −ϵ
¯x0 +ϵ
¯
= f (x 0 )θ(x − x 0 )¯¯ = f (x 0 )[θ(ϵ) − θ(−ϵ)]
x 0 −ϵ
= f (x 0 )[1 − 0] = f (x 0 ). (1.338)
d [θ(x − x 0 )]
δ(x − x 0 ) = . (1.339)
dx
All these discussions can be generalized to many variables. For example,
in two dimensions and Cartesian coordinates
n −n[(x−x0 )2 +(y−y 0 )2 ]
δn (rr −rr 0 ) ≡ δn (x − x 0 , y − y 0 ) = e
π
n −n(x−x0 )2 −n(y−y 0 )2
= e e = δn (x − x 0 )δn (y − y 0 ) (1.340)
π
and the integral of δn over the entire x y-plane equals to one. In this way,
which is zero everywhere except at the point which makes all two of its argu-
ment arguments zero, in which case it is infinite, i.e., (see Fig. 1.16)
Figure 1.16: As n gets larger and larger, the two-dimensional Gaussian expo-
nential in Eq. (1.340) approaches to the two-dimensional Dirac delta function.
For the left bump, n = 400; for the middle bump, n = 1000; and for the right
spike n = 4000.
and the integral of δn over the entire x y z-plane equals to one. In this way,
which is zero everywhere except at the point which makes all three of its ar-
gument arguments zero, in which case it is infinite, i.e.,
Z1
〈 f |g 〉 = d x f ∗ (x)g (x). (1.349)
−1
Z1
〈 fk | fl 〉 = d x | f k (x) − f l (x)|2 → 0 (1.350)
−1
lim 〈 f | f k 〉 = 0 (1.352)
k→∞
does not exist. In this way, we arrive to the conclusion that the space of contin-
uous functions form a linear space, but it is not complete with respect to the
inner product defined in Eq. (1.349) since in the limit k → ∞, the sequence
{ f k }∞
k=1
does not give an element of the vector space considered. There is
something missing in the vector space: discontinuous functions!
Then, to have L 2 (a, b) or L 1 (a, b) as complete spaces with respect their
inner products, we need to include discontinuous functions like the one in
1.12 A PPENDIX : T HE L EBESGUE INTEGRAL 75
But then comes the problem: How to calculate the inner product of these
functions? In other words, how to determine
Zb
d x | f (x)|2 (1.354)
a
Zb
d x | f (x)| (1.355)
a
in case of f (x) ∈ L (a, b) with f (x) being such discontinuous functions. Here
is precisely where the name of Lebesgue becomes relevant! To determine such
integrals the familiar notion of the Riemann integral needs to be generalized
so as to be able to integrate such rebellious functions. This generalization is
achieved by introducing the Lebesgue integral, which is equal to the Riemann
integral for functions that are integrable in the conventional sense. In the fol-
lowing we simply provide the basics of the Lebesgue method for integration
to give to the reader only a very general idea of what a Lebesgue integral is.
The principal difference between the integrals of Riemann and Lebesgue
may be illustrated by pictures. Figure 1.17 shows a positive-continuous func-
tion f defined on an interval −∞ < a ≤ x ≤ b < ∞, subdivided as for Riemann
1.12 A PPENDIX : T HE L EBESGUE INTEGRAL 76
in which x k0 is any point between x k−1 and x k , and verify that this sum ap-
proaches a limit, namely the Riemann integral
Zb
d x f (x), (1.358)
a
where
Zb Zb
d x f (x) = inf[U (P, f )], d x f (x) = sup[L(P, f )]. (1.362)
a a
1.12 A PPENDIX : T HE L EBESGUE INTEGRAL 77
f(x)
Note that the upper Riemann integral of f is always greater than or equal to
the lower Riemann integral. When the two are equal to each other, we say
that f is Riemann integrable on [a, b], and we call this common value the Rie-
mann integral of f . Let us then determine the integral of the function f (x)
in Eq. (1.353) (see Fig. 1.19). If we partition the domain of this function, then
each subinterval will contain both rational and irrational numbers. Thus, the
supremum on each subinterval is 1 and the infimum on each subinterval is 0.
Then
Zb Zb
d x f (x) = b − 1 and d x f (x) = 0, (1.363)
a a
1.12 A PPENDIX : T HE L EBESGUE INTEGRAL 78
so the upper and lower Riemann integrals are different: this function is not
Riemann integrable.
What to do then with this kind of functions? Lebesgue simply turned the
Riemann’s recipe for integration on its side and subdivided the range of the
function instead of the domain. The idea, as indicated by the different shad-
ings in Fig. 1.20, being to lump together the points at which the function takes
on (approximately) the same values. This would appear to be a perfectly triv-
ial modification, but it has far-reaching consequences (however, those details
are beyond the scope of this Appendix)! Lebesgue’s recipe tells you first to
subdivide the vertical axis by a series of points
Zb
d x f (x) (1.366)
a
Zb
d x| f (x)|2 , (1.367)
a
2
I NTRODUCTION TO PARTIAL
DIFFERENTIAL EQUATIONS
HENRI POINCARÉ
d 4u d 2u
+ + u 2 = cos(x) (2.1)
d x4 d x2
is a differential equation for the function u(x) depending on a single (inde-
pendent) variable x, while
∂u ∂2 u ∂2 u
= 2 + 2 −u (2.2)
∂t ∂x ∂y
81
82
u t = u xx + u y y − u. (2.3)
by
1 ∂2 u ∂2 ∂2 ∂2
∇2u = , ∇2 = + + . (2.4)
c 2 ∂t 2 ∂x 2 ∂y 2 ∂z 2
∂u
∇2 u =
κ∇ . (2.5)
∂t
This equation describes the temperature u(rr , t ) in a region containing
no heat sources or sinks. It also applies to the diffusion of a chemical
that has a concentration u(rr , t ). The constant κ is called the diffusivity.
The equation is clearly second order in the three spatial variables, but
first order in time.
∇ 2 u = 0, (2.6)
ħ 2 ∂u
− ∇ u + V (rr )u = i ħ . (2.7)
2m ∂t
This equation describes the quantum mechanical wavefunction u(rr , t )
of a non-relativistic particle of mass m moving in the force field pre-
scribed by the (real) potential function V (rr ). While the solution u is
complex-valued, the independent variables t , x, representing time and
space, remain real. As in case of the diffusion equation it is second order
in the three spatial variables and first order in time.
2.1 I NITIAL CONDITIONS AND BOUNDARY CONDITIONS 84
How many solutions does a partial differential equation have? In general, lots.
Even ordinary differential equations have infinitely many solutions. Indeed,
the general solution to a single nth order ordinary differential equation de-
pends on n arbitrary constants. The solutions to partial differential equa-
tions are yet more numerous, in that they depend on arbitrary functions. Very
roughly, we can expect the solution to an nth order partial differential equa-
tion involving m independent variables to depend on n arbitrary functions of
m − 1 variables. But this must be taken with a large grain of salt−only in a few
special instances will we actually be able to express the solution in terms of
arbitrary functions.
An ordinary or a partial differential equation will provide a unique solu-
tion to a physical problem only if the initial or the starting value of the so-
lution is known. We refer to this as the boundary conditions. For ordinary
differential equations, when time is involved, boundary conditions amount
to the specification of one or more properties of the solution at an initial time;
that is why for ordinary differential equations involving time one speaks of
initial conditions. For example,
dy
= a y(t ), a ∈ R, (2.8)
dt
is a differential equation, i.e., and equation involving both y(t ) and d y/d t and
is an ordinary differential equation. Its solution is given by
y(t ) = be at , b ∈ R, (2.9)
where each value of b defines a different solution. How to know that Eq. (2.9)
is indeed a solution? We just need to substitute Eq. (2.9) into Eq. (3.129) and
check if Eq. (3.129) is satisfied:
· ¸ · ¸
d at at at
be = abe = a be , (2.10)
dt
which is precisely Eq. (3.129). So, as we have mentioned, the first thing that
we have observed is that the solution of a differential equation need not to be
unique. The set of all solutions of a differential equation is called general so-
lution. To fix a unique specific solution we need to impose some conditions.
For example, we might require that the solution of Eq. (3.129) obey an initial
condition such as y(0) = 1. If we impose such condition, using Eq. (2.9), we fix
a particular solution with b = 1.
2.1 I NITIAL CONDITIONS AND BOUNDARY CONDITIONS 85
ferential equation, the initial conditions, and the boundary conditions leads
to an initial-boundary value problem.
An additional consideration is that, besides any smoothness required by
the partial differential equation within the domain, the solution and any of
its derivatives specified in any initial or boundary condition should also be
continuous at the initial or boundary point where the condition is imposed.
For example, if the initial condition specifies the function value u(0, x) for
a < x < b, while the boundary conditions specify the derivatives ∂u ∂x
(t , a) and
∂u
∂x
(t , b) for t > 0, then, in addition to any smoothness required inside the do-
main {a < x < b, t > 0}, we also require that u be continuous at all initial points
(0, x), and that its derivative ∂u∂x
be continuous at all boundary points (t , a) and
(t , b), in order that u(t , x) qualify as a solution to the initial-boundary value
problem.
L[u] = 0, (2.15)
∂2 ∂2 u ∂2 v
L[u + v] = (u + v) = + = L[u] + L[v],
∂x 2 ∂x 2 ∂x 2
∂2 ∂2 u
L[cu] = 2 (cu) = c 2 = cL[u], (2.17)
∂x ∂x
2.2 L INEAR AND NONLINEAR EQUATIONS 88
which are valid for any functions continuously differentiable twice, u, v and
any constant c. The corresponding homogeneous linear differential equation
L[u] = 0 is
∂2 u
= 0. (2.18)
∂x 2
Let’s consider now
As we can see
The defining attributes of linear operators (2.3.1) imply the key properties
shared by all homogeneous linear differential equations: If u 1 , . . . , u k are so-
lutions to a common homogeneous linear equation L[u] = 0, then the linear
combination, or superposition, u = c 1 u 1 +· · ·+c k u k is a solution for any choice
of constants c 1 , . . . , c k . This is called the superposition principle. Indeed,
L[v] = f , (2.24)
tions like the wave equation, the Laplace equation, etc., is of the form
dnu d n−1 u du
L[u] = a n (x) n
+ a n−1 (x) n−1
+ · · · + a 1 (x) + a 0 (x)u = 0, (2.28)
dx dx dx
which is an homogeneous, linear, ordinary differential equation with variable
coefficients, i.e., the a i , i = 1, 2, . . . , n appearing in Eq. (2.28) depend on x. The
general solution of Eq. (2.28) will contain n arbitrary constants, which can be
determined if n boundary conditions are also provided.
To determine u(x), since we have c i , i = 1, 2, . . . , n, arbitrary constants that
may be determined if n boundary conditions are provided, we need to find n
solutions, u 1 (x), u 2 (x), . . . , u n (x) and, due to the linearity of the differential
equation, construct the linear superposition
over the domain in question, for any set of constants c 1 , c 2 , . . . , c n , except for
the trivial case c 1 = c 2 = · · · = c n = 0. A statement equivalent to the above equa-
tion, which is perhaps more useful for the practical determination of linear
independence can be obtained by repeatedly differentiating Eq. (2.30) n − 1
times in all, to obtain n simultaneous equations for c 1 , c 2 , . . . , c n :
on the domain considered is non-zero, then the only solution to Eq. (2.31)
is the trivial solution c 1 = c 2 = · · · = c n = 0, thus, the n functions u 1 (x), u 2 (x),
u n (x) are linearly independent on the domain. This determinant W (u 1 , u 2 , . . . , u n )
is called the Wronskian of the set of functions. Note, however, that, in gen-
eral, the vanishing of the Wronskian does not guarantee that the functions are
linearly dependent. For example, if we consider the functions u 1 (x) = x and
u 2 (x) = |x|, we have that ddux1 = 1 and ddux2 = x/|x|. Then
¯ ¯
¯ x |x| ¯ x 2
¯
W (u 1 , u 2 ) = ¯ ¯
x ¯= − |x| = |x| − |x| = 0, (2.34)
1 |x| |x|
Then we might conclude that u 1 (x) = x and u 2 (x) = x/|x| are not linearly inde-
pendent, and that such conclusion is valid no matter the interval considered
for x. Note, however, that ddux2 = x/|x| does not exist at x = 0, and if we consider
a domain where x = 0 is part of it, the vanishing of W (u 1 , u 2 ) does not imply
that u 1 and u 2 are linearly dependent. However, it is possible to demonstrate
that if u i (x) are solutions to an nth order ordinary linear differential equation,
which is our case, and the Wronskian W (u 1 , u 2 , . . . , u n ) vanishes, then {u i }ni=1
is a linearly dependent set of functions. Moreover, if the Wronskian does not
vanish for some value of x, then it does not vanish for all values of x, in which
case an arbitrary linear combination of the u i (x) constitutes, as stated before,
the most general solution to the nth order ordinary linear differential equa-
tion.
