2

THE FOURIER TRANSFORM

In this section, we develop the Fourier transform and its inverse. The Fourier transform can be thought of as a continuous form of Fourier series. A Fourier series decomposes a signal on [–π, π] into components that vibrate at integer frequencies. By contrast, the Fourier transform decomposes a signal defined on an infinite time interval into a λ-frequency component where λ can be any real (or even complex) number. Besides being of interest in their own right, the topics in this chapter will be important in the construction of wavelets in the later chapters.

2.1 INFORMAL DEVELOPMENT OF THE FOURIER TRANSFORM

2.1.1 The Fourier Inversion Theorem

In this section we give an informal development of the Fourier transform and its inverse. Precise arguments are given in the Appendix (see A.l).

To obtain the Fourier transform, we consider the Fourier series of a function defined on the interval –lxl and then let l go to infinity. Recall from Theorem 1.20 (with a = l) that a function defined on –lxl can be expressed as

where

c02e002

If f is defined on the entire real line, then it is tempting to let l go to infinity and see how the preceding formulas are affected. Substituting the expression for αn into the previous sum, we obtain

c02e003

Our goal is to recognize the sum on the right as the Riemann sum formulation of an integral. To this end, let c02ie001. We obtain

(2.1)c02e004

Let

c02e005

The sum in Eq. (2.1) is

c02e006

This term resembles the Riemann sum definition of the integral c02ie002.As l converges to ∞, the quantity Δλ. converges to 0 and so Δλ becomes the dλ in the integral c02ie003. So (2.1) becomes

c02e007

As l c02ie004 ∞, Fl (λ) formally becomes the integral c02ie005. Therefore

c02e008

or

(2.2)c02e009

We let c02ie006(λ) be the quantity inside the parentheses, that is,

c02e010

The function c02ie006(λ) is called the (complex form of the) Fourier transform of f. Equation (2.2) becomes

c02e011

which is often referred to as the Fourier inversion formula, since it describes f(x) as an integral involving the Fourier transform of f.

We summarize this discussion in the following theorem.

Theorem 2.1 If f is a continuously differentiable function with c02ie007 dt < ∞, then

(2.3)c02e012

where c02ie006(λ) (the Fourier transform of f) is given by

c02e013

The preceding argument is not rigorous since we have not justified several steps including the convergence of the improper integral c02ie008. As with the development of Fourier series (see Theorem 1.28), if the function f(x) has points of discontinuity, such as a step function, then the preceding formula holds with f(x) replaced by the average of the left- and right-hand limits, that is,

(2.4)c02e014

Rigorous proofs of these results are given in Appendix A.

The assumption c02ie009 in Theorem 2.1 assures us that the improper integral defining c02ie006, that is,

c02e015

is absolutely convergent. Often, this assumption doesn’t hold. However, in nearly all of the cases in which we are interested, it is not a difficulty if the improper integral is only conditionally convergent. See Example 2.2.

Comparison with Fourier Series. The complex form of the Fourier transform of f and the corresponding inversion formula are analogous to the complex form of the Fourier series of f over the interval –l ≤ xl:

(2.5)c02e016

where

c02e017

The variable λ in the Fourier inversion formula, Eq. (2.3), plays the role of c02ie010 in Eq. (2.5). The sum over n from –∞ to ∞ in Eq. (2.5) is replaced by an integral with respect to λ from –∞ to ∞ in Eq. (2.3). The formulas for c02ie011 and c02ie006(λ) are also analogous. The integral over [–l, l] in the formula for c02ie011 is analogous to the integral over (–∞,∞) in c02ie006(λ). In the case of Fourier series, c02ie011 measures the component of f which oscillates at frequency n. Likewise, c02ie006(λ) measures the frequency component of f that oscillates with frequency λ. If f is defined on a finite interval, then its Fourier series is a decomposition of f into a discrete set of oscillating frequencies (i.e., c02ie011, one for each integer n). For a function on an infinite interval, there is a continuum of frequencies since the frequency component, c02ie006(λ), is defined for each real number λ. These ideas will be illustrated in the following examples.

2.1.2 Examples

Example 2.2 In our first example, we will compute the Fourier transform of the rectangular wave (see Figure 2.1):

c02e018

Figure 2.1. Rectangular wave.

c02f001

Now, we have c02ie012. Since f is an even function, f (t) sin (λt) is an odd function and its integral over the real line is zero. Therefore the Fourier transform of f is reduced to

c02e019

A graph of c02ie006 is given in Figure 2.2.

As already mentioned, the Fourier transform, c02ie006(λ), measures the frequency component of f that vibrates with frequency λ. In this example, f is a piece wise constant function. Since a constant vibrates with zero frequency, we should expect that the largest values of c02ie006(λ) occur when λ is near zero. The graph of c02ie006 in Figure 2.2 clearly illustrates this feature.

