0 and j (af) df-L = a ff df-L. 3. f v g and f 1\ g are upperfunctions. Proof. Choose two generating sequences {4>n} and { 1/111 } for f and g, respec 1.
tively. ( 1 ) Clearly, {f/1, + 1/111} is a sequence of step functions, and f/>11 a.e. holds. The result now follows by observing that
+ 1/111 t f + g
(2) Straightforward. (3) Note that both {f/>11 v 1/111} and { f/1, 1\ 1/J, } are sequences of step functions. Now, f/>11 1\ 1/ln t f 1\ g a.e. and lim J f/>11 1\ 1/111 df-L < lim J f/>11 df-L < oo, show that f 1\ g is an upper function. To see that f v g is an upper function, observe first that f/>11 v 1/111 t f v g a.e.
holds, and then use the identity
f/>u V 1/J, to obtain Jf/>11 V
=
f/>11 + 1/111 - f/>u
1/111 df-L = Jf/>11 d f-L + J 1/111 d f-L
-
1\
1/111
Jf/>11 1\ 1/111 df-L. This implies
This finishes the proof of the theorem. The next theorem states that the integral is a monotone function on U.
•
Chapter 4: THE LEBESGUE INTEGRAL
164
Theorem 21.5. If f and
are upper functions such that f f g dJ.L holds. In particular, if f E U satisfies f g
J f dJ.L > f f dJ.L > 0.
2:: g >
a.e., then 0 a.e., then
Proof. Let {>,} and { l/f,} be generating sequences for f and g, respective]y.
Then >11 A 1/1, t g a. e. holds, and so { >, A 1/1n } is also a generating sequence for g. By Theorem 17.3, we have J >, d J.L > J
and the proof is finished.
•
It should be noted that if f is an upper function such that f 2:. 0 a.e., then there exists a sequence of step functions { 1/1n } satisfying l/f, > 0 a.e. for each n and + l/f, t f a.e. To see this, notice that if
Theorem 21.6. Let f : X
� IR. be a function. If there exists a sequence {f,,}
of upper functions such that fn t f a.e. and lim J fn d J.L < oo, then f is an upperfunction and J f dJ.L = lim f f,, dJ.L. Proof. For each i choose a sequence { i, it follows that J /; dJ.L = lim,_oc J
J fn dJ.L = J 1/ln dJ.L J f dJ.L,
and the proof is complete.
lim 11-+-0C
=
II
The integral satisfies an important convergence property for decreasing se quences. It is our familiar order continuity property of the integral.
Theorem 21.7. If {fn } is a sequence of upperfunctions such that f, .J, 0 a.e.,
then lim J f,, dJ.L = 0 holds.
For each n choose a step function , is an
Proof. Let E
a.e. and
j(f,,
>
0.
Section 21: UPPER FUNCTIONS
165
=
upper function). Let 1/111 /\�'= 1 k. Now, the almost everywhere inequalities
0 < ft,
- = 1/111
ft, -
II
=
=
II
II
II
V([t, -
i=l
<
imply
Thus,
for all n
> k, which shows that f ft, dJ.L ,{, 0
•
holds.
Finally, we mention that U is not in general a vector space since it fails to be closed under multiplication by negative real numbers. An example of this type is presented in Exercise 2 of this section.
EXERCISES 1. Let L be the collection of all step functions ¢ such that there exist a finite number of sets A 1 , , A, in S all of finite measure and real numbers a 1 , , a11 such that 2:7=• a; XA;. Show that L is a function space. Is L an algebra of functions? ¢ [HINT: Use Exercise 14 of Section 12.] 2. Consider the function f : 1R ---+ 1R defined by f(x) = 0 if x � (0, l], and f(x) .,fo if x E (11!1 , �] for some n. Show that f is an upper function and that - f is not an upper function. [HINT: A step function is necessarily bounded.] 3. Compute J f d).. for the upper function f of the preceding exercise. 4. Verify that every continuous function f : [a, b] ---+ IR is an upper function-with respect to the Lebesgue measure on [a, b]. 5. Let A be a measurable set, and let f be an upper function. If XA � f a e , then show that ,u*(A) < oo. 6. Let f be an upper function, and let A be a measurable set of finite measure such that a � f(x) � b holds for each x E A. Then show that
=
• • •
• • •
=
..
Chapter 4: THE LEBESGUE INTEGRAL
166
a. f XA is an upper function, and b. 7.
a 11 *(A) � j f XA d11 � hJ-L*(A).
Let (X, S, J-L) be a finite measure space. and let f be a positive measurable function. Show that f is an upper function if and only if there exists a real number M such that J ¢ dJ-L � M holds for every step function ¢ with ¢ � f a.e. Also, show that if this is the case, then
If d
J.L
8.
=
sup
{I
¢ dJ.L: ¢ is a step function with ¢ �
Show that every monotone function to the Lebesgue measure on [a, b] .
f
:
[a. b]
-+
f a.e.
J
.
1R is an upper function-with respect
22. INTEGRABLE FUNCTIONS It was
observed before that the collection U of all upper functions is not a vector
space. However, if we consider the collection of aii functions that can be written as
an almost
everywhere difference of two upper functions, then this set is a function space. The members of this collection are the Lebesgue integrable functions. The details will be explained in the following.
Definition 22.1. A function f : X � IR is called Lebesgue integrable (or simply integrable) ifthere exist two upperfunctions u and v such that f = u - v a.e. holds. The Lebesgue integral (or simply the integral) of f is defined by
It should be noted that the value of the integral is independent of the representa tion of f as a difference of two upper functions. Indeed, if f = u - v = u I - VI a.e. with u, LII , v, and vI all upper functions, then u + vI = u I + v a.e. hoi ds, and by Theorem 2 1 .4( 1) we have f u df.L + f VI dJL = f u 1 dJL + J v dJ-L. Therefore,
j U dJ.L - j V df.L
=
j UI dJ-L - j VI dJ.L.
An integrable function is necessarily measurable, and every upper function is Lebesgue integrable. Also, it is readily seen that if a function f . is Lebesgue integrable and g is another function such that f = g a.e., then g is also Lebesgue integrable and J g dJ-L = J f df.L holds.
Historical Note: The above introduction. of the Lebesgue integral is a modification of a method due to P. J. Daniell [5]. Daniell's 2 general approach to integration starts with a function space L on some nonempty set X , together with an "integral" : L -+ 1R is said to be an ime�:ral if
I
I
on L . The function
2P. J. Daniell (1 889-1946), a Briiish mathematician. He worked in functional analysis and the theory of integration.
167
Section 22: INTEGRABLE FUNCTIONS
1. 2. 3.
l (a¢ + {31/1) = a/ (¢) + {31(1/1) for all fX, f3 E IR and ¢, 1/1 E L, /(¢) > 0 whenever ¢ :;:: 0, and whenever {¢11} C L satisfies ¢11 (x) .l, 0 for each x E X, then I (¢11 ) .l,
0.
A function u : X IR is called an upper function if there exists a sequence { c/J11} C L with ¢11(x) t u(x) for all x E X and lim /(¢11) < oo. As in the proof of Theorem 17 .5, we can show that lim /(c/J11) is independent of the "generating" sequence {¢11 }. The real number I (u) = lim I (¢11) is the integral of u. Finally, f : X -+ IR is said to be integrable i f there exist two upper functions u and v with f = u v. The integral of f is then defined by �
-
l(f) = l(u) - l(v).
Here our approach to the Lebesgue integral can be considered as a "measure theoretical
Daniell method." The set of integrable functions has all the expected nice properties.
Theorem 22.2. The collection of all Lebesgue integrable functions is a func
tion space.
Proof. Let f and g be two integrable functions with representations f = u - v a. e. and
g
=
u 1 - v1
a. e. Then the almost everywhere identities
J+g ct.f af
j+
=
(u + u 1 ) - (v + vJ),
= Cf.ll - Cf.V if =
=
ct. > 0,
[(-a)v] - [(-a)u] if + (u - v) = u v v - v
a <
0,
and
express the above functions as differences of two upper functions. This shows that the collection of all integrable functions is a function space. • Thus, if f is integrable, then If I is also an integrable function. In particular, it follows from the last theorem that a function f is Lebesgue integrable if and only if and f- are both integrable. The next result describes the linearity property of the integral. Its easy proof follows directly from Definition 22.1 and is left as an exercise for the reader.
j+
Theorem 22.3. Iff and g are two integrable functions, then
1 (ct./ + f3g) dJL 11 dJL + f31 g dp. = ct.
holdsfor all a. f3 E IR.
Every positive integrable function is necessarily an upper function.
Chapter 4: THE LEBESGUE INTEGRAL
168
Theorem 22.4. If an integrable function f satisfies f > 0 a.e.. then f is an upperfunction.
=u
v a.e. holds. Since each u and v is the almost everywhere limit of a sequence of step functions, there exists a sequence { 1{1,} of step functions such that 1{1n � f a.e. Since f > 0 � f a.e. also holds. a.e., it follows that By Theorem 17.7, there exists a sequence {s,} of simple functions satisfying 0 < s, t f a.e. Now, for each n let ¢, = s, 1\ Then {¢,} is a sequence of step functions such that 0 < ¢, t f a.e. holds. To complete the proof, we show that {j ¢, d J.L} is bounded. Indeed, from ¢, + v < f + v < u a.e. and Theorem 21.5, it follows that J ¢, dJ.L + J v dJ.L < J u d /1-. and therefore, J ¢, d J.L < J u dJ.L J v dJ.L < oo holds for all n. The proof of the theorem is • now complete.
Proof. Choose two upper functions u and v such that f
-
t:
-
If f is an integrable function, then by Theorem 22.4, t + and .r- are both upper functions, and so, f = f+ - f- is a decomposition of f as a difference of two positive upper functions. In particular,
(This formula is usually the one used by many authors to define the Lebesgue integral.) As an application of the preceding theorem, we also have the following useful result.
Theorem 22.5. If f is an integrable function, then for evet)' E > 0 the mea
) > E} hasfinite measure.
surable set {x E X : l f(x l
Proof. Fix E > 0, let A
=
{x E X : lf(x)l > E}, and note that EXA < l f l holds. Now, � I f I is an integrable function, in fact, by Theorem 22.4 an upper function. Let {¢,} be a sequence of step functions such that ¢, t � I f I a.e. Then {¢, 1\ XA } is a sequence of step functions such that ¢, 1\ XA t XA a.e. holds. Thus, by Theorem 17 .6,
and the proof is finished.
•
Measurable functions "sandwiched" between integrable functions are integrable.
Section 22: INTEGRABLE FUNCTIONS
169
Theorem 22.6. Let f be a measurablefunction. If there exist two integrable
functions h and g such that h < f < g a.e., then f is also an integrable function. Proof. Writing the given inequality in the form 0 :::: f - h :::: g - h a.e., we see that we can assume without loss of generality that 0 < f :::: g a.e. holds. By Theorem 22.4, g is an upper function. Pick a sequence { ¢n} of step functions such that 0 < ¢, t g a. e. By Theorem 17.7 there exists a sequence { 1/r,} of simple functions such that 0 < 1/r, t f a.e. holds. But then {¢, 1\ 1/r, } is a sequence of step functions such that ¢, 1\ 1/r, t f a.e., and J ¢k 1\ Vrk dJJ- :::: lim J ¢, d!J- = J g dJJ- < oo for all k. Hence, f E U, and so, f is an integrable function. • More properties of the integral are included in the next theorem:
Theorem 22.7. For integrable functions f and g we have the following:
1 . J Ill d!J- = 0 if and only if f = 0 a.e. 2. If f > g a.e., then J f d!J- > J g d!J- . 3. If f dJJ-1 < J Ill dJJ-. Proof. ( 1) Clearly, if f = 0 a.e., then J Ill dJJ- = 0 holds. On the other hand, assume that J Ill dJJ- = 0. Since, by Theorem 22.4, 1 /1 is an upper function, there exists a sequence {¢,} of step functions such that 0 < ¢, t If I a.e. holds. By Theorem 21.5, it follows that J ¢n d!J0 for each n, and so, ¢, = 0 a.e. for each n. Thus, 1/1 = 0 a.e., and so f = 0 a.e. (2) Since f-g > 0 a. e., it follows from Theorem 22.4 that f-g is an upper func tion. But then, Theorem 2 1 .5 implies that J f dJJ- - J g dJJ- = j(J - g) dJJ- > 0, so that J f d!J- > J g dJJ- holds. • (3) The conclusion follows from (2) and the inequality -1/1 < f < 1 /1 . =
The reader has probably noticed that the functions we have considered so far are real-valued. It is a custom, however, to allow a function to assume infinite values, provided that the set of all points where the function equals -oo or oo is a null set. The reason for this is that neither the integrability character nor the value of the integral of a function changes by altering its values on a null set. Moreover, assigning any value to the sum of two functions at the points where the form oo - oo occurs does not affect the integrability and the value of the integral of the sum function (as long as the set of points of all such encounters has measure zero). If one does not want to deal with functions assuming infinite values (up, of course to null sets), then one may change the infinite values to finite ones (for instance, change all the infinite values to zero) without loosing anything regarding integrability. When an extended real-valued function f is said to define an inte grable function, it will be meant that f assumes the infinite values (or it is even
Chapter 4: THE LEBESGUE INTEGRAL
170
undefined) on a null set, and that if finite values are assigned to these points, then f becomes an integrable function. To summarize the preceding: •
Functions that are almost everywhere equal have the same integrability properties and can be considered as identical.
We continue with a theorem of B. Levi3 describing a basic monotone property of the integral.
Theorem 22.8 (Levi). Assume that a sequence {fn} of integrable functions satisfies /11 < f11+ 1 a.e. for each n and lim J /11 dJL < oo. Then there exists an illtegrablefunction f such that !11 t f a.e. (and hence, J fn d JJ. t J f dJL
holds).
Proof. Replacing { /11} by { f, - f1 } if necessary, we can assume without loss of generality that j,, > 0 a.e. hoids for each n. Aiso, an easy argument shows that we can assume that 0 < f, (x) t holds for all x E X . Let I = lim J fn dJL < oo. For each x E X let g(x) = lim .fn (x) E IR*, and consider the set E=
{x
E X : g(x) =
oo}.
Clearly, E = n:1 [U:dx e X : f,(x) > i}] holds, and so E is a measurable set. Next, we shall show that JL * (E) = 0. By Theorem 22.4, each /11 is an upper function. Thus, for each i there exists a sequence {¢J, } of step functions such that 0 < ¢� t n f; a.e. holds. For each n let '1/1n = V;'= 1 ¢j,, and note that { l/J, } is a sequence of step functions such that '1/111 t g a.e. and lim J l/J, dJL = lim J /11 d J.L = I . In particular, for each k the sequence of step functions { l/J, A f.: XE} satisfies '1/111 A k XE t f.: XE a. e. From Theorem 17 .6, it follows that JJ.*(E) < oo and kJL* (E) < lim J l/Jn dJL = I < oo for each k. Hence, J.L*(E) = 0. Now, define f : X -+ IR by f(x) = g(x) if x fl. E and f(x) = 0 if x E E. Then f11 t f a.e. holds, and the result follows from Theorem 21.6. II The series analogue of the preceding theorem is presented next.
Theorem 22.9. Let {/11} be a sequence of non-negative integrable functions 3 Beppo Levi (1 875-1961}, an halian malhemaiician. His main comribUiions were in algebraic 1opology, malhemaiical logic, and analysis.
Section 22: INTEGRABLE FUNCTIONS
171
such that L.: 1 J fn dJ.L < oo. Then Lr:: J f,, defines an integrable function and
Proof. For each n let g, .L�'= 1 f;, and note that each g, is an integrable function such that g, t L: J f; a.e. holds. Now, by Levi's theorem, L: J f,, =
defines an integrable function, and
holds.
•
The next result is known in the theory of integration as Fatou 's4 lemma.
Theorem 22.10 (Fatou's Lemma). Let {f:,} he a sequence ofintegrablefunc tions such that J,, > Oa.e. for each n and lim inf J f, dJ.L < oo. Then lim inf f,
defines an integrable function, ·and
I
lim inf f,, d 11-
I
< lim inf f, d J.t.
Proof. Without loss of generality, we can suppose that f,(x) > 0 holds for all x E X and all n. Given n, define g,(x) inf{f;(x) : i > n} for each x E X. Then each g, is a measurable function, and 0 < g, < f, holds for all 11. Thus, by Theorem 22.6, each g, is an integrable function. Now, observe that g, t and lim J g, dJ.t =s lim inf J f,, df..L < oo holds. Thus, by Theorem 22.8, there exists an integrable function g such that g11 t g a.e. holds. It follows that g lim inf f,, a.e., and therefore, lim inf fn defines an integrable function. Moreover, =
=
I
lim inf f,, dJ1
and the proof is finished.
=
Ig dJ.t
I
I
�� g, dJ.L < lim inf j, dJ.t,
= ��
•
We are now in the position to state and prove the Lebesgue dominated conver gence theorem-the cornerstone of the theory of integration. 4 Pierre Joseph Louis Fatou ( 1878-1929), a French mathematician. Besides his work in analysis, he also studied the motion of planets in astronomy.
Chapter 4: THE LEBESGUE INTEGRAL
172
Let { fn} be a sequence of integrable functions satisfying IfnI g a.e. for all nintegrable and somefunction fixed i12and tegrable function g. If f, f a.e., then f defines an lim jfn dj.L J lim fn dj.L jJ dj.L. Proof. Clearly, 1!1 < g a.e. holds, and the integrability of f follows from Theorem 22.6. Observe next that the sequence {g - fn} satisfies the hypotheses of Fatou 's lemma and moreover, lim inf(g - fn) - f a.e. Thus, Jg dJ-L - J f dJ.L = J(g - f) dJ.L Jlim inf(g - fn) dJ.L < lim inf J(g - fn) dj.L J g dj.L � lim sup J fn dj.L, Theorem 22.1 1 (The Lebesgue Dominated Convergence Theorem). < �
=
11-HXJ
=
II-+ 00
=
g
=
=
and hence,
J fn dj.L < J f dJ-L. Similarly, Fatou 's lemma applied to the sequence {g fn} yields Jg dJ.L + J f dJ-L J(g + f) dJ-L J lim inf(g + J,, ) dJ.L < lim inf J(g + fn) dJ-L J g dJ.L + lim inf J fn dj.L , lim sup
+
=
=
=
and so,
J f dJ.L < lim inf J fn dJ-L. Therefore, lim J fn dJ.L exists in and lim Jfn dJ.L J f d J.L holds. IR.,
=
•
The next theorem characterizes the Lebesgue integrable functions in terms of some given property. It is usually employed to prove that all Lebesgue integrable functions possess a given property.
Let S, J.L) be a measure space and let (P) be a property which may or may not be possessed by an integrable function. Assume that:
Theorem 22.12.
(X,
Section 22: INTEGRABLE FUNCTIONS
173
1. Iff and g are integrable functions with property (P ) , then f + g and af
for each a E IR also have property (P). 2. If f is an integrable function such that for each E > 0 there exists an integrable function g with property (P) satisfying J If - g l d�J- < E, then f has property (P ). 3. For each A E S with �J-(A) < oo, the characteristic function XA has property
(P).
Then every integrable function has property (P). 11- * (A) < oo. So, there exists a disjoint sequence {A,} of S such that A = U:1 A11• Put 8, = U;=l Ak for each n and note that 8, t A. From X B" = L�=l XAt• (3) and (1), we see that x8" has property (P) for each n. Since f IXA - x8, l d�J- = IJ-*(A) - �J-(8,,) -7 0, it follows from (2) that XA likewise has property (P). Next, assume that A is an arbitrary measurable set of finite measure and let E > 0. Then there exists a a -set 8 of finite measure such that A c 8 and IJ-*(8) < JL* (A) + E. This implies f IXA - XB I d�J- = 11-*(8) - �J-*(A) < E . From the above discussion and (2), we infer that XA satisfies property (P). Proof.
Assume first that
Now, from
A
is a a-set with
( 1 ) we see that every step function satisfies property (P).
then, it follows from
(2) that every upper function satisfies property (P).
But
Since
every integrable function is the difference of two upper functions, invoking
(1)
once more, we infer that indeed every integrable function satisfies property (P).
•
It is easy to see that for every subset
E of X, the collection SE
of subsets of E (called the restriction ofthe semiring S to
=
{EnA : A e S }
E) is a semiring of subsets
E is a measurable subset of X, then 11-* restricted to SE is also That is, (E, S£ , 11-*) is a measure space for every measurable subset
of E. If, in addition, a measure.
E of X. Also, a straightforward verification shows that the measurable subsets of (E, S£, 11-*) are precisely the subsets of E that are measurable subsets of X; see Exercise If
E
7 of Section 15.
is a measurable subset of
integrable over
X , then a function
f:E
-7
IR is said to be
E if f is integrable with respect to the measure space (£, SE, 11-* ).
f can be extended to all of X by assigning the values f (x) = 0 if x ¢ E. Then f so defined is an integrable function over X, and in this case fx f d�J- = JE f dJL holds. A function f : X -7 IR is said to be integrable over a measurable subset E of X if the function f XE is integrable over X, or equivalently, if f restricted to E is integrable with respect to the measure space (E, S£, JL*). In this case, we shall write J fXE d�J- = J£ f dJL. Of course, the domain of
The simple proof of the next result is left as an exercise for the reader.
174
Chapter 4: THE LEBESGUE INTEGRAL
Theorem 22.13. Every integrable function f is integrable over every mea
surable subset of X. Moreover,
holds for every measurable subset E of X . The rest of this section deals with an observation concerning infinite Lebesgue integrals. If ¢ = :L7= 1 a; XA; is the standard representation of a positive simple function ¢, then the sum :L7= 1 a;p,* (A;) makes sense as an extended non-negative real number. If :L;'=t a; p,* (A;) = oo, then it is a custom to write J ¢ dp, = oo and say that the Lebesgue integral of¢ is infinite. Assume now that f : X -+ IR� is a function where there exists a sequence {¢11} of positive simple functions such that ¢11 t f a.e. hol4s. Then lim J ¢11 dp, exists as an extended reai number, and it can be seen easily that lim J ¢n d;.,;, is independent from the chosen sequence {¢11 }. In the case that lim J
In this manner, we can assign an "integral" to a much larger class of measurable extended real-valued functions. One advantage ofthe above extension ofthe integral is that a number of theorems can be phrased without Lebesgue integrability assumptions on the functions. For instance, Fatou's lemma can be stated as follows: •
If {f,,} is a sequence of measurable functions satisfying fn > 0 a.e. for each n, then
J
lim inf f,, d J1 < lim inf
J fn d
J..L
holds-where, of course, one or both sides of the inequality may be infinite.
Section 22: INTEGRABLE FUNCTIONS
175
EXERCISES 1. Show by a counterexample that the integrable functions do not fonn an algebra. 2. Let X be a nonempty set, and let 8 be the Dirac measure on X with respect to the point a (see Example 13.4). Show that every function f : X -+ IR is integrable and that J f d8 = j(a). 3. Let J.L be the counting measure on IN (see Example 13.3). Show that a function f : IN -+ IR is integrable if and only if L::,, lf(n)l < oo. Also, show that in this case J f dJ.L = L:,t f(n). 4. Show that a measurable function f is integrable if and only if I f I is integrable. Give an example of a nonintegrable function whose absolute value is integrable. 5. Let f be an integrable function, and let IE 11} be a sequence of disjoint measurable subsets of X. If E = u:,, £,, then show that
1£ f dJ.L 11=f1 1E,f dJ.L . =
6. 7. 8.
Let f be an integrable function. Show that for each € > 0 there exists some 8 > 0 (de pending on €) such that If£ f dJ.LI < € holds for all measurable sets with J.L*(E) < 8. [HINT: Note that I/ l A n t lfll Show that for every integrable function f the set (x E X : f(x) =I= 0} can be written as a countable union of measurable sets of finite measure-referred to as a a-finite set. Let f : IR -+ IR be integrable with respect to the Lebesgue measure. Show that the function g : [0, oo) -+ IR defined by g( t )
9.
10.
= sup
{ j l f(x
+ y)
- f(x) I d>..(x) : IYI
:::
t}
for t =:: 0 is continuous at t = 0. [HINT: Use Theorem 22 . 12.] Let g be an integrable function and let If�,} be a sequence of integrable functions such that If, I < g a.e. holds for all n. Show that if f, 4 f, then f is an integrable function and lim J I f, - !I dJ.L = 0 holds. [HINT: Combine Theorem 19.4 with the Lebesgue dominated convergence theorem.] Establish the following generalization of Theorem 22.9: If (/11 } is a sequence of inte grable functions such that L::,1 J If, I d J.L < oo, then L::, 1 f, defines an integrable function and
[HINT: By Theorem 22.9, the series g = L::,1 If,,! defines an integrable func tion and IL�= I f, l ::; g a.e . holds for each k. Now, use the Lebesgue dominated convergence theorem.] 11. Let f be a positive (a.e.) measurable function, and let ; i l e; = J.L*
Chapter 4: THE LEBESGUE INTEGRAL
176
Let ( fn) be a sequence of integrable functions such that 0 ::::; f,+ 1 ::::; fn a.e. holds for each n. Then show that fn ..J, 0 a.e. holds if and only if J /11 dJ.L ..J, 0. 13. Let f be an integrable function such that f(x) > 0 holds for almost all x. If is a measurable set such that fA f dJ.L = 0, then show that J.L*(A) = 0. 14. Let (X, S, J.L) be a finite measure space and let f : X � 1R be an integrable function satisfying f(x) > 0 for almost all x. If 0 < e ::::; J.L*(X), then show that
12.
A
inf 15.
A
J.L*(E) :::: e
I
> 0.
v(A) = fA! dJ.L
(X, A, v) is a measure space.
If A 11 denotes the a -algebra of all v-measurable subsets of X, then A 5; A 11 • Give an example for which A ::j: A 11• If J.L*((x E X : /(x) = 0)) = 0, then A = A11• If g is an integrable function with respect to the measure space (X, A . v), then fg is integrable with respect to the measure space (X, S, J.L) and
c. d.
17.
E All and
Let f be a positive integrable function. Define v : A � [0, oo) by for each E A . Show that: a. b.
16.
It f dJ.L : E
1g dv 1gf dJ.L . =
(A Change of Variable Fonnula) Let be an interval ofiR., and let f : -+ 1R be an in tegrable function with respect to the Lebesgue measure. For a pair of real numbers a and b with a ::j: 0, let J = ((x -b)/a : x E I). Show that the function g : J � 1R de fined by g(x) = f(ax +b) for x E J is integrabie and that f1 f d>.. = Ia I f1 g d).. holds. [HINT: Use Theorem 22.12.] Let (X, S, J.L) be a finite measure space. For every pair of measurable functions f and g let
I
I
- I 1 +I fI f- -g lgl dJ.L.
d(f' g) -
Show that (M, d) is a metric space. Show that a sequence Un) of measurable functions (i.e., (/n) 5; M) satisfies /11 � f if and only if lim d(f,, /) = 0. Show that (M, d) is a complete metric space. That is, show that if a sequence (/n) of measurable functions satisfies d(/11, fm) -+ 0 as n, m � oo, then there exists a measurable function f such that lim d(fn, /) = 0.
a. b. c.
18.
!I
f
=
Let
:
1R � 1R be a Lebesgue integrable function. For each finite interval f1 f d>.. and E1 = {x E I : f(x) > /J ). Show that
>.ln
{ (f - /J)d>.. . 11{ I f - /J i dJ.. = 2 1£,
19.
Let f : [0, oo) � 1R be a Lebesgue integrable function such that for each t :::: 0. Show that f(x) = 0 holds for almost all x.
I
J� f(x)dJ..(x)
let
=
0
Section 23: THE RIEMANN INTEGRAL AS A LEBESGUE INTEGRAL 20.
177
Let (X, S, J.l) be a measure space and let f, !1, h, . . . be non-negative integrable functions satisfying j,, -+ f a.e. and lim J /11 df.l = J f df.l. If E is a measurable set, then show that
n-+limoo }E{ !" df.l = }£{ f df.l.
21.
22.
If a Lebesgue integrable function f : [0, 1] -+ IR satisfies Jd x2" j(x) d>..(x) = 0 for each n = 0, 1 , 2, . . . , then show that f = 0 a.e. [HINT: Since the algebra generated by {1. x2} is uniformly dense in C[O, 1], we have
Jd g(x)f(x)dx = 0 for each g E C[O, 1].] 11 11 1 For each 11 consider the partition {0, 2-1 , 2 · 2- , 3 2- , , (211 - 1) 2-11 , 1 } the interval [0, 1 ] and define the function r11 : [0, 1 ] -+ IR by r11 (1) = - 1, and r11(x) = ( - 1 )"- I for (k - 1 )2 _, :=: x < k2 -n and each k = 1 , 2, . . . , 211 • ·
a. b.
lim
of
lot r11(x)j(x)d>..(x)
=
0.
[HINT: For (b) use Theorem 22.12.] Let { E11} be a sequence of real numbers such that 0 < E11 < 1 for each n. Also, let us say that a sequence { A11) of Lebesgue measurable subsets of [0, 1 ] is consistent with the sequence { E,} if I..(A11) = E11 for each n. Establish the following properties of !En}: a.
b. 24.
·
Draw the graphs of r 1 and r2. Show that if f : [0, 1 ] -+ IR is a Lebesgue integrable function, then
n-+oo
23.
• • •
The sequence {E, ) converges to zero if and only if there exists a consistent sequence {A,) of measurable subsets of [0, 1 ] such that L::,.1 XA, (x) < oo for almost allx. The series I:: 1 E11 converges in IR if and only if for each consistent sequence {A,) of measurable subsets of [0, I ] we have L::,.1 XA, (x) < oo for almost all x .
Let (X, S. p.) be a finite measure space and let f : X -+ IR be a measurable function. a. b.
Show that if f" is integrable for each n and lim J f" dJ.l exists in IR, then 1/(x)l � 1 holds for almost all x. 1 11 If / 1 is integrable for each 11, then show that J / dJ.l = c (a constant) for 11 = 1, 2, . . . if and only if f = XA for some measurable subset A of X.
23. THE RIEI\'IANN INTEGRAL AS A LEBESGUE INTEGRAL It will be shown in this section that the Lebesgue integral is a generalization of the Riemann5 integral. We start by reviewing the definition of the Riemann integral. 5Georg Friedrich Bernhard Riemann (1 826-1866), a German math�matician, one of the greatest mathematicians of all time. Although his life was short, his contributions left the impact of a real genius. He made path-breaking contributions to the theory of complex functions. space geometry, and mathematical physics.
Chapter 4: THE LEBESGUE INTEGRAL
178
For simplicity, the details will be given for functions of one variable, and at the end it will be indicated how to carry out the same results for functions of several variables. Unless otherwise specified, throughout our discussion, f will be a fixed bounded real-valued function on a closed interval [a, b] of JR. A collection of points P {x0, x1 , , x,} is called a partition of [a, b] if
= = Xo < X1 < · · < = b •
a
holds. Every partition P tervals
= {x0, x1,
•
•
X11
·
•
•
•
, x,} divides [a, b] into the n closed subin
=
=
The length of the largest of these subintervals is called the mesh of P and is denoted by I P I ; that is. I P I max{x; - x; _ 1 : i 1 , . . . , n } . A partition P is said to be finer than another partition Q if Q c P holds. If P and Q are partitions, then P U Q is also a partition that is finer than both P and Q . For a partition P {xo, x1 , , x,} of [a, b], let
=
m; = inf{j(x) : x
=
•
E
•
•
M;
[x; _ 1 , x;]} and
= sup{f(x) : x
E
[x;_ 1 , x; ]}
for each i 1 , . . . , n. Then the lower sum S.(f, P) of f corresponding to the partition P is defined by S*(f, P)
n
= Li= l m (x - X;-1), ;
;
and similarly, the upper sum S* (f, P) of f by S* (f, P )
= Li=l "
M;(X;
-
Xi-1 ) .
Clearly, S*(f, P) < S*(f, P) holds for every partition P of [a, b]. Lemma 23.1.
then
Ifa partition P isfmer than anotherpartition Q (i.e., Q c P),
S* (f, Q) < S* (f, P) and S*(f, P) < S* (f, Q). Proof. We show that S* (f, Q) < S (f, P) holds. The other inequality can be * proven in a similar manner.
Section 23: THE RIEMANN INTEGRAL AS A LEBESGUE INTEGRAL
179
To establish the inequality, it is enough to assume that P has only one more point , x,} and P = Q U {t }. Then there exists than Q, say t. So, let Q (x0, x 1 , some k ( 1 < k < n) such that Xk-1 < t < x�.;, and thus, P = {xo, x 1 , . . . , Xk- l , t, XJ.:, , x11 } . Let c 1 = inf{f(x ) : x E [xk_1 , t] } and c2 = inf{f(x) : x E [t, xk]} . Observe that both m�.: < c1 and m �.: < c2 hold. Therefore, =
•
•
•
•
•
•
S* (f, Q) = <
II
l: m; (x; - x; - 1 ) = l: m; (x; - x;-J .) + mdxk - x�.: -1) i=l
i�k
L m;(x; - X;- t) + c1 (t - x�.:-d + c2(x�.: - t) i#
holds, and the proof is finished.
Lemma 23.2.
•
For every pair ofpartitions P and Q, we have
Proof. From Lemma 23.1, it follows that * S*(f, P) < S*(f, P U Q) < S (f, P U Q) < S* (f, Q) ,
as claimed.
•
The preceding lemma states that every upper sum is an upper bound for the collection of all lower sums of f, and similarly, every lower sum is a lower bound for the collection of all upper sums. Thus, if the lower Riemann integral of f is defined by /* (f) = sup{ S*(f, P) : P is a partition of [a , h]} .
and the upper Riemann integral of f by ! * (f) = inf{S*(f, P ) : P is a partition of [a, b] } ,
Chapter 4: THE LEBESGUE INTEGRAL
180
then
S.(J, P)
<
holds for every pair of partitions
I.(/ ) < / * (f) < S* (f, Q ) P and Q of [a, b].
Definition 23.3. A boundedfunction f : [a, b] -+ IR is called Riemann inte grable if I.(/) = I * (f). In this case, the common value is called the Riemann integral off and is denoted by the classical symbol f(x)dx. Historical Note:
Riemann's definition of the integral is
J:
a generalization of Eudoxus'
method of exhaustion as was used by Archimedes6 in his computation of the area of a circle. Interestingly, the value of the limit of the areas of the inscribed (or circumscribed) polygons that were employed by Archimedes to compute the area of the circle, was also called by him the integral
(ro rrav).
It should be historically correct to call the Riemann integral the
Eudoxus-Archimedes integral or the Eudoxus-Archimedes-Riemann integral (or even the
Archimedes-Riemann integral).
A characterization for the Riemann integrability of a function, known as Rie mann's criterion, is presented next.
Theorem 23.4 (Riemann's Criterion). A bounded function f : [a, b]
IR is Riemann integrable if and only iffor every E > 0 there exists a partition P of [a, b] such that S* (f, P) - S.(J, P) < E holds. -+
Riemann integrable and let E > 0. Then let I = J: j (x) dx. Then there exist t wo partitions P1 and P2 of [a, b] such that I - s.cJ, PI ) < E and S* (f, P2) - I < E . Then (by Lemma 23.1), the partition P = P1 U P2 satisfies
Proof. Assume
that
f
is
S* (J, P) - S* (f, P)
< =
<
S* (f, P2) - S* (f, P1) [S * (J, P2) - I ] + [I - S* (f, P1 )] E + E = 2E.
6Archimedes of Syracuse (287-212 BC), a Greek mathematician and inventor. He was the mo;t celebrated mathematician of antiquity and perhaps the best mathematician of all times. He used EudOxus' method of exhaustioo to compute areas and volumes very successfully. In his classic work, TheMeasurement of the Circ:/e, he established that a circle has the same area as a right triangle having ooe leg equal to the radius Of the circle and the other equal to the circumference of the circle, and that the VOlume of a sphere is four times the VOlume Of a right cooe with radius and height equal tO the radius of the sphere.
Section 23: THE RIEMANN INTEGRAL AS A LEBESGUE INTEGRAL
181
Conversely, if the condition is satisfied, then since 0
< /* (f) - I*(f) < S* (f, P) - S* (f, P)
holds for every partition P of [a, b], we have O < l*(f) - l* (f) < E for all E > 0. 0, or l (f) Hence, l * (f) - l* (f) l*(f), which shows that f is Riemann * integrable. • =
T
for i
{ t1 ,
=
•
1,
, x,} be a partition of [a, b].
A collection of points , t,} is said to be a selection of points for P if x; _ 1 < t; < x; holds . . . , n. We write
Now, let P =
{x0, x1 ,
=
•
=
•
•
•
•
Rt
;
f(t; )(x; - x - J ), L i= l II
=
and call it (as usual) a Riemann sum associated with the partition P. The following theorem ofJ.-C. Darboux7 presents some powerful approximation formulas for the Riemann integral. The theorem can be viewed as an abstract formulation of Eudoxus' exhaustion method.
Theorem 23.5 (Darboux). Let f : [a, b]
-r
JR. be Riemann integrable, and
let ( P11 } be a sequence ofpartitions of [a, b] such that lim
11--+oo
S* (f, P, )
=
lim
11-+-00
S*(f,
P,)
=
lim I P, l
lim Rt(P11 , T11 )
=
0.
