12
Vector Spaces in Physics 8/6/2015 10 - 1 Chapter 10. Fourier Transforms and the Dirac Delta Function A. The Fourier transform. The Fourier-series expansions which we have discussed are valid for functions either defined over a finite range ( /2 /2 T t T , for instance) or extended to all values of time as a periodic function. This does not cover the important case of a single, isolated pulse. But we can approximate an isolated pulse by letting the boundaries of the region of the Fourier series recede farther and farther away towards , as shown in figure 10-1. We will now outline the corresponding mathematical limiting process. It will transform the Fourier series, a superposition of sinusoidal waves with discrete frequencies n, into a superposition of a continuous spectrum of frequencies . As a starting point we rewrite the Fourier series, equation 9-39, as follows: /2 /2 () 1 n n i t n n T i t n T ft Ce n C f te dt T (10-1) The only change we have made is to add, in the upper expression, a factor of n for later use; 1 1 n n n is the range of the variable n for each step in the summation. We now imagine letting T get larger and larger. This means that the frequencies Figure10-1. Evolution of a periodic train of pulses into a single isolated pulse, as the domain of the Fourier series goes from [-T/2, T/2] to [-, ].

Chapter 10. Fourier Transforms and the Dirac Delta Functionbland/courses/385/downloads/vector/ch10.pdf10 - 1 Chapter 10. Fourier Transforms and the Dirac Delta Function A. The Fourier

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Vector Spaces in Physics 8/6/2015

10 - 1

Chapter 10. Fourier Transforms and the Dirac Delta Function

A. The Fourier transform.

The Fourier-series expansions which we have discussed are valid for functions either defined over a

finite range ( / 2 / 2T t T , for instance) or extended to all values of time as a periodic function.

This does not cover the important case of a single, isolated pulse. But we can approximate an

isolated pulse by letting the boundaries of the region of the Fourier series recede farther and farther

away towards , as shown in figure 10-1. We will now outline the corresponding mathematical

limiting process. It will transform the Fourier series, a superposition of sinusoidal waves with

discrete frequencies n, into a superposition of a continuous spectrum of frequencies .

As a starting point we rewrite the Fourier series, equation 9-39, as follows:

/ 2

/ 2

( )

1

n

n

i t

n

n

T

i t

n

T

f t C e n

C f t e dtT

(10-1)

The only change we have made is to add, in the upper expression, a factor of n for later use;

1 1n n n is the range of the variable n for each step in the summation. We now imagine

letting T get larger and larger. This means that the frequencies

Figure10-1. Evolution of a periodic train of pulses into a single isolated pulse, as the domain of the Fourier series goes

from [-T/2, T/2] to [-, ].

Vector Spaces in Physics 8/6/2015

10 - 2

2

n

n

T

(10-2)

in the sum get closer and closer together. In the large-n approximation we can replace the integer

variable n by a continuous variable n, so that

n

n

n n

C C n

n

n dn

(10-13)

We thus have

/ 2

/ 2

( )

1

i n t

n

Ti n t

T

f t C n e dn

C n f t e dtT

(10-4)

Next we change variables in the first integral from n to 2 n

nT

:

/ 2

/ 2

( )2

1

i t

T

i t

T

Tf t C e d

C f t e dtT

(10-5)

Now define

2

Tg C

(10-6)

This gives

/ 2

/ 2

1( )

2

1

2

i t

T

i t

T

f t g e d

g f t e dt

(10-7)

Finally, we take the limit T , giving the standard for m for the Fourier transform:

1

( )2

i tf t g e d

Inverse Fourier Transform (10-8)

1

2

i tg f t e dt

Fourier Transform (10-9)

There are a lot of notable things about these relations. First, there is a great symmetry in the roles

of time and frequency; a function is completely specified either by f(t) or by g(). Describing a

function with f(t) is sometimes referred to as working in the "time domain," while using g() is

referred to as working in the "frequency domain." Second, both of these expressions have the form

of an expansion of a function in terms of a set of basis functions. For f(t), the basis functions are

Vector Spaces in Physics 8/6/2015

10 - 3

1

2

i te

; for g(), the complex conjugate of this function,

1

2

i te

, is used. Finally, the

function g() emerges as a measure of the "amount" of frequency which the function f(t)

contains. In many applications, plotting g() gives more information about the function than

plotting f(t) itself.