This example also illustrates the inversion formula when there is a jump discontinuity. By Eq. (2.4), we have that

c02e020

Figure 2.2. Fourier transform of a rectangular wave.

c02f002

The improper integral above is conditionally convergent, rather than absolutely convergent, because c02ie013. Even so, the inversion formula is still valid.

Example 2.3 Let

c02e021

(see Figure 2.3).

Since f is an even function, only the cosine part of the transform contributes:

c02e022

The preceding integral is left as an exercise (sum the two identities

c02e023

with u = 3t and v = λt and then integrate). The result is

c02e024

Figure 2.3. Plot of cos(3f).

c01f003

The graph of c02ie006 is given in Figure 2.4.

Note that the Fourier transform peaks at λ = 3 and –3. This should be expected since f(t) = cos(3t) vibrates with frequency 3 on the interval –πttπ.

Figure 2.4. Fourier transform of cos(3f).

c02f004

Example 2.4 Let

c02e025

Since f is an odd function, only the sine part of the transform contributes (which is purely imaginary). The transform is

c02e026

Example 2.5 The next example is a triangular wave whose graph is given in Figure 2.5. An analytical formula for this graph is given by

c02e027

Figure 2.5. Triangular wave.

c02f005

This function is even and so its Fourier transform is given by

c02e028

This integral can be computed using integration by parts:

c02e029

The graph of c02ie006 is given in Figure 2.6.

Note that the Fourier transforms in Examples 2.4 and 2.5 decay at the rate l/λ2 as λ c02ie004 ∞ which is faster than the decay rate of 1/λ. exhibited by the Fourier transforms in Examples 2.2 and 2.3. The faster decay in Examples 2.4 and 2.5 results from the continuity of the functions in these examples. Note the parallel with the examples in Chapter 2. The Fourier coefficients, an and bn, for the discontinuous function in Example 1.9 decay like l/n, whereas the Fourier coefficients for the continuous function in Example 1.10 decay like 1/n2.

Figure 2.6. Fourier transform of the triangular wave.

c02f006

2.2 PROPERTIES OF THE FOURIER TRANSFORM

2.2.1 Basic Properties

In this section we set down most of the basic properties of the Fourier transform. First, we introduce the alternative notation

c02e030

for the Fourier transform of f. This notation has advantages when discussing some of the operator theoretic properties of the Fourier transform. The Fourier operator c02ie014 should be thought of as a mapping whose domain and range is the space of complex-valued functions defined on the real line. The input of T is a function, say f, and returns another function, c02ie014[f] = c02ie006 , as its output.

In a similar fashion, we define the inverse Fourier transform operator as

c02e031

Theorem 2.1 implies that c02ie014–1 really is the inverse of c02ie014

(2.6)c02e032

because

c02e033

Some properties of the Fourier transform and its inverse are given in the following theorem. Other basic properties will be given in the exercises.

Theorem 2.6 Let f and g be differentiable functions defined on the real line with f (t) = 0 for large |t|. The following properties hold.

1. The Fourier transform and its inverse are linear operators. That is, for any constant c we have

c02e034

2. The Fourier transform of a product of f with tn is given by

c02e035

3. The inverse Fourier transform of a product of f with λn is given by

c02e036

4. The Fourier transform of an nth derivative is given by

c02e037

(here, f(n) stands for the nth derivative of f.

5. The inverse Fourier transform of an nth derivative is given by

c02e038

6. The Fourier transform of a translation is given by

c02e039

7. The Fourier transform of a rescaling is given by

c02e040

8. If f (t) = 0 for t < 0, then

c02e041

where c02ie015[f] is the Laplace transform of f defined by

c02e042

Proof. We prove each part separately.

1. The linearity of the Fourier transform follows from the linearity of the integral as we demonstrate in the following:

c02e043

The proof for c02ie014[cf] = cc02ie014[f] is similar as are the proofs for the corresponding facts for the inverse Fourier transform.

2 and 3. For the Fourier transform of a product of f with tn, we have

c02e044

Using

c02e045

we obtain

c02e046

The corresponding property for the inverse Fourier transform is proved similarly.

4 and 5. For the Fourier transform of the nth derivative of f, we have

c02e047

Now, we integrate by parts:

c02e048

with dv = f(n) and u = e–iλt. As we will see, this process has the effect of transferring the derivatives from f(n) to e–λt. Since f vanishes at +∞ and –∞ by hypothesis, there are no boundary terms. So we have

c02e049

Note that integration by parts has reduced the number of derivatives on f by one (f(n) becomes f(n–1)). We have also gained one factor of ik. By repeating this process n – 1 additional times, we obtain

c02e050

as desired. The proof for the corresponding fact for the inverse Fourier transform is similar.

