B
(Figure 6.9). Here we can take the density f of a to be the derivative of a where it exists (which is everwhere except at A and B), and
,i
104
6. Integral Calculus
a(x) 1
_ _ _ _+-_-JLo A
I I I I I I LI- - ~ x
B
FIGURE 6.9
r---r----r f{x)
l/(B-A)
FIGURE 6.10
define f (arbitrarily) to be zero at A and B, giving
~X <
f (x) = { 0 'I B-A'
A or x > B, If A < x < B
(Figure 6.10). The previous theorem dealt with differentiation of an integral that has a variable interval of integration. It is also important to be able to differentiate functions of the form get)
= [f(t,x)dx,
where the variable t appears in the integrand, not in the interval of integration. It is natural to ask whether we can find g' (t) by interchanging the order of the differentiation and integration operations. Thus g'(t)
= ~[[f(t,X) dx] = [ ~f(t,x)dx,
6.3. Integration and Differentiation
105
where in the right-hand integral, x is held constant while differentiation is carried out with respect to t. Simple examples suggest that this is correct. Consider, for example, get)
= (I sin(2t + 3x) dx = [_~ cos(2t + 3X)]X=1 k 3 x~ 1 1 = - 3 cos(2t + 3) + 3 cos2t.
By direct differentiation, g/ (t) hand, ~
1 1
= ~ sin(2t + 3) -
-a sin(2t + 3x) dx =
o at
=
1
~ sin 2t. On the other
1
2 cos(2t + 3x)dx
0
]X=l
2 - sin(2t + 3x) [ 3 x=o
2
= - sin(2t + 3) 3 = g/(t)
2 - sin 2t, 3
as expected. The general theorem, which tells us that this process, called IIdifferentiation under the integral:' is legitimate, is as follows: Theorem 6.3.2 (Differentiation Under the Integral) Let 1 and J be any intervals. Let the real-valued function f(t, x) be such that (i) f(t, x) is defined for all t E J, X E I; (ii) f(t, x) is integrable with respect to x over I; (iii) For each t E J, :/(t, x) exists a.e. on I; (iv) For each closed subinterval J* C J, there exists a function A : I --+ IR such that A is integrable over I, and I ~f(t, x)1 < A(X) for all x E I and t E J*. Then for each t E J, a/atf(t, x) is integrable with respect to x over I, and
~ [ [f(t, x) ax] = [:/Ct, x) ax. Example 6-3-3: Consider the function get) = JoT( In(l + t cos x) dx, where -1 < t < 1. Note that since I cosxl < 1 for all x, we have that It cosxl < It I < 1
106
6. Integral Calculus
for all t E (-1, 1). Thus 1 + t cos x > 0 for all x and all t and so both In(1 + t cos x) and
a
-a In(1 + t cos x)
+ tcosxl
(-1, 1),
cosx
= 1 + t cosx
t are continuous for all x and all t E (-1,1). Take any closed subinterval [a, b] C max{la/, lb/}. We have that for all x and all t and so 11
E
(-1, I), and let k = E [a, b], It cosxl < k < I,
> l-Itcosxl > 1 - k >
o.
Thus cosx 1 <-I + t cos x - I - k
for all x and all t E [a, b]. Hence all the conditions of Theorem 6.3.2 are satisfied in this case, where f(t, x) = In(1 + t cos x), I = [0, JT], J = (-1, I), and A(X) = 1/(1 - k) for each J* = [a, b] C J. By Theorem 6.3.2 we have therefore 1C 1C a cos x ax g' (t) = -In(1 + t cosx) ax = o at 0 1 + t cos x
1
1
for all t
E
(-1, 1). Note first that g'(O)
1ft
E
=
1"
cosxdx = [sinx]~
= O.
(-1,1) is not zero, then 't g()
=
{1C ~ (1
Jo
l1
--
t
t
1C ( 1-
0
= ~ (JT _ [ t
X-I) ax
+ t cos 1 + t COSX
1) ax (0 -.Jt)tan(XI2))]X=1C) . x=o
1 + t cosx
2
.J1 -
arctan
1 - t2
t2
(The final equality is not obvious, but it can be verified using any reasonably comprehensive table of integrals.) Now, as x -* JT-, tan(xI2) -* 00, and so (1 - t) tan(x/2) ..:..-_;=:::::=:;=---..:.. -* Jl-t 2
.i
00
6.3. Integration and Differentiation
as x --+
j(-,
107
and thus arctan ((1- t)tan(XI2)) --+ j( .Jl - t 2 2
as x --+
j(-.
Also, arctan(O) = 0, and so () = t1(j( - --;.J=1j===:;::t 2
g /(t)
for t
f. 0, t
E
(-1,1). Thus we have finally
/ x> _ {
g( )-
Sinceg(O)
7(1 -
Jl1_t2)'
0,
if t E (-1, 1) and t if t = 0.
#
0,
= J;(lnl)dx = J; Odx = 0, we have that for alIt E (-1,1) get)
= get) -
g(O)
=
l'
g(x)dx
by the fundamental theorem of calculus; thus,
get)
(1 _ 1 ) dx. Jo x .Jl - x
= {t j(
2
Now (using a table of integrals) we have that
f (~ x
1 ) dx = In Ixl + In 1 + .Jl - x 2 x.Jl -x x = In
(note that 1 + .Jl - x2 >
get)
= j([ In(1 + -/1
x(1
+ .Jl - x2 ) x
for -1 < t < 1.
In(l
+C
+C
°for Ixl < 1). Hence, -
X2)]~ = j( (In(1 + -/1
- t 2) -In
and so finally
1"
2
+ tcosx)dx = :;rIn ( + ~l 1
2
t
)
2) ,
108
6. Integral Calculus
In the most general case, a function defined by an integral may have the variable appearing both in the limits of integration and in the integrand, for example get) =
i
t2
e'" ax.
Such a function can be differentiated by using a combination of the fundamental theorem of calculus and differentiation under the integral. Consider an integral of the form get) =
1
f(w(t), x) dx,
l(t)
where l(t) is the closed interval [u(t), vet)]. Here we assume that for some interval 10, u, v, w : 10 -* lR are differentiable functions of t such that for some interval h we have wet) E II for all t E 10 . We assume also that for each t E 10, few, x) satisfies the conditions of Theorem 6.3.2 for w ElI andx E l(t). Let h(u, v, w) = ~u,v]f(w, x) dx. Then get) = h(u(t), vet), w(t)) , so by the chain rule / g (t)
=
ah du au dt
+
ah dv av dt
ah dw
+ aw at
du
dv
= -few, u) dt + few, v) dt +
({
a ) dw awf(w,x)dx at'
J[U,V]
where the fundamental theorem of calculus has been used to find ah/ au and ah/ av, and differentiation under the integral to find ah/aw. In particular, if we take wet) = t for t E 10, then dw/dt = I, and we obtain the following result, often referred to as Leibniz's rule:
d[l
dt
[u(t),v(t)]
f(t, x) dx
]=
~ -f(t, u(t))-d t
1
+
~ + f(t, v(t))-d t
a
(6.1)
-f(t,x)dx. [u(t),v(t)] at
If f happens to be independent of t, then we obtain the important special case -d dt
[1
]
du f(x) dx = -f(U(t))-1 [u(t),v(t)] at
.i
dv . + f(v(t))-d
t
(6.2)
6.3. Integration and Differentiation
109
tz etx ax for t > 1. Here 1 = [1, (0), Consider the function get) = 0 and since wet) = t, we have also II = [1, (0). Take any t > I, and let [a, b] be any subinterval of[1, (0). Then
h
for all w obtain g' (t)
E
[a, b] and x 'd
E
[t, t 2 ], so we can apply Leibniz's rule to
1 + {[~'" t - 1" ~
d
= _et(t) -(t) + et(t2) -(t2 ) + dt
= -e"
dt
+ 2tet'
xe tx ax
t
e; ax} integrating by parts
etx tz} = -e~ + 2tei3 + { [te i3 - etz ]- [t2"1 2
3
= -et + 2tet + {i3 te = 3te i3 -
~
[et2" - t2"e~] } i3
e -
2e~ - t12 ( ei3 - etz) .
In this particular case we can check by evaluating get) directly: tx
get) = [ te
2
Jtt
= 1 ( ei3 - etz) ;
t
thus,
t
2 2 g ,(t) = 1 ( 3t 2 ei3 - 2te t ) - t12 ( ei3 - et ) 3 2 2 = 3te t - 2e t - -1 ( ei3 - et )
t2
as before. E~ple
6-3-4: Consider the differential equation
'
110
6. Integral Calculus
with initial conditions yeO) = CI, y/(O) = C2. Assume that g is continuous on the interval [0, (0). For all t > 0, define yet) =
CI
+ czt +
l'
(t - x)g(x)dx.
Clearly, yeO) = CI· The function few, x) = (w - x)g(x) certainly satisfies conditions (i) and (ii) of Theorem 6.3.2, for w E [0, (0) and x E [0, t] (t > 0). Further, for each t E [0, (0) we have that
a
awf(w, x) = Ig(x)/
(independent of w) is continuous and therefore integrable over [0, t]. It follows by Leibniz's rule that yet) = Cz + (t - t)g(t) + = Cz +
Thus y/(O) = calculus that
C2;
l'
l'
g(x)dx
g(x) dx.
also, we have by the fundamental theorem of
y" (t) = get), and so the function yet) defined above is the solution of the given initial value problem. As a final comment on the relationship between the integral and the derivative, we point out a serious gap in the Lebesgue theory. Recall that an antiderivative (sometimes called an indefinite integral) of a function f is a function F such that F' = f. It turns out that there exist functions that are lIintegrable" in the sense of having an antiderivative at all points of a certain interval but are not Lebesgue integrable on that interval. An example of such a function is given in Section 10.1, where we revisit this matter.
Exercises 6-3: 1. Use L'Hospital's rule to find erf(t) (a) lim and (b) t---+O+
t
, i
lim t erfc(t).
t---+oo
6.3. Integration and Differentiation
III
2. By writing erf(x) as 1 x erf(x) and integrating by parts, show that
(t
Jo
erf(x) dx
=t
1
erf(t) - ,ji(1 - e-
t2 ).
3. Use a table of integrals to show that
(Jr __ 1 _ dx _
---;:::;;::]'(===
Jo
t - cosx - Jt 2 - 1 for all t > 1. By differentiating both sides of this equation with respect to t, evaluate
~
{Jr
Jo
1
(t - cos X)2 dx.
4. Given thatg(t) = J~ sin(x-t)dx, findg'(t) by using Leibniz's rule. Check by evaluating get) directly and then differentiating. 5. Find g/(t) if get)
=
j
t 21
-
x
t
sin(tx) dx,
where t > 1. 6. Assuming that g is continuous on the interval [0, (0), show that the function yet)
1
1
(t
= k C2 sin(kt) + CI cos(kt) + k Jo
g(x) sin {k(t - x)} dx,
for t > 0, satisfies the differential equation
d2y dt 2
+ ~y =
get),
where k > 0, together with the initial conditions yeO) = CI, y/(O) = C2. 7. Assuming that g is continuous on the interval [0, (0), show that the function yet) = -1
n!
I
t
(t - x)ng(x) dx
0
satisfies the differential equation dn+ly dt n+l = get) together with the initial conditions yeO) = 0, yCI)(O) = 0, ... ,ycn)(O)
= O.
Double and Repeated Integrals
CHAPTER
Lebesgue-Stieltjes integrals of functions of more than one variable can be defined using the same approach as was used in Section 4.5 for functions of one variable. For the sake of simplicity we will discuss only functions of two variables. The process for functions of more than two variables is completely analogous.
7.1
Measure of a Rectangle
We define a rectangle to be a set ofthe form II x 12 C JR2, where II and h are intervals. For monotone increasing functions aI, a2 : JR --+ JR we define the al x a2-measure of II x h, denoted by fJ-al xa2 (II X 12 ), by fJ-al xa2 (II
X h)
= fJ-al (II) X fJ-a2 (h).
For example, if al and a2 are the functions defined in Exercises 4-1, problems 1 and 2, respectively, then fJ-al
((0, 1))
fJ-a2 ((0,
= 1-
1)) = 0,
e- I ,
=3fJ-a2 ([0, 1)) = I,
fJ-al ([0,
1))
e- I ,
113
114
7. Double and Repeated Integrals
and therefore JL CX I XCX 2((O, 1) x [0,1)) = 1 - e- l , JLcxIXCX2([O, 1) x [0,1)) = 3 - e- l .
JL CX I XCX 2((O, 1) x (0,1)) = 0,
7.2
Simple Sets and Simple Functions in Two Dimensions
A simple set in]R2 is a subset of]R2 that can be expressed as the union of a finite collection of disjoint rectangles. Just as for simple sets in ]R, we can define the measure of a simple set in ]R2. If aI, a2 : ]R --+ ]R are monotone increasing functions and S is a simple set of the form m
S = U(Ilj
X
hj),
j=l
where h,l x 12,1, h,2 X h,2, ... , 11,m X 12,m are pairwise disjoint rectangles, then the al x a2-measure of S is defined by m
JLcxIXCX2(S) = LJLcxl xcx2(I1j x h,j)' j=l
The properties of simple sets in ]R2 and their measures are the same as those described in Section 5-3 for simple sets in ]R. We can now define simple functions of two variables by analogy with step functions of one variable (see Sections 2-5 and 4-4). We could continue to use the term "step functions:' but customary usage restricts this term to functions of one variable. • A function 8 : ]R2 --+ ]R is a simple function if there is a simple set n
S = U(11 ,}0 x h ,}0) j=l
8(x,y) = {
0,
if (x, y)
E
]R2 - S.
7.4. Repeated Integrals and Fubini's Theorem
115
The set S is called the support of8. The properties of step functions given in Section 2-5 carry over without difficulty to simple functions. If aI, a2 : lR -+ lR are monotone increasing functions, we define the generalized "volume" A CX1 xCXz (8) in a way exactly analogous to the definition of A cx (8) in Section 4-4.
7.3
The Lebesgue-Stieltjes Double Integral
Let S be a subset of lR2 and let f : S -+ lR be a function. We extend the definition of f to lR2 by defining f(x, y) to be zero if (x, y) E lR2 - S. Let aI, a2 : lR -+ lR be monotone increasing functions. The Lebesgue-Stieltjes double integral of f, denoted by
L2 f f de
IJlj
x 1Jl2),
is defined by a process that is almost word-for-word the same as that used for the single-variable integral in Section 4-5. The only change is that a-summable step functions on I are replaced by al x a2summable simple functions on ]R2. All the elementary properties analogous to those proved in Sections 5-1 and 5-2 carry over and are proved in the same way, and the same goes for the convergence theorems of Section 5-3 and the definitions of measurable functions and measurable sets given in Section 5-4. In practice, the evaluation of double integrals is invariably done, as in elementary calculus, by converting them to repeated integrals.
7.4
Repeated Integrals and Fubini's Theorem
Let f : lR2 -+ lR be a function. For any y E lR we define the singlevariable function f(·, y) : x -+ f(x, y), and for any x E lR we likewise define the function f(x,·) : y -+ f(x, y). Let aI, a2 be monotone
116
7. Double and Repeated Integrals
increasing functions. If for eachy E lR, fIRf("y)001 exists, then this defines a function 12 : y --+ f IR f(·,y)OO 1. If fIR 12 002 exists, we call this a repeated integral of f and write it as fIR fIR f 001 da 2. If for each x E lR, fIRf(x,') 002 exists, we define fI : x --+ fIRf(x,') 002, and if fIR fI 001 exists, it gives us the repeated integral of f with the opposite order of integration, written fIR fIR f 002 001. In most cases calculation shows that fIR fIR f 001 002 and fIR fIR f 002 001 have the same value, but this is not always the case. Consider, for example, the improper Riemann repeated integrals 1 1 1 11 ---.-;:.X - Y dx dy and X - Y 3 dy dx. 0 0 (x + y) 1o 0 (x + y)3
11
We have
1 11 1 o
0
Y (x + y)3 dxdy X -
111(1 1 = + 0
-
[1 1 + (1 1 + 1
+
Y
JX=l dy
(x + y? x=o
(x
y)
o
1
Y + +--Y (1 + y? y y
1
-1
0
=
(x
0
2Y ) y? - (x + y)3 dxdy
1
1
=
o (1 1
+ y?
J1
1 1)
~
dy 1
= [ 1+yo=-2'
However, 1 11 ---:X - Y dydx = 1 o 0 (x + y)3
11 + 11 1 + 1
0
1
0
Y - x dxdy (interchanging x andy)
(x 1
= -
o
0
y)3
•
x-y dxdy (x y)3
1
2'
and so the two repeated integrals have different values. It turns out, though we shall not prove it, that this cannot happen if either repeated integral is absolutely convergent. We can easily verify that this condition does not hold in our example. We
7.4. Repeated Integrals and Fubini's Theorem
117
y 1 r----."
y>x y<x --f---~~X
o
1
FIGURE 7.1
investigate
1 11 1 + o
0
Ix -YI ---dxdy (x y)3
by splitting the region ofintegration into two parts, one where y < x and one where y > x (cf. Figure 7.1). Then,
118
7. Double and Repeated Integrals
y
y=x 2 ----\--
FIGURE 7.2
and so this integral (and likewise J; Jo (~.;~13 dy dx) does not converge. The fundamental theorem that relates double and repeated integrals using the absolute convergence condition is called Fubini's theorem: l
Theorem 7.4.1 (Fubini's Theorem) If aI, a2 : ]R -+ ]R are monotone increasing functions and f : ]R2 -+ al x a2-measurable, then the existence ofanyone ofthe integrals
LJ Ifl
dCetl
x
et2), LL Ifl det det2, LL Ifl det2det l
]R is
l,
implies the existence and equality ofthe integrals
LJfdCetl x et2), LLf detldet2 ' LLf det2detl . In practice, the functions that arise are almost always measurable, so Fubini's theorem justifies the use ofrepeated integrals to evaluate double integrals, provided that one of the repeated integrals is absolutely convergent. The details can be very messy, so we will confine ourselves to one example. Example 7-4-1: Let 8 = {(x, y) : < X <2, f : ]R2 -+ ]R be defined by
°
° < Y < x} (see Figure 7.2) and let
x ) = { 1 +xy, if (x,y) E 8, f( ,y 0, if(x,y)E]R2-8.
7.4. Repeated Integrals and Fubini's Theorem
119
Let CXl and CX2 be the functions defined in Exercises 4-1, problems 1 and 2, respectively. We will evaluate fIR2 f f d(CXl x CX2) by evaluating the repeated integral fIR fIRf da l da 2 , and check by evaluating fIR fIR f da 2 da l . Since the integrand is nonnegative, our procedure will be justified (by Fubini's theorem) ifone ofthe repeated integrals exists.
We have that • I
f(·,Y)(x) =
{
0,
+ xy,
if 0 < y < x < 2, otherwise;
hence, h(}j) = LfCo,y)dJY.1
={
~YI2](1
+ xy) dab
0,
if 0
1 (CXl(O+) -CXl(O-)) + fo21~(3 - e-X)dx,
ify
= 0,
if 0 < y < 2, otherwise ify = 0,
0,
otherwise
3 -e-2 ,
ify = 0,
0,
otherwise.
120
7. Double and Repeated Integrals
Therefore,
LLf
da, da 2
=
Lh
da 2
= (3 - e-2)(CX2(0+) - CX2(0-))
+
1
d
1 o
(-e- 2 - 3ye- 2 + e-Y + y2 e - y
+( -4e-2 + 3e- 1 )(CX 2(l +) + ],
2
+ ye-Y)-(l )dy dy
CX2(l-))
(-e- 2 - 3ye- 2 + e-Y + y2 e- y
1
+ O(cx2(2+) = (3 - e- 2 )(1)
= 3 + ge- 1
cx2(2-))
+ (-4e- 2 + 3e- 1 )(3)
13e-2 .
-
We have that
I [(x, .)(y) =
{
0,
+ xy,
if
°< y < x < 2,
otherwise;
hence,
ftCx) =
L
f(x, 0) da 2
= { ~O,X](l 0,
+ xy) 002,
ifO<x<2,
otherwise
d
+ ye-Y)-(4)dy dy
7.4. Repea ted Integ rals and Fubin i's Theo rem
1(CX 2(0+) - CX2(0-))
+ 10 (1 +XY)~(1)dy,
1 (CX2(0+) - CX2(0-))
+ 101 (1 +xY)~(1)dy
+ (1 + X)(CX2(l +) -
cx2(1-)) + It(l
x
121
if 0 < X < I,
+ XY)~(4) dy, . ifl< x<2 ,
1(CX 2(0+) - CX2(0-))
+ fa1 (1 + xy) ~ (1) dy
+ (1• + X)( CX2(l +) -
CX2(l-))
+ (1 + 2x)(CX2(2+) 0,
Ther efore ,
cx2(2-)),
other wise
I,
ifO< x< I,
4+3 x,
if 1 < x < 2,
6 +7x,
ifx = 2,
0,
other wise.
-
d (4) dy + h2 (1 + XY)dy
if X = 2,
122
7. Double and Repeated Integrals
Exercise 7-4: Let 8 = {(x, y) : defined by
°<
t(x
X
<2,
°< Y < 2 - x} and let t :
]R2
--+ ]R be
) = {csin y , if (x,y) E 8, ,y 0, if (x,y) E]R2 - 8.
Let (Xl and (X2 be the functions defined in Exercises 4-1, problems 1 and 2, respectively. Verify that
LLf
da l da 2 =
LLf
da2 da l.
The Lebesgue Spaces V CHAPTER
There are many mathematical problems for which the solution is a function of some kind, and it is often both possible and convenient to specify in advance the set of functions within which the solution is to be sought. For example, the solution to a first-order differential equation might be specified as being differentiable on the whole real line. The set of functions differentiable on the whole real line has the useful property that sums and constant multiples of functions in the set are also in the set. In fact, this set of functions has the structure of a vector space, where the I/vectors" are functions. Beyond the algebraic properties associated with vector spaces, many problems are solved by use ofseries or sequences offunctions, and it is desirable that any I/limits" also be in the set. Some ofthe most useful sets of functions have this property. We have seen in Section 3-3 that the limit ofa sequence ofRiemann integrable functions does not necessarily yield a Riemann integrable function, and this signals that the sets defined using the Riemann integral may not be suitable for many applications. In contrast, sets defined using the Lebesgue integral have the desirable I/limit properties:' There are a number of sets of functions that are vector spaces that are of importance in subjects such as differential and integral equations, real and complex function theory, and probability theory,
123
124
8. The Lebesgue Spaces IJ'
along with the fields of applied mathematics where these subjects playa significant role. Some of the most important of these function sets make use of the Lebesgue integral in their definitions, and so it is appropriate to discuss them here. In this chapter we aim to give the reader an overview of some of these function sets. We neither go into all the technical details nor attempt a comprehensive survey. References are given where more detail can be found if desired.
8.1
Normed Spaces
The reader has probably encountered the concept of a finitedimensional vector space. These spaces are modeled after the set of vectors in lR.n . Vector spaces, however, can be defined more generally and need not be finite-dimensional. Indeed, most the vector spaces of interest in analysis are not finite-dimensional. Recall that a vector space is a nonempty set X equipped with the operations of addition 1+' and scalar multiplication. For any elements f,g, h in X and any scalars a, fJ these operations have the following properties: (i) f+g EX; (ii) f + g = g + f; (iii) f + (g + h) = Cf + g) + h; (iv) there is a unique element 0 (called zero) inX such that f +0 = f for allf EX; (v) for each element f E X there is a unique element (-f) EX such that f + (-f) = 0; (vi) af EX; (vii) aCf + g) = af + ag; (viii) (a + fJ)f = af + fJf; (ix) (afJ)f = a(fJf); (x) 1· f = f· For our purposes, the scalars will be either real or complex numbers. We shall use the term complex vector space if the scalars are complex numbers when there is some danger of confusion.