Example - the Fourier transform of the square pulse. Let us consider the case of an isolated

square pulse of length T, centered at t = 0:

1,

( ) 4 4

0 otherwise

T Tt

f t

(10-10)

This is the same pulse as that shown in figure 9-3, without the periodic extension. It is

straightforward to calculate the Fourier transform g():

/ 4

/ 4

4 4

1

2

1

2

1 1

2

sin4

2 2

4

i t

t

T

i t

t T

T Ti i

g f t e dt

e dt

e ei

T

T

T

(10-11)

Here we have used the relation sin2

i ie e

i

. We have also written the dependence on in the

form sin

sincx

xx

. This well known function peaks at zero and falls off on both sides, oscillating

as it goes, as shown in figure 10-2.

B. The Dirac delta function (x).

The Dirac delta function was introduced by the theoretical physicist P.A.M. Dirac, to describe a

strange mathematical object which is not even a proper mathematical function, but which has many

uses in physics. The Dirac delta function is more properly referred to as a distribution, and Dirac

played a hand in developing the theory of distributions. Here is the definition of (x):

1)(

0,0)(

dxx

xx

(10-12)

Vector Spaces in Physics 8/6/2015

10 - 4

Isn't this a great mathematical joke? This function is zero everywhere! Well, almost everywhere,

except for being undefined at x=0. How can this be of any use? In particular, how can its integral

be anything but zero? As an intellectual aid, let's compare this function with the Kronecker delta

symbol, which (not coincidentally) has the same symbol:

1

,1

,0

3

1

i

ij

ijji

ji

(10-13)

There are some similarities. But the delta function is certainly not equal to 1 at x = 0; for the

integral over all x to be equal to 1, (x) must certainly diverge at x = 0 In fact, all the definitions

that I know of a Dirac delta function involve a limiting procedure, in which (x) goes to infinity.

Here are a couple of them.

The rectangular delta function

Consider the function

Figure10-2. The Fourier transform of a single square pulse. This function is sometimes called the sync function.

Vector Spaces in Physics 8/6/2015

10 - 5

0

1/ x2

( ) lim

0 x2

aa

xa

(10-14)

This function, shown in figure 10-3, is a rectangular

pulse of width a and height h = 1/a. Its area is equal to

( ) 1A f x dx h a

, so it satisfies the integral

requirement for the delta function. And in the limit that

a 0, it vanishes at all points except x = 0. This is one

perfectly valid representation of the Dirac delta

function.

The Gaussian delta function

Another example, which has the advantage of being an

analytic function, is

.

2

22

0

1 1( ) lim

2

x

x e

(10-15)

The function inside the limit is the Gaussian function, 2

221 1

( )2

x

g x e

(10-16)

in a form often used in statistics which is normalized so that

1)(

dxxg , and so that the standard deviation of the

distribution about x=0 is equal to . A graph of the

Gaussian shape was given earlier in this chapter; the width

of the curve at half maximum is about equal to 2. (See

figure 10-4.) It is clear that in the limit as goes to zero

this function is zero everywhere except at x = 0 (where it

diverges, due to the factor 1

), maintaining the

normalization condition all the while.

Properties of the delta function

By making a change of variable one can define the delta function in a more

general way, so that the special point where it diverges is x = a (rather

than x=0):

x

)

g(x)

Figure 10-4. The Gaussian function,

becoming a delta function in the limit

0 .

x

a 1/a

f(x)

Figure 10-3. Rectangular function,

becoming a delta function in the limit a 0.

Vector Spaces in Physics 8/6/2015

10 - 6

1)(

,0)(

dxax

axax

(10-17)

Two useful properties of the delta function are given below:

( ) ( ) ( )f x x a dx f a

, (10-18)

( ) '( ) '( )f x x a dx f a

, (10-19)

Here the prime indicates the first derivative.

The property given in equation (10-18) is fairly easy to understand; while carrying out the integral,

the argument vanishes except very near to x=a; so, it makes sense to replace f(a) by the constant

value f(a) and take it out of the integral. The second property, Eqn. (10-19), can be demonstrated

using integration by parts. The proof will be left to the problems.