6 and 7. The translation and rescaling properties can be combined into the following statement:

(2.7)c02e051

This equation can be established by the following change of variables argument:

c02e052

where the second equality follows from the change of variables s = bta (and so t = (s + a)/b and dt = c02ie016). Rewriting the exponential on the right as

c02e053

we obtain

c02e054

8. The final part of the theorem, regarding the relationship of the Fourier and Laplace transforms, follows from their definitions and is left to the exercises. This completes the proof. I

Example 2.7 We illustrate the fourth property of Theorem 2.6 with the function in Example 2.4:

c02e055

whose Fourier transform is

c02e056

The derivative of f is 3 cos 3t on –π ≤ t ≤ π, which is just a multiple of 3 times the function given in Example 2.3. Therefore from Example 2.3 we have

(2.8)c02e057

The fourth property of Theorem 2.6 (with n = 1) states

c02e058

This expression for c02ie006′(λ) agrees with Eq. (2.8).

Example 2.8 In this example, we illustrate the scaling property:

(2.9)c02e059

If b > 1, then the graph of f (bt) is a compressed version of the graph of f. The dominant frequencies of f (bt) are larger than those of f by a factor of b. This behavior is illustrated nicely by the function in Example 2.4:

c02e060

whose graph is given in Figure 2.7. The graph of f(2t) is given in Figure 2.8. Note the frequency f (2t) is double that off.

Increasing the frequency of a signal has the effect of stretching the graph of its Fourier transform. In the preceding example, the dominant frequency of f (t) is 3, whereas the dominant frequency of f (2t) is 6. Thus the maximum value of |c02ie006(λ)| occurs at λ = 3 (see Figure 2.9) whereas the maximum value of the

Figure 2.7. Plot of sin(3t) –πtπ.

c02f007

Figure 2.8. Plot of sin(6t) –π/2 ≤ t ≤ π/2.

c02f008

Figure 2.9. Plot of Fourier transform of sin(3f ).

c02f009

Fourier transform of f (2t) occurs at λ = 6 (see Figure 2.10). Thus the latter graph is obtained by stretching the former graph by a factor of 2. Note also that the graph of c02ie006(λ/2) is obtained by stretching the graph of c02ie006(λ) by a factor of 2.

This discussion illustrates the following geometrical interpretation of Eq. (2.9) in the case where b > 1: Compressing the graph of f speeds up the frequency

Figure 2.10. Plot of Fourier transform of sin(6f).

c02f010

and therefore stretches the graph of c02ie006(λ). If 0 < b < 1, then the graph of f(bt) is stretched, which slows the frequency and therefore compresses the graph of c02ie006(λ)

2.2.2 Fourier Transform of a Convolution

Now we examine how the Fourier transform behaves under a convolution, which is one of the basic operations used in signal analysis. First, we give the definition of the convolution of two functions.

Definition 2.9 Suppose f and g are two square-integrable functions. The convolution of f and g, denoted f * g, is defined by

c02e061

The preceding definition is equivalent to

c02e062

(perform the change of variables y = tx and then relabel the variable y back to x).

We have the following theorem on the Fourier transform of a convolution of two functions.

Theorem 2.10 Suppose f and g are two integrable functions. Then

(2.10)c02e063

(2.11)c02e064

Proof. To derive the first equation, we use the definitions of the Fourier transform and convolution:

c02e065

We write c02ie017. After switching the order of integration, we obtain

c02e066

Letting s = tx, we obtain

c02e067

The right side can be rewritten as

c02e164

which is c02ie018, as desired.

Equation (2.11) follows from Eq. (2.10) and the inverse formula for the Fourier transform as follows:

c02e068

This completes the proof of the theorem.

2.2.3 Adjoint of the Fourier Transform

Recall that the adjoint of a linear operator T : V c02ie004 W between inner product spaces is an operator T* : W c02ie004 V such that

c02e069

In the next theorem, we show that the adjoint of the Fourier transform is the inverse of the Fourier transform.

Theorem 2.11 Suppose f and g are square integrable. Then

c02e070

Proof. We have:

c02e071

(by switching the order of integration). The second integral (involving g) is c02ie019 therefore

c02e072

as desired.

2.2.4 Plancherel Theorem

The Plancherel formula states that the Fourier transform preserves the L2 inner product.

Theorem 2.12 Suppose f and g are square integrable. Then

(2.12)c02e073

(2.13)c02e074

In particular,

(2.14)c02e075

Proof. Equation (2.12) follows from Theorems 2.11 and 2.1 as follows:

c02e076

as desired. Equation (2.13) can be established in a similar manner. Equation (2.14) follows from (2.12) with f = g.