8.1. Normed Spaces
125
Example 8-1-1: The set of vectors {(Xl, xz, ... ,xn ) : Xk E JR., k = 1,2, ... , n} is denoted by JR.n. Let x = (Xl, XZ, ... ,xn) and y = (YI, Yz, ... , Yn) be vectors in JR.n. If addition is defined by
x
+Y=
(Xl
+ YI, Xz + Yz,···, Xn + Yn)
and scalar multiplication by ax = (axI, axz, ... , axn )
for any a E JR.,. then JR.n is a vector space. Similarly, the set Cn = {(ZI,ZZ, ... ,zn): Zk E C, k = 1,2, ... , n} of complex vectors is a complex vector space when addition is defined by z
+W =
(ZI
+ WI, Zz + Wz, .. ·, Zn + wn)
for any vectors z = (ZI, Zz, ... ,zn), multiplication by
W
= (WI, Wz, ... , wn), and scalar
where a E C. The vector spaces JR.n and cn are essentially the prototypes for more abstract vector spaces.
Example 8·1·2: Let C[a, b] denote the set of all functions f : [a, b] ~ JR. that are continuous on the interval [a, b]. If for any f, g E C[a, b], addition is defined by
if + g)(x) = f(x) + g(X) , and scalar multiplication by (af)(x) = af(x)
for a
E
JR., then it is not difficult to see that C[a, b] is a vector space.
Example 8·1·3: Let.e I denote the set of sequences {an} in JR. such that the series L~I Ian I is convergent, and define addition so that for any two elements A = {an}, B = {b n}, A
+B =
{an
+ bn},
126
8. The Lebesgue Spaces IJ'
and scalar multiplication so that
Then .e I is also a vector space. The above examples show that the elements in different vector spaces can be very different in nature. More importantly, however, there is a significant difference between a vector space such as lR.n and one such as C[a, b] having to do with Iidimension." The space ~n has a basis: Any set of n linearly independent vectors in lR. n such as el = (1,0, ... ,0), e2 = (0, 1, ... ,0) ... , en = (0,0, ... ,1) forms a basis. The concept of dimension is tied to the number of elements in a basis for spaces such as lR.n , but what would be a basis for a space like C[a, b]? In order to make some progress on this question we need first to define what is meant by a linearly independent set when the set itself might contain an infinite number of elements. We say that a set is linearly independent if every finite subset is linearly independent; otherwise, it is called linearly dependent. If there exists a positive integer n such that a vector space X has n linearly independent vectors but any set of n + 1 vectors is linearly dependent, then X is called finite-dimensional. If no such integer exists, then X is called infinite-dimensional. We will return to the question of bases for certain infinite-dimensional vector spaces in Chapter 9. A subspace of a vector space X is a subset of X that is itself a vector space under the same operations of addition and scalar multiplication. For example, the set of functions f : [a, b] ~ lR. such that f is differentiable on [a, b] is a subspace of C[a, b]. Given any vectors Xl, X2, ••• ,Xn in a vector space X, a subspace can always be • formed by generating all the linear combinations involving the Xk, i.e., all the vectors of the form C¥IXI + C¥2X2 + ... + C¥nXn, where the C¥k'S are scalars. Given any finite set 8 C X the subspace of X formed in this manner is called the span of 8 and denoted by [8]. If 8 ex has an infinite number of elements, then the span of 8 is defined to be the set of all finite linear combinations of elements of 8. Vector spaces of functions such as C[a, b] are often referred to simply as function spaces. After the next section we shall be concerned almost exclusively with function spaces, and to avoid rep-
8.1. Normed Spaces
127
etition we shall agree here that for any function space the operations of addition and scalar multiplication will be defined pointwise as was done for the space C[a, b] in Example 8-1-2. Vector spaces are purely algebraic objects, and in order to do any analysis we need to further specialize. In particular, basic concepts such as convergence require some means of measuring the distance" between objects in the vector space. This leads us to the concept of a norm. A norm on a vector space X is a real-valued function on X whose value at f E X is denoted by IIf II and that has the fonowing properties: II
(i) IIfll > 0; (ii) Ilf II
= 0 if and only iff = 0;
(iii) lIafll = laillfll; (iv) IIf + gil < IIf II
+ IIg II
(the triangle inequality).
Here, f and g are arbitrary elements in X, and a is any scalar. A vector space X equipped with a norm II . II is called a normed vector space. E~ple 8-1-4: For any x E lR.n let
II . II e be defined by
IIxll e
= {(xr + (x2i + ... + (xni} 1/2 .
Then II . II e is a norm on lR.n . This function is called the Euclidean norm on lR.n . Another norm on lR.n is given by
E~ple 8·1·5:
The function II . II 00 given by IIflloo = sup If (x) I xE[a,b]
is well-defined for any f E C[a, b], and it can be shown that II ·1100 is a norm for C[a, b]. Alternatively, since any functionf in this vector space is continuous, the function If Iis Riemann integrable, and thus
------------.-,----r-------------.- .
128
8. The Lebesgue Spaces Ll
y I I
g
I
If I
- - - - ~ - - lif- glloo I I I I I
.. X
--r----'-a-----------..Lb--~
FIGURE 8.1
the function II . IIR given by
IIfliR =
t
If(x)1 ax
is well-defined on C[a, b]. It is left as an exercise to show that II . IIR is a norm on C[a, b]. The above examples indicate that a given vector space may have several norms leading to different normed vector spaces. For this reason, the notation (X, 11·11) is often used to denote the vector space X equipped with the norm II . II. Once a vector space is equipped with a norm II . II, a generalized distance function (called the metric induced by the norm II . II) can be readily defined. The distance d(f, g) of an element f E X from another element g E X is defined to be d(f,g) = Ilf - gil.
The distance function for the normed vector space (lR. n, II . lie) corresponds to the ordinary notion of Euclidean distance. The distance function for the normed vector space (C[a, b], II . 1100) measures the maximum vertical separation of the graph of f from the graph of g (Figure 8.1). Convergence can be defined for sequences in a normed vector space in a manner that mimics the familiar definition in real anal-
8.1. Normed Spaces
129
ysis. Let (X, II - II) be a normed vector space and let {fn} denote an infinite sequence in X. The sequence {fn} is said to converge in the norm if there exists an f E X such that for every E > 0 an integer N can be found with the property that IIfn - f II < E whenever n > N. The element f is called the limit of the sequence {fn}, and the relationship is denoted by limn - HXl fn = f or simply fn ~ f. Note that convergence depends on the choice of norm: A sequence may converge in one norm and diverge in another. Note also that the limit f must also be an element inX. In a similal spirit, we can define Cauchy sequences for a normed vector space. A sequence ifn} in X is a Cauchy sequence (in the norm II - II) if for any E > 0 there is an integer N such that
IIfm - fnll <
E
whenever m > Nand n > N. Cauchy sequences playa vital role in the theory of normed vector spaces. As with convergence, a sequence ifn} inX maybe a Cauchy sequence for one choice of norm but not a Cauchy sequence for another choice. It may be possible to define any number of norms on a given vector space X. Two different norms, however, may yield exactly the same results concerning convergence and Cauchy sequences. Two norms II -lla and II -lib on a vector space X are said to be equivalent if there exists positive numbers a and f3 such that for all f EX,
allflla < Ilfllb < Pllflla. If the norms II ·lIa and 1I·llb are equivalent, then it is straightforward to show that convergence in one norm implies convergence in the other, and that the set of Cauchy sequences in (X, II . lIa) is the same as the set of Cauchy sequences in (X, II . lib). Equivalent norms lead to the same analytical results. Identifying norms as equivalent can be an involved process. In finite-dimensional vector spaces, however, the situation is simple: All norms defined on a finite-dimensional vector space are equivalent. Thus the two norms defined in Example 8-1-4 are equivalent. The situation is different for infinite-dimensional spaces. For example, the norms II·IIR and 11·1100 defined on the space C[a, b] in Example 8-1-5 are not equivalent. We elucidate further this comment in the next section, when we discuss completeness.
130
8. The Lebesgue Spaces Ll
Exercises 8-1: 1. Let Q denote the set of rational numbers. Show that Q is a vector space, provided that the scalar field is the rational numbers. 2.
ea) Prove that the function II ·IIR defined on C[a, b] in Example 8-1-5 satisfies the conditions of a norm. (b) Suppose that the set C[a, b] is extended to R[a, b], the set of all functioris f : [a, b] ~ lR. such that IfI is Riemann integrable. Show that II . IIR is not a norm on R[a, b].
3. Let Cn[a, b] denote the set of functions f : [a, b] ~ lR. with at least n continuous derivatives on [a, b]. Show that the functions 11·111,00 and II . 111,1 defined by IIfll1,oo = sup Ife x ) I + sup I['ex)!, xE[a,b] xE[a,b] IIfl11,l =
1
[a,b]
{lfex)1
+ l['ex)l}
dx,
are norms on the space C 1 [a, b]. 4. The number ~ can be approximated by a sequence {an} of rational numbers. Let So = {I, 2, ... , g} and choose an as the largest element of So such that a5 < 2. Since 12 = 1 < 22 = 4, we have an = 1. Let Sl = {1.1, 1.2, ... , log} and choose a1 as the largest element of Sl such that < 2. Thus, a1 = 1.4. Let S2 = {1.41, 1.42, ... , 1.49} and choose a2 as the largest element in S2 such that a~ < 2. This gives a2 = 1.41. Following this procedure for the general n show that the resulting sequence {an} must be a Cauchy sequence.
ar
5. Suppose that II . lIa and II . lib are equivalent norms for the vector space X. Prove that the condition allflla < IIfllb < ,Bllflla,
where ex and ,B are positive numbers, implies that there exist positive numbers y and 8 such that yllfllb < IIflla < 811fllb.
8.2. Banach Spaces
8.2
131
Banach Spaces
The definitions for convergence and Cauchy sequences for the normed vector space (lR. n, II . lie) correspond to the familiar definitions given in real analysis. Various results such as the uniqueness of the limit can be proved in general normed vector spaces by essentially the same techniques used to prove analogous results in real analysis. The space (lR. n, II . lie), however, has a special property not inherent in the definition of a normed vector space. It is well known that a sequen<;e in (lR., II . lie) converges if and only if it is a Cauchy sequence. This result does not extend to the general normed vector space. It is left as an exercise to show that every convergent sequence in a normed vector space must be a Cauchy sequence. The converse is not true. The following examples illustrate the problem for finiteand infinite-dimensional spaces. E~ple
8-2-1: The set Q of rational numbers, equipped with the Euclidean norm II . II e restricted to the rational numbers, is a normed vector space, provided that the scalar field is the rational numbers (Exercises 8-1, No.1). It is well known that the number ~ is not a rational number. The sequence {an} defined in Exercises 8-1, No.4, is a Cauchy sequence, which in the normed vector space (lR., 1I·lIe) can be shown to converge to the limit ~. This sequence is also a Cauchy sequence in Q, but it cannot converge to an element in Q and is therefore not convergent in Q. E~ple
8-2-2: Consider the normed vector space (C[ -I, 1], II ·11 R) and the sequence of functions {fn} defined by
fn(x) = {
if -1 -< x -< a, if a < x < 1/2 n , if 1/2 n < x < 1.
The function in is depicted in Figure 8.2, and it is clear that in C[ -I, 1] for all positive integers n.
E
132
8. The Lebesgue Spaces Ll
fn(x)
---'-:----t-~--_.x
-1
FIGURE 8.2
Now, Ifni = fn, and IIfnllR
=
L:
fnCx)dx
= 1 + 2;+1;
therefore, for any m > N, n > N, we have that 1 IIfm - fnllR = 2m+ 1
-
1 1 2n+ 1 < 2N
•
Given any E > 0, any positive integer N such that N > -log2 E suffices to ensure that IIfm - fnllR < E whenever n > Nand m > N. The sequence is thus a Cauchy sequence. It is clear geometrically that fn approaches the function f defined by f(x) = {I, if -1 < x < 0, 0, if 0 < x < 1.
Indeed, for any fixed Xo E [-1, 1] the sequence of real numbers {fn(XO)} converges to f(xo) (in the II . lie norm), i.e., ifn} is pointwise convergent to f. The function f, however, cannot be a limit in C[-1,1] for {fn} because f E C[-1,1]. Although the pointwise limit function f cannot be a limit in C[-I, 1] for {fn}, this does not extinguish the possibility that there is some other function g E C[-I, 1] that is the limit. We will show now that no such g exists. Suppose, for a contradiction, that fn ~ g E C[ -1,1] in the II . IIR norm. Then I
=
L
If(x) - g(x)1 dx
=
L
lif(x) - fnCx))
+ ifn(x) -
g(x)) I dx
8.2. Banach Spaces
<
=
L
Ifn(x) - f(x) I ax +
2n~1 +
= If
L
133
L
Ifn(x) - g(x)1 ax
Ifn(x) - g(x)1 ax
+ I g•
The quantity If can be made arbitrarily small by choosing n sufficiently large. By hypothesis fn ~ g in the II . IIR norm, so that I g can be made arbitrarily small for n large. Now, I < If + I g , and I does not depend op n. This implies that I = o. Since g E C[ -I, 1] and f is continuous on the intervals [-1, 0), (0, 1], the condition I = 0 implies that f = g for all x E [-1,0) and x E (0,1]. Therefore, limx~o- g(x) = 1 and limx~o+ g(x) = 0, so that limx~og(x) does not exist, contradicting the assumption that g is continuous on the interval [-1, 1]. The Cauchy sequence lfn} therefore does not converge in the II . IIR norm. Suppose that the space C[ -I, 1] is equipped with the II . II 00 norm defined in Example 8-1-5 instead ofthe II·IIR norm. If {h n } is a Cauchy sequence in the 11·1100 norm, then {h n } converges pointwise to some limit function h. The difference here is that because {h n } is a Cauchy sequence in the II . II 00 norm, it can be shown that {h n } in fact converges uniformly to h, and a standard result in real analysis implies that a uniformly convergent sequence of continuous functions cC?nverges to a continuous function. In other words, the limit function h must be in C[ -I, 1]. It is left as an exercise to verify that the sequence lfn} defined in this example is not a Cauchy sequence in the II ·1100 norm. A normed vector space is called complete if every Cauchy sequence in the vector space converges. Complete normed vector spaces are called Banach spaces. In finite-dimellsional vector spaces, completeness in one norm implies completeness in any norm, since all norms are equivalent. Thus, spaces such as (lR. n , 1I·lIe) and (lR. n , II . liT) are Banach spaces. On the other hand, Example 82-1 shows that no norm on the vector space Q can be defined so that the resulting space is a Banach space. For finite-dimensional vector spaces, completeness depends entirely OIl; tpe vector space; for infinite-dimensional vector spaces Example 8.-2-2 shows that
134
8. The Lebesgue Spaces Y
completeness depends also on the choice of norm. The space (C[ -I, 1], 11·1100) is a Banach space, whereas, the space (C[ -I, 1], II·IIR) is not. Ifthe norms 1I·lla and II· lib are equivalent, then the corresponding normed vector spaces are either both Banach or both incomplete, since the set of Cauchy sequences is the same for each space and convergence in one norm implies convergence in the other. The two norms II . IIR and II . II 00 on C[ -I, 1] are evidently not equivalent. Given a set 8 C X, if (X, II . II) is a normed vector space, a new subset 8 called the closure can be formed by requiring that f E 8 if and only if there is a sequence lfn} of vectors in 8 (not necessarily distinct) such that fn -+ f (in the norm II . II). If 8 = 8, then 8 is called a closed set. The subset 8 is called complete if every Cauchy sequence in 8 converges to a vector in 8. For Banach spaces the concepts of completeness and closure are linked by the following fundamental result: Theorem 8.2.1 Let (X, II . II) be a Banach space and 8 C X. The set 8 is closed if and only if 8 is complete. In particular, if 8 is a subspace, then it forms a normed vector space (8, 11·11) in its own right, and if it is closed, the above theorem indicates that (8, 11·11) is a Banach space. This observation leads to the following corollary: Corollary 8.2.2 Let (X, II . II) be a Banach space and 8 Banach space.
C
X. Then ([S],
II . II) is also a
Exercises 8-2: 1. Let (X,
II . II) be a normed vector space and suppose that
{an} is
a sequence in X that converges in the norm to some element a EX. Prove that {an} must be a Cauchy sequence. 2. Let lfn} be the sequence defined in Example 8-2-2. Prove that lfn} is not a Cauchy sequence in the normed vector space (C[-I, 1], II . 1100).
i---·-----------·--
_ _ _ _ _ _ _·_ _· - - - , - - " i ' T i
8.3. Completion of Spaces
8.3
135
Completion of Spaces
If a normed vector space is not complete, it is possible to Ilexpand" the vector space and suitably redefine the norm so that the resulting space is complete. In this section we discuss this process, and in the next section we apply the result to get a Banach space with a norm defined by the Lebesgue integral. Before we discuss the main result, however, we need to introduce a few terms. In functional analysis, a function T : X -+ Ythat maps a normed vector space X to a normed vector space Y is called an operator. •
E~ple
8-3-1: Let the set CI[a, b] and the norm II . 111,00 be as defined in Exercises 8-1, No.3. Every function in CI[a, b] has a continuous derivative. 1fT is the operator corresponding to differentiation d/ dx, then T maps every element in CI[a, b] to a unique function continuous on the interval [a, b]. Thus T maps CI[a, b] into C[a, b]. The definition of an operator is not norm dependent, but for illustration, we can regard T as mapping the space (Cl[a, b], 11·111,00) into the space (C[a, b], 11·1100).
Much of functional analysis is concerned with the study of operators. For a general discussion the reader can consult [25]. Here, we limit ourselves to a special type of operator that preserves norm. An operator T from the normed vector space (X, II . IIx) into the normed vector space (Y, II . II y) is called an isometry if for all Xl, X2 E X IIXI - x211x =
II TxI
- TX2I1y·
In essence, an isometry preserves the distance between points in X when they are mapped to Y. An isometry must be one-to-one. If there exists an isometry T : X -+ Y such that T is onto (so the inverse T- I : Y -+ X exists), then the normed spaces (X, II· IIx) a.nd (Y, II . II y) are called isometric. If two spaces are isometric, then completeness in one space implies completeness in the other. E~ple
8-3-2: The differentiation operator of Example 8-3-1 is clearly not an isometry from (Cl[a, b], II . 111,00) into the space (C[a, b], II . 1100)' since in
136
8. The Lebesgue Spaces LJ
general,
Ilf - glll,co
= sup If(x) - g(x) I + sup If'(x) - g'(x) I
xE[a,b]
xE[a,b]
> sup If'(x) - g'(x) I = xE[a,b]
11f' - g'llco.
Example 8-3-3: Let H(D(c; r)) denote the set of all functions holomorphic (analytic) in the closed disk D(c; r) = {z E C : Iz - cl < r}, r > O. This set forms a complex vector space, and the function II . lie defined by
IIflle = sup If(z) I zED(e;r)
provides a norm for the space. In fact, it can be shown that (H(D(c;r)),II·lIe)isaBanachspace.LetT¢,bbetheoperatormap ping H(D(a; r)) to H(D(a - b; r)) defined by T¢,bf = ei¢f(z - b),
where ¢ E lR is a constant. The operator T¢,b is a one-to-one and onto mapping from (H(D(a; r)), II . lIa) to (H(D(a - b; r)), II . II a-b), and since sup leitPf(z - b) - eitPg(z - b)1 zED(a-b;r) sup If(z - b) - g(z - b)1 ZED(a-b;r) = sup If(z) - g(z) I = IIf - gila, ZED(a;r)
IIT¢,bf - T¢,bglla-b =
the operator is also an isometry. The two normed spaces are thus isometric. Given an incomplete normed vector space (X, 1\.1\), it is natural to enquire whether the space can be made complete by enlarging the vector space and extending the definition of the norm to cope with the new elements. The paradigm for this process is the completion of the rational number system Q to form the real number system R. This example has two features, which, loosely speaking, are as follows:
8.3. Completion of Spaces
137
(i) the completion does not change the value of the norm where it was originally defined, Le., IIrll e = Ilrll e , for any rational number r; (ii) the set Q is dense in the set lR (cf. Section 1-1). The first feature is obviously desirable: We wish to preserve as much as possible the original normed vector space, and any extended definition of the norm should not change the value of the norm at points in the original space. The second feature expresses the fact that the extension of tne set Q to the set lR is a IIminimal" one: Every element added to Q is required for the completion. We could have Ilcompleted" Q by including all the complex numbers to form the complex plane C, which is complete, but this is overkill. The completion ofthe rational numbers serves as a model for the general completion process. Feature (i) can be framed for general normed vector spaces in terms of isometries. In order to discuss feature (ii) in a general context, however, we need to introduce a general definition of density. Let (X, II . II) be a normed vector space and W ex. The set W is dense in X if every element of X is the limit of some sequence in W. Density is an important property from a practical viewpoint. If W is dense in X, then any element in X can be approximated by a sequence in W to any degree of accuracy. For example, the sequence {an} of Example 8-2-1 consists purely of rational numbers and can be used to approximate ,J2 to within any given (nonzero) error. A fundamental result in the theory of normed vector spaces is that any normed vector space can be completed. Specifically, we have the following result: Theorem 8.3.1 Given a normed vector space (X, II . Ilx), there exists a Banach space (Y, II . II y) containing a subspace (W, II .II y) with the following properties: (i) (W, II . II y) is isometric with (X, II . Ilx); (ii) W is dense in Y. The space (Y, II . II y) is unique except for isometries. In other words if (Y, II . II y) is also a Banach space with a subspace (W, II . II y) haVing properties (i) and (ii), then (Y, II . II y) is isometric with (Y, II . II y).
138
8. The Lebesgue Spaces Y
The space (Y, 1I·lIy) is called the completion of the space (X, 1I·lIx). The proof of this result would lead us too far astray from our main subject, integration. We refer the reader to [25] for the details. Exercises 8-3: 1.
(a) Suppose that Z is dense in W, and W is dense in Y. Prove that Z is dense in Y.
(b) Suppose that the completion of (X, 1I·lIx) is (Y, 1I'lIy) and that P is dense in X. Prove that (Y, II . II y) is also the completion of (P, II . IIx). 2. Let P[a, b] denote the set of polynomials on the interval [a, b], and let PQ[a, b] denote the set of polynomials on [a, b] with rational coefficients. Prove that PQ[a, b] is dense in P[a, b]. 3. Weierstrass's theorem asserts that any function in C[a, b] can
be approximated uniformly by a sequence of polynomials, Le., P[a, b] is dense in C[a, b] with repect to the II . 1100 norm. Use Exercises 8-3, No.2, to deduce that any function in C[a, b] can be approximated uniformly by a sequence in PQ[a, b].