Then
1bf(x)dx. a
In particular, if a sequence ofpartitions { P,} satisfies is a selection ofpoints for P11 , then
n--+oo
=
lim I P, l
0
=
and T,,
1bf(x)dx. a
l f(x) l < c holds for all x e [a, b]. Let E > 0. By Theorem 23.4, there exists a partition P {xo, x 1 , , Xm} of [a, b] such that S*(f, P) - S* (f, P) < E. Choose n0 such that Proof. Choose a constant c
> 0 such that
=
I Pn l
f
< 2cm --
7 Jean-Gastin
and
IP111
•
•
•
< min{XJ - xo, X2 - XJ , . . . , Xm - Xm- J }
Darboux ( 1842-1917), a French mathematician. He was a geometer who used his
geometric intuition to solve various problems in analysis and differential equations.
Chapter 4: THE LEBESGUE INTEGRAL
182
for all n
> n0. Fix n > n0, and let P, = { t0, I J ,
MJ = s p{ j(x) M; = s p{ j(x) u
u
:x :
x
• • .
E [XJ-J . Xj]}
for
[t;_1, t;]}
for
E
, tk}.
Put
j = 1, 2, . . . m, i = l , . . . , k. .
The definitions of mr and m; are analogous (replace the sups by infs). Then
0<
=
f." f(x) dx - S,(f, P,) < S*(f, P,) - S,(f, P,) k
(M; - m;)(t; - t;_1) = V + W, L l i= (M;
where V is the sum of the terms -m; )(t; - 1;_1) for which [t;_1, t; ] lies entirely in some subint�".'al of the partition P , a.Tld W is the sum of the remaining terms. The sums V and W are estimated separately. We estimate V first. Note that V = 'E1 + · · · + 'Em, where each 'E1 is the sum of the terms m;)(t; - 1;_1) for which [1;_1, t;] c [x1_h x1] holds. But if m{ holds. Also, the sum of the [1;_1, t;] c [x1_ h x1 ], then M; - m; < lengths of those subintervals of the partition P, that lie inside [xJ-l, x1] never exceeds Xj - Xj-l · Thus, "£ j < r - mr)(Xj - Xj-t) holds, and hence
(M; -
MJ -
(M
Ill
v < L (Mf - mr )<xj - Xj- l ) = S*(j. P) - S* (f, P) < E . j=l
Now, we estimate W. Let (M; - m;)(ti - 1;_1) be a term of the sum W. Since !Pn l < min{x1 - Xj-l : j = 1 , . . . . m}, there exists exactly one j with 1 < j < n such that x1_1 < t;-t < x1 < t; < XJ+l · Thus, the sum W has at most m terms, and since E
E
=(M; - m;)(t; - t;_1) < 2c !P,! < 2c · 2em m it follows that
W < m(Ejm) = E . Thus,
0< for all n
f." j(x) dx
- S,(f,
!'•J < V + W < < + < = 2£
> no. That is, lim s (/, P,) = I: f(x) dX.
* h 0 ,:::: S*(j, P, ) - Ia j(x) dx < V + W < 2E holds for all n > no, and so, lim S*(f, P,) = I: j(x) dx. The last part follows immediately from the Similarly,
Section 23: THE RIEMANN INTEGRAL AS A LEBESGUE INTEGRAL
183
inequalities
S* (f, P11) < Rf(P,. T,, ) < S* (J, P,). •
The proof of the theorem is now complete.
We are now ready to establish that the Lebesgue integral is a generalization of the Riemann integral. Here, " f is Lebesgue integrable" means that f is integrable with respect to the Lebesgue measure.
EveryRiemann integrableftmction f : [a, b] integrable, and in this case the two integrals coincide. That is,
Theorem 23.6.
---+
Proof. For each n let
I f dA = {f(x)dx. P11 = {x0, x 1 , . . . , x2, } (b - a )2-1 ;
2"
=
¢, = L m iX[x;_1,x;)
i= l
is Lebesgue
=
be the partition that divides [a, b] that is, x; a+ i (b-a)2-11 •
1 into 2 subintervals all of the same length Let
JR.
and
2"
1/J, = Li=l M;X[x;_1,x;)• M; = {f(x) x
sup : E [xi 1 • x;]}. where m ; inf {f (x) : x E [Xi - 1 • X;]} and Clearly. {¢11 } and { 1/111} are two sequences of step functions satisfying the propetries cf>��(x) t< f(x) < 1/111(x) ,J, for each x E [a, b). Now, if ¢,(x) t g(x) and 1/J,(x) ,J, h(x), then by Theorem 22.6, both functions g and h are Lebesgue integrable and g(x) < f(x) < h(x) holds for all x E [a, b). Also, by definition, J c/J11 d>... S*(f, P11) and J 1/ln d>... S* (f, P11 ). Therefore, since 1/J,(x) - cj>11(x) ,J, h(x) - g(x) > 0, it follows that
= o < Jch - g) d>... = lcl/111 II = 11-+00 S*(f, P11) - 11-+00 S*(J. P11) .• h=g=f f lim
-HXJ
lim
= II-+00 11/J,
cfJ,)d>... = lim
lim
= 0,
l
c/J, d>... II-+00
d>... - lim
=
where the last equality holds true by virtue of Theorem 23.5. This implies h - g 0 a.e and hence, a. e. holds. In particular, c/J, t f a.e. and 1/111 ,J, f a.e. hold, which show that is Lebesgue integrable-in fact, an upper function. Finally,
I f d>... = l lim
11-+oo
c/>11 d>... = lim
n-oo
and the proof of the theorem is finished.
S* (f, P, )
= 1"
f(x) dx.
u
•
Chapter 4: THE LEBESGUE INTEGRAL
184
The next theorem, due to H. Lebesgue and G. Vitali, characterizes the Riemann integrable functions in terms of their discontinuities. (The almost everywhere relations are considered with respect to the Lebesgue measure.)
Theorem 23.7 (Lebesgue-Vitali). A bounded functioll f : [a, b]
--+
IR is
Riemanll il7tegrab/e if and only if it is colltinuous almost everywhere. Proof. For each ll, let Pn. ¢n. and 1/f, be as they were introduced in the proof of Theorem 23. 6. Assume first that f is Riemann integrable. Then a glance at the proof of Theorem 23.6 guarantees the existence of a (Lebesgue) null subset A of [a , b] such that ¢,(x) t f(x) and 1/f,(x) -!- f(x) for all x ¢ A. Clearly, D = A U ( U: 1 P,) has Lebesgue measure zero, and we claim that f is continuous on [a, b] \ D. To see this, let s e [a , b] \ D and E > 0. Pick some n with f(s) - ¢n(s) < E and 1/f,(s) - f(s) < E . Then there exists some subinterval [x;_ 1 , X;] of the partition P;; such that. \' e (x; -1 , Xi). Clearly, ¢,(s) = m; and 1/f,(s) = M;. Therefore, if x e (x;-lt x;), then - E < m; -
f(s)
<
f(x) - f(s)
<
M; - f(s)
< E.
Since (x;_ 1 , x;) is a neighborhood of s, the last inequality shows that f is contin uous at s. This establishes that f is continuous almost everywhere. For the converse, assume that f is continuous almost everywhere. Let s # b be a point of continuity of f. If E > 0 is given, then choose some 8 > 0 such that
f(s) - E
<
f (x)
<
f (s) + E
for all x e [a, b] with lx - sl < 8. Pick some k so that IPkl < 8. Then for some subinterval [x;-1 , X;] of Pkt we must have s e [x;_1 , x; ). In particular, lx - s l < 8 must hold for all x e [x;_1 , x; ], and so, from (•) we get
j(s) - E
<
m; = ¢k(s)
<
j(s) + E.
Since ¢,(s) t, it easily follows that ¢,(s) t j(s). Similarly, 1/f,(s) -1- f(s) holds. Since f is continuous almost everywhere, we conclude that ¢, t f a.e. and 1/f, -1- f a.e. both hold. This shows that f is Lebesgue integrable. Moreover, by the Lebesgue dominated convergence tfieorem we have
S*(f, Pn) =
J¢, d/, t J f d).
and
S*(f, P,) =
J1/J, d), -1- J f d)..
Section 23: THE RIEMANN INTEGRAL AS A LEBESGUE INTEGRAL
185
Thus, lim[S*(f, Pn) - S* (f, P,)] 0, and-so, by Theorem 23.3, the function f is Riemann integrable. The proof of the theorem is now complete. • =
An immediate consequence of Theorem 23.7 is the following:
The collection ofall Riemann integrablefunctions on a closed interval is a fwzction space and an algebra offtmctions.
Theorem 23.8.
It is easy to present examples of bounded Lebesgue integrable functions that are not Riemann integrable. Here is an example:
Example 23.9. Let f : [0, 1] � IR be defined by f(x) = 0 if :c is a rational number and f (x) = 1 if x is irrational (in other words, f is the characteristic function of the irrationals of [0, 1]). Then f is discontinuous at every point of [0, 1], and thus, by T heorem 23.7, f is not Riemann integrable. On the other hand, f = 1 a.e. holds (since the set of rational numbers has Lebesgue measure zero), and so, f is Lebesgue integrable. Also, note that • f f d>.. = 1 holds. It follows from Theorem 23.7 that if a function f : [a, b] � JR. is Riemann integrable, then f restricted to any closed subinterval of [a, b] is also Riemann integrable there. Moreover, by the same theorem, if two functions f and g are Riemann integrable on [a, c] and [c, b], then the function h : [a, b] � JR., defined by h(x) = f(x) ifx E [a, c] , and h(x) = g(x) ifx E (c, b], is Riemann integrable. Clearly, by Theorem 23.7, every continuous function on a closed interval is Riemann integrable. To compute the Riemann (and hence, the Lebesgue) integral of a continuous function, one usually uses the fundamental theorem of calculus, one form of which is stated next. Since any "reasonable" calculus book contains a proof of this important result, its proof is omitted. (See also Exercise 6 at the end of the section.) The fundamental theorem of calculus is due to I. Newton8 and independently to G. Leibniz.9
Theorem 23.10 (The Fundamental Theorem of Calculus). For a continu
ousftmction f : [a, b] � JR. we have the following:
8Isaac Newton ( 1642-1727). a great British mathematician, physicist, astronomer, and philoso pher. He discovered the law of gravity and was one of the founders of calculus. His pioneering original contributions to mathematics and science revolutionized the modem scientific app roach. 9Gottfried Wilhelm Leibniz ( 1646-1716), a prominent Gennan mathematician and philosopher. He was a person with "universal" scientific interests. Besides his philosophical and metaphysical contributions, he contributed substantially to mathematics, mathematical logic, and physics. Together with Newton, he is considered the founder of calculus.
Chapter 4: THE LEBESGUE INTEGRAL
186
If A : [a, b] --+- IR is an area function off (i.e., A(x) = I;rf(t)dt holds for all x E [a, b] and some fixed c E [a, b]), then A is an antiderivative of f. That is, A'(x) = f(x) holdsfor each x E [a, b]. 2. If F : [a, b] --+- IR is an amiderivative off, i.e., F '(x) = f(x) holds for each x e [a, b], then 1.
{f(x)dx = F(b) - F(a).
In a conventional way, the integral I: f(x)dx is defined to be - I: f(x)dx; that is, I: f(x)dx = - I <x)dx. Also, I: f(x)dx is defined to be zero. By doing so, the useful identity
!J
idf(x)dx = f .f(x)dx rd f(x)dx c
·
c
+
Jc·
Jc
holds regardless of the ordering between the points c, d, and e of [a, b]. We now indicate how to extend the above results to functions ofseveral variables. In the general case, the interval [a, b] is replaced by a cell J = [a 1 , b1] x · · · x [an, b,], and its Lebesgue measure is A.(J) = n7=1 (b; - a;). A partition P of 1 is a set of points of the form P = P1 x · · · x Pn, where P; is a partition of [a;, b;] for each i = 1, . . . , 12. Clearly, any partition P divides J into a finite number of subcells. Now, if f : J --+- IR is a bounded function and the partition P divides J into the subcells 11 , , lt. then we define again the numbers •
•
•
m; = M; =
. . . , x,) E J; }, sup { f(xt , . . . , Xn) : (X t , . . . , x,) E J; }. inf{ f(xt, . . . , Xn ) : (xt,
The lower and upper sums corresponding to the partition by the formulas
k
S* (f, P) = L m;A.(J;) i=l
and
P are defined as before k
S *(f, P ) = L M;A.(J;), i=l
respectively. The lower Riemann integral of f is defined (as before) by
I*(f) = sup( S*(f, P ) : P
is a partition of
J},
Section 23: THE RIEMANN INTEGRAL AS A LEBESGUE INTEGRAL
187
and the upper Riemann integral of f by. /*(f)
=
inf{S*(f,
P) : P
is a partition of J}.
As in the one-dimensional case, -00 < / (/) < / * ( f) < 00 *
=
holds. The function f is called Riemann integrable if / (f) /*(f). This * common number is called the Riemann integral of f and is denoted by
All the results given in this section are valid in this general setting. Their proofs parallel the ones presented here, and for this reason we leave them as an exercise for the reader.
EXERCISES 1. Let
f : [a, b]
� JR. be Riemann integrable. Show
1tl
1! !d 1 = f(x) dx + f(x) dx
every closed subinterval of [a, b ]. Also, show that ,
2.
c:
I!
holds for every three pon i ts c, d, and e of [a, b]. Let f : [a. b] � JR. be Riemann integrable. Then show that
1b f(x)dx U
3.
f(x) dx
-- L. f (a + - a ) . f" b =1 1 11-+00 = lim
u-Hx:>
b-a n
l=l
f,,(x)dx
u
5. 6.
i (b
11
)
n
Let (f,, ) be a sequence of Riemann integrable functions on [a, b] such that {/11 } converges uniformly to a function f. Show that is Riemann integrable and that lim
4.
that f is Riemann integrable on
For each n . let j�, : [0, 1] � JR. be defined by
f(x)dx.
u
/11 (x) =
"j�:.' for all x E [0, 1 ]. Then
show that lim f f,,(x)dx = � · [HINT: Use integration by parts.] Let f : [a. b] � JR. be an increasing function. Show that f is Riemann integrable. [HINT: Verify that f satisfies Riemann's criterion.] (The Fuudamental Theorem of Calculus) If f : [a , b] � JR. is a Riemann integrable function, define its area function A : [a, b] � JR. by A(x) = J�\- f(t)dt for each x E [a, b]. Show that
d
Chapter 4: THE LEBESGUE INTEGRAL
188
a. A is a uniformly continuous function. b. If f is continuous at some point of [a, b], then A is differentiable at and 1 A ( ) = j{c) holds. c. Give an example of a Riemann integrable function f whose area function A is differentiable and satisfies N =F f.
c
c
c
[HINT: For part (c) use the function defined in Exercise 7 of Section 9.] 7. (Arzela) Let {/11} be a sequence of Riemann integrable functions on [a, b] such that lim f,,(x) = f(x) holds for each x E [a, b] and f is Riemann integrable. Also, assume that there exists a constant M such that lf,,(x)l < M holds for all x E [a, b] and all Show that
n.
11-HXJ
lim
f(x) dx. J{b a fn(x) dx = J{b a
Determine the lower and upper Riemann integrals for the function of Example 23.9. 9. Let C be the Cantor set (see Example 6.15). Show that xc is Riemann integrable over [0, 1], and that Jd xc dx = 0. 10. Let 0 < E < 1 , and consider the E-Cantor set CE oi iO, ij. Show thai xcf is not Riemann integrable over [0. 1]. Also, determine I.(xcf ) and / *(xcf ). [HINT: Show that the set of all discontinuities of xcf is CE.] 11. Give a proof of the Riemann integrability of a continuous function based upon its uniform continuity (Theorem 7.7). 12. Establish the familiar change of variable formula for the Riemann integral of contin uous functions: If [a, bl---4 [c, d] ....L. IR are g differentiable ( e , g ha.li a derivative), 8.
i
ously
.
.
iab
continuousftmctions with continu continuous then
f(g(x))l(x)dx =
lg(b) g(a)
f(u)du.
13. Let f : [0, oo) � IR be a continuous function such that limx-oo f(x) = 8. Show that limn-oc J; f( x) dx = a8 for each a > 0. 14. Let f : [0, oo) � IR be a continuous function such that f(x + J) = f(x) for all x :::. 0. If g : [0, 1] � IR is an arbitrary continuous function, then show that
n
of {0, J]. Let f : [0, 1] � [0, oo) be Riemann integrable on every closed subinterval 1 Show that f is Lebesgue integrable over [0, J] if and only if limE�O fe f(x) dx exists in IR. Also, show that if this is the case, then J f d)... = limE�O JE1 f(x) dx. 16. As an application of the preceding exercise, show that the function f : [0, J] � IR defined by j(x) = xP if x E (0, J] and f(O) = 0 is Lebesgue integrable if and only if p > J. Also, show that if f is Lebesgue integrable, then 15.
-
f
j d)... =
l +�-. p
-
17. Let f : [0, l ] � IR be a function and define g : [0, J] � IR by g(x) = ef<x>.
Section 23: THE RIEMANN INTEGRAL AS A LEBESGUE INTEGRAL
f
189
g.
a. Show that if is measurable (or Borel measurable), then so is b. If f is Lebesgue integrable, is then 8 necessarily Lebesgue integrable? c. Give an example of an essentially unbounded function which is continuous on 2, . . . . (A function 1] such that /" is Lebesgue integrable for each n is "essentially unbounded," if for each M > the set {x e l] : f(x)l > M} has positive measure.)
(0,
18.
19.
f = 1,
0
f
[0,
[HINT: For consider the function f(x) = x-! .] Let f : [ l] -+ IR be Lebesgue integrable. Assume that f is differentiable at x 3 and Show that the function IR defined by for = x-1 x e and is Lebesgue integrable. ] x [c. -+ IR be a continuous function. Show that the Riemann integral Let off can be computed with two iterated integrations. That is, show that
(b) 0, f(O) (0, 1]= 0. g(O) = 0 f : [a, b d]
t 1d
f(x ,
y)dxdy
I
g: [0, 1] -+
=
g(x)
=0 f(x)
t [1d f(x, y) dy J dx = 1" [{f(x, y)dx J dy.
Generalize this to a continuous function of n variables. ] -+ IR are two continuous functions such ] -+ IR and g : 20. Assume that < < :5 for E b]. Let that E b] and A= E IR2
[a, b
f : [a, b g(x) x [a, f(x)
g(x)} .
a.
{(x, y)
:x [a,
f(x) y
Show that A is a closed set-and hence, a measurable subset of IR2. b. If lz : A IR is a continuous function, then show that h is Lebesgue integrable over A and that
21.
-+
1 hd'J.. = 1ab [ f g(x)lz(x, y)dy] dx. A
f(x)
Let -+ IR be a differentiable function-with one-sided derivatives at the end-points. If the derivative !' is uniformly bounded on then show that f' is Lebesgue integrable and that
f : [a, b]
22. Let /, 8 :
[a, b],
r
J[a.b]
!'
d'J... = f(b)- f(a).
[a. b] -+ IR be two Lebesgue integrable functions satisfying lx f(t)d'J.. (t) :::= lx g(t)d'J.. (t) for each E [a, b]. If ¢ : [a, b] -+ IR is a non-negative decreasing function, then show that the functions ¢f and cpg are both Lebesgue integrable over [a, b] and that x
they satisfy
lx cp(t)f(t) d'J.. (t) 1x cp(t)g(t) d'J..(t) <
[a, b]. [IDNT: ao a1
for all x E Prove it first for a decreasing function of the form ¢ 'Lf=l c; X[a;-�oa;)• where < < · · · < is a partition of b], and then use the fact that there
ak
[a,
=
190
Chapter 4: THE LEBESGUE INTEGRAL exists a sequence {¢n} of such step functions satisfying ¢11 (1) t ¢(t) for almost all t in [a, b]; see Exercise 8 of Section 2 1 .]
24. APPLICATIONS OF THE LEBESGUE INTEGRAL
If a function oo) --+ 1R (where, of course, E 1R) is Riemann integrable on every closed subinterval of [a, oo then its improper Riemann integral is defined by
f: [a,
)
a
,
}a
r)O lim rr f(x) dx' Ja f(x) dx = r-HX>
provided that the limit on the right-hand side exists in IR. The existence of the prior limit is also expressed by saying that the (improper Riemann) integral exists. Similarly, if - oo, --+ 1R is Riemann integrable on every closed subintervai of -oo, then f�00 [(x) dx is defined by
f : ( a] ( a], la f(x)dx ,._!�! 00 1° f(x)dx
J:O f(x) dx
=
-oo
whenever the limit exists in
Jboo f(x) dx
r
IR.
also exists for each
It should be clear that if
b>a
and
fa00
f(x) dx
exists, then
['"f(x)dx J."f(x)dx + {"f(x)dx. =
on eve1y Assume that f: i s Riemann integr a ble closed subinterval of [a, oo). Then faxf(x) dx exists if and only iffor eve1fory > 0 ther e exi s ts some M > 0 (depending on such that I J; f(x) dxl all s, t > M. Proof. I Jcr: f(x) dx M>0 r > M. s, t M, I/ -J�-f(x)dxl rf(x) dx J>(x) dx - t f(x) dx I - }[' f (x) d I - f (x) d j.fa a Theorem 24.1.
[a. oo) --+ 1R
<E
E)
E
Assume that
=
exists. Pick a real number
< E holds for all
that
If
2:.
such
then
= <
X +
X < 2E.
Conversely, assume that the condition is satisfied. If {a11} is a sequence of [a, oo) such that lim x} is = oo, then it is easy to see that the sequence
an
{j:n f(x) d
191
Section 24: APPLICATIONS OF THE LEBESGUE INTEGRAL
A f(x) dx IR. {b,} = [a, = B = J:" f(x) dx. b, !. a, f. Al - Bl < A - a f(x)dx a, f(x)dx B - f.h,a f(x)dx , A - BI < E E > A = B, fa00 f(x) dx
Cauchy. Thus, lim J:" of oo) with limb, oo; let
exists in lim
Now, let be another sequence From the inequality
+
+
it is easy to see that I holds for all 0. Thus, limit is independent of the chosen sequence. This shows that and the proof is finished.
and so the exists,
•
f(x) J: l f (x)l dx dxl s t, Lemma 24.2. If a function f: a, oo) -+ IR is Riemann itttegcable otz every closed subitzterval of[a, oo) and face 1/(x)l dx exists thetz J;� f(x) dx also exists and 1fa00 f(x) dxl faoo l f(x)l dx. f fa00 f (x) dx fa00 I f (x) I dx f(x) = si�I x = oo). J000 si�x dx = I }�00 si-�xl dx. From the preceding theorem, and the inequality < we have the following:
11:�
<
for
[
<
The converse of this lemma is false. That is, there are functions exist but for which improper Riemann integrals to exist. A well-known example is provided by the function I ) over [0, We shall see later that /(0) does not exist. To see the latter, note that
whose fails (with
but
2 1br -I x.l dx >- - 1 lsmxldx. . = krr X I
sin
br
(k-l)rc
holds for each
k
rc
(k-l)rr
k. Therefore,
J."rc -I x.l dx - L 1kir .xl d.'l. > -2 :L - . k X X 7C 0
sin
_
!sin
"
k=l
(k-l)rr
"
I
k= l
J000 lsi�- xl dx
which shows that does not exist in IR. If the improper Riemann integral of a function exists, then it is natural to ask whether the function is, in fact, Lebesgue integrable. In general, this is not the case. However, if the improper Riemann integral of the absolute value of the function exists, then the function is Lebesgue integrable. The details follow.
Let f: [a, oo) IR be Riemantz itztegcable otz every closed subinterval of [a, oo). Thetz f is Lebesgue itUegrahle ifand only ifthe improper Theorem 24.3.
-+
Chapter 4: THE LEBESGUE INTEGRAL·
192
Riemann integral J000 lf(x)l dx exists. Moreover, in this case
f is Lebesgue integrable over [a, oo). Then j+ is also Lebesgue integrable over [a, oo). Let {rn} be a sequence of [a, oo) such that lim rn = oo. For each n, let f,,(x) = j+(x) if x E [a, rn] and fn(X) = 0 if x rn . Then lim f,,(x) = j+(x) and O < f,,(x) < j+(x) hold for all x E [a, oo). Proof. Assume that
>
Moreover, by Theorem 23.6, {/11 } is a sequence of Lebesgue integrable functions such that J f,, dJ... = J�·" j+(x) dx. Thus, by the Lebesgue dominated convergence theorem
I f+ dJ... = 11-00 j f,, dJ... = n-oo 1,." j+(x)dx. lim
lim
0
This shows that J000 j+(x)dx exists and J000 j+(x)dx = J j+ dJ.... Similarly, j000 f-(x)dx exists and j000 f-(x)dx = J f- dJ... . But then, it follows from f = j+ f- and If I = j+ + f- that both improper Riemann integrals J000 f(x) dx and fa00 1 /(x)l dx exist. Moreover, -
{' J(x)dx = J f d
A
and
["' i f(x)l dx = j fl d . l
For the converse, assume that the improper Riemann integral
A
J000l/(x)l dx
n exists. Clearly, lim J;+ l f (x) l dx = j000l/(x)l dx. Define f,,(x) = 1/(x)l if x E [a, a+n] and fn(x) = Oifx a+ for each n . Note thatO < /11 (x) t 1/(x)l holds for all x a. Since j,, is Riemann integrable on [a, a + ], fn is an upper n + function on [a, oo) satisfying J f,, dJ... = J; fn (x)dx = faa+n lf(x)l dx. Thus, >
>
n
n
by Theorem 21.6, 1/1 is an upper function (and hence, Lebesgue integrable) such that f i J I dJ... = foc l f (x) l dx. The Lebesgue integrability of f now follows from a Theorem 22.6 by observing that f is a measurable function, since it is a measurable function on every closed subinterval of [a, oo). The proof of the theorem is now • complete. The next result deals with interchanging the processes of limit and integration.
Let (X, S, J,L) be a measure space, let J be a subinterval of IR, a d let f : X x J � IR be a function such that f(·, t) is a measurable function for each t E J. Assume also that there exists an integrable function g Theorem 24.4. n
193
Section 24: APPLICATIONS OF THE LEBESGUE INTEGRAL
such that for each t E J we have lf(x, 01 < g(x) for almost all x. Iffor some accumulation point to (including possibly ±oo) of J there exists a function h such that lim,_,0 f(x, t) = h(x) exists in JR.for almost all x, then h defines an integrable function, and lim
,_,0
j f(x, t) d�J-(x) = J lim f(x, t) d�J-(.x) = Jh d!J-. (--1' (1)
� JR. satisfies the hypotheses of the theorem. Let {t,} be a sequence of J such that lim t11 = to. Put h,(x) = f(x, t,) for x E X, and note that l h, I < g a.e. holds for each n and h, � ll a. e. By f:X
Proof. Assume that the function
x J
Theorem 22.6, each It, is integrable. Moreover, by the Lebesgue dominated convergence theorem, h defines an integrable function and ,_oc
lim
j f(x, t,) d!J-(x)
11-oo
= lim
J h, d!J- Jh d!J=
holds, from which our conclusion follows.
•
If f : X x (a, b) � JR. is a function and to E function at to is defined by
D,11 (x, t )
(a, b), then its difference quotient
- f(x, t) - f(x, to) t - to
-
-----
for all x E X and all t E (a, b) with t #- to. To make the function D,0 defined everywhere, we let D,0(x, to) = for all x E X. As usual, lim,--..,0 D,0(x, t), whenever the limit exists, is called the partial derivative of f with respect to t at the point (x, t0) and is denoted by ex, t0). That is,
0
¥-
. D,0(x, t) = ltm . f(x, t) - f(x, to) . of to) = ltm -(x, at t - to ,_,0
r-ra
The next result deals with differentiation of a function defined by an integral.
Let (X, S, /.L) be a measure space and let f : X x (a, b) � IT·� be afimction snell that f(·, t) is Lebesgue integrablefor eve1y t E (a, b). Assume that for some to E (a, b) the partial derivative 1, (x, to) exists for almost all x. Suppose also that there exists an integrable function g and a neighborhood V of to such thatfor each t E V we have ID,0(x, t)l < g(x)for almost all E X. Then 1 . r, (·, to) defines an integrable fimction, and Theorem 24.5.
x
Chapter 4: THE LEBESGUE INTEGRAL
194
2.
thefunction F : (a, b) --+ IR defined by F(t) = J f(x, t) dJ.L(X) is differ entiable at to and
, faatt (x, to)dJ.L(x).
F (to) =
r,(x, to) 0 at each point x where the partial derivative does not exist. Then limr r Dr0(X, t) fr (x, to) holds for almost all x. Thus, by Theorem 24.4, �� (·, to) defines an integrable function and Proof. Put
=
-+
o
F(t) - F(to) t - to
=
=
=
J f(x, t)t -- tof(x, to) dJ.L(X) J Dr0(X, t) dJ.L(X) J ir (x, to)dJ.L(X) �
as t -+ to. This shows that F is differentiable at to and that F'(to) (x) holds.
=
frr (x, to) dJ1.
•
There is a criterion for testing the boundedness of the difference quotient function D111(x, t) by an integrable function that is very useful for applications. It requires the existence of a neighborhood V of to satisfying the following two properties:
1.
2.
the partial derivative (x, t) exists for all x and all t e V , and there exists a non-negative integrable function g such that for each t E V, we have I <x, t) l < g(x) for almost all x.
%f
o/i
Indeed, if the previous two conditions hold, then by the mean value theorem, it easily follows that for each t E V we have ID10(x, t)l < g(x) for almost all x. Next, as applications of the last two theorems we shall compute a number of classical improper Riemann integrals.
Theorem 24.6 (Euler).l0
We have J000e-x� dx
=
�.
Proof. The existence of the improper Riemann integral follows from the in
equalities 0 < e-x2 < e-x for x > 1 and J;:x:� e-x dx = e- 1 • Also, since e-x2 for all x , the improper Riemann integral is a Lebesgue integral over [0, oo)
.
>0
10Leonhard Euler ( 1 707-1 783), a great Swiss mathematician. He was one of the most prolific writers throughout the history of science. He is considered (together with Gauss and Riemann) as one of the three greatest mathematicians of modem times.
Section 24: APPLICATIONS OF THE LEBESGUE INTEGRAL
For the computation of the integral, c�nsider the real-valued functions on defined by
195
IR
The derivatives of the previous functions will be determined separately. For f we use the fundamental theorem of calculus and the chain rule. Thus, f'(t) =
,
2e-t·
For g observe that
it 0
,
e-x· dx.
holds for every x E [0, I] and every t E IR in any bounded neighborhood of some fixed point to. The constant M depends, of course, upon the choice of the neighborhood of t0 . By Theorem 24.5, and the fact that the Lebesgue integrals are Riemann integrals, we get
for all t
E IR.
Substituting u = xt (for t :f: 0), we obtain g'(t) =
.,
-2e_1_
11 0
e
-11,
du =
for each t E IR. Thus, J'(t)+ g'(t) = 0 holds for each t holds for all t . In particular, c = f(O) + g(O) =
-2e_1• ,
11 0
e-x- dx .,
E IR, and so, j(t)+ g(t) = c (a constant)
11 1 0
dx
., + x-
- -4
,
and so j(t) + g(t) = :f for each t . Now, observe that for each x and t we have 2 e -t (
I +.r2)
1 + x2
<
1
+ x2 '
---
1
Chapter 4: THE LEBESGUE INTEGRAL
196
and clearly lim 1 -+oo
Thus, by Theorem 24.4, lim1_. 7C
4
-
00
e-12(l+x2) I + x2
g(t) =
0.
=
Consequently,
= lim f(t) + lim g(t) = l-+ l-400 00
2
from which it follows that J000 e-x dx =
Theorem 24.7. For each t 100
e
0.
.;;
(
1 00 0
e-x- dx ,
)2
,
•
.
JR. we have
..fiC e-··,.2 cos(2xt) dx = _ e-1 2 . � 0 2 F(t) = [000 e-.r cos(2xt) dx for all t e JR.
2
Since le-x cos(2xt)l < e-x- holds for all x and t, it follows that the improper Riemann integral F(t) exists and, moreover, is a Lebesgue integral over [0, oo). Now,
Pro.?f· Let
� [e-x2 cos(2xt) ] ar
=
2 < 2x e-x = g(x) l -2xe-x2 sin(2xt)l -
holds for all x > 0 and t . Hence, since the function g is positive on [0, oo) and the improper Riemann integral [000 g(x) dx exists (its value is I), g is Lebesgue integrable over [0, oo). Therefore, by Theorem 24.5 (and the remarks after it), it follows that
ar
2
2
F'(t) = roo � [e-x cos(2xt)] dx = -2 roo xe-x sin(2xt)dx. lo lo
Integrating by parts, we obtain
2
-2 rocxe -x sin(2xt) dx = lo
- 2t e-x2 sin(2x t) 100 0
2
2
roo e -x cos(2xt) dx lo
= -2t roo e -x cos(2xt) dx. · lo
Thus, F'(t) = -2t F(t) holds for all t . Solving the differential equation, we get e-12 , as and so, F(t) = F(t) = F(O)e-12 • By Theorem 24.6, F(O) = claimed. •
./{,
./{
Section 24: APPLICATIONS OF THE LEBESGUE INTEGRAL For the next result, the value of
si�x
at zero will be assumed to be 1 .
Theorem 24.8. If t > 0, then
1oc · 0
holds.
197
smx . t - e-.\ dx X
=
- - arctan t
rr
2
Proof. Fix some to > 0. The cases to > 0 and to
=
0 will be treated separately.
Case I. Assume to > 0. For each fixed t > 0, note that le-x' si�x I :5 e--� t holds for all x > 0. Thus, the improper Riemann integral exists as a Lebesgue integral over [0, oo). Let e-xt si�x dx for t > 0. Then F satisfies the following properties: F(t) =
J000
lim F(t)
1-+00
and
F
I
(t) = -
1
1 + t2
=
0,
for each t > 0.
To see (*), note that if g(x) = 1 for x E [0, 1 ] and g(x) = e-x for x > 1, then g is Lebesgue integrable over [0, oo) for each t > 0 and le-x' si-�x I < g(x) holds for all x > 0 and all t > I. On the other hand, lim,-+oo e-x t si�x = 0 holds for all
x > 0, and so, by Theorem To establish (**)
24.4, lim,-+oc F(t) note first that ..! a [e-x' sinx] x
=
0.
= -e-x' sin x holds for all x > 0 t and that for each fixed a > 0, the inequality le-x' si�x I < e-ax holds for each •
-
•
e-xt sinx dx for all t > a t 2:, a and all x > 0. By Theorem 24.5, F'(t) = (and all a > 0), where the last equality holds since the improper Riemann integral is a Lebesgue integral. Thus, F'(t) = e-xt sinx dx holds for all t > 0. Since from elementary calculus we have
J;
J000
lor0 -·
. x dx e .\ t sm
=
-
e-rt(t sinr + cosr) + -1 + t-., 1 + t2 '
by letting r ----* '>· we get F'(t) = Integrating (**) from t to to yields
1,!
F(t0) - F(t)
and by letting t
----*
oo
f 1 dx ., = tn
= -
for t > 0.
1
+ x-
it follows that F(t0)
=
arctan t - arctan t0,
� - arctan t0•
198
Chapter 4: THE LEBESGUE INTEGRAL
Case II.
Assume to = 0.
s�i x s�i J000 .Y
In this case,
is not Lebesgue integrable over [0, oo) However, the improper
Riemann integral parts yields
[' sinx dx =
ls
_
X
1
X
_
t cosx dx = cos s _ cos t
s ls X2
1 sin x 1 I -- dx < - + - + s t s x
oo sin x x
_
t
S
[
By Theorem 24. 1 • ro
dx = 0.
dx exists. Indeed, if 0 < s < t, then an integration by
cos x
and thus,
J
.
t cos x dx,
ls X2
[
1 dx 2 - = -. s x2 s
n-+oo Jruu+ l sinx
dx exists. In particular note that lim
e-.n si�.;·
'
,
_ .
Now, foreachn define f.1(t � = J�, dx for t > 0, andnotethatthereiation 1 lfn(n)I .:S f� e-.m dx = t-��-��- < * implies lim fn (n) = 0. By Theorem 24.5, we have
_ _, 1 1 11 (t ) _
1u o
e
-.n
. . sm x d.\
.
-
_
· n + cos n ) - 1 e- 111 (t sm , I + t,
and so, lim11-.. 00 J.:(t) = - 1�1 1 holds for all t > 0. Also,
Jn
I + (1 + t)e-1 1 -fl(t)l < 1 + t2 -
holds for each t > 0, and the dominating function for the sequence { f1: } is Lebesgue integrable over [0, oo). Let g11 = /1:Xro.11], and note that {g11} is a sequence of Lebesgue integrable for 1 functions over [0, oo). Also, since lg11 1 < 11.: and lim11_,.00 g11(t) = t > 0, the Lebesgue dominated convergence theorem yields lim
u-+oo
10n
J.:(t)dt = lim 11 -+oo
I
10x
g11(t ) dt = -
1
- �1
10oo
Since f� J.:(t)dt = j.1(n) - J.,(O) and lim J., (n) = Finally, letting n � oo in the identi�
n10 +l X
sinx
-- dx =
we easily get
1
si .,-
11 sinx - dx +
0 X
J000 � dx = I ·
1 "
0,
dt 7r 2 = -- . 2 I +t
we get lim /.1 (0)
11+ 1 sinx -- dx = /11 (0) +
X
u1 +l n
2
-
I·
sin x
-- dx, X
•
Section 24: APPLICATIONS OF THE LEBESGUE INTEGRAL
199
EXERCISES 1.