Remark. The equation c02ie020 is analogous to Eq. (1.41), and is also referred to as Parseval’s equation. It has the following interpretations. For a function, f, defined on [–π, π], let c02ie014(f)(n) be its nth Fourier coefficient (except with the factor c02ie021 instead of 1/2π):

c02e077

Equation (1.41) can be restated as

c02e078

which is analogous to (2.14) for the Fourier transform with l2 and L2[–π, π] replaced by the L2 norm on the entire real line. As in the case with Fourier series, Plancherel’s Theorem states that the energy of a signal in the time domain, c02ie022, is the same as the energy in the frequency domain c02ie023

2.3 LINEAR FILTERS

2.3.1 Time-Invariant Filters

The Fourier transform plays a central role in the design of filters. A filter can be thought of as a “black box” that takes an input signal, processes it, and then returns an output signal that in some way modifies the input. One example of a filter is a device that removes noise from a signal.

From a mathematical point of view, a signal is a function f : R c02ie004 C that is piecewise continuous. A filter is a transformation L that maps a signal,f, into another signal c02ie024. This transformation must satisfy the following two properties in order to be a linear filter:

  • Additivity: L[f + g] = L[f] + L[g]
  • Homogeneity: L[cf] = cL[f], where c is a constant.

There is another property that we want our filter L to have. If we play an old, scratchy record for half an hour starting at 3 pm today and put the signal through a noise reducing filter, we want to hear the cleaned-up output, at roughly the same time as we play the record. If we play the same record at 10 am tomorrow morning and use the same filter, we should hear the identical output, again at roughly the same time. This property is called time invariance. To formulate this concept, we introduce the following notation: For a function f(t) and a real number a, let fa(t) = f(ta). Thus fa is a time shift, by a units, of the signal f.

Definition 2.13 A transformation L (mapping signals to signals) is said to be time-invariant if for any signal f and any real number a, L[fa](t) = (Lf)(ta) for all t (or L[fa] = (Lf)a). In words, L is time-invariant if the time-shifted input signal f(ta) is transformed by L into the time-shifted output signal (Lf)(ta). (See Figure 2.11.)

Example 2.14 Let l(t) be a function that has finite support (i.e., l(t) is zero outside of a finite t-interval). For a signal f, let

c02e079

This linear operator is time-invariant because for any aR

c02e080

Thus, (Lf)(ta) = L[fa)](t) and so L is time-invariant.

Not every linear transformation has this property, as the following example shows.

Figure 2.11. L is time-invariant if the upper and lower outputs are the same.

c02f011

Example 2.15 Let

c02e081

On one hand, we have

c02e082

On the other hand, we obtain

c02e083

Since L[fa](t) and (Lf)(ta) are not the same (for a ≠ 0), L is not time-invariant. ¦

The next lemma and theorem show that the convolution in Example 2.14 is typical of time-invariant linear niters. We start by computing L(eiλt).

Lemma 2.16 Let L be a linear, time-invariant transformation and let λ be any fixed real number. Then, there is a function h with

c02e084

Remark. Note that the input signal eiλt is a (complex-valued) sinusoidal signal with frequency λ. This lemma states that the output signal from a time-invariant filter of a sinusoidal input signal is also sinusoidal with the same frequency.

Proof. Our proof is somewhat informal in order to clearly explain the essential ideas. Let hλ(t) = L(eiλt). Since L is time-invariant, we have

(2.15)c02e085

for each real number a. Since L is linear, we also have

c02e086

Thus

(2.16)c02e087

Comparing Eqs. (2.15) and (2.16), we find

c02e088

Since a is arbitrary, we may set a = t yielding hλ(0) = eiλthλ(t); solving for hx(t), we obtain

c02e089

Letting c02ie033(λ) = hλ(0)/c02ie025 completes the proof.

The function c02ie033(λ) determines L. To see this, we first use the Fourier Inversion Theorem (Theorem 2.1)

c02e090

Then we apply L to both sides:

(2.17)c02e091

The integral on the right can be approximated by a Riemann sum:

(2.18)c02e092

Since L is linear, we can distribute L across the sum:

(2.19)c02e093

As the partition gets finer, the Riemann sum on the right becomes an integral and so from (2.17), (2.18), and (2.19), we obtain

c02e095

Even though the preceding argument is not totally rigorous, the result is true with very few restrictions on either L or the space of signals being considered (see the text on Fourier analysis by Stein and Weiss (1971) for more details). We summarize this discussion in the following theorem.

Theorem 2.17 Let L be a linear, time-invariant transformation on the space of signals that are piecewise continuous functions. Then there exists an integrable function, h, such that

c02e096

for all signalsf.

Physical Interpretation. Both h(t) and c02ie033(λ) have physical interpretations. Assume that h(t) is continuous and that δ is a small positive number. We apply L to the impulse signal

c02e097

whose graph is given in Figure 2.12. Note that c02ie026. Applying L to fδ, we obtain

c02e098

Figure 2.12. Graph of fδ.

c01f012

Since h is continuous, h(tτ) is approximately equal to h(t) for |τ| ≤ δ. Therefore

c02e099

Thus h(t) is the approximate response to an input signal which is an impulse. For that reason, h(t) is called the impulse response function.