8.4
The Space L 1
Having made our brief foray into functional analysis, we are now ready to return to matters directly involved with integration. Example 8-2-2 shows that the normed vector space (C[a, b], II . IIR) is not complete. We know, however, that this space can l]e completed, but it is not clear exactly what kinds of functions are required to complete it. In this regard, the norm itself can be used as a rough guide. Clearly, a function f need not be in C[a, b] for the Riemann integral of IfI to be defined. This observation suggests that perhaps the appropriate vector space would be R[a, b], the set of all functions f : [a, b] ~ lR such that IfI is Riemann integrable. This Ilexpansion" of C[a, b] to R[a, b] solves the immediate problem, since the sequence {fn} in Example 8-2-2 would converge to a function f E R[a, b], but it opens the floodgates to sequences such as that defined in Section
i i
8.4. The Space L1
139
3-3-1 that do not converge to functions in R[a, b]. Although II . IIR is not a norm on R[a, b] (Exercises 8-2, No. 2(b)), this problem can be overcome. Any hopes of using R[a, b] to complete the space, however, are dashed by Example 3-3-1, because this example indicates that (R[a, b], II . IIR) is not complete. Recall that Example 4-3-1 motivated us initially to seek a more
general integral to accommodate functions such as f(x) = •
{I,0,
I, if x IS IrratIonal or x = 0, 1. ~fx ~s ~atio.nal, x#- 0,
Eventually, we arrived at the Lebesgue integral. The function f plays a role in the completion of (R[a, b], II . IIR) analogous to that played by the number ,J2 in the completion of (Q, II . lie). The Lebesgue integral essentially leads us to the appropriate space and isometry for the completion of (R[a, b], II . IIR) (and (C[a, b], II . IIR)). Let A l [a, b] denote the space of all functions f : [a, b] ~ lR that are (Lebesgue) integrable on the interval [a, b] and let II . III be the function defined by IIfll1
= [
If (x) Idx.
J[a,b]
The set Al[a, b] forms a vector space, but II . III is not a norm on it because there are nonzero functions g in A1[a, b] such that IlglII = 0, Le., ifg = a.e. then IIglII = 0. Functions that fail to be norms solely because they cannot satisfy this condition are called seminorm8, and the resulting spaces are called seminonned vector spaces. Notions such as convergence and Cauchy sequences are defined for seminormed vector spaces in the same way they are defined for normed vector spaces. The problem with the seminorm on A 1 [a, b] is not insurmountable. The essence of the problem is that IlglI = IIfll whenever f = g a.e. (Theorem 5.2.3 (iii)). The set Al[a, b], however, can be partitioned into equivalence classes based on equality a.e. Let L 1 [a, b] denote the set of equivalence classes of Al[a, b]. An element F of L 1 [a, b] is thus a set of functions such that iff1,fz E F, tq~nfi = fz a.e. Since any element f of F can be used to represent the equivalence
°
140
8. The Lebesgue Spaces Y
class, we use the notation F = [f].l Addition is defined as
[f] + [g] = [f + g], and scalar multiplication as
= [af].
a[f]
The set L 1[a, b] forms a vector space, and if II . 111 is defined by
1I[f]1I =
{
ira,b]
If(x) Idx,
then (Ll[a, b], II . 111) is a normed vector space. The candidate for the completion of the space (C[a, b], II·IIR) (and the space (R[a, b], II . IIR)) is the space (Ll[a, b], II . III). In the notation of the previous section, we have X = C[a, b], II . IIx = II . IIR, Y = Ll[a, b], and II . lIy = II . III. Let W = {[f] E Ll[a, b] : [f] contains a function in C[a, bU, and let T be the operator that maps a function f E C[a, b] to the element [f] E W. Now, every element of C[a, b] has a corresponding element in W, and no equivalence class in W contains two distinct functions in C[a, b]; therefore, T is a one-to-one, onto operator from C[a, b] to W. Moreover, Theorem 4.6.1 implies that
IITfl1I
= 11[f]11I = { If(x) Idx ira,b]
.t
If(x)ldx=
IlfllR,
so that T is an isometry. The space W is thus isometric with C[a, b]. To establish that (Ll[a, b], II· III) is the completion of (C[a, b], II ·IIR) it remains to show that W is dense in Ll[a, b] and that GLl[a, b], II . III) is a Banach space. We will not prove that W is dense in Ll[a, b]. The reader is referred to [37] for this result. We will, however, sketch a proof that (Ll[a, b], II . III) is complete. Theorem 8.4.1 The normed vector space (Ll[a, b],
II . III) is a Banach space.
Although this is standard notation, there is some danger of confusion with the notation used for the span that takes sets as arguments
1
i i
8.4. The Space L1
141
Proof We prove that the seminormed space CAl[a, b], II . III) is complete. The completeness of (Ll[a, b], II . III) then follows upon identification of the functions with their equivalence classes in Ll[a, b]. Let lfn} be a Cauchy sequence in (A l[a, b], II . III). Given any E > 0 there is thus an integer N such that IIfn - fm II < E whenever m > Nand n > N. In particular, there is a subsequence lfnk} of lfn} with the property that
Let m
gm
=L
Ifnk+l - fnk!,
k=l
and let g = limm~oo gm denote the pointwise limit function. Note thatg(x) need not be finite for all x E [a, b]. Let [a, b] = h UIz, where II denotes the set of all points such thatg(x) < 00. We will show that Iz must be a null set. Now, gm E Al [a, b] and
IIgmllJ
=
1
[a,b]
Igm(x)1 ax <
t1 k=l
m
= L
m
IIfn k+l
k=l
-
[fn.+l(X) - fn.(x) I
[a,b]
fnklll < L k=l
ax
1 2k < 1.
The sequence {gm} is a monotone sequence of functions in Al[a, b], and the above inequalities indicate that limm~oo IIgmlll < 1. The monotone convergence theorem (Theorem 5.3.1) implies that g E Al[a, b] and IIgmlll -+ IIglll; hence, IIglll < 1. Since IIglll is finite, g(x) < 00 a.e., and so the set Iz must be null. The series 00
fnl (x)
+ L(fnk+l (x) -
fnk(x))
k=l
must therefore be absolutely convergent for almost all x. This series thus defines a function f, the pointwise limit, almost everywhere. Eventually, f will be identified with an equivalence class in Ll[a, b], so the fact that f is defined only a.e. is not a real concern.
, i
142
8. The Lebesgue Spaces Y
We have shown that fnk ~ f; we need to show that fn ~ f in the 11·111 seminorm and thatf E AI[a, b]. Since ifnI is a Cauchy sequence, for any E > 0 there is an integer N such that IIfn - fmllI
=
1
Ifn(x) - fm(x) I
[a,b]
ax < E
for any m > N, n > N. Let k be sufficiently large so that nk > Nand let m = nk. Then for n > N,
o<
lim (inf
k--*oo
1
m;::k [a,b]
Ifn(x) - f nm (x) I
ax)
<
E,
and so Fatou's lemma (Lemma 5.3.2) implies that for n > N, Ifn -fl is integrable over [a, b] and ~a,b] Ifn(x) - f(x) Iax < E. Therefore, fn - f E Al[a, b], and so f E A1 [a, b]. Moreover, IIfn - fill ~ 0 as n ~ 00, so that the sequence ifnI converges to f in the II . III seminorm. The space (AI [a, b], II . 111) is thus complete. The completeness of this space implies the completeness of the space (LI[a, b], II . III), since each Cauchy sequence in LI[a, b] can be represented by a Cauchy sequence in Al[a, b]. 0
8.5
The Lebesgue Spaces I!
The norm defined for the space LI[a, b] is a Iinatural" choice in applications where the average magnitude of a function is of conspicuous importance. The function
IIflll =
1
[a,b]
If(x) Iax
is the continuous analogue of the norm II . II T defined in Example 81-4 for R n . If we seek a continuous analogue for the Euclidean norm in R n we are led to the function II . liz defined by
IIfllz =
1 If(x)I ax
l/Z
Z
{
[a,b]
}
,
8.5. The Lebesgue Y
143
and more generally, if we seek a continuous analogue to a general mean norm for R n ,
for p
IIxll = {IXIIP + IX21P + ... + IxnlP} lip > I, we are led to a function II . lip defined by IIfll p
={
1
[a,b]
lip If(x)IP
ax }
In this manner, vector spaces for which these functions define norms or seminoTInS' come into prominence. The space (LI[a, b], 11·111) serves as a prototype for all the Lebesgue spaces. Let AP[a, b], 1
1
If(x)IP
[a,b]
ax < 00.
Now, IIfll p = IIgllp for any f,g E AP[a, b] such that f = g a.e., so we know that II . lip is at best a seminorm for AP[a, b]. This problem can be easily remedied by using equivalence classes. A more serious concern is that AP[a, b] may not even be a vector space. In particular, if f, g E AP[a, b], it is not clear that f + g E AP[a, b]. Moreover, it is not obvious that II . lip will satisfy the triangle inequality. As it turns out, the sets AP[a, b] are vector spaces and II· lip is a seminorm on them for 1
1
1I[f]lIp = { If(x)IP ax } [a,b] Theorem 8.5.1 (Holder's Inequality) Let F E Il[a, b] and G E Lq[a, b], where 1 < P < Then FG E LI[a, b] and
00
and IIp+ IIq = 1.
_ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _...,..-_--,'.,..1- - - - - - - - - - - - - - - - - - - - - . "
144
8. The Lebesgue Spaces IY
Theorem 8.5.2 Let 1 < P < and
00
(Minkowski's Inequality)
and suppose that F, G E V[a, b]. Then F + G E V[a, b]
The proofs of these inequalities can be found in most texts on functional analysis, e.g., [38]. In the Holder inequality, the product is the pointwise product of functions, i.e., if F = rf], G = [g], then FG = rfg], where (fg)(x) = f(x)g(x). That V[a, b] is a vector space and II . lip defines a norm on it follows from Minkowski's inequality. As with the space (L1[a, b], II . III), the normed vector spaces (V[a, b], II . lip) are complete. This result is a generalized version of the classical Riesz-Fischer theorem. Theorem 8.5.3 The normed vector spaces (V[a, b], P < 00.
II . lip) are Banach spaces for 1
<
A detailed proof of this result can be found in [17] and [18]. The proof for the case 1 < P < 00 is similar to that for the case p = 1. Essentially, the civilized behavior of the Lebesgue integral (as manifested in the monotone convergence theorem) is responsible for completeness. The Lebesgue integral thus yields an entire family of Banach spaces. To simplify notation, we shall refer to the Banach space (V[a, b], II . lip) simply as V[a, b] unless there is some ambiguity regarding the norm. These Banach spaces are collectively referred to as the Lebesgue or V spaces. We also follow the common (and convenient) practice of blurring the distinction between AP[a, b] and V[a, b] by treating elements of V[a, b] as functions. We trust the reader to make the correct technical interpretation ana to remember in this context that "f = g" means f = g a.e. Suppose that f E L 2 [a, bJ. The constant function g = 1 is also in L 2 [a, b], and therefore Holder's inequality implies that the function f ·1 = f is in L1[a, b]. In addition, we have that
IIflli
<
IIfll2111112 = (b - a)1/2l1fIl2'
This observation shows that L 2 [a, b] C L1[a, b]. The calculation works" because p = q = 2 in Holder's inequality and g = 1 is /I
----------------r-----,-,
T ' " " i-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-• • • •-
•..
8.5. The Lebesgue V
145
integrable on any interval of finite length. We can repeat this argument for the general P > 1 because g E Lq[a, b] for any q. Thus, V[a, b] C LI[a, b] for all P > 1. The next result indicates that if 1
E
VZ [a, b]. Then
IIfll p l < IIfll p z(b - a)IIPI-IIPz, and consequently f E VI [a, b]. • Proof If PI = Pz, the result is trivial. Suppose that 1
Jra,b]
IIf(x)IPIl
k
ax ==
[ If(x)IPZ ax < Jra,b]
00,
since f E VZ[a, b], and therefore If(x)IPI E Lk[a, b]. Holder's inequality with P = k, q = m, and g = 1 E Lm[a, b] yields the inequality
and
Consequently,
J
(lIfll p PI < (lIfll p zYI (b - ai-PIIPZ,
o
and the inequality follows.
Two positive numbers P and q are called conjugate exponents if IIp+ IIq = 1. Holder's inequality suggests that the Banach spaces V[a, b] and Lq[a, b] are related if P and q are conjugate exponents. In fact, the spaces are intimately related. A linear functional is an
, i
146
8. The Lebesgue Spaces IY
operator 1 from a normed vector space (X, 11·11) to the scalar field or C) of the vector space such that for any fI, fz E X
OR
l(fi + fz) = l(fi) + l(fz), and for any scalar ex
°
The functional 1 is bounded if there is a number c > such that 11 (f) I < c IIf II for all f EX. The norm of a bounded functional 1 is defined as the smallest number c such that 11 (f) I < c IIf II for all f E X and denoted by 111 II. Since ex = is a legitimate choice for a scalar, we see that
°
1(0 . fI)
= 1(0) = 01(fl) = 0,
Le., 1(0) = 0. The norm of 1 is thus given by
11111
= sup 11(f) I [EX
(any choice of c >
°satisfies 101
f#O
<
IIfll
cIlOIl)·
Exam.ple 8-5-1:
Let 1(f) =
1
k(x)f(x) dx,
[a,b]
where k E LZ[a, b], and let X = L2[a, b]. Now, 1 is evidently a linear functional, and Holder's inequality implies that for any f E LZ[a, b],
11 (f) I <
1
[a,b]
•
Ik(x)f(x) I dx = IIkflll < IIkllzllfllz;
thus,
11111 < IIkllz·
°
In fact, 11111 = IIkll z. To see this, note first that if IIkllz = 0, then k = a.e., and therefore kf = a.e. for any f E LZ[a, b]; hence, 1(f) = for all f E LZ[a, b] and 11111 = 0. If IIkll z -I- 0, choose f = k/llkll z E LZ[a, b].
°
°
8.5. The Lebesgue If
147
Now,
11er)I ==
1
[a,b]
1 == IIkll z
Since IIfllz
k(x) k(x)
ax
1
ax == IIkll z.
IIkllz
[a,b]
k
Z
(X)
== 1 for this choice, we see that 11111 = IIkll z.
The above example is typical of bounded linear functionals on the V[ a, b] spaces for 1 < P < 00. The following representation theorem expresses the situation: Theorem 8.5.5 Let p and q be conjugate exponents with 1 < P < 00, and suppose that 1 : V[a, b] ~ 1R is a bounded linear functional. Then there exists a unique element g E Lq[a, b] such that
ler) = for all f
E
1
f(x)g(x) ax,
[a,b]
V[a, b]. Moreover;
11111 == IIgll q . The proof of this fundamental result can be found in [37]. In the more general context of functional analysis, the dual space of a normed space (X, II •II) is defined to be the set of all bounded linear functionals on X. The above theorem indicates that Lq[a, b] is the dual space of V[ a, b] when p and q are conjugate exponents with 1 < P < 00. Conjugate exponents and dual spaces bring to the fore two exceptional cases. Each Banach space V[a, b] is paired with its dual space Lq[a, b], where lip + llq == I, 1 < P < 00. The conjugate exponent ofp = 2 is q = 2, and so LZ[a, b] is its own dual space. The space LZ[a, b] is clearly the only member of the V[a, b] spaces with this property, and it turns out that LZ[a, b] has several other special properties not shared by the other V[a, b] spaces. The space LZ[a, b] is special because it forms what is known as a Hilbert space. We postpone further discussion about the special properties of LZ[a, b] until the next chapter.
148
8. The Lebesgue Spaces L!
The other exceptional case is the space L1[a, b]. There is no conjugate exponent for p = I, yet there are certainly bounded linear functionals defined on Ll[a, b]. What then is the dual space of L1[a, b]? If Ik(x)1 < M < 00 for all x E [a, b], then the functional J defined by J(f)
= (
Jra,b]
k(x)f(x) ax
is a bounded linear functional, since IJ(f) I < M (
Jra,b]
If(x) I ax = Mllflll.
In fact, the condition that Ik(x)1 < M < 00 for all x E [a, b] can be relaxed to Ik(x)\ < M < 00 a.e. in view of the Lebesgue integral. This example serves as a guide to what might be expected as a dual space for Ll[a, b]. Let f : [a, b] -+ ~ be a measurable function. A number J-L is an essential upper bound for IfI if If(x) I < J-L a.e. If IfI has an essential upper bound, then it can be shown that there exists a least such bound. This leastbound is denotedby ess sup IfI. Let A 00 [a, b] denote the set of all measurable functions on [a, b] such that ess sup IfI < 00 and define the function II . II 00 by
IIflloo = ess sup Ifl· As with the AP[a, b] spaces, A 00 [a, b] is a vector space and II . 1100 is a seminorm on it. Let LOO[a, b] denote the set of equivalence classes of A 00 [a, b] modulo equality a.e. Then (LOO[a, b], 11·1100) is a normed vector space. The following theorem summarizes some of the properties of LOO[a, b]: Theorem 8.5.6 (i) (LOO[a, b], II . 1100) is a Banach space. (ii) The dual space of L1[a, b] is LOO[a, b]. (iii) Iff E L1[a, b] andg E LOO[a, b], thenfg E L1[a, b] and
IIfglll
<
IIflllllglloo
(an extension of the Holder inequality). (iv) Iff E LOO[a, b], then f E V[a, b] for all 1 < p <
00.
8.5. The Lebesgue V
149
The proofs of parts (i) and (ii) can be found in [33]. The proofs of parts (iii) and (iv) are left as exercises. Up and q are conjugate exponents with p > I, then the dual space of V[a, b] is Lq[a, b] and vice versa. In the language of functional analysis, the V spaces are reflexive for p > 1. In contrast, the dual space of V'O[a, b] is not L 1 [a, b]. The dual space of V'O[a, b] turns out to be a space of measures. More details concerning the dual space of V'O[a, b] can be found in [34]. The space L 1 [a, b] is thus unusual among the V[a, b] spaces in that it is not reflexive. We have t4us far been concerned with V spaces where the interval of integration is bounded. The definitions for V spaces can be extended to include unbounded intervals of integration. If I c ~ is any interval, the space V(l) consists of the set of functions f : I -+ R (Le., equivalence classes) such that
IIfll p =
Ii
If(x)IP
ax}
lip
<
00.
All the results stated for the V[a, b] spaces carry over to the general V(l) spaces with the notable exception of the inclusion results. The proof of Theorem 8.5.4 relies crucially on the fact that the interval is bounded. If the interval is not bounded, then results such as L 2 (l) c L 1 (I) are not valid. For example, consider the functions fI and f2 defined by fi(x) hex)
I =I
=
1/,JX., if 0 < x < I, 0, if x > I, 0,
llx,
ifO<x 1.
Now, fI E L 1 (O, 00), but f2 E L 1 (O,OO); alternatively, fI E £2(0,00) but fz E L 2(O,oo). Hence L2 (O, 00) does not contain L 1 (O, 00), and L 1 (O, 00) does not contain L2 (O, 00). Exercises 8-5: 1. Suppose that the function r positive on the interval [a, b].
[0, 1] -+ ~ is continuous and
150
8. The Lebesgue Spaces If
(a) Prove that for 1
(
r(x)lf(x)IP
( 1[a,b]
)
1~
ax
is a norm on V[ a, b].
(b) Prove that the norm rll . lip is equivalent to the norm II . lip. 2. Let 1 < PI < pz and let I be a bounded interval. Prove that IIfll pI < 1 + IIfllp 2 and from this deduce that V2(I) C VI (I).
3. The Fredholm integral operator K is defined by (Kf)(x)
= ( k(x,~)f(~)~, 1[0,1]
where k : [0,1] X [0,1] -+ ~ is a given function. Suppose that k is bounded in the square [0, 1] X [0,1]. Prove that the operator K maps functions in £z[O, 1] to functions in £z[O, 1]. 4. Prove parts (iii) and (iv) of Theorem 8.5.6.
5. Iff
E
£<'°[0,1], prove that
IIfll1 < IIfllz < ... < IIflln < IIflln+1 < ... < IIflloo.
8.6
Separable Spaces
A Banach space (X, II . II) is separable if X contains a countable set
that is dense in X in the II . II norm. The Banach space (~, II . lie), for example, is a separable space. To establish this clairp. we note that the rational numbers form a countable set that is dense in the real numbers. Separability is a desirable property, because the existence of a countable dense set simplifies problems concerning the representation of functions (e.g., bases for spaces) and the approximation of functions. In the Hilbert space theory discussed in Chapter 9, separability is a key ingredient in establishing the existence of an orthonormal basis (cf. Theorem 9.2.4). Fortunately, the V spaces are separable.
8.6. Separable Spaces
Theorem 8.6.1 The V spaces are separable for 1
151
00.
The proof of this result can be found in [17]. We limit ourselves here to a brief informal discussion of the ideas underlying the proof and focus on sets that are dense in the V spaces. These sets are ofinterest in their own right for the purpose of approximation. The construction of the Lebesgue-Stieltjes integral in Chapter 4 brings to the fore one class of functions that must be dense in any V space, viz. the set of functions consisting of x-summable step functions (Section 4.5). This density relationship is used to prove the Riemann-Lebesgue theorem (Theorem 9.3.4) in the next chapter. Unfortunately, this set is not countable. There are, of course, other sets dense in the V spaces. The next result (which we state without proof) provides asample of such a set. Theorem 8.6.2 Let ego (~) denote the set offunctions f : ~ -+ ~ with derivatives of all orders and with supports offinite length. The set Cgo~) is dense in each V(~) space for 1 O. Now, f has a support of finite length, and hence there is some number fJ > 0 such that f = 0 outside the interval [-fJ, fJ]. Certainly f (restricted to the interval [-fJ, fJ]) is in the space C[ -fJ, fJ], and thus there is a polynomial PE such that E
sup IPE(X) - f(x) I < (2fJ)l/p
xE[-P,Pl
(8.1)
Moreover, we know from Exercises 8-3-2 that the set PQ [ -fJ, fJ] of polynomials with rational coefficients is dense in P[ - fJ, fJ], and hence we can find a polynomial with rational coefficients satisfying inequality (8.1). In fact, we can find a polynomial qE with rational coefficients such that inequality (8.1) is satisfied and qE(-fJ) = qE(fJ) = O. Now the polynomial qE canbe extended to a functiongE = qE' X[-P,Pl
152
8. The Lebesgue Spaces £P
defined on ~. Here X[-P,P] is the characteristic function defined by
X_
ex) =
[ P,P]
{I,
if x E [-,8, ,8], 0, otherwise.
Thus, IIgE - flip = { ( IgE(X) - f(x)I Pdx}l/P =
JJR
<
{i
IgE(X) - f(x)I Pdx}l/P
[-P,P]
sup IgE(X) - f(x) I(2,8)l/P <
E.
XE[-P,P]
Now, the set of rational numbers is countable, and this can be used to establish that the set PQ t-,8,,8] is also countable. If we define the set r = {g : g = q . X[-~,P] for some q E PQ [ -,8,,8] and some ,8 = 1,2, ...}, then it can be shown that r is a countable set in Cgo~). The above inequality indicates that r is dense in Cgo(R), and hence it is dense in V(~). Thus, V(~) contains a countable dense set.
8.7
Complex LP Spaces
In this section we pause to make a modest generalization ofV spaces to include complex-valued functions of a real variable. A fuller treatment of these spaces can be found in [17]. Let I c ~ be some interval and let f : I ~
We wish to construct spaces analogous to AP(!) and V(l). The natural generalization of the norm function is the function II· lip defined by
lIfllp =
{i
If(t)[P dt} liP,
__
._~-----------------_.
..- .... -
8.7. Complex V Spaces
153
where for 1 = Ref - ilmf (the complex conjugate), Ifl 2 = fl. The next theorem shows that IflP is integrable if and only if IRefl P and IImfl P are both integrable. The proof rests on the inequalities IRefl < If!' IImfl < If!' and If I < IRefl + IImf!' and is left as an exercise. Theorem 8.7.1 Let f : I ---+ C be a complex-valued function defined on an interval I c ~ and suppose f = Ref + Hmf, where Ref and Imf are real-valued measurable functions. Then, for 1
l
lf (t), Pdt
<
00
if and only if
1
IRe f(t)I P dt <
00
and lIImf(t)IP dt <
00.