1. All roots are real and distinct. In this case, the n solutions to Eq. (2.35)
are u i (x) = e λi x , i = 1, 2, . . . , n, and the related Wronskian would be not
zero since all λi are different to each other. Thus, the solution u(x) is
given by the linear superposition of all the u i (x), i.e.,
u(x) = c 1 e λ1 x + c 2 e λ2 x + · · · + c n e λn x . (2.39)
2. Some roots are complex. If all a i coefficients are real and one of the
roots of Eq. (2.38) is complex, say λR + i λI , with λR , λI ∈ R, then its
complex conjugate λR − i λI is also a root. In this case, when combining
them into a linear superposition, we will have that
c 1 e (λR +i λI )x + c 2 e (λR −i λI )x = e λR x [c 1 e i λI x + c 2 e −i λI x ]
= e λR x [c 1 {cos(λI x) + i sin(λI x)}
+ c 2 {cos(λI x) − i sin(λI x)}]
= e λR x [(c 1 + c 2 )cos(λI x) + i (c 1 − c 2 )sin(λI x)]
≡ e λR x [αcos(λI x) + βsin(λI x)]
= e λR x Acos(λI x + ϕ) or e λR x B sin(λI x + η),
(2.40)
3. Some roots are repeated. If, for example, the root λ1 occurs k times
(k > 1), then we have not found n linearly independent solutions of
Eq. (2.35). We must find k − 1 further solutions that are linearly inde-
pendent of those already found and also of each other. Interestingly, by
direct substitution into Eq. (2.35), if e λ1 x is a solution,
are also solutions, it is easily shown that they, together with the solu-
tions already found, form a linearly independent set of n functions. In
this way,
The above argument can be easily extended if more than one root is
repeated. For example, suppose as before that λ1 is a k-fold root of
Eq. (2.38) and, further, that λ2 is an l -fold root (both k, l >1). Then, the
solution u(x) reads
d 2u du
2
−2 + u = 0. (2.44)
dx dx
Using as solution for this equation the function Ae λx , we get the polynomial
λ2 − 2λ + 1 = 0. (2.45)
The above equation has as root λ1 = 1, which occurs twice. Thus, e λ1 x and
xe λ1 x are two linearly independent solutions and
as expected.
thus, we have two roots, λ1 = 2i and λ2 = −2i . Then, using Eq. (2.40), the
solution u(x) is given by
where α, β are arbitrary constants which can be determined from the bound-
ary conditions. We can check our solution:
du
= −2αsin(2x) + 2βcos(2x),
dx
d 2u
= −4αcos(2x) − 4βsin(2x) = −4u, (2.51)
d x2
as expected.
d 2u du
2
+ P (x) +Q(x)u = 0, (2.52)
dx dx
and its solution can be written as
The solutions u 1 (x) and u 2 (x) are linearly independent, thus, as we saw, the
Wronskian
¯ ¯
¯ u1 u2 ¯ d u2 du 1
W (x) = ¯ d u1 du2 ¯¯ = u 1
¯ − u2 6= 0. (2.54)
dx dx dx dx
2.3.2 VARIABLE COEFFICIENTS : SERIES SOLUTION 96
So far we have always assumed that u(x) is a real function of a real variable
x. However, this is not always the case, and we are going to broaden our dis-
cussions in this section by generalizing u(x) to a complex function u(z) of a
complex variable z. We thus consider the second-order linear homogeneous
equation
d 2u du
+ P (z) +Q(z)u(z) = 0, (2.55)
d z2 dz
where differentiation with respect to z is treated in a way analogous to ordi-
nary differentiation with respect to a real variable x. We limit our considera-
tions to the cases where the functions P (z) and Q(z) are analytic in a certain
domain R, except at an enumerable number of points of R where these func-
tions may have isolated singularities.
If at some point z = z 0 the functions P (z) and Q(z) are finite and can be
expressed as complex power series about z 0
X
∞ X
∞
P (z) = P n (z − z 0 )n , Q(z) = Q n (z − z 0 )n , (2.56)
n=0 n=0
then P (z) and Q(z) are said to be analytic at z = z 0 , and this point is called
an ordinary point of the ordinary differential equation. If, however, P (z) or
Q(z), or both, diverge at z = z 0 , then it is called singular point of the ordinary
differential equation. Even if an ordinary differential equation is singular at a
given point z = z 0 , it may still possess a non-singular solution at that point.
In fact, the necessary and sufficient condition for such a solution to exist is
that (z − z 0 )P (z) and (z − z 0 )2Q(z) are both analytic at z = z 0 . Singular points
that have this property are called regular singular points, whereas any singu-
lar point not satisfying both these criteria is called an irregular or essential
singularity.
Sometimes z 0 might not be finite and we might need to determine the
nature of the point |z| → ∞. This can be done by simply substituting w = 1/z
into the differential equation and investigating the behavior at w = 0.
d 2u du
(1 − z 2 ) 2
− 2z + l (l + 1)u = 0, (2.57)
dz dz
where l is a constant. Let us show that z = 0 is an ordinary point, z = ±1 and
|z| → ∞ are regular singular points of the equation. First, Eq. (2.57) can be
written as
d 2u 2z d u l (l + 1)
− + u = 0, (2.58)
dz 2 1 − z2 d z 1 − z2
2.3.2 VARIABLE COEFFICIENTS : SERIES SOLUTION 97
2z 2z l (l + 1) l (l + 1)
P (z) = − = − , Q(z) = = . (2.59)
1 − z2 (1 + z)(1 − z) 1 − z2 (1 + z)(1 − z)
It is then clear that P (z) and Q(z) are both analytic at z = 0 and both diverge at
z = ±1. Then, z = 0 is an ordinary point of the differential equation and z = ±
are singular points. At z = 1,
2z 1−z
(z − 1)P (z) = , (z − 1)2Q(z) = l (l + 1) (2.60)
1+z 1+z
and they are both analytic at z = 1. Hence, z = 1 is a regular singular point.
Similarly, at z = −1, both (z + 1)P (z) and (z + 1)2Q(z) are analytic, thus, z = −1
is another regular singular point of the equation.
Next, letting w = 1/z,
du du dw 1 du du
= =− 2 = −w 2 ,
dz dw dz z dw dw
µ ¶ µ ¶
d 2u d 2 du dw d 2 du
= −w = −w
d z2 d z dw dz dw dw
µ 2 ¶ µ ¶
2 du 2d u 3 du d 2u
= −w − 2w −w =w 2 +w . (2.61)
dw dw2 dw dw2
d 2u 3 du
w 2 (w 2 − 1) + 2w + l (l + 1)u = 0. (2.63)
dw2 dw
2w l (l + 1)
P (w) = , Q(w) = . (2.64)
w2 − 1 w 2 (w 2 − 1)
At w = 0, P (w) is analytic but Q(w) diverges, and so the point |z| → ∞ is a sin-
gular point of the Legendre’s equation. However, wince wP (w) and w 2Q(w)
are both analytic at w = 0, |z| → ∞ is a regular singular point.
1) S ERIES SOLUTIONS ABOUT AN ORDINARY POINT. 98
X
∞
u(z) = an z n . (2.65)
n=0
Moreover, it can be shown that such a power series converges for |z| < R,
where R is the radius of convergence and is equal to the distance from z = 0 to
the nearest singular point of the ordinary differential equation. At the radius
of convergence, however, the series may or may not converge.
Since every solution of Eq. (2.55) is analytic at an ordinary point, it is al-
ways possible to obtain two independent solutions of the form (2.65) from
which the general solution
can be constructed.
Using Eq. (2.65),
du X ∞ X∞
= na n z n−1 = (n + 1)a n+1 z n ,
d z n=0 n=0
d u X
2 ∞
n−2
X∞
= n(n − 1)a n z = (n + 2)(n + 1)a n+2 z n . (2.67)
d z 2 n=0 n=0
Substituting the above expressions into Eq. (2.55) and requiring that the co-
efficients of each power of z sum to zero, we obtain a recurrence relation ex-
pressing each a n in terms of the previous a r , 0 ≤ r ≤ n − 1. In some cases we
may find that the recurrence relation leads to a n = 0 for some n greater than a
value N , for one or both of the two solutions u 1 (x) and u 2 (x). In such a case,
the series solution becomes a polynomial, thus, the solution would converge
for all finite z.
E XAMPLE 2.5. Let’s determine the series solution about z = 0 of the differential
equation
d 2u
+ u = 0. (2.68)
d z2
1) S ERIES SOLUTIONS ABOUT AN ORDINARY POINT. 99
For this equation to be satisfied we need that the coefficients of each power of
z vanishes separately, thus,
an
(n + 2)(n + 1)a n+2 + a n = 0 =⇒ a n+2 = − , n ≥ 0. (2.70)
(n + 2)(n + 1)
Using this equation, for a given a 0 , we can calculate the even coefficients, i.e.,
a 2 , a 4 , a 6 , etc., while for a given a 1 , we can determine the odd coefficients, i.e.,
a 3 , a 5 , a 7 , etc. Two independent solutions can be obtained by setting either
a 0 or a 1 to zero and choosing the other coefficient equal to 1. For example, if
we set a 0 = 0 and choose a 1 = 1, all the even coefficients a 2n , n = 0, 1, 2, . . . are
equal to zero. For the even coefficients, using Eq. (2.70)
a1 1 a3 a1 1
a3 = − = − , a5 = − = = , ...
3·2 3! 5 · 4 5 · 4 · 3 · 2 5!
(−1)n
a 2n+1 = , n = 0, 1, 2, . . . (2.71)
(2n + 1)!
a0 1 a2 1 1
a2 = − = − , a4 = − = = , ...
2 2! 4 · 3 4 · 3 · 2 4!
(−1)n
a 2n = , n = 0, 1, 2, . . . (2.73)
(2n)!
X∞ (−1)n
u 2 (z) = z 2n . (2.74)
n=0 (2n)!
Note that both series converge for all z, as might be expected since Eq. (2.68)
possesses no singular point, except |z| → ∞. Interestingly, the series in Eqs. (2.72)
1) S ERIES SOLUTIONS ABOUT AN ORDINARY POINT. 100
and (2.74) correspond to the series expansion of sin(z) and cos(z), respec-
tively, around z = 0. Then, we can write the solution of Eq. (2.68) as
Solving the above example was quite straightforward and the resulting se-
ries were easily recognized and written in closed form, i.e., in terms of elemen-
tary functions. But this is not usually the case. Another simplifying feature of
the previous example was that we obtained a two-term recurrence relation
relating a n+2 and a n , so that the odd- and even-numbered coefficients were
independent of one another. This is also not usually the case and, in general,
the recurrence relation expresses a n in terms of any number of the previous
a r , 0 ≤ r ≤ n − 1.
Then,
· ¸
X
∞
(n + 2)(n + 1)a n+2 − 2(n + 1)na n+1 + (n(n − 1) − 2)a n z n = 0. (2.81)
n=0
In this way, given a 0 and a 1 , we can get any other coefficient a n . From Eq. (2.118),
one obvious solution is, for example, a n = a 0 for all n , since we will get
which is satisfied for any value of n. Choosing then a 0 = 1, we find the solution
X
∞ X
∞
u 1 (z) = an z n = zn = 1 + z + z2 + z3 + . . . (2.85)
n=0 n=0
n=0: 2a 2 − 2a 0 = 0,
n=1: 3a 3 − 2a 2 − a 1 = 0,
n=2: 4a 4 − 4a 3 = 0,
n=3: 5a 5 − 6a 4 + a 3 = 0,
..