We have already seen that c02ie027. Thus up to a constant factor, c02ie033(λ) is the amplitude of the response to a “pure frequency” signal eiλt; is called the system function.

2.3.2 Causality and the Design of Filters

Designing a time-invariant filter is equivalent to constructing the impulse function, h, since any such filter can be written as

c02e100

by Theorem 2.17. The construction of h depends on what the filter is designed to do. In this section, we consider filters that reduce high frequencies, but leave the low frequencies virtually unchanged. Such filters are called low-pass filters.

Taking the Fourier transform of both sides of Lf = f * h and using Theorem 2.10 yields

c02e101

A Faulty Filter. Suppose we wish to remove all frequency components from the signal f that lie beyond some cutoff frequency λc. As a natural first attempt, we choose an h whose Fourier transform is zero outside of the interval –λc ≤ λ ≤ λc:

(2.20)c02e102

(the choice of constant c02ie021 is for convenience in later calculations). Since c02ie028 is zero for |λ| > λc, this filter at least appears to remove the unwanted frequencies (above λc) from the signal f. However, we will see that this filter is flawed in other respects.

The impulse response function corresponding to the system function hXc is easy to calculate:

c02e103

c02e104

Therefore

(2.21)c02e105

Now we filter the following simple input function:

c02e106

Think of ftc as a signal that is “on” for t between 0 and tc and off at other times. The effect of this filter on the signal ftc is

c02e107

where c02ie029. A plot of c02ie030 with tc and λc both 1 is given in Figure 2.13. Note that the graph of the output signal is nonzero for t < 0, whereas the input signal, ftc(t), is zero for t < 0. This indicates that the output signal occurs before the input signal has arrived! Clearly, a filter cannot be physically constructed to produce an output signal before receiving an input signal. Thus, our first attempt at constructing a filter by using the function hλc is not practical. The following definition and theorem characterize a more realistic class of filters.

Causal Filters

Definition 2.18 A causal filter is one for which the output signal begins after the input signal has started to arrive.

The following result tells us which filters are causal.

Figure 2.13. Graph of ± {Si(f) – Si(f – 1)}.

c01f013

Theorem 2.19 Let L be a time-invariant filter with response function h (i.e., Lf = f * h). L is a causal filter if and only if h(t) = 0 for all t < 0.

Proof. We prove that if h(t) = 0 for all t < 0, then the corresponding filter is causal. We leave the converse as an exercise (see exercise 8). We first show that if f (t) = 0 for t < 0, then (Lf)(t) = 0 for t < 0. We have

c02e108

where the lower limit in the integral is 0 because f (τ) = 0 when τ < 0. If t < 0 and τ ≥ 0, then t – τ < 0 and so h(tτ) = 0, by hypothesis. Therefore, (Lf)(t) = 0 for t < 0. We have therefore shown that if f (t) = 0 for t < 0, then (Lf)(t) = 0 for t < 0. In other words, if the input signal does not arrive until t = 0, the output of the filter also does not begin until t = 0.

Suppose more generally that the input signal f does not arrive until t = a. To show that L is causal, we must show that Lf does not begin until t = a. Let g(t) = f (t + a). Note that the signal g(t) begins at t = 0. From the previous paragraph, (Lg)(t) does not begin until t = 0. Since f (t) = g(ta) = ga(t), we have

c02e109

Since (Lg)(τ) does not begin until τ = 0, we see that (Lg)(ta) does not begin until t = a. Thus (Lf)(t) = (Lg)(ta) does not begin until t = a, as desired.

Theorem 2.19 applies to the response function, but it also gives us important information about c02ie033(λ). By the definition of the Fourier transform:

c02e110

If the filter associated to h is causal, Theorem 2.19 implies h(t) = 0 for t < 0 and so

c02e111

We summarize this discussion in the following theorem.

Theorem 2.20 Suppose L is a causal filter with response function h. Then the system function associated with L is

c02e112

where c02ie031 is the Laplace transform.

Example 2.21 One of the older causal, noise-reducing filters is the Butterworth filter (Papoulis, 1962). It is constructed using the previous theorem with

c02e113

where A and a are positive parameters. Its Fourier transform is given by

c02e114

(see Exercise 10). Note that c02ie033(λ) decays as λ c02ie004 ∞, thus diminishing the high frequency components of the filtered signal c02ie032.