Let A~(I) denote the set of functions f : I -+ C such that IIfllp < 00. The above result asserts that f E A~(I) if and only if Ref E AP(I) and Imf E AP(I). For 1 < P < 00 it can be shown that the function II . lip satisfies the Holder and Minkowski inequalities. In particular, the sets A~(I) are complex vector spaces and II . lip is a seminorm. As with the AP(!) sets, we can form a normed vector space by partitioning A~(I) into equivalence classes. Two functions belong to the same equivalence class if If - gl = 0 a.e. Let L~(I) denote the resulting set of equivalence classes. The spaces L~(I) share the same properties as the V(l) spaces. For example, if 1 I, then the dual space of L~(l) is the space Lt(I). If J is a bounded linear functional from L~(l) to C, then there is a unique g E Lt(I) such that
J(f) =
1
f(fJk(t) dt
for all f E L~(I). The space LC'(!) can be defined in a manner analogous to that used to define Loo(I), and this space is the dual of Lt(I).
154
8. The Lebesgue Spaces l j
The Lt(!) spaces often arise in applications involving line integrals in the complex plane. Recall that if y is some curve in C represented parametrically by some piecewise smooth function z(t) = x(t) + iy(t) for t E [to, tI], then the line integral of a function F defined on y is given by
1
F(z) dz
y
=
i
tl
F(z(t))z'(t) dt.
to
The definition of a line integral can be extended using the Lebesgue integral in an obvious way, and thus line integrals can be defined for a more general class of functions. Note that a curve y may be parametrized any number of ways. For example, the variable t may be replaced by any smooth function g : [so, sI] -* [to, tIl to give a new parametrization z(s) = z(g(s)), provided that g'(s) ¥= 0 for all s E [so, sI]. If g'(s) > 0 in the interval [so, sI] then the parametrization is orientation preserving, and it is well known that the value of the line integral in the Riemann setting is invariant under smooth orientation preserving reparametrizations. In the Lebesgue setting, Theorem 6.2.1 ensures that this is also the case. Thus, provided that the orientation of y is specified, we can use the notation f y F(z)dz without ambiguity. In this context we may also use the notation Lt(y) unless a specific parametrization of y is required. A particularly interesting case is that in which y is a simple closed curve (no self-intersections) and F(z) is ho1omorphic (analytic) in the region enclosed by y but not necessarily on y itself. In the next section, we investigate a class of spaces known as Hardy spaces and show that the Lt(!) spaces are closely related to spaces of ho1omorphic functions.
8.8
The Hardy Spaces HP
Hardy spaces are complex normed vector spaces of functions ho1omorphic on a region of the complex plane. These spaces are closely related to the Lebesgue spaces and share the same properties. The functions in a Hardy space are quite different in nature from elements in an Lt(!) space. Aside from the fact that elements in the
8.8. The Hardy Spaces HP
155
latter set are equivalence classes, the underlying functions in an L~(I) space need not be continuous. In contrast, the functions in a Hardy space are infinitely differentiable. That these spaces should be so closely linked is remarkable. Hardy spaces play an important role in complex function theory and harmonic analysis. These spaces also arise in applications such as control theory. The discussion in this section presumes some knowledge of complex analysis, but it is primarily a descriptive account and no proofs are given. For a basic reference on complex analysis the reader is directed to [2], [10], or [39]. A much fuller discu$sion of Hardy spaces along with the proofs of various results in this section can be found in [20], [21], and [37]. Let 0 C C denote a region (Le., a nonempty, connected open set) and let A(O) denote the set of functions holomorphic on O. It is clear from the elementary properties of holomorphic functions that A(O) is a vector space. A candidate for a norm on A(O) is the function II . II 00 defined by
IIflloo = sup !f(z)l. zeQ
Although this function satisfies the conditions for a norm such as the triangle inequality, it is not finite for all f E A(O). There are functions in A(O) With singularities on the boundary ao of O. If the set A(O) is restricted to the functions f for which Ilflloo < 00, then the resulting space is a normed vector space. Let ROO(O) = if E A(n) :
IIflloo <
oo}.
The set ROO (0) is still a vector space, and now II . II 00 is a norm on it. In fact, the space (ROO(O), II . 1100) is a Banach space. The value of IIflioo for a given f E ROO(O) is less tractable than the analogous function defined fot e[a, b] (Example 8-1-5) owing to the two-dimensional nature of the region 0 and the fact that 0 is an open set. A fundamental result in complex analysis known as the maximum modulus principle, however, mitigates these problems. Let 0 c C be a bounded region with boundary ao and suppose that f E A(O) and also that f is continuous in the set Q = 0 u ao. Under these conditions, the maximum modulus theorem implies that the function IfI assumes its maximum value on the boundary ao and
----------------------_
......
156
8. The Lebesgue Spaces IJ'
that a nonconstant function in A(O) cannot have local maxima for IfI in o. The maximum modulus principle cannot be applied directly to all the functions in ROO (0), sinoe functions in this set need not be continuous on 0 U ao. Nonetheless, the norm IIflloo of a function in ROO(O) reflects the values of If(z) I as z approaches the boundary of the region, and this suggests that the norm can be calculated using a limiting argument. The Riemann mapping theorem implies that arbitrary regions such as S can be conformally mapped to a unit disk in the complex plane. We can thus limit our investigati~nsmostly to the study of functions holomorphic on the unit disk centred at the origin. Let D(O; r) = {z E C : Izl < r}, D(O; r) = {z E C : Izl < r}, aDr = {z E C : Izl = r}, and for simplicity, denote DCa; I), D(O; I), aD l by D, D, and aD, respectively. If a < r < I, then, taking 0 = D(O; r), the maximum modulus principle implies that sup If(z) I = sup If(z)l. zeD(O;r)
The function
ze8Dr
Moo is defined by Moo(r,f) = sup If(reUP)1,
and it is evident from the maximum modulus principle that for a fixed f E ROO (D) the function Moo is monotonic increasing in rand that sup If(z) I = Moo(r,f)· zeD(O;r)
We thus expect that for any f lim
r-+l-
E
HOO(D),
Mooer , f) = IIflloo,
and in fact, this relationship can be proved formally. The function Moo suggests a ,generalization for norms analogous to the IY norms. Let Mp be the function defined by Mp(r,f)
= { .2..1· 2n
[0,211"]
lip
If(rei
}
,
-----------------~---------------_._-_.-.
__
.
8.8. The Hardy Spaces HP
157
and let HP(D) be defined as HP(D) =
if E A(D) : sup Mp(r, f)
< oo}.
r< 1
It can be shown that for 1
the Mp functions satisfy H6lderand Minkowski-type inequalities. Specifically, if p and q are conjugate exponents with p > I, then for any functions f E HP(D), g E Hq(D), 00
MI(r,fg) < Mp(r,f)Mq(r, g),
and ifh
E HP(rJ),
then Mp(r,f + h) < Mp(r,f)
+ Mp(r, h).
If P > I, the function II . lip defin¢d by
IIfllp =
SlilP Mp(r, f) r<: 1
is therefore a norm on HP(D). (Note that we need not form equivalence classes because the functions involved are holomorphic, and thus iff = g a.e., thenf = g.) As with Moo(r, f), it can be shown that for all f E HP (D), lim Mp (r 1 f) = Ilfll p .
r-+I-
The normed vector spaces (HP(D), II . lip) are called Hardy or HP spaces. As with the Lebesgue spaces, we refer to the normed vector space (HP(D), II . lip) as HP(D). 'Iihe Hardy spaces share the same properties as the Lebesgue spaces!. The next theorem gives a sample of some of these properties. Theorem 8.8.1 (i) Holder's Inequality: If 1 < P < exponent ofp, then for any f fg E HI (D) and
IIfglh
and q is the conjugate E HP (D), g E Hq (D) we have that
<
00
IIfllp lIgll q .
(ii) Minkowski's Inequality: For any f, g E HP (D) with P > 1,
IIf +gllp < IIfll p + IIgllp · (iii) Completeness: The normed vector spaces (HP(D), Banach spaces. .
II . lip)
are
158
8. The Lebesgue Spaces l j
::s P2
(iv) Inclusion: For any 1
<
00,
HOO(D) c HP2(D) c HP2(D).
The two inequalities in the above. theorem are direct consequences ofthe corresponding inequalities involving the Mp functions. A proof of completeness can be found in [21]. The inclusion result is established using the same approach as used to prove a similar result in Theorem 8.5.4. lt is not a coincidence that the HP spaces mimic the IY spaces. These spaces are closely related through the l'boundaryvalues" ofthe functions in HP(D). Cauchy's integral formula is a remarkable manifestation that a holomorphic function is determined by its boundary values. Recall that for any f E A(b(O; r)), r > 0, the Cauchy integral formula states that fez)
=~ ( 2m
few) dw - z
)aDr w
for any z E D(O; r). Here, the circle aDr is oriented anticlockwise. The value of f on the boundary aDr thus determines the function f uniquely in the interior D(O; r). We cannot apply Cauchy's integral formula directly on the boundary of D because f need not be holomorphic on aD, but we can still aJj>ply the formula on any boundary aDr when 0 < r < 1. This suggests that any function f E HP(D) can be lIidentified" with a function i defined on aD a.e. by some limiting process as r -+ 1- . Let fr(z) :; f(rz).
-
.
If 0 < r < I, then fr(z) is holomorphic on D and the Cauchy integral formula implies that fr(z) =
~
1.
fr(~)~.
2m. aD ~ - z
Intuitively, we should be able to define a function f : aD -+ C by considering the function lim r--+ 1- fr. The next result shows that this approach does lead to an isometry from the HP(D) spaces to the Lt(aD) spaces. The notation for the norms for these spaces is the
8.8. The Hardy Spaces HP
159
same by convention. Here, we shall use the notation II . IIHP, and II . Ilv to distinguish the norm f~r WeD) from the norm for Lt(aD). Theorem 8.8.2 L..,et 1
IIfIlHP. IIf - fr IIv
E
HP(D). Then there is an element
= o. fez) =
~
1 f(~)~.
2m aD ~ - z
The above result indicates that there is an isometry from HP (D) to Lt(aD). These spaces are Iljot isometric, since the set Lt(aD) contains elements with no conesponding members in HP(D). Let ,!!p(aD) C Lt(aD) denote the set of elements If] E Lt(aD) such that f corresponds to some function f E HP(D). The complex normed vector spaces (1-lP(aD), II . IIv) and (HP(D), II . IIHP) are isometric. Now, it can be shown that the space (1-lP (aD) , II . IIv) is a closed subspace of the Banach space C~t(aD), II . Ilv), and this means that (1-lP(aD), II . IIv) is also a Banach space. Since (1-l,P(aD), II . IIv) is isometric to (HP(D), II . IIHP), the ]atter space must also be a Banach space. Properties ofthe HP(D) spaces such as completeness are generally proved by establishing the analogous results for the 1-lP(aD) spaces, and in this sense the Hardy spaces "inherit" the qualities of the IY spaces. The HP(D) spaces can be generalized to sets of functions holomorphic in arbitrary regions ot the complex plane via conformal transformations. Perhaps the most frequently encountered Hardy spaces aside from the HP (D) spaces are the spaces of functions holomorphic on a half-plane. Specifically, the Hardy spaces of functions analytic on the right half-plane no = {z E C : Rez > o} are of interest owing to (among other things) their connection with the Laplace transform. The Hardy $paces in the half-plane share most of the properties of the HP(D) spaces. Indeed these properties are sometimes established by using a Mobius transformation such as
w-l w+l
z=--
160
8. The Lebesgue Spaces IJ'
This transformation maps points ZED to points w E no; the circle aD is mapped to the imaginary axis. If g E A(D), then the function f defined by few) =g
(W-l) w+l
(8.2)
is in A(no), and in this manner we can IItransport" some of the properties of the HP(D) spaces. We return to this comment at the end of the section. Let Mp(x,n =
{!
r
p
If(x + iY)IP dy
,
and define the function II . lip by
IIfllp
= supMp(x,f)· x>l!l
The Hardy spaces HP(n o) are defined as HP(n o) =
if
E
A(n o) : IIfllp < oo}.
In contrast with the set D, the set no is not bounded in the complex plane, and this complicates matters because along any line Ix = {z E C : Re z = x, x > O} the function fx defined by fx(y)
= r(x + iy)
must be in IY(Ix ). Moreover, we cannot appeal directly to the maximum modulus principle to evaluate IIfllp because the region is unbounded and the function may have a singularity at the point at infinity. The bound for Mp ( x, f) for large x is as important as that for small x for the general f E A(n o). It turns out, however, that the requirement that IIfllp < 00 forces f to tend uniformly to zero as z tends toward the point at infinity along any path inside any fixed half-plane of the form no = {z E C : Rez > 0 > O}. It can thus be shown that Mp ( x, f) is a decreasing function of x and that
As with the HP(D) spaces, the boundary values of functions in HP(n o) can be mapped to elements in an IY space. The
8.9. Sobolev Spaces
Wk,p
161
Lebesgue space associated with HP(n o) is the space L~(lR), and the corresponding Cauchy (Poisson) integral for this case is x ( f(it) fx(Y) = 1C fRo x2 + (y _ t)2 dt.
For each x > 0 the function f" is in L~ (lR), and
IIf" - fllv -+ 0 as x -+ 0+. Moreover, we have that IIf IIHP = IIlllv. As remarked earlier, the sets D and no are related by a Mobius transformation, and we can expect that the sets of functions in one space can be used to generate the functions in the other space. Ifg E HP(D) , then the function f as defined in equation (8.2) is certainly in A(n o), but it is not clear whether f E HP(no) and whether all the functions in HP(n o) can be generated by functions in HP(D). The final result of this section gives a concrete connection between the two spaces (cf. [21 D. Theorem 8.8.3 Letg E A(D) and define f as in equation (B. 2). lfp > 1, thenf E HP(n o) if and only if there is a function G E HP(D) such that g(z)
= (1 -
z)2/P G(z).
Equivalently, the function g is in HP(D) ifand only if there is a function F E HP(n o) such that few)
8.9
= (1 + w)2/p F (w).
Sobolev Spaces
Wk,p
Sobolev spaces are another class of function spaces based on the Lebesgue integral. These function spaces have found widespread applications in differential equations and feature norms that not only IImeasure" the modulus of a function but also the modulus of its derivatives in an IY setting. In this section we give a briefglimpse
iii
162
8. The Lebesgue Spaces IJ'
of these important spaces and hopefully whet the reader's appetite for a more serious study. A detailed account of the theory can be found in [1 ]. Let Cl (1) denote the set offunctions f mapping the intervall C lR to lR such that f' exists and is continuous for all x E 1. The set Cl (1) forms a vector space, and the function II . Ih,oo defined by
IIflh,oo = sup If(x)1 + sup I['(x) I xeI
xeI
is a norm. The space (Cl(l), II . Ih,oo) can be shown to be complete. More generally, we can define functions II . Ih ,P by
and investigate the correspondingnormed vector spaces (Cl(l), IIflh ,P ). For 1
8.9. Sobolev Spaces
Wk,p
163
by k
IIfllk,P =
L
IIfCJ)lI p ,
j=O
then the Sobolev space (Wk,p(I), II . IIk,P) is defined as the completion of the space (ek(I), II . IIk ,P)'
Hilbert Spaces 2 andL CHAPTER
9.1
Hilbert Spaces
Hilbert spaces are a special class ofBanach spaces. Hilbert spaces are simpler than Banach spaces owing to an additional structure called an inner product. These spaces playa significant role in functional analysis and have found widespread use in applied mathematics. We shall see at the end of this section that the Lebesgue space £2 (and its complex relative H 2 ) is a Hilbert space. In this and the next section, we introduce some basic definitions and facts concerning Hilbert spaces of immediate interest to our discussion of the space £2. Further details and proofs of the results presented in these sections can be found in most books on functional analysis, e.g., [25]Let X be a real or complex vector space. An inner product is a scalar-valued function (., .) on X x X such that for any f, g, hEX and any scalar a the following conditions hold: (i) if, f) > 0; (ii) if, f) = 0 if and only if f = 0; (iii) if + g, h} = if, h} + (g, h); (iv) if, g) = (g,f); (v) (af,g)
= a if, g). 165
166
9. Hilbert Spaces and L 2
A vector space X equipped with an inner product (., .) is called an inner product space and denoted by (X, (., .)). Note that in general, (f,g) is a complex number; however, condition (i) indicates that (f,f) is always a real nonnegative number. Note also that conditions (iii) and (iv) imply that (f,g + h)
= (f,g) + (f, h).
In general, (f, pg) = p(f, g) # f3(f, g), and consequently the inner product is not in general a bilinear function. The special case arises frequently in applications that X is a real vector space and (., .) is realvalued. These spaces are referred to as real inner product spaces, and for these spaces the inner product is a bilinear function.
Example 9·1·1: en and Rn Let X = en and for any z = (Zl, Z2, ... ,zn), W = (WI, W2, ... , w n) en define (., .) by
E
n
(z, w)
= LZjWj. j=l
Then it is straightforward to verify that (.,.) is an inner product on en. Similarly, if X = Rn and for any x = (Xl, X2, ... ,xn), y = (Y1, Y2, ... ,Yn) E Rn the function (., .) is defined by n
(x, y)
= LXjYj, j=l
then (., .) defines an inner product on Rn. The definition of the inner product is modeled after the familiar inner product (dot product) defined for Rn.
Example 9·1·2: -e 2 Let -e 2 denote the set of complex sequences {an} such that the series L~=l Ian 12 is convergent. If addition and scalar multiplication are defined the same way as for the space -e 1 in Example 9-1-3, then-e 2 is a vector space. Suppose that a = {an}, b = {b n } E -e 2, and let Cn = max (an , b n). Then the series L~=l Icn l2 is convergent, and hence the series L~l anb n is absolutely convergent. An inner product on this
9.1. Hilbert Spaces.
167
vector space is defined by 00
(a, b)
=L
anbn .
n=l
Let (X, (., .)) be an inner product space and let II . II : X -+ R be the function defined by IIfll
= J (f,f)·
We will show that II ·11 as defined above is a norm hence justifying our notation. The,conditions for the inner product ensure that II· II meets the requirements for a norm except perhaps the triangle inequality. In order to establish the triangle inequality we need the following result, which is of interest in its own right: Theorem 9.1.1 (Schwarz's Inequality) Let (X, (., .)) be an inner product space. Iff, g
E
X, then
I(f, g) I < IIf illig II· The proof of this inequality is left as an exercise. Now, IIf + gll2 = if + g,f + g) = (f,f) + (f,g) + (f,g) + (g,g) = IIfll 2 + 2Re (f,g) + IIgll 2 2 2 < IIfll + 21(f,g)1 + IIgll , and Schwarz's inequality implies that 2 IIf + gll2 < IIfl1 + 211fllllgll
+ IIgll 2
= (lIfll + IIglli· Thus, IIf + gil < IIfll
+ IIgll,
and hence II . II defines a norm on X. Given any inner product space (X, (., .)) we can construct a normed vector space (X, II . II). The function II . II is called the norm induced by the inner product. The normed vector space mayor may not be complete. If (X, II . II) is a Banach space, then the inner
168
9. Hilbert Spaces and £2
product space (X, (., .)) is called a Hilbert space. A Hilbert space is thus an inner product space that is complete in the norm induced by the inner product. The inner product spaces (Rn, (., .)) and (en, (., .)) are examples of finite-dimensional Hilbert spaces. Although we do not show it here, the inner product space (£2, (., .)) of Example 9-1-2 is an infinite-dimensional Hilbert space. It is of interest to enquire whether a given normed space (X, II . II) can be identified with an inner product space (X, (., .)). In other words, given a norm 11·11 on X can an inner product (., .) be defined on X such that II . II corresponds to the norm induced by (., .)? Suppose that the norm II . II can be obtained from some inner product (., .). Then for any f,g EX,
IIf + gll2 + IIf - gll2 = if + g,f + g) + if -
g,f - g)
+ (f,g) + (g,f) + (g,g) +(f,f) + (f, -g) + (-g,f) + (g,g) 2 2 2 (lIfll + IIg1l ) ;
= (f,f) =
thus, if a norm II . II can be obtained by some inner product, then it must satisfy the equation
IIf + gll2 + IIf - gll2 = 2 (lIfll 2+ IIgI1 2). This condition is also sufficient. This equation is called the parallelogram equality. The name comes from an elementary relation in plane geometry. If x and yare two vectors in R2 that are not parallel, then they can be used to define a parallelogram. Here, the quantities IIxll and IIYII correspond to the lengths of the vectors x and Y respectively and hence the lengths of the sides of the parallelogram. The quantities IIx + YII and IIx - Y II correspond to the lengths of the diagonals. The parallelogram equality is useful for (among other things) showing that certain norms cannot be obtained from an inner product. Example 9-1-3: Consider the normed vector space (C[a, b], 11·1100) defined in Example 8-1-5. We shall use the parallelogram equality to show that the norm II . 1100 on C[a, b] cannot be obtained by an inner product. Suppose, for contradiction, that the norm II . II 00 can be obtained by an inner product. Then the parallelogram equality must be satisfied for any
_ _ _ _ _ _~ - - - - - - - - - - - - - - _ . -••_., •.
I
9.1. Hilbert Spaces
choice of f, g
E
169
C[a, b]. Let f and g be the functions defined by
x-a . -a
f(x) = I, g(x) = b
Now, IIflloo = I, IIglloo = I,
IIf+glloo= sup 1+ xE[a,b]
x-a =2, b- a
and
IIf - glloo = sup 1xE[a,b]
x-a = 1; b- a
consequently,
IIf+gll~ + IIf-gll~ = 4+ 1= 5, and
2 (lIfll~ + IIgll~) = 4. As the parallelogram equality is not satisfied for these functions, the normed vector space (C[a, b], 11·1100) cannot be obtained from an inner product space. The parallelogram equality can be used to show that the only lY space that might arise from an inner product space is L 2 • Theorem 9.1.2 The only Lebesgue norm II .lip that can be obtained from an inner product is the L 2 norm II . 112. Proof We first establish the result for lY[ -I, 1] spaces. A modest change of the proofleads to the result for general lY[a, b] spaces and lY(l) spaces where I is an unbounded interval. Suppose that the norm II . lip on lY[ -I, 1] can be obtained from an inner product. Then the parallelogram equality indicates that
IIf + gil; + IIf - gil; = 2(lIfll; + IIgll;) for all f, g
E
lY[ -I, 1]. Consider the functions f and g defined by
f(x)
= 1 + x,
i I
g(x)
=1-
x.
9.1. Hilbert Spaces
1 71
Now,
/ p+ 1 8 (P) = 1 - P + 1 - log(p + 1) = - log(p + 1) < 0, and thus 8 is a monotonic strictly decreasing function of p. Since 8(1) = 1 - 2 log 2 < 0, it follows that 8(P) < 0 for all p > 1 and consequently that Q/(P) < 0 for all p > 1. This means that Q is a monotonic strictly decreasing function of p, and therefore the equation Q(P) = 0 can have at most one solution. Therefore, the parallelogram equality is satisfied only if p = 2. • A slight modification of the functions f, g for the general interval [a, b] leads to the proof of the result for the V[a, b] spaces. If I is not a bounded interval, then a suitable restriction of the functions can be used to establish the result for the V(1) spaces. For example, suppose that 1= (-00,00). Then we can choose the functions 1 +x,
ifxE[-I,I], otherwise,
g(x) = { 1 - x, 0,
if x E [-1,1], otherwise,
[(x)
={
0,
and the proof then is exactly the same.