. (2.87)
Note that this solution is valid for all finite values of z. In this way, the solution
of Eq. (2.77) is given by
c1
u(z) = c 1 u 1 (z) + c 2 u 2 (z) = + c 2 (1 − z)2 . (2.89)
1−z
The linear independence of u 1 and u 2 is obvious but can be checked by com-
puting the Wronskian. Using Eq. (2.54) (and changing x to z),
1 1
W (z) = [−2(1 − z)] − (1 − z)2 = −3 6= 0, (2.90)
1−z (1 − z)2
where the exponent σ is a number that may be real or complex and where
a 0 6= 0 (since, if it were otherwise, σ could be redefined as σ + 1 or σ + 2, etc.,
so as to make a 0 6= 0). Such a series is called a generalized power series or
Frobenius series. As in the case of a simple power series solution, the radius
of convergence of the Frobenius series is, in general, equal to the distance to
the nearest singularity of the ordinary differential equation.
Let’s define
S(z) ≡ zP (z), T (z) ≡ z 2Q(z), (2.98)
such that in terms of S(z) and T (z), Eq. (2.52) can be written as
d 2 y S(z) d y T (z)
+ + 2 y = 0. (2.99)
d z2 z dz z
Substituting now Eq. (2.97) in (2.99), since
dy X ∞
= (n + σ)a n z n+σ−1 ,
d z n=0
d2y X ∞
= (n + σ)(n + σ − 1)a n z n+σ−2 , (2.100)
d z 2 n=0
we get
X
∞ X
∞ X
∞
(n + σ)(n + σ − 1)a n z n+σ−2 + S(z) (n + σ)a n z n+σ−2 + T (z) a n z n+σ−2 = 0.
n=0 n=0 n=0
(2.101)
2) S ERIES SOLUTIONS ABOUT A REGULAR SINGULAR POINT. 104
which is valid for any z. Setting z = 0, all terms in the sum with n > 0 vanish
and we get
The linear independence of these two solutions follows from the fact that
u 2 /u 1 is not a constant since σ2 − σ1 is not an integer. Then, the general solu-
tion is given by
Note that σ1 and σ2 can be complex numbers where σ2 = σ∗1 . In such a case,
σ1 − σ2 = σ1 − σ∗1 = 2i Im[σ1 ], which is purely imaginary , thus, σ1 − σ2 cannot
be equal to an integer, as required.
2) S ERIES SOLUTIONS ABOUT A REGULAR SINGULAR POINT. 105
d 2u du
4z 2
+2 + u = 0. (2.107)
dz dz
The above equation can be written as
d 2u 1 du 1
2
+ + y = 0. (2.108)
dz 2z d z 4z
If we compare the latter equation with (2.55), we can identify P (z) and Q(z),
1 1
P (z) = , Q(z) = . (2.109)
2z 4z
Clearly, z = 0 is a singular point of the differential equation, but since
1 z
S(z) = zP (z) = , Q(z) = z 2Q(z) = (2.110)
2 4
are finite at z = 0, z = 0 is a regular singular point. We then use a Frobenius
series as in Eq. (2.97) to determine u(z). From Eqs. (2.104) and (2.110),
µ ¶
1 1
σ(σ − 1) + σ = 0 =⇒ σ σ − = 0, (2.111)
2 2
which has roots σ1 = 1/2 and σ2 = 0. Since these roots do not differ by an
integer, we expect to find two independent solutions to Eq. (2.108). From
Eq. (2.102),
X∞ · 1 z
¸
(n + σ)(n + σ − 1) + (n + σ) + a n z n = 0, (2.112)
n=0 2 4
a0 1 a1 1 (−1)n
a1 = − =− , a2 = − = , ... , an = , n = 0, 1, 2, . . .
3·2 3! 5 · 4 5! (2n + 1)!
(2.116)
a0 1 a1 1 (−1)n
a1 = − =− , a2 = − = , ... , an = , n = 0, 1, 2 . . .
2·1 2! 4 · 3 4! (2n)!
(2.119)
X∞ (−1)n z z2
u 2 (z) = zn = 1 − + − . . .
n=0 (2n)! 2! 4!
p 2 p 4
z z p
= 1− + − · · · = cos( z). (2.120)
2! 4!
The linearly independence of Eqs. (2.117) and (2.120) can be checked by cal-
culating the Wronskian for these solutions. Using Eq. (2.54)
· ¸ · ¸
p 1 p p 1 p
W (z) = sin( z) − p sin( z) − cos( z) p cos( z)
2 z 2 z
1
= − p 6= 0. (2.121)
2 z
d u2 d u1
W (z) = u 1 (z) − u 2 (z) . (2.123)
dz dz
Zz
W (u)
u 2 (z) = u 1 (z) du . (2.125)
u 12 (u)
dW d u 1 d u 2 d 2 u2 d u2 d u1 d 2 u1 d 2 u2 d 2 u1
= + u1 − − u 2 = u 1 − u 2 . (2.126)
dz dz dz d z2 dz dz d z2 d z2 d z2
In this way
µ ¶ Z
dW W
= −P (z)d z =⇒ ln = − d zP (z)
W W0
R
=⇒ W (z) = W0 e − d zP (z)
, (2.128)
d 2u du
z(z − 1) 2
+ 3z + u = 0. (2.130)
dz dz
By writing the above equation as
d 2u 3 du 1
+ + u=0 (2.131)
dz 2 z − 1 d z z(z − 1)
and comparing with Eq. (2.55), we can identify P (z) and Q(z),
3 1
P (z) = , Q(z) = . (2.132)
z −1 z(z − 1)
As we can see, z = 0 is a singular point, but since
3z z
S(z) = zP (z) = , T (z) = z 2Q(z) = , (2.133)
z −1 z −1
2) S ERIES SOLUTIONS ABOUT A REGULAR SINGULAR POINT. 109
are finite there, it is a regular singular point and we expect to find at least
one solution in the form of a Frobenius series. Using Eq. (2.104), since S(0) =
T (z) = 0, we get
σ(σ − 1) = 0, (2.134)
thus, we have the roots σ1 = 1 and σ2 = 0. Since the roots differ by an integer
(unity), it may not be possible to find two linearly independent solutions of
Eq. (2.131) in the form of Frobenius series. We are guaranteed, however, to
find one such solutions corresponding to the larger root, σ1 . Using Eq. (2.102)
for σ = σ1 and (2.133), we get
∞ ·
X 3z z
¸
(n + 1)n + (n + 1) + a n z n = 0. (2.135)
n=0 z − 1 z − 1
X∞ · ¸
(n + 1)n(z − 1) + 3z(n + 1) + z a n z n = 0,
n=0
X∞ · ¸
=⇒ {(n + 1)(n + 3) + 1}z − n(n + 1) a n z n = 0, (2.136)
n=0
n +1
(n + 1)a n−1 − na n = 0 =⇒ an = a n−1 . (2.138)
n
Setting a 0 = 1, we get
3 4
a 1 = 2a 0 = 2, a 2 = a 1 = 3, a 3 = a 2 = 4, ..., a n = n + 1, n = 0, 1, 2, . . .
2 3
(2.139)
Using Eq. (2.97), one of the solutions for Eq. (2.131) is then
X
∞
u 1 (z) = z (n + 1)z n = z(1 + 2z + 3z 2 + . . . ). (2.140)
n=0
2) S ERIES SOLUTIONS ABOUT A REGULAR SINGULAR POINT. 110
In this way, since the above equation is valid for an arbitrary value of z, we get
that
n
an = a n−1 . (2.145)
n −1
Since we require a 0 6= 0, we see that the above expression produces an a 1
which is infinite, thus, within this method, we can not get a second solution u 2
related to the root σ2 . We consider then the Wronskian method to determine
u 2 (z). Using Eq. (2.129) and substituting u 1 (z) and P (z),
Zz
z (1 − u)4 − Ru d v 3
u 2 (z) = du e v−1
(1 − z)2 u2
Zz
z (1 − u)4 −3ln(u−1)
= du e
(1 − z)2 u2
Zz · ¸
z u −1 z 1
= du 2 = ln(z) + . (2.146)
(1 − z)2 u (1 − z)2 z
2.4 S EPARATION OF VARIABLES 111
We can now calculate the Wronskian of u 1 (z) and u 2 (z) to show, as expected,
that the two solutions are linearly independent. In fact, the Wronskian has
already been evaluated as W (u) = e −3ln(u−1) = (u−1)−3 , i.e., W (z) = 1/(z−1)3 6=
0. In this way, the general solution to Eq. (2.131) is given by
µ · ¸¶
z 1
u(z) = c 1 u 1 (z) + c 2 u 2 (z) = c 1 + c 2 ln(z) + . (2.147)
(1 − z)2 z
X
∞
u 1 (z) = z σ an z n . (2.148)
n=0
1 ∂2 u
∇2u = . (2.150)
c 2 ∂t 2
2.4.1 C ARTESIAN COORDINATES 112
For the present, we are going to work in Cartesian coordinates and assume a
solution of the form (2.149), and later on we will work in other coordinate sys-
tems, like spherical or cylindrical. In Cartesian coordinates, Eq. (2.150) takes
the form
∂2 u ∂2 Y ∂2 u 1 ∂2 u
+ + = . (2.151)
∂x 2 ∂y 2 ∂z 2 c 2 ∂t 2
Substituting Eq. (2.149), we get
d2X d 2Y d2Z 1 d 2T
Y Z T + X Z T + X Y T = X Y Z . (2.152)
d x2 d y2 d z2 c2 dt2
If we now divide the above equation throughout by u = X Y Z T , we obtain
1 d2X 1 d 2Y 1 d2Z 1 1 d 2T
+ + = . (2.153)
X d x2 Y d y 2 Z d z2 c2 T d t 2
In this way, of the four terms in the equation, the first one is a function of x
only, the second of y only, the third of z only and the right-hand side a func-
tion of t only, and yet there is an equation connecting them. This can only
be so for all x, y, z and t if each of the terms does not in fact depend upon
the corresponding independent variable but is equal to a constant such that
Eq. (2.153) is satisfied. Let us make the choice,
1 d2X 1 d 2Y 1 d2Z 1 1 d 2T
2
= −l 2 , 2
= −m 2 , 2
= −n 2 , 2 2
= −µ2 . (2.154)
X dx Y dy Z dz c T dt
Then, from Eq. (2.153), the relation between the four constants l , m, n and µ
is given by
µ2 = l 2 + m 2 + n 2 . (2.155)
These constants are called separation constants. The important point to no-
tice is that by assuming a separable solution, the partial differential equation
(2.151), which contains derivatives with respect to the four independent vari-
ables all in one equation, has been reduced to four separate ordinary differen-
tial equations (2.154), which are connected through four constant parameters
that satisfy the algebraic equation (2.155).
The general methods for solving ordinary differential equations (see Sec. 2.3.1)
show that the solutions of equations (2.154) are given by
X (x) = Ae i l x + B e −i l x = A 0 cos(l x) + B 0 sin(l x),
Y (y) = C e i m y + De −i m y = C 0 cos(m y) + D 0 sin(m y),
Z (z) = E e i nz + F e −i nz = E 0 cos(nz) + F 0 sin(nz),
T (t ) = Ge i cµt + He −i cµt = G 0 cos(cµt ) + H 0 sin(cµt ), (2.156)
2.4.1 C ARTESIAN COORDINATES 113
E XAMPLE 2.9. Use the method of separation of variables to obtain for the one-
dimensional diffusion equation
∂2 u ∂u
κ = , (2.159)
∂x 2 ∂t
a solution that tends to zero as t → ∞ for all x.
First, in this case, we have only two independent variables x and t , thus,
we assume a solution of the form
1 d2X 1 1 dT
= . (2.161)
X dx 2 κ T dt
Since the left-hand side is a function of x only, and the right-hand side is a
function of t only, Eq. (2.161) implies that each side must equal a constant.
For convenience, we choose
1 d2X 1 1 dT
= −λ2 , = −µ2 , (2.162)
X d x2 κ T dt
and, from Eq. (2.161), we have the relation
µ2 = λ2 (2.163)
2.4.2 S UPERPOSITION OF SEPARATED SOLUTIONS 114
is also a solution for any constants a i , provided that the λi are the allowed
values of the separation constant λ given the imposed boundary conditions.