Consider the signal given by

c02e115

whose graph is given in Figure 2.14. We wish to filter the noise that vibrates with frequency approximately 40. At the same time, we do not want to disturb the basic shape of this signal, which vibrates in the frequency range of 2–4 By choosing A = α = 10, c02ie033(λ) is close to c02ie033(0) = c02ie021 for |λ| ≤ 4; but c02ie033(λ)| is small (less than 0.1) when λ ≥ 40. Thus filtering by h with this choice of parameters A and α should preserve the low frequencies (frequency ≤ 4) while damping the high frequencies (> 40). A plot of the filtered signal (f * h)(t) for 0 ≤ tπ is given in Figure 2.15. Most of the high-frequency noise has been filtered. Most of the low-frequency components of this signal have been preserved.

Figure 2.14. Graph of et/3 (sin 2t + 2sin4t + 0.4sin2t sin40t).

c02f014

Figure 2.15. Graph of the filtered signal from Figure 2.14.

c02f015

2.4 THE SAMPLING THEOREM

In this section we examine a class of signals (i.e., functions) whose Fourier transform is zero outside a finite interval [–Ω, Ω]; these are (frequency) band-limited functions. For instance, the human ear can only hear sounds with frequencies less than 20 kHz (1 kHz = 1000 cycles per second). Thus, even though we make sounds with higher pitches, anything above 20 kHz can’t be heard. Telephone conversations are thus effectively band-limited signals. We will show below that a band-limited signal can be reconstructed from its values (or samples) at regularly spaced times. This result is basic in continuous-to-digital signal processing.

Definition 2.22 A function f is said to be frequency band-limited if there exists a constant Ω > 0 such that

c02e116

When Q is the smallest frequency for which the preceding equation is true, the natural frequency v := c02ie034 is called the Nyquist frequency, and 2v = c02ie035 is the Nyquist rate.

Theorem 2.23 (Shannon-Whittaker Sampling Theorem). Suppose that c02ie036 (λ) is piecewise smooth and continuous and that c02ie037, where Ω is some fixed, positive frequency. Then f = c02ie014–1[c02ie036] is completely determined by its values at the points tj = jπ/Ω, j = 0, ±1, ±2, .... More precisely, f has the following series expansion:

(2.22)c02e117

where the series on the right converges uniformly.

This is a remarkable theorem! Let’s look at how we might use it to transmit several phone conversations simultaneously on a single wire (or channel). As mentioned earlier, phone conversations are band-limited. In fact the dominant frequencies are below 1 kHz, which we will take as our Nyquist frequency. The Nyquist rate v = c02ie035 is then double this, or 2 kHz; so we need to sample the signal every c02ie038 millisecond. How many phone conversations can we send in this manner? Transmission lines typically send about 56 thousand bits per second. If each sample can be represented by 7 bits, then we can transmit 8 thousand samples per second, or 8 every millisecond, or 4 every half millisecond. By tagging and interlacing signals, we can transmit the samples from four conversations. At the receiver end, we can use the series in Eq. (2.22) to reconstruct the signal, with the samples being f(c02ie038j) for j an integer and time in milliseconds (only a finite number of these occur during the life of a phone conversation–unless perhaps a teenager is on the line). Here is the proof of the theorem.

Proof. Using Theorem 1.20 (with a = Ω and t = λ), we expand c02ie036 (λ) in a Fourier series on the interval [–Ω, Ω]:

c02e118

Since c02ie036 (λ) = 0 for |λ| ≥ Ω, the limits in the integrals defining ck can be changed to –∞ and ∞:

c02e119

By Theorem 2.1,

c02e120

If we use this expression for ck in the preceding series, and if at the same time we change the summation index from k to j = –k, we obtain

(2.23)c02e121

Since c02ie036 is a continuous, piecewise smooth function, the series (2.23) converges uniformly by Theorem 1.30. Using Theorem 2.1, we obtain

c02e122

Using Eq. (2.23) for c02ie006 and interchanging the order of integration and summation, we obtain

(2.24)c02e123

The integral in Eq. (2.24) is

c02e124

After simplifying Eq. (2.24), we obtain Eq. (2.22), which completes the proof.

The convergence rate in Eq. (2.22) is rather slow since the coefficients (in absolute value) decay like 1/j. The convergence rate can be increased so that the terms behave like 1/j2 or better, by a technique called over sampling, which is discussed in exercise 14. At the opposite extreme, if a signal is sampled below the Nyquist rate, then the signal reconstructed via Eq. (2.22) will not only be missing high-frequency components, but will also have the energy in those components transferred to low frequencies that may not have been in the signal at all. This is a phenomenon called aliasing.

Example 2.24 Consider the function f defined by

c02e125

We can calculate f(t) = T x[/] by computing

c02e126

The last equality can be obtained by integration by parts (or by using your favorite computer algebra system). The plot of f is given in Figure 2.16.