Show that Ia� _2, .\
0
holds for
2. 3. 4.
0. 1x. 2, . . . . e-1 2 ! .[if
11 =
Show that J0� Show that f(x) = Show that
dx
=
(2n)! e-xl = 2-1111! dx
for each
Let f : [0, oo)
---+
tion. Show that
!au( +
6. 7. 8.
1
> 0.
1
r :.... ll
) e-2x "
if and only if lim
(x
dA.
=
l.
(x)
=0
for each t > O.
Show that the improper Riemann integrals, J000 cos(x2)dx and J000sin(x2)dx (which as the Fresnel
11
integrals) both exist. Also, show that cos(x2) and sin(x2) are not Lebesgue integrable over (0, oo). are known
roo 5'�:x f(·, t
Show that JO Let
'
dx =
!f.
t
(X, S, J-L) be a measure space, T a metric space, and f : X t) is a measurable function for each
that for each
function F :
T
F(t) = is a continuous function. Show that
0,
holds for each t >
0.
fx
T ---+ IR a function.
E T and f(x,
let F(t) = f000
----
X
·
) is a continuous
t(x , t)d J-L(x),
laoo e-x _ e-xt 0
For each t >
x
X. Assume also that there exists an integrable function g such E T we have lf(x, t)l ::::; g(x) for almost all x E X. Show that the ---+ IR, defined by
function for each x E
10.
f
dx = l .
fj + t) x-+oo
1
Assume that
9.
fo 2
-
(0, oo) be a continuous, decreasing, and Lebesgue integrable func
x-+oo ) 1xf(s) x ds = 0 lim
t
·
� is Lebesgue integrable over [ 1 , oo) and that J lim 11 -+00 0
5.
? -
dx = ln t
t���1 dx.
a.
Show that the integral exists as an improper Riemann integral and as a Lebesgue
b.
Show that F has a second-order derivative and that F"(t) + F(t) =
integraL
each t >
0.
1 1 Augustin Jean Fresnel (
f holds for
1788-1827) was French physicist who worked extensively in the field of optics. Using the integrals that bear his name, he was the first to demonstrate the wave conception of light.
a
200
11.
12.
Chapter 4: THE LEBESGUE INTEGRAL
cosx)dx,
Show that the improper Riemann integral, J02 ln(t and that it is also a Lebesgue integral. Also, show that lt
0.
holds for all t > Show that for each t >
0,
the improper Riemann integral
100
Lebesgue integral and that
The Gamma function for
J: x�;����1>
dx
0
exists as a
x d = -(l -e-1). x(l + x2) t>0 = la00x1 - le-x dx.
o
13.
exists for each t >
sin t
--=-
-
x
rr
2
is defined by an integral as follows:
f(t)
a.
b. c. d.
e.
Show that the integral
exists as an improper Riemann integral (and hence, as a Lebesgue integral). Show that r( �) = .fii. Show that f(t + 1 ) = t f(t) holds for all t > and use this conclusion to for n = 1 , 2, . . . . establish f(n + = Show that r is differentiable at every t > and that
1) n!
0,
0 f'(t) = la00 x1- 1 e-·'· lnxdx
holds. Show that r has derivatives of
all
r<" > (t) = holds for 14.
Let
b.
=
1,
fooo
2, . . . and all t >
f: [0, 1] � IR F : [0, 1] �
function
a.
n
order and that x ' - l e-x (lnx)n dx
0.
be a Lebesgue integrable function and define the real-valued IR by
F(t)
=
Jd f(x) sin(xt) d>.(x).
Show that the integral defining· F exists and that F is a unifonnly continuous function. Show that F has derivatives of all orders and that
201
Section 25: APPROXIMATING INTEGRABLE FUNCTIONS and
Fl2 •- 1 l ( t)
for n = 1 , 2, . . . and each c. Show that F = 0 (i.e.,
=
(- I )"
fo
i .r2n - l f(.r) cos(.
t E [0, 1]. F(t) 0 for all t E [0, 1]) if and only if f =
=
0 a.e.
25. APPROXIIVIATING INTEGRABLE FUNCTIONS The following type of approximation problem is commonly encountered analysis. •
m
Given an integrable function f, a family ofF of integrable functions, and E > 0, determine whether there exists a function g in the collection F such that f lf - gl d/1- < E .
By the definition of an integrable function, the following result is immediate.
Theorem 25.1. Let f be an integrable function and E a
step function ¢ such that
J If - ¢1 d/1- < E.
>
0. Then there exists
Let us denote by L the collection of all step functions ¢ for which there exist sets A1, A, E S all of finite measure, and real numbers a 1 , , a, (all depending on ¢) such that ¢ = E�'=1 ai X �; holds. Then L is a function space; see Exercise 1 of Section 21. . . . •
•
•
•
Theorem 25.2. Let f be an integrable function, and let E
>
0.
Then there
exists a function ¢ E L such that J If - ¢ 1 dJ1. < E . Proof. Let F denote the collection of all integrable functions f such that for each E > 0, there exists some function ¢ L such that J If -¢1 d11- < E. It should be clear that F is a vector space such that XA E F holds true for each A E S. Now, a glance at Theorem 22.1 2 guarantees that F consists of all integrable functions .
E
•
The next result deals with the approximation of integrable functions by continu ous functions. Recall that if X is a topological space and f : X -7 lR is a function, then the closure of the set {x E X : f(x) # 0} is called the support of f (denoted by Supp f). If the support of f happens to be a compact set, then f is said to have
compact support.
Theorem 25.3. Let X be a Hausdorff locally compact topological space, and
let f.L be a regular Borel measure on X. Assume that f is an integrable function
202
Chapter 4: THE LEBESGUE INTEGRAL
with respect to the measure space (X, B, Jl). Then given E > 0, there exists a continuousfunction g : X -+ IR with compact support suclz that J If -g I d f..L < E.
Proof. By Theorem 25.2, we can assume without loss of generality that f is
XA
the characteristic function of some Borel set of finite measure. So, assume f = for some Borel set A with J.L(A) < oo. Since Jl is a regular Borel measure, there exist a compact set K such that K c A and Jl(A) - J.L(K) < E (see Lemma 18.5) and an open set V such that A c V and J.L(V) - J.L(A) < E. By Theorem 10.8, there exists a continuous function g : X -+ [0, l ] (and hence, g is Borel measurable) with compact support such that g(x) = l for each x E K and Supp g c V. Clearly, g is integrable and IX A - g l < holds. Therefore,
Xv - XK J IXA - g
l d J.L <
Jcxv -
XK ) dJ.l = J.L( V) - Jl(K) < 2E ,
and the proof of the theorem is finished.
•
The following theorem describes an important property of the Lebesgue in tegrable functions on JR. It is usually referred to as the Riemann-Lebesgue
Lemma.
Theorem 25.4 (Riemann-Lebesgue). Iff : IR -+ IR is Lebesgue integrable,
then
lim
11-HX>
j
t(x) cos(nx) d}..(x)
= 11-+ lim
00
j
t(x) sin(nx) d).. (x)
= 0.
Proof. Note first that the inequality lf (x) cos(nx) l < lf(x) l for all x , com
bined with Theorem 22.6, shows that the function f(x) cos(nx) is a Lebesgue integrable function for each n. Also, by Theorem 25.2, in order to establish Thus, let the theorem, it suffices to consider functions of the form f = In this case, the Lebesgue integrals are Riemann integrals, and f=
f
Xla.h)·
f(x) cos(nx) d)..(x)
=
as n -+ oo. Similarly, Iim
1h "
Xla.h)·
cos(nx) dx
=
l
n
-
l sin(n b)
f f(x) sin{nx) d)..(x) = 0.
-
sin(n a) l <
2
n
-+
0
•
A sequence of integrable functions { fn } is said to converge in the mean to some function f if lim f l fn - fl dJ.l = 0 holds.
Section 25: APPROXIMATING INTEGRABLE FUNCTIONS
203
Let { f,,} be a sequence.ofintegrablefunctions. Iff is an inte grablefunction such that lim Jlf, - !I df..L = 0, then there exists a subsequence {fkn } of {/,} such that [k, � f a.e. holds. Proof. Let E > 0. For each n let En = {x E X : lf,(x) - f (x ) l > E } . and note that each £11 is a measurable set of finite measure (see Theorem 22.5). Now, since EX£, < If,, - !I holds for each n, it follows that EJ..L* (£11) < flfn - !I df..L also holds for each n. Thus, lim J.L.*(£11) = 0, and so, j,, .J!:....:,. f. The conclusion Theorem 25.5.
•
now follows immediately from Theorem 1 9.4.
A sequence of integrable functions that converges in the mean to some function need not converge pointwise to that function. An example of this situation is provided by the sequence { /11 } of Example 1 9.6. EXERCISES 1.
Let f : 1R -4 1R be a Lebesgue integrable function. Show that lim 1 �00
lim j t(:c)sin(xt) dA.(x) = J t(:c)cos(xt)dA.(x) = 1-+00
0.
2. A function f : 0 -+ 1R (where 0 is a nonempty open subset of 1R") is said to be a C00-function if f has continuous partial derivatives of all orders. a. Consider the function p : 1R -+ 1R defined by p(x) = exp[xLt ] if lxl < I and = 0 if lxl > 1. Show that p is a C00-function such that Supp p = [-1, 1]. (Induction and L' Hopital's rule are needed here.) x b. ForE > and a E 1R show thatthe function f(x) = p( �a ) is also aC00-function with Supp f = [a - €, a + €].
p(x)
12
O
3.
Let [a. b] be an interval, E > 0 such that a + € < b - €, and p as in the previous exercise. Define lz : 1R -+ 1R by h (x ) = I: p( '-;x ) dt for all x E 1R. Then show that
a.
Supph C [a - E, b + E],
b. h(x) = c (a constant function) for all :c E [a + E, b - €], c. lz is a C00-function and h(l ) (x) = I: l,�, p('-;X ) dt holds for all .r E 1R, and t/z satisfies 0 < f(x) < for all X E 1R, f(x) = 1 for all d. the C00-function f xE €, b - E], and I IX[a ,l') - /1 dA. < 4€. l
[a +
=
1
4. Let f : 1R -4 1R be an integrable function with respect to the Lebesgue measure, and let € > 0. Show that there exists a C00-function g such that Ill - gl dA. < E. [HINT: Use Theorem 25.2 and the preceding exercise.] 12 Guillaume-Franc;ois-Antoine de L'H6pital
(1661-1704), a French mathematician. He is mainly
remembered for the familiar rule of computing the limit of a fraction whose numerator and denominator tend simultaneously to zero (or to ±oo).
Chapter 4: THE LEBESGUE INTEGRAL
204 5.
The purpose of this exercise is to establish the following general result: Iff : IR" � IR E> ! If for i = I , . . . , n , and put I = Choose € > 0 such that a. Let E - E for each Use Exercise 3 to select for each a C00-function f; : IR � IR such that 0 � f;(x) � 1 for all x, f;(x) = 1 if E E, - E], and Supp - E, E]. Now define II : IR" � IR by h(xl, . , Xn ) = 0�'= 1 f;(x;). Then show that h is a C00-function on IR" and that
is an illtegrablefunction (with respect to tbe Lebesgue measure) aud 0, tben there exists a C00-function g such that gl dA < €. a; < b; Of_1(a;, b;). a; + < b; i. x i [a+ b; f; � [a; b; + .. jlxt -hldA < [D(b; -a; + 2e) - D(b; - a;)] . 2
b. Let f : IR" � IR be Lebesgue integrable, and let E > 0. Then use part (a) to show that there exists a C00-function with compact support such that
g jlf - g j d), < €. a regular Borel measure on IR.n , f : IR" IR a J-L-integrable Junction, and 6. ELet J-L0. beShow that there exists a C00-function g : IRn IR such that fl f - g I dJ-L < E. let E 0. Show that there Let f [a, b] IR be a Lebesgue integrable function, exists a polynomial p such that f l f - pi d). < E, where the integral is considered, of course, over [a, b]. [HINT: Use the Stone-Weierstrass approximation theorem.] � �
>
7.
:
26.
and
�
>
PRODUCT MEASURES AND ITERATED INTEGRALS
Throughout this section, (X, S, J..L) and (Y, :E , v) will be two fixed measure spaces. The product semiring S ® :E of subsets of X x Y is defined by S ® :E
= {A x B : A E S
and
B
E :E}.
The members of S ® :E are called rectangles. The above collection S ® :E is indeed a semiring of subsets of X follows immediately from the identities
x Y.
This
(A x B) n (A1 x B1 ) = (A n A1 ) x (B n B1 ), and 2. (A x B) \ (A1 x Bd = [(A \ A1 > x B] u [(A n A J ) x (B \ B1 )], and the fact that A \ A 1 and B \ B 1 can be written as finite disjoint unions 1.
of members of S and :E, respectively. The product se miring was discussed in s ome detail in Section 20. Now, define the set function J..L x v : S ® :E � [0, oo] by J..L x v(A x B) = J..L(A) · v(B) for each A x B E S ® :E (keep in mind that 0 oo = 0). This set function is a measure on the product semiring S ® :E, called the product measure of J..L and v. The details follow. ·
·
Section 26: PRODUCT MEASURES AND ITERATED INTEGRALS
Theorem 26.1.
The setfimction J.t x J.t x v(A
x B)
v.: S
® :E
= J.t(A)
--*
[0,
205
oo] defined by
· v(B)
for each A x B E S ® :E is a measure. (This measure can be thought of as the generalization of the familiarformula "Length x Width" for computing the area ofa rectangle in elementary geometry!) Proof. Clearly, J.t x v(0) 0. For the a -additivity of J.t x v, let A x B E S ® :E and let {A, x B, } be a sequence of mutually disjoint sets of S ® :E such that A
X B
= u:l A,
X
=
B,. It must be established that
J.t(A) · v(B)
= L J.t(A,) 0:::
II= I
· v ( B,) .
Obviously, (*) holds true if either A or B has measure zero. Thus, we can assume that J.t(A) > 0 and v(B) > 0. Since = 2::1 we see that
XA,x B, •
XAx B
00
XA(x) XB(Y) = LXA,(x) · XB,(y) ·
n=l
holds true for all x and y. For each fixed y E B let K = {i E IN : y E B;} and note that L;e K for each x E X. Since the collection {A; : i E K } LieK J.t(A; ). Therefore, is pairwise disjoint (why?), the latter implies J.t(A)
XA(x )
=
XA; (x )
=
= L J.t(A; )XB/Y) = LJ.t(A ) XB,(Y) J.t(A,) = n. J.t(A,)
j.t(A) . XB(Y)
00
iE K
,
.
(**)
11=1
0 does not alter the sum in (*) or holds for all y E Y. Since a term with (**), we can assume that > 0 for all Now, if both A and B have finite measures, then by using Theorem 22.9 and integrating (**) term by term, we see that (*) is valid. On the other hand, if either A or B has infinite measure, then I::1J.t(A,) · v ( B,) = oo must hold. Indeed, if the last sum is finite, then, by Theorem 22.9, J.t(A)x8(y) defines an integrable function, which is impossible. Thus, in this case (*) holds with both sides infinite, and the proof is finished. • The next few results will unravel the basic properties of the product measure J.t x v . As usual (J.t x v)* denotes the outer measure generated by the measure space (X x Y, S ® :E, J.t x v) on X x Y.
Chapter 4: THE LEBESGUE INTEGRAL
206
Theorem 26.2.
tben
(Jl
x
If
A and B Y are measurable sets offmite measure, v)*(A B) = p.* v*(A B) p.*(A) v*(B). :E {A1 1 B1 1 } A B u:)AII BlAl. B 26.1, p.* v* cX
c
x
x
x
·
=
Proof. Clearly, S ® C Ap. ® Av holds. Now, let x be a sequence X c of s ® :E such that X Since, by Theorem X x E Ap. ® Av. it follows from is a measure on the semiring Ap. ® A v and Theorem 1 3.8 that 00
X
00
p.* v*(A B) 1L=1 p.* v*(A11 B11) 1L=1 v(A11 B11), <
x
x
and so,
x
x
x
=
x
< (p.
J1. x
J.l* v*(A B) v)*(A B). E 0 {An} L:: {B,}L::,v(B,J A v*U: p(A,) p.*(A)+E 1 1 A11, B U:1 B11, (B) +E. A B u:, u�=I AII Bm (Jl v)*(A B) n=lL m=lLJl v(A, B11 ) Ll =l m=LJ1(A,) v(B111) [�JL(A., )] · [� v(B,)] lJL*(A)+ <] · [v•(B) + <]
On the other hand, if c :E, with C and < so X
X
is given, choose two sequences c such that Then, X c X
>
X
00
:Xl
x
x
<
X
=
0.
That is,
x
=
(p. v)*(A B) p.*(A) v*(B) (p. v)*(A B) v*(A B) x
Therefore,
x
holds, and
00
·
I
<
=
for ali E >
00
<
c S and
x
<
=
p.* x
·
x
p.* x
v*(A B). x
holds, as required.
•
It is expected that the members of Ap. ® Av are J1. x v-measurable subsets of X x Y; that is, Ap. ® Av c A11-xv holds. The next theorem shows that this is actually the case.
If is a Jl.-measurable subset of X and a v-measurab/e subset of Y, then x is a J1. x v-measurable subset of X x Y. Theorem 26.3.
Proof. Assume
f.L(C)
·
v(D)
A A B A B
B
E Av. and fix C x D E S® :E with p. x E All-, < oo. In order to establish the J1. x v-measurability of
v(CA B,
x D) = x by
Section 26: PRODUCT MEASURES AND ITERATED INTEGRALS
207
Theorem 15.2, it suffices to show that
(J-L x v)*((C x D) n (A x B)) + (J-L x v)*((C
x D) n (A x B/) < J-L x v(C x D).
If f.1. x v(C x D) = 0, then the previous inequality is obvious (both sides are zero). So, we can assume JJ-(C) < oo and v(D) < oo. Clearly,
(C (C
x
X
D) n (A D) n (A
x
B) X B)c
= (C n A)
x
= [(C n A e) u [(C
(D n B), and
X
n Ae)
(D n B)] u [(C n A)
X
(D n Be)]
X
(D n Be)]
hold with every member of the previous union having finite measure. Now, the subadditivity of (J.L x v)* , combined with Theorem 26.2, yields
(J-L
X
v)*((C
X
D) n (A
X
B)) + (J-L
X
v)*((C
X
D) n (A
< J.L*(C n A) · v*(D n B ) + J.L*(C n Ae) v*(D n B)
X
B)e)
·
+ J.L*(C n A) · v *(D n Be) + J-L*(C n A e) · v*(D n Be)
= [J.L*(C n A) + J.L*(C n A e)] · [v*(D n B ) + v*(D n Be)] = J.L(C) · v(D)
=
J.L x v(C
x
D),
as required.
•
In general, it is not true that the measure JJ-* x v * is the only extension of J.L x v from S ® :E to a measure on AJ.L ® A v . However, if both (X, S, J.L) and (Y, :E, v) are a-finite measure spaces, then (X x Y, S ® I;, J-L x v) is likewise a a-finite measure space, and therefore, by Theorem 15.10, f.l* x v * is the only extension of J.L x v to a measure on the semiring AJ.L ® Av. Moreover, in view of AJ.L ® A v C AJ.Lxv and the fact that (J.L x v)* is a measure on AJ.Lxv, it follows in this case that (J.L x v)* = J-L* x v * holds on AJ.L ® A v . We now tum our attention to measurability properties of arbitrary subsets of X x Y. Let us recall a few definitions from Section 20. If A is a subset of X x Y Ax of A is the subset of Y defined by and x E X, then the
x-section
Ax = {y E Y : (x, Similarly, i f y E Y , then
y) E A } .
the y-section A Y o f A is the subset of X defined by AY = {x E X : (x, y) E A } .
The geometrical meanings of the x- and y-sections are shown i n Figure 4. 1 . Regarding sections of sets, we have the following identities.
Chapter 4: THE LEBESGUE INTEGRAL
208 )'
X
X
FIGURE 4.1.
The proofs of the previous identities are straightforward, and they are left as an exercise for the reader. The next theorem demonstrates the relationship between the J.L x v-measurable subsets of X x Y and the measurable subsets of X and Y, and it is a key result for this section. Recall that an extended real-valued function f that is undefined on a set of measure zero is said to define an integrable function if there exists an integrable function g such that f = g almost everywhere. That is, if arbitrary values are assigned to f on the points where it is undefined or attains an infinite value, then f becomes an integrable function. (The value of the integral does not depend, of course, upon the choices of these values.)
Let E be a J.L x v-measurable subset of X x Y satisfying (J.L x v)*(E) < oo. Then for J.,L-almost all x the set Ex is a v-measurable subset of Y, and the function x 1--7 v*(Ex) defines an integrable function on X such that
Theorem 26.4.
(J.L x
v)*(E) =
fx v*(Ex) dJ.,L(X).
Similarly,for v-almost all y, the set EY is a J.,L-measurable subset of X , and the function y J.L *(E Y) defines an integrable function on Y such that 1--7
(J.L x v) *(E) =
l J.L*(EY) dv(y).
209
Section 26: PRODUCT MEASURES AND ITERATED INTEGRALS
By the symmetry of the situatiQn, it suffices to establish the first for mula. The proof goes by steps.
Proof.
= A B E S :E. = B x xE EX,A =
Step I. Assume
x
E
if Clearly, Ex subset of Y for each
® and Ex and
(j) if x fj; A. Thus, Ex is a v-measurable
holds for all x E X. Since (!J- x v)"'(E) = (!J- x v)(A x B) = !J-(A) · v(B) < oo, two possibilities arise:
n
a.
Both A a d B have finite measure.
In this case, (*) shows that x 1--+ v*(Ex) is an integrable function (actually, it is a step function) satisfying
b.
l
v*(Ex) dfJ-(X) =
I
v(B)XA df-L = fJ-(A)
·
v(B) = (/-L x v)*(E).
Either A or B has infinite measure.
In this case, the other set must have measure zero, and so, (*) guarantees Thus, 1--+ v*(Ex) defines the zero function, v(E.. J = 0 for !J--almost all and hence
l
x.
x
== 0
v*(E.J dfJ-(X)
(/-L x v)*(E).
Step II. Assume that E is a a-set of S ® :E. Choose a disjoint sequence {E11} of S ® :E such that E U: 1 En. In view of Ex U:1 (E11).,. and the preceding step, it follows that Ex is a measurable subset of Y for each x E X. v *(Ex) and :L�'= t v((E;)x) for each x E X and all Now, define f(x)
=
n . By Step I, each
= J,,
If,, = t;II jr df-L
=
J,,(x) =
defines an integrable function and
X
= t;f-L II
v((E;)_,J d!J-(X)
X
v(E;)
t
(/-L
=
X
v)*(E) < oo.
Since {(E,).� } is a disjoint sequence of :E, we have v*(Ex) I:: 1 v((En)x). and so, fn(x) t f(x) holds for each x E X. Thus, by Levi's theorem (Theorem 22.8), f d�fines an integrabl� _f�nctio�. and
{ v*(Ex) dfJ-(X) = I f df-L =
lx
lim
ll-+oo
If,,
df-L =
� fJ8
X
v(E;)
=
(/-L
X
v)*(E).
210
Chapter 4: THE LEBESGUE INTEGRAL
Step l/1. Assume that E is a countable intersection of a-sets offmite measure. Choose a sequence {En} of -sets such that E n: l En, (Jl X \) )*(E I ) < 00, and En+ l C E11 for all n.
=
a
=
For eachn, letg11(X) = O if v*((En)...- ) = oo and g11(X) v*((En)x ifv*((E11)x < oo. By Step II, each g11 is an integrable function over X such that g11 d J1 = (Jl X v)*(E,) holds. In view of Ex = n:.l (En)x. it follows that E_,. is a \) measurable set for each x E X . Also, since v*((E1)x) < oo holds for jl-almost all x, it follows from Theorem 15.4 that gn(x) v*((E,Jx) {. v*(Ex) holds for J.L-almost all x. Thus, x H- v*(Ex) defines an integrable function and
J
=
}r v*(Ex) dJl{X) = nlim -oo X
! gil dJ1 = lim (Jl n-oo
X
v)*(En)
= (Jl
X
v)*(E),
where the last equality holds again by virtue of Theorem 1 5.4. Step N: Assume that E is a null set, i.e., (Jl x v)*(E) = 0. Arguing as in the proof of Theorem 15. 1 1 , we see that U"iere exists a measurable set G, which is a countable intersection of a-sets of finite measure, such that E c G and (Jl x v)*(G) = 0. By Step Ill, fx v*(Gx)dJ1(x) = (Jl x v)*(G) = 0, and so, by Theorem 22.7( 1), v*(Gx) 0 holds for jl-almost all x . In view of E.,. c G.,. for all x , we must have v*(Ex) 0 for jl-almost all x. Therefore, Ex is v-measurable for J.L-almost all x and x H- v*(Ex) defines the zero function. Thus,
= =
i v*(Ex) dJl(X) =
0 = (Jl
x
v)*(E).
e a Choose a v-measurable set F that is a countable intersection of -sets all of finite measure such that E F and (Jl v)*(F) = (Jl v)*(E). Let G = F \ E. Then G is a null set, and thus, by Step v*(G ) = 0 holds for jl-almost all x. Therefore, Ex is v-measurable and v*(Ex) = v*(Fx) holds for Jl-almost all x. By x H- v*(Fx) defines an integrable function, and so, x Step v*(Ex) defines Step V. The g ne l case. J1 x
r
c
x
IV,
x
ITI,
H-
an integrable function and (J.L x
v)*(E)
= (Jl
x
v)*(F)
= i v*(Fx) dJ.L(X) = i v*(Ex) djl(X)
holds. The proof of the theorem is now complete.
�
x
a
•
Now, let f : X x Y IR be a function. Then for each fixed x E X, the IR defined by f-:(Y) = f (x , symbol f-: will denote the function fx : Y for all y E Y. Similarly, for each y E Y the notation fY denotes the function IR defined by fY(x) = f(x, y) for all x E X. JY : X
�
�
y)
Section 26: PRODUCT MEASURES AND ITERATED INTEGRALS
Definition 26.5. Let f : X x Y
-+ JR. be afunction.
211
Then the iterated integral
JJf df.J.dv is said to exist l/ 1 . fY is an integrable function avec X for v-almost ally, at1d 2. the function g (y)
= f f>' dJ.L = l f(x, y) djJ.(X)
defines all imegl·able function over Y. The value of the itemted integml JJf df.J.d v is computed by starting with the innermost integration and by continuing with the second as follows:
holding y fixed while tile integmtioll over X is pelformed. The meaning of the iterated integral
JJf dvdf.J.
is analogous. That is,
If E is a J.L x v-measurable subset of X x Y with (J.L x v )*(£) < oo, then it is readily seen from Theorem 26.4 that both iterated integrals JJX E df.J.dv and JJ XE dvdf.J. exist, and that
ff
X£
dj.LdV
= ff
XE
dvdjJ.
=f
XE
d(J.L
X
v)
= (f.J.
X
v)*(£).
Since every J.L x v-step function is a linear combination of characteristic functions of fJ. x v-measurable sets of finite measure, it follows from the prior observation that if ¢ is a fJ. x v-step function, then both iterated integrals JJ¢ dJJ.d v and JJ¢ d vdfJ. exist and, moreover,
JJ¢ dJ.Ldv = JJ¢ dvdf.J. = J¢ d(J.L x
v).
The previous identities regarding iterated integrals are special cases of a more general result known as Fubini 's 1 3 theorem. 13Guido Fubini (
1879-1943),
an Italian mathematician. He made important contributions in analysis,
geometry, and mathematical physics.
212
Chapter 4: THE LEBESGUE INTEGRAL
Let f : X x Y -+ Then both iterated integrals exist and Theorem 26.6 (Fubini).
Jf
d(p, x
v)
IR be a
p, x
v-integrablefunction.
= JJ f dp,dv = JJf dvdp,.
Proof. Without loss of generality we can assume that f (x, y) > 0 holds for all
x and y.
Choose a sequence {t/>n } of step functions such thatO < holds for all x and y. Thus,
t/>11(x, y) t f(x , y)
By Theorem 26.4, for each n the function
8n(X) = J(t/>n).r dv = i
t/>11 (X , y) dv(y)
defines an integrable function over X; and clearly, 11 (x ) t holds for p,-almost all x. But then, by Levi's Theorem 22.8, there exists a p,-integrable function g : X -+ IR such that g11 (x ) t g(x) p,-a.e. holds. That is, there exists a p,-null subset A of X such that dv t g(x) < oo holds for all x ¢. A. Since (t/>n).\. t fx holds for each x, it follows that fr is v-integrable for all x ¢. A and
g
f{¢n}r
g"(x) = J(t/>ll)x dv = i t/>"(x,
y) dv(y) t
i fx dv
holds for all x ¢. A. Now, (*) combined with Theorem 21.6 implies that the function defines an integrable function such that
Similarly, J f d(p,
X
v)
= JJfdp,dv,
x � frfx dv
and the proof is complete.
•
The existence of the iterated integrals is by no means enough to ensure that the Y [0, 1], function is integrable over the product space. For instance, let X 2 2 2 I. (the Lebesgue measure), and f(x, y ) p, v (x - y )j(x + y2)2 if
= =
=
= =
Section 26: PRODUCT MEASURES AND ITERATED INTEGRALS
(x, y) =I= (0, 0) and f (0, 0)
=
213
0. Then, it is �asy to see that
JJ f djJ.dv = : -
and
JJ f dvdjJ. = :
.
Fubini 's theorem shows, of course, that f is not integrable over [0, I ] x [0, I]. However, there is a converse to Fubini 's theorem according to which the exis tence of one of the iterated integrals is sufficient for the integrability of the function over the product space. The theorem is known as Tonelli's theorem, and this result is frequently used in applications.
14
Theorem 26.7 (Tonelli). Let (X, S, J.l.) and (Y, :E , v) be two a-finite measure
spaces, and let f : X x Y -+ IR be a J.1. x v-measurable function. If one of the iterated integrals JJ Ill djJ.dv or JJ If I dvdJJ. exists, then the function f is JJ. x v -integrable--and hence, the other iterated integral exists and
J f d(J.l. x v) = JJ f djJ.dv = JJ f dvdjJ..
Proof. We can assume without loss of generality that f(x, y) > 0 holds for
all x and y. S ince (X, S, J.l.) and (Y, :E, v) are a-finite measure spaces, it is easy to see that the product measure space is also a a-finite measure space. Choose a sequence {A,} of JJ. x v-measurable sets such that (JJ. x v) * (A,) < oo for each n and A, t X x Y. By Theorem 17.7, there exists a sequence {1/t,} of J.1. x v simple functions such that 0 < 1/t,(x, y) t f(x, y) holds for all x and y. Let ¢, 1/t, · XA" for each n. Then {¢,} is a sequence of J.1. x v-step functions such that 0 < cp,(x, y) t f(x, y) holds for all x and y. Now, assume that JJ f djJ.dv exists. This means that for v-almost all y, the integral f(x, y) djJ.(x) exists and defines a v-integrable function. From ¢, (x. y) t f(x, y) it follows that ¢"(x, y) djJ.(x) t J f(x, y) dJJ.(x) holds for v-almost all y. But then, by applying the Lebesgue dominated convergence theo rem, we get
=
J
J
J>,. d(p. v) = i [i
< oo.
v-upper function and that This shows that is a holds. The rest of the proof now follows immediately from Fubini 's theorem.
The Fubini and Tonelli theorems are usually referred to as "the method of computing a double integral by changing the order of integration." 14 Leonida Tonelli ( 1885-1946), an IIalian maihemaiician. He contribuled Io measure Iheory, Ihe Iheory of iniegraiion. and Io calculus of variaiions.
Chapter 4: THE LEBESGUE INTEGRAL
214
In general, it is a difficult problem to determine whether or not a given function IR is J.1 x v-measurable. However, in a number of applications the f:X x Y J.1 x v-measurability off can be established from topological considerations. For instance, if X Y = IR and J.1 = v =the Lebesgue measure, then it should be the product measure J.1 x v on IR2 is precisely the Lebesgue measure clear that 2 on IR . Therefore, every continuous real-valued function on IR.2 is necessarily J.1 x v-measurable. For more about the joint measurability of functions, see also Section 20.
�
=
EXERCISES 1.
Let (X, S, J.L) and (Y, �. v) be two measure spaces, and let A x B E
AJJ. ® Av.
a. Show that J.L*(A) · v*(B) ::: (J.L x v)*(A x B). b. Show that if J.L*(A) · v*(B) ::p 0, then (J.L x v)*(A x B) = J.L*(A) · v*(B). c. Give an example for which (J.L x v)*(A x B) ::p f.L*(A) · v*(B).
2. Let (X,.S, J.L) and (Y, �. v) be two cr-finite measure spaces. Then show that
(J.L x v)*(A
x
B) = J.L*(A) · v*(B)
holds for each A x B E AJJ. ® Av. 3. Let (X, S, J.L) and (Y, �. v) be two measure spaces. Assume that A and B are subsets of X and Y, respectively, such that 0 < J.L*(A) < oo, and 0 < t.•*(B) < oo. Then show that A x B is f.L x 11-measurable if and only if both A and B are measurable in their corresponding spaces. Is the above conclusion true if either A or B has measure zero? 4. Let (X, S, J.L) and (Y, �. v) be two cr-finite measure spaces, and let f : X x Y � IR be a Jl x v-measurable function. Show that for J.L-almost all x , the function f'" is a v-measurable function. Similarly, show that for v-almost all y, the function JY is J.L-measurable. 5. Show that if f(x, y) = (x2 - y2) j(x 2 + y2)2 , with /(0, 0) = 0, then
6. 7.
fo' [ fo'
]
/(x , y) dx dy = -
and
fo' [ fo 1
f(x ,
}
y) d)
x=
:
Let X = Y = IN, S = � = the collection of all subsets of IN, and f.L = v = the counting measure. Give an interpretation of Fubini 's theorem in this case. Establish the following result, known as Cavalieri's 1 5 principle. Let (X, S, J.L) and (Y, �. v) be two measure spaces, and let E and F be two J.L x v-measurable subsets of X x Y of finite measure. If v*(Ex) = v*(F.t) holds for J.L-almost all x. then (J.L
8.
:
x
v)*(E) = (J.L
x
v)*(F).
For this exercise, 'A denotes the Lebesgue measure on IR. Let (X, S, J.L) be a cr-finite measure space, and let f : X � IR be a measurable function such that j(x) > 0
15Bonaventura Cavalieri ( 1 589-1647). an Italian mathematician. He was a geometer who worked on problems regarding volumes of solids and wrote several monographs on this subject.
Section 26: PRODUCT MEASURES AND ITERATED INTEGRALS holds for all
E X. Then show that
x IR: 0 < y f(x)}, J1. x x IR. = ((x, y) x IR: 0 < f(x)} x x IR (J.l x x = ((x, f(x)) :x } J1. x x IR. x f dJ.l x =0 Y -+ IR g -+ IR f x Y -+ IR f(x, = )h f J1. x j fd(J.l x v) = (LgdJJ.) ([ h dv) .
a.
b.
c.
d. e.
9.
x
215
f
The set A = ((.Y, y) E X � called the ordinate set of , is a A.-measurable subset of X The set B EX is a J1. A.-measurable subset of � y X and A.)*(A) = (JJ. A.)*(B) holds. E X , is a The graph of f, i.e., the set G A.-measurable subset of X holds. If f is JJ.-integrable, then (JJ. A.)*(A) = J If f is JJ.-integrable, then (JJ. A.)*(G) holds.
be a JJ.-integrable function, and let Jz : be a v-integrable Let : X function. Define : X by y) g(x (y) for each x and y. Show that is v-integrable and that ·
10.
1r
rl oo (1r e-.\y. . x dx ) dy = d o --;- x 0 < E < r. E -+ o+ r -+ 00 loco --dx = -. f(x, ye-( l +x2>.v2
Use Tonelli's theorem to verify that sinx
sm
f
f
holds for each By letting steps) give another proof of the fonnula
11.
Show that if
0
y) =
and
sinx X
(and justifying your
rr
2
for each x and y, then
Use the previous equality to give an alternate proof of the fonnula
roo .J7i oJ e dx = --z · -x
12.
Show that
Iaoo (fo r > 0.
r
holds for all
e-xy
2
2
.
dx) dy = Ia (Ia e-xy2 dy) dx r -+ oo roo :c = � d 2 . lo ../X f000 cJ-jr d:c = J2ii ,
sinx
By letting
oo
sin x
show that
sinx
In a similar manner show that 13.
/2.