0
It is not difficult to see that for any interval I, £2(1) is a Hilbert space.
We know from Chapter 8 that this space is complete with respect to the II . 112 norm, and the inner product (', .) defined by (f, g) = [f(X)g(x)
ax
for all f, g E £2(1) induces this norm. The complex space £~(1) is also a Hilbert space with an inner product (', .) defined by (f,g)
= [f(t)g(t)dt.
The conjugate exponent p = q = 2 is special throughout the function spaces based on the Lebesgue integral. One can show, for example, that the only Hilbert space among the Hardy spaces is H 2 , and a similar statement can be made about the Sobolev spaces. These spaces have found widespread use and have many special properties not enjoyed by the other V (HP, Wk,P) spaces, p # 2, because they
---,---,~
I
170
9. Hilbert Spaces and £2
These functions are (in equivalence classes) in V[ -I, 1] for all p > 1. Now,
1 =1 + gll~ = 1 gll~ 1
IItll~ IIgll~
and
lit
lit -
=
[-1,1]
2P+I P
+1
,
2P+I
1
(l-xfax=
[-1,1]
-1
1 1
P+1
,
1
11
[-1,1] 2P+ I
+ x + (1 -
x)IP ax
=
,
=
[-1,1]
= 2P+ I
=
(1 +xfax
-1
11-x1Pax
=
1
1 = 1
11 +xlP ax
=
11 + x - (1- x)IP ax =
2P ax
-1
1
12xlP ax
-1
2P+I xP ax = - o p+l
1 1
The parallelogram equality implies that
{2P+ I)2IP+
(2P+
P I 2P+I )2I = 2 {( 2P+I )2/P + )2/P} ( p+l p+l p+l
,
Le.,
i/
(p + 1
p
-
3 = O.
Note that p = 2 is a solution to this equation. We will show that this is the only solution for p > 1. Let Q(P) = (p + 1)2/p
-
3.
Then Q'(P) =
2
p2(p+ 1)
(p + li/p s(p)
,
where S(P) = p - (p + 1) log(p + 1).
--.-..--.. - I
172
9. Hilbert Spaces and £2
are Hilbert spaces. In the next section we investigate some special properties of Hilbert spaces. For the remainder of this chapter we shall denote an inner product space (X, (', .)) simply by X, unless there is some danger of confusion, and the norm II . II on X will always be the norm induced by the inner product unless otherwise stipulated. Exercises 9-1: 1. Verify that the function (', .) defined in Example 9-1-2 satisfies
the conditions for an inner product. 2. Let X be an inner product space and suppose f, g E X with g # O. Prove that l(f,g/lIgll)I < IIfll and use this to prove the Schwarz inequality. 3. Let X be an inner product space. Prove that for any elements f,g, hEX, IIh - fll
2
+ IIh -
gll2
1
= -llf -
gll2
+ 211h -
1 -(f + g)1I 2.
2 2 This relation is known as the Appolonius identity.
9.2
Orthogonal Sets
The paradigm for a finite-dimensional inner product space is the space R n discussed in Example 9-1-1. In this space, the inner product can be used to measure the angle between two vectors, Le., (x, y) = IIxllllYII coscP, where cP is the angle between the vectors x and y. This geometrical idea can be extended to infinite-dimensional real inner product spaces, but there is no satisfactory extension to general inner product spaces. It turns out, however, that the magnitude of the angle between elements in an infinite-dimensional real vector space is of limited interest in the general theory with one important exception. Recall that for x, Y E Rn the relationship (x, y) = 0 means geometrically that the vector x is orthogonal to the vector y. This concept of orthogonality can be readily extended to general inner product spaces. Let (X, (', .)) be any inner product space. 'TWo elements f, g E X are said to be orthogonal if (f, g) = o.
i
I
9.2. Orthogonal Sets
173
This relationship is denoted by f 1.. g. As we shall see, orthogonality plays an important part in Hilbert space theory. Example 9-2-1:
Consider the inner product space £2[ -Jr, Jr], and let fn denote the functions defined by fn(x) = sin(nx) for n = 1,2, .... Now, ifn, fm) =
=
1
[-1l',1l']
L:
fn(x)fm(x) ax
sin(nx) sin(mx) dx.
Integration by parts indicates that 1 ]1l' +n j1l' cos(nx)cos(mx)ax ifn,fm)= [ --sin(nx)cos(mx) m -1l' m_1l'
= -nj1l' cos(nx) cos(mx) ax, m -1l'
and integration again by parts yields ifn, fm)
= !!- {[~ sin(mx) COS(nx)]1l' + !!-j1l' sin(nx) sin(mx) ax} m
=
m
-1l'
m_1l'
(:)2 ifn,fm).
If n # m, then 1 - (n/m)2 n = m, then (fn,fm)
=
llfn1l
# 0, and consequently 2
=
L:
2
sin (nx) dx
ifn,fm) = 0. If
= Jr.
We therefore have that fn 1.. fm unless m = n. Suppose that f and g are elements in the inner product space X. Then
If f 1.. g then If,g) = 0, and we thus have an extension of Pythagoras's theorem:
174
9. Hilbert Spaces and £2
We pause here to introduce two terms applicable to a general vector space. Let X be a vector space, and let Y and Z be subspaces of X. If for each f E X there exist elements g E Y and h E Z such that f = g + h, then we say that X is the vector sum of Y and Z, and denote this relationship by X = Y + Z. If, in addition, the elements g E Y and h E Z are determined uniquely for every f EX, then we say that X is the direct sum of Y and Z, and write X = Y ffi Z. Given a set S C X, where (X, (., .)) is an inner product space, another set S1. can be formed by taking all the elements of X that are orthogonal to every element of S, i.e., the set S1. = if EX: if, g) = o for all g E S}. The set S1. is called the orthogonal complement of S. Example 9-2-2: Let el, e2, e3 be three linearly independent vectors in lR3 such that (ej, ek) = 0 forj, k = 1,2,3 unlessj = k. If S = {v E lR3 : v = ael +be2 for some a, b E }R} (i.e., the span of tel, e2}), then S1. = {w E lR3 : W = ce3 for some C E lR}. Since the set {el, e2, e3} forms a basis for }R3, we have that }R3 = S ffi S1. . Example 9-2-3: Let.e 2 be the inner product space defined in Example 9-1-2, and let el E f} be the sequence {I, 0, 0, ...}. Let S be the subspace defined by S = {a E f} : a = ael for some a E C}, and define the set B by B = {b = {b n } E .e 2 : b I = OJ. Then (a, b) = 0 for all a E Sand all b E B, and consequently B C S1.. In fact, B = S1., for if there is a sequence c = {cn} with CI # 0, then (a, c) = aCI, and aCI is not zero for all a E C. Since any sequence d = {tin} can be written as {dl , 0, 0, ...} + {O, d2 , d3 , ••• }, and this representation is unique, we have that .e 2 = S ffi S1. . The inner product space]R3 and the set S ofExample 9-2-2 demonstrate two geometrical properties both of which extend to any finite dimensional inner product space. First, given any vector x E }R3 there is a unique vector v E S at a minimum distance from x, i.e., IIx - vII < IIx - vII for all v E S \ {v}. Second, any vector x E }R3 can be represented in the form x = v + w, where v E Sand W E S1. are uniquely determined, Le., R 3 = SffiS1.. Any subspace of}R3 has these properties, and one can enquire whether these properties persist in
----------.-...------------_._-_
.. ,....-
,
9.2. Orthogonal Sets
175
the infinite-dimensional case. The space £2 in Example 9-2-3 demonstrates the second property, and it is straightforward to show that the set S in this example also has the minimum distance" property. Unfortunately, these properties do not carry over to the general inner product space. Under the crucial assumption of completeness, however, these properties do extend to infinite-dimensional inner product spaces. II
Theorem 9.2.1 Let S be a closed subspace of the Hilbert space X and. let f EX. Then • there exists a unique g E S at a minimum distance from f. Note that if S is a closed subspace, then S is a Hilbert space in its own right. Theorem 9.2.2 (Projection Theorem) Let S be a closed subspace ofthe Hilbert space X. Then S..L is also a closed subspace in X, and. X = S E9 S..L. Moreover, iff E X is decomposed into the form f = g + h, where g E S and h E S..L then g is the unique element in S closest to f. The projection theorem provides a key to understanding bases in Hilbert spaces. The study of bases in general Banach spaces is of limited value, but there is a rich and useful theory for bases in Hilbert spaces. As our main focus will eventually be on the space L 2 , which is separable, we limit our general discussion to bases for separable Hilbert spaces. In Example 9-2-2, consider the set K = {ell e2, e3}' The set K..L consists of all the vectors x E lR3 such that (x, ek) = 0 for k = 1,2,3, buttheonlyvectorwiththispropertyisx = O;thusK..L = O.Now, the set K forms a basis for lR3 , and this is characterized by the condition K..L = O. Motivated by this observation, we can extend the concept of an orthogonal basis to general (separable) Hilbert spaces. Let N be a set of elements in the Hilbert space X. If N..L = 0, then N is called a total set1 inX. If M is a countable set inX and (a, b) = 0 for all a, b E M, a =1= b, then M is called an orthogonal set in X. If, in addition, lIall = 1 for all a E M, then M is called an orthonormal I
IThese sets are also called "complete" in the literature. We avoid this term, since "complete" in this context has no connection with "complete" as used in Banach space theory.
176
9. Hilbert Spaces and £2
set in X. For our purposes, we are interested primarily in the case where M has an infinite number of elements but is countable. Let M = {4>nl be an orthonormal set in the Hilbert space X, and let f E X. The numbers if, 4>n} are called the Fourier coefficients of f with respect to M, and the series L~=l if, 4>n}4>n is called the Fourier series of f (with respect to M). Now, for any mEN, m
o < IIf -
m
L if, 4>n}4>n 11 n=l
2
= if -
m
m
L if, 4>n}4>n, f - L if, 4>n}4>n} n=l n=l m
= if,f} - if, Lif, 4>n}4>n} - (Lif, 4>n)4>n,f}
n=l
n=l
m
m
+ (L if, 4>n}4>n, L n=l
if, 4>n)4>n}
n=l m
= IIfll
2
-
m
+ II L
2 L (f, 4>n) if, 4>n} n=l
2
n=l
m
=
if, 4>n}4>n 11
m
IIfll 2 - 2 L Iif, 4>n}1 2 + II Lif, 4>n}4>nIl 2 , n=l
n=l
and Pythagoras's equation implies that m
m
II Lif,4>n}4>nIl = L 1Iif,4>n}4>nIl 2 2
n=l
n=l m
= L
Iif, 4>n} 12 114>n 11 2
n=l m
= L Iif, 4>n}1 n=l
2
and therefore we have m
o < IIf -
L if, 4>n}4>nIl n=l
2
m
=
IIfll
2
-
L 1(f,4>n}1 n=l
2
.
,
9.2. Orthogonal Sets
177
Hence, m
L /(f, rPn)/2 < I/fl/ 2,
n=l
and since the sequence {am} = {L::=1 I (f, rPn) 12 } is a monotonic increasing sequence bounded above, we have that {am} converges to some a E JR and Bessel's inequality holds: 00
L I(f, rPn)/2 < I/fl/ 2. n=l
Bessel's inequality indicates that the Fourier series of f with respect to M is absolutely convergent (in the norm). This means that the Fourier series is convergent (in the norm) and that its sum is independent of the order in which terms are added. Thus, the series L~=l (f, rPn)rPn converges to some functionPMf EX. Now, iff E [M]..L (= M..L), all the Fourier coefficients off are zero, and thus P Mf = O. Alternatively, iff E [M], then it can be shown thatPMf = f. For the general f EX, the set [M] is a closed subspace in the Hilbert space X, and the projection theorem indicates that f = PMf + h, where PMf E [M] and h~M]..L. The function PMf is thus the Ilclosest" approximation in [M] to f. An orthonormal set M = {rPn} in the separable Hilbert space X is called an orthonormal basis of X if for every f EX, 00
f
= L(f, rPn)rPn. n=l
If M is an orthonormal basis, then m
lim "(f, rPn) rPn £...J
m-+oo
= f,
n=l
and using the derivation ofBessel's inequality (and continuity of the norm function) we have that
178
9. Hilbert Spaces and £2
We thus arrive at Parseval's formula 00
2
IIfll =
L I(f, rPn) 1 2
•
n=l
Suppose now that f E M.1. Since M is an orthonormal basis of X, f = L~l (f, rPn}rPn and (f, rPn) = for all rPn E M; consequently, by Parseval's formula we have that IIfll = 0, and the definition of a norm implies that f = O. In this manner we see that if M is an orthonormal basis, then M.1 = 0, i.e., M is a total orthonormal set. Yet another implication of M being an orthonormal basis is that [M] = X. This follows from the projection theorem, since [M].1 c M.1 = 0, and so
°
X
--
--.1
--
--
= [M] E9 [M] = [M] E9 0 = [M].
It is interesting that the above implications actually work the
other way as well. For example, if M is an orthonormal set in X such that Parseval's formula is satisfied for all f EX, then it can be shown that M is an orthonormal basis. In summary we have the following result:
Theorem 9.2.3 Let X be a separable Hilbert space and suppose that M = {rPnl is an orthonormal set in X. Then the following conditions are equivalent: (i) M is an orthonormal basis; (ii) IIf11 2 = L~l 1(f, rPn) 12 for all f E X; (iii) M is a total set; (iv) [M] = x. It can be shown that every separable Hilbert space has an orthonormal basis. Formally, we have the following result: Theorem 9.2.4 Let X be a separable Hilbert space. Then there exists an orthonormal basis for X. The proof of this result is based on a Gram-Schmidt process analogous to that used in linear algebra to derive orthonormal bases. Parseval's formula is remarkable in this context because it essentially identifies all separable Hilbert spaces with the space e2 . Now, e2 is a separable Hilbert space, and given an orthonormal basis M = {rPnl
i i
9.2. Orthogonal Sets
179
on a general separable Hilbert space X there is a mapping T : X -+ e2 defined by the Fourier coefficients if, cPn}. In other words, T maps f E X to the sequence {if, cPn}} E e2 . Parseval's formula shows that T is an isometry from X to e2 . On the other hand, given any sequence {an} E e2 , the function defined by L~I ancPn is in X, since X is a Hilbert space and the series is convergent. We thus have that any separable Hilbert space is isometric to the Hilbert space e2 . In fact, an even stronger result is available, viz., all separable complex Hilbert spaces are isomorphic to e2 . This means that at an Iialgebraic level" a complex separable Hilbert space is indistinguishable from the space e2 , i.e., at this level there is only one distinct space. A similar statement is true for real Hilbert spaces, where the space e2 is replaced by its real analogue. In passing we note that any Hilbert space X possesses an Ilor_ thogonal basis," though it may not be countable. The cardinality of the set is called the Hilbert dimension of the space. If X and Xare two Hilbert spaces both real or both complex with the same Hilbert dimension, then it can be shown that X and X are isomorphic. This is a generalization ofthe situation in finite-dimensional spaces, where, for example, all n-dimensional real Hilbert spaces are isomorphic to JRn. Exercises 9-2:
1. Show that for any integers m, n
1
cos(mx)cos(nx)dx =
[-Jr,Jr]
{](o, ,
ifm =1= n, if m = n,
and
1
sin(mx) cos(nx) ax = 0.
[-Jr,Jr]
2. The first three Legendre polYnomials are defined by Po(x) = I, PI (x) = x, and P2(X) = ~(3x2 - 1). (a) Show that the set P onL 2 [-I,I].
=
{Po, PI, P2} forms an orthogonal set
(b) Construct an orthonormal set M from the set p' and find the Fourier coefficients for the function f(x) = ex.
180
9. Hilbert Spaces and £2
3. The Rademacher functions are defined by rn(x) = sgn(sin(Zn.rrx)) for n
= I, Z, ... , where sgn denotes the signum function sgn C) x
={
-I, I,
ifx < 0, if x > 0.
(a) Show that the set M = {rnl is an orthogonal set on £2[0,1].
(b) Iff is definedbYf(x) = cos(ZJl'X), showthatf J.. rnforalln = I, Z, ... and deduce that M cannot form a total orthogonal set. 4. Let X be an inner product space and let M = {rPnl be an orthonormal basis of X. Show that for any f, g EX, 00
if, g}
= L if, rPk} (g, rPk). k=1
9.3
Classical Fourier Series
We saw in Section 9-1 that £2 is a Hilbert space, and we know from Theorem 8.6.1 that all IY spaces are separable. Theorem 9.2.4 thus implies that the space £2 must have an orthonormal basis. It turns out that the classical Fourier series (i.e., trigonometric series) lead to an orthonormal basis for £2[a, b]. In this section we study classical Fourier series and present some basic results with little detail. There are many specialized texts on the subject ofFourierseries such as [15] and [45], and we refer the reader to these works for most the details. A particularly lively account of the theory, history, and applications of Fourier series can be found in [Z4]. For convenience we focus primarily on the space £2[ -Jr, Jr] and note here that the results can be extended mutatis mutandis to the general closed interval. A classical Fourier series is a series of the form 1
Zao +
L (an cos(nx) + bnsin(nx)), 00
n=l
(9.1)
9.3. Classical Fourier Series
181
where the an's and bn's are constants. We know from Example 9-2-1 that for any nonzero integers m, n,
1
sin(mx) sin(nx) ax = {O,
[-Jr,Jr]
]"(,
~ffm =1= n,
1
m - n.
We also know from Exercises 9-2, No. I, that for all integers m, n,
1
cos(mx) cos(nx)
[-Jr,Jr]
ax =
{' 0
if m =1= n,
]"(,
if m
= n,
and
1
sin(mx) cos(nx) ax
= o.
[-Jr,Jr]
In addition, it is evident that for any integer n,
1
sin(nx) ax
=
[-Jr,Jr]
Let
1
cos(nx) ax
= o.
[-Jr,Jr]
= {rPn} and \II = {'lfrn} denote the sets of functions defined by rPn(x)
=
cos(nx) ~
,
'lfrn(x)
=
sin(nx) ~
,
for n = I, 2, .... Then the above relationships indicate that the set S = {11.J2]r} U U \II forms an orthonormal set in £2[-]"(,]"(]. The Fourier coefficients of a function [ E £2[_]"(,]"(] with respect to S are given by
1) .J2]r 11 (, .J2]r f, -
= -
(f, rPn) = (f, 'lfrn) =
[-Jr,Jr]
1 ~1
~
'\/ ]"(
[-Jr,Jr]
'\/ ]"(
[-Jr,Jr]
[(x)
ax ,
[(x) cos(nx) ax, [(x) sin(nx)
ax,
and the Fourier series is (9.2)
\
,
182
9. Hilbert Spaces and £2
which is equivalent to expression (9.1) with the familiar coefficient relations ao
.!.1 .!.1 = .!.1 =
an = bn
f(x) dx,
re
[-Jr,3l']
re
[-Jr,Jr]
re
[-Jr,Jr]
f(x) cos(nx) dx, f(x) sin(nx) dx.
Fourier series can be expressed in atidier, more symmetric, form using the relation einx = cos(nx) + i sin(nx). Using this relation, the series (9.1) can be written in the form 00
inx
'L...J " cn e
(9.3)
,
-00
where the
Cn
are complex numbers defined by Cn
The set of functions B
= -1
1
2re [-Jr,Jr]
= {f3n},
.
f(x)e znx dx.
where
f3n(x) =
einx
$'
forms an orthonormal set for L~[ -re, re], and for any f E L~[ -re, re] (and consequently for any f E L 2 [ -re, reD the corresponding Fourier series is 00
PBf =
L(f, f3n)f3n.
(9.4)
-00
Now, for any f E L 2 [-re, re], Bessel's inequality guarantees that the Fourier series (9.2) (and (9.3)) converges in the \I . 112 norm to some function Psf E L 2 [ -re, re]. The central question here is whether the set S forms an orthonormal basis for L 2 [-re, re], so that Psf = f a.e. In fact, it can be shown that S forms an orthonormal basis for L 2 [ -re, re]. If we combine this fact with Theorem 9.2.3, we have the following result:
9.3. Classical Fourier Series
183
Theorem 9.3.1 Let f E L 2 [-re, re] and let
k=n
=
L
if, 13k}13k·
k=-n
Then: • (i) IIsn - fll2 -+ 0 as n -+ 00; (ii) Parseval's relation is satisfied:
Ilfll~ = 1if, 1}1 2+
00
L (I if, rPn}1
2
+ 1if, 1/Ik}1 2 )
n=l
re
00
2 2 2 = _a 20 + re "Ca L....Jn + bn)
n=l 00
00
= L 1if, f3n}1 = 2re L Icn l2 . 2
-00
-00
From the last section we know that all separable Hilbert spaces are isomorphic to the Hilbert space £2 and Parseval's relation is a manifestation of this relationship. This observation leads to the following result: I
Theorem 9.3.2 CRiesz-Fischer) Let {c n } E £2. Then there is a unique function f E L 2 [-re , re] such that the Cn are the Fourier coefficients for f. Note that Ilunique" in the above theorem means that the sequence {c n } determines an equivalence class of functions modulo equality a.e. An immediate consequence ofTheorem 9.3.1 and the projection theorem (Theorem 9.2.2 ) is the following result: Theorem 9.3.3 Let f E L 2 [-re, re]. Then for any E > 0 there exists a positive integer n and a trigonometric polynomial ofdegree n, say On = L~-n qkf3k such that liOn - fl12 < E. Moreover; among the trigonometric polynomials of
184
9. Hilbert Spaces and £2
degree n, the closest approximation to f in the II . 112 norm is that for which the qk correspond to the Fourier coefficients. In other words, trigonometric polynomials can be used to approximate any function in L 2 [-re, re], and the Fourier coefficients provide the best approximation among such polynomials in the II . 112 norm. Theorem 9.3.1 guarantees that the Fourier series will converge in the II ·112 norm, but this is not the same as pointwise convergence, and an immediate question is whether or not a Fourier series for a function f E L 2 [ -re, re] converges pointwise to f. More explicitly, for a fixed x E [-re, re] does the sequence of numbers {sn(x)} converge, and if so does sn(x) -+ f(x) as n -+ oo? The answer to this question is complicated, and much of the research on Fourier series revolved around pointwise convergence. Some simple observations can be made. First, the orthonormal basis defining the Fourier series is manifestly periodic, so that if sn(x) converges for x = -re, it also converges for x = re, and s( -re) = sere). The existence of a Fourier series does not require thatf( -re) = f(re), so pointwise convergence will fail at an endpoint unless f satisfies this condition. More generally, we see that any two integrable functions f and g such that f = g a.e. produce the same Fourier coefficients, so that for a specific function f, we expect that the best generic situation would be that sn(x) -+ f(x) a.e. We shall not Ilplumb the depths" of the vast results concerning pointwise convergence of Fourier series; however, we will discuss two results of interest in their own right that make crucial use of the Lebesgue integral. Prima facie, it is not obvious that the coefficients an and bn in the Fourier series have limit zero as n -+ 00. Given that arguments such as x = 0 lead to series such as L~=o an, this is clearly a concern. An elegant result called the Riemann-Lebesgue theorem resolves this • concern and is of interest in its own right. We state the result here in a form more general than is required for the question at hand.