Figure 2.16. Graph off.

c02f016

Figure 2.17. Graph of partial series in Sampling Theorem with í2 = 1.

c02f017

Since c02ie006 (λ) = 0 for |λ| > 1, the frequency Ω from the Sampling Theorem can be chosen to be any number that is greater than or equal to 1 With Ω = 1, we graph the partial sum of the first 30 terms in the series given in the Sampling Theorem in Figure 2.17; note that the two graphs are nearly identical.

2.5 THE UNCERTAINTY PRINCIPLE

In this section we present the Uncertainty Principle, which in words states that a function cannot simultaneously have restricted support in time as well as in frequency. To explain these ideas, we need a definition.

Definition 2.25 Suppose f is a function in L2(R). The dispersion of f about the point aR is the quantity

c02e127

The dispersion of f about a point a measures the deviation or spread of its graph from t = a. This dispersion will be small if the graph of f is concentrated near t = a, as in Figure 2.18 (with a = 0). The dispersion will be larger if the graph of f spreads out away from t = a as in Figure 2.19.

Another description of the dispersion is related to statistics. Think of the function c02ie039 as a probability density function (this nonnegative function has integral equal to one, which is the primary requirement for a probability density function). If a is the mean of this density, then Δa f is just the variance.

Figure 2.18. Small dispersion off5.

c02f018

Figure 2.19. Larger dispersion of f1.

c02f019

Applying the preceding definition of dispersion to the Fourier transform off gives

c02e128

By the Plancherai Theorem (Theorem 2.12), the denominators in the dispersions of f and c02ie006 are the same.

If the dispersion of c02ie006 about λ = α is small, then the frequency range of f is concentrated near λ = α.

Now we are ready to state the Uncertainty Principle.

Theorem 2.26 (Uncertainty Principle). Suppose f is a function in L2(R) which vanishes at +∞ and –∞. Then

(2.25)c02e129

for all points aR and α ∈ R.

One consequence of the Uncertainty Principle is that the dispersion of f about any a (i.e., Δa f) and the dispersion of the Fourier transform of f about any frequency α (i.e., Δαc02ie006) cannot simultaneously be small. The graph in Figure 2.18 offers an intuitive explanation. This graph is concentrated near t = 0, and therefore its dispersion about t = 0 is small. However, this function changes rapidly, and therefore it will havejarge-frequency components in its Fourier transform. Thus, the dispersion of c02ie006 about any frequency value will be large. This is illustrated by the wide spread of the graph of its Fourier transform, given in Figure 2.20.

For a more quantitative viewpoint, suppose

c02e130

Figure 2.20. Large-frequency dispersion of c02ie0365

c02f020

The graphs of fs for s = 5 and s = 1 are given in Figures 2.18 and 2.19, respectively. Note that as s increases, the exponent becomes more negative and therefore the dispersion of fs decreases (i.e., the graph of fs becomes more concentrated near the origin).

The Fourier transform of fs is

c02e131

(see exercise 6). Except for the constant in front, the Fourier transform of fs has the same general negative exponential form as does fs. Thus the graph of c02ie006s has the same general shape as does the graph of fs. There is one notable difference: The factor s appears in the denominator of the exponent in c02ie006s instead of the numerator ofthe exponent (as is the case for fs). Therefore, as s increases, the dispersion of c02ie006s also increases (instead of decreasing as it does for fs). In particular, it is not possible to choose a value of s that makes the dispersions of both fs and c02ie006s simultaneously small.

Proof of Uncertainty Principle. We first claim that the following identity holds:

(2.26)c02e132

Here, α and a are real constants. Written out in detail, the left side is

c02e133

After using the product rule for the first term and then simplifying, the result isf, which establishes Eq. (2.26).

Note that Eq. (2.26) remains valid after dividing both sides by the constant ||f|| = ||f||L2. Since the L2 norm of f/||f|| is one, we may as well assume from the start that ||f|| = 1 (just by relabeling f/||f|| as f.

Now take the L2 inner product of both sides of Eq. (2.26). The result is

(2.27)c02e134

Both terms on the left involve integrals (from –∞ to ∞). We use integration by parts on the first integral on the left and use the assumption that f (–∞) = f (∞) = 0. The result is

(2.28)c02e135

(the details of which we ask you to carry out in exercise 9).

From Eq. (2.28) and the triangle inequality, we obtain

c02e136

Now apply Schwarz’s inequality (see Theorem 0.11) to the preceding two inner products:

c02e137

Next, we apply the Plancherel formula (Theorem 2.12) and the fourth property of Theorem 2.6 to obtain

c02e139

Combining this equation with the previous inequality, we get

c02e140

Since ||f ||L2 = 1 = ||c02e006 ||L2, we have

c02e141

Therefore, squaring both sides of the preceding inequality yields

c02e142

which completes the proof of the Uncertainty Principle. (Note: This proof is based on that given for the quantum mechanical version by H. P. Robertson, Phys. Rev. 34, 163-164 (1929)).