Using the conclusions of the preceding exercise (and an appropriate change of vari able), show that the values of the Fresnel integrals (see Exercise 6 of Section 24) are
� loco sm(x-)dx = fooo cos(x )dx = --. 0
.
?
0
2
4
Chapter 4: THE LEBESGUE INTEGRAL
216
J..L = the Lebesgue measure on [0, 1], and v = the counting measure on [0, 1]. Consider the "diagonal" fl. = {{x, x) : x E X} of X x Y. Show that:
14. Let X = Y = [0, 1],
a. b. c.
fl. is a J..L x v-measurable subset of X x Y , and hence, Xt. is a non-negative J..L x v measurable function. Both iterated integrals JJ Xt. dJ..Ldv and JJ Xt. dvdJ..L exist. The function x t. is not J..L x v-integrable. Why doesn't this contradict Tonelli's theorem?
15. Let f : IR � IR be Borel measurable. Then show that the functions f(x + y) and f(x - y) are both >.. x >..-measurable. [HINT: Consider first f = x v for some open set V .]
C HAPTER
5
__ _ _ _ _ _ _ _ _ _ _ _ _
NORMED SPACES AND Lp-SPACES The algebraic theory of vector spaces has been an integral part of modem mathe matics for some time. In analysis, one studies vector spaces from the topological point of view, taking into consideration the already existing algebraic structure. The most fruitful study comes when one attaches to every vector a real number, called the
norm of the vector. The norm can be thought of as a generalization of
the concept of length. A normed space that is complete (in the metric induced by its norm) is called a Banach 1 space.
A variety of diverse problems from different branches of mathematics (and science in general) can be translated to the framework of Banach space theory and solved by applying its powerlul techniques. For this reason the theory of Banach spaces is in the frontier of modem mathematical researc h. This chapter presents a brief introduction to the theory of normed and Banach spaces. After developing the basic properties of normed spaces, the three comer stones of functional analysis are proved-the principle of uniform boundedness, the open mapping theorem, and the Hahn-Banach theorem.2 Then a section is devoted to the study of Banach lattices; that is, Banach spaces whose norms are compatible with the lattice structure of the spaces. As we shall see, many Banach
lattices are actually old friends. Finally, the classical, Lp-spaces are investigated,
and the theory of integration is placed in its appropriate perspective.
27. NORMED SPACES AND BANACH SPACES A real-valued function
11·11
defined on a vector space
satisfies the following three properties:
X
is called a
norm
if it
1 Stefan Banach ( 1892-1945), a prominent Polish mathematician. He is the founder of the contem
porary field of functional analysis. 2Hans Hahn (1879-1934), an Austrian mathematician and philosopher. He contributed decisively to functional analysis, general topology, and the foundations of mathematics.
217
Chapter 5: NORMED SPACES AND Lp·SPACES
218 I.
2. 3.
=
0 if and only if x ll x ll > 0 for each x e X , and ll x ll Ia ! · ll x ll for all x E X and a e JR.; ll ax ll llx + Yll < ll x ll + II YII for all x, y E X.
=
=
0;
Property (3) is referred to as the triangle inequality, and it is equivalent to the inequality ll x - Yll .:S ll x - zll + liz - Yll
for all x, y, and z in X . A vector space X equipped with a norm 11·11 is called a normed ''ector space, or simply a normed space. To avoid trivialities, the vector spaces will be tacitly as sumed to be different from {0}. Also, they will be assumed to be real vector spaces. On a normed space X, a metric is defined in terms of the norm 11·11 via the function d (x, y) ll x - y 11. From the properties of the norm, it is a routine matter to verify that d ( ·, ·) is indeed a metric on X . We shaH caH this metric on X t.'ie metric induced by the norm. A sequence {x,} in X is said to converge in norm to x if lim llx - x, II 0; that is, if {x,} converges to x with respect to the metric induced by the norm. In view of the triangle inequality,
=
=
l ll x II - ll y Il l
<
ll x - Yll
holds for all x and y in X. This readily implies that the norm considered as a function x r+ llx II for X into 1R is uniformly continuous. Also, by the triangle inequality it is easy to prove, but important to observe, that if x, x and y, -+ y in X and a, -+ a in JR., then x, + y, -+ x + y and a,x, -+ ax hold. A subset A of a normed space is said to be norm bounded (or simply bounded) if there exists some M > 0 such that ll x II < M holds for each x e A . Every Cauchy sequence {x,} of a nonned space is bounded. Indeed, to see this choose k such that ll x, - Xm II < I for all n, m > k, and let M max { I + ll x; ll : I < i < k}. Then the inequality ll x1111 < I + ll xk II for n > k implies ll x, II < M for all n. A nonned space X that is complete with respect to the metric induced by its norm is called a Banach space. In other words, X is a Banach space if for every Cauchy sequence {x,} of X there exists an element x e X such that lim ll x, - x II 0. Thus, Banach spaces special examples of complete metric spaces. Some examples of Banach spaces presented next.
�
=
are
are
=
=
o=�l= l Xf)� for each X = Example 27.1. The vector space R" with the norm llxll (x t , . . . , x11) E R" is a Banach space. This nonn is called the Euclidean norm, and it gives
the Euclidean metric. Example 27.2.
=
•
Let X be a nonempty set, and let B(X) be the vector space of all bounded real-valued functions defined on X. Then llfii"Xl sup{lf(x)l: x E X} for each f E B(X)
219
Section 27: NORMED SPACES AND BANACH SPACES
defines a norm on B (X), called the sup norm .. A glance at Example 6.12 guarantees that • the vector space B(X) with the sup norm is a Banach space.
. . .) such Example 27 3 Let e 1 denote the collection of all real sequences x = (..q , that :L� 1 lx, I < oo. It is easy to see that e 1 is a vector space under the algebraic operations .
x2 ,
.
x + y = (..q
+ Yl ,
x2 + Y2· . . . )
and
ax = (c:u 1 , ax2, . . ) . .
Moreover, if for every X E e I we define llx Il l = L� I lx, I' then 11·11 I is a norm on e I . The verification of the norm properties are left for the reader. We show next that f. 1 is actually a Banach space. To this end, let {x, J be a Cauchy sequence of e 1 . That is, for every E > 0 there exists some k such that llx, Xm Il l < E holds for all n, m > k. It follows that there exists some M > 0 such that Ux, II 1 � M for each n. Let x, = x2, . . . ) for each n. The relation
-
Cxl' ,
j x}' - xj" I � L jxj' - xj'j = llx, - x,ll 1 i=l 00
<E
for 1z, m > k implies that for each fixed i, the sequence of real numbers {xj' } is a Cauchy sequence. Let x; = lim,_.., 00 xj' for each i. Since for all n > k and each p we have p
p
p
p
i=l
i=l
i=l
i=l
L lx; I < L jx; - x}' I + L jxj' I � L lx; -xj' I + llx, II 1
(
� 1 1-+00 1=1 � lx;n - xj' I
= 1 lim
)
+
llx, II 1
< E + M < oo ,
= (.q ' X2, . . . ) E e I . Also, since :Lf= I lx;' - xj' l � llx, - Xm II I < E for all n , m > k, we have 'Lf=1 lx; - xjl = limm-+ oo C'Lf=l l xj - xjl) < E for all p and all n > k. Hence, llx - x11 11 1 < E holds for all n > k, so that {x11} converges to x in e 1 . That it follows that X
is, e I is a Banach space.
•
Example 27.4. Consider an interval [a, b] and a natural number k. Let Ck [a, b] denote the collection of all real-valued functions defined on [a, b] which have a continuous kth order derivative on [a, b] (with right and left derivatives at the endpoints). Clearly, c k [a, b] with the pointwise algebraic operations is a vector space. Moreover, if for each f E C k [a. b] we let
11!11 = llflloo + llf ' llco + · · · + llf(k)lloo. k then 11·11 defines a norm under which c [a, b] is a Banach space. k It is straightforward to verify that 11·11 is indeed a norm. To see that C [a, b] is complete k under 11·11, let {j;,} be a Cauchy sequence of C [a, b]. Then, it is easy to see that there exist continuous functions go, g 1 , . . . , 8k such that for each i = 0, l , . . . , k the sequence
Chapter 5: NORMED SPACES AND Lp·SPACES
220
of functions
{/�0} converges uniformly to
holds for eachx E from (*) that
[a,
gi.
Now, if 1 ::; i ::; k, then
b]. Thus, by the uniform convergence of the sequence {f,�0} it follows
gi- l (x) =gi-l (a) + 1xg;(t) dt holds for each x E (for 1
::;
=
[a, g;_ 1 (x) =
b]. By the fundamental theorem of calculus,
gi-l
is differentiable
g;(x) for each E b]. In particular, note i < k), and 1 , . . . , k. Therefore, go E b] and lim ll fn o ll = 0 holds, so that
for i is a Banach space.
x [a,
Ck[a,
-
g
that g; = gg> Ck[a,
b]
•
Two norms ll · ll t and 11·11 on a vector space X are said to be equivalent if there 2 exist two constants K > 0 and M > 0 such that
holds for each x E X. The reader should stop and verify that two norms are equivalent if and only if they generate the same open sets. In particular, note that if two norms on a vector space X are equivalent, then X is a Banach space with respect to one of them if and only if X is a Banach space with respect to the other. Also, observe that if II· II 1 is equivalent to II ·11 2 and II · 112 is equivalent to II · 113. then ll · l l t is equivalent to II· 113· In a normed vector space, all open balls "look alike." More precisely, any two open balls are homeomorphic. Indeed, if B(a, r) is an arbitrary open ball, then it is easy to see that the function x � a + rx is a homeomorphism from B(O, 1) onto B(a, r). For this reason, the open ball B(O, 1) = {x : llxll < 1 } plays an important role in the study of normed spaces and is called the open unit ball of the space ( and likewise {x: ll x II < 1 } is called the closed unit ball of the space).
Example 27.5.
It is instructive to consider various norms on 1R2. For v
define llvi!J = lxl llvll 2
+ IYI.
= ( ·2 + y2)! , x
llvlloo = max{lxl, jyl},
=
(x, y) E 1R2
Section 27: NORMED SPACES AND BANACH SPACES
l v II
I
..
221
=
FIGURE 5.1. The Unit Balls of Various Norms
where a and b are two, fixed, posi tive real numbers. The reader should stop and verify that the previous functions are norms and that they are, in fact, all equivalent. The geometric shape of the closed unit ball for each norm is illustrated in Figure 5.1. •
The fact that all norms in the preceding example were equivalent is not acci dental.
In afinite dimensional vector space all norms are equivaient. Proof. JR". l ·l z l x l 2 = (Xf · · x;) � . l ·llxz�= (x 1, x,) = x1e1 · x,e11 e11
Theorem 27.6.
Without loss of generality, we can assume that the finite dimensional denote its Euclidean norm; that is, + · + space is Let Also, let 11·1 1 be another norm on lR2 . It suffices to establish that 11·11 is equivalent to If + ··+ • • • , (where the are the standard basic unit vectors), then by the triangle inequality we have II
e X l xl = L ; ; = i l Thus, if M =
E�'=1
l edl.
then
Chapter 5: NORMED SPACES AND Lp-SPACES
222 holds for all x E the inequalities
IR". This establishes one-half of the desired inequality. In addition, l llx ii - IIYII I < llx - yll < M ll x - y lb
show that the function x r+ llx II from IR" with the Euclidean norm into IR is (uniformly) continuous. Let S = {x E IR": llx 112 = 1 } be the "unit sphere" for the Euclidean norm. Then S is closed and bounded, and hence, by the Heine-Borel Theorem 7 .4, S is compact for the Euclidean norm. In particular, the continuous function x r+ ll x II attains a minimum value on S, say at x0• Thus, llxll > llxoll holds for all x E S. Let K = llxoll· Since llxoll2 = 1, it follows that xo =I= 0, and so K = llxoll > 0. Now, if x E IRn is nonzero, then llxll/llxll2 = llxfllxlbll > K , and so
K llx ll2 < llx II holds for all x E IR". This establishes the other half of the desired inequality and !I completes the proof of the theorem. As an application of the last theorem, let us establish the following useful result:
Theorem 27.7.
is closed.
Eve1y finite dimensional vector subspace of a normed space
Proof. Let Y be a finite dimensional vector subspace of a normed space X.
Then Y can be identified (linearly isomorphically) with some IR". Hence, by Theo rem 27 .6, the norm of X restricted to Y must be equivalent to the Euclidean norm. In particular, Y is a complete metric space, and so (by Theorem 6 . 1 3) Y is closed. •
The structure of the open (or closed) unit ball reflects the topological and geo metrical properties of all balls. For example, a normed space is locally compact if and only if its closed unit ball is compact. (Recall that a topological space is locally compact if each point has a neighborhood whose closure is compact.) The next theorem tells us that the finite dimensional normed spaces are the only locally compact normed spaces.
Theorem 27.8. A
dimensional.
normed space is· locally compact if and only if it is finite
Proof. If the normed space is finite dimensional, then by Theorem 27 .6, its
norm is equivalent to the Euclidean norm. By Theorem 7.4, the closed unit ball is compact with respect to the Euclidean metric, and from this it follows (how?) that the space is locally compact.
Section 27: NORMED SPACES AND BANACH SPACES
223
For the converse, assume that X is a locally compact nonned space. Since all < 1} must E X: closed balls are homeomorphic, the closed unit ball V , X11 E V such that V C U�'= 1 B(x;, � and let Y be be compact. Choose XJ, X11 }. We shall show that X = Y, and this the linear subspace generated by will establish that X is fin ite dimensional. Assume by way of contradiction that X =F Y . Thus, there exists some EX the vector subspace Y is closed, it follows with x0 � Y . Since (by Theorem E Y} > 0 (also, see Exercise 1 of Section 10). that Y ) = inf x < Y). Then for some 1 < i < n we have Pick some E Y with belongs to Y , it follows that B(x; , � ). Since + •
•
•
{x1,
•
•
•
= {x
,
llxll ) .
x0
27.7)
d(x0,
{ll o yll: y llxo - yll 2d(xo, y (xo - y)fllxo - Yll E y llxo - yllx;
xo - y x; = ll xo - (y + llxo - yllx;)ll > d(xo, Y ) > -1 - llxo - Y ll 2 llxo - Yll llxo - Yll -
1 ->
---
2
which is impossible. This contradiction completes the proof of the theorem.
•
EXERCISES 1.
Let
X
be a normed space. Then show
sphere
2.
{.t =
E X: llx II =
d(x, y) llx - .vii).
X be
a.
Show that the mappings .t 1-+
a normed vector space. Fix
If A and B
3.
4.
X Let X
Let
a E X and a nonzero scalar a. .
a+
a. A + f3B
is an open set.
be a normed vector space, and let B
Show that B
= {x
E X: ll.tll � 1}.
a. and {3 are nonzero scalars,
E X: l .t II < be its open 1}
unit ball.
let {.t11 } be a sequence of X sue� that lim Xn = .t 1 holds. If y1 = n- (XJ + · · · + x11) for each then show that limy11 = x. (See also
of Section 4.) 5. Assume that two vectors x
n,
and y in a normed space
Then show that
7.
= {.t
isms.
be a normed space, and
Exercise 1 1
6.
morph
x and x 1-+ ax are both homeo
are two sets with either A or B open and
then show that
if its unit induced metric
is a Banach space if and only
1 } is a complete metric space (under the
Let
b.
that X
satisfy llx + yll = Ux ll + IIY II ·
ll a.x + /3yll = a.llxll + f311Y II holds for all scalars a. > 0 and {3 � 0. Let
X
be the vector space of all real-valued functions defined on [0, 1] having contin
11/11 = 1/(0)1 + 11/'llco is a norm on X that is equivalent to the norm 11/lloo + 11/' ll oo· A series L:�1xn in a normed space is said to converge to x if limllx - E7=1x; ll = uous first-order derivatives. Show that
0. As usual, we write summable if L�=
x = L:� 1 x11• A series L:�1xn
1 llxnll < oo holds.
Show that a normed space
X is
is said to
be absolutely
a Banach space if and only if every absolutely
summable series is convergent.
8.
Show that a closed proper vector subspace of a normed vector sp�ce is nowhere dense.
Chapter 5: NORMED SPACES AND Lp·SPACES
224
9.
Assume that f: [0, 1] ---+- IR is a continuous function which is not a polynomial. By Corollary we know that there exists a sequence of polynomials {Pn } that converges uniformly to f. Show that the set of natural numbers
11.6,
{k E lN': k = degree of Pn for some n } is countable. [mNT: Show that the vector subspace of ] consisting of all polynomials of degree less than or equal to some m is closed.] This exercise describes some classes ofimportant subsets of a vector space. A nonempty subset of a vector space X is said to be:
C[O, 1
10.
A
=-A;
symmetric, if x e A implies -x e A, i.e., if A convex, if x, y E A implies /..x + - A.)y E A for all 0 =::: /.. =::: two vectors x, y E A the line segment joining x and y lies in A; c. circled (or balanced) if x E A implies /..x e A for each l A. I :::
a. b.
(1
1,
i.e., for every
1.
Establish the following: 1.
ii.
m.
iv.
11.
A circled set is symmetric. A convex and symmetric set containing zero is circled. Anonemptysubset B ofavectorspaceisconvex ifandonly ifaB+bB = (a+b)B holds for all scalars a ::;: 0 and b > 0. If is a convex subset of a normed space, then the closure A and the interior A 0 of A are also convex sets.
A
This exercise describes all norms on a vector space X that are equivalent to a given norm. So, let (X, 11·11) be a normed vector space. Let A be a bounded convex symmetric subset of X having zero as an interior point (relative to the topology generated by the norm 11 · 11). Define the function PA: X ---+- IR by PA (X)
=
inf(/.. > 0:
X E /..A } .
Establish the following: a. b. c. d.
The function p A is a well-defined not:m on X. The norm PA is equivalent to 11·11 , i.e., there exist two constants > 0 and K > 0 such that C llx II < pA (x) =::: K llx II holds true for each x E X. i.e., {x e X : pA (x) < 1 } = A. The closed unit ball of p A is the closure of Let 11 1 · 11 1 be a norm on X which is equivalent to 11·11 , and consider the norm bounded nonempty symmetric convex set B = (x e X: lllxlll ::: 1 } . Then zero is an interior point of B and lllx Il l = pB (x) holds for each x e X .
C
A,
28. OPERATORS BETWEEN BANACH SPACES In this section, we shall study linear operators (or transformations). Recall that a function T: X ---+ Y between two vector spaces is called a linear operator (or simply an operator) if T(ax + {Jy) = aT(x) + {JT(y) holds for all x, y E X, and all a, fJ E R. Observe that every linear operator T satisfies T(O) = 0.
Section 28: OPERATORS BETWEEN BANACH SPACES
225
If X and Y are two normed spaces, then th_e symbol 11·11 will be used (as usual) to denote the norm on both spaces.
Y be a linear operator between two normed Let T: spaces. Then the operator norm of T is defined by Definition 28.1.
X �
li T II = sup{IIT(x) ll: llxll = I } .
If
isthen finite,T isthencalled ians called a bounded operator (and, of course, unbounded operator). T
li T II liT II = oo,
if
Observe that
IIT(x)ll < II T II · IIxll holds for all x E X. To see this, note that if x f= 0, then the vector y = x/llxll satisfies II II = l, and so
y
IIT(x)ll llxll
_
T
( ) x
llxll
= IIT(y )ll < li T II
holds. In particular, it follows from the previous inequality that li T II = sup{II T (x) ll: llxll < 1 } . Next, some concrete examples of operators are presented.
X X.
X X 1,
Example 28.2. Let = C[O, 1] with the sup norm. Define T : � by T(f)(x) = xf (x) for each f E X and x E [0, 1]. Clearly, T is a linear operator such that IIT(/)1100 < II f II 00 for each f E This shows that II T II .5 1 (in fact, II T II = ) and so T is a bounded operator. •
The next example presents
be
an unbounded
operator.
Example 28.3. Let X the vector space of all real-valued functions on [0, 1] that have continuous derivatives with the sup norm. Also, let Y = C[O, 1] with the sup norm. Define D: � Y by D(f) = f' (the differential operator). It is easy to see that D is a linear operator. However, liD II = oo. Indeed if f, (x) = xn, then 11/11 lloo = 1 and II D(/11)1100 = sup{11x11- 1 : x E [0. 1]} = n hold for each n, implying II D II = oo. Hence, D is an unbounded • operator.
X
,
The next example is drawn from the important class of operators known as "integral operators."
Example 28.4. Let [a, b] be a (finite) closed interval, and let K : [a, b] x [a, b] � lR be a continuous function. Consider the vector space C [a, b] with the sup norm, and then define
226
Ch apter 5: NORMED SPACES AND Lp·SPACES
T: �[a, b] --+ C[a, b] by
T(j)(x)
=
ih
K(x, y)f(y)dy
for each f E C[a, b]. (The unifonn continuity of K on [a, b] x [a, b] can be invoked here to verify that indeed T(j) E C[a, b] for each f E C[a, b].) For obvious reasons, the operator T is called an integral operator and the function K is referred to as the kernel of T. Clearly, T is a linear operator. On the other hand, if
M
=
sup {IK(x , y)l: (x, y) E [a, b] x [a, b]} < oo,
then the estimate IT(j)(x)l ::: M(b - a)llfll oo shows that IIT(f)ll oo ::: M(b - a)llfll oo· Thus, II T II ::: M (b - a) < oo, and so is a bounded operator. It is worth verifying that T also satisfies the following important property:
T
•
If B = {/ E C[a, b]: 11/lloo < 1 } is the open unit ball of C[a, b], tlzeu T(B) is a compact subset ofC[a, b].
To see this, observe first that T(B) is closed and bounded. Thus, by the Ascoli-Arzela Theorem 9.10, we need only to show that T(B) is an equicontinuous subset of C[a, b]. To see this, fix xo E [a, b], and let E > 0. Since (by Theorem 7 .7) K is unifonnly continuous, there exists some 8 > 0 such that IK(x l , y) - K(x2, y)l < E holds whenever lxt - x2l < 8. Therefore, if x E [a, b] satisfies lx - xo I < 8 and f E B , then
IT(j)(x) - T(j)(xo)l
=
lib
I
[K(x . y) - K (xo . y)]f(y) dy ::: (b - a)E.
This shows that T(B) is equicontinuous at xo. Since xo is arbitrary, T(B) is equicontinuous everywhere. Thus, T(B) is compact. The preceding property is expressed simply by saying that T is a compact operator. In general, an operator T: X --+ Y between two Banach spaces is said to be a compact • operator if T(B) is a compact subset of Y (where B is the open unit ball of X).
Our next example is borrowed from the theory of differential equations. Example 28.5. Let Ck[a, h] be the Banach space of Example 27. Consider C[a, b] with the sup norm ll·lloo. and fix k functions po, Pl . . . . , Pk-l in C[a. b]. Now, define L : Ck[a, b] --+ C[a, b] by L(y) for each y E Ck[a, b).
=
k y + Pk - lYC -l ) + · · · + Pl y' + POY
Section 28: OPERATORS BETWEEN BANACH SPACES
227
Clearly, L is a linear (differential) operator. On the other hand, if
M=
l + IIPolloc +
IIP 1 lloo + · · · + IIPk - l lloo.
then it is easy to see that
IIL(y)ll oc � M[llylloo + 11/k - l >lloo + · · · + lllll oo + II Y II oo l = Mllyll holds. This shows that L is a bounded operator. It is also interesting to observe that L is onto. This follows from the standard existence theorem of solutions to an ordinary linear differential equation. • The bounded operators are precisely the continuous linear operators. The next theorem clarifies the situation.
For a linear operator T : X Y between two normed spaces, thefollowing statements are equivalent: 1 . T is a bounded operator. 2. There exists a real number M > 0 such that II T (x)ll < Mllxll holds for all
Theorem 28.6.
X
�
E X.
is continuous at zero. 4. is continuous. Proof. (1) => (2) We have seen before that IIT(x) ll 3.
T T
< IIT II · IIxll holds for all x E X. Thus, if T is a bounded operator, then (2) holds for any choice of the real number M with M > liT 11 . (2) => (3) Clearly, liml!x, II = 0 combined with the inequality IIT(x, ) ll < M ll x,ll implies lim!I T(x11)11 = 0. That is, T is continuous at zero. (3) => .(4) The (uniform) continuity of T follows immediately from the iden tity IIT(x) - T(y)ll = II T(x - y)ll and the simple fact that lim x, = x holds in a normed space if and only if lim(x, - x) = 0. (4) => ( 1 ) Assume by way of contradiction that liT II = oo. Then there exists a sequence {x,} of X with ll x11 11 = 1 and IIT(x11)11 > n for each n. Let y, = ·�· for each n, and note that l!y, ll = fr implies limy, = 0. But then, by the continuity of T, we must have lim T(y,) = 0, contrary to IIT(y,)ll = n 1 11 T(x, )ll > 1 for each n. Thus, II T II < oo holds, and the proof of the theorem is complete. • -
Let X and Y be two normed spaces. Then the collection of all bounded linear operators from X into Y will be denoted by L(X, Y). By Theorem 28.6, it follows that L(X, Y ) under the algebraic operations ( S + T)(x) = S (x) + T (x)
and
(aT)(x) = aT(x)
Chapter 5: NORMED SPACES AND Lp·SPACES
228
is a vector space. In fact, L(X, f) is a normed vector space under the operator norm II T II = sup{ I T (x) II: llx II = 1}. The details follow.
Let X and f be two normed spaces. Then L(X, f) is a normed vector space. Moreover, if f is a Banach space, then L(X, f) is likewise a Banach space. Proof. We verify first that T II T II = sup{ I T (x) II: llx II = I } is a norm on L(X, f). Clearly, from its definition liT II > 0 holds for all T E L(X, f). Also, the inequality IIT(x)ll < IIT II · IIxll shows that liT II = 0 if and only if T = 0. The proof of the identity llaT I = Ia I · liT I is straightforward. For the triangle inequality, let S, T E L(X, f), and let x E X with llxll = 1. Then Theorem 28.7.
r+
II(S + T)(x)ll = IIS(x) + T(x)ll
<
IIS(x)ll + IIT(x)ll
<
liS II + li T II
holds, which shows that liS + T II < liS II + liT 11. Thus, L(X, f) is a normed vector space. Now, assume that f is a Banach space. To complete the proof, we have to show that L(X, f) is a Banach space. To this end, let {Tn} be a Cauchy sequence of L(X, f). From the inequality IIT"(x) - Tm(x)ll < I Tn - Tm ll · llxll, it follows that for each x E X, the sequence {T (xn)} of f is Cauchy and thus convergent in f. Let T (x) = lim T11 (x) for each x E X, and note that T defines a linear operator from X into f. Since {Tn} is a Cauchy sequence, there exists some M > 0 such that iiTnll < M for all n. But then, the inequality IIT11(x)ll < IITnll · llxll < M IIxll, coupled with the continuity of the norm, shows that II T (x) II < M llx II holds for all x E X. Therefore, by Theorem 28.6, T E L(X, f). Finally, we show that lim T11 = T holds in L(X, f). Let E > 0. Choose k such that IIT11 - Tm II < E for all n, m > k. Now, the relation
for all n,
m > k implies
IIT(x) - Tn(x)ll
m-+oo IITm(X) - Tn(x)ll
= lim
for all n > k and x E X. That is, we have therefore, lim Tn = T holds in L(X, f):
liT - Tn ll
< <
Ellxll E for all n > k, and
A subset A of L(X, f) is said to be pointwise bounded if for every x subset {T(x): T E A} off is norm bounded.
•
E X the
Section 28: OPERATORS BETWEEN BANACH SPACES
A L(X, liT I < M
If a subset that
of Y) is norm bounded (i.e., if there exists some holds for each then i"n view of the relation
T E A,
T E A), I T(x)ll < I T I · I xl < Mll xll A
229
M
>
0 such
is· pointwise bounded. That is, every norm for each it follows that bounded set of operators is pointwise bounded. The converse of this statement is also true, provided that is a Banach space. The result is known as "the principle of uniform boundedeness," or as "the Banach-Steinhaus theorem," and it is a very powerful theorem.
X
Let X be a Ba be a normed space. Then a subset of L(X, Y) is norm space, and let nach bounded if and only if it is pointwise bounded. Proof. L(X, A c L (X, n = {x E X: II T (x)ll < n T E A} A X = u:, X 6.18, kE yE r 0 l x - Yll < r E X E A, = r, rx) x l ( y = l x l Yl l y rx E rii T(x)ll = II T (rx)ll = I T(y rx) - T(y)ll < I T (y rx)ll I T(y)ll < 2k. -1 = M E X < 2kt · x I T (x)l l l x l = A T E A. L(X, Y). li T I < M Theorem 28.8 (The Principle of Uniform Roundedness). Y
We already know that every norm bounded subset of Y) is point Y) be pointwise bounded. wise bounded. For the converse, let Start by observing that for each the set for all
E,
En is norm closed. Also, since is pointwise bounded, it follows that and Theorem there exists some holds. In view of the completeness of implies with (Ek)0 =f. 0. Choose such that Ek and > X Ek. with Now, let T and let l . Since it + follows that + Ek. and so
+
Thus,
holds for all
+
+
holds for all with l , and therefore, That is, is a norm bounded subset of
•
In applications, the next result is an often-utilized special case of the "Principle of Uniform Boundedness."
Let X be a Banach space, Y a normed space, and {T } a n E X, then T i s a sequence of L(X, I f T,(x) = T(x) holds for each x bounded operator. Corollary 28.9.
Y).
lim
Chapter 5: NORMED SPACES AND Lp-SPACES
230
The hypothesis that X is a Banach space in Theorem 28 . 8 is essential. The next example clarifies the situation.
Example 28.10.
Consider the vector space
terms are eventually zero. Then
X
consisting of all (real) sequences whose
X equipped with the sup norm is a normed space. On the other hand, the reader can verify easily that X is not a Banach space. Now, for each n define J,, : X -+ JR by II
fn (X) = L kXk k=l x = (XJ, x2 ) in X. From lf,,(x)l � (Lk=J k)llx !100, it follows that {f11} C L(X, JR). Moreover, lim f,,(x) = L:� 1 kxk holds for each x E X. (Observe that the series has in fact a finite number of nonzero terms.) Thus, {f,,} is pointwise bounded. However, it so that {fn} is not a norm bounded sequence. is easy to see that llfn II 2:. n hold for each for each
• . . .
n,
Hence, in general the Principie of Uniform Boundedness does not hoid true if X is not
•
assumed to be complete.
Another useful variation of the Principle of Uniform Roundedness (which is also due to S. Banach and H. Steinhaus) is known as the "Principle of Condensation of Singularities," and is stated next.
Theorem 28.11 (The Principle of Condensation of Singularities). For each
pair of natural numbers n and m , let Tn.m: X � Y be a bounded oper ator from a Banach space X to a normed space Y . If for each m we have lim sup11 IIT".m I = oo, then the set
lx
E X: lim sup i i
n-+oo
Tn,m (x) ll = oo for each
m
J
is a dense subset of X. Proof. For every n, m , and
Anmk
= {x
E X:
k,
let
IITp,m(X)II >
k
for some
p > n}.
Clearly, Anmk is an open set and we claim that it is also dense. We must show that if x e X and r > 0, then B(x, r) n Anmk # (/). So, let x e X and r > 0. such that By the Uniform Boundedness Principle there exists a vector
x0
lim sup ii
p-+00
Tp.m (Xo) ll = oo .
. Section 28: OPERATORS BETWEEN BANACH SPACES
Pick some E
>
0 such that E ll .toll <
231
r. From
it easily follows that either I!Tp.,(x)ll > k or else IITp,,(x + Exo) ll > k for some p > 11. That is, either X E B (x , r) n A nmk or X + EXo E B(x, r) n Anmk· Hence, 8 (x , r) n A,mk =/= 0, and so A,mk is an open dense set. Now, notice that the Baire Category Theorem 6. 17 guarantees that
n{Anmk:
II, m ,
k
=
I , 2, . . . }
is a dense set. Clearly, any point in this set satisfies lim sup11 IIT,_,(x) ll = oo for
�I
•
m.
Our next objective is to establish another important result of analysis known as the "open mapping theorem." It asserts that a surjective continuous linear operator between two Banach spaces is an open mapping (in the sense that it carries open sets onto open sets). To establish this result we need a lemma.
X
X -+
Let and Y be two Banach spaces, and let T : Y be an onto continuous linear operator. If zero is an interior point of a subset A of then zero is also an interior point of T(A). {rx : x E V} is Proof. Let V {x E llx II < 1}, and observe that r V the closed ball with center at zero and radius r. Since zero is assumed to be an interior point of A, there exists some r > 0 with r V c A. By the linearity of T, we must have T(r V) rT(V) c T(A). Hence, to establish the result it suffices Lemma 28.12.
=
X:
X,
=
=
to show that zero is an interior point of T(V). Clearly, U� 1 n V , and since T is an onto linear operator, Y U: 1 nT(V) also holds. By Theorem 6.18, there exists some k such that kT(V) has a nonempty interior. Since kT(V) kT(V), it follows that T(V) has an interior point. That is, there exist some y0 E T(V) and r > 0 such that B(y0, 2r) c T(V). Now, if y E Y satisfies IIY II < 2r , then y - Yo E T(V). Therefore,
X=
=
=
Y
= (y -
Yo) + Yo
E
T(V) + T(V)
c 2T(V) ,
where the last inclusion follows easily from the identity V + V {y E Y: IIYII < r } c T(V). By the linearity of T , it follows that
{y E Y:
IIYII < r 2-" }
c 2-"T(V)
=
=
2 V. That is,
T(2-"V)
holds for each n. Now, let y E Y be fixed such that IIYII < r 2- 1 . Since y e T(2 - 1 V), there exists some x 1 E 2- 1 V such that II Y- T (xJ) II < r 2-2 . Now proceed inductively. Assume that x, has been selected such that x, E 2-"v and IIY - :L;'=• T(x;)ll < r2-"- 1 .
Chapter 5: NORMED SPACES AND Lp·SPACES
232
Clearly, y - 2::�'- 1 T(x;) with IIY - 'L�';:{ T(x;)ll llx, ll < 2-" and y-
-- -- •
-
E T(2 " 1 V), and so there exists some Xn+1 E 2-" 1 V < r2 2 Thus, a sequence {x,} is selected such that "
II
L T (x; ) 1=1
hold for all n. Next, define
= II =
s,
ll sn+p - Sn
=
Y-
x1 +
· · ·
r( t ) 1= 1
+ x, for each n, and note that
II+P
L
i=n+1
x;
X;
.:S
ll+p
L llx; II < 2-"
i= n+1
=
for all n and p shows that {sn} is a Cauchy sequence. Let x lim s, in X. Then llxll < I:: : 1 llxn II < l (i.e., x E V), and by the continuity and linearity of T, we see that lim T(sn) T(x) = n-oo That is, y E T (V), and so {y now complete.
=
"
"" T(x1)
11-00 �
lim
i=1
=
y.
E Y: IIYII < �} c T (V ) . The proof of the lemma is
•
The open mapping theorem is stated next. The result is due to S. Banach.
Let X and Y be two Ba nach spaces, and let T: X Y be a bounded linear operator. If T is onto, then T is an open mapping (and hence, if in addition T is one-to-one, then it is a homeomorphism). Proof. Let 0 be an open subset of X, and let y e T(O). Pick a vector x E 0 T(x), and note that y T(O) T(x - 0) holds. Now observe such that y that zero is an interior point of x 0, and hence, by Lemma 28.12, zero is also an interior point of T(x - 0) y - T (0). This implies that y is an interior point of T (0). Since y is arbitrary, T ( 0) is an open set, and the proof of the theorem is Theorem 28.13 (The Open Mapping Theorem). -+
=
complete.
=
= -
•
=
Consider two normed spaces X and Y. Then a norm can be defined on X x Y by l!(x, y )ll llx II + lly ll. This norm is called the product norm. You should stop and check that indeed this function satisfies the properties of a norm. Some other (frequently used) norms equivalent to the product norm are
!l (x , y ) ll
=
? I
(ll x ll2 + llyll-)2
Section 28: OPERATORS BETWEEN BANACH SPACES and
ll (x , y) ll = max{llxll ,
233
II Y II}.
It should be noted that l im (x, , y,) = (x , y) holds in X x Y with respect to the product norm if and only if lim x, = x in X and lim y, = y in Y both hold. Moreover, if both X and Y are Banach spaces, then
x Y with the product nonn is likewise a Banach space. Unless otherwise specified, the Cartesian product of two normed spaces will be considered as a nonned space under its product nonn. The next example illustrates an application of the open mapping theorem to differential equations . Example 28.14.
X
Consider the differential equation
I Y(k) + Pk- l Y(k -J ) + · · · + P I Y + POY = q
·
(l)
where po, Pl· . . . , Pk-l and q are fixed continuous functions on a (finite) closed interval [a, b]. If y is a solution of (1) and c E [a, b], then the k real numbers y(c), y'(c), . . . , y
What we would like to demonstrate here (by using the open mapping theorem) is that the solutions of ( 1) depend continuously on their initial values. That is, we would like to show that "small perturbations" in the initial values cause "small perturbations" to the solutions. To do this, we need to translate the problem to the framework of Banach spaces. Consider the Banach space Ck [a, b] of Example 27. Clearly, its norm
controls the "sizes" of the functions and their derivatives. Also, consider the bounded linear operator L: C k [a. b] � C[a, b] of Example 28; that is,
L(y) = y + Pk- ly
[a, b] and define T: c k [a, b] � C[a, b] X IRk by T(y) = (L(y), y(c), y'(c), . . . , y
for each y E ck [a, b]. Then it is easy to check that T is linear and continuous. On the other hand, the existence and uniqueness of the solutions of (1) with prescribed initial values guarantees that T is one-to-one and onto. Thus, by the open mapping theorem, T - I (which exists and is linear) must be continuous.