Theorem. 9.3.4 (Riemann-Lebesgue) Let I C JR be some interval and f E L 1 (f). If {An} is a sequence of real numbers such that An -+ 00 as n -+ 00, then
1
{(x) caS(AnX) ax ---> 0 and
as n -+
00.
1
{(x) sin(AnX) ax ---> 0
9.3. Classical Fourier Series
185
Proof The proof of this result is of particular interest because the convergence theorems of Chapter 5 are of little help, and we must return to the definition ofthe integral itself. We sketch here the proof for the cosine integral. Suppose first that I = [a, b] and that f : I ~ 1R is bounded on I. If M denotes an upper bound for IfI, then [
f(x) COS(AnX) dx < M
J[a,b]
cos(Anx)dx
[ J[a,b]
=M
l
b cOS(AnX)d<
M = -lsin(Anb) - sin(Ana)1
An
2M <-
- An .
Since An ~ 00, as n ~ 00, we have that ~a,b] f(x) COS(AnX) dx ~ O. This calculation shows in particular that the theorem holds for any step function having a support of finite length. Now, the Lebesgue integral is defined in terms of a sequence of a-summable functions, with a = x, and by definition these functions have supports of finite length. If f E L 1 (!), then we know that there is a sequence ()j of x-summable functions such that
1,lf(X) - OJ(x)1 d< --+ a asj
~ 00.
Choose any E > 0, and select aj sufficiently large so that [If(x) - ()j(X) I dx <
J1
E.
2
Now, 1,f(X) cOS(AnX)d< 1,(f(x) - OJ(x) + OJ(x)) cOS(AnX) d< < 1,(f(X) - OJ(X)) COS (AnX) d<
+
1,OJ(X)cOS(AnX)d<
186
9. Hilbert Spaces and £2
< IIf(X) - OJ(x)1 ax + 10j(X) COS(AnX) ax < ;
+
l 0j(X) COS(AnX) ax
.
From the above discussion we know that l0j(X) COS(AnX) ax -*
as n
~ 00
a
for any j, so that there is an integer N such that E
< 2
whenever n > N. Therefore,
I
E
[f(x) cos(Anx)dx < 2
E
+2 =
E
whenever n > N, and by definition this means that 1 [(x) COS(AnX) ax -*
as n
~
a
o
00.
The Riemann-Lebesgue theorem with I = [-re, re] and An = n shows that the Fourier coefficients tend to zero as n ~ 00. This theorem is also crucial in proving another notable result known as the Riemann localization theorem, which we shall not prove. Theorem. 9.3.5 (Riemann Localization) Let f E L 1 [ -re, re] and x E [-re, re]. Then for any fixed 8 with 0 < 8 < re, sn(x)
~
f(x) if and only if
•
lim ~ [ f(x + t) + f(x - t) sin((n + ~ )t) dt = n~oo 2re 1[0,0] t 2
o.
Here it is assumed that f has been extended periodically to a function on 1R so that f is defined for arguments x ± t that may not be in [-re, re]. What is interesting in the above theorem is that the number 8 can be arbitrarily small, and this means that the pointwise convergence of the Fourier series depends only on the values that f
if
9.3. Classical Fourier Series
187
assumes in a small neighborhood of x. Given that the Fourier coefficients in the series defining Sn depend on the values f assumes in the entire interval [-re, re], this result is remarkable. We can study Fourier series outside the comfortable space 2 L [-re, re]. Naturally, we lose the results that rely on L 2 [ -re, re] being a Hilbert space such as Parseval's relation, but in the larger space L 1 [ -re, re] the Fourier coefficients are still well-defined, and results such as the Riemann-Lebesgue theorem are still valid. We even have the following uniqueness result: •
Theorem. 9.3.6 Iff, gEL1 [ - re, re] have the same Fourier coefficients, then f = g a. e. What is needed, however, is some result that shows that the partial sums of the Fourier series for a function f E L 1 [ -re, re] converge in the 11·111 norm to f, but this is where things go wrong. Iff E L 1 [ -re, re], then in general we do not have that IIsn - fill ~ 0 as n ~ 00. It can be shown, however, that for 1 Un
=-
n
n-1
LSk, k=l
I\un - fllI ~ 0 as n ~ 00 for any f E L 1 [ -re, re]. The quantity Un is called the Cesaro mean of the partial sum sequence {sn}. This is a weaker convergence result, since Un effectively measures an averaged partial sum. The sequence {un} may converge even if {sn} diverges, and if {sn} converges to some s, then {un} also converges to some function s E L 1 [ -re, re]. In the space L 1 [ -re,n] this is the sharpest result we can get. In fact, it can be shown that there are functions in L1 [-re, re] such that {sn(x)} diverges a.e. Though most of the above results concerning pointwise convergence were established by the early twentieth century, it was not until the 1960's that Carleson [7] proved that is f E L 2 [ -re, re] then {sn(x)} ~ f(x) a.e. From this perspective, the space L 2 [-re, re] is the natural space in which to study Fourier series. Exercises 9-3: 1. Let f(x)
= x on the interval [-re, re].
(a) Determine the Fourier series for f.
188
9. Hilbert Spaces and £2
(b) Show that the series obtained by differentiating term by term the Fourier series in part lea) is a divergent series. 2. The complex Fourier series for a function f E £2[ -A, A], A > 0, is given by 00
PMf(x)
= LCneirurxlA, -00
where en
= ~ i>(x)e-inm
for n = 0, ±l, ±2, ... , and it can be shown that the set M _ irurx1 {e A} sUitably normalized forms an orthonormal basis for £2[-A,A].
(a) Find the complex Fourier series for the function g : ~ lR defined by
[-!,!]
g(x)
= { ~l,
if _1 < x < 0 2 ' if 0 < x <
!.
(b) Use Parseval's relation and the Fourier series in 2(a) to prove that 1
00
]"(2
~ (2n + 1)2 = 8' 3. Suppose thatf has a continuous derivative on the interval [-]"(, ]"(], and let an, bn denote the Fourier coefficients off. Use integration by parts to prove the Riemann-Lebesgue theorem for this special case.
9.4
The Sturm-Liouville Problem
Theorem 9.2.4 indicates that for any interval I the Hilbert space £2(1) must have an orthonormal basis. If I is bounded, then the trigonometric functions provide such a basis. Theorem 9.2.4 does not, however, preclude the existence of other orthonormal bases, and there are, in fact, any number of orthonormal bases available. If
i i
9.4. The Sturm-Liouville Problem
189
I is not bounded, then the trigonometric functions are not even in the space L 2 (1), and the classical Fourier expansions are no longer
valid. Nonetheless, Theorem 9.2.4 guarantees the existence of an orthonormal basis for any interval I. In practice, many of the orthonormal bases for L 2 (1) arise as solutions to boundary value problems of the Sturm-Liouville type. A regular Sturm-Liouville problem consists in determining a solution y to a differential equation of the form
;
(r(x):)
+ (q(x) + J..p(x))y = a
(9.5)
on some interval [a, b], satisfying boundary conditions of the form k1y(a)
+ k2y'(a) = 0,
lly(b)
+ l2y'(b) = O.
(9.6)
Here, r, q, andp are given functions, J... is a complex parameter, and the k/s and ~'s are constants such that + k~ =f:: 0, + l~ =f:: O. In addition, it is assumed for the r¢gular Sturm-Liouville problem that the functions r andp are nonzero in the interval [a, b]. Sturm-Liouville problems are important in theory and applications of differential equations. As a consequence, this class of boundary value problems has been studied exhaustively for some 150 years. Relatively accessible accounts of the theory can be found in [5], [9], [11], [23], and [40]. In addition, many applications-oriented texts such as [3], [8], and [26] contain short summaries of the theory and various applications. We do not attempt here to replicate the general theory in any detail or depth, but merely focus on the Sturm-Liouville problem as a Ilmachine" for producing orthogonal sets in L 2 • Equation (9.5) is often written in the abbreviated form
kr
[,y = -J...p(x)y,
lr
(9.7)
where the linear operator [, is defined by
Cy =
;
(r(x):)
+ q(x)y.
(9.8)
If Y E G2[a, b], r E G'[a, b], and q E G[a, b] then Ly E G[a, b], and in particular £y E L 2 [a, b]. For linear operators such as [, there exists another operator [,* called the (Hilbert) adjoint. The adjoint operator
190
9. Hilbert Spaces and L 2
satisfies the relation (9.9) for all Yl, Y2 E L 2[a, b], which alsc:> satisfy the boundary conditions (9.6). The peculiar form of the operator £ for the Sturm-Liouville problem ensures that the operator is self-adjoint, Le., £ = £*. This can be verified directly as follows: (£Yl, Y2) = {
ira,b]
WI, £Y2)
(Y2(X)£Yl (x) - Yl (X)£Y2(X))
ax
=
la'bl; ((r(x)
=
[rex) (Y2(X) ;Yl(X) - Yl(X~ ;Y2(X)) J:
=
o.
;Yl(X)) Y2!CX) -
(r(x) ;Y2(X)) Yl(X)) ax
The final equality follows from the condition thatYl andY2 satisfy the same boundary conditions (9.6). (The specific details can be found in [26]). As we shall see shortly, the self-adjointness of £ is the key to obtaining orthogonal sets. For any value A E C, the Sturm-Liouville problem always has one obvious solution, viz., the triVial solution Y = o. Generically, this is the only solution available for a given value of A, but there may be some values of A for which nontrivial solutions exist. These values are called eigenvalues, arid the associated nontrivial solutions are called eigenfunctions. The set of all eigenvalues for the problem is called the spectrum2
i
i
9.4. The Sturm-Liouville Problem
191
The regular Sturm-Liouville problem has fairly tractable spectral properties, as the next theorem illustrates.
Theorem 9.4.1 Supposethatr E C1 [a,b],p,q E C[a,b], andthatr(x) > Oandp(x) > 0 for all x E [a, b]. Then: (i) the spectrum for the Sturm-Liouville problem is an infinite but countable set; (ii) the eigenvalues are all real; (iii) to each eigenvalue there c(J)rresponds precisely one eigenfunction (up to a'constant factor), i.e., the eigenvalues are simple; (iv) the spectrum contains no finite accumulation points. Thus, under the conditions of the above theorem, the SturmLiouville problem always has an infinite (but countable) number of eigenfunctions. The spectrum is a countable set, so we can regard it as a sequence V.n} and impose the condition IAn I < IAn+ll for all n. Since there are no finite accumlll1ation points, we see that IAnI ~ 00 as n ~ 00. Suppose that the conditions of Theorem 9.4.1 are satisfied, and that Am and An are distinct eigenvalues with corresponding eigenfunctions Ym, Yn, respectively. Then, £Ym = -AmP(X)Ym, £Yn
= -r-AnP(x)Yn,
and therefore Ym£Yn - Yn£Ym = (Am - An)P(X)YmYn.
The above equation indicates th~t
1
(Ym(X)£Yn(X) - Yn(X)£Ym(X)) dx
[a,b]
= (Ym, £Yn) - (Yn, £Ym)
= (Am -
An)
1
P(X)Ym(X)Yn(X) dx.
[alb]
Now, the eigenfunctions are real, so that (Ym, £Yn) since £ is self-adjoint, (Am - An)
1
P(X)Ym(X)Yn(x}dx
[a,b]
i i
= (£Yn, Ym) -
= (£Yn,Ym), and (Yn, £Ym)
= O.
192
9. Hilbert Spaces and L 2
By hypothesis, the eigenvalues are distinct (Am =f. An), and the above calculation shows that [
ira,b]
P(X)Ym (X)Yn (X) dx =
o.
(9.10)
If P = I, the above equation im(plies that Ym ..L Yn for m =f. n, and thus the eigenfunctions are orthgonal. By hypothesis we have that p(x) > 0, so that in any event the set of functions {.JPYn} is orthogonal. As the eigenfunctions are by definition nontrivial solutions, we know that IIPYn 112 =f. 0, so that this set of functions can always be normalized to form an orthonormal set in L 2 [a, b]. The Sturm-Liouville problem thus produces eigenfundtions from which orthonormal sets can be derived. Given a continuous function p positive on the interval [a, b], it is always possible to define anothet norm for L 2 [a, b], viz., pllYII2 = { [
ira,b]
P(X)y2(X)dx}1/2
(9.11)
Since P is positive and continuous on [a, b], there exist numbers Pm and PM such that 0 < Pm < p(x) tS PM for all x E [a, b]. This implies that PmllyII2 < pllYII2 < PMIIYII2,
so that the p 1\·112 norm is equivalent to the 1\·112 norm. Consequently, the vector space L 2 [a, b] equippe
= [ P(X)YI (X)Y2(X) dx, ira,b]
then it is readily verified that (', .}p is an inner product and thatpll . 112 is the norm induced by this inn~r product. Thus the inner product space (L 2 [a, b], (., '}p) is a Hilbert space. We use the (standard) notation L 2 ([a, b],p) to denote this Hilbert space, with the abbreviation L 2 [a, b] for the space L 2 ([a, b], 1).. If equation (9.5) satisfies the conditions of Theorem 9.4.1, the above arguments indicate that the normalized eigenfunctions
i i
9.4. The Sturm-Liouville Problem
193
{Yn/p llYn 1121 form an orthonormal set in the Hilbert space L 2 ([a, bJ,p). In fact, this set forms an orthonormal basis for L 2 ([a, b],p).
Theorem 9.4.2 Suppose that equation (9.5) sati$fies the conditions of Theorem 9.4.1. Then the collection of normalized eigenfunctions forms an orthonormal basis for the Hilbert space L 2 ([a, b],p). Example 9-4-1: Fourier Sine Series
Consider the differential equation yl/(x) + AY
= 0,
(9.12)
with boundary conditions yeO)
= 0,
yen)
= o.
(9.13)
If A < 0, then the general solutibn to the differential equation (9.12) is y(x) = Ae Rx
+ Be- Rx ,
where A and B are constants. The boundary conditions (9.13), however, indicate that A = B = 0, so that only the trivial solution is available in this case. Thus, this problem does not have any negative eigenvalues. If A = 0, then the general solution is Y(x) ::;: A~
+ B,
where A and B are constants. Again, the boundary conditions imply that A = B = 0, so that A = 0 lcannot be an eigenvalue. If A > 0, then the general solution is y(x) = A cos( vlIx) + B sine,..fix),
where A and B are constants. 'the condition yeO) = 0 implies that A = 0, and the condition yen) ::::::l 0 implies that B sine Jin) =
o.
(9.14)
Equation (9.14) is satisfied for BI =f. 0 only if A = n 2 for some integer n, and in this case equation (9.12) has the nontrivial solution Yn(x) :;::: sin(nx).
i I
(9.15)
194
9. Hilbert Spaces and L 2
The set {n 2 } corresponds to the spectrum. Hence from Theorem 9.4.2 we know that the set {Yn/IIYnIl21 =l:: {Yn.J2hr} forms an orthonormal basis for L2 [0, n'). Example 9-4-2: Mathieu Functions Consider the differential equation Y"
+ (A -
2B c'os(2x))y
= 0,
(9.16)
along with the boundary conditions (9.13). Equation (9.16) is called Mathieu's equation, and B is some fixed number. Here, rex) = p(x) = 1 and q(x) = -2B cos(2x). Note ~hat when B = 0 equation (9.16) reduces to equation (9.12). Now, unlike the previous example, we cannot solve equation (9.16) in clbsed form, and it is clear that the eigenvalues will depend on the parameter B. Nonetheless, the above results indicate that for any B, there is a set {An (B)} of eigenvalues with corresponding eigenfunction;s that when normalized will yield an orthonormal basis for L 2 [O, Jr]. The solutions to equation (9.16) are well-known special functions call~d (appropriately enough) Mathieu functions. The intricate details cdncerning these functions can be found in [28] or [41]. Suffice it h~re to say that corresponding to the spectrum {An(B)}, Mathieu's eqllationhas eigenfunctions sen (x, B) that are periodic with period 2Jr artd reduce to sine functions 3 when B = o. The Mathieu functions {s~n(x, B)} thus form an orthogonal basis for L 2 [O, Jr]. The Sturm-Liouville problem; can be posed under more general conditions. These generalizatlions lead to bases for L2 that are widely used in applied mathematics and numerical analysis. The generalizations commonly made Gorrespond to either: (i) relaxing the conditions on p and r at the endpoints so that these functions may vanish (or even be discontinuous) at x = a or x = b (or both); or (ii) posing the problem on an unbounded interval. The general solutions to the differential equations with these modifications are usually unbounded on the interval, and the homo3The notation sen comes from Whit!taker and Watson [41] and denotes "sine-elliptic." There are "cosine-elliptic" functions cen with analogous properties.
i i
9.4~
The Sturm-Liouville Problem
195
geneous boundary conditions (9.6) are often replaced by conditions that ensure that the solution i~ bounded, or that limit the rate of growth of the function as x appr~aches a boundary point of the interval. These generalized versions of the Sturm-Liouville problem are called singular Sturm-Liouville problems. The theory underlying singular Sturm-Liouville problems and the corresponding results are more complicated than those fd>r the regular Sturm-Liouville problems. For example, the spectruIill may consist of isolated points or a continuum, and not every poirtt in the spectrum need correspond to an eigenvalue. The singular Sturm-Liouville problem is studied in some depth in [9] and [40]. More general references such as [5] and [11] give less detailed but clear, succinct accorunts of the basic theory. We content ourselves here with a few exanp.ples that lead to well-known bases for L 2 . The special functions atising in these examples have been studied in great detail by numer
Consider the Legendre differenti4l equation
~ ((l-~)~) Hy =
0
(9.17)
on the interval (-1,1). Note thctt rex) = 1 - x2 is zero at x = ±1, so that this equation leads to a sin~lar Sturm-Liouville problem. The general solution to equation (9.!17) can be found by using a power series method, which seeks soltltions of the form 00
y(x) :::::
L anxn,
(9.18)
n=O
where the an's are constants (cf~ [5] for details of the method). Substituting power series (9.18) into the Legendre differential equation and equating the coefficients of:xn to zero for n = 1,2, ... yields the recursive relation 1'1r(n + 1) - A n a +2 = + l)(n + 2)'
(n
Once ao and al are specified, the'\above relation determines the other an's uniquely. Specifically, the recursive relation defines two linearly
196
9. Hilbert Spaces and L 2
independent solutions Ye and Yo ;corresponding to the choices ao = I, al = 0 and ao = 0, al = I, respectively. HereYe is an even solution and Yo is an odd solution. Suppose now that we require the solution to be bounded as x ~ ±l. If A # n(n + I), we can apply the ratio test to the series defining Ye (or,yo) to establish that the radius of convergence is I, and it can be shqwn thatYe (and Yo) are unbounded in the interval (-1, 1). We thus n~ed A = n(n + 1) for some positive integer n to get bounded solutions. The eigenvalues for the problem are thus An = n(n + 1). The corrresponding eigenfunctions are the polynomials formed by the trunc~ted series for Ye and Yo (modulo a scaling factor). Specifically, the eigenfunctions Pn corresponding to An are defined by A1
.
'"'" m ' (2n - 2m)! n-2m Pn(x) = m=O -1) 2n m. In m. n 2m )'x , .
'J _ )'( _
L.,c
where M = n/2, or (n -1)/2, whichever is an integer. For example, PO(x) = I, P2(x)
PI (x)
= ~(3X2 -
1), .P3(x)
= x, = ~(5x3 -
3x).
The peculiar form of the polynomial coefficients is standard: The polynomials have been scaled so!that Pn (1) = 1 for all n. This corresponds to choosing the last coefficient Cn in the polynomial P n to be C n
=
(2n)! 2n (n!)2
-+1":""-''':''-
and using the recursive relation to get the lower-order coefficients. The functions Pn are called the £~gendre polynomials. Note that the " proof of orthogonality follows imnilediately from the self-adjointness of the operator. The boundary values for the Pk do not matter, because r(l) = r( -1) = O. Thti Legendre polynomials, suitably normalized, form an orthonorma] basis for £2[-1,1]. ,
Example 9-4-4: Hermite Polynomials
The Hermite differential equation is d 2y dy - 2 - 2x---+ dx
ax + AY = O.
(9.19)
9.4. The Sturm-Liouville Problem
197
This equation is not in the seU-adjoint form of equation (9.5), but since -d ( e_X2dY) -
ax
and e- X
2
=1= 0
ax
for all x
E
2 2y Y) =c _x (d- 2 xd2
ax
lR, equatton (9.19) is equivalent to
d ( -x2dy~
ax
ax'
e
ax j + Ae
-x2
(9.20)
Y = O.
The singular Sturm-Liouville pI10blem consists in finding solutions yto equation (9.20) on the intervfil (-00,00) such that ly(x)1 does not grow exponentially as x ~ ±oo. :Substituting the power series (9.18) into the equivalent equation (9.19) yields the recursive relation
an+2 = (n
2n-A
+l)(n + 2) an·
As with the Legendre equation,! this recursive relation defines an even and an odd solution. If A tj; 2n, then the ratio test indicates that the series converges for all x: E JR. The solutions in this case will have exponential growth. In or¢ier to meet the growth condition, we must therefore have that A= 2n for some n = 0,1,2, .... If A = 2n, then the series (9.18) refluces to a polynomial of degree n. The eigenfunctions are commotily given in the form Ho(x)
= I, 2
Hn(x) = (_I)n C ::n
(e-
x2
) ,
and called Hermite polynomials. The first few Hermite polynomials are
H3(X) = 8x3 - 12x, H 4 (x) = 16x4
-
48x2
+ 12.
Note that for the self-adjoint equation (9.20), we have that p(x) = 2 e-x , and therefore for m =1= n,
1
(-00,00)
e-x2 Hn(x)Hm(x)
dx =
o.
198
9. Hilbert Spaces and L 2
It can be shown that the Hermit¢ polynomials form an orthogonal basis for L2(~, e- xl ). If en is defin¢d by
en(x)
=
J2
1 n
then it can be shown that the L2(~, e- X2 ).
1
n!',fii s~t
e-x2 /2Hn(x) ,
{en} is an orthonormal basis for
Example 9-4-5: Laguerre Polynomials A basis for L2(0, 00) can be deriKred from the Laguerre differential equation d 2y x -2 dx
+ (1 -
dy
~)i
dx
+ Ay =
0
'
(9.21)
which in self-adjoint form is (9.22) The singular Sturm-Liouville prqblem consists in solving equation (9.22) on the interval (0,00) subje:ct to the conditions limx.-+o+ y(x) < 00, limx.-+ oo e-Xy(x) = O. As with the Hermite equation, the growth conditions lead to a restriction the values of A. For the Laguerre differential equation, the eigenv~ues are An = n for n = 1,2, .... The standard representation for 1lhese eigenfunctions is
0*
Lo(x) Ln(x)
= I, ~. d n
= ---+1' ax (xne- X). n! n
The functions L n are polynomials of degree n; they are called Laguerre polynomials. The first few Laguerre polynol11ials are L 1(x)
=
L3(X)
=1-
L~(x)
= 1-
2x + ~X2,
3x + ~X2 - ~x3, L i4 (x)
=1-
4x + 3x2 - ~x3 - 214X4.
1 - x,
,
The Laguerre polynomials fortm an orthogonal basis for the space L 2 ((O, 00), e- X ) The above examples are a s1ffiall sample of the many IIspecial functions" in mathematics that c¢lrrespond to bases for L 2 • The list of
i
I
9.5. Other Bases for £2
199
well-known special-function bases for L2 is extensive. Aside from the references given above, the read~r is also directed to the monograph of [19], which lists most the bases for L 2 in common use and gives another perspective on total set$ for L 2 . Exercises 9-4: 1. Using the properties of the inner product and self-adjoint opi
erators, prove that a Sturrq.-Liouville problem satisfying the conditions of Theorem 9.4.1 must have only real eigenvalues. 2. Solve equation (9.12) subject to the boundary conditions yeO) = 0, y'(1) = 0 and determine the eigenvalues and corresponding
eigenfunctions. 3. Verify by direct calculation that the Hermite polynomials Ho, HI,
and H2 of Example 9-4-4 fotm an orthogonal set in the space L2CIR, e- X2 ) .