EXERCISES

1. Let

c02e143

Show that

c02e144

where m is an integer and λ ≠ m. Hint: Sum the two identities

c02e147

Use this integral to show

c02e148

as indicated in Example 2.3.

2. Let

c02e149

Compute c02ie006 (λ) (i.e., provide the details for Example 2.4).

3. (Note: A computer algebra system would help for this problem). Let

c02e150

Show that f and f′ are continuous everywhere (pay particular attention to what happens at t = –2, – 1, L and 2). Draw a careful graph off. Compute its Fourier transform and show that

c02e151

Plot the Fourier transform over the interval –10 ≤ λ ≤ 10. Note that the Fourier transform decays like c02ie040 as λ c02ie004 ∞. Note that in exercise 1 the given function is discontinuous and c02ie006 (λ) decays like c02ie041 as λ c02ie004 ∞; in exercise 2, f is continuous, but not differentiable and c02ie006 (λ) decays like 1/λ2. Do you see a pattern?

4.Suppose f ∈ L2(–∞, ∞) is a real-valued, even function; show that c02ie006 is real-valued. Suppose fL2(–∞, ∞) is a real-valued, odd function; show that c02ie006 (λ) is pure imaginary (its real part is zero) for each λ.

5.Let

c02e152

Show that

c02e153

6. Let fs(x) = √sesx2. Show that

c02e154

Hint: After writing out the definition of the Fourier transform, complete the square in the exponent, then perform a change of variables in the resulting integral (a complex translation) and then use the fact that c02ie042.

7.Establish the parts of Theorem 2.6 that deal with the inverse Fourier transform. Establish the relationship between the Fourier transform and the Laplace transform given in the last part of this theorem.

8.Suppose L is a time invariant filter with a continuous response function h (see Theorem 2.19). Show that if L is causal then h(t) =0 for t < 0. Hint: Prove by contradiction—Assume h(t0) ≠ 0 for some t0 < 0; show that by choosing an appropriate f which is nonzero only on the interval [0, δ] for some very small δ > 0, that L(f)(t) > 0 for some t < 0 even though f(t) = 0 for all t < 0 (thus contradicting causality).

9. Using integration by parts, show that Eq. (2.28) follows from Eq. (2.27).

10. Let

c02e155

where A and a are parameters as in Example 2.21. Show that

c02e156

11. Consider the signal

c02e157

Filter this signal with the Butterworth filter (i.e., compute (f * h)(t)) for 0 ≤ t ≤ π. Try various values of A = α (starting with A = α = 10). Compare the filtered signal with the original signal.

12. Consider the filter given by f c02ie004 f * h, where

c02e158

Compute and graph c02ie033(λ) for 0 ≤ λ ≤ 20. for various values of d. Suppose you wish to use this filter to remove signal noise with frequency larger than 30 and retain frequencies in the range of 0 to 5 (cycles per 2π interval). What values of d would you use? Filter the signal

c02e159

with your chosen values of d and compare the graph of the filtered signal with the graph off.

13. The Sampling Theorem (Theorem 2.23) states that if the Fourier transform of a signal f is band-limited to the interval [–Ω, Ω], then the signal can be recovered by sampling the signal at a grid of time nodes separated by a time interval of π/Ω. Now suppose that the Fourier transform of a signal, f, vanishes outside the interval [ω1 ≤ λ ≤ ω2]; develop a formula analogous to Eq. (2.22) where the time interval of separation is 2π/(ω2ω1). Hint: Show that the Fourier transform of the signal

c02e160

is band-limited to the interval [–Ω, Ω] with Ω = (ω2ω1)/2 and then apply Theorem 2.23 to the signal g.

14. (Oversampling) This exercise develops a version of the sampling theorem which converges faster than the version given in Theorem 2.23. Complete the following outline.

(a) Suppose f is a band-limited signal with c02ie006 (λ) = 0 for |λ| ≥ Ω Fix a number a > 1. Repeat the proof of Theorem 2.23 to show that

c02e161

(b) Let c02ie043 be the function whose graph is given by Figure 2.21. Show that

c02e162

(c) Since c02ie044. Use Theorem 2.1, Theorem 2.6, and the expressions for c02ie006 and ga in parts a and b to show

(2.29)c02e163

Since ga(t) has a factor of t2 in the denominator, this expression for f(t) converges faster than the expression for f given in Theorem 2.23 (the nth tem behaves like l/n2 instead of l/n). The disadvantage of Eq. (2.29) is that the function is sampled on a grid of /(aΩ) which is a more frequent rate of sampling than the grid /Ω (since a > 1) used in Theorem 2.23. Thus there is a trade-off between the sample rate and the rate of convergence.

Figure 2.21. Graph of ga.

c02f021
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.147.76.89