Chapter 5: NORMED SPACES AND Lp·SPACES
234
r- 1
means that given E > 0 th ere exists some 8 > 0 such that 1 ur- 1 (/, x) - r (g , z)ll < E if (f, x) and (g. z) in C[a, b] X IRk satisfy II! - glloo + llx - zlloo < 8. Translating this statement back to the differential equation (1), we get the The continuity of
following: Suppose that
Yl and Y2 are two solutions of ( I ) satisfying the initial conditions
and
I Y2Cc) = fJ 1 , Y2(c) If
Ia; - {3; 1
<8
holds for each
i
=
=
fJ2 . . . . , y2(k-1)(c) = f3k ·
1, . . . , k, then
(k I ) (k I ) bi - Y2 II oo + II YI1 - Y2I II oo + · · · + II Y1 - - y2 - II oo < E. That is, the solutions of the differential equation values.
( I ) depend continuously on their initial •
Y is a function, then the subset G = {(x, T(x)): x E X} of X x Y is called the graph of T . Now, if X and Y are vector spaces and T is a linear operator, then G is a vector subspace of X x Y. Moreover, if T is a bounded operator, then it is a routine matter to verify that G is a closed subspace of X x Y . Recall that if T : X -+
The converse of this last statement is true for Banach spaces. The result is known as "the closed graph theorem," and it has numerous applications in analysis.
Let X and Y be llvo Banach spaces, and let T: X -+ Y be a linear operator. If the graph of T is a closed subspace of X x Y, then T is a bounded operator. Proof. Since the graph G of T is a closed subspace of the Banach space X x Y, it is a Banach space in its own right. The function (x, T(x)) x is a Theorem 28.15 (The Closed Graph Theorem).
1-+
linear operator from
G onto X that is clearly one-to-one and bounded. (It is bounded
ll(x, T(x))ll .) By the open mapping theorem, x is a homeomorphism. That is, x (x, T(x)) is a continuous (x, T(x)) function from X onto G. Now, observe that (x, T(x)) -+ T(x) is continuous. Therefore, x T(x) from X into Y is continuous since it is a composition of because
llxll
<
r+
ll xll + IIT(x)ll
=
r+
1-+
two continuous functions, and the proof of the theorem is complete.
•
EXERCISES 1.
Y be two Banach spaces and let T: X --+ Y be a bounded linear operator. Show that either T is onto or else T(X) is a meager set.
Let
X
and
Section 29: LINEAR FUNCTIONALS
235
2. Let X be a Banach space, T: X � X a bounded operator, and I the identity operator on X. If II T II < l , then show that I T is invertible. 1 [HINT: Show that (I - T)- = L:�0T".] -
3.
On C[O, I] consider the two norms llfll :o = sup{ l f(x)l: x E [O, l ] }
and
ll fll t =
fo
1
lf(x)l dx .
Then show that the identity operator / : (C[O, l] . ll·lloo) � (C[O, l]. l l · ll t ) is continu ous, onto, but not open. Why doesn't this contradict the open mapping theorem? 4. Let X be the vector space of all real-valued functions on [0, l ] that have continuous derivatives with the sup norm. Also, let Y = C [O, l ] with the sup norm. Define the mapping D : X � Y by D(f) = f' . Show that D is an unbounded linear operator. Show that D has a closed graph. Why doesn 't the conclusion in (b) contradict the closed graph theorem? 2 Consider the mapping T : C [O, I ] � C [O, I ] defined by Tf(x ) = x f(x) for all f E C[O, l ] and each x E [0, l]. a. b. c.
5.
6.
a. Show that T is a bounded linear operator. b. If / : C [O. l ] � C[O, l ] denotes the identity operator (i.e., !(f) = f for each f E C[O. l]), then show that II/ + Til = l + li T II.
Let X be a vector space which is complete in each of the two norms ll·ll t and 11 · 11 2· If there exists a real number M > 0 such that llx 11 1 .::: M llx 1!2 holds for all x E X, then show that the two norms must be equivalent. 7. Let X, Y, and Z be three Banach spaces. Assume that T: X � Y is a linear operator and S: Y � Z is a bounded, one-to-one linear operator. Show that T is a bounded operator if and only if the composite linear operator S o T (from X into Z) is bounded. [IDNT: Use the closed graph theorem.] 8. An operator P: V � V on a vector space is said to be a projection if ?2 = P holds. Also, a closed vector subspace Y of a Banach space is said to be complemented if there exists another closed subspace Z of X such that Y ffi Z = X. Show that a closed subspace of a Banach space is complemented if and only if it is the range of a continuous projection. [HINT: Use the closed graph theorem.]
29. LINEAR FUNCTIONALS
In this section, we shall discuss the basic properti es of continuous linear function als. Recall that a linear operator f: X � .IR, where X is a vector space, is called a linear functional on X. One of our objectives is to show that for every normed vector space X there are enough bounded linear functionals to separate the points of X . That is, for every p air of distinct vectors x and y of X, there exists a bounded linear functional on X such that f(x) i= f( y). This result (as well as many others) rests upon a classical
Chapter 5: NORMED SPACES AND Lp·SPACES
236
result known as the "Hahn-Banach theorem," which is one of the cornerstones of modern analysis. We shall discuss this theorem next. A mapping p: X -+ IR., where X is a vector space, is called a sublinear mapping if it satisfies the following two properties: a. b.
p(x + y) < p(x) + p(y) for all x, y e X; and p(ax) = ap(x) for all x e X and a > 0.
The crux of the proof of the Hahn-Banach theorem lies in the following lemma:
Let p be a sublinear mapping on a vector space X, Y a vector subspace of X, and xo rl. Y. If f is a linear functional on Y such that f (x) < p(x) boldsfor all .x e Y , then f can be extended to a linearfunctional g on tbe l'ector subspace Z generated by Y and x0 satisfying g(.x) < p(x)for all x e Z. Proof. Clearly, Z = {x + ax0: x e Y and a e IR}. Assume that g is a linear functional on Z that agrees with f on Y. Then g(x + axo) = f(.x) + ag(xo) holds for all x e Y and a e IR., and so g is determined uniquely by the value g(xo). Let c = g(x0). Thus, every real number c gives rise to a linear functional on Z that agrees with f on Y. Our objective is to show that there exists some value of c such that Lemma 29.1.
f(x) + ac < p(x + a.xo) holds for all x e Y and a e IR.. For a > 0 and x e Y , the inequality (*) is equiva lent to c < p(a- 1 x + xo) - f(a- 1 x) and to - f(a- 1 x) - p(-a- 1 .x - x0) < c for .x e Y and a < 0. Certainly, these inequalities [and hence, (*)] will be satisfied by a choice of c for which
- f(x) - p(-x - xo) < c < p(x + .xo) - f(x) holds for all .x e Y. If .x, y e Y, then the relations
f(y) - f(x) show that - f(x) Consequently, if
=
f(y - x) < p(y - x) = p(y + xo + (-.x - .xo)) < p(y + .xo) + p(-x - xo)
- p(-x - xo) < p(y + xo) - f(y)
s = sup{- f(x)- p(-x -xo): x
E Y}
and
holds for all
t = inf p( {
y+
x,
y e
Y.
.xo) - f(y): y E Y } ,
then s and t are both real numbers and s < t. But then any real number c such that s ::: c < t satisfies (**), and hence, (*). This completes the proof of the • lemma.
Section 29: LINEAR FUNCTIONALS
237
The classical Hahn-Banach theorem is stated next. This theorem is the heart of modern analysis, and it has far-reaching applications.
Theorem 29.2 (Hahn-Banach). Let p be a sublinear mapping on a vector
space X, and let Y be a vector subspace of X. If f is a linear functional on Y such that f(x ) < p(x) holds for all x E Y , then f can be extended to a linear functional g on all of X satisfying g(x) < p(x)for every x E X. Proof. Let C be the collection of all pairs (g, Z) such that Z is a vector subspace
of X containing Y and g is a linear functional on Z satisfying g(x) = f(x) for all x E Y and g(x) < p(x) for all x E Z. The collection C is nonempty since (f, Y) E C. Define an order relation on C as follows: (g2, Z2) > (g1 , Z 1 ) whenever Z 1 c Z2 and g2(x) = g1 (x) for all x E Z 1• It is easy to verify that > is indeed an order relation on C. Now, consider a chain {(g;. Z;): i E / } ofC (i.e., for every pair i, j E I either (g;, Z;) > (gj. Zj) or (gj, Zj) > (g;, Z;) holds). Let Z = U ; e 1 Z;, and note that Z is a vector subspace of X. Now, define g: Z � JR. by letting g(x) = g;(x) if x E Z;. Since {(g;, Z;): i E / } is a chain, the value of g(x) is independent of the chosen index i. Clearly, g is a linear functional. In addition, g(x) = f(x) holds for all x E Y, and g(x) < p(x) for each x E Z. Thus, (g, Z) E C, and clearly, (g, Z) > (g;, Z;) holds for all i E I. Therefore, every chain of C has an upper bound in C. By Zorn 's lemma, C has a maximal element, say (g, Z). To complete the proof, it suffices to show that Z = X. Indeed, if Z # X, then there exists some x0 E X with x0 ¢ Z. Let M be the vector subspace generated by Z and x0• By Lemma 29. I , there exists a linear functional h on M such that h(x) = g(x ) for allx E Z and h(x) < p(x ) forallx E M . But then (h, M) E C, and (h, M) > (g, Z) holds with (h, M) # (g, Z), contrary to the maximality property • of (g, Z). Hence, Z = X holds, and the proof of the theorem is complete. The next three theorems are applications of the Hahn-Banach theorem. The first result tells us that a continuous linear functional defined on a subspace can be extended to a continuous linear functional on the whole space with preservation of its original norm.
Theorem 29.3. Let Y be a vector subspace of a normed space X, and let f
be a continuous linearfunctional on Y . Then f can be extended to a continuous linear functional g on X such that ll g II = II f 11. = sup { l f(y ) l : y E
Y
and IIYII < 1 } < oo. Define p: X � JR. by p(x) = llfll · ll xll for each x E X. Note that p is a sublinear mapping on X such that f(x) < p(x) holds for all x E Y . By the Hahn-Banach theorem there exists a linear extension g of f to all of X such that g(x) < p(x) holds for all x E X . This implies lg(x)l < llfll· llxll for all x E X, and so llgll < llfll. On the
Proof. Let llf ll
Chapter 5: NORMED SPACES AND Lp-SPACES
238
other hand,
llf ll = sup {lf(y)l : y E Y and llyll < 1 } < sup { lg(x)l : x E X and llx II < 1 } = llgll also holds, so that II g II =
II f 11. Thus, g is a desired extension off to all of X.
•
For a nonned space X, the Banach space L(X, lR) is called the norm dual of X and is denoted by X * . That is, X * consists of all functions f: X � 1R that are both linear and continuous. The members of X * are called the continuous linear functionals on X. The norm dual X * plays an important role in the theory of Banach spaces, and part of this theory deals with the properties and structure of X * . The vectors of a nonned vector space are always separated by its continuous linear functionals.
Theorem 29.4. If x is a vector in some normed space X, then there exists a
continuous linear functional f on X such that llfll = 1 and f(x) = llxll. In particular, X * separates the points of X. Proof. If x = 0, then any f E X * with llfll = 1 satisfies f(x) = llxll. (When
X =P {0 }, Theorem 29.3 guarantees X * =P {0 } .) Thus, assume x # 0. Let Y = {ax: a E lR}, the vector subspace generated by x. Then, the formula g(ax) = a llxll defines a continu�ms linear functional on Y such that llgll = 1 and g(x) = llxll. By Theorem 29.3, there exists a continuous linear extension f of g to all of X with ll f ll = llgll = 1. Clearly, f(x) = llxll, as desired. To see that X * separates the points of X, let x , y E X with x =j:. y. By the preceding, there exists some f E X * such that f(x) - f(y) = f(x - y) • llx - y II # 0. Hence, f(x) =j:. f(y) holds, and the proof is complete. A vector always can be separated from a closed subspace that does not contain
it by a continuous linear functional. The details are included in the next result.
Theorem 29.5. Let Y be a vector subspace of a normed vector space X, and
let xo ¢ Y. Then there exists some f .E X * such that f(x) = 0 for all x E and f(xo) = 1 .
Y
Proof. Since xo ¢ Y, there exists some r > 0 such that llx - x0 II > r holds for all x E Y. Let Z = {x + axo : x E Y and a E lR}, the vector subspace generated
Section 29: LINEAR FUNCTIONALS by Y and xo. Now, define functional and
f: Z
� lR by
f(x
+ axo) =
239
a. Then f is a linear
holds for all x E Y and a E lR. It follows that f is a continuous linear functional Also, f(x) = 0 holds for each x E Y and on Z whose nonn does not exceed f(xo) = l . Now, apply Theorem 29.3 to extend f to a continuous l inear functional on X. •
r-1.
Let X be a nonned space. As we have seen in Theorem 28.7, the nonn dual X* [ = L(X, lR)] of X is always a Banach space. In particular, the nonn dual (X*)* of X* is likewise a Banach space. This Banach space is called the second dual of X and is denoted by X**. That is, X** = (X*)*. Every vector x E X gives rise to a continuous linear functional -� on X* via the fonnula
S:(f) = f(x) for all
f E X*. Indeed, clearly, S: is linear on X*, and the inequality IS:(f) l = l f(x) l
<
ll fll · llxll = ll.xll · ll f ll
shows that .t E X** and n.t-11 < llxll . On the other hand, by Theorem 29.4 there exists some f E X* with ll f ll = I and f(x) = ll.xll . Thus, llxll = f(x) = 1-t (f) l < 11 t ll and so 11-� 11 = llxll holds for all x E X. The mapping x H- -� (from X into X**) is called the natural embedding of X into its second dual X**. Summarizing the previous, we have the following theorem: -
,
The natural embedding x H- S: ofa normed space X into its second dual X** is a norm preserving linear operator (and hence, X can be considered as a subspace of X**).
Theorem 29.6.
-+
A linear operator T: X Y between nonned spaces that satisfies II T (x) II = llxll for all x E X is called a linear isometry. In this tenninology the preceding theorem can be phrased as follows: The natural embedding x H- -� is a linear
isometry.
In general, x H- S: is not surjective, and hence, X (when embedded in X**) is, in general, a proper subspace of X**. If X is not a Banach space, then x H- .t· cannot be onto, simply because X** is a Banach space. If the natural embedding of a Banach space X into its second dual X** is onto, then X is called a reflexive Banach space, and this is denoted by X** = X. The properties of the natural
Chapter 5: NORMED SPACES AND L,-SPACES
240
X**
X
embedding of into its second dual are utilized in many applications, some of which are illustrated below. then it is easy to see that its If Y is a vector subspace of a normed space closure Y is also a vector subspace (why?). In particular, if is a Banach space, then Y is the completion of Y. These observations will be used in the proof of the next theorem.
Theorem 29.7.
X,
X
The completion ofa normed space is a Banach space.
X be a normed space. Consider X embedded in X ** by its natural embedding. Then X ** is a Banach space and induces on X its original norm. Thus, X is the completion of X, which is likewise a Banach space. Proof. Let
•
The next result can be viewed as a"dual" of the Uniform Boundedness Principle.
Let A be a subset ofa normedspace X such that { f(x): x E A } is a bounded subset of JR. for each f E X*. Then A is a norm bounded subset of Theorem 29.8.
X.
X embedded in X ** . Then A as a subset of X ** is pointwise bounded on the Banach space X * . Thus, by the Principle of Uniform Boundedness (Theorem 28.8), A is norm bounded in X ** . It follows that A is also norm bounded X, and the proof is finished. Proof. Consider
•
in
EXERCISES I. Let I: X � JR. be a linear functional defined on a vector space X. The kernel of I is
the vector subspace
2.
Ker I = 1- 1 ((0}) = (x E X : l(x) = 0}. If X is a nonned space and I: X � JR. is nonzero linear functional, establish the following: a. I is continuous if and only if its kernel is a closed subspace of X. b. I is discontinuous if and only if its kernel is dense in X. Show that a linear functional I on a nonned space X is discontinuous if and only if for each a E X and each r > 0 we have I (B(a, r)) = (l(x): !Ia - xll < r} = JR.
3. Let 1. l1 , h, . . . , In be linear functionals defined on a common vector space X. Show
that there exist constants AI, , ).., satisfying I = 'L7 1 ')..;I; (i.e., I lies in the linear span of II . . . ' In ) if and only if nr= I Ker I; c Ker I. Prove the converse of Theorem 28.7. That is, show that if X and Y are (nontrivial) nonned spaces and L(X, Y) is a Banach space, then Y is a Banach space. .
4.
•
•
=
Section 29: LINEAR FUNCTIONALS
241
[HINT: Let {y,) be a Cauchy sequence c;>f Y. Pick f e X* with f # 0, and then consider the sequence {T, ) of L(X , Y) defined by T,(x) = f(.t)y,.] The Banach space B(JN) is denoted by 1!.':)0. That is, 1!.00 is the Banach space consisting of all bounded sequences with the sup nonn. Consider the collections of vectors
5.
co
=
{x
c = {x 6.
=
=
(.\: J , x2. X3, . . . ) E
f.oo: x, � 0 ). and
(.q , x2 • .t3, . . . ) e 1!. 00 :
lim x, exists in
JR. ).
Show that co and c are both closed vector subspaces of 1!.00• Let c denote the vector subspace of f.cx:. consisting of all convergent sequences (see Exercise 5 above). Define the limit functional L: c � JR. by L(x)
=
L(x1, x2 . . . . ) = lim 11-+00 x, ,
and p: eoo � JR. by p(X) = p(XJ, X2,
.
•
.
) = limsupx11•
Show that L is a continuous linear functional, where c is assumed equipped with the sup nonn. b. Show that p is sublinear and that L(x) = p(x) holds for each x e c. c. By the Hahn-Banach Theorem 29.2 there exists a linear extension of L to all of f.:o (which we shall denote by L again) satisfying L(x) � p(.\:) for all .t E 1!.00 • Establish the following properties of the extension L: 1. For each x E 1!.00, we have a.
lim infx, � L(x) � limsupx11 • 11-+00 /1-+00
n. m.
7.
L is a positive linear functional, i.e., x =:: 0 implies L(x) L is a continuous linear functional (and in fact II L II
=
1
:::,
0.
).
Generalize Exercise 6 above as follows. Show that there exists a linear functional £im: eoo � JR. satisfying the following properties: a. .Cim is a positive linear functional of nonn one. b. For each x (x,, x2, . . . ) E 1!.00, we have XJ + -t2 + · · · + x, 1 . x 1 + x2 + · · · + x, .c · tmmf . � 1m( x) � 1.tm sup 11-+oo n n 11-+oo In particular, .Cim is an extension of the limit functional L. ) e 1!. 00, we have c. For each x = (XJ, x2 =
.
• . . .
£im(.q , X2, X3 , . . ) = .Cim(X2, X3, X4, . . . ) . .
Any such linear functional .Cim is called a Banach-Mazur3 limit. [HINT: Define p : 1!.00 � JR. by p(x) = lim sup x1 +x,; and note that p is sublinear satisfying L(.r) = p(x) for all x e c.] 8. Let X be a nonned vector space. Show that if X* is separable (in the sense that it contains a countable dense subset), then X is also separable. .
3 Stanislaw
.
+x,
Mazur ( 1905-1981 ), a Polish mathematician and a close collaborator of Stefan Banach.
He made important contributions to functional analysis and probability theory.
Chapter 5: NORMED SPACES AND Lp·SPACES
242
[HINT: Let {ft , h, . . . } be a countable dense subset of X*. For each n choose x, E X with l lx,ll = 1 and !f,,(x,)l 2:: � !If, II, and let Y be the closed subspace generated by {xi, x2 }. Use Theorem 29.5 to show that Y = X.] 9. Show that a Banach space X is reflexive if and only if X* is reflexive. [HINT: If X =I= X**, then by Theorem 29.5 there exists a nonzero F E X*** such that F(x) = O for all x E X.] 10. This problem describes the adjoint of a bounded operator. If T: X .....:,. Y is a bounded operator between two normed spaces, then its adjoint is the operator T*: Y* � X* defined by (T* f)(x) = j(T x) for all f E Y* and all x E X. (Writing ll(x) = (x, the definition of the adjoint operator is written in "duality" notation as . . . .
·
h},
(Tx, j}
=
(x, T* f}
for all f E Y* and all x E X.) a. Show that T*: Y* � X* is a well-defined bounded linear operator whose norm coincides with that ofT , i.e., II T* II = II T 11 . b. Fix some g E X* and some u E Y and define S: X � Y by S(x) = g(x)u. Show that S is a bounded linear operator satisfying II S II = II g 11 · 11 u 11. (Any such operator S is called a rank-one operator.) c. Describe the adjoint of the operator S defined in the preceding part (b). d. Let A = [aij] be an m x n matrix with real entries. As usual, we consider the adjoint operator A* as a (bounded) linear operator from R" to Rm . De�cribe A*.
30. BANACH LATTICES As we have seen, both the continuous functions and the measurable functions have a natural ordering under which they are lattices. For this reason, it is important to consider Banach spaces that are also lattices. Let us begin by reviewing a few of the basic facts that will be needed about vector lattices from Section 9. equipped with an A partially ordered vector space is a real vector space order relation .:::. that is compatible with the algebraic structure as follows:
1. 2.
If x .:::. If x >
y, then x +
y, then ax
z
>
.:::.
X
holds for all z E holds for all a .:::. 0.
y+
ay
z
X.
and its members The set x+ = (x E x > 0) is called the positive cone of are called the positive vectors of Clearly, the sum of two positive vectors is again a positive vector. A partially ordered vector space is called a vector lattice (or a Riesz4 space) if for every pair of vectors x, y E both_ sup(x, y) and inf{x, y) exist As usual, sup{x, y) is denoted by x v y and inf{x, y) by x 1\ y. That is, x v y = sup{x, y) and x 1\ y = inf{x, y ) .
X:
X. X X
X,
4Frigyes (Fred�ric) Riesz (1880-1956), a dislinguished Hungarian maihemaiician. His name is closely associa1ed wilh Ihe developmem of funclional analysis in 1he tiTSI half of Ihe Iwentieth cemury.
Section 30: BANACH LATTICES
243
In a vector lattice, the positive part, the �egative part, and the absolute value of a vector x are defined by x+ = x V O, x- = (-x) v O,
and !x! = x V (-x),
respectively. By Theorem 9.1 the following identities hold: x = x+ - x-
and
!x! = x + + x -.
A number of useful inequalities are stated in the next theorem.
Theorem 30.1. If x, y, and z are vectors in a vector lattice, then the following
inequalities hold: l. 2. 3. 4. 5.
!x + y ! < l x l + l y l ; l lx l - l y l l < !x - y ! ; lx + - y+ l < !x - y ! ; !x v z - y v zl < !x - y!; !x 1\ z - y 1\ z ! < !x - y l .
<
lx l and y < I Y I . it follows that x + y < lx l + I Y I · Similarly, -(x + y) < lx l + ! y l holds, and thus
Proof. ( l ) From x
!
!x + y ! = (x + y) v [-(x + y) J < x l + I Y I (2) By ( l ) we seethat lx l = !x -y+y ! < !x-y l + ! yl. and so !x ! - l yl < lx -y !. Similarly, ly l - l x l < !x - y!. and hence, l lx i - I YII < lx - y !. (3) Note that x+ = ! <x + !x l ). Then to establish the required inequality, use ( 1) and (2) as follows: 1 1 1 ! x+ - y+ ! = l ) - (Y + l y l) = l(x - y) + ( !x ! - ! y l ) ! !x (x + z z l
1 I < 2 1x - Y l + 2 1 1xi - I Y I I < !x - y !. (4) Note that xv
z -
y v z = [(x - :) v 0 + z]
- [(y - z) v 0 + z] = (x - z)+ - (y - z)+
holds. (See the identities in Section 9 before Theorem !x v z - y v rl = !(x - z) +
9. 1.) Thus, by (3) we have
- (y - z) + l < ! (x - z) - (y - :) 1
=
!x - yl.
244
Chapter 5: NORMED SPACES AND Lp·SPACES
(5) Since x 1\ z (4) that
= -[(-x ) v (
-
z
)] and y 1\ z =
-
[(-y) v (-z)], it follows from
lx 1\ z - y 1\ z I = I (-y ) v (-z) - (-x) v (-z) I � I - y - (-x ) I = lx - y I, •
and the proof of the theorem is complete.
Let X be a vector lattice. A subset A of X is called order bounded if there exists an element y e X such that lx I < y holds for all x e A. A linear functional on X is said to be order bounded if it carries order bounded subsets of X onto order bounded subsets of JR. That is, a linear functional f: X � 1R is order bounded if for every y e x + there exists some M > 0 such that lf(x)l < M holds for all x E X with lxl � y. A linear functional f on X is called positive iff(x) > 0 holds for each x e X + . Clearly, every positive linear functional is order bounded. As we shall see in the L.,_ .. ._:••-- -- - d!U---- ,.., rl l:no".- fU"""�:,..... nJ nrrle" hounrl nPVt t!hPQrPnl PUPI"1.1 aue \..,1\;;t vv 11Llc;;ta a.� a 1J1C1 Cll\,;C . -�J of two positive linear functionals. The collection of all order bounded linear functionals on a vector lattice X is denoted by x- and is called the order dual of X. Obviously, x- (under the usual algebraic operations) is a vector space. Moreover, if we define f > g whenever f(x) > g(x) holds for all x E x+, then it is easy to see that x- equipped with > is a partially ordered vector space. In actuality, x- is a vector lattice, as the next result of F. Riesz shows. -n.
·-
• -···,
""
V'a �
.& v
a
u ····�LU.
"'"'""
11\.,LaVIIQI \.f(UI
is a vector lattice, then its order dual x- is likewise a vector lattice. Moreover,
Theorem 30.2 (F. Riesz). If X
f + (x) = sup {f (y): 0 < y < x}, f- (x) = sup{- f (y): 0 < y < x}, and lf l(x) sup {f (y ) : lyl < x }
=
holdfor each f E X- and a/1 X E x + . Proof. In view of the identities
f v g = (f - g)+ + g and f 1\ g = -[(- f) v (-g)], we establish that x- is a vector lattice by proving that f+ exists for each f E x-. To this end, let f E x-. Define g: x+ � 1R by
g(x) = sup {f(y):
0
x}
Section 30: BANACH LATTICES
245
for each x E x+. The supremum i s finite since f is order bounded. Clearly, g(x) > 0, g(x) > f(x), and g(ax) = ag(x) hold for all x E x+ and a > 0. We claim that g(x + y) = g(x) + g(y) holds for all X ' y E x + . To see this, let X ' y E x+. If ll and u satisfy 0 < u < X and 0 < u < y ' then 0 < ll + u < X + y holds, and so f(u) + f(u) = f(u + u) < g(x + y); consequently, g(x) + g(y) < g(x + y). On the other hand, i f O < z < x + y, then let u = x A z, u = z - u , and note (by using the identities in Section 9 before Theorem 9. 1 ) that 0< u
=z-
x A z = z + (-x) v (-z) = (z - x) v 0 < y v 0 = y.
Therefore, 0 < ll < x and 0 < u < y. This implies f(:) = f(u + u) = f(u) + f(u) < g(x) + g(y), from which itfollows that g(x+y) < g(x)+g(y). That is,g(x+y) = g(x)+g(y). Now, for arbitrary x E X define g(x) = g(x+) - g(x-). Note that if x = u - u with u , u positive, thenx++u = u +x - , andso g(x+)+g(u) = g(u)+g(x-) holds by the additivity of g on x+. Therefore, g(x) = g(x+) - g(x-) = g(u) - g(v), and so the value of g(x) does not depend upon the particular representation of x as a difference of two positive vectors. The preceding observation, combined with ag(x) for all x , y E x+ and a > 0, implies g(x + y) = g(x) + g(y), and g(ax) that g is a linear functional on X. Finally, we show that g = J + holds in X"'. Indeed, if h is another positive linear functional satisfying f < h, then since f(y) < h(y) < h(x) holds for all 0 < y < X, it follows that g(x) < h(x) for each X E x + . Therefore, g is the least upper bound of f and zero. That is, g = J+ holds in X"'. This shows that X"' is a vector lattice and that J+(x) = sup{f(y): 0 < y < x} holds. The formula for f- follows from f- = (-f)+, and the formula for the absolute value follows from I f I = J+ + f-. II
=
The formula lfl(lxl) = sup{f(y): IYI < lx l } implies the following useful in equality lf(x)l < lfl(lxl), for all f E and all x E X. In view of the identity f v g = (f - g)+ + g and Theorem 30.2, the following identities hold:
X"'
f v g(x) = sup{f(y) + g(x - y): 0 < y < x}
Chapter 5: NORMED SPACES AND Lp-SPACES
246 and
f 1\ g(x) = inf{/(y) + g(x - y):
0
< y < x}
for all f, g E x- and x E x + . Sometimes it is important to know that the "dual" formulas of Theorem 30.2 are also true. More precisely we have the following:
X X
Let be a vector lattice, and let f be a positive linearfunc tional. Then for eveI)' x E the following identities hold: f(x+) = sup{g(x): g E x- and 0 < g < /}, f(x-) = sup {-g(x): g E x- and 0 < g < / } , and / ( lx l) = sup{g(x): g E x- and l gl < /}. Proof. We establish the formula for f(x+ ). Note first that ifO < g � f holds, then g(x) < g(x + ) < f(x+ ), and so sup{g(x): g E x- and 0 < g < /} < f(x + ). For the reverse inequality consider the function p: -+ IR defined by p(u) = f(u+). It is easy to see that p is a sublinear mapping on X such that p(u) > 0 holds for each u e Now, let Y = {ax: a e IR } , and define h: Y -+ IR by h(ax) = af(x+ ). Obviously, h(u) < p(u) holds for all u E Y, and so, by the Hahn-Banach Theorem 29.2, h can be extended linearly to all of so as to preserve the inequality h (u) < p(u) for all u e Next, observe that if u > 0, + then h(u) < p(u) = f(u ) = f(u) holds, and moreover, -h(u) = h( -u) < p(-u) = f((-u)+) = /(0) = 0. That is, 0 < h < f holds. Therefore, f(x+ ) = h(x) < sup{g(x): g E x- and 0 < g < /}, so that Theorem 30.3.
·
X
X.
X
X.
f(x+ ) = sup{g(x}:
g e x- and 0
< g < f}.
The other two formulas now follow easily from the relations f(x-) and /(l x l ) = f(x+ ) + f(x-).
X
= f (( -x)+ )
•
A norm 11 ·11 on a vector lattice is said to be a lattice norm whenever lx I < I y I in X implies ll x I I < II y 11- A normed vector lattice is a vector lattice equipped with a lattice norm. If a normed vector lattice is complete, then is referred to as a Banach lattice. Let be a normed vector lattice. Then it should be clear that llx II = ll lx I ll holds for all x E Also, in view of Theorem 30. 1 , the following inequalities are valid for all x, y E
X X. X:
llx+ - y+ ll < llx - y ll
X
and
X
ll lx i - IYI II < IIx - y ll .
Section 30: BANACH LATTICES
247
In particular, they imply that the mappings _x H- x+ and x H- lx I (both from X into X) are uniformly continuous. It is interesting to observe that most of the normed spaces one encounters in analysis are normed lattices or Banach lattices. Some examples follow.
Example 30.4. Let X be a nonempty set. The collection of all real-valued bounded func tions defined on X is denoted by B(X). Then B(X) is a vector lattice (in fact a function space) under the ordering f ::: g whenever j(x) ::: g(x) holds for all x e X. Also, under the sup norm 11/lloo sup{lf(.t)l: x e X}, the vector lattice B(X) is a Banach lattice. See Example 6.12. •
=
Example 30.5. Let X be a topological space. Denote (as usual) by Cc(X) the vector space of all continuous real-valued functions on X that have compact support. In other words, f e C(X) belongs to Cc(X) if and only if the set {.t e X : f(x) =I= 0} has compact closure. (When X is compact, Cc(X) = C(X).) Then Cc(X) under the pointwise ordering (i.e., f ::: g if f(x) ::: g(x) for all x e X), and the sup norm is a normed vector lattice. • Example 30.6. Consider C [0, 1] with the pointwise ordering, and define a norm by II f II = J� 1/(.t)l dx. Then C[O, I ] under this norm is a normed vector lattice, but not a Banach
.
��
Example 30.7. Let e 1 denote the vector space of all (real) sequences x = (.ta, x2, . . .) such that L�l lx/1 I < 00. With the pointwise algebraic and lattice operations e I is a function L� dx11 I. the vector lattice e 1 becomes a Banach lattice space. Under the norm llx II (see Example 27.3). •
=
A vector subspace Y of a vector lattice X is called a vector sublattice of X if for every pair x and y of Y, the elements x v y and x A y belong to Y . That is, Y is a vector sublattice of X if it is closed under the lattice operations of X. A vector subspace V of a vector lattice X is said to be an ideal whenever lxl < lyl and y E V imply x E V . In view of the identity x v y !<x + y + lx - yl), it follows that every ideal is a vector sublattice. Let X be a normed vector lattice and let A be an order-bounded subset of X. Pick some y E X such that lxl < y for each x E A, and consequently, llx II < llyll holds for all x E A. That is, every order-bounded set is norm bounded. It follows from this that every continuous linear functional on X carries order-bounded subsets of X onto bounded subsets of JR. In other words, the norm dual X* is a vector subspace of the order dual X"'. The next result shows that X* is, in actuality, an ideal of x-.
=
The norm dual X* ofa normed vector lattice X is an ideal of the order dual x- (and hence, X* is a \'ector lattice in its own right). Proof. Assume 1 / 1 < lgl holds in x- with g E X* and f E x-. We must
Theorem 30.8.
show that f E X*.
Chapter 5: NORMED SPACES AND Lp-SPACES
248
Let x E X satisfy ll x ll = < ll x ll = L Therefore,
IIYII
L Then for every y E X with
l f(x)l :S l fl(lx l) < lgl(lxl)
= sup {g(y) :
IYI
< lx l we have
IYI :S lx l } :S llgll ,
where the equality in the middle holds by virtue of Theorem 30.2. This shows that • f E X* and ll f ll < llgll . The preceding arguments also show that for a normed vectorlattice X the rehltion l f l < l g l in X* implies llfll :S llgll . That is, X* is a normed vector lattice. Since X* is in addition a Banach space (Theorem 28.7), the following result should be immediate.
The norm dual ofa normed vector lattice is a Banach lattice.
Theorem 30.9.
If X is a normed vector lattice, then Theorem 30.8 shows that X* c X"' holds. However, the norm dual of a Banach lattice always coincides with its order duaL
Theorem 30.10.
IfX is a Banach lattice, then X*
=
X"' holds.
Proof. Let f E X"'. Assume by way of contradiction that f is not continuous.
That is, assume that llf ll = sup {l f(x) l: ll x ll = 1 } = oo. Then there exists a sequence of vectors {x,} such that ll x, ll = 1 and l f(x11) 1 > 11 3 for all 1z. 2 Let = .L�= 1 k- 1xk!. Then
Yn
IIYn+p - Yn ll =
n+p L k-2 1xk l k=n+l
holds for all n and p, and hence, {y, } is a Cauchy sequence of X. Since X is complete, there exists some y E X with lim = y. Clearly, 0 < and we claim that 0 < y, < y holds for each n. < Indeed, note first that the inequalities
Y11
Yn
Yn+t.
- y)+ll < for all n and p. Therefore, 0 < (y, - y )+ = 0 holds for each n. But then, y, - y < implies 0 < < y for all n. Now, since n-2 1x, l < < y, it follows that
II (Yn
Yn
limp-+xiiY,+p - y ll = 0, and so (Yn - y) v 0 = (y11 - y)+ = 0
Yn
for each n, which is impossible. Thus,
f is continuous, and so X* = X"'.
•
Section 30: BANACH LATTICES
249
vector lattice, and let Y be a vector sublattice of X . We Let X be a normed . already know that Y is a vector subspace of X , and in view of the continuity of x r; x+, it is easy to see that Y is a vector sublattice of X . We also know that the natural embedding x r; -� of a normed space into its second dual is linear and norm preserving. If X is a normed vector lattice, then x r; -� is in addition lattice preserving. Indeed, for x E X and 0 < f e X*, Theorems 30.2 and 30.3 applied consecutively in connection with the fact that X * is an ideal of x- give
(.r)+(f)
= sup{.�(g): g
0 � g < f} = sup{g(x): g e X* and 0 < g < f} = sup{g(x): g e x- and 0 < g < f} e X * and
= f(x+ ) = ;+(f).