C. Application of the Dirac delta function to Fourier transforms

Another form of the Dirac delta function, given either in k-space or in -space, is the following:

0

0

( )

0

( )

0

1( )

2

1( )

2

i k k x

i x

e dx k k

e dx

. (10-20)

We will not prove this fact, but just make an argument for its plausibility. Look at the integral (10-

20), for the case when k = k0. The exponential factor is just equal to 1 in that case, and it is

clear that the integral diverges. On the other hand, if k is not equal to k0, it is plausible that the

oscillating nature of the argument makes the integral vanish. If we accept these properties, we can

interpret the Fourier transform as an expansion of a function in terms of an orthonormal basis, just

as the Fourier series is an expansion in terms of a series of orthogonal functions. Here is the picture.

Basis states

The functions

tiee

2

1)(ˆ . (10-21)

constitute a complete orthonormal basis for the space of ''smooth'' functions on the interval

t . We are not going to prove completeness; as with the Fourier series, the fact that the

expansion approximates a function well is usually accepted as sufficient by physicists. The

orthonormality is defined using the following definition of an inner product of two (possibly

complex) functions u(t) and v(t):

vdtuvu

*, . (10-22)

where u* represents the complex conjugate of the function u(t). Now the inner product of two basis

states is

Vector Spaces in Physics 8/6/2015

10 - 7

)(

2

1

ˆˆˆ,ˆ*

dtee

dteeee

titi (10-23)

(The proof of the last line in the equation above is beyond the scope of these notes - sorry.) This is

the equivalent of the orthogonality relation for sine waves, equation (9-8), and shows how the Dirac

delta function plays the same role for the Fourier transform that the Kronecker delta function plays

for the Fourier series expansion.

We now use this property of the basis states to derive the Fourier inversion integral. Suppose that

we can expand an arbitrary function of t in terms of the exponential basis states:

.

degtf ti)(2

1)( (10-24)

Here represents a sort of weighting function, still to be determined, to include just the right

amount of each frequency to correctly reproduce the function f(t). This is a continuous analog of

the representation of a vector A

in terms of its components,

.

jjeAA ˆ

(10-25)

The components Ak are obtained by taking the inner product of the k-th basis vector with A

:

.

AeA kk

ˆ (10-26)

If the analogy holds true, we would expect the weighting function (the ''coordinate'') to be

determined by

dttfe

dttfe

tfeg

ti

ti

ti

)(2

1

)(2

1

)(,2

1

*

(10-27)

This is exactly the inverse Fourier transformation which we postulated previously in equation (10-

9). We can now prove that it is correct. We start with the inner product of the basis vector with f(t):

dtdege

dttfetfe

titi

titi

)(2

1

2

1

)(2

1)(,

2

1

(10-28)

Here we have substituted in the Fourier expansion for f(t), changing the variable of integration to

’ to keep it distinct. Now we change the order of the integrations:

Vector Spaces in Physics 8/6/2015

10 - 8

dtedgtfe titi

2

1)()(,

2

1 (10-29)

and next use the definition of the delta function, giving

)(

)'()(

2

1)()(,

2

1

g

dg

dtedgtfe titi

(10-30)

as we had hoped to show.

Functions of position x

The Fourier transform can of course be carried out for functions of a position variable x, expanding

in terms of basis states

.

ikxeke2

1)(ˆ (10-31)

Here

2k (10-32)

is familiar as the wave number for traveling waves. In terms of x and k, the Fourier transform takes

on the form

1( ) ( )

2

ikxf x g k e dk

x

k

Inverse Fourier Transform

(10-33)

1

( ) ( )2

ikxg k f x e dx

Direct Fourier Transform (10-34)

D. Relation to Quantum Mechanics

The expressions just above may evoke memories of the formalism of quantum mechanics. It is a

basic postulate of quantum mechanics that a free particle of momentum p is represented by a wave

function

,

px

iikx

p x Ae Ae (10-35)

where 2

h and h is Planck's constant and

2k , giving back the deBroglie relationship

between a particle’s momentum and wavelength,

Vector Spaces in Physics 8/6/2015

10 - 9

.

p

h (10-36)

The wave function p x corresponds to a particle with a precisely defined wavelength, but whose

spatial extent goes from x = - to x = . The Fourier transform thus represents a linear

superposition of wave functions with different wavelengths, and can be used to create a ''wave

packet'' f(x) which occupies a limited region of space. However, there is a price to pay; now the

wave function corresponds to a particle whose wavelength, and momentum, is no longer exactly

determined. The interplay between the uncertainty in position and the uncertainty in momentum is

one of the most famous results of quantum mechanics. It is often expressed in terms of the

uncertainty principle,

2

px . (10-37)