Note that if the boundary conditions allow any of the separation constants to
be zero, then the form of the general solution is normally different and must
be deduced by returning to the separated ordinary differential equations.
The value of the superposition approach is that a boundary condition, say
that u(x, y) takes a particular form f (x) when y = 0, might be met by choosing
the constants a i such that
X
f (x) = a i X λi (x)Yλi (0). (2.168)
i
In general, this will be possible provided that the functions X λi (x) form a com-
plete set (as do the sinusoidal functions of Fourier series).
2.4.2 S UPERPOSITION OF SEPARATED SOLUTIONS 115
Figure 2.1: A semi-infinite metal plate whose edges are kept at fixed tempera-
tures.
In this case, we are asked to find the steady-state temperature, which corre-
sponds to ∂u/∂t = 0, and so, we are left with the two-dimensional equation
∂2 u ∂2 u
+ = 0, (2.170)
∂x 2 ∂y 2
1 d2X 1 d 2Y
+ = 0, (2.171)
X d x2 Y d y 2
which implies
1 d2X 1 d 2Y
= −l 2 , = −m 2 , (2.172)
X d x2 Y d y2
2.4.2 S UPERPOSITION OF SEPARATED SOLUTIONS 116
with the separation constants l 2 = −m 2 [from Eq. (2.171)]. In the current prob-
lem, we have to satisfy the boundary conditions u(x, 0) = u(x, b) = 0. A sinu-
soidal expression for Y (y) seems then more appropriate than an exponential
form. Furthermore, we also require u(∞, y) = 0, thus, an exponential form for
X (x) seems more convenient. Then, we write
1 d2X
= m2 =⇒ X (x) = Ae mx + B e −mx ,
X d x2
1 d 2Y
2
= −m 2 =⇒ Y (y) = C cos(m y) + Dsin(m y), (2.173)
Y dy
and, thus,
for some constants B k [note that the term k = 0 is identically zero, thus, we
have omitted in the sum in Eq. (2.177)]. Using the remaining boundary con-
dition u(0, y) = f (y), we see that the constants B k must satisfy
µ ¶
X∞ kπy
f (y) = B k sin . (2.178)
k=1 b
This is clearly a Fourier series expansion of f (y)! For Eq. (2.178) to hold, how-
ever, the continuation of f (y) outside the region 0 ≤ y ≤ b must be an odd
2.4.2 S UPERPOSITION OF SEPARATED SOLUTIONS 117
periodic function with period 2b (see Fig. 2.2). We also see from the figure
that if the original function f (y) does not equal zero at either of y = 0 and
y = b then its continuation has a discontinuity at the corresponding point(s).
Nevertheless, as discussed in Chapter 1, the Fourier series will converge to the
mid-points of these jumps and hence tend to zero in this case. If, however,
the top and bottom edges of the plate were held not at 0◦ C but at some other
non-zero temperature, then, in general, the final solution would posses dis-
continuities at the corners x = 0, y = 0 and x = 0, y = b.
Often, the principle of superposition can be also used to: (1) write the so-
lution to problems with more complicated boundary conditions as the sum of
solutions to problems that each satisfy only some part of the boundary con-
dition but when added together satisfy all the conditions. (2) For dealing with
inhomogeneous boundary conditions. In general, inhomogeneous boundary
conditions can cause difficulties and it is usual to attempt a suitable change
of variables and transform the problem into an equivalent homogeneous one.
But we are not going to enter into these details.
In cylindrical coordinates, see Fig. 2.3, the position of a point in space hav-
ing Cartesian coordinates x, y, z can be expressed in terms of ρ, ϕ, z,
and ρ ≥ 0, 0 ≤ ϕ ≤ 2π and −∞ < z < ∞. In this way, the vector position r of the
point can be written as
r = ρcosϕii + ρsinϕjj + zk
k. (2.183)
e ρ = cosϕii + sinϕjj ,
e ϕ = −sinϕii + cosϕjj ,
ez =k. (2.184)
These three unit vectors, like the Cartesian unit vectors i , j and k form a ba-
sis. Plane polar coordinates correspond to a situation where the point P is on
the x y plane, thus, they can be obtained as particular case of the cylindrical
coordinates for which z = 0.
In case of spherical coordinates, the position of a point in space with Carte-
sian coordinates x, y, z can be expressed in terms of r , θ, ϕ (see Fig. 2.3), where
∂2 ∂2 ∂2
∇2 = + + . (2.187)
∂x 2 ∂y 2 ∂z 2
The last term depends only on z, and the first and second, taken together,
depend only on ρ and ϕ. Taking the first separation constant to be k 2 , we find
1 d2Z
= k 2,
Z d z2
µ ¶
1 d dP 1 d 2Φ
ρ + + k 2 = 0. (2.193)
Pρ dρ dρ Φρ 2 d ϕ2
Note that when writing the above equation in term of exponentials, we are
implicitly assuming that k 2 > 0. If not, we can always define k 2 ≡ −λ2 , with
λ > 0, and the exponentials in Eq. (2.194) would produce a solution in terms
on sin(λz) and cos(λz).
2.5.1 C YLINDRICAL COORDINATES 121
1 d 2Φ
= −m 2 , (2.196)
Φ dϕ 2
µ ¶
d dP
ρ ρ + (k 2 ρ 2 − m 2 )P = 0. (2.197)
dρ dρ
The equation in the azimuthal angle ϕ has the familiar solution
For physical situations defined for the entire range of ϕ, (ρ, ϕ, z) and (ρ, ϕ +
2π, z) represent the same physical point. Then u, and thus, Φ, must be single-
valued and so not change when ϕ increases by 2π. This implies that m must
be an integer. In this way, cos(mϕ + m2π) = cos(mϕ) and sin(mϕ + m2π) =
sin(mϕ). Negative values of m will not give rise to any new solutions, so they
are not included in the range of m. In the particular case m = 0, Eq. (2.196)
produces as solution
Φ(ϕ) = C 0 ϕ + D 0 . (2.199)
(2.207)
2.5.1 C YLINDRICAL COORDINATES 123
which is a convergent series: the ratio test for the convergence of series yields
¯ ¯
¯ a 2(n+1) v 2(n+1) ¯
lim ¯ ¯ ∼ lim 1 v 2 = 0, (2.208)
n→∞ ¯ a 2n v 2n ¯ n→∞ n 2
The factor m!2m is a constant factor which can be reabsorbed in the arbitrary
constant multiplying the solution P 1 (v) when constructing the general solu-
tion P (v). In this way, we can consider as solution related to the root σ1
Note that contrary to P 1 (v), P 2 (v) is not well behaved at v = 0 due to the pres-
ence of u in the denominator of the integrand. Although the above procedure
produces a second solution for the Bessel equation, it is not the customary
procedure. It turns out to be more common to use as second solution the
combination
J m (v)cos(mπ) − J −m (v)
Ym (v) = , (2.213)
sin(mπ)
which is called the Bessel function of the second kind or Neumann function
(more details related to the Bessel functions will be studied in Chapter 3). In
this way, we have obtained that
or, in terms of ρ,
Figure 2.4: A conducting cylindrical can whose top has a potential given by
V (ρ, θ) with the rest of the surface grounded.
given by Eq. (2.216). Since the bottom face of the can is grounded, u(ρ, ϕ, 0) =
0 for arbitrary ρ and ϕ. This means that
yielding to
Za µ ¶
4 x 0n
Cn = µ ¶ d ρ ρV (ρ)J 0 ρ , (2.227)
a
a 2 J 12 (x 0n )sinh x0na h 0
2.5.2 S PHERICAL COORDINATES 127
To calculate this integral, we need to use the following property of the Bessel
functions
Z
d x x m J m−1 (x) = x m J m (x), (2.229)
and we get
4V0
Cn = µ ¶. (2.231)
x 0n h
x 0n J 1 (x 0n )sinh a
Therefore
µ ¶ µ ¶
J 0 xa0n ρ x 0n
sinh a z
X∞
u(ρ, z) = 4V0 µ ¶ (2.232)
n=1 x 0n
x 0n J 1 (x 0n )sinh a h
µ ¶
1 d dΘ 1 d 2Φ
sinθ + = −λ. (2.237)
Θsinθ d θ dθ Φsin2 θ d ϕ2
Equation (2.236) is a homogeneous equation,
d 2R dR
r2 2
+ 2r − λR = 0, (2.238)
dr dr
which can be reduced, by the substitution r = e t , and writing R(r ) = S(t ), to
d 2S d S
+ − λS = 0. (2.239)
dt2 dt
This has as solution
S(t ) = Ae λ1 t + B e λ2 t . (2.240)
λ21 + λ1 − λ = 0,
λ22 + λ2 − λ = 0. (2.242)
R(r ) = Ar λ1 + Br λ2 , (2.246)
λ = l (l + 1). (2.247)
Now, considering Eq. (2.237), multiplying through by sin2 θ and using Eq. (2.247),
it too takes a separated form:
· µ ¶ ¸
sinθ d dΘ 2 1 d 2Φ
sinθ + l (l + 1)sin θ + = 0. (2.249)
Θ dθ dθ Φ d ϕ2
1 d 2Φ
= −m 2 (2.250)
Φ d ϕ2
dµ d dµ d d
= −sinθ = −(1 − µ2 )1/2 =⇒ = = −(1 − µ2 )1/2 , (2.253)
dθ dθ dθ dµ dµ
d 2M 2µ d M l (l + 1)
− + M = 0. (2.257)
d µ2 1 − µ2 d µ 1 − µ2
In this way, multiplying Eq. (2.257) by the factor (1 − µ2 ) we get the recurrence
relation
∞ ·
X
¸
(n + 2)(n + 1)a n+2 (1 − µ ) − 2µ(n + 1)a n+1 + l (l + 1)a n µn = 0, (2.259)
2
n=0
In this way
(n − 2)(n − 1) − l (l + 1)
an = a n−2
n(n − 1)
(l − n + 2)(l + n − 1)
=− a n−2 , n = 2, 3, . . . , (2.261)
n(n − 1)
Considering a 0 = 1 and a 1 = 0, all a n coefficients with odd values of n are zero,
while for the even values of n we get
l (l + 1) l (l + 1)
a2 = − a0 = − ,
2 2
(l − 2)(l + 3) l (l + 1)(l − 2)(l + 3) l (l − 2)(l + 3)(l + 1)
a4 = − a2 = = ,
4·3 4! 4!
(l − 4)(l + 5) l (l + 1)(l − 2)(l + 3)(l − 4)(l + 5) l (l − 2)(l − 4)(l + 5)(l + 3)(l + 1)
a6 = − a4 = − = ,
6·5 6! 6!
..
.
(−1)i l !! (l + 2i − 1)!!
a 2i = , i = 0, 1, 2, . . . (2.262)
(2i )! (l − 2i )!! (l − 1)!!
where we have introduced the double factorial, which is given by j !! = j ( j −
2)( j −4)( j −6) · · · . In this way, we have found one solution for Eq. (2.257), which
is
X
∞ (−1)n l !! (l + 2n − 1)!! (2n)
M 1 (µ) = µ . (2.263)
n=0 (2n)! (l − 2n)!! (l − 1)!!
By applying the ratio test to these series,
¯ ¯
¯ a 2(n+1) µ2(n+1) ¯
lim ¯ ¯ = µ2 , (2.264)
n→∞ ¯ a 2n µ2n ¯
which converges for |µ| < 1, and so the radius of convergence is unity, which,
as expected, is the distance to the nearest singular point of Eq. (2.257).
Similarly, if we choose a 0 = 0 and a 1 = 1, all coefficients a n where n is an
even number are zero, while for the case of n being an odd number we obtain
(l − 1)(l + 2) (l − 1)(l + 2)
a3 = − a1 = − ,
3·2 3!
(l − 3)(l + 4) (l − 1)(l − 3)(l + 4)(l + 2)
a5 = − a3 = ,
5·4 5!
(l − 5)(l + 6) (l − 1)(l − 3)(l − 5)(l + 6)(l + 4)(l + 2)
a7 = − a5 = − ,
7·6 7!
..
.
(−1)i (l − 1)!! (l + 2i )!!
a 2i +1 = , i = 0, 1, 2, . . . (2.265)
(2i + 1)! (l − 2i − 1)!! l !!