;+ holds for each x e X, and this shows that x r; .t preserves That is, (.t)+ the lattice operations. Thus, X "sits" in X** with its norm, algebraic, and lattice structures preserved. Since the closure of X in the Banach lattice X** is the com pletion of X, it follows from the preceding that the completion of X is a Banach lattice. Summarizing, we have the following result: =
Theorem 30.11.
The completion ofa normed vector lattice is a Banach lattice.
A linear isometry T: X """* Y between two normed vector lattices that satisfies T (x v y) = T (x) v T (y) for all x, y e X is called a lattice isometry. Two normed vector lattices X and Y are said to be isomorphic if there exists a lattice isometry from X onto Y . In other words, X and Y are isomorphic if there exists a mapping from X onto Y that preserves all structures: the norm, the algebraic, and the lattice. An operator T: X """* Y between two partially ordered vector spaces is called positive if T (x) > 0 holds for all x > 0. When X is a Banach lattice and Y a normed vector lattice, then every positive operator from X into Y is necessarily continuous. (To see this, repeat the arguments of the proof of Theorem 30.1 0.) Here is an example of a classical positive operator: Example 30.12 (The Laplace Transform). In this example, we shall denote by dt the Lebesgue measure on [0, oo). Consider the operator .C: L 1 ([0, oo)) -+ Ch([O, oo)) defined by £f (s =
)
fooo
e-sr f{t)dt, s ::: 0,
where Ch([O, oo)) denotes the Banach lattice of all (unifonnly) bounded continuous func tions on [0, oo). Since le-sr f{t)i � 1/(t)l holds for all s, t ::: 0 and for each s ::: 0 the function e-sr f(t) is measurable, it follows from Theorem 22.6 that the function e-sr f(t)
Chapter 5: NORMED SPACES AND L,-SPACES
250
is Lebesgue integrable over [0, oo) for each s ::: 0. Moreover, it should be clear that I.Cf(s)l =::: 11/111 for each s ::: 0 so that .Cf is a uniformly bounded function . In addition, f(t) holds for each so ::: 0, it follows from Theorem 24.4 since lim -+ o f(t) = that
s s e-st
lim s-+so
e-sot
.Cf(s) = =
s-Ho }0rooe-st fo00e-sot lim
j(t)d =
}0roos-+limsoe-st
f(t)d t
f(t)dt = .Cf(so),
which shows that .C f is indeed a bounded continuous function. (The preceding argument also shows that lims-+oo .Cf(s) = 0.) It is now a routine matter to verify that .C is a lin ear positive operator-and hence, a continuous operator. The operator .C is known as the Laplace5 transform and plays an important role in applications. It is useful to notice that the Laplace transform is also one-to-one. To see this, assume that some function f 1 ([0, oo)) satisfies .Cf(s) = f(t) dt = 0 for each s ::: 0. l Making thechangcofvariab!ex c:-1 , we get Jd xs - f(- lnx) dx = 0 foraii s ::;: 0. In par ticular, we have Jd xk f( - lnx)dx =0 for all k 0, 1 , 2, . . . . This implies /( - lnx)=O for almost all x E [0, 1] (see Exercise 21 of Section 22). Using the fact that the function In x carries null sets to null sets (see Exercise 9 of Section 18), we infer that f = 0 a.e. Thus, .C is one-to-one. •
J000e-st
EL
=
=
We shall close this section with a remarkable convergence property (due to P. P. Korovkin6) about sequences of positive operators on C[O, 1]. Korovkin's result demonstrates the usefulness of the order structures. If T: C [0, 1] � C [0 , 1] is an operator, then (for simplicity) we shall write Tf instead of T (/). For our discussion below, C [0, 1] will be considered equipped with the sup norm 11 · 1100• For instance, lim fn = f in C [0, 1] will mean limllf, - f lioc = 0, i.e., that {/,} converges uniformly to f. Also, 1, x, and x 2 will denote the three functions of C [0 , 1] defined by 1(t) = 1, x(t) = t, and x 2 {t) = t 2 for each t e [0, 1].
Let {T,} be a sequence of positive operators from C[O, I] into C[O, 1]. /f lim Tnf = f holds when f equals 1, x, and x 2 , then lim Tn f = f holds for all f E C [0, I)Proof. Let f e C [0, I]. Our objective is to show that given E > 0 there exist constants C 1 , C2 , and C3 (depending on E) such that Theorem 30.13 (Korovkin).
5 Pierre Simon Marquis de Laplace ( 1749-1827), a distinguished French mathematician, physicist,
and astronomer. He made fundamental contributions to many fields including electricity, magnetism, planetary motion, and the theory of probability, 6Pavel Petrovich Korovkin ( 191 3-1985), a Russian mathematician. He worked in approximation theory.
Section 30: BANACH LATTICES
251
If this is done, then our hypothesis implies that II T11 f - f ll oo < 2E must hold for all sufficiently large n. Therefore, this will establish that lim T11 f = f holds for all f E C[O, 1 ] . To this end, start by observing that for each t in the interval [0, 1 ] the function 0 < g, E C [O, I ] defined by g,(s) = (s - t i satisfies g, = x2 - 2tx + t 2 1. Since each T,, is positive, T,,g, is a positive function. In particular, we have ., - 2tT11X + r-T111)(t) ., 0 < T11 g,(t) = (T,x=
(T11x 2 - x2)(t) - 2t (T,x - x)(t) + t2(T111 - l )(t) .,
(1)
.,
< II T,,x- - x-lloo + 2IIT��x - x lloo + UT,,l - llloo for each t E [0, 1]. Now, let M = II/ 1100, and let E > 0. By the uniform continuity of f on [0, 1 ] , there exists some 8 > 0 such that -E < f(s) - f(t) < E holds whenever s, t E [0, 1] satisfy Is - t I < 8. Next, observe that 2M 2M ., -E - 02 (s - t)-., < f(s) - f(t) < E + 02 (s - t)-
(2)
holds for all s, t E [0, 1]. Indeed, if Is - t l < 8, then (2) follows from -E < f(s) - f(t) < E. On the other hand, if Is - t I > 8, then (2) follows from the inequalities 2M ., - 2M 02 (s - t )2 < -2M < f(s) - f(t) < 2M < 02 (s - !)-. Since each T,, is positive and linear, it follows from (2) that 2M 2M -ET11 1 - [ l T11g1 < Tn f - f(t)T, l < ET,,l + T,g,. / 02 Next, let K
=
2 M /82
(3)
and evaluate (3) at t to get
I [ T,, J - f(t) T,,l ](!)I < ET,l( t) + K T11g,(t) =
E + E[ (T, l - 1)(!) ) + KT,,g,(t)
< E + E liT,, I - l lloo + K T,g,(t).
In particular, note that
I(T, f - f)(t)l < I [T, f - f(t)T,l](t) l + lf(t) l · I(T, l - l)(t) l < E + KT,, g,(t) + (M + E)I!T,,l - l lloo
(4)
252
Chapter 5: NORMED SPACES AND Lp-SPACES
holds for each t E
[0, 1]. Thus, by taking into account (1), it follows from (4) that •
This completes the proof of the theorem.
EXERCISES 1.
Let X be a vector lattice, and let f: x+ � [0, oo) be an additive function (that is, f(x + y) = f(x) + f(y) holds for all X, y E X+). Then show that there exists a unique linear functional g on X such that g(x) = f(x) holds for all x E x+. [IDNT: Use the arguments oftheproofofLemma 18.7 to show first that f(rx) = rf(x) holds for each x E x+ and each rational number r ::::. 0. Then define g (x) = f (x+) f(x-) for each x E X.] 2. A vector lattice is called order complete if every nonempty subset that is bounded from above has a least upper bound (also called the supremum of the set). Show that if X is a vector lattice, then its order dual x- is an order complete vector lattice. [HINT: If A is a nonempty set of positive linear functionals such that f .:s g holds in x- for each f E A, put h(x) = sup{(vj'=l/;)(x): /; E A } for each X E x+. Use the p-eceding exercise to show that h extends to a positive linear functional and that = sup A.] 3. Show that the collection of all bounded functions on [0, 1] is an ideal of IR[O, 11. Also, show that C [O, I] is a vector sublattice of IR[O, I ] but not an ideal. 4. Let X be a vector lattice. Show that a norm 11·11 on X is a lattice norm if and only if it satisfies the following two properties:
h
a. If 0 :=:: x :=:: y, then llx II :=:: II y II, and b. llx II = ll lx I ll holds for all x E X.
=
Show that in a normed vector lattice X, its positive cone x+ is a closed set. [HINT: x+ {x E X : x- = 0}.] 6. Let X be a normed vector lattice. Assume that {xn } is a sequence of X such that x, :=:: Xn 1 holds for all n. Show that if lim x, x holds in X, then the vector x is + the least upper bound of the sequence {xn } in X. In symbols, Xn t x holds. [HINT: Observe that Xu+p x, ::::. 0 for all n and p and use the conclusion of the preceding exercise.] 7. Assume that x, --+ x holds in a Banach lattice and let {En} be a sequence of strictly positive real numbers, i.e., En > 0 for each n. Show that there exists a subsequence {xkn } of {.x, } and some positive vector u such that lxkn - x i :=:: E,,u holds for each n. [HINT: Pick a subsequence {y,} of {xn} satisfying II Y11 - xll < E,2-" for each n and let u :L:,.1 (En)- 1 l y11 - xl. N'ow, use the preceding exercise to conclude that (En)- 1 1Yn - x l .:S u holds for each n.] 8. Let T : X � Y be a positive operator between two normed vector lattices. If X is a Banach lattice, then show that T is continuous. [HINT: If is not continuous, then there exist a sequence {x11} of X and some E > 0 satisfying x11 � 0 and II Tx11 1l :2:. E for each n. By the preceding exercise, the-:e exists 5.
=
-
=
T
Section 30: BANACH LATTICES
9. 10.
253
a subsequence {y11 } of {x11 } and some u E .x+ satisfying IY111 .::;: �u for each n. Now, note that I I TY11 ll .5 t II T u II holds for each n .] Show that any two complete lattice norms on a vector lattice must be equivalent. [HINT: Apply the previous exercise.] The averaging operator A: l00 -+ l00 is defined by
A (x)
=
(
x1,
x1 + x2 XI + x2 + x3 2
,
3
,···,
x1 + x2 + · · + x" ·
n
)
, · · ·
for each x = (x1. x2, . . . ) E l00 • Establish the following: a. A is a positive operator. b. A is a continuous operator. c. The vector space
{.
.. · ·�2· .. · · · ) E e00 •. V = -� = (.�1 is a closed subspace of f00• Is V
=
{ x, +x2+11 ·+x, } ..
converges in
IR.}
e:r:;?
11. This exercise shows that for a normed vector lattice X, its norm dual X* may be a proper ideal of its order dual x- . Let X be the collection of all sequences {x,} such thatx, = 0 for all but a finite numberofterms (depending on the sequence). Show that: a. b. c. 12.
13. 14.
Determine the norm completion of the normed vector lattice of the preceding exercise. Determine the norm completion of the normed vector lattice of Example 30 when X is a Hausdorff locally compact topological space. Let X and Y be two vector lattices, and let T: X -+ Y be a linear operator. Show that the following statements are equivalent: a. b. c. d.
15.
X is a function space. X equipped with the sup norm is a normed vector lattice, but not a Banach lattice. If f: X -+ IR is defined by f(x) = 2::�1 nx11 for each x = {xn} E X , then f is a positive linear functional on X that is not continuous.
T(x v y) = T(x) v T(y) holds for all x, y E X. T(x 1\ y) = T(x) 1\ T(y) holds for all x , y E X . T(x) 1\ T(y) = 0 holds in Y whenever x 1\ y = 0 holds in X. I T(x)l = T(lxD holds for all x E X.
(A linear operator T that satisfies the preceding equivalent statements is referred to as a lattice homomorphism.) Let f00 be the Banach lattice of all bounded real sequences, that is, f00 = B(JN), and let {r1, r2, . . . } be an enumeration of the rational numbers of [0, 1]. Show that the mapping T : C[O, 1] -+ f00 defined by T(f) = (f(r1), j(r2), . . ) is a lattice isometry that is not onto. Let X be a norrned vector lattice. Then show that an element x E X satisfies x 2: 0 if and only if f(x) 2: 0 holds for each continuous positive linear functional f on X. [HINT: For the "if'' part use the second formula of Theorem 30.3 to obtain f(x-) = 0 for each continuous positive linear functional f.] .
16.
Chapter 5: NORMED SPACES AND Lp-SPACES
254
Let X be a Banach lattice. If 0 =:: x E X, then show that
17.
ll x ll = sup{f(x): 0
:5
f E X* and
11/11 =
1).
Assume that cp: [0, 1 ] --+ 1R is a strictly monotone continuous function and that T: C[O, 1] --+ C[O, 1 ] is a continuous linear operator. If T(cpf) = cpT(f) holds for each f E C[O, 1] (where cpf denotes the pointwise product of cp and f). Show that there exists a unique function h E C[O, 1] satisfying T(f) = hf for all f E C[O, 1]. Iff E C[O, 1], then the polynomials
18.
19.
where ('/) is the binomial coefficient defined by (k ) = k!(,7�k)!, are known as the Bernstein7 polynomials of f. Show that if f E C[O, 1 ] , then the sequence {811) of Bemstein polynomials of f converges unifonniy to j. [HINT: Consider the sequence { T, ) of positive operators defined by T,f(x) =
20. 21.
31.
{; ( ) "
n
k
f
(k ) -;;
.
xk ( l - x)"-k .
and apply Korovkin 's theorem.] Let T: C[O, 1] --+ C[O, l ] be a positive operator. Show that if Tf = f holds true when f equals 1, x, and x2, then T is the identity operator (that is, Tf = f holds for each f E C[O, 1]). (Korovkin) Let {T, } be a sequence of positive operators from C[O, 1] into C[O, 1 ] satisfying T, 1 = 1 . If there exists some c E [0, 1] such tha1 lim T,g = 0 holds for the function g(t) = (t -c)2, thenshow that lim T, f = f(c) · 1 holdsforall f E C[O, 1].
Lp-SPACES
Our attention is now turned from the study of general normed spaces to function spaces. Many of the classical spaces in analysis consist of measurable functions, and most of the important norms on such spaces are defined by integrals. The theory of integration enables us to study the remarkable properties of these spaces. Here the classical Lp-spaces will be considered. As we shall see, they are special examples of Banach lattices. Throughout this section (X, S, 11-) '!Viii be a fixed measure space, and unless otherwise specified, all properties of functions will refer to this measure space. It Sergei Nalanovich Bemslein (1 880-1968), a Russian malhemalician. He comribu1ed proximalion of funclions and probabilily lheory 7
.
10
1he ap
Section 31: Lp·SPACES
255
P If I p> Then the collection of all measurable Definition 31.1. Let 0 < p < functions f for which l fiP is integrable will be denoted by L p(JJ.). X L p(J..L ) Lp(X) Lp(JJ.) Lp(X, S, J..L). f E Lp(JJ.), ctf L E
is important to keep in mind that if measurable for each 0.
f
is a measurable function, the
is also
oo.
to be indicated, then will be If clarity requires the measure space denoted by or even by is a vector space. Indeed, if It is easy to see that then clearly, E p(J..L) holds for all ct JR. On the other hand, the elementary inequality among the real numbers
L f E L < f- < If I , lfl , f f < < lfl f+ + L p(J..L). f E Lp(J..L) Lp(J.L. ) I f li p (! IJIP dJl r > 0 l ctf l p f. flip E L L p-norm f p I I l i f p(J..L) ct LE IR. lctl I f I p Lemma 31.2. If 0 < ).. < then
shows that p(J..L ) is closed under addition.8 Moreover, if and 0 imply that inequalities 0 to In other words, is a vector lattice. For each let
p(J..L ), then the belong and
I
=
is called the of Obviously, The number and and hold for all To obtain additional properties of the p-norms, we need an inequality.
=
·
·
1,
holds for eve/}' pair of nonnegative real numbers a and b. Proof. a b a1 > 0 b >xJ...0. x 0. f'(x) f: [0, xJ...-IRI ) > f. f x 1. x ajb x > 0. f(1) 0 < + f..x
The inequality is trivial if either or equals zero. Hence, assume and Consider the function oo) ---* defined by f(x) = - ).. + )..x Then , and so x = 1 is the for = f..( l It follows that attains its minimum at = Thus, only critical point of Now, let = to obtain the - x;. holds for all = 1 - ).. • desired inequality. 8To verify this inequality. note that lal
Ia + hi
S
(laiP)ii =:: (la lfl I
=
+ lbl'')fi , and so I
la.J + I/JI =:: 2(1al�' + lhlfl) fi . I
Chapter 5: NORMED SPACES AND Lp·SPACES
256
An important inequality between L P -nonns, known as Holder's9 inequality, is stated next.
Let 1 < p < oo and 1 < q < oo be such that * + � = 1. If f E L p(J.L) and g E Lq (J.L), then fg E L 1 (J.L) and
Theorem 31.3 (Holder's Inequality).
Proof. If f = 0 a.e. or g = 0 a. e. holds, then the inequality is trivial. So, assume
f -:j:. O a.e. and g -:j:. O a.e. Then II f li p > 0 and llgllq > 0. Now, apply Lemma 3 1 .2
with
1
).. = - , a = (lf(x)l/ ll f llp)P , p
and
l f(x )g(x)l < 1 l f(x ) I P II f II II II - P (II f II P )P
---P· g q
By Theorem 22.6,
+
b = (lg(x ) lfllgllq )q
lg(x) l q . q (IIg II q )q 1
fg E L 1 (J.L), and by integrating, we get _
_
_;;,!_Ifg_l _dJ.L
< _!._ + � = 1 . ll f llp . llgllq - p q
That is,
•
J lfgl dJ.L < II f liP · llg llq . as claimed.
For the special case p = q = 2, Holder's inequality is known as the Cauchy Schwarz 1 0 inequality; see also Theorem 32.2. The triangle inequality of the func 1 tion 11·11 P is referred to as the Minkowski 1 inequality. The details follow.
Let 1 < p < pair f, g E L p(J.L) tlze following inequality holds: Theorem 31.4 (Minkowski's Inequality).
oo.
Then for every
90no Ludwig HOlder ( 1 859-1937), a German malhemalician. He worked in group 1heory and geome1ry. He also con1ribu1ed 10 philosophical mallers concerning 1he foundalion of ma1hemaiics. 1 0 Hermann Amandus Schwarz ( 1 843-1921 ): a German maihemaiician. He worked in complex anal ysis and made several con1ribu1ions 10 1he 1heory of minimal surfaces. 11 Hermann Minkowski ( 1864-1 909), a German ma1hemalician. He s1udied eXIensively 1he geome1ric properties of convex se1s. His ideas in ma1hemaiical physics coniribUied greally 10 1he crea1ion of 1he 1heory of re1a1ivi1y.
Section 31: Lp-SPACES
257
Proof. For p = 1 the inequality is clearly true. Thus, we can assume 1 <
p <
oo. Let 1 < q < oo be such that l + l = L We already know that if f and /belcing to L p (J.J- ), then f + g likewise belongs to L p (J.J- ) . Next, observe thatsince (p - l )q = p, itfollowsthat l f + g iP- I E Lq(J.J-) . Thus, by Theorem 3 1 .3 both functions p 1/1 · I f + gl - t
p l g l · If + gl - l
and
belong to L 1 (J.J-) and we have the inequalities ·
(j
j
If I · If + g
j
p lgl · If + gl -t dJ.J- < llgllp · < II ! + gllp) � .
iP- '
dJL < l l f ll "
)
I f + g i
= 11/llp · ( llf + g llp)" , 1!.
I
'
p So, from I f + g! P = If + g l l f + gl -l < C l / 1 + lgl)lf + gj P l , we get
J
j
If + g i P d).l. <
J
p 1[1 . If + gl - l dj.J.
p lgl · If + gl -t dJ.J-
f f < 11/llp ·
The proof of the theorem is now complete.
•
Summarizing the preceding discussion: If 1 < p < oo, then a. b. c.
11 / llp > 0. lla/llp = lal · 11/llp , and II! + g llp < 11/llp + llgllp
hold for all f, g E L p (J.J-) and ex E JR. Obviously, by Theorem 22.7, 11 / llp = 0 if and only if f = 0 a.e. holds. Thus, unfortunately, the function ll · llp on L p(J.J- ) fails to satisfy the norm requirement
Chapter 5: NORMED SPACES AND Lp-SPACES
258
11/llp =
f=
0 imply 0. To avoid this difficulty, it is customary to call two that functions of L equivalent if they are equal almost everywhere. Clearly, this ), and introduces an equivalence relation on becomes a norm on the P for 1 < p < oo, is a normed space equivalence classes. In other words, L if we do not distinguish between functions that are equal almost everywhere. This means that in reality consists of equivalence classes of functions, but this should not pose a problem. In actual practice, the equivalence classes are relegated to the background, and the elements of L are thought of as functions (where two functions are considered identical if they are equal almost everywhere). An important advantage of the identification of functions that are equal almost everywhere is the following: A function of can assume infinite values or even be left undefined on a null set-since by assigning finite values to these points, the function becomes equivalent to a real-valued function of satisfies Also, it should be clear that if is a measurable function and E L < < a.e., then E is holds. In other words, (J..L) and (u' .....- iTi 'T'ha. a.+" e , +" 1 _::: y · ., l .,Hico ·-· · r l· " L L·l · C ,-.. ) I� i1 11o l11C::U C::L:tU a e jJ ..._ \J\.1 ca�11 and, in fact, a Banach lattice, as the next result of F. Riesz and E. Fischer12 shows.
p(J..L)
L p(J..L p(J..L),
11·11
L p(J..L)
p(J..L)
L p(J..L)
l g l 1/1
u
uLu
., ,...,...... .., uvuu •
g
g Lp
... .. .o .o ..,.. ..,.ovt
... lVt
1
Theorem 31.5 (Riesz-Fischer).
lattice.
Proof. Let
llgll p 11/lLJplp ;
-�
L.
- - -
r
L p(J..L). f p(J.L. )
:_
-
_
_
_�
ll·ll p
If 1 < p < oo, then L p(J..L) is a Banach
(f,,}
be a Cauchy sequence. By passing to a subsequence if nec < 2-" essary, we can assume without loss of generality that holds for each n. We must establish the existence of some E such that 0. l im 0 and f,,_ J ! for n > 2. Then Let
l f g f,=, l p = g · · + If,, fd 1 / d + 1 /:?. 1 + · , = 0 < g, t and
l /,,+1 - f,,llp f Lp(J..L)
-
holds for all n . By Levi's Theorem 22.8, there exists some 0 < t a.e. From
g, g
1 1,,+k - 1, 1 =
l +k
I: ct; - .ti- 1_)
i=ll+l
n+k
< L i=ll+l
12Emst Sigismund Fischer ( 1 875-1954), an Austrian
g Lp(J..L) E
such that
IJi - fi - l l = gn+k - g, ,
mathema ticia n. He in Ge nnany and studie d o rthononnal se quences o f fun ctio ns.
spent his s cie ntific career
Section 31: Lp-SPACES
259
it follows that { f,, } converges pointwise a. e. to some function f. Since
I f, I = !1 +
II
L(/; - /; - d i=2
< g, < g a.e.,
it follows that I f I < g a. e. holds, and hence, f E L P (JJ-). Now, in view of If - f, l < 2g a.e. and liml f,, - f i P = 0, the Lebesgue dominated convergence theorem implies lim II f - f,, ll P = 0, and the proof is finished. • A glance at the preceding proof reveals also the following interesting property of convergent sequences in L p·Spaces.
!fa sequence { f,,} c L p(JJ-), where 1 < p < oo, satisfies lim II / f,, II P = 0, then there exist a subsequence { !k, } of { f, } and some g E L p(JJ-) such that !k, ---* f a.e. and I lk, I < g a.e.for each n.
Lemma 31.6.
In general, it is not true that lim II f - f,, ll P = 0 implies f, ---* f a.e . For instance, the sequence {f,,} of Example 19.6 satisfies limii J,, II p = 0 (for each 1 < p < oo), but { f,, (x)} does not converge for any x E [0, l ] . Also, it is easy to construct an example of a sequence in an L P -space that converges pointwise to some function of the space, but fails to converge in the X L p-nonn. For instance, consider JR. with the Lebesgue measure and f,, for each n. Then j, (x) � 0 holds for each x E JR. and f,, E Lp(lR) for all n and all 1 < p < oo. On the other hand, II /, II P = 1 holds for each n and 1 < p < oo, and so {/, } does not converge to zero with respect to any Lp-norm. The next useful result gives a condition for pointwise convergence to imply norm convergence in L p-spaces. =
Theorem 31.7. Assume 1 < p < oo. Let f E L p(JJ-) and let {f,,} be a sequence
of L p(JJ-) such that J,,
�
f a.e. If lim II /n il p = II f II P' then lim II f - !n il p = 0.
Start by observing that (a + b)P < 2 P -1 (aP + b P ) holds for each pair of nonnegative real numbers a and b. Indeed, for p = 1 the inequality is trivial. On the other hand, if 1 < p < oo, then the convexity of the function g(x) = x P (x > 0) implies
Proof.
Chapter 5: NORMED SPACES AND Lp·SPACES
260
and hence, (a + b)P < 2P- 1 (a P + b P) holds. In particular, for each pair of real numbers a and b we have
Thus, 0 < 2r- I (I J,, l P + If I P) - I fn - f I P a.e., and by applying Fatou 's le�ma (Theorem 22.1 0) and using the assumption lim I [11 I P dJ-L = J If I P dJ-L, we get 2P
j
lfiP dJL =
f
J
lim [2P- l (lf, I P + lfiP) - I f,, - fiP] dJ,L n -+ oo
/ f lfiP
< lim inf 11-+00 = 2p -
l
[2P- I (If�� I P + l f iP) - lfn - fiP] dJL
r
dJL + 2p- l lim
11-+00
l JriJ�, - fi P dJ-L
+ iim inf _ 11-+00 = 2P
j
lfi P dJl - lim sup 11-+00
/
f J
l f,IP dJL
I t, - fiP dJL.
Now, since J lfiP dJ,L < oo, the last inequality yields lim sup J I f,, - fiP dJ,L < 0. Hence, lim sup J lfn - fiP dJ,L = lim inf I f,, - fiP dJL = 0, so that lim
tl-+00
/
J
If,, - / I dJl = 0. •
Therefore, lim II [11 - f II = 0 holds, as required. P
A real number M is said to be an essential bound for a function f whenever lf(x)l < M holds for almost all x. A function is called essentially bounded if it has an essential bound. Therefore, a function is essentially bounded if it is bounded except possibly on a set of measure zero. The essential supremum of a function f is defined by
11 / ll oo = inf{M > 0: lf(x)l < M holds for almost all x } . If f does not have any essential bound, then it is understood that II f lloo = oo. Observe that lf(x)l < 11 /ll oo holds for almost all x. The following properties are easily verified, and they are left as exercises for the reader. I.
2.
If f = g a.e., then 11 /ll oo = ll g ll x.· II f II ';)C > 0 for each function f, and II f II 00 =
0 if and only if f = 0 a.e.
Section 31: Lp-SPACES 3.
4. 5.
261
llaflloo = lal · 11/lloo for all a E IR. II/ + gllx < 11/lloo + llglloo · If 1/1 < lgl, then 11/lloo < llglloo·
The collection of all essentially bounded measurable func tions is denoted by L00(JJ-).
Definition 31.8.
Here again, two functions are considered identical if they are equal almost everywhere. It should be obvious that with the usual algebraic and lattice operations L00(JJ-) is a vector lattice. Moreover, according to the above listed properties, L00 (JJ-) equipped with 11· 11 00 is a normed vector lattice that is actually a Banach lattice.
L00(JJ-) is a Banach lattice. Proof. Let {f,,} be a Cauchy sequence of L00(JJ.). We have to show that there exists some f E Lx(JJ.) such that lim ll / - J,, ll 00 = 0. Since for each pair m and n we have l f,,(x)-f,,(x)l < II J,, - fnr Ileo for almost all x, it follows that there exists a null set A such that l f (x) - fm (x)l < ll f,, - fm lloo holds for all m and n and all x rt A. But then lim fn(x) = f(x) exists in lR for all x rt A, and moreover, f is measurable and essentially bounded. That is, f E Lco (JJ.). Now, let E > 0. Choose such that llf,, - fm ll oo < E for all n , .m > k. Since 1/(x) - f,,(.x)l = lim,4-oo lf,(x) - f,,(x)l < E holds for all x ¢. A and n > it follows that II/ - f,, II < E for each n > This shows that lim II/ - f,, ll oo = 0, Theorem 31.9.
,
k
co
and the proof of the theorem is complete.
k.
k,
•
It is easy to verify that each step function belongs to every L p-space. Moreover, the collection of all step functions forms a vector sublattice of every L p -space. In addition, by Theorem 25.1 this vector sublattice is norm dense in L 1 (JJ.). The next result tells us that actually the vector lattice of step functions is norm dense in every L p(JJ.) with 1 < p < oo.
For eve1y I < p < oo, the collection ofall stepfunctions is
Theorem 31.10.
norm dense in L p(JJ-). Proof. Let 0 < f E Lp(JJ.). By Theorem 17.7, there exists a sequence of simple functions such that 0 < c/J, t f a.e. Clearly, each c/J, is a step function and (/ - c/J,)P .!- 0 a.e. holds. By the Lebesgue dominated convergence theorem we .get II! - c/J, II (j If - c/J,IP dJJ.)fi .!- 0. P Since every function of L p(JJ-) can be written as a difference of two positive functions of L p(JJ-), it easily follows that the step functions are norm dense in • Lp(JJ-).
{c/J,}
I
=
Chapter 5: NORMED SPACES AND Lp·SPACES
262
In case the measure is a regular Borel measure, the continuous functions with compact support are also norm dense in each Lp(J.L) for 1 < p < oo. The details are included in the next theorem.
Let J.L be a regular Borel measure on a Hausdorff locally compact topological space X. Then the collection of all comiuuous functions with compact support is norm dense in Lp(J.L) for every 1 < p < oo. Theorem 31.11.
Proof. Clearly, every continuous function with compact support belongs to
each L p(J.L). Now, let 1 < p < oo, f E L p(J.L), and E > 0. We must show thatthere exists some continuous function g with compact support such that II f - g II < E. P By Theorem 3 1 . 1 0 it suffices to assume that f = XA , where A is a measurable set such that J.L*(A) < oo. By Theorem 25.3, there exists a continuous function g : X --+ [0, 1] with compact support suchthat f iXA -:- g l dJ.L < 2-PfP. (Notethat i XA - g l < 2 holds.)But then II XA - cllv =
(/ (/
< 2
r (/ ) I
lxA - g i P d jL
=
p l lxA - c i · IXA - g l - dJL
I
IXA - g l dJL
and the proof is finished.
;;
r I
< 2 · r ' · E = E. •
Consider JR. equipped with the measure J.L that assigns to every subset of JR. the value zero, that is, J.L = 0. Then J.L is a regular Borel measure, and obviously, any two functions on JR. are equal J.L-almost everywhere. Thus, in this case, all functions on JR. can be identified with the zero function, a situation that is not very useful. Therefore, it is desirable to deal with regular Borel measures for which distinct continuous functions are not equivalent. To do this, we need to know where the measure is "concentrated" in the space.
Let J.L be a regular Borel measure on a Hausdorff locally compact topological space X . Then there exists a unique closed subset E of X with thefollowing two properties: 1 . J.L(Ec) = 0, and 0. 2. if V is an open set such that E n V # (/), then J.L(E n V )
Theorem 31.12.
Proof. Let 0 = U { V : V is opeo and J.L(V)
=
>
0}. Clearly, 0 is an open set,
and we claim that J.L(O) = 0. To see this, let K be a compact subset of 0. From the definition of 0 it follows that there exist open sets V1 , , V,, all of measure zero, such that K c U?= 1 V; . •
•
•
Section 31: Lp-SPACES
263
Hence, 11-(K) 0. Our claim now follows from Jl-(0) = sup{,u(K): K compact and K C 0}; see Definition 1 8.4. Now, let E oc. Then E is a closed set and ,u(Ec) = /1-( 0) = 0. On the other hand, if V is an open set such that E n V =F (/), then 11-CE n V) > 0 must hold true. Otherwise, if 11-CE n V) 0 holds, then ,u(V) = 11-(E n V) + 11-CEc n V) = 0 also holds, implying V c 0 = Ec, contrary to E n V =F (/). For the uniqueness of E assume that another closed set F satisfies ( I ) and (2). From it follows at once that Fe c 0, and so E = oc C F holds. On the other hand, since 11-CO n 0, it follows from (2) that 0 n F (/). Hence, c oc = E, so that F = E, and the proof is finished. •
= =
(I)
F
=
F) =
=
The unique set E determined by Theorem 3 1 . 12 is called the support of ,u and is denoted by Supp fl-. That is, Supp 11E. If we think of the measure space as a set over which some material has been distributed, then Supp 11- represents the parts of the set at which the material has been placed. IR", then If for example, the support of the zero measure is the empty set, while the support of the Lebesgue measure A satisfies SuppA = IR". Let 11- be a regular Borel measure on a Hausdorff locally compact topological space with Supp ,u = Then two continuous real-valued functions and holds for all x E This satisfy a.e. if and only if (x ) = on follows immediately by observing that if holds for some a E then f(x) =F g(x) holds for all x in some nonempty open set V (a neighborhood of for a.e. to be true. In particular, it follows Since 11-(V) > 0, it is impossible I defines a lattice norm on the function space of that II (j all continuous real-valued functions on with compact support. In general, equipped with an p-norrn is not a Banach lattice. However, by Theorem 31.11 the following result should be immediate.
= How "large" can the support ofa regular Borel measure be? X = X
X
f=g
X.
f li P = lf lP dfl-)li L
fX. g X, a).
ff(a) =P g(a) g(x) f=g Cc:(X), X
C,(X)
Let be a regular Borel measure on a Hausdor ff locally compact topological space X with =X. Then for each 1 < < the completion ofC,(X) with the L p-norm is the Banach lattice L L X = (0, f(x) x � 0<x <1 f(x) 0 x 1 L (J1) , L2(J1 ). 1 1 g(x) 0 0 < x < g x L (J1-). L2(J1-), L (X, S, J.1.) Theorem 31.13.
11-
p
Supp 11-
oo,
p(Jl-).
In general, the p-spaces are not "comparable." As an example, let oo) with the Lebesgue measure. Then the function = if and = if > belongs to but it does not belong to On the other 1 and (x) = hand, the function = if if x > 1 belongs to but not to 1 Two comparison results of the p-spaces are presented next. The first one is for the case that is a finite measure space. -
Chapter 5: NORMED SPACES AND Lp-SPACES
264
Let (X, S, JL) be a finite measure space, and assume that 1 < p < q < oo. Then Lq(JL) c L p(JL) holds. Proof. Clearly, in this case L00(JL) c Lp(JL) holds for each 1 < p < oo. Thus, assume 1 < p < q < oo. Let r = q I p > 1 , and then choose s > 1 such that : + � = 1 . If f E Lq(JL), then clearly, lfiP E L,.(JL). Since the constant function 1 belongs to LsCJL), it follows from Theorem 3 1.3 that IJIP = IJIP · 1 E L1 (JL). That is, f E L p(JL), and Theorem 31.14.
•
the proof of the theorem is complete.
Lq(JL) c L p(JL) holds, then Lq(JL) is an ideal of
It should be observed that if the vector lattice
L P(JL).
Some important examples of
L p-spaces are provided by considering the count
ing measure on IN. In this case, the functions on and integration is replaced by summation. These 1 . ....tho.T.-. VLII""I 0 p .&II A b" "'"'n"'oS .1-J p=.::� pu'-""" t ""A u..&IU the" Lll J are denot LeU J "\,..
IN are denoted as sequences, L p-spaces are called the little A.., :+ Q
• ._,....... Y VI U.:J1 .l.l
..... ......
..,. }J
.....
00
th.,....
x 1 (x 1 , x2, ) such that E: 1 lxn IP < oo, and in this case llxllp = cr::l lxn iP)r. Similarly, loo is the vector space of all bounded •
l P consists of all sequences
•
•
-.....
t
Lll""ll
•
sequences with the sup norm.
L
The lP -spaces, unlike the general p-spaces, are always comparable. Note the contrast between the next theorem and the preceding one.
Theorem 31.15.
inclusion is proper. Proof.
If
1
Observe that if
x
=
(x1 , x2,
oo,
•
•
•
then
lp
c lq holds. Moreover, the
) belongs to some lp-space with 1 <
p < oo, then {xn} must be a bounded sequence (actually, convergent to zero), and hence, x E l00• That is, l C l00 holds for all 1 < p < oo. P ) E lp· Since E:1 lx11 1P < oo, Thus, assume 1 < q < oo. Let x = (x1, x2, q there exists some k such that lxn I < 1 for all n > k. This implies lxn l < lxn IP q for all n > k, and this shows that E: 1 lxnl < oo. Therefore, x E lq. and hence, lp c lq. I For the last part note that if we let Xn = for all n, then x = (x 1 , x2, . . . ) E lq .
but x
n - -,
¢. lp.
•
.