We can see how this plays out in a practical example by

creating a wave packet with the weighting function

2 20

2( )

k k aa

g k e

. (10-38)

In quantum mechanics the probability distribution is obtained

by taking the absolute square of the wave function. Here g(k)

is the wave function in “k-space,” with a corresponding

probability distribution given by

2 20

2 20

2

2

2

( ) ( )

1

k k a

k k a

P k g k

ae

ae

, (10-39)

plotted in figure 10-5. Comparing to our previous expression for the normalized Gaussian with

standard deviation ,

2

221 1

( )2

x

g x e

, (10-40)

we see that P(k) is a Gaussian probability distribution which is normalized to unity and which has a

standard distribution (of k, about k0, of 2

1

a; that is,

.

2

1

akk . (10-41)

In terms of the corresponding particle momentum p,

,

k 2

1

ak

P(k)

k0

Figure 10-5. Probability function

2 2

0/ exp ( )P k a k k a

Vector Spaces in Physics 8/6/2015

10 - 10

2a

kp

(10-42)

where we have taken the rms deviation of the momentum distribution about its average value to

represent the uncertainty in p. We see that this distribution can be made as narrow as desired,

corresponding to determining the momentum as accurately as desired, by making a large.

Now, what is the corresponding uncertainty in position? We have to carry out the Fourier

transform to find out:

dkea

dkeea

dkekgxf

kxa

ikk

a

ikx

akk

ikx

2

20

2

220

2

2

2

2

1

2

1

)(2

1)(

(10-43)

We will work with the quantity in square brackets, collecting terms linear in k and performing an

operation called completing the square:

422

0

2

0

22

0

2

0

22

0

22

0

2

0

2

2

0

2

0

2

2

2

0

//2/

///2

/22

axaxikkaixkk

kaixkaixkaixkkk

kaixkkkkxa

ikk

(10-44)

The factor 22

0 / aixkk is the “completed square.” Substituting into the previous equation

gives

2

2

0

220

2

220

2/

2

422

0

2

0

22

0

2

2

1

//2/

2

1

)(2

1)(

a

x

xikaixkk

a

ikx

akk

ikx

eedkea

axaxikkaixkk

dkeea

dkekgxf

(10-45)

The quantity in square brackets is the integral of a Gaussian function and is equal toa

2; so we

have

2

2

0 21

)( a

x

xikee

axf

(10-46)

Vector Spaces in Physics 8/6/2015

10 - 11

This is a famous result: the Fourier transform of a Gaussian is a Gaussian! Furthermore, the wave

function in ''position space,'' f(x), also gives a Gaussian probability distribution:

2

2

1

)()(2

a

x

ea

xfxP

(10-47)

This is a normalized probability distribution, plotted in figure 10-6, with an rms deviation for x,

about zero, of

2

axx (10-48)

So, the product of the uncertainties in position space and k-space is

2

1

2

1

2

a

akx (10-49)

and

2

kxpx (10-50)

The Gaussian wave function satisfies the uncertainty

principle, as it should.

The uncertainty principle requires the product of the

uncertainties of any two “conjugate variables” (x and px, or E

and t, for examples) to be greater than or equal to /2, for

any possible wave function. The Gaussian wave function is a

special case, the only form of the wave function for which

the lower limit given by the uncertainty principle is exactly

satisfied.

Problems

Problem 10-1. Fourier Transform of a Rectangular Spectral Function.

Calculate the (inverse) Fourier transform f(x), given by eq. (10-33), for the spectral function g(k)

given by

11

( ) 2

0 otherwise

kg k a

.

x 2

ax

P(x)

0

Figure10-6. Probability function

2

2

1)( a

x

ea

xP

Vector Spaces in Physics 8/6/2015

10 - 12

Make a plot of the result, f(x) vs x. Indicate the values of x at the zero crossings.

Problem 10-2. Theorem Concerning the Dirac Delta Function. Use integration by parts to

prove the theorem given in eq. 10-19,

)()( afdxaxxf

(Here f' represents the derivative of f with respect to its argument, and ' represents the derivative

of with respect to its argument.)

Problem 10-3. Fourier Transform of a Gaussian. Calculate the Fourier transform (eq. 10-34)

of the wave function 2

20 2

1( )

x

ik x af x e ea

and verify that the result is

2 2

0

2( )

k k aa

g k e

.