2.5.2 S PHERICAL COORDINATES 132
In this way, we have a second solution to Eq. (2.257), which is linearly inde-
pendent to the M 1 (µ),
X∞ (−1)n (l − 1)!! (l + 2n)!! 2n+1
M 2 (µ) = µ . (2.266)
n=0 (2n + 1)! (l − 2n − 1)!! l !!
As in case of M 1 (µ), the above series is convergent for |µ| < 1. Hence, the
general solution to Eq. (2.257) is given by
l (l + 1) − l (l + 1)
a l +2 = a l = 0, (2.268)
(l + 1)(l + 2)
i.e., the series terminates and we obtain a polynomial solution of order l ,
which is finite for any value of µ. In particular, if l is even, then M 1 (µ) reduces
to a polynomial, whereas if l is odd the same is true of M 2 (µ). In each case,
the other series does not terminate and therefore converges only for |µ| < 1.
These solutions (suitably normalized) are called the Legendre polynomials
of order l ; they are written as P l (µ) and it is conventional to normalize them
such that P l (1) = 1.
According to whether l is even or odd, we define the Legendre functions
of the second kind as Q l (µ) = αl M 2 (µ) or Q l (µ) = βl M 1 (µ), where
or, equivalently,
where we have absorb the constant C in the other constants. As before, a gen-
eral solution may be obtained by superposing solutions of this form for the
allowed values of the separation constants (l in case of axial symmetry). As
mentioned above, if the solution is required to be finite on the polar axis, then
F = 0 for all values of the separation constants. Although we have not solved
Eq. (2.257) for general values of m, when m 6= 0, one simply replaces P l (cosθ)
and Q l (cosθ) by the associated Legendre functions P lm (cosθ) and Q lm (cosθ),
which are given by
d |m| d |m|
P lm (x) = (1 − x 2 )|m|/2 P l (x), Q lm (x) = (1 − x 2 )|m|/2 Q l (x) (2.274)
d x |m| d x |m|
where 0 ≤ |m| ≤ l and P l0 (x) = P l (x), Q l0 (x) = Q l (x). In this way, the general
solution of the Laplace’s equation in spherical coordinates is given by
X
u(r, θ, ϕ) = [A l m r l + B l m r −(l +1) ][C l m cos(mϕ) + D l m sin(mϕ)]
l ,m
× [E l m P lm (cosθ) + F l m Q lm (cosθ)] (2.275)
or, equivalently,
½
−T0 , if − 1 ≤ µ < 0,
T (a, µ) = (2.279)
T0 , if 0 < µ ≤ 1.
Z1
2
d µP n (µ)P m (µ) = δnm , n, m = 0, 1, 2, . . . . (2.281)
2n + 1
−1
2.5.2 S PHERICAL COORDINATES 135
Z1
1 2n + 1
An = n d µP n (µ)T (a, µ)
a 2
−1
· Z0 Z1 ¸
1 2n + 1
= n d µ(−T0 )P n (µ) + d µ(T0 )P n (µ)
a 2
−1 0
· Z0 Z1 ¸
2n + 1
= T0 − d µP n (µ) + d µP n (µ) . (2.282)
2a n
−1 0
Next, the first integral can be related to the second one by using the following
property of the Legendre polynomials
Z0 Z0 Z1 Z1
d µP n (µ) = (−d µ)P n (−µ) = d µP n (−µ) = (−1)n d µP n (µ). (2.284)
−1 +1 0 0
Z1 ¸
2n + 1 n
An = T0 [1 − (−1) ] d µP n (µ)
2a n
0
2n + 1 1
0, if n is even,
= T0 R (2.285)
2a n 2 d µP 2k+1 (µ), if n = 2k + 1,
0
Z1
(−1)k (2k)!
d µP 2k+1 (µ) = . (2.286)
22k+1 k!(k + 1)!
0
2.5.2 S PHERICAL COORDINATES 136
A 2n = 0,
Z1
2(2n + 1) + 1
A 2n+1 = 2 T0 d µP 2n+1
2a 2n+1
0
n
(−1) (4n + 3)(2n)!
= T0 , (2.287)
22n+1 n!(n + 1)!a 2n+1
with n = 0, 1, 2, . . . . Substituting these expressions in Eq. (2.277), we arrive to
the final answer
∞ (−1)n (4n + 3)(2n)! µ r ¶2n+1
X
T (r, θ) = T0 2n+1 n!(n + 1)!
P 2n+1 (cosθ). (2.288)
n=0 2 a
Note that, in general, if the temperature on the surface of the sphere had
been given as a function of θ and ϕ, then we would have P lm (cosθ) instead
of P l (cosθ) and we would have had to consider a double series summed over
l and m, since the solution would not have been axially symmetric.
of Eq. (2.276), the terms 1/r (l +1) can be neglected when r → ∞ and to satisfy
V (r → ∞, θ) = 0, we need that A l = 0 for all l . So we can write the potential as
X∞ B
n
V (r, θ) = P (cosθ).
n+1 n
(2.289)
n=0 r
∂u
∇2 u =
κ∇ . (2.292)
∂t
This equation describes the temperature u(rr , t ) in a region containing no heat
sources or sinks; κ > 0 is a real constant characterizing the medium in which
heat is flowing. The separation of variables u(rr , t ) = T (t )R(rr ) yields
∂ dT
∇2 [T (t )R(rr )] =⇒ R(rr )
[T (t )R(rr )] = κ∇ ∇2 R.
= κT (t )∇ (2.293)
∂t dt
Dividing both sides by T (t )R(rr ), we obtain
1 dT 1
= κ ∇ 2 R ≡ −κλ. (2.294)
T dt R
2.6.1 H EAT- CONDUCTING ROD 138
Zb µ ¶
2 mπ
Bm = d x f (x)sin x . (2.304)
b b
0
Thus, if we know the initial temperature distribution on the rod, i.e., the func-
tion f (x), we can determine the temperature distribution of the rod for all
time. For instance, if the initial temperature distribution of the rod is uniform,
say u 0 , then
Zb µ ¶ · µ ¶¸¯b
2u 0 mπ 2u 0 b mπ ¯
Bm = d x sin x = − cos x ¯¯
b b b mπ b 0
0
· ¸
2u 0
= 1 − (−1)m . (2.305)
mπ
It follows that the odd m’s survive, and if we set m = 2n + 1, we obtain
4u 0
B 2n+1 = (2.306)
π(2n + 1)
and
∞ e −[(2n+1)πκ/b]2 t · ¸
4u 0 X (2n + 1)π
u(t , x) = sin x . (2.307)
π n=0 2n + 1 b
This distribution of temperature for all time can be obtained numerically for
any heat conductor whose κ is known. Note that the exponential in the sum
causes the temperature to drop to zero (the fixed temperature of its two end
points) eventually. This conclusion is independent of the initial temperature
distribution of the rod, as Eq. (2.302) indicates.
2.6.2 H EAT CONDUCTION IN A RECTANGULAR PLATE 140
∂2 R ∂2 R
+ + λR = 0. (2.308)
∂x 2 ∂y 2
A separation of variables, R(x, y) = X (x)Y (y), and its usual procedure leads to
the following equation:
1 d2X 1 d 2Y
+ + λ = 0. (2.309)
X d x2 Y d y 2
1 d2X 1 d 2Y
≡ −µ, ≡ −ν, (2.310)
X d x2 Y d y2
d2X d 2Y
+ µX = 0, + νY = 0, λ = µ + ν. (2.311)
d x2 d y2
Z x 0 +L µ ¶ µ ¶ L for r = p = 0,
πr x πpx L
d xcos cos = for r = p > 0, (2.318)
x0 L L 2
0 for r 6= p,
Z x 0 +L µ ¶ µ ¶ 0 for r = p = 0,
πr x πpx L
d xsin
L
sin
L
=
2 for r = p > 0, (2.319)
x0
0 for r =
6 p,
where r and p are integers greater or equal to zero and x 0 is arbitrary (but
finite), we can find the coefficients A mn :
Za Zb µ ¶ µ ¶
4 nπ mπ
A mn = dx d y f (x, y)sin x sin y . (2.320)
ab a b
0 0
µ ¶
d 2P 1 d P m2
+ + λ − 2 P = 0. (2.323)
d ρ2 ρ d ρ ρ
The solution of the last (Bessel) equation, which is well defined for ρ = 0 and
vanishes at ρ = a is, as we already saw in Sec. 2.4
µ ¶
x mn p x mn
P (ρ) = C J m ρ , with λ= and n = 1, 2, . . . , (2.324)
a a
Multiplying the three solutions and summing over the two indices yields the
most general solution
µ ¶ · ¸
X
∞ X
∞ x mn −κ(x mn /a)2 t
u(ρ, ϕ, t ) = Jm ρ e A mn cos(mϕ) + B mn sin(mϕ) .
m=0 n=1 a
(2.326)
which is basically identical to Eq. (2.224). Therefore, the coefficients are given
by expressions similar to Eq. (2.225). In the case at hand, we get
Z2π Za µ ¶
2 x mn
A mn = d ϕ d ρρ f (ρ, ϕ)J m ρ cos(mϕ),
πa 2 J m+1
2
(x mn ) a
0 0
Z2π Za µ ¶
2 x mn
B mn = d ϕ d ρρ f (ρ, ϕ)J m ρ sin(mϕ). (2.328)
πa 2 J m+1
2
(x mn ) a
0 0
2.7 T HE S CHRÖDINGER EQUATION 143
Za µ ¶
4 x 0n
A n ≡ A 0n = d ρ ρ J0 ρ , (2.330)
a 2 J 12 (x 0n ) a
0
ħ2 2 ∂u
− ∇ u + V (rr )u = i ħ , (2.331)
2µ ∂t
ħ2 2 dT
−T ∇ R + V (rr )(RT ) = i Rħ . (2.333)
2µ dt
1 ħ2 2 1 dT
− ∇ R + V (rr ) = i ħ . (2.334)
R 2µ T dt
The left-hand side is a function of position alone, and the right-hand side is
a function of time alone, and since r and t are independent variables, as we
2.7.1 Q UANTUM PARTICLE IN A BOX 144
have already discussed in previous sections, the only way that Eq. (2.334) can
hold is for both sides to be constant, say E :
1 ħ2 2 ħ 2
− ∇ R + V (rr ) = E =⇒ − ∇ R + V (rr )R = E R (2.335)
R 2µ 2µ
and
1 dT dT iE
iħ =E =⇒ = − dt. (2.336)
T dt T ħ
The solution of the time part is easily obtained, since it can be integrated di-
rectly:
d2X d 2Y d2Z
+ λX = 0, + σY = 0, + νZ = 0, (2.340)
d x2 d y2 d z2
with
Since we cannot find the particle outside the box, we impose the boundary
conditions
R(0, y, z) = R(a, y, z) = 0 =⇒ X (0) = 0 = X (a),
R(x, 0, z) = R(x, b, z) = 0 =⇒ Y (0) = 0 = Y (b),
R(x, y, 0) = R(x, y, c) = 0 =⇒ Z (0) = 0 = Z (c), (2.342)
in this way, |u(x, y, z, t )|2 d xd yd z, which is related to the probability density
of finding the particle between (x, y, z) and (x + d x, y + d y, z + d z) at some in-
stant t , would be zero outside the box (this is not true for a particle inside a
finite potential well, in which case the particle has a nonzero probability of
tunneling out of the well). From Sec. 2.3.1, the general solution of each of the
equations in (2.340), together with the about boundary conditions, lead to the
following solutions:
µ ¶ µ ¶
nπ nπ 2
X n (x) = A n sin x , λn = , for n = 1, 2, . . . , (2.343)
a a
µ ¶ µ ¶
mπ mπ 2
Ym (y) = B m sin y , σm = , for m = 1, 2, . . . , (2.344)
b b
µ ¶ µ ¶2
lπ lπ
Zl (z) = C l sin z , νl = , for l = 1, 2, . . . , (2.345)
c c
where the multiplicative constants have been suppressed. Using the values
found for λn , σm and νl together with Eqs. (2.337) and (2.341), the time solu-
tion has the form
·µ ¶ µ ¶ µ ¶2 ¸
−i E nml t /ħ ħ2 nπ 2 mπ 2 lπ
Tnl m (t ) = D nl m e , where E nml = + + .