•
p and q in [1, oo] are called conjugate exponents if l + l = 1. We adhere to the convention 1 oo = 0, so that 1 and oo are conjugate �xpo�ents. Let p and q be two conjugate exponents. If g E Lq(JL), then it follows from Theorem 31.3 that fg E L1(JL) for each f E Lp(JL). Therefore, for each fixed g E Lq(JL) a real-valued function Fg can be defined on L p(JL) by 1\vo numbers
I
Section 31: Lp-SPACES
265
Fg
is a linear functional and, in fact, as the next result for all f E L p(Jl). Clearly, shows, a bounded linear functional.
Let I < p < oo, let q be its conjugate exponent, and let g E Lq(Jl). Then the linear functional defined by
Theorem 31.16.
1
F.c:
l gl q·
I
for f E L p(Jl) is a bounded linearfunctional on L p(Jl) satisfying F.c: II = Proof. First, we consider the case p = oo and q 1. From IF.c:{/ )1 < 11 /ll oo for each f E Loo(JJ.), it follows that F.c: is a bounded linear func < g 1! 1 holds. On the other hand, let f = S gn g, where tional and that > 0 and Sgn Sgn g(x) = I if = -1 if g(x) < 0. Then f belongs Therefore, to L 00(J1 ) and satisfies = I and = J dJJ. = =
l gl 1 ·
I Fg I I g(x) I /lloo
g(x) Fg(/)
lgl
I FNow, .c: l = wel gllconsider l· I < p < oo. By Holder's inequality
l g Il l ·
I F.c: l
< holds for all f E Lp(Jl). Hence, F.c: is a bounded linear functional, and holds. Now, let f = Sgn g . Clearly, f is a measurable.function, and fP= = holds, so that f E L p(Jl). Since fg = it follows that
ll g l q l g iP
That is,
lglq- l
l g l q,
I Fgll llgllq· Thus, I Fgll l gl q holds, and the proofis complete. >
=
•
The preceding theorem shows that for each 1 < p < oo a linear isometry g � F.c: can be defined from Lq(Jl) into L ; (JJ.), the norm dual of Lp{Jl). Observe that this isometry is also lattice preserving. Indeed, if 0 < f E L p(Jl ) , then by Theorem 30.2
(F,)+ (f)
=
=
sup( Fg(h):
0
<
I fg+ dJ1 Fg (f) =
+
f) .
= sup
{ Ihg dw 0
< /! <
f
}
Chapter 5: NORMED SPACES AND Lp-SPACES
266
Therefore, isometry.
(F.�)+ = F,�+
holds, which implies that g
H-
F.�
is also a lattice
Is every bounded linear functional on L p(J.J-) representable, as in Theorem 31 .1 6, I < p < oo. by aful1ctiol1 ofLq (J.J-)?
This is a classical result The answer is yes if of F. Riesz. A proof of this theorem, as well as some of its applications, is deferred until Section 37. Therefore, L ; (J.J-) and L q (J.J-) can be considered (under the above isomorphism) as identical Banach lattices. This is usually expressed by saying that for 1 the norm dual of L p(J.J-) is Lq(J.J-); in symbols, L ;(J.J-) Lq(JJ-). When the lattice isometry g Hfrom L1 (p.) to L�(p.) is rarely onto. The following example will clarify the situation.
< pp<=oooo
=
Fg
Let (X, S, 11-) be a measure space such that there exists a disjoint sequence of measurable sets {E,} with /1-*(En ) > 0 for each n and X = £11• Let L be the
Example 31.17.
U�1
collection of all real-valued functions f defined on X that are constant on each E,, assuming
on each
En
the value
sublattice of Loo(/1-).
f(E11), and for which
lim f(E,) exists in
Now, define a linear funciional F on L
R. Clearly, L is a vector
by F(f) = lim f(E,) for each f E L. It is clear holds for all f E L, and so is a continuous linear functional. By � can be extended to Loo(/1-) with preservation of its original norm. Denote Theorem 29.3, this extension y again. We claim that cannot be represented by a function of L (/1-). To see this, assume by way of contradiction that there exists some E Ll (/1-) satisfying F(f) = d11- for all E Loo(/1-). Let Gil E;)C and f,, XG,. Then {f,,} is a sequence of L, and = CU�'=I · ---+ 0, it F(f,) = I holds for each On the other hand, because � and follows from the Lebesgue dominated convergence theorem that J j,,g d11- -+ 0, which is impossible. � Fg from L 1 (/1-) to L � (JJ.) is not onto. Therefore, the lattice isometry • that
f
IF(f)l 1F/lloo b F F
= n.
F
1
g
fg lf,,gl F(f,l,)gl fng =
g
J
Later, we shall see (Theorem 37. 10) that if the measure is cr-finite. then the norm dual of L (J.J-) coincides with L00(J1 ) . The representation theorem for the bounded linear functionals on the l r-spaces can be proved directly.
1
Let 1 < p < oo, and let f be a continuous lineaz'[uzzctimzal ozsuclzz thatThen thez·e exists a uzzique y = (yl, ) lq (wlzere * � = f(x) l: x,y, holds for evezy x = (x1, ) lp. Proof. n, e, nth x = (x1 , x2, ) lp, Theorem 31.18. lp.
Y2, . . .
E
+
1)
00
=
It= I
x2, . . . E
For each let be the sequence having the value one at the .. . E then coordinate and zero at every other. Clearly, if
267
Section 31: Lp-SPACES
lim l x - :L�'= 1 x;e; l r = 0. Thus, f(x) = Z:::: 1 x,f(e,) holds. Let = f(e,) (y1, y2, . . . iq. If for each To complete the proof, we have to show that y = , then ly,l = l f(e,)l 11/11. so that y E ioo. let a, = y, · l y ,l q -2 if y, # 0 and a, = 0 if y, = 0. Now, for Then l a ,l P = l y, l q = a,y1 holds for all Moreover, Yn
n.
p
I
<
I < p < oo,
)
=
1
E
n.
t IYdq t,a; y; = t,a;J(e;) = f ( t, a1e1) =
<
l f l · t,a;e;
=
p
I
l fll ( t. l a;IP r
<:L�'= 1 l y;lq) ,. = <:L�'=1 ly;lq );; < 11!11 holds for. each n. This y (yl, Y2· . . .) belongs to iq and f(x) = :L:1xny,
Thus, implies that
I
-I
< oo
I
•
=
For a given normed space, it is often useful to have a characterization of its compact subsets. The Ascoli-Arzela theorem provided such a criterion for the compact subsets of (X)-spaces. Next, we shall characterize the compact sub sets of the Banach spaces L p([O, 1]). To do this, we need some preliminary discussion. Every function E L p([O, I]) will be considered defined on all oflR by =0 d>... instead of if rt [0, 1 ] . Also, for simplicity, we shall write If I < p < oo, then for E Lp([O, I ] ) and > 0 we define
C
f
t
h
f
f,1(t)
t
J: f(x) dx
[ =2h f(x) dx I
f (t) �a.b)f
t+ll
t-Il
for each E [0, I]. Note that the integral exists, since by Theorem 3 1 . 14 we have Lp([O, 1]) c L 1 ([0, I]). = f X(t,-ll.t,+ll> � Each f1, is a continuous function. Indeed, if t, � then < the Lebesgue dominated conver fX
t, g,
lg,l If I,
,_. oo
lim j,,(t,)
[ [ =2h ,_.oo f(x) dx = 2h f(x) dx = 1
tn+ll
lim
1
t+l1
-
t,-lr
t-Il
j, (t).
Chapter 5: NORMED SPACES AND Lp·SPACES
268
In particular, note that since C[O, I ] for each h > 0.
c L p([O, I]), it follows that [11
L p([O, I])
Let I < p < oo, and let f E Lp([O, I ]) Thenfor each h > 0 the continuous function [11 satisfies I a. 1/J,(t )l < (2h)-;; I f li for all t E [0, I ], and P b. I /J, I < II f I p . Proof. If p > I , then choose I < q < oo with l p + ! = I and apply HOlder's .
Lemma 31.19.
.
E
.
.
p
mequa1 1ty to get
q
p ' +h r llfh(t)IP = (2h)P } 1 · f(x) dx ,_11 f'+h 1 dx ) i < � lf(x)IP dx \2rt )P \ / t-h
I
( lt-h
[t + h lf(x)IP dx. ,
=
I
·
j'+h
2h t-h
Therefore,
[
1 t+h dx 1/J,(t)IP < 2h t-h lf(x)IP
(1)
holds for each 1 < p < oo and all t E [0, 1]. Also, ( I ) is obviously true for p and thus, statement (a) follows immediately. On the other hand, it follows from ( I ) that
h {1 lfh(t)IP dt < 2h l{1o [ ] lo [y
= 1,
'+ f J,_h lf(x)IP dx dt
I _
=
(2)
Here we have used the substitution x = t + y; see Exercise 16 of Section 22. Since the function f(t + y) is a Lebesgue measurable function on 1R2 (see Exercise 15 of Section 26), it follows from Tonell.i's Theorem 26.7 that
[ [ [;, 1/(t + y)IP dy ] dt t [{ 1/(1 y)IP dt ] dy < 2h { lf(x)IP dx . =
+
Section 31: LrSPACES
Thus,
269
(2) implies
1 ' lfh (tli' dt 1 ' Jf(x)J P dx, <
so that
ll f11 lip
<
•
II f liP holds.
A. N. Kolmogorov 13 characterized the compact subsets of Lp([O, 1])-spaces as follows:
Let l < p < oo, and let A be a closed and bounded subset of L p([O, 1]). Then the following statements are equivalent: I . The set A is compact lfor the L "-norm). 2. For each E > 0 there exists some 8 > 0 such that II f - /J, II P < E holds for all f E A and all 0 < h < 8. Proof. ( l ) ==} (2) Let E > 0. Since (by Theorem 3 1 . 1 1 ) C [0, l] is dense (for the L p -nonn) in L p([O, 1]) and A is compact, it is easy to see that there exist continuous functions fJ , . . . I f,, such that A c u:.'= I B(f;, E). By the uniform continuity of each f; , there exists some 8 > 0 such that If; (t) f;(x)l < E holds for each l < i < n whenever t, x E [0, l ] satisfy lx - t I < 8. In particular, if 0 < h < 8, then Theorem 31.20 (Kolmogorov).
lf;(t) - (f; )h (t)l =
I
2h
lr+h [J;(t) - f;(x)] dx t-h
<E
holds. Thus, (j;)h ll p < E for each l < i < n and all 0 < h < 8. Now, if f E A, then choose I < i < 11 with f E B(f; , E). By Lemma 3 1 . 1 9 we have f1r - (J;)" li p < f - f; liP < E. Therefore,
II /; -
I
I
holds for all
f E A and all 0 < h
< 8. (2) ==} (1) According to Theorem 7.8, it is enough to show that A is totally bounded (for the L p-nonn). To this end, let E > 0. Fix some h > 0 such that f - fh P < E holds for all f E A. Next, choose M > O with ll f ll " < M forall f E A . Thenby Lemma 3 1 . 1 9 i t follows that
II
II
13 Andrey Nikolayevich Kolmogorov (1903-1987), a prominent Russian mathematician. He is the founder of modem probability theory. In addition, he studied the theory of turbulent flows and worked on dynamical systems in relation to planetary motion.
270
Chapter S: NORMED SPACES AND Lp-SPACES
holds for all t E [0, 1 ] and all f E A. Let A11 = {/1111: f E A}, where
!hh( t) =
1 2h
i
t+h
t-Il
/11 (x) dx.
Clearly, l .fi,11(t)l < K holds for all t E [0, 1 ] and f E A, and hence, A11 is a uniformly bounded set. Next, we claim that the set of continuous functions A11 is equicontinuous. To see this, note that if f E A and t < s, then
-
< <
1
2 /l
-1 2h
h is+ Ji, (x) dx - i s -h.Ji,(x) dx 1-11 t +h rs-11 , r s+h i l .fi,(x) l dx + l f,(x)l dx 1-h t+h
l
J
K 1 [2K(s - t)] = h(s - t) 2h
J
holds, and this shows that A11 is an equicontinuous set. Now, by the Ascoli-Arzela theorem, A11 is a totally bounded subset of C [O, 1] (for the sup norm). Choose functions /1 , . . . , f,, E A such that for each f E A there exists some 1 < i < n with II j,, - (/; )11, lloo < E. In particular, note that
II f - J; II P � II f - Ji, II P + II J, - Ji,l, II P + II .fi," - J; II P < 2E + II .fi," - J; II P < 2E + ll .fi,, - (j; ),,ll p + II (/; )"" - ( J;)hll p + II(J;), - /; li p
< 5E.
Thus, A is totally bounded (for the L p-nonn), and the proof is finished.
•
EXERCISES 1.
Let f E
L p(J.L), and let f > 0. Show that J.L*({x E X:
lf(x)i ::: E)) ::; f- p
f
l fi P dj.L .
be a sequence of some L p (J.L)-space with 1 ::; p < oo. Show that if lim llf, - flip = 0 holds in Lp(J.L), then {f,) converges in measure to f. [HINT: Use the preceding exercise.] 3. Let (X, S, J.L) be a measure space and consider the set
2. Let { f,)
E
=
{x A : A E A11. with J.L*(A) < oo ) .
271
Section 31: Lp·SPACES
Show that E is a closed subset of L 1 (J.L) (and hence, a complete metric space in its own right). Use this conclusion and the identity /;.(A b. B) = J IXA - XB I dJ.L to provide an alternate sol ution to Exercise 12(c) of Section 14. 4. Show that equality holds in the inequality of Lemma 3 1 .2 if and only if a = b. Use this to show that if f E Lp(J.L) and g E Lq(J.L), where 1 < p < oo and + = I , then J lfgl dJ.L = 11/llp llgllq holds if and only if there exist two constants C 1 and C2 (not both zero) such that C1 lflP = C2 lglq holds. 5. Assume that J.L*(X) = I and 0 < p < q � oo. If f is in Lq (J.L), then show that 11/l lp � 11/ll q holds. [HINT: Use Holder's inequality.] 6. Let f E L 1 (J.L) n L00(J.L). Then show that:
� �
·
a. f E L p(J.L) for each I < p < oo, b. If J.L*(X) < oo, then limp-+oo 11/llp = llflloo holds. [HINT: For (b), let E > 0. Then E = {x E X: lf(x)l > 11/lloo - E } has positive measure, and (11/lloo - E) · X E � 1!1 holds.] 1 7. Let f E L 2 [0. I ] satisfy 11/112 = I and ]0 f(x) d)... (x) > a > 0. Also, for each f3 E R, let EtJ = {x E [0, I]: f(x) > {J}. If O < f3 < a , show that )... (£f3) > ({3 -a)2. 8. Show that for I < p < oo each e p is a separable Banach lattice. 9. Show that f.'X is not separable. [HINT: Consider the collection of all sequences having zeros and ones as their entries.] 10. Show that Lco([O, I ]) (with the Lebesgue measure) is not separable. [HINT: Consider the set {X[O.x}: 0 < x < I }.] 11. Let X be a Hausdorff locally compact topological space, and fix a point a E X. Let J.L be the measure on X defined on all subsets of X by J.L( A ) = l if a E A and J.L( A) = 0 if a ¢. A. In other words, J.L is the Dirac measure (see Example 13.4). Show that J.L is a regular Borel measure and that S upp J.L = {a}. 1 12. If g E C [a, b] and f E L1 [a, b], then a. show that the function F: [a, b] --+- R defined by F(x) tinuous, and b. establish the following "Integration by Parts" formula:
b1
g(x)f(x) d)... (x)
=
g(x)F(x)
b b -1
=
J�r f(t) d)...(t) is con
g'(x)F(x)dx.
u
[HINT: For (a) use Exercise 6 of Section 22. For (b) use Theorem 25.3 in connection with Lemma 3 1 .6 and the Lebesgue dominated convergence theorem.] 13. Let J.L be a regular Borel measure on R". Then show that the collection of all real valued functions on R" that are infinitely many times differentiable is norm dense in Lp(J.L) for each 1 < p < oo. [HINT: See Exercise 5 of Section 25.] Let (X, S, J.l.) be a measure space with J.L*(X) = 1. Assume that a function f E L 1 (J.L) 14. satisfies f(x) 2: M > 0 for almost all x. Then show that ln(f) E L 1 (J.L) and that J In(/) dJ.L < ln{j f dJ.L) holds. [HINT: Let t = f(x)/11! 1 1 1 in the inequality l - f � ln t < t - 1, and integrate.] oo. 15. Show with an example that Theorem 3 1 .7 is false when p =
272
16.
Chapter 5: NORMED SPACES AND Lp·SPACES This exercise shows that Theorem 3 1 . 16 is false when p = 1 and presents a necessary and sufficient condition for the mapping g 1-+ Fg from L00(J.L) into L j(J.L) to an isometry.
be
Show that for each g E L 00 (J.L), the linear functional FK (f) = J fg dJ.L for f E L I (J.L) is a bounded linear functional on L I (J.L) such that II Fcll � llgll 00 holds. b. Consider a nonempty set X and J.L the measure defined on every subset of X by J.L((/)) = 0 and J.L(A) = oo if A =I= (/). Then show that L I (J.L) = {0} and L00(J.L) = B(X) [the bounded functions on X] and conclude from this that g E L00(J.L) satisfies I Fg II = llg lloo if and only if g = 0. c. Let us say that a measure space (X, S, J.L) has the finite subset property whenever every measurable set of infinite measure has a measurable subset of finite positive measure. (A measure with the finite subset property is also called a iocally finite measure.) Show that the linear mapping g 1-+ Fg from L00(J.L) into Lj(J.L) is a lattice isometry if and only if (X, S, J.L) has the finite subset property. a.
17. Let (X, S, J.L) be a measure space. Assume that there exist measurable sets £ 1 , . . . , En such that 0 < J.L(E;) < oo for 1 < i < n, X = Ui I E;, and each £; does not contain
=
any proper nonempty measurable set. Then show that L6o(J.L) = L1 (J.L); that is, show that g 1-+ Fg from L I (J.L) toL6o(J.L)is onto. [HINT: If F E L6o(J.L), let c; = F (XE; ), g = [ci/ J.L*(E; )] XEI ' and then show that F = Fg holds.] 18. Let (X, S, J.L) be a measure space, and let 0 < p < 1.
Lf=l
a. b.
19.
·
11·11
Show by a counterexample that p is no longer a norm on L p(J.L). For each f, g E Lp(J.L) let d(f, g) = fl f - g! P dJ.L = (II/ - gllp)P . Show that d is a metric on L p(J.L) and that L p(J.L) equipped with d is a complete metric space.
[HINT: For the triangle inequality, observe that (a + b)P � aP + bP holds for every pair of non-negative real numbers a and b.] Let (X, S, J.L) a finite measure space. Then show that Theorem 3 1 . 1 0 is true for p = oo. That is, show that the step functions are norm dense in Lx(J.L).
be
20. If K is a compact subset of a metric space X, then show that there exists a regular 21. 22.
Borel measure J.L on X such that Supp J.L = K . If {fn} is a norm bounded sequence of L2(J.L), then show that J,,jn ---+ 0 a.e. Let (X, S, J.L) a measure space such that J.L*(X) = 1 . If f, g E L 1 (J.L) are two positive functi�ns satisfying f(x)g(x) � 1 for almost all x, then show that
be
23. Consider a measure space (X, S, J.L) with J.L*(X) = 1, and let f, g E L2(J.L). If
ff dJ.L = 0, then show that
Section 31: Lp·SPACES
273
24. If two functions f. g E L3 (Jl.) satisfy IIIII3 = ilg 113 = J I2g dJL = l , then show that g = III a.e. 25. For a function I E L 1 (JL) n Lz(JL) establish the following properties: a. I E L p(JL) for each 1 ::: p ::: 2, and b. limp- !+ III lip = III II! ·
26.
Assume that the positive real numbers ct 1 , a, satisfy 0 < ct1 < 1 for each i and , J,, are positive integrable functions on some measure space, "Li'= l ct; = 1 . If II · then show that 1 a. I� I:2 I,�n E L 1 (JL), and 1 b. J I�' I:2 Ir�'' dJL ::: (III! II! )a (II hill )a! · · · (ll frr ll l )an · ,
• • •
. . .
•
•
•
•
•
•
27. Let (X, S. JL) be a measure space and let {A11} be a sequence of measurable sets
satisfying 0 < JL*(A11) < oo for each n and lim Jl.*(A,) = 0. Fix l < p < oo and let I g11 = [JL*(A11)]-;j XA, for each n, where + = l. Prove that lim J Ig11 dJL = 0 for each I E Lp(JL). 28. Let (X, S. JL) be a measure space such that Jl. *(X) = 1 . For each 1 < p < oo define the set
� i
{
Cp = I E LJ (JL): Show that for each 0 < € <
j
III dJL =
I and
J IJIP dJL = 2 ) .
I there exists some 8p > 0 such that
JL*({x E X : II(x)l > €}) � 8p
29.
for each I E Cp. Let (X, S, JL) be a measure space and let 1
:::
p < oo and 0 < 11 < p.
Show that the nonlinear function 1/f: Lp(JL) � L e. (JL), where 1/f(J) = IIII7, is norm contmuous. b. If f,, � I and g11 � g hold in Lp(JL), then show that a.
.
,,
1 IJIP-171g 1 17 dJL. I J,,IP- 7Ign 1 17 dJ1. = 1lim 1 -+00 Let T : L p(J.I. ) � Lp(JL) be a continuous operator, where l < p < oo, and let 0 .s 11 ::: p. Show that:
J
30.
a.
J
If I E Lp(JJ.), then IIIP-171TIll7 E L 1 (JJ.) and
j
p 1 II -17 1TII11 dJL < IITII17 ( 11 IIIp )P .
b. If for some I E Lp(JJ.) with IIIIIp ::: 1 we have f i iiP-171TII'� dJL = 1171117 , then ITII = IITIII II .
31. Let (X, S. J.l.) be a measure space and let I E Lp(JL) for some l ::: p < oo. Show that the function g : [0, oo) � [0, oo) defined by p g(t ) = pt - l JL*({ x E X : II(x) l � t})
Chapter 5: NORMED SPACES AND Lp-SPACES
274
is Lebesgue integrable over [0, oo) and that
J IJIP dj.t
=
r
lro.oo)
g(t)dA(t) = p { 00t p-l J-L*({x E X : lf(x)l ::: t})dt .
lo
Let (X, S, p.) be a measure space and let f: X � JR. be a measurable function. If J-L*({x E X : lf(x)l 2: t}) � e-1 for all t 2: 0, then show that f E L p (J-L) holds for each 1 � p < oo. 33. Consider the vector space of functions
32.
{ f d>. = 0}. }JR. n Show that for each 1 < p < oo the vector space E is dense in L p(IR"). Is E dense in E=
34.
{f: JR." � IRI f
is a C00-function with compact support and
L 1 (JR.")? Let (0, oo) be equipped with the Lebesgue measure, and let 1 < p < oo. For each f E L p (A) let T(f)(x) = x- 1 fX. for x > 0. Then show that T defines a one-to-one bounded linear operator from L p((O, oo)) into itself such that liT II = [HINT: Assume that 0 � f E Cc((O, oo)). Then integrate by parts, and use Holder's inequality to get
J
(IIT(f)llp )P
=
pS.
flo JO' (�l lorxf(t) dt ) p dx =
1
1_P
(
lo lo
100 f(x)[ T(f)(x) ]P-l dx ( < _ foo T(f)(x) ]q
=
p p-
--
1
0
I
[
·
- _L1 11!11r · ( I IT(f)llr) � . p-
Now, use Theorem 3 1 . 1 1 . For
j11 (x) .=. x(n-'-l )p-'
if O
)
{ 00 { .xf(t) dt p d(x1-P)
liT II
<x< 1
use the sequence of functions and j,, (x) = 0 if x 2: 1.]
{!11 } defined by
CHAPTER
6
_ _ _ __ _ _ _ _ _ _ _ _ _
HILBERT SPACES The notion of norm can be thought of as a generalization of Euclid's concept of length. So, normed or Banach spaces can be viewed as finite or infinite dimensional analogues of the classical two- or three-dimensional spaces of Euclidean geometry. In the two- or three-dimensional spaces of Euclidean geometry there is, however, one more important ingredient that describes the length of a vector: it can be obtained by means of the familiar inner product. That is, the length of an arbitrary vector x is given by
llxll = �. where x · y denotes the inner product of the vectors x and y. In general, the inner productx · y = 2:7=t X; y; of two vectors x, y E lR11 can be considered as a function of two variables that satisfies certain linearity and homogeneity properties. Besides using the inner product to evaluate the norm of a vector, · the inner product also allows us to compute angles between vectors and express projections of vectors along lines and planes. For this reason, although all norms on lR.n are equivalent (Theorem 28.8), the Euclidean norm is the one that is more suitable for physical applications. Normed spaces whose norms are obtained (as above) by means of an inner product are called innerproduct spaces. If they are also complete, they are referred to as Hilbert 1 spaces. Naturally, these spaces inherit the most important properties of Euclidean spaces, and for physical applications they are the most suitable infinite dimensional analogues of Euclidean spaces. This chapter presents a brief introduction to Hilbert spaces. In the first section we study the basic properties of an inner product, and in the second section we · introduce Hilbert spaces. We discuss orthonomal bases in the third section. The final (fourth) section deals with some natural applications of Hilbert spaces to Fourier analysis. In particular, it discusses a few classical convergence properties 1 David Hilbert ( 1862-1943), a Gennan mathematician. He was a towering influential mathematical figure of modem times. In his 1900 address at the International Congress of Mathematicians. he proposed a list of 23 mathematical problems that have since stimulated and shaped the direction of present mathematical research.
275
276
Chapter 6: HILBERT SPACES
of Fourier series. Fourier analysis is today one of the basic tools of every engineer and applied scientist working in linear systems, antennas, mechanical vibrations, optics, bioengineering, various random processes, boundary value problems, and the design of filters used in the field of image and signal processing. Since the natural setting of Hilbert spaces is within the framework of complex vector spaces, in this chapter we shall deal mainly with complex vector spaces. The complex conjugate of a complex number ex = a + bz will be denoted by ex, i.e., ex = a - bz.
32. INNER PRODUCT SPACES An inner product on a real vector space X is a real-valued function of two variables
(·, ·): X
x
X � JR. such that:
(·, ·) is linear in tbefirst variable, i.e., (o:x + f3y, z) = o:(x, z) + f3(y, z)for ail x, y, z E X and aU reai numbers ex and {3; 2. (·, ·) is symmetric, i.e., (x, y) = (y, x)for all x, y E X; and 3. (x, x) > Ofor each x E X and (x, x) 0 if and only ifx 0. 1.
=
=
From (2) and ( l ), we see that every inner product on a real vector space is also linear in the second variable. A real inner product space is a real vector space equipped with an inner product. In case X is a complex vector space, a complex-valued function of two variables (·, ·): X x X � C is an inner product on X if a. b. c.
=
( ·) is linear in thefirst variable, i.e., (o:.x + f3y, z) o:(x, z) + f3(y, z)for all x, y, z E X and all complex numbers ex and {3; (x, y) (y, .x )for all x, y E X; and 0 ifand only ifx 0. (x, x) > Ofor each x E X and (x, ·,
=
.x) =
=
A complex inner product space is a complex vector space equipped with an inner product. It is easy to see that the inner product of a complex inner product space is additive in the second variable and "conjugate homogeneous" in the second variable. That is, • •
(x, y + z) = (x, y) + (x, z)for all x, y, z E X, and (x, o:y) = o:(x, y)for all x, y E X and all ex E C.
Every real vector space X can be "extended" to a complex vector space by means of its "complexification." If we define formally
Xc = X EB zX
= {x zy: +
.x ,
y E X},
Section 32: INNER PRODUCT SPACES
277
and equip it with the algebraic operations
(X + lJ) + + lJt) = X +X t + l(y + yt), (a + zf3)(x + zy) = ax - f3y + z(f3x + ay),
and
(-� 1
then X is a complex vector space-called the complexification of X . The reader should stop and verify that X with these operations is indeed a complex vector space. Also, it should be clear that the vector space X can be identified with the E X } of Xc . vector subspace Now, if X is a real inner product space with complexification X then we can define a function X x X c � C via the formula c
c
{x + zO: x ( ·}: (x + zy, Xt + zyt } = (x, X t ) + (y, Yt) + z[(y, xt ) - (x, yt )]. c.
0,
c
It is a routine matter to verify that ( 0 } is indeed an inner product on X 0,
(x + zO, Xt + zO} = (x, Xt)o ( ·, 0}
c.
and that
( ·)
That is, the inner product is an extension of the inner product from X to its complexification X Since most interesting results regarding inner product spaces are associated with complex inner product spaces and since every real inner product space can be extended (as above) to a complex inner product space, from now on, unless otherwise stated, all inner product spaces will be assumed to be complex inner product spaces. In particular, if X is a real inner product space, then we shall denote for brevity its complexification X by X again. In case we consider a real inner product space, we shall refer to specifically as a "real inner product space." Here are two examples of inner product spaces: c·
0,
c
The classical Euclidean spaces lR" provide the simplest examples of real inner product spaces. The inner product in lRn is defined by (x, y) L:�'= 1 x; y; for all vectors x = (.q , . x,) and y y,). Its complexification is simply C" whose inner product is defined by Example 32.1.
0
0 ,
= (Yt,
=
0 0 0 ,
II
(x, y) = L X; Yi
i= l
for all x, y E C" . The second example of an inner product space is the vector space C [a. b] of all con tinuous real-valued functions on a closed interval [a, b]. The inner product is given by the formula (f, g) J: f (x )g(x) dx Its complexification is the vector space of all continuous
=
0
Chapter 6: HILBERT SPACES
278
complex-valued functions defined on [a, b] and its inner product is given by
•
for all continuous functions f, g: [a, b) -+ C. In an inner product space X ,
the norm of a vector
x E X is defined by
llxll = J(x , x). Our immediate objective is to show that the preceding formula indeed defines a norm on X. To do this, we need to establish the classical Cauchy-Schwarz inequality.
Theorem 32.2 (The Cauchy-Schwarz Inequality). lfx vec ors
tra!}' t
in an inner product space, then
andy are two arbi
l(x, y)l < llxii ii Y II·
Proof. Let x and y be two vectors in an inner product space. If x
then the inequality is obvious. So, we can assume each E IR we have
A
0< =
=
0 or y = 0,
x # 0 and y # 0. Then, for
(x + AY, x + AY)
(x, x) + A(X, y) + A(X, y) + A2(y, y)
llxll2 + 2ARe(x, y) + A211yll2 < llxll2 + 2AI(x, y)l + A2II Y II2• A
Now, notice that the last expression is a quadratic in which is always non-negative. This means that its discriminant is non-positive. That is,
Hence,
•
l(x, y)l < llxll llyll, as claimed.
We are now ready to establish that the formula
norm.
Theorem 32.3.
If X
prod
llx II = J<x, x) defines indeed a
t
is an inner uc space, then theformula llxll = J<x. x )
defines a norm on (called the norm induced by the inner product). X
279
Section 32: INNER PRODUCT SPACES
llx II > 0. for each x E C and x E X , then
Proof. Start by observing that
and only if x
=
0. Now, if a
and so I axil = l ct. l inequality to get
llx II = 0 if
llxll. Finally, for the triangle inequality use the Cauchy-Schwarz llx + Yll 2
This implies
E X and that
y,
(x + x + y ) = (x, x) + (x , + x) + = (x , x) + 2 Re(x , + < llxll2 + 21(x, y )l + IIYII 2 < llxll 2 + 211xii iiYII + IIYII2 = (llxll + llyll)2. =
y) (y, (y, y) y) (y, y)
llx + II .:S llx II + lly 11. and the proof is complete.
•
y
The inner product is a jointly continuous function . .
The inner pr o duct is a jointly continuous function . with respect to the induced nprm. That is, llxu - xll -+ and llYn - Yll -+ ==> (xu. Yu ) -+ (x , y).
Lemma 32.4.
0
0
Proof. The conclusion follows from the inequality
l(xu. Yu ) - (x, y ) l = l(xu -X, Yu ) +(x, Yu - y ) l and the fact that if II Yu
-
y II -+
0, then
<
IIYull ll xu -xll + llxii iiYu - Yll.
{ II Yull} is a bounded sequence.
•
A well known theorem of Euclidean geometry states that the sum of the squares of the sides of a parallelogram equals the sum of the squares of its diagonals. This fact is also true in any inner product space and is known as the
law.
Theorem 32.5 (The Parallelogram Law).
inner product space, then
parallelogram
If x and y are two vectors in an
Chapter 6: HILBERT SPACES
280
Proof. If x and y are two vectors in an inner product space, then note that
llx + yll2 + llx - y ll2
=
(x + y, x + y) + (x - y, x - y) = [(x , x) + (x, y) + (y, x) + (y, y)] + [(x, x) - (x, y) - (y, x) + (y, y)] 2(x, x) + 2(y, y) 2( 11x ll 2 + ll y ll 2),
= =
as claimed.
•
For a given norm 11·11 on a vector space X there exists at most one inner product on X that induces the norm 11 ·11 . To see this, assume that two inner products (·, ·) and {·, ·} on X induce 11· 11 . That is, assume that (x, x) {x , x } llx ll2 for all x E X. In particular, we have
= =
(x + y, x + y)
and therefore
= (x, x) + (x, y) + (y, x) + (y, y)
=
=
{x + y, x + y} {x, x} + {x, y} + {y, x } + {y, )'} ,
(x, y) + (y, x) = {x, y} + {y, x} for all x and y. If the vector space X is real, then the latter easily implies
(x, y) = {x, y}
for all x and y. If X is a complex vector space, then replacing x by z x in (*), we get z (x, y) - r (y, x) = z {x, y } - z {y, x} or
x=
(x, y) - (y, )
{x, y} - {y, x}
=
=
2 {x, y}, or (x, y) {x , y} for for all x and y. Adding this to (*) yields 2(x, y) all x and y. This establishes that a given norm is the induced norm of at-most one inner product Which norms are induced by inner products? Remarkably, as we shall see next, the ones that satisfy the parallelogram law. norm 11· 11 on a vector space is induced by an inner product if and only if it satisfies the parallelogram law, i.e., zf and only
Theorem 32.6. A
if
Section 32: INNER PRODUCT SPACES
281
holdstheforcaseallofvectors x andy. Moreover, the inner product ( ·, ) that induces 11 · 11 a real vector space is given by in ·
and in the case of a complex vector space by (x, y) = 4
Proof. Assume that a real Banach space X satisfies the parallelogram law. We
shall show that the formula 1
(x, y) = 4 (llx + yll- - l x -Yll2), x (x, x) = 11 · 11 . l 1 2 >0 x = 0. (x, y) = (y, x) '
is an inner product which induces the norm Clearly, for each Also, it should be clear that 0 if and only if and for all and y. Next, we shall show that the function (·, ·) is additive in the first variable. To do this, note first that
x (x,x x) = 4(u + v, w) +4(u - v , w) = ( l u + v + wll2 - l u + v- wll2) 2 + (l l u - v + wll2 - l u -v - wll ) = (llu + w + vll2 + l u + w- vll2) -(llu - w + vll2 + l u - w- vll2) = (21lu + wll2 + 211vl 2)-(211u - w!l2 + 21 1 vll2) = 2(11u + wll2 - l u - wll2) = 8(u, w). all u, v, w (u + v , w) + (u - v, w) = 2(u, w). � (x+y), v = 4 <x-y) , (2u, w)=2(u, w). =u v w=z (x, z) + (y, :) = (u + v, z) + (u - v , z) = 2(u, z) = (2u, z) = (x + y, z), ( ) 18.7, (rx, y) = r(x,y) r x, y X. ( ), Thus, for
When and
E X, we have
Now, letting u =
(••) yields in (••), we get
which is the additivity of in the first variable. we can establish that Now, as in the proof of Lemma as defined above, holds for each rational number and all E Since ·, ·
·,
·
282
Chapter 6: HILBERT SPACES
is a jointly continuous function (relative to the norm 1 1 · 11 ), it easily follows that (ax, y) = a(x, y) for all a E JR. and all x, y E X. This completes the proof of the real case. We leave the complex case as an exercise for the reader. • In an inner product space the notion of orthogonality is introduced by means of the inner product. Two vectors x and y are said to be orthogonal, in symbols x l. y, if (x, y) = 0. A set of vectors S in an inner product space is said to be an orthogonal set if 0 ¢ S and any two distinct vectors of S are orthogonal. An orthogonal set consisting of unit vectors is called an orthonormal set. Notice that if S is an orthogonal set, then the set of vectors {xI llx II: x E S} is automatically an orthonormal set. In the case of a real inner product space, we can also define the angle between two vectors. Indeed, if x and y are two non-zero vectors in a real inner product space, then (in view of the Cauchy-Schwarz inequality) there exists a unique angle f} satisfying 0 < f} < 180° and cos f}
=
(x, y) _____ llxll llyll
The angle f} is called the angle between the vectors x and y. So, in the real case, two vectors are orthogonal if and only if the angle between the vectors is (we also define the angle between the zero vector and any other vector to be One of the olde�t and most famous theorems in mathematics is the Pythagorean2 theorem. It states that the square of the hypotenuse of a right triangle equals the sum of the squares of its two legs. This result also carries over to inner product spaces.
goo goo).