9.5
Other Bases fot £2
The Sturm-Liouville problem cfln be used to generate numerous bases for L2 spaces; however, aJjly solution to equation (9.5) must be at least twice differentiable, and consequently, any basis derived from a Sturm-Liouville problem '!must consist of IIsmooth" functions with at least two derivatives. Giyen the nature of the functions in L 2 , one expects that bases consis~ing of nonsmooth functions should also be available. In this section we present a brief example of such a basis for the space L2 [0, 1]. The Haar functions hn : [0, 1] ~ 1R are defined by :
2k12 if x
,
E [i-1 i-ll2) T'~'
O,otherwise,
i i
200
9. Hilbert Spaces and £2
where l = 1,2, ... , 2k and k k = 0 and l = I, so that h 2 (x)
Ifn = 3, then k
=
= 0,1, .... For example, if n = 2, then
{
1, if x
E
[0, ~),
-1" if x
E
[!' 1].
= 1 and l = I, and thus 21/ 2 , ifx
E
[0 , 1) 4 '
_2 L12 , if x E [14' 1] 2 ' P,ifXE(!,I].
Ifn
= 4, then k = I, l = 2, and
p,
if x
E
[0, !),
if x
E
[12' ~] 4 '
- 21112 , if x
E
(~, 1].
21~2
,
The support of the Haar function h2k+t is the interval [(l 1)/2 k ,l/2 k ]. Figure 9.1 depicts thd functions h s through h 8 • The set of Haar functions H ~ {hnl forms an orthonormal basis for £2[0, 1], The proof that H is a t?tal set can be found in [19]. Here, we show that H is an orthonorm*l set in £2[0, 1]. The normality of the Haar functions is simple to ~stablish. Evidently, IIhl ll2 = 1; if n = 2k + l, then
Hence IIh n ll 2 = 1 for all n = 1,2,. ~ .. To establish orthogonality note first that hI 1. hn for all n = 2,31, .... Suppose that k is fixed, and
i
i
9.5. Other Bases for £z
hs(x) 2
h6(X) 2
I
-
1 1 1 1
0
-2
1/4
1 1
1
1
1
1
1
1
I
1
>-x
0
1
1
1
1
1
1/2
1
1 1
1/4
1
t
1
1
I
I
I
1
1
>
X
1
-2 ----
h8(X)
2 ----
-
2 --------
1
1 1 1
-2
1
1
h7(X)
0
201
1/2
1
1
I
1
1
3/4
1
1
1
1
1
1
I
-----
1
1
1
1
X
0
3/4
1 1 1 1
1
X
1 1
1
1
-2 --------
1
FIGulrn 9.1
consider the functions h 2k+l for.e: = 1,2, ... , 2k . The support ofh2k+l is the interval Ik,l = [(.e -I)/2 k , .e/2 k ], and if.e 1 t:: .e 2 , the set Ik,ll nh,lz contains at most one point. Therefore, h 2k+l l (x) h 2k+l z (X) = 0 a.e., so that
r
J[O,l]
h 2k+l l (X)h 2k+l z (x)dx .
=
o.
Consequently, we have that h2k111 ..L h 2k+l z for all.e 1 t:: .e 2 . Suppose now that k1 > k2 and consider; the functions h2kl+ll, h2kz+l z with supports Ikl,lll Ikz,lz' respectively., Since k1 > k2 , the length of Ikl,ll is at most half that of hz,lz' and sin:ce any endpoint in a support for a Haar function must be of the form m/2 n for m, n = 0, I, 2 ..., the set Ikl,ll n Ikz,lz will be one of the follpwing: (i) the empty set; !
(ii) a set consisting of a single (iii) the set hl,ll.
~ndpoint
i
i
of Ikl,ll;
202
9. Hilbert Spaces and L 2
Cases (i) and (ii) indicate that h2kItlI (x)h 2k2+l z(x) = 0 a.e. and therefore h 2kI+l l ..L h 2k2+l z . For case (cJ, note that in hIll' the function h 2k2+l z does not change sign; henc~, ,
[, [0,1]
h2" H, (x)h2"+l, (x)
ax l= ±2k,12
I
(
h2"
+l, (x)
ax
,J1kI,ll
;:: O.
Thus, for case (iii), h 2kI+l l ..L h 2k2+fz' The above arguments indicate that hm ..L hn for m =j:. n, and the set H is consequently orthonormal. Series of Haar functions may ~e used to represent measurable functions. If f : [0,1] ~ lRe is firlite a.e. on the interval [0,1] and measurable, then there exists a semes of the form L~=l Clnhn(x) that converges a.e. on the interval [0, 1~ to f. In passing we note that the Rademacher functions rn of Exercises 9-2-3 do not form a basis for £2[0, 1];i however, another set offunctions called the Walsh functions W n cari be formed from the products of Rademacher functions. The set ofWalsh functions {wnl forms a basis for £2[0,1], and these functions hare applications in probability and communication theory. For a short discussion of these functions the reader is referred to [19].
i i
Epilogue CHAPTER
10.1
Generalizations of the Lebesgue Integral
The Lebesgue-Stieltjes integral may be generalized in any number of ways to accommodate, for example, higher dimensions. Rather than pursue extensions of this nature, we choose to describe informally and briefly a generalization of the Lebesgue integral whose origins go back to Newton. Newton regarded the integral of a function as being the antiderivative of that function. Formally, a function [ : [a, b] --* lR is said to be Newton integrable on the interval [a, b] if there exists a differentiable function F defined on [a, b] such that F'(x) = [(x) for all x E [a, b]. The definition of the Newton integral is evidently very limiting: not even step functions, strictly speaking, have Newton integrals. On the other hand, there are functions that have Newton integrals but are not Lebesgue integrable. Consider, for example, the function F defined by 2
F(x) = { x sin(l/x 0
2 )
if x =1= 0,
if x = o.
203
204
10. Epilogue
Now, F is evidently differentiable at every point in the interval (0,1]. At x = 0 the definition of the derivative can be used to establish that F is differentiable there as well. Consequently, the function f defined by f(x) = F'(x) is Newton integrable in the interval [0, 1]. It can be shown, however, that this function is not Lebesgue integrable on [0,1]. It is interesting to note that the improper Riemann integral also exists for f on the interval (0, 1]. The essence of the problem with the Lebesgue integral is that the theory is restricted to absolutely convergent integrals. Recall that if a function f is Lebesgue integrable, then the function IfI must also be Lebesgue integrable. This seems quite a harsh restriction, and it filters out conditionally convergent Riemann integrals as well as certain Newton integrals. Newton's definition of an integral is very natural to the student of elementary calculus and useful, particularly in fields such as differential equations. The class of Newton integrable functions is not contained by the class of Lebesgue integrable functions, and this awkward situation was soon realized by mathematicians. Within fifteen years of Lebesgue's pioneering work two mathematicians arrived independently at a generalization of the Lebesgue integral that would mend the awkward gaps in the definition where improper Riemann integrals and/or Newton integrals exist, but the Lebesgue integral does not. In 1912, A. Denjoy made a generalization directly from the Lebesgue integral. The definition of the Denjoy integral proved a complicated affair, and as a result, some of its potential for applications and generalizations was lost. In 1914, O. Perron devised another integral that would also remedy the problem. Rather than start with the Lebesgue integral, Perron devised a definition based on upper and lower integrals defined by function~ whose derivatives are respectively greater than and less than the given function. Roughly, a function A is said to be a major function off on the interval [a, b] if its derivative A' satisfies A'(x) > f(x) for every x E [a, b]. A function B is said to be a minor function for f if - B is a major function for -f. A functionf is said to be Perron integrable on the interval [a, b] if f has both major and minor functions, and
-00
< inf{A(b) - A(a)} = sup{B(b) - B(a)} <
00.
10.2. Riemann Strikes Back
205
Here, the infimum is taken over all major functions of f on [a, b], and the supremum is taken over all minor functions of f on [a, b]. The common value of the infimum and supremum is defined to be the Perron integral of f on [a, b]. Note that Perron's extension immediately includes the class of Newton integrable functions. This is because the primitive F of a Newton integrable function f is at the same time a major and a minor function. Although the Perron integral patched up the gaps with the Lebesgue integral, it suffered from the same problems as the Denjoy integral. The Perron integral and the Denjoy integral are defined very differently, and for a time it was thought that they characterized different functions as Ilintegrable!' As it turns out, however, these integrals are equal, Le., f is Denjoy integrable if and only if it is Perron integrable. The integral is now commonly referred to as the Denjoy- Perron integral.
~
10.2
Riemann Strikes Back
The reader who has followed these chapters on the LebesgueStieltjes integral is doubtless convinced of the vast superiority of the Lebesgue integral over the humble Riemann integral. Indeed, the Riemann integral is denigrated by many authors as merely a mathematical object of Ilhistorical" interest. At best, the integral is used as a pedagogical tool to introduce a Ilrigorous definition" of the integral in an elementary course in analysis. For example, in the 1930s Norbert Wiener [42] wrote in the introduction to his book on the Fourier integral,lIHowever, the Riemann integral is of relatively little importance in the theory of Fourier series and integrals, save as the classical cidinition applying to continuous and Istep-wise continuous' functions!' It is certainly true that the Lebesgue integral has won many resounding victories over the Riemann iritegral in fields such as Fourier analysis; it is also true that the Riemann integral suffers analytical deficiencies that are absent with the Lebesgue integral, but the victory is not complete, and good ideas (once accepted) are difficult to extinguish completely in mathematics. There is still an important realm where the Riemann integral reigns, viz.,
206
10. Epilogue
conditionally convergent integrals. The (improper) Riemann integral does not suffer the same restrictions as the Lebesgue integral regarding absolute convergence, and this annoying fact perhaps motivated Lebesgue and some of his contemporaries (e.g., Denjoyand Perron) to search for a more all-embracing definition of an integral. It is interesting to note that though many mathematicians have joined in the funeral chorus for the Riemann integral, the study of the Riemann integral has never really left the curriculum of mathematicians, owing to certain pedagogical advantages it has over the Lebesgue integral, and the need for conditionally convergent integrals in applications. Riemann's approach to integration was as revolutionary as Lebesgue's; Riemann effectively divorced the integral from differentiation and brought the geometrical properties of the integral into focus. In the late 1950s, the Riemann approach was vindicated by R. Henstock and J. Kurzweil. Working separately, these mathematicians extended the Riemann integral to include Lebesgue integrable functions. Even more impressive, their Riemann-based definition of an integral also captured the more general Denjoy-Perron integral. The resulting integral is now called the Henstock-Kurzweil integral, and it has since been shown to be equivalent to the DenjoyPerron integral,I i.e., f is Denjoy-Perron integrable if and only if it is Henstock-Kurzweil integrable. The key to their success in the extension was replacing Riemann's uniformly fine partitions of the integration interval with locally fine partitions. The use of locally fine partitions can be readily motivated by numerical examples. Consider, for example, the function f defined by f(x) = X-I sin(l/x) in some interval [E, 100], where Eis some small positive number. The graph of f shows that the function oscillates rapidly in the interval [E,l], but it then decreases steadily to 0 in the interval [1,100]. If we wished to approximate efficiently and accurately the integral lIt is tempting now to refer to the common integral as the Denjoy-Perron-
Henstock-Kurzweil integral, but one needs some patience in reading and typing such an appellation, and thus we eschew it. Often, if some proof requires the integral, then it is referred to as the Denjoy-Perron integral or the HenstockKurzweil integral, depending on which definition is to be used for the purposes of the proof.
10.3. Further Reading
207
of f over the interval [E, 100], we would be inclined to use a much finer partition of the interval [E, 1] than in the interval [I, 100]. Indeed, given some freedom, we would probably devise even more refined partitions for subintervals near the endpoint at x = E. The Henstock-Kurzweil approach allows the partition refinements in the integration interval to be nonuniform in the limit and thus Ilex_ tra fine" where needed. Although the Henstock-Kurzweil integral is equivalent to the Denjoy-Perron integral, the Riemann-type definition of the former makes it a more tractable concept and thus more amenable to applications and generalizations. In passing we note that these general integrals do not lead immediately to complete function spaces. The extension of the Lebesgue integral to include conditionally convergent integrals (the DenjoyPerron integral) or the generalization of the Riemann integral to include the Lebesgue integral (the Henstock-Kurzweil integral) opens floodgates that neither Riemann nor Lebesgue can close without help.
10.3
Further Reading
Our approach to the Lebesgue-Stieltjes integral in this book has been pragmatic and arguably ostrich-like. Major results such as the monotone convergence theorem were stated without proof, and we seldom entertained generalizations or abstractions. The purist might rightfully claim that this approach is demeaning to the subject; however, there is little harm in viewing some of the cornucopia before a substantial investment in further study is made (if desired). Moreover, the Lebesgue-Stieltjes integral is no longer the exclusive preserve of mathematicians: It is an important tool in any subject that uses integrals. At any rate, there are numerous excellent texts that cover the Lebesgue-Stieltjes integral in depth resplendent with proofs and abstractions. Some elementary texts pitched approximately at the same level as this book include those by Pitt [3D], [31], Priestley [32], and Weir [43], among others. These books develop the Lebesgue integral from the central concept of measure (which we did not emphasize). Pitt's books also cover applications of the Lebesgue-Stieltjes integral
208
10. Epilogue
to geometry, harmonic analysis, and probability theory. Priestley's book is a particularly lively account of the integra12 with many practical comments and examples. There are several advanced accounts of integration theory available. Classic specialist references include Halmos [14] and Thylor [38], but many advanced analysis books such as Royden [36] and Rudin [37] also cover integration in depth. In addition, since the Ll spaces loom large in functional analysis, most books on this subject devote some time to the Lebesgue integral. Riesz and Nagy [35], for example, devote nearly a quarter of their book to the Lebesgue and Stieltjes integrals; other authors such as Hutson and Pym [22] and Yosida [44] give concentrated but general accounts ofthe theory. The reader is encouraged to explore these references as curiosity or the need for more refined details dictates. Some of the reference cited above discuss the Ilpost Lebesgue" integrals of Denjoy et al. These integrals have formed a nucleus of specialist literature somewhat apart from the normal texts on integration. Lee [27] and Pfeffer [29] provide a quite accessible reference on the Henstock-Kurzweil integral and its equivalence to the Denjoy-Perron integral. Gordon [13] also provides a basic self-contained account of these integrals along with the Lebesgue integral. The reader is also directed to the article by Bartle [4] for an introductory account. Finally, the history of integration is interesting in its own right. Most accounts (such as ours) contain fleeting glimpses of the development of the integral. Hawkins [16] traces the history of the Riemann and Lebesgue integrals along with some early applications of the Lebesgue integral.
2Where else would you find yetis and nonmeasurable functions discussed in the same paragraph?
Appendix: Hints and Answers to Selected Exercises.
Exercises 1-1 2. For example, -J2+(1--J2) = 1 is rational, while -J2+-J2 = 2-J2 is irrational; (-J2)( -J2) = 2 is rational, while (-J2)(1 +-J2) = -J2+2 is irrational.
3. Hint: Can you be sure that b is an element of S*? 4. (a) Hint: Look at the proofthat the set ofall integers is countable.
(b) Hint: Ifthe set ofall irrational numbers were countable, what would that imply about the union of the set of all irrational numbers and the set of all rational numbers? Exercises 1-3 1.
Let lub and glb denote the least upper bound and the greatest lower bound respectively. (a) lub = 5, glb = 0, both are in the set.
°
(b) lub = 5, glb = 0, 5 is in the set but is not. (c) lub = +00, glb = -00, neither is in the set. (d) lub =
t, glb =
0, neither is in the set.
209
210
Appendix: Hints and Answers to Selected Exercises
(e) lub = ,.;2, glb = -,.;2, neither is in the set.
= 4, glb = 3, 4 is in the set but 3 is not. lub = +00, glb = 0, neither is in the set.
(f) lub (g)
3. Hint: Show that sup 82 is an upper bound of 8 1 , and that inf 82 is a lower bound of 81 . 4. Hint: For (a), show that c(sup 8) satisfies the definition of sup 8*, in other words, show that c(sup 8) is an upper bound of 8* and that c(sup 8) < B for any upper bound B of 8*. Follow a similar strategy to show that c(inf 8) = inf 8*, and also for part (b).
Exercises 2-1 1. Hint: Tb choose a sequence {an} such that
an t M, consider sep-
arately (i) M finite, (ii) M = 00. For (i), choose al E 8 such that M - 1 < al < M (using Theorem 1.3.2). Then choose a2 E 8 such that max{M - ~,al} < a2 < M, a3 E 8 such that max{M - ~,a2} < a3 < M, and so on. For (ii), choose al E 8 such that al > 1. (Why is this always possible?) Then choose a2 E 8 such that a2 > max{2, all, a3 E 8 such that a3 > max{3, a2}, and so on. Use a similar approach to choose a sequence bn such that b n ,J, m. Exercises 2-4 3. Hint: GivenE > O,choosensuchthat1/(n+1) < E.Itfollowsfrom the definition off that 0 < x < lin =} 0 < f(x) < 1/(n + 1) < E.
Exercises 2-6 1. Hint: Prove that If+(x) - g+ (x) I < If(x) - g(x)1 for all x E I by considering the four cases f(x) > 0 and g(x) > 0, f(x) > 0 and g(x) < 0, f(x) < 0 and g(x) > 0, f(x) < 0 and g(x) < O.
2. Similar to 1.
3. Hint: For each x E I, If (x) I = If(x) - g(x) + g(x) I < If(x) - g(x) I+ Ig(x) I, therefore If(x)-g(x)1 > If(x)I-lg(x)l. Interchangingf and g gives If(x) - g(x) I > Ig(x) I - If(x)l· Since IIf (x) I - Ig(x) II must equal either If (x) I - Ig(x) I or Ig(x) I - If(x)I, the result follows.
Appendix: Hints and Answers to Selected Exercises
211
Exercises 2·7 1. Hint: Suppose f has bounded variation on I. Choose a point a E I,
and let x be any point in I. Denote by Ix the closed interval with endpoints a and x. Use the fact that {Ix} is a partial subdivision of I, together with the definition of bounded variation, to obtain the required result. 2. Hint: For fg, use n
n
If(bj)g(bj) - f(tlj)g(tlj) I = L
L j==l
•
If(bj)g(bj) - f(bj)g(aj)
j==l
+f(bj)g(tlj) - f(aj)g(aj) I n
If(bj)llg(bj) - g(tlj) I
n
+L
Ig(aj)llf(bj) - f(aj) I,
j==l
and then use the result of Exercise 1 to obtain bounds for If(bj) I and Ig(tlj) I that are independent ofj. 3. Hint: In view of Exercise I, for the first part you need prove only that if sup{f(x) : X E I} and inf{f(x) : X E I} are both finite (and f is monotone on 1), then f has bounded variation on I. Using the notation for partial subdivisions introduced earlier, you can assume without loss of generality that al < b l < a2 < b2 < ... < an < bn.
If f is monotone increasing on I, then f(al) < f(b l ) < f(a2) < f(b 2) < ...
and therefore for any partial subdivision of I we have n
L j==l
n
If(bj) - f(tlj) I = L(f(bj) - f(tlj)) j==l n
n
< L(f(bj) - f(aj))
f(bj- l ))
j==2
j==l
= f(b n) -
+ L(f(aj) -
f(al)
< sup{f(x) : X
E
I} - inf{f(x) : x
E
I}.
212
Appendix: Hints and Answers to Selected Exercises
A similar argument shows that iff is monotone decreasing, then n
L If(bj) - f(aj) I < f(al) - f(b
n)
j=l
< sup{f(x) : X
E
I} - inf{f(x) : X
E
I}.
The first part then follows easily, and the second part is a straightforward application of the results already proved.
4. Hint: Use the construction descnbed in the proof of Theorem 2.7.2.
5. (a) Hint: Use the fact that Ix sin(l/x)I < Ixl for all x ;/=
o.
6. (b) Hint: If an interval I has finite endpoints a, b, then for any Xl,X2 E I we have IXf - x~1 = IXI + x211xI - x21 < (2 max{lal, Ibll)lxI - x21.
(c) Hint: Letf(x) = x 2 , and assume thatf is absolutely continuous on I = (-00,00). Then by definition (choosing E = 1 in the definition), there exists a 8 > 0 such that Vs(f, 1) < 1 for all partial subdivisions S of I for which the sum of the lengths of all the constituent intervals is less than 8. Considering in particular partial subdivisions consisting of a single interval [n, n + 8/2] (where n is a positive integer) leads to the desired contradiction.
Exercises 3-1 1. Suppose that x E [a, b]. Then x Elk for some 1k E P and x Elk' for some 1k' E pl. Now, pI is a refinement of P, so that 1k' elk. The result follows from the general inequalities inf{f{x) : X E 1k'} > inf{f(x) : X Elk} and sup{f(x) : X Elk'} < sup{f(x) : X Elk}. 2. Since the partition Q = p U pI is a refinement of both P and pI, Lemma 3.1.2 implies that Sp(f) > SQ(f) and§.o.(f) > §,p,(f). Since SQ (f) > SQ (f) for any partition Q, we have that
Sp(f) > SQ(f) > and the lemma thus follows.
~(f)
2: Sp,(f),
Appendix: Hints and Answers to Selected Exercises
213
Exercises 4-1 1. (b) Jla((O, 1)) = l-e-l, Jla([O,I]) = 3 -e- 1 , Jla((-I,I))4 - e-l, Jla([O,OD = 2, Jla(( -00,1)) = 00, Jla((O,oo)) = I, Jla([O,oo)) = 3. 2. (b) Jla([-I,2)) = 4, Jla((I,oo)) = 2, Jla((-oo,4)) Jla((O,2]) = 5, Jla((~, ~)) = 3, Jla([I, 3]) = 5, Jla((1, 3))
= = 2.
6,
Exercises 4-2 ifx < A, if x > A.
1. a(x) = {
0.,
2. a(x) =
0, Un,
I,
{ I,
if x < AI, if Ai < x < Ai+l (i = 1,2, ... , n - I), if x > An.
Exercises 4-3 1. (a) S U T = [1,8). S n T = (2,3) U (4,5] U (6, 7]. S - T = [1,2] U (5,6] U (7,8).
(b) S U T
= [1,4] U [5,8).
SnT
= (2,3) U [6, 7].
S- T
= [5,6).
(c) S U T = (1,4] U [5, 7). S n T = [2,2] U (5,6). S - T = (1,2) U [5,5]. Exercises 4-4 1. Aa(O)
= O.
2. Aa(O) = O. 3. Aa(O) = 3. 4. Aa(O) = ~.
5. 0 is not a-summable. Exercises 4-5
2. Hint: The difficulty with this one is that it is too easy! Since La*(lfl) = L a *(/) = 0, you can just take On to be the zero function on [0,1], for each n = 1,2, ....
214
Appendix: Hints and Answers to Selected Exercises
Exercise 4-6 (a) Let n be the integer part of c, i.e., the largest integer not exceeding c. Then ifn
1,
(c Jo
fdx
1
<2(n+l)'
where n is the integer part of c. (c) Hint: Show that if n is the integer part of c and c > I, then c
1,o
1
1
1
c- n
2
3
n
n+l
IfI dx = 1 + - + - + ... + - +
.