2µ a b c
(2.346)
The solution of the Schrödinger equation that is consistent with the boundary
conditions is, then,
µ ¶ µ ¶ µ ¶
X
∞
−i E nml t /ħ nπ mπ lπ
u(x, y, z, t ) = A nml e sin x sin y sin z . (2.347)
n,m,l =1 a b c
The constants A l mn ≡ A n B m C l D nml are determined by the initial shape u(x, y, z, 0).
In fact, setting t = 0, multiplying by the product of the three sine functions in
the three variables, using the orthogonality properties of the trigonometric
functions, and integrating over appropriate intervals for each coordinate, we
obtain
Za Zb Zc µ ¶ µ ¶ µ ¶
8 nπ mπ lπ
A nml = d x d y d zu(x, y, z, 0)sin x sin y sin z .
abc a b c
0 0 0
(2.348)
2.7.1 Q UANTUM PARTICLE IN A BOX 146
As we mentioned earlier,
µ ¶
ħ2 π2 n 2 m 2 l 2
E nl m = + + (2.349)
2µ a 2 b 2 c 2
represents the energy of the particle which depends on three positive inte-
gers (n, l , m). Each set of three positive integers (n, l , m) represents a quantum
state of the particle. For a cube, a = b = c, the energy of the particle is
ħ2 π2 2 2 2 ħ2 π2
E nl m = (n + m + l ) = (n 2 + m 2 + l 2 ), (2.350)
2µa 2 2µV 2/3
where V = a 3 is the volume of the box. The ground state, i.e., the one with low-
est energy, is (1, 1, 1), has energy 3ħ2 π2 /(2µV 2/3 ), and is nondegenerate (only
one state corresponds to this energy). However, the higher-level states are de-
generate. For instance, the three distinct states (1, 1, 2), (1, 2, 1), and (2, 1, 1)
all correspond to the same energy, 6ħ2 π2 /(2µV 2/3 ). The degeneracy increases
rapidly with larger values of n, m and l .
Note that Eq. (2.350) can be written as
Thus the density of states, i.e., the number of states per unit volume, is then
µ ¶
N π 2µ 3/2 3/2
N = = E nml . (2.353)
V 6 ħ2 π2
3
O RTHOGONAL POLYNOMIALS
AND SPECIAL FUNCTIONS
and the coefficients f k were obtained from the corresponding inner product
of L ω2 (a, b), i.e.,
Zb
f k = 〈e k | f 〉 = d xω(x)e k∗ (x) f (x). (3.2)
a
In turns out that sine and cosine functions are not the only family of func-
tions that are orthogonal and complete. In this chapter we introduce a class of
orthogonal polynomials, the so-called classical ones, which are of particular
importance in physical applications and permits also the expansion of f (x)
in terms of e k (x). This includes, for example, the Legendre polynomials that
we introduced in Chapter 2. A way of obtaining these orthogonal polynomials
is to start with the monomials 1, x, x 2 , . . . , x n , which are not orthogonal, but
form a base in L ω2 (a, b) according with the Stone-Weierstrass theorem, and
apply the Gram-Schmidt orthogonalization process to these monomials, i.e.,
X 〈x k |e
k−1
j〉
e k (x) = x k − e j (x), k = 1, 2, . . . , (3.3)
j =0 〈e j |e j 〉
149
3.1 G ENERALIZED R ODRIGUES FORMULA 150
where
Zb
〈p k |F n 〉 = d xω(x)p k (x)F n (x) = 0, for k < n. (3.6)
a
Since F n (x) and p k (x) are real, from now onwards, we omit the complex con-
jugation involved in their inner product.
L ET ’ S PROVE IT ! Before starting with the formal proof of the above statements,
first, we need two realize that
Thus,
· ¸
d n dω n n −1 d s d p (≤k)
[ω(x)s (x)p (≤k) ] = s (x)p (≤k) + nω(x)s (x) s p (≤k) +
dx dx dx dx
·½ ¾ ¸
n−1 d s d p (≤k)
= ω(x)s F 1 (x) + (n − 1) p (≤k) + s .
dx dx
(3.9)
d
[ω(x)s n (x)p (≤k) ] = ωs n−1 p (≤k+1) , (3.10)
dx
where
½ ¾
ds d p (≤k)
p (≤k+1) ≡ F 1 (x) + (n − 1) p (≤k) + s . (3.11)
dx dx
m n
(ii) All the derivatives d [ω(x)s
d xm
(x)]
with m < n vanish at x = a and x = b.
Indeed, from Eq. (3.7), putting k = 0 and p (≤0) ≡ p 0 = 1, we get
d m [ω(x)s n (x)]
= ω(x)s n−m (x)p (≤m)
d xm
= [ω(x)s(x)]s n−m−1 (x)p (≤m) . (3.12)
Since ω(a)s(a) = ω(b)s(b) = 0, the right hand side in Eq. (3.12) vanishes
at x = a and x = b when n > m. In the case of an infinite interval, it can
be shown that ω(x)s(x) vanishes at infinity faster than any polynomial.
Let us now first proof the orthogonality condition in Eq. (3.6). The proof
3.1 G ENERALIZED R ODRIGUES FORMULA 152
This shows that each integration by parts transfers one differentiation from
ω(x)s n (x) to pk(x) and introduces a minus sign. Thus, after k integrations by
parts, we get
Z Z
k d k p k d n−k [ω(x)s n (x)]
d xω(x)p k (x)F n (x) = (−1) dx . (3.14)
d xk d x n−k
Since the kth derivative of a polynomial of degree k is a constant, we can take
d k p k /d x k outside the integration, i.e.,
Z k Z
k d pk d n−k [ω(x)s n (x)]
d xω(x)p k (x)F n (x) = (−1) d x
d xk d x n−k
k Z · n−k−1 ¸
k d pk d d [ω(x)s n (x)]
= (−1) dx
d xk dx d x n−k−1
· ¸¯
k
k d pk d
n−k−1
[ω(x)s n (x)] ¯¯x=b
= (−1) ¯ = 0. (3.15)
d xk d x n−k−1 x=a
| {z }
considering property (ii)
Note that n − k − 1 ≥ 0 because k < n, so that the last line of the equation is
well-defined.
Next, we need to proof that F n (x) is a polynomial of degree precisely equal
to n. Let’s see this. First, from Eq. (3.7), putting k = 0, m = n and p (≤0) = p 0 = 1,
we have
d n [ω(x)s n (x)] 1 d n [ω(x)s n (x)]
= ω(x)p (≤n) , or F n (x) = = p (≤n) . (3.16)
d xn ω(x) d xn
Then, we can write
In this way,
Zb
〈F n | f n 〉 = d xω(x)[F n (x)]2
a
Zb Zb
= d xω(x)p (≤n−1) F n (x) + k n(n) d xω(x)x n F n (x). (3.18)
a a
The left-hand side of Eq. (3.18) is a positive quantity because both ω(x) and
[F n (x)]2 are positive, and the first integral on the right-hand side vanishes
from Eq. (3.15), since k ≤ n − 1. Therefore, the second term on the right-hand
side in Eq. (3.18) cannot be zero. In particular, k n(n) 6= 0, and F n (x) is of degree
n.
where λn = K 1 k 1(1) n + σ2 n(n − 1), with k 1(1) being the coefficient of x in F 1 (x)
and σ2 the coefficient of x 2 in s(x).
R EALLY ? In case you are curious to know who Eq. (3.20) arises, here we pro-
vide a proof. Since F n (x) is a polynomial which in the interval [a, b] is orthog-
onal with weight ω(x) to any polynomial of degree n, i.e.,
Zb
d xω(x)F n (x)p (<n) = 0, (3.21)
a
where λin are some numbers. Multiplying both sides of Eq. (3.23) by F m and
integrating, we get, using the orthogonality property of F m
Zb · ¸
d d Fn
d xF m s(x)ω(x) = −λ(m)
n hm , (3.24)
dx dx
a
with
Zb
2
hm = d xω(x)F m (x). (3.25)
a
Writting,
· ¸ · ¸
d d Fn d d Fn d Fm d Fn
F m (x) s(x)ω(x) = F m (x)s(x)ω(x) − s(x)ω(x) ,
dx dx dx dx dx dx
(3.26)
and using that s(x)ω(x) vanishes at the ends of the integration interval, the
left-hand side of Eq. (3.24) yields for m < n,
Zb · ¸ Zb
d d Fn d Fn d Fm
d xF m s(x)ω(x) = − d xω(x)s(x)
dx dx dx dx
a a
Zb µ · ¸¶
1 d d Fm
= d xω(x)F n (x) s(x)ω(x) = 0,
ω(x) d x dx
a
(3.27)
where we have used that F n (x) is orthogonal to any polynomial of degree < n.
Comparing with Eq. (3.24), we arrive to λ(m) (n)
n = 0 for m < n. Defining λn ≡ λn
for simplicity, we can write Eq. (3.23) in the form
· ¸
d d Fn
s(x)ω(x) = −ω(x)λn F n (x). (3.28)
dx dx
3.2 C LASSIFICATION 155
Zb · ¸ Zb · ¸
d Fn d {s(x)ω(x)} d F n d 2 Fn
d xF n (x) s(x)ω(x) = d xF n (x) + s(x)ω(x)
dx dx dx d x2
a a
Zb · ¸
d Fn d 2 Fn
= d xω(x)F n (x) K 1 F 1 (x) + s(x) ,
dx d x2
a
(3.29)
where we have used Eq. (3.19) for n = 1. Because of the orthogonality property
of the polynomials F n , only the nth power of x in the nth degree polynomial
in the square brackets contributes to the integral. If k n(n) is the coefficient of
x n in F n (x), the coefficient of x n in F 1 (x) ddFxn is given by nk 1(1) k n(n) . Similarly,
if σ2 is the coefficient of x 2 in s(x), then the coefficient of x n in sd 2 F n /d x 2 is
given by σ2 n(n − 1)k n(n) . In this way, we get
Zb · ¸ · ¸ Zb
d Fn (1)
d xF n (x) s(x)ω(x) = K 1 nk 1 + σ2 n(n − 1) d xω(x)F n (x)k n(n) x n
dx
a a
· ¸
= K 1 nk 1(1) + σ2 n(n − 1) h n . (3.30)
Comparing this latter equation with Eq. (3.24), and since the only non-zero
value for λ(m)
n is when m = n, we get
(1)
λn ≡ λ(n)
n = K 1 k 1 n + σ2 n(n − 1). (3.31)
3.2 C LASSIFICATION
Let us now investigate the consequence of various choices of s(x). We start
with F 1 (x), which, according to Eq. (3.19)
For nonzero B , the only way that this equality can hold is for α to be negative
and for a 2 and b 2 to be infinite. Since a < b, we must take a = −∞ and b = +∞.
In this way, Eq. (3.35) can be written as
2 2
ω(x)s = B e −|α|x +βx
= B e −(|α|x −βx)
. (3.37)
so we have
½ · µ ¶¸2 ¾
β2 p β
ω(x)s = B e 4|α| exp − |α| x − (3.39)
2|α|
where γ ≡ K 1 k 1(1) /σ1 , ρ ≡ K 1 k 1(0) /σ1 − K 1 k 1(1) σ0 /σ21 , and B is A modified by the
constant of integration. Using Eq. (3.5),
which give a = −σ0 /σ1 , ρ > 0, γ < 0, i.e., γ = −|γ| < 0, and b = +∞. We now
redefine the variable x → |γ|(σ0 +σ1 x)/σ1 , such that, in terms of the new vari-
able Eq. (3.40) can be written as
µ ¶
x xσ1 ρ −x+|γ|σ0 /σ1
ω(x) σ1 = B e
|γ| |γ|
µ ¶
xσ1 ρ |γ|σ0 /σ1 −x
=B e e . (3.42)
|γ|
In this way
µ ¶ρ−1
|γ|σ0 /σ1 σ1
ω(x)x = B e x ρ e −x ≡ B̃ x ρ e −x , (3.43)
|γ|
³ ´ρ−1
with B̃ = B e |γ|σ0 /σ1 σ 1
|γ|
an arbitrary constant which can be reabsorbed in
the normalization constant K n . Alternatively, since we will use a convention
to define K n , we can simply redefine B̃ = 1 and write
ω(x)x = x ρ e −x , (3.44)
3.2 C LASSIFICATION 158
such that
ω(x) = x ν e −x , (3.46)
where ν > −1 since ρ > 0. Note that after redefining the variable x, the limit
a = −σ0 /σ1 becomes 0, while b = ∞ remains unaltered. In this way, we have
found
Using now Eqs. (3.19) and (3.47), we can get F n (x). The resulting polynomials
are called Laguerre polynomials and are denoted by L νn (x).