If in an inner product space the vectors x1, X2 , . . . , x, are pailwise orthogonal, then Theorem 32.7 (The Pythagorean Theorem).
2Pylhagoras of Samos (ca 58�500 BC), a greal Greek philosopher and malhemalician. Born on lhe Greek island of Samos, he is crediled with many discoveries in malhemalics and science and was lhe firsllo associale number lheory wilh musical sounds. Pylhagoras moved lo lhe Greek cily of Crolona in soulhem llaly (a well known cily in ancienllimes for ils medical school and for many greal aihleles who won in the ancienl olympic games) and crealed lh� famous Pylhagorean school, which was in essence a secrel brolherhood. Following Pylhagoras' maxim "everylhing was nol lo be lold lo everybody," whal was done and 1aughl among lhe members of lhe brolherhood was kepi in profound secrecy from lhe ou1Side world. Due lo lhis secrecy, the exlend of knowledge and discoveries of the Pylhagorean school is fragmenled and highly speculalive.
Section 32: INNER PRODUCT SPACES
283
, x11 are pairwise orthogonal in an inner pro Proof. If the vectors x 1 , x2 , duct space, i.e., (X;, Xj) for -j:. j, then note that
=0 i ( II II ) II = = 8 II •
2
B
x;,
f;
xj
•
•
(x;, Xj) f;
II = B = �II (x;, x;)
as claimed.
2 llx; ll , •
An immediate consequence of the Pythagorean theorem is that non-zero orthog onal vectors are linearly independent.
In an inner product space any set of non zero pairwise or thogonal vectors is linearly independent. Proof. x 1 , x2, , x11 'L�'= 1 A;X; = 0. AIXI, A2X2, A2X11 'L7= 1 l A; 1 2 llx; 11 2 = A; = 0 i, x1 , x2 . . . , X11 'L7= 1 A;X; 11 2 = 0. Corollary 32.8.
•.. are non-zero pairwise orthogonal vectors , and let Since the vectors are also pair wise orthogonal, it follows from the Pythagorean theorem that This implies for each so that the vectors II • are linearly independent.
Assume that
•
.
.
Recall that if {A; };e/ is a family of non-negative real numbers, then the symbol L ; e/ A; denotes the supremum of the collection of all possible finite sums of the A;. That is, if F denotes the collection of all finite subsets of then
I,
where the supremum is taken in the extended real numbers.
Lemma 32.9. {Ai };eJ A; Lief A; < oo, L; e 1 A; Ai > �}, then Fn is a finite set (why?) and moreover, Proof. If F11 Ai > 0} U:1 F11 • Hence, A; > 0} is at-most countable. •
Ithen f # i0s holds a family of non-negative real numbers and for at most countably many indices i (in which case is either a finite sum or a series). = { i E I: = {i E I: {i E I:
One of the most basic properties regarding orthonormal sets is Bessel 's3 in equality. As we shall see later on, almost every aspect regarding the structure of orthonormal sets depends upon the next result. 3Friedrich Wilhelm Bessel ( 1784-1846), a notable Gennan astronomer and mathematician. By introducing new methods, he detennined the positions of several stars and planets quite accurately. He is also well known for a special class of functions which is today the indispensable tool of every scientist working in applied mathematics, physics, or engineering.
Chapter 6: HILBERT SPACES
284
Theorem 32.10 (Bessel's Inequality).
vectoz·s in an inner product space, then
If {x; };e/ is an oz·thozwl"malfamily of
L l(x, x;)l2 < llxll2 ie/
holdsfor each vector x. In particulaz; for each x all but an at most countable number of the (x, x;) vanish. Proof. Fix some vector x in an inner product space and let x 1 , x , . . . , Xn be 2
a finite orthonormal set. Then we have
II
II
n
II
- llxll2 - L(x, x;)(x;, x) - L(x, (x, Xj)Xj) + L L((x, x;)x;, (x, Xj)Xj) j=l i=l j=l i=l n n - llxf - L l(.x, x; )l2 - L l(.x, Xj )l2 + L l(x, x; )l2 i=l j=l i=l
1
II
llxll2 - L l(.x, Xi)l2. i=l Hence, L:�1 l(x, x;)l2 < ll.x ll2 . Now, if {x; }ie/ is an arbitrary orthonormal set, then the above conclusion easily implies L; e1 l(x, xi)l2 < llx 11 2 . In particular, • (x, x;) :f:. 0 must hold true for at most countably many indices i.
There is a standard procedure called the Gram4-Schmidt5 orthogonalization process for converting a countable (or .a finite) set of linearly independent vectors into an orthogonal set of vectors. As usual, let us use the symbol Span{z 1 , • • • to denote the vector subspace generated by the vectors 1 , . . . , Zn .
z
,
z11}
Let x1 , x , be a sequence of lineaz·ly independent vectors in an izmez· product space. If2we define the sequence ofvectoz·s y1 , y2 inductively by Theorem 32.11 (The Gram-Schmidt Orthogonalization Process). .
.
•
• . . .
� Yl = X) and Yn+l = Xn+l - L., i=l
(xll+l• y; ) y; for n = Ij y,. j1 2
1 , 2, . .
.
,
4Jorgen Pedersen Gram (1850-1916), a Danish mathematician. He taught at the University of Copenhagen and was the chairman of the Danish Insurance Council. 5 Erhard Schmidt (1 876-19�9), a German mathematician. He was a student of David Hilbert and worked on integral equations and the geometry of Hilbert spaces.
285
Section 32: INNER PRODUCT SPACES
then we have the following: 1 . The set {Yt , yz, . . . } is orthogonal, i.e., y, # Ofor each n and (y; , yj) = 0 for i # j, and 2. Span{Xt , Xz, . . . , Xu} = Span {Yt , yz, . . . , Ynl for n = 1 , 2, . . . . Moreover, the vectors Yt, Yz, . . . are (aside of scalar factors) uniquely determined in the sense that if Zt , zz, . . . is another sequence of orthogonal vectors satisfying Span{Xt , xz, . . . , X11} = Span{zt , Zz, . . . , Z11} for each n, then for each n there exists a 1wn-zero scalar A11 such that z11 = A11 Yn . Proof. First, we shall prove ( 1) and (2) by induction. For n = 1 , the claims ( 1 ) and (2) are obvious. So, for the induction step, assume that the vectors y1 , • • • , Yn
(as constructed by the above method) are all non-zero, pairwise orthogonal and satisfy
(t)
Span{Xt , Xz, . . . , Xn} = Span{yt , yz, . . . , Ynl· Now, de fine the vee tor I < j < n we have
Yn+ l = Xn+ l
-
"'II �i=l
(X,+J,J;)
IIYill2 y;
and note that 1or each &
are
so that the vectors Yt , . . . , y11, Yn+ l are pairwise orthogonal. Now, using (t), the • de�-mear1y m . l = Yn l + '\'II (Xu+ I 2 . 1 y Xn 1"dent"t + + �i=l IIYd.y;) l y; an d that X t , . . . , Xn+l pendent, we easily infer that Yn+ l # 0 and Span{Xt , xz, . . . , Xn, Xn+tl = Span{yt, yz, . . . , y11, Yn+tl- This completes the induction and the proof of ( 1 ) and (2). To establish the uniqueness of the sequence {y11}, assume that another sequence {z11} of pairwise orthogonal vectors satisfies .
Span{xt . X2, . . . , xn} = Span{yt , yz, . . . , yn} = Span{z t , Zz, . . . , zu} for each n. Clearly, z11 # 0 for each n. We shall prove by induction that for each n there exists a non-zero scalar A11 such that z11 = A11Y11 • For n = 1 , the conclusion is obvious. So, assume that there exist non-zero scalars A t , . . . , A11 such that z; = A;Y; for each i = 1 , . . , n. From .
Span{yt , yz, . . . , Yn. Yn + t l = Span{zt, Zz, . . . , Zn, Zn+d,
Chapter 6: HILBERT SPACES
286
1
we can write Z n+l = 'L-�':1 CiYi with c,+l # 0. Now from Zn+l - Cn+lYn +l = 'L�'= 1 ciYi and the fact that the vectors z,+l and Yn+l are both orthogonal to Span{y,, Y2· . . . , y, }, it easily follows that z,+l - c,+I Y,+ I l. :,71 - c, +lY,+ l · Thus, z,+l - Cn+l Yn+l = 0 or Zn+l = Cn+l Yn+l· This completes the induction and the proof of the theorem. •
EXERCISES 1. 2.
Let
c 1 , c2
, en ben (strictly) positive real numbers. Show that the function of two
variables(·, ·): IR" on •
.
.
.
IR.n .
x
IR.n � IR, defined by (x, y) = L:f=l CiXi Yi, is an inner product
Let (X, (· , ·)) be a real inner product vector space with complexification X c. Show that the function { · , ·): X c x Xc � C defined via the formula
{X + 1 )' , X I + 1 )'I ) = (X, .q ) + ()' , )'I ) + 1 [(y, X I ) - (X. )' J )].
A ,..,.. ..h,..... hot •h .., ,.. .....,. :.., . .-�..:1 h •I.e :-- - ---..l -� \I ' 7 ) " C• � a ro.Jv1 "'""" ,.." t""' L 'e l1V111l '''d U\..o\..U UJ L11 � '""e1 plUUu""L '"'"" ""' y v ' "0 v X h· on ';nm"'r "r"duc• ••
•
on X c is given by
llx + 'YII = Jcx , x) + (y , y) = Cllxll 2 + IIY if) ! .
3.
Let n be a Hausdorff compact topological space and let J1. be a regular Borel mea- ·
sure on n such that Supp J1.
defined by
= n. Show that the function ( ·): C(Q) X C(Q) � IR, · ,
(f, g) = £ tg dj1.,
is an inner product. Also, describe the complexification of C(Q) and the extension of
4.
the inner product to the complexification of C(Q).
Show that equality holds in the Cauchy-Schwarz inequality (i.e.,
if and only if
5. 6.
7.
8. 9.
x and y are linearly dependent vectors.
j(x. y)l = llxll llyll)
llxll = sup{ j(x. y)j: IIYII = 1}. holds if and only if llx + y 11 2 =
lfx is a vector inan innerproduct space, then.showthat
x ..L y llx ll2 + llyl! 2 • Does llx + yll 2 = llxll2 + IIYII 2 in a complex inner product space imply
Show that in a real inner product space
X j_ y?
{xn } in an inner product space satisfy l!xll . Show that x, � x.
Assume that a sequence
llxn II �
S be an orthogonal subset of an inner product complete orthogonal subset C such that S c C . Let
(x,.
x) � ll x 112
space. Show that there exists a
Show that the norms of the following Banach spaces cannot be induced by inner products: a. b. c.
10.
and
The norm
llxll = max{ lx t l , lx2 j, . . . , lxn l l on lR.n .
The sup norm on C[a, b].
The L p-norrn on any
Prove Theorem 32.6
L p(J.L)-space for each 1
for complex normed spaces.
� p � oo with p :;f 2.
Section 32: INNER PRODUCT SPACES
287
Let X be a complex inner product space and let T: X ---+ X be a linear operator. Show that T = 0 if and only if (T X '( ) = 0 for each X E X. Is this result true for real inner product spaces? 12. If {xn } is an orthonormal sequence in an inner product space, then show that Iim(x11 , y) = 0 for each vector y. 13. The orthogonal complement of a nonempty subset A of an inner product space X is defined by
11.
•
Al.
=
•
{.x E X: x .l.. y for all y E A } .
We shall denote (AJ..)l. by AJ..J.. . Establish the following properties regarding orthog onal complements:
Al. is a closed subspace of X, A c Al.l. and A n Al. = {0}. b. If A £ B , then B l. £ A l.. c. Al. A l. = [.C(A)]l. [.C(A)] l., where .C(A) denotes the vector subspace generated by A in X. d. If M and N are two vector subspaces of X, then Ml.l. + Nl.l. c (M + N)J.J.. e. If M is a finite dimensional subspace, then X = M E9 M l. . a.
=
14.
=
Let V be a vector subspace of a real inner product space X. A linear operator L: V is said to be symmetric if (Lx. y) = (x, Ly) holds for all x, y E V. a.
---+
X
Consider the real inner product space C [a, b] and let
V
=
.,
{ f E C -[a, b]: f(a) = f(b) = 0}.
Also, let p E C 1 [a, b] and q E C[a, b] be two fixed functions. Show that the linear operator L: V ---+ C[a, b], defined by
L (f)
=
(pj' )' + qf,
is a symmetric operator. Consider JR" equipped with its standard inner product and let A: JR" ---+ JR" be b. a linear operator. As usual, we identify the operator with the matrix A = [aij] representing it, where the jth column of the matrix A is the column vector Ae j . Show that A is a symmetric operator if and only if A is a symmetric matrix. (Recall that an n x n matrix B = [bij ] is said to be symmetric if bij = bji holds for all i and j.) c. Let L: V --+ X be a symmetric operator. Then L extends naturally to a linear operator L: Vc = {x + 1y: x, y E V } ---+ Xc via the formula L(x + 1y) = Lx + 1 Ly. Show that L also satisfies (Lu, v) = (u, Lv) for all u , v E Vc and that the eigenvalues of L are all real numbers. Show that eigenvectors of a symmetric operator corresponding to distinct eigen d. values are orthogonal. 15.
L:�'= 1
Xi Yi for all Let ( · , ·) denote the standard inner product on JR", i.e., (x, y) = x. y E JR". Recall that an n x n matrix A is said to be positive definite if (x, Ax) > 0 holds for all non-zero vectors x E JR". Show that a function of two variables ( ·, · } : JR" x JR" ---+ 1R is an inner product on JR" if and only if there exists a unique real symmetric positive definite matrix A such
Chapter 6: HILBERT SPACES
288 that
(x, y) = (x, Ay)
holds for all x, y E IR.n . (It is known that a symmetric matrix is positive definite if and only if its eigenvalues are all positive.)
33.
HILBERT SPACES
We start by introducing the notion of a Hilbert space.
Definition 33.1. A Hilbert space is an inner product space which is complete
under the norm induced by its inner product.
Again, we distinguish Hilbert spaces into real and complex. If space, then its complexification induced nonn is given by
1
He H EB H =
H
is areal Hilbert
is a complex Hilbert space whose
Here are two classical examples of Hilbert spaces:
Example 33.2.
The real Banach space f.2 is a real Hilbert space under the inner product
defined by
(x, y) = L: x, yn 00
n=l
.
for all x = (xt . x2 • . . . ) and y = ()'J , Y2· . . . ) in f.2. Its complexification, which we shall denote by f.2 again, consists of all square summable sequences of complex numbers. In this case, the inner product satisfies 00
(x, y) = L X,Yn· n=l
for all x =
(xt. x2
• . . .
) and y = (yt . Y2· . . .) in f.2.
•
Example 33.3.
The real Banach space L2(1J.) is a real Hilbert space. Its complexification, which we shall denote again by L2(1J.), consists of all complex valued measurable functions such that 1/1 2 dJ.L < oo holds (where, as usual, two functions are identical if they are equal almost everywhere). Then L2(J.L) is a complex Hilbert space with the inner product defined by
J
j
(J, g) = f(x)g(x) dJ.L. •
Section 33: HILBERT SPACES
J-L)
Example 33.4.
289
p:
Let (X, S, be a measure space and let X � (0, oo) be a measurable function-called a weight function. Then the collection of measurable functions
L2(p) L2(p) � by (f,g) fpfgdJ-L, then (·, ·) is a well-defined inner product. Indeed, since f E L2(p) is equivalent to having .JPf E L2(J-L), it follows from Holder's inequality that is clearly a real vector space. Now, if we define (·, ·):
JR.
x
=
2 pi ( ) pfg dJ-L Pifl2 l f g ( d dJ-L) I f J-LI f I
:::
2
2 I
<
00,
L2(p) L2(p) L2 p) ( L2(v)
is and so (-. ·) is well-defined. We leave it as an exercise for the reader to verify that a Hilbert space. As usual, we shall denote the complexification of by again. for the measure is exactly the Hilbert space The reader should also notice that 1\.IJ. � [0, oo] defined by
L2(p)
v:
v(A) ip(:r)dJ-L(:c) =
for each
AEA
•
w
The norm completion of an inner product space is a Hilbert space.
The norm completion X of an inner product space i s a of space. Moreover, Hilbert X and two sequences and satisfy and in X, then Theorem 33.5. X11 � x
X
if x, y e Yn 4 y
{x11 }
{y11}
X
(x , y) = lim (x,, y11 ). n--+oo
Proof. By Theorem 29.7 we know that X is
Banach space. We leave it as an exercise for the reader to verify that the formula ( · , · ) as defined above is a well-defined inner product that induces 11·11 on X. • a
Recall that in a normed space the distance from a vector x to a set A is defined by
d(x,
A ) = inf{llx - yll: y e A } .
Chapter 6: HILBERT SPACES
290
Our next important result asserts that a closed convex subset of a Hilbert space contains a unique vector of smallest distance from a given vector.
If A is a non-empty closed convex subset of a Hilbert space H, then for each x E H there exists a unique y E A such that
Theorem 33.6.
d(x, A) = llx - y ll . Proof. Let A be a non-empty closed convex subset of a Hilbert space H, let
x E H, and put d = d(x, A). Choose a sequence {y,} c A such that d = n-+oo lim llx - y, ll .
We claim that the sequence {y,} is a Cauchy sequence. To �ee this, note first that by the Parallelogram law, we have
llYn - Ym 11 2
ll(y, - X) - (Ym - x)ll 2 = 2 11Yn - x ll 2 + 211Ym - x ll 2 - ll(y, - x) + ( Ym - x)f y, Ym 2 = 211Yn - x ll 2 + 2 11Ym - x ll =
4
Since A is convex, it follows that
y,�y"'
xf.
1 ;
E A, and so d < II v,�y"' - x 11 . Therefore,
This shows that {Yn} is a Cauchy sequence. The completeness of H guarantees that {y,} converges to some vector y in H. Since A is closed, y E A and, clearly, d = d(x, A) = llx - y ll . For the uniqueness of y ' assume that another vector z E H satisfies nx - z II = d(x, A). Since A is a convex set, !CY + z) E A, and by the Parallelogram Law
d2
<
=
1
x - 2 (y + z)
2
1
4 11 cx - y) + (x - z)ll
=
2
� [211x - Yll 2 + 211x - zlf - ll(x - y) - (x - z)ll 2] 1
= -(2d2 + 2d2 - liz - yll 2 ) 4
.
1 = d2 - 4 ll z - y f < d2 • This implies
! II y - z 11 2 = 0 or y = z, and the proof is finished.
•
291
Section 33: HILBERT SPACES
Now, let X be an inner product space. If A is a non-empty subset of X, then the orthogonal complement A.l of A is the set of all vectors that are orthogonal to every vector of A. That is, A.l = {x E X: x y for all E A}. From tl)e linearity and continuity of the inner product it should be clear that A is always a closed subspace of X satisfying A.l = (A ).l and A n A.l = {0}; see also Exercise 13 of Section 32. Remarkably, when X is a Hilbert space and A is a closed subspace, then A together with A.l span the whole Hilbert space. Theorem 33.7. If M is a closed subspace of a Hilbert space H, then we have H = M EB M.i. Proof. Since M n M .l = {0}, it suffices to show that every vector x E H can written in the form x = y + z with y E M and z E M.i . So, letx E X. Since M is a closed and (as a vector subspace) convex set, there exists (by Theorem 33.6) a unique vector y E M such that d(x, M) ll x 11 . Let z = x - y and note that x y + z. We shall finish the proof9 by establishing that z M. To this end, let w E M and write (z , w) = r e' with 0. Then, for each real A we have j_
y
.l
be
=
-
=
y
.l
r >
liz II�.,
<
., ., liz + Ae 19 w ll- = li z II- + 2rA + A- ll w ll�. .,
.,
.:S
This implies 2rA + A2 11w ll 2 0 for each real A, and so 2r -A II w ll 2 for each A < 0. Hence, < 0, and therefore = 0. Consequently, (z, w) 0. In a Hilbert space the dense subspaces have a nice characterization in terms of an orthogonality condition. Corollary 33.8. A vector subspace M of a Hilbert space is dense if and only if the zero vector is the one and only vector orthogonal to M. Proof. Let M be a vector subspace of a Hilbert space H and let be its norm closure. Assume first that M = H and let x .l /vf. Pick a sequence {xn} C M such that x11 -+ x and note that 0 = (x11 , x) -+ (x, x) implies (x, x) = 0, or x = 0. For the converse, assume that M.l = {0}. Since (M).l = M.l, it follows from Theorem 33.7 that = M EB {0} = M EB(M).l = H, and soMis dense in H. Since the inner product is ajointly continuous function, it follows that every vec tor in an inner product space X defines a continuous linear functional f\': X C >
r
r
•
=
/vi
/vi
y
•
-+
Chapter 6: HaBERT'SPACES
292
via the formula
/y(x) = (x, y). X X
If is a Hilbert space, then (as we shall see next) all continuous linear functionals on are of this form.
If H is a Hilbert space and f: H C is a conti nuous linear functional, then there exists a unique vector y e H such that f(x) = (x, y) holdsfo all X E H. Moreover, we have f = y Proof. Let f: H C be a continuous linear functional on a Hilbert space and let M be its kernel, i.e., M = Ker f = f- 1 ({0}) = {x E H: f(x) = 0}. Since f is a continuous linear functional, it follows that M is a closed subspace. If M H, then y = 0 satisfies f(x) = (x, y) = 0 for each x e H. So, we0can assume that M is a proper closed subspace of H. Then, there exist some x e H with . . Now, notice f(x0) = and (by Theorem 33.7) some non-zero vector w E M.1 that ifx E H, then x - f(x )xo E M, and so(x- j(x)xo , w) = O or f(x)(xo, w) = (x, w). This implies f(w)(xo, w)=(w, w) 0, and so (xo , w) '# 0. Now, notice that the vector y = wj(w, xo ) satisfies f(x) = (x, y) for all x e H. Forthe uniqueness notethatif(x, y) = (x, y1)foreachx E H , then (x , y-y1) = 0 or 0 for each x E H , and so, by letting x = y - Y1, we get (y - Y1 , y - Y1) Theorem 33.9· (F. Riesz).
�
II
r
II II·
II
�
=
I
>
=
y = y1 • Finally, to establish the norm equality, note first that the Cauchy-Schwarz inequality
lf(x)l = ICx, y)l < l x li l Y
II
implies II II ::5 II y On the other hand, if y # 0, then letting = y I II y II, we see • = IIYII· that = ICY/IIYII . y)l = IIYII· Thus, = 1 and so >
l xl
f
1. 1 /1 lf(x)l
x
l fll
fy
If H is a Hilbert space, then Theorem 33.9 shows that a function y 1--7 where y) can be defined from H onto H * . In view of the properties
fy(x) = (x,
fy + fz = fy+z• afy = fay
and
l fyll =
IIYII,
it easily follows that this mapping is a "conjugate" linear isometry from H onto
Section 33: HILBERT SPACES
293
H * . By means of this isometry, we can establish that Hilbert spaces are reflexive Banach spaces.
Corollary 33.10. Proof. Let
Every Hilbert space is reflexive.
H be a Hilbert space and let F : H*
functional. Define ¢: H
� C be a continuous linear � C via the formula ¢(y) = F(/y). From
¢(y + z ) = F( /y :) = F(/y + f:) = F( /y ) + F(fz) = ¢(y) + ¢(z ) , + ¢(ay ) = F( /ay) = F(ii/y) = aF(/y) = a F( /y) = a cp(y) , and l¢(y ) l = IF( /y)l = I F(/y)l < II FII II /y ll = II FII IIyll, we see that ¢ E H*. So, by Theorem 33.9, there exists a unique x E X such that (y, x) = ¢(y) = F (/y) for all y E H. This implies .t(/y) = /y(x ) = (x, y ) = F(/y) for each y E H, and so F = .i. This shows that the natural embedding x � .i of H into its double dual H** is onto, and consequently H is a reflexive Banach space. • We shall close this section with a few more facts concerning orthogonal and orthonormal sets in Hilbert spaces. Let us say that an orthogonal set S of an inner product space is complete if x ..L s for each s e S implies x = 0. In view of Corollary 33.8, it should be noted that an orthogonal set S in a Hilbert space is complete if and only if the vector space generated by S is dense.
Every inner product space has a complete orthogonal set (and hence, it also has a complete orthonormal set).
Theorem 33.11.
Proof. The conclusion follows from Zorn's lemma by observing that the col lection of all orthogonal sets, ordered by inclusion, has a maximal element. To finish the proof, notice that an orthogonal set is maximal if and only if it is complete. • It is useful to note that if S is a complete orthogonal subset of a real Hilbert space H, then S is also a complete orthogonal set in the complexification Hc = H ffi z H of H. Indeed, if some x + z y satisfies x + z y ..L S, then (x + z y, s + z 0) = (x, s) + z (y, s) = 0 for all s E S, and so (x, s) = (y, s) = 0 for each s E S . Since S is complete, the latter implies x = y = 0, and so x + z y = 0. Now, assume that the vector subspace generated in a Hilbert space H by is the a countable collection of vectors X i , x2, . . . is dense in H . If y1, y2, orthogonal sequence of vectors produced by applying the Gram-Schmidt orthog onalization process to the sequence X i , x2 , then the orthogonal set {yi , y2, } is automatically complete. Indeed, if y ..L y, for each n, then y ..L x11 for each n,
M
,
•
•
•
•
•
•
•
.
.
Chapter 6: HILBERT SPACES
294 and so
H,
y _i M. Since M is dense in
we get y
_l M =
H,
and thus y
=
0. This
observation is used extensively in constructing complete orthogonal (or complete orthonormal) sequences of vectors. The next result illustrates this point.
[a, b] -+ (0, oo) is a measurable weight } c L2CP ). Also, let a sequence Po, PI, P2, . . . function such that of non-zero polynomials be such that a. each Pn is of degree n, and b. (Pn, Pm ) = J:p(x)Pn(x)-= P� m� (x�) dx = 0 for n ;f. m . Then the orthogonal sequence Po, PI , P2, . . . is complete and coincides (aside ofscalarfactors) with the sequence of orthogonalfunctions ofL�(p) that is ob tained by applying the Gram-Schmidt orthogonalizationprocess to the sequence of linearly independentfunctions { 1, x, x2, x3, } . Theorem 33.12.
Assume that
{1, x, x2 , x3,
.
p:
•
.
.
Proof.
•
•
We shall prove first that
holds for each n. For n = 0, the claim is obvious. So, assume that (*) is true for some n and let P be a polynomial of degree n + 1. Since Pn+l is a non
zero polynomial of degree n + I , there exists some non-zero scalar Cn+l such that P -Gu+l Pn +l is a polynomial of degree less than or equal to n, i.e., P -Cn + l Pn+l E Span{1, x , x2 , . . . , x n }. So, from (*), we can write P - cu+ I Pn+l = L�=O ci Pi , or
P
=
E7:ci ci Pi. This shows that 2 Span {1, x , x ,
•
•
.
n I , x , xu + } C Span {Po , P,, P2, . . . , P, , Pn + d·
Since the reverse inclusion is obvious, the validity of (*) for n + I follows. Thus, (*) is true for each n. Now, a glance at Theorem 32. 1 1 guarantees that (aside of scalar factors) the sequence of polynomials P0 , P1 , P2, is the one •
•
•
obtained by applying the Gram-Schmidt orthogonalization process to the sequence 2 {1, x , x , }. •
•
•
Finally, in order establish that to show that the linear subspace
Po, P1 , P2 ,
M
•
•
•
is a complete sequence, it suffices
generated by
{ 1, x, x2 ,
•
.
.
} is dense in
L 2(p ).
M c C[a, b] c L2(p) and that M (by Corollary I 1 .6) is dense with respect to the sup norm in C[a, b]. This implies (how?) that C [a, b] To see this, note first that
L2(p). Next, notice that dense in L2(p), and so M is
is dense in the norm induced by the inner product in
Theorem 3 1 . 1 1 guarantees (how?) tliat dense in L 2(p). This shows that
C[a, b]
Po, PI , P2,
is . . . is a complete orthogonal sequence
in L2(p).
The following example illustrates the preceding theorem:
•
Section 33: HILBERT SPACES
295
Let Lz([ - l, l]) be the Hilbert. space of all square integrable functions on the closed interval [- 1 , 1] equipped with the Lebesgue measure. Then the functions 1, x, x2, are linearly independent. If the Gram-Schmidt orthogonalization process is applied to these polynomials, then the polynomials of the resulting orthogonal sequence are known as Legendre6 polynomials. We claim that, aside of scalar factors, the Legendre polynomials are given by the formula J.;'n (x2 - 1)2, the nth-order derivative of the polyno mial (x2 - l )". It is a custom to normalize the polynomials at l , in which case the Legendre polynomials are given by Example 33.13. .
.
•
l d" --1 (x2 - 1)11 • Pn(X) = -1 1 11! 2 dx 1 Clearly, each Legendre polynomial P11 is of degree 11 (since it is the nth-order derivative of a polynomial of degree 211) and satisfies the condition P11( 1 ) = 1. By Theorem 33. 12, in order to prove that the Legendre polynomials result from the Gram Schmidt orthogonalization process, we need only to show they are mutually orthogonal. Pick n and m such that n > m. To verify that P11 is orthogonal to P111, it suffices show that f�1 xk P11(x) dx = 0 for all k < n. The proof of this claim is based upon the observation that the function (x2 - 1)11 = (x - 1)" (x + 1)11 and all of its derivatives of order less than or equal to n - 1 vanish at the· points ±1. So, integrating by parts k times (where k < 11) and using the preceding observation, we get
1 1 xk P11(x)dx -I
-
d" 1-11 xk dx" (x2 - 1)" dx ( - l )kk! 1 d" -k 1 ----:- (x2 - lY' dx 2" n! - 1 dx 11-k
1 --
__
2"n!
I
(- l)kk! dll -k- 1 .., (x- - l)" 2"n! dxll-k- 1 -1
I
=
0.
Therefore P11 ..L P111 holds for all n =/= m. By Theorem 33.12, ( P11 } is a complete orthogonal sequence in the Hilbert space Lz([ -1, l ]). Finally, it is not difficult to see that the norms of the Legendre polynomials are given by Therefore, the sequence of polynomials Qo. Q 1 . Q2 , . . . , where II Pn l l =
J21�1 •
QII (x)
=
F[I I
n + I d" \"2 ( 2 dxII .
l_ _ 211 n.
_
I)1
1 .
is a complete orthonormal sequence of the Hilbert space L2([ - 1 , 6Adrien-Marie
r
I]).
•
r
Legendre ( 1752-1833), a F ench mathematician. His majo contribution was on
elliptic integrals that provided the analytical tools for modem mathematical physics. He spent over 40 years of his life attempting to prove Euclid's parallel postulate.
Chapter 6: HILBERT SPACES
296
EXERCISES 1. 2.
Verify that the inner product space L2(p) of Example 33 is a Hilbert space. Show that the Hilbert space L2[0, oo) is separable. 3. Let { 1/111 } be an orthonormal sequence of functions in the Hilbert space L2[a, b] which is also uniformly bounded. If {an} is a sequence of scalars such that an 1/111 -+ 0 a.e., then show that lim a11 = 0. 4. Let {¢11} be an orthonormal sequence of functions in the Hilbert space L2[- l , 1]. Show that the sequence of functions { 1/1n}, where
(x = Y,, )
( 2 )4 ( (x -2-)) ' b-a
¢n
2
b -a
-
b+a
is an orthonormal sequence in the Hilbert space L2[a, b]. 5. Show that the function (-, ·): X x X -+ C defined in Theorem 33.5 is a well-defined inner product on X that induces the norm of X. 6. Show that the closed unit ball of l2 is not a norm compact set. 7. Show that the Hilbert cube (the set of all x = (XJ , x2 , . . . ) e £2 such ihat lxn I ,::::: t holds for all n) is a compact subset of l2. 8. Show that every subspace M of a Hilbert space satisfies M = M .L.L . 9. For two arbitrary vector subspaces M and N of a Hilbert space, establish the following: a. b.
10. 11.
(M + N).L = M.L n N.L , and if M and N are both closed, then (M n N).L
= M.L + N.L .
Let X be an inner product space such that M = M .L.L holds for every closed subspace M. Show that X is a Hilbert space. Consider the linear operator V: L2[a, b] -+ L2[a, b] defined by
f(x ) = 1xf(t)dt.
V
12.
Show that the norm of the operator satisfies II V II =:: b - a. Let {x11 } be a norm bounded sequence of vectors in the Hilbert space l2, where x11 = (x�1 , x2_, x3, . . . ). If for each coordinate k we have lim11--+oc xJ: = 0, then show that lim (x , y) 11-+00 ,
13.
=0
holds for each vector y E l2. Let H be a Hilbert space and let {xn } be a sequence satisfying lim (x11, y) = (x, y) 11-+00
for each
y E H. Show that there exists a subsequence {xkn } of {xn } such that Jim n-+oo
}
II
X - - L.J �Xk·' = 0. n
i=l
Section 33: HILBERT SPACES 14.
Let p: [a, b] --+ (0, oo) be a measurable and �ssentially bounded function and for each n = 0, I , 2, . . . let P11 be a non-zero polynomial of degree n. Assume that
1 bp(x)P11(x)Pm(x) dx =
15.
297
0
for
n -:/:: m.
Show that each P, has n distinct real roots all lying in the open interval (a, b). In Example 33 we defined the sequence Po, P1, Pz, . . . of Legendre polynomials by the formulas
I dll P11(x ) = -- -- (x.2 l )II . 2" n ! dx " We also proved that these are (aside of scalar factors) the polynomials obtained by applying the Gram-Schmidt orthogonalization process to the sequence of linearly independent functions { 1, x, x2, . • . } in the Hilbert space L2([-1, 1]). Show that for each n we have -
P11 ( I) = 16.
17.
1
II P" II =
and
v�. z;;+I
Let {Ta }aeA be a family of linear continuous operators from a complex Hilbert space X into another complex Hilbert space Y. Assume that for each x E X and each y E Y, the set of complex numbers {(Ta(x), y): ex E A } is bounded. Show that the family of operators {Ta laeA is uniformly ·norm bounded� i.e., show that there exists some _ constant M > 0 satisfying II Ta II .::= M for all ex E A. Let {¢11} be an orthonormal sequence in a Hilbert space H and consider the operator T: H -+ H defined by 00
T(x) = L cx"(x, ¢n )¢", 11=1
18.
where {cx11} is a sequence of scalars satisfying lim cx11 = 0. Show that T is a compact operator. Assume that T, T*: H --+ H are two functions on a Hilbert space satisfying (Tx, y)
=
(x, T*y)
for all x, y E H. Show that T and T* are both bounded linear operators satisfying liT II = II T* II 19.
and
II TT* II = II T 11 2 •
Show that if T: H -+ H is a bounded linear operator on a Hilbert space, then there exists a unique bounded operator T*: H -+ H (called the adjoint operator of T) satisfying (Tx, y) = (x, T*y)
for all x, y E H. Moreover, show that II T II = liT* II.
Chapter 6: HILBERT SPACES
298
34. ORTHONORMAL BASES Recall that an orthogonal set S in an inner product space is said to be orthonormal if Us II = l holds for each s e S. A complete orthonormal set is a maximal orthonormal set. By Theorem 33.1 1, we know that every inner product space has a complete orthonormal set.
Definition 34.1. A complete orthonormal set in a Hilbert space is known as an orthonormal basis. An orthonormal basis will be denoted by {e; };e/ . That is, a family of vectors and only if l.
2.
{e; };e1 of a Hilbert space is an orthonormal basis if
(e;, ej) = �ii• where �ij is Kronecker's7 delta (defined by �ij = and �ij = 0 if i :f: j); and (x, e;) = 0 for all i implies x = 0.
l if i
=
j
The orthonormal sets that are bases are characterized as follows:
For an orthonormal family {e; };e1 of vectors in a Hilbert space the following statements are equivalent. l . The family {e; };e / is an orthonormal basis. 2. If x .l e; for each i e I , then x = 0. 3. For each vector x we have (x, e;) :f: 0for at-most countably many indices i and x = L;e1 (x, e;)e;, where the series converges in the norm. 4. For each pair of vectors x and y we have (x, e;) :f: 0 and (y, ei) :f: 0 for at-most countably many indices and (x, y) = L;e1(x, e;)(y, e;). 5. (Parseval's8 Identity) For each vector x we have llxll 2 = Li e/ l(x , e;) l 2 . Proof. ( l ) =? (2) This is an immediate consequence of the definition of the Theorem 34.2.
orthonormal basis. (2) =? (3) Fix some vector x . By Bessel's inequality (Theorem 32.1 0), we know that (x, e;) :f: 0 for at most countably many indices i. Let i 1 , i2 , be an enumeration of the indices for which (x, e;) :f: 0. (Here we consider the case where e;) =I= 0} is countable; the finite case can be dealt in a the set of indices {i e I : similar manner.) From Bessel's inequality, it also follows that 2::1 l(x, ei.t) 1 2 < •
•
•
(x,
llx 112 < oo. 7Leopold Kronecker ( 1823-1891 ), a German mathematician. He