Exercises 5-1 2. Hint: Tb prove that max{f(x),g(x)} = f(x) + (g - n+(x) for all x E I, consider separately the cases f(x) > g(x) and f(x) < g(x); similarly, for min{f(x), g(x)}. 3. Hint: Use Theorem 5.1.5(ii) and Theorem 4.5.6. 4. Hint: Use Theorem 5.1.4 and Theorem 5.1.3. Exercises 5-2 1. Hint: Use the same approach as was used in Example 4-5-2. 3. Hint: Use the fact that g = 1-f, where f is the function defined in Section 3.2, or use the fact thatg = 1 a.e. to show that ~O,l] g dx = 1.
4. Hint: Define g* on I by g*(x)
=
{g(X), if g(x) < f(x), f(x), if g(x) > f(x), '
so that g = g* a.e. and g* < f on I. 5. Hint: Show that XSUT < Xs
Exercises 5-3 2. (a) f = o.
+ XT on JR.
Appendix: Hints and Answers to Selected Exercises
215
Exercises 5-4 1. Hint: To obtain a sequence 81 , 821 ... ofa-summabIe step functions on lR such that lim n -+ oo 8n = 1 on lR, define 8 ( x) n
2. (b) J.La([O,2))
= -I,
=
{I,0,
J.La([O, 1])
if -n <. x < n , otherwIse.
=
0, J.La([1, 1])
= -1,
J.La((1, 2)) =
-l.
(c) 3.
Exercises 6-1 1. (i) 3(1 - e- l ), (ii) 4 - 3e- l , (iii) 4 - 3e-1, (iv) 4 - 3e- l (v) 8 + e- 2 , (vi) 3. 2. (b) (i) 61, (ii) e + e2 + e3 + e4 + 2e s , (iii) I, (iv) 2.
+ e- 2 ,
3. (a) ~(A + B).
(b) 0, lin L~=l Ai· if x < AI, (c) a(x) = I:;=lPj, if Ai < x < Ai+l (i { I, if x > An. Mean = I:?=lPiAi. 0,
= 1,2, ... , n -
I),
Exercises 6-3 1. limt-+o+ erf(t)lt
5.
t sin(t
3
) -
Exercise 7-4
= 21 -ft.
t sin(t
limt-+oo terfc(t)
= O.
2 ).
Each repeated integral has the value 9 sin 1 +4 sin 2.
Exercises 8-1 1. (a) Properties (i), (iii), and (iv) are straightforward to establish. The real problem is showing that property (ii) is satisfied. Evidently, if f = 0, then IIfllR = o. Suppose now that IIfllR = 0 but that f =j:. o. Since f is not identically zero on the interval [a, b], there is some number c E [a, b] such that If(c)I > O. Since f is continuous on [a, b] this means there is some interval [a ,,8] C [a, b] containing c such that If(x) I > 0 for all
216
Appendix: Hints and Answers to Selected Exercises
I:
It
x E [a, ,8]. Now IIfllR = Ifex)1 ax > If(x)1 ax > 0, which contradicts the hypothesis that IIfllR = O. Therefore, II . IIR satisfies property (ii).
(b) Hint: Let c be any number in the interval [a, b] and let f be the function defined by f(x) = 0 if x i= c and f( c) = 1. What is the norm of this function? 3. Hint: Properties (i) and (ii) follow from the inequalities IIfllI,oo > IIflloo and IIfll1,l > IIfIIR. Property (iv) follows from the inequality If(x) + g(x) I < sUPxE[a,bj If(x) I + sUPxE[a,b] Ig(x)l. 4. Hint: ISn+1 - Snl < IlIOn, and ifn > m, then ISn - Sml = I(SnSn-1) + (Sn-1 - Sn-2) + ... + (Sm+1 - Sm)l. sUPxE[a,bj
5. Since IIfllb < ,8l1flla we can choose y = 1/,8. Similarly, we can choose 8 = Ila.
Exercises 8-2 1. Suppose that an ~ a as n ~ 00. Then, for any E > 0 there is an N such that lIam - all < E/2 whenever m > N. Now, lIam - all = lI(am - an) + (an - a) II > lIam - an II - lIan - aiL and thus
lIam - an II - lIan - a II < lIam
E
-
a II < 2'
so that ifn > N, then
lIam - an II -
E
E
2 < 2'
Thus, for any E > 0 there is an N such that lIam - anll < whenever n, m > N.
E
2. Let n be any positive integer. Note that SUPXE[-l,lj Ifn(x) fn+1(X)1 is achieved at x = I/2 n+1, where fn+1CX) = 0 and fn(x) = 1 - 2m- Cm+1) = ~. Thus IIfn(x) - fn+l(X) II 00 = ~ for all n, and {in} cannot be a Cauchy sequence.
Exercises 8-3 1. (a) Let y be any element in Y and choose any E > O. Since W is dense in Y, there is a il; E W such that 1Iil; - YII < E/2. Similarly, since Z is dense in W, there is a z E Z such that
Appendix: Hints and Answers to Selected Exercises
217
liz - wll < E/2. Consequently, for any Y E Y and any E > 0 there is a z E Z such that liz - YII < E, i.e., Z is dense in Y. (b) Use part (a) and the definition of completion. 2. Hint: Use the fact that the set of rational numbers is dense in the
set of real numbers.
Exercises 8-5 1. (a) Hint: To show the triangle inequality rllf+gll p < rllfllp+rllgllp apply the Minkowski inequality with F = rl/Pf and G = rllpg.
(b) See the discussion after equation (9.11). 3. Since f E £2[0, 1], we have that f E £1[0, 1]; since k is bounded, there is an M < 00 such that Ik(x, Y)I < M for all (x, y) E [0, 1] X [0, 1]. Hence,
r
I(KD(x) I <
Ik(x,
J[O/I]
~)llf(~)~ < Mllflll,
and consequently
IIKfll~ < 4. Part(iii): Suppose that f
1,
r M2I1flli~=M2I1flli.
J[O/I]
E
£1 [a, b] and that g E VXl[a, b]. Then
If(x)g(x) I dx. < IIglloo
[a/b)
1,
If(x)1 dx.
[a/b)
= IIglloollflh¥ <
00,
and therefore fg E £1 [a, b]. Part(iv): Hint: First establish the inequality
1,
If(x)IP dx. <
~~
IIfll~(b -
a).
.
5. Hint: Note that
{
1,
[a,b]
I/P
If(x)IP dx.
}
{
<
1,
[a,b]
IIfll~ dx.
} lip
.
218
Appendix: Hints and Answers to Selected Exercises
Exercises 9-1 1. Let h
= g/lIgll. Then
o<
IIf - if, h)hll 2 = if - if, h)h,f - if, h)h) 2 = IIfll - if, h)(h,f) - if, h) if, h) + if, h) if, h) 2 = IIfll - Iif, h) 12 ;
thus, Iif, h)1 < IlflL and hence Iif, g)-I < IIfllllgll forg then the inequality follows immediately.)
-:f: o. (Ifg = 0,
3. Hint: Use the parallelogram equality, or verify by direct calculation.
Exercises 9-2 2. (a) For example,
(b) 1b normalize Po(x) , note that IIPo 11 2 let 4>o(x)
= Jipo(x) = Ji.
If
= f~l 12 dx = 2; thus,
Similarly, IIPl ll 2
=
~, II P211 2
=
~,
so let epI (x) = PI (x), ep2(X) = !fP2(X). The Fourier coefficients are given by (e!', epn) for n = 0, 1,2. For example,
a2
=
(e!', ep2)
= f~l e!'~(3X2 -1)dx = !f(e - ~).
4. Since M is a total orthonormal set, Parseval's formula is valid. Using the notation ak = if, epk), bk = (g, epk) we thus have that IIfll 2 = L~I lakl 2 and IIgll 2 = L~I Ibk l2 . In addition, we also have that 00
IIf + gll2 =
:E
2 lak + bkl ,
k=l
IIf + igll
2
00
=
:E lak + ibkl2. k=l
Now, Ilf + gll2 = IIfll
2
+ 2Re if, g) + IIgll 2,
Appendix: Hints and Answers to Selected Exercises
219
so that IIfll
2
+ 2Re if, g) + IIgll
2
00
= L
lak + bkl2
k=l 00
=L
lakl
2
k=l
=
IIfll
2
00
00
k=l
k=l
+ 2 LRe(akbk) + L 1~12 00
+2L
Re (akbk)
+ IIg1l 2;
k=l consequently, Re if, g) = 'L:l Re (akbk). A similar argument applied to IIf + igll indicates that 1m if, g) = 'L:l 1m (akbk), and hence the result follows.
Exercises 9-3 1. (a) PMx
k
~ sin(2x) + sin(3x) 'L~oo Cn e2mrix , where
= 2(sin(x) -
2. (a) PMg(X)
=
_ { -;;;i,
n C -
(b) IIgll
2
0,
i sin(4x) + ....
if n is odd,
if n is even.
= f~~~2g2(X)dx = 1 = L~oo Icnl2 = 2'L:1 ((2k~1)Jr)2.
3. Since f' is continuous on the interval [-Jr, Jr], there exist numbers M and M' such that If (x) I < M and 1f'(x)1 < M' for all x E [-Jr, Jr]. Now, Jr f(x)cos(Anx)dx j -Jr
= 2- [sin(Anx)]~Jr An
_jJr 2-!,(X)Sin(An X)dx, -Jr An
and therefore
i:rCX) cosC1..nx)dx <
I:nl (lI/CX) sinC1..nX)]~nl +
< _1_ (2M - IAnl
i:
I['Cx) sinC1..nX)I
dx)
+ 2JrM/) .
Since IAnl -* 00 as n ~ 00, f~Jrf(x)cos(Anx)dx-* 0 as n ~ 00. The limit for the sine integral can be established using the same arguments.
220
Appendix: Hints and Answers to Selected Exercises
Exercises 9-4 1. Suppose A is an eigenvalue for the operator £. Then £y = -Apy, and (£y, y) = (-Apy, y). If £ is self-adjoint, then (£y, y) = W, £y);
thus, (£y,y)
= -A{Py,y) = W, £y) = W, -Apy) = -AWY,y)·
Since p is a real-valued function, we must have that A = A, i.e, A is real.
((2n~1)Jr)2 for n = 0,1,2 ...; corresponding eigenfunctions are epn = sin ((2n~1)Jrx).
2. The eigenvalues are An
=
References
[1] Adams, R.A., Sobolev Spaces, Academic Press, 1975.
[2] Ahlfors, L., Complex Analysis, 2nd edition, McGraw-Hill Book Co., 1966. [3] Arfken, G., Mathematical Methods for Physicists, 2nd edition, Academic Press, 1970. [4] Bartle, R.G., I~ return to the Riemann integral", Amer. Math. Monthly, 103 (1996) pp. 625-632 [5] Birkhoff, G. and Rota, G., Ordinary Differential Equations, 4th edition, John Wiley and Sons, 1989. [6] Bromwich, T.A., An Introduction to the Theory ofInfinite Series, Macmillan and Co., 1926. [7] Carleson, L. IIConvergence and growth of partial sums of Fourier Series" Acta Math., 116, (1966) pp. 135-157 [8] Churchill, R.V:, Fourier Series and Boundary Value Problems, 2nd edition, McGraw-Hill Book Co., 1963. [9] Coddington, E.A. and Levinson, N., Theory of Ordinary Differential Equations, McGraw-Hill Book Co., 1955. [10] Conway, J.B., Functions of One Complex Variable I, 2nd edition, Springer-Verlag, 1978.
221
222
References
[11] Courant, R. and Hilbert, D., Methods of Mathematical Physics, volume I, John Wiley and Sons, 1953. [12] Dunford, N. and Schwartz, J.T, Linear Operators, parts I, II, III, John Wiley and Sons, 1971. [13] Gordon, R.A., The Integrals of Lebesgue, Denjoy, Perron, and Henstock, American Math Soc., 1994 [14] Halmos, :P.R., Measure Theory, Springer-Verlag, 1974. [15] Hardy, G.H. and Rogosinski, WW., Fourier Series, 3rd edition, Cambridge University Press, 1956 [16] Hawkins, T., Lebesgue's Theory of Integration, Its Origins and Development, The University of Wisconsin Press, 1970. [17] Heider, L.J. and Simpson, J.E., Theoretical Analysis, W.E. Saunders Co., 1967. [18] Hewitt, E. and Stromberg, K., Real and Abstract Analysis, Springer-Verlag, 1969. [19] Higgins, J.R., Completeness and Basis Properties of Sets of Special Functions, Cambridge University Press, 1977. [20] Hille, E., Analytic Function Theory, volume II, Ginn and Co., 1959. [21] Hoffman, K., Banach Spaces of Analytic Functions, Prentice-Hall, 1962. [22] Hutson, V. and Pym, J.S., Applications of Functional Analysis and Operator Theory, Academic Press, 1980. [23] Ince, E.L., Ordinary Differential Equations, Longmans, Green and Co., 1927. [24] Korner, T.W., Fourier Analysis, Cambridge University Press, 1988. [25] Kreyszig, E., Introductory Functional Analysis with Applications, John Wiley and Sons, 1978. [26] Kreyszig, E., Advanced Engineering Mathematics, 4th edition, John Wiley and Sons, 1979. [27] Lee, P., Lanzhou Lectures on Henstock Integration, World Scientific, 1989. [28] McLachlan, N.W., Theory and Application ofMathieu Functions, Oxford University Press, 1947. ~
References
223
[29] Pfeffer, W.F., the Riemann Approach to Integration: Local Geometric Theory, Cambridge University Press, 1993. [30] Pitt, H.R., Integration, Measure and Probability, Oliver and Boyd, 1963. [31] Pitt, H.R., Measure and Integration for Use, Oxford University Press, 1985. [32] Priestley, H.A., Introduction to Integration, Oxford University Press, 1997. [33] Pryce, J.D. Basic Methods of Linear Functional Analysis, Hutchinson and Co., 1973. [34] Richtmyer, R.D. Principles of Advanced Mathematical Physics, volume I, Springer-Verlag, 1978. [35] Riesz, F. and Nagy, B., Functional Analysis, Frederick Ungar Publishing Co., 1955. [36] Royden, H.L., Real Analysis, Macmillan and Co., 1963. [37] Rudin, W., Real and Complex Analysis, 2nd edition, McGraw-Hill Book Co., 1974. [38] Thylor, A.E., General Theory of Functions and Integration, Blaisdell Publishing Co., 1965. [39] Titchmarsh, E.C., The Theory of Functions, 2nd edition, Oxford University Press, 1939. [40] Titchmarsh, E.C., Eigenfunction Expansions, Part I, 2nd edition, Oxford University Press, 1962. [41] Whittaker, E.T. and Watson, G.N., A Course ofModem Analysis, 4th edition, Cambridge University Press, 1952. [42] Wiener, N., The Fourier Integral and Certain ofIts Applications, Cambridge University Press, 1933. [43] Weir, A.J., Lebesgue Integration and Measure, Cambridge University Press, 1973. [44] Yosida, K. Functional Analysis, 6th edition, Springer-Verlag, 1980. [45] Zygmund, A., Trigonometric Series, volumes I and II, Cambridge University Press, 1959.
Index
almost everywhere (a.e.), 76 anti-derivative, 110 Appolonius identity, 172 Banach space, 133 Lebesgue, 144 separable, 150 Sobolev, 162 HP spaces, 157 Bessel's inequality, 177 bound essential upper, 148 greatestlovver, 8 least upper, 8 Cauchy integral formula, 158 Cesaro mean, 187 characteristic function, 76, 152 conjugate exponents, 145
continuity one-sided, 19 continuous absolutely, 36 convergence in the norm, 129 of a double series, 15 of a sequence, 11 of a series, 13 of an improper integral, 44 pointwise, 12 uniform, 133 Denjoy integral, 204 dense, 2 density of distnbution, 103 divergence of a series, 13 proper, 15 dominated convergence theorem, 80
225
226
Index
-------------------------==~
dual space, 147 eigenfunctions, 190 eigenvalues, 190 error function, 102 complenaenta~,
103 extended real number system, 6 Fatou's Lenama, 80 Fourier coefficients, 176 complex series, 182 series, 176 series,classical, 181 sine series, 193 Fredholm integral operator, 150 Fubini's theorem, 118 function a-measurable, 82 characteristic, 76, 152 error, 102 Lipschitz, 38 major, 204 monotone, 20 negative part, 28 null, 75 positive part, 28 simple, 11 5 step, 24 strictly increasing, 97 function spaces, 126 functional bounded, 146 linear, 145 norm, 146 fundamental theorem of calculus, 102 Holder's inequality, 143 HP spaces, 158
Haar functions, 199 Hardy spaces, 157 Henstock-Kurzweil integral, 206 Hermite differential equation, 197 polynomials, 197 Hilbert adjoint operator, 190 dimension, 179 Hilbert space, 168 improper integral, 44 absolutely convergent, 46 conditionally convergent, 46 convergence of, 44 infimum, 8 inner product, 165 inner product space, 166 real, 166 integral Denjoy, 204 Henstock-Kurzweil, 206 improper, 44 indefinite, 110 Newton, 203 Perron, 205 Riemann, 39 Riemann-Darboux,42 interval closed, 7 open, 7 irrational numbers, 2 jump discontinuity, 19 Laguerre differential equation, 198 polynomials, 198 Lebesgue integral definition, 66
Index - - - - - - - - - - - - - - -227
generalization of Riemann, 69 Lebesgue-Stieltjes integral change of variable, 97 definition, 66 differentiation under the integral, 105 double, 115 first mean value theorem, 75 integration by parts, 100 linearity of, 74 • repeated, 116 Legendre differential equation, 195 polYnomials, 179, 196 Leibniz's rule, 108 limit as x -+ ±oo, 18 for a function, 16 of a sequence, 11 one-sided, 16 Lipschitz function, 38 Mathieu equation, 194 functions, 194 maximum modulus principle, 155 measure of a rectangle, 113 of an interval, 50 probability, 52 Minkowski's inequality, 144 HP spaces, 158 monotone convergence theorem, 79 Newton integral, 203 norm definition, 127 equivalent, 129, 130
Euclidean, 127 induced by inner product, 167 of a partition, 43 normed vector space closed subset of, 134 complete, 133 complete subset of, 134 completion of, 137 definition, 127 dual space, 147 isometric, 135 nul function, 75 set, 76 operator adjoint, 190 definition, 135 Fredholm integral, 150 isometry, 135 resolvant, 190 self-adjoint, 190 orthogonal complement, 174 definition, 172 set, 175 orthonormal basis, 177 set, 176 ostrich, 207 parallelogram equality, 168 Parseval's formula, 178 partial subdivision, 30 partition locally fine, 206 norm of, 43 of an interval, 41 refinement of, 41 Perron integral, 205
228 probability density, 103 discrete distnbution, 54 distribution function, 52 measure, 52 uniform distribution, 53 Pythagoras's theorem, 173 Rademacher functions, 180 random variable, 52 mean, 96 rational numbers, 1 rectangle, 113 Riemann -Lebesgue theorem, 184 localization theorem, 186 Mapping Theorem, 156 theorem on derangement of series, 15 Riemann integral definition, 39 Schwarz's inequality, 167 seminorm, 139 separable, 150 sequence admissible, 60 Cauchy, 129 convergence, 11 monotone, 12 series term by term integration, 80 set a-finite, 55 a-measurable, 82 closure of, 134 complete, 175 countable, 3 dense, 137 nul, 76
-=In=d::.:ex=
orthogonal, 175 orthonormal, 176 simple, 55, 114 total, 175 signum function, 180 simple function, 115 Sobolev space, 162 span, 126 spectrum, 190 step function, 24 a-summable, 56 Sturm-Liouville problem regular, 189 singular, 195 sum direct, 174 vector, 174 support of a Haar function, 201 of a step function, 25 of simple function, 115 supremum, 8 variation functions ofbounded, 32 total, 30 vector space complex, 124 definition, 124 finite-dimensional, 126 infinite-dimensional, 126 normed, 127 subspace, 126 vector spaces seminormed, 139 Walsh functions, 202 Weierstrass, 138 yeti, 208
-----------------,----..,..-----------------'----'--'"
Undergraduate Texts in Mathematics (continued/rom page ii)
HiltonfHoltonlPedersen: Mathematical Reflections: In a Room with Many Mirrors. Iooss/Joseph: Elementary Stability and Bifurcation Theory. Second edition. Isaac: The Pleasures of Probability. Readings in Mathematics. James: Topological and Uniform Spaces. Janich: Linear Algebra. Janich: Topology. Kemeny/Snell: Finite Markov Chains. Kinsey: Topology of Surfaces. Klambauer: Aspects of Calculus. Lang: A First Course in Calculus. Fifth edition. Lang: Calculus of Several Variables. Third edition. Lang: Introduction to Linear Algebra. Second edition. Lang: Linear Algebra. Third edition. Lang: Undergraduate Algebra. Second edition. Lang: Undergraduate Analysis. LaxIBursteinlLax: Calculus with Applications and Computing. Volume 1. LeCuyer: College Mathematics with
APL. LidllPilz: Applied Abstract Algebra. Second edition. Logan: Applied Partial Differential Equations. Macki-Strauss: Introduction to Optimal Control Theory. Malitz: Introduction to Mathematical Logic. MarsdenlWeinstein: Calculus I, II, III. Second edition. Martin: The Foundations of Geometry and the Non-Euclidean Plane. Martin: Geometric Constructions. Martin: Transformation Geometry: An Introduction to Symmetry. MillmanlParker: Geometry: A Metric Approach with Models. Second edition. Moschovakis: Notes on Set Theory.
Owen: A First Course in the Mathematical Foundations of Thermodynamics. Palka: An Introduction to Complex Function Theory. Pedrick: A First Course in Analysis. Peressini/SullivanlUhl: The Mathematics of Nonlinear Programming. PrenowitzJJantosciak: Join Geometries. Priestley: Calculus: A Liberal Art. Second edition. ProtterlMorrey: A First Course in Real Analysis. Second edition. ProtterlMorrey: Intermediate Calculus. Second edition. Roman: An Introduction to Coding and Information Theory. Ross: Elementary Analysis: The Theory of Calculus. Samuel: Projective Geometry. Readings in Mathematics. ScharlauJOpolka: From Fermat to Minkowski. Schiff: The Laplace Transform: Theory and Applications. Sethuraman: Rings, Fields, and Vector Spaces: An Approach to Geometric Constructability. Sigler: Algebra. Silverman/Tate: Rational Points on Elliptic Curves. Simmonds: A Brief on Tensor Analysis. Second edition. Singer: Geometry: Plane and Fancy. Singerffhorpe: Lecture Notes on Elementary Topology and Geometry. Smith: Linear Algebra. Third edition. Smith: Primer of Modern Analysis. Second edition. Stanton/White: Constructive Combinatorics. Stillwell: Elements of Algebra: Geometry, Numbers, Equations. Stillwell: Mathematics and Its History. Stillwell: Numbers and Geometry. Readings in Mathematics. Strayer: Linear Programming and Its Applications.
Undergraduate Texts in Mathematics Thorpe: Elementary Topics in Differential
Valenza: Linear Algebra: An Introduction
Geometry. Toth: Glimpses of Algebra and Geometry. Readings in Mathematics. Troutman: Variational Calculus and Optimal Control. Second edition.
Whyburn/Duda: Dynamic Topology. Wilson: Much Ado About Calculus.
to Abstract Mathematics.