Similarly, we can obtain the weight and the interval of integration for the
case when s(x) is of degree 2. The result, with the corresponding redefini-
tion of variables and parameters, is ω(x) = (1 + x)µ (1 − x)ν , with µ, ν > −1,
s(x) = 1−x 2 , a = −1 and b = +1. The polynomials found are called Jacobi poly-
µ,ν
nomials are are denoted by P n (x). The Jacobi polynomials are themselves
divided into other subcategories depending on the values of µ and ν and for
historical reason, and because they play important roles in applications, they
are named differently. For example, in case of µ = ν = 0, we have ω(x) = 1, and
the obtained polynomials, represented as P n (x), are called Legendre polyno-
mials; For µ = ν = ∓1/2, we have ω(x) = (1 − x 2 )∓1 and we have the Chebyshev
polynomials of the first and second kind, respectively. Strictly speaking, the
definition of each of the preceding polynomials contains also the specifica-
tion of the normalization constant K n in the generalized Rodrigues formula,
something which is called standardization; this standardization will be spec-
ified later on, but you should be aware that different books can use different
standardizations for the same polynomials.
Z1 Z1 ¯
1 2 ¯¯1
0= d xP 1 (x)P 0 (x) = d x(ax + b) = ax ¯ + 2b = 2b. (3.48)
2 −1
−1 −1
3.3 R ECURRENCE RELATIONS 159
So one of the coefficients, b, is zero. To find the other one, we need some
standardization procedure. We standardize P n (x) by requiring that P n (1) = 1
for any value n. For n = 1 this yields a · 1 = 1, or a = 1, so that P 1 (x) = x.
We can calculate P 2 (x) similarly: write P 2 (x) = ax 2 + bx + c, impose the
condition that it would be orthogonal to both P 1 (x) and P 0 (x), and enforce
the standardization procedure. All this will yield
Z1 Z1
2 2
0= d xP 2 (x)P 0 (x) = a + 2c, 0= d xP 2 (x)P 1 (x) = b, (3.49)
3 3
−1 −1
Taking the inner product of both sides of this equation with F m (x), we get
Zb (n+1) Zb
k n+1
d xω(x)F n+1 (x)F m (x) − d xω(x)xF n (x)F m (x)
a
k n(n) a
X
n Zb
= aj d xω(x)F j (x)F m (x). (3.53)
j =0
a
3.3 R ECURRENCE RELATIONS 160
Using the orthogonality relation (3.21), the first integral on the left-hand side
vanishes as long as m ≤ n; the second integral vanishes if m ≤ n − 2, since in
this case xF m (x) is a polynomial of degree n − 1. In this way, we have
X
n Zb
aj d xω(x)F j (x)F m (x) = 0 for m ≤ n − 2, (3.54)
j =0
a
but, by Eq. (3.21), the integral in the sum is zero unless j = m. Therefore, the
sum reduces to
Zb
am d xω(x)[F m (x)]2 = 0 for m ≤ n − 2. (3.55)
a
a m = 0, for m = 0, 1, . . . , n − 2, (3.56)
This is the recurrence formula we are looking for; there remains to find the
constants a n and a n−1 . First, if we multiply Eq. (3.57) by ω(x)F n−1 (x) and in-
tegrate, using the orthogonality property (3.21), the integral of the first term
on the left-hand side as well as the integral of the first term on the right-hand
side are both zero, thus, we have
· (n+1) ¸ Zb
k
− n+1 d xω(x)xF n (x) = h n−1 a n−1 . (3.58)
k n(n) a
where h n−1 is given by Eq. (3.24). Now, notice that due to the orthogonality
relation (3.21)
Zb Zb Zb
2
hn = d xω(x)[F n (x)] = d xω(x)F n (x)F n (x) = d xω(x)F n (x)[k n(n) x n ].
a a a
(3.59)
3.3 R ECURRENCE RELATIONS 161
(n+1) Zb
k n+1
h n−1 a n−1 = − d xω(x)xF n (x)F n−1 (x)
k n(n) a
(n+1) (n) Zb
k n+1 k n−1
=− d xω(x)F n (x)k n(n) x n
k n(n) k n(n) a
k (n+1) (n)
k n−1
= − n+1 hn . (3.60)
k n(n) k n(n)
In this way,
(n+1) (n)
h n k n+1 k n−1
a (n−1) = − . (3.61)
h n−1 k n(n) k n(n)
Other recurrence relations can be obtained from Eq. (3.50). For example,
if we differentiate both sides of Eq. (3.50) and use Eq. (3.20), we get
· ¸
d Fn d [ω(x)s(x)]
2ω(x)s(x)αn + αn + ω(x)λn (αn x + βn ) F n
dx dx
− ω(x)λn+1 F n+1 (x) + ω(x)γn λn−1 F n−1 (x) = 0. (3.64)
d ω(x) dω
A n (x)F n (x) − λn+1 (αn x + βn ) F n+1 (x) + γn λn−1 (αn x + βn ) F n−1 (x)
dx dx
d F n+1 d F n−1
+ B n (x) + γn D n (x) = 0, (3.67)
dx dx
where
· ¸
d 2 [ω(x)s(x)] dω
A n (x) = (αn x + βn ) 2ω(x)αn λn + αn + λn (αn x + βn )
d x2 dx
d [ω(x)s(x)]
− α2n ,
dx
d [ω(x)s(x)]
B n (x) = αn − ω(x)(αn x + βn )(λn+1 − λn ),
dx
d [ω(x)s(x)]
D n (x) = ω(x)(αn x + βn )(λn−1 − λn ) − αn . (3.68)
dx
All these recurrence relations seem to be very complicated. However, com-
plexity is the price we pay for generality! When we work with specific orthog-
onal polynomials, the equation simplify considerably. For instance, for Her-
mite and Legendre polynomials, Eq. (3.65),
d Hn d Pn
= 2nHn−1 (x), and (1 − x 2 ) + nxP n (x) − nP n−1 (x) = 0. (3.69)
dx dx
Also, applying Eq. (3.66) to Legendre polynomials gives
d P n+1 d Pn
−x − (n + 1)P n (x) = 0, (3.70)
dx dx
and Eq. (3.67) yields
d P n+1 d P n−1
− − (2n + 1)P n (x) = 0. (3.71)
dx dx
3.4 T HE CLASSICAL POLYNOMIALS 163
Zb Zb
2
hn = d xω(x)[F n (x)] = d x ω(x)[k n(n) x n + . . . ]F n (x)
a a
Zb Zb · ¸
1 d n [ω(x)s n (x)] k n(n) d d n−1 [ω(x)s n (x)]
= k n(n) d xω(x)x n
= d xxn
K n (x)ω(x) d xn Kn dx d x n−1
a a
¯ Zb
k n(n) n d n−1 [ω(x)s n (x)] ¯¯b k n(n) d (x n ) d n−1 [ω(x)s n (x)]
= x ¯ − K dx (3.72)
Kn d x n−1 a n dx d x n−1
a
The first term of the last line is zero by the property (ii) of Sec. 3.1. It is clear
that each integration by parts introduces a minus sign and shifts one differ-
entiation from ω(x)s n (x) to x n . Thus, after n integrations by parts and noting
0 n
d n [x n ]
that d [ω(x)s
dx 0
(x)]
= ω(x)s n
(x) and d x n = n!, we obtain
Zb
(−1)n k n(n) n!
hn = d xω(x)s n (x). (3.73)
Kn
a
n x2 d n −x 2
Hn (x) = (−1) e [e ]. (3.74)
d xn
2
It is clear that each time e −x is differentiated, a factor −2x is introduced. The
2
highest power of x is obtained when we differentiate e −x n times. This yields
2 2
(−1)n e x (−2x)n e −x = 2n x n , thus, k n(n) = 2n .
To obtain k n(n−1) , it is helpful to see whether the polynomial is even or odd.
Substituting −x for x in Eq. (3.74), we get Hn (−x) = (−1)n Hn (x), which shows
that if n is even (odd), Hn is an even (odd) polynomial, i.e., it can have only
even (odd) powers of x. In either case, the next-higher power of x in Hn (x) is
not n − 1 but n − 2. Thus, the coefficient of x n−1 is zero for Hn (x), and we have
p
k n(n−1) = 0. For h n , we use Eq. (3.73) to obtain h n = π2n n!.
Next we calculate the recurrence relation of Eq. (3.50). We can readily cal-
culate the constants needed: αn = 2, βn = 0, γn = −2n. Then substitute these
in Eq. (3.50) to obtain
d 2 Hn d Hn
2
− 2x + 2nHn (x) = 0. (3.76)
dx dx
1 d n [x ν e −x x n ] 1 −ν x d n [x n+ν e −x ]
L νn (x) = = x e . (3.77)
n!x ν e −x d xn n! d xn
To find k n(n) we note that differentiating e −x does not introduce any new pow-
ers of x, but only a factor of −1. Thus, the highest power of x is obtained by
leaving x n+ν alone and differentiating e −x n times. This gives
If ν is not an integer (and it need not be), the integral on the right-hand side
cannot be evaluated by elementary methods. In fact, this integral occurs so
frequently in mathematical applications that it is given a special name, the
gamma function. We are not going to discuss this function in this course and
at this point we simply note that
Z∞
Γ(z + 1) ≡ d x x z e −x , Γ(n + 1) = n!, n ∈ N, (3.82)
0
and write h n as
Γ(n + ν + 1) Γ(n + ν + 1)
hn = = . (3.83)
n! Γ(n + 1)
The relevant parameters for the recurrence relation can be easily calculated
using Eq. (3.63):
1 2n + ν + 1 n +ν
αn = − , βn = , γn = − . (3.84)
n +1 n +1 n +1
Substituting these in Eq. (3.50) and simplifying yields
(n + 1)L νn+1 (x) = (2n + ν + 1 − x)L νn (x) − (n + ν)L νn−1 . (3.85)
(−1)n d n [(1 − x 2 )n ]
P n (x) = . (3.87)
2n n! d xn
and take the nth derivative of the highest power of x. This yields
(−1)n d n [(−x 2 )n ] 1 d n [x 2n ]
k n(n) x n = =
2n n! d xn 2n n! d x n
1
= n 2n(2n − 1)(2n − 2) . . . (n + 1)x n . (3.89)
2 n!
If we multiply and divide the above expression by n!, and take a factor of 2 out
of all terms in the numerator, the even terms yield a factor of n! and the odd
terms give a Γ function, such that we can write
µ ¶
1
Γ n+2
(n) n
kn = 2 µ ¶ . (3.90)
1
n!Γ 2
The integral appearing on the right-hand side of the preceding equation can
be evaluated by repeated integration by parts to get
Z1 m Z1
2 n(n − 1) . . . (n − m + 1)
d x(1 − x 2 )n = d xx 2m (1 − x 2 )n−m
3 · 5 · 7 . . . (2m − 1)
−1 −1
µ ¶
2Γ 12 n!
= µ ¶. (3.92)
1
(2n + 1)Γ n + 2
In this way,
2
hn = . (3.93)
2n + 1
Next, we need αn , βn and γn for the recurrence relation. Using Eq. (3.63),
µ ¶ µ ¶
1
(n+1)
n+1
2 Γ n +1+ 2 n!Γ 12
k n+1 2n + 1
αn = (n) = µ ¶ = µ ¶= , (3.94)
kn 1 1 n +1
(n + 1)!Γ 2 2 Γ n+2
n
where we have used the property Γ(z + 1) = zΓ(z). We also have βn = 0, since
k n(n−1) = 0 = k n+1
(n)
and γn = −n/(n + 1). In this way, the recurrence relation is
given by
Now we use K 1 = −2, P 1 (x) = x, thus, k 1(1) = 1, and σ2 = −1, since s(x) =
1 − x 2 as we saw in Sec. 3.2, to obtain λn = −n(n + 1). Then, we obtain the
following differential equation from Eq. (3.20)
· ¸
d 2 d Pn
(1 − x ) = −n(n + 1)P n (x), (3.96)
dx dx
1) J ACOBI POLYNOMIALS
µ,ν
• Nomenclature: P n (x).
• Constants:
Γ(2n + µ + ν + 1) n(ν − µ)
k n(n) = 2−n , k n(n−1) = kn ,
n!Γ(n + µ + ν + 1) 2n + µ + ν
2µ+ν+1 Γ(n + µ + 1)Γ(n + ν + 1)
hn = . (3.98)
n!(2n + µ + ν + 1)Γ(n + µ + ν + 1)
• Rodrigues formula:
n · ¸
µ,ν (−1)n −µ −ν d µ+n ν+n
P n (x) = (1 + x) (1 − x) (1 + x) (1 − x) . (3.99)
2n n! d xn
• Differential equation:
µ,ν µ,ν
d 2Pn d Pn
(1 − x 2 ) 2
+ [µ − ν − (µ + ν + 2)x]
dx dx
µ,ν
+ n(n + µ + ν + 1)P n (x) = 0. (3.100)
• A recurrence relation:
µ,ν
2(n + 1)(n + µ + ν + 1)(2n + µ + ν)P n+1 (x)
· ¸
2 2 µ,ν
= (2n + µ + ν + 1) (2n + µ + ν)(2n + µ + ν + 2)x + ν − µ P n (x)
µ,ν
− 2(n + µ)(n + ν)(2n + µ + ν + 2)P n−1 (x). (3.101)
2) G EGENBAUER POLYNOMIALS
• Nomenclature: C nλ (x).
µ ¶
Γ n+λ+ 12 Γ(2λ)
• Standardization: K n = (−2) n! n µ ¶.
Γ(n+2λ)Γ λ+ 21
• Constants:
µ ¶
p 1
πΓ(n + 2λ)Γ λ + 2
2n Γ(n + λ)
k n(n) = , k n(n−1) = 0, hn = . (3.102)
n! Γ(λ) n!(n + λ)Γ(2λ)Γ(λ)
3) C HEBYSHEV POLYNOMIALS OF THE FIRST KIND 169
• Rodrigues formula:
µ ¶
n 1
(−1) Γ(n + 2λ)Γ λ + 2 n · ¸
λ 2 −λ+1/2 d 2 n+λ−1/2
C n (x) = µ ¶ (1 − x ) (1 − x ) .
1 d xn
2 n!Γ n + λ + 2 Γ(2λ)
n
(3.103)
• Differential equation:
d 2C nλ dC nλ
(1 − x 2 ) − (2λ + 1)x + n(n + 2λ)C nλ (x) = 0. (3.104)
d x2 dx
• A recurrence relation:
λ
(n + 1)C n+1 (x) = 2(n + λ)xC nλ (x) − (n + 2λ − 1)C n−1
λ
(x). (3.105)
• Nomenclature: Tn (x).
• Rodrigues formula:
n · ¸
(−1)n 2n n! 2 1/2 d 2 n−1/2
Tn (x) = (1 − x ) (1 − x ) . (3.106)
(2n)! d xn
• Differential equation:
d 2 Tn d Tn
(1 − x 2 ) 2
−x + n 2 Tn (x) = 0. (3.107)
dx dx
• A recurrence relation:
• Nomenclature: Un (x).
• Rodrigues formula:
n · ¸
(−1)n 2n (n + 1)! 2 −1/2 d 2 n+1/2
Un (x) = (1 − x ) (1 − x ) . (3.109)
(2n + 1)! d xn
• Differential equation:
d 2Un dUn
(1 − x 2 ) 2
− 3x + n(n + 2)Un (x) = 0. (3.110)
dx dx
• A recurrence relation:
As in case of the Fourier expansion, we can use the classical orthogonal poly-
nomials to write an arbitrary function f (x) ∈ L ω2 (a, b) as a series of these poly-
nomials. If we denote a complete set of orthogonal polynomials by {C k (x)}∞ k=0
,
with the corresponding vectors being represented by |C k 〉, the vectors
1
|e k 〉 = p |C k 〉, k = 0, 1, 2, . . . (3.112)
hk
where h k is given by Eq. (3.24), form an orthonormal basis in L ω2 (a, b). In this
way, if | f k 〉 are the vectors related to the function f (x) ∈ L ω2 (a, b), we can write
X
∞
|f 〉 = f k |e k 〉. (3.113)
k=0
where
Zb Zb
1
f k = 〈e k | f 〉 = d xω(x)e k∗ (x) f (x) = p d xω(x)C k∗ (x) f (x). (3.114)
hk
a a
3.5 E XPANSION IN TERMS OF ORTHOGONAL POLYNOMIALS 171
Zb ¯ ¯2
¯ X
∞ ¯
lim ¯
d xω(x)¯ f (x) − f k e k (x)¯¯ = 0. (3.116)
n→∞
a k=0
Then, we have
X
∞
f (x) = f k e k (x) (3.117)
k=0
Z1
1
fn = p d xP n (x)δ(x). (3.119)
hn
−1
Using Eq. (3.93) and the properties of the Dirac delta function
r
2n + 1
fn = P n (0). (3.120)
2
and, thus,
X∞ 2n + 1
δ(x) = P n (0)P n (x). (3.121)
n=0 2
For odd values of n, the above expression give zero, because P n (x) is an odd
polynomial. This is to be expected because δ(x) is an even function of x, i.e.,
3.6 G ENERATING FUNCTIONS 172
δ(x) = δ(−x) = 0 for x 6= 0. To evaluate P n (0) for even values of n, we use the
recurrence relation (3.95) for x = 0:
n −1
(n + 1)P n+1 (0) = −nP n−1 (0) =⇒
|{z} P n (0) = − P n−2 (0). (3.122)
n
n→n−1
(n − 1)(n − 3) . . . (n − 2m + 1)
P n (0) = (−1)m P n−2m (0). (3.123)
n(n − 2)(n − 4) . . . (n − 2m + 2)
(2m − 1)(2m − 3) . . . 3 · 1
P 2m (0) = (−1)m P 0 (0)
2m(2m − 2) . . . 4 · 2
2m(2m − 1)(2m − 2) . . . 3 · 2 · 1
= (−1)m
[2m(2m − 2) . . . 4 · 2]2
(2m)! (2m)!
= (−1)m m 2
= (−1)m 2m . (3.124)
[2 m!] 2 (m!)2
X
∞
g (x, t ) = a n t n F n (x), (3.126)
n=0
case of the Legendre polynomials and show that the functions P n (x) defined
by the equation
X
∞
g (x, t ) = (1 − 2xt + t 2 )−1/2 = t n P n (x), (3.127)
n=0
satisfy the Legendre equation and, thus, are the Legendre polynomials (note
that, in this case, the constants a n in Eq. (3.126) are all equal to 1). In this way,
from Eq. (3.127), the nth coefficient of the Taylor series expansion of g (x, t ) =
(1 − 2xt + t 2 )−1/2 about t = 0 is the Legendre polynomial P n (x). Specifically,
· µ ¶¸¯
1 ∂n 1 ¯
P n (x) = p ¯ , (3.128)
n! ∂t n ¯
1 − 2t x + t 2 t =0
and g (x, t ) = (1−2xt + t 2 )−1/2 is, thus, the generating function of the Legendre
polynomials.
Let’s start then. First, we differentiate the defining equation (3.127) with
respect to x and get
X∞ dP
t (1 − 2xt + t 2 )−3/2 =
n n
t . (3.129)
n=0 d x
X
∞
(x − t )(1 − 2xt + t 2 )−3/2 = nP n (x)t n−1 . (3.130)
n=0
X
∞ X∞ dP
n n
t P n (x)t n = (1 − 2xt + t 2 ) t , (3.131)
n=0 n=0 d x
or, equivalently
X
∞ X∞ dP
n n X∞ dP
n n+1 X
∞ dP
n n+2
P n (x)t n+1 = t − 2x t + t
n=0 n=0 d x n=0 d x n=0 d x
X∞ ·dP ¸
d P n d P n−1 n+1
n+1
= − 2x + t . (3.132)
n=0 dx dx dx
d P n+1 d P n d P n−1
P n (x) = − 2x + . (3.133)
dx dx dx
3.6 G ENERATING FUNCTIONS 174
X∞ dP
n n X∞ X∞
(x − t ) t =t nP n (x)t n−1 = nP n (x)t n , (3.134)
n=0 d x n=0 n=0
or, equivalently
· ¸
X
∞ d P n d P n−1 n X ∞
x − t = nP n (x)t n . (3.135)
n=0 dx dx n=0
d P n d P n−1
x − = nP n (x). (3.136)
dx dx
Eliminating d P n /d x between Eqs. (3.133) and (3.136) gives the further result
d P n+1 dP n
(n + 1)P n (x) = −x . (3.137)
dx dx
If we now take Eq. (3.137) with n replaced by n − 1 and add x times (3.136)
to it, we obtain
d Pn
(1 − x 2 ) = n[P n−1 (x) − xP n (x)]. (3.138)
dx
Differentiating now both sides with respect to x and using Eq. (3.136), we find
·µ ¶ ¸
2 d 2Pn d Pn d P n−1 d Pn
(1 − x ) − 2x =n −x − P n (x)
d x2 dx dx dx
= n[−nP n (x) − P n (x)] = −n(n + 1)P n (x), (3.139)
so the P n (x) defined by Eq. (3.127) satisfy the above differential equation,
which is precisely the Legendre equation [see Eq. (2.257)].
In the following we list the generating functions for the others classical
orthogonal polynomials and the corresponding value a n of Eq. (3.126):
2
• Hermite Hn (x): g (x, t ) = e −t +2xt
, a n = 1/n!.
because the first m terms of the first series have Γ functions in the denomi-
nator with negative integer (or zero) arguments. Now in the second series, we
3.7 S PECIAL FUNCTIONS : B ESSEL FUNCTIONS 176
2m
J m−1 (x) + J m+1 (x) = J m (x),
x
d Jm
J m−1 (x) − J m+1 (x) = 2 . (3.144)
dx
Combining these two equations, we obtain
m d J m (x)
J m−1 (x) = J m (x) + ,
x dx
m d Jm
J m+1 (x) = J m (x) − . (3.145)
x dx
We can use these equations to obtain new, and more useful, relations. For
example, by differentiating x m J m (x), we get
· ¸
d d Jm
x J m (x) = mx m−1 J m (x) + x m
m
dx dx
· ¸
m m d Jm
=x J m (x) + = x m J m−1 (x). (3.146)
x dx
· µ ¶¸¯ Zb
dg d f ¯¯b 2 2
ρ f −g = (k − l ) d ρρ f (ρ)g (ρ). (3.150)
dρ d ρ ¯a
a
In all physical applications, a and b can be chosen to make the left-hand side
vanish. Then, substituting for f and g in terms of the corresponding Bessel
functions, we get
Zb
2 2
(k − l ) d ρρ J m (kρ)J m (l ρ) = 0. (3.151)
a
Zb
d ρρ J m (kρ)J m (l ρ) = 0, if k 6= l . (3.152)
a
To complete the orthogonality relation, we must R also address the case when
2
k = l . This involves the evaluation of the integral d ρρ J m (kρ), which, upon
the change of variable x = kρ, reduces to
Z
1 2
2
d xx J m (x). (3.153)
k
Integrating by parts and using the recurrence relations for the Bessel func-
tions, it is possible to show that
Z · ¸
1 2 1 2 2 1 2 2 1 2 d Jm 2
d xx J m (x) = x J m (x) − m J m (x) + x , (3.154)
k2 2 2 2 dx
3.7 S PECIAL FUNCTIONS : B ESSEL FUNCTIONS 178
for all m ≥ 0 and, by Eq. (3.143), for all negative integers too. It is customary
to simplify the right-hand side of Eq. (3.156) by choosing k in such a way that
J m (ka) = 0, i.e., that ka is a root of the Bessel function of order m. If we denote
x mn the nth root of J m (x) (in general, there are infinitely many roots)
x mn
ka = x mn =⇒ k= , n = 1, 2, . . . , (3.157)
a
and if we use Eq. (3.145), we obtain
Za µ ¶ · ¸2
2 x mn ρ 1 2
d ρρ J m = a J m+1 (x mn ) . (3.158)
a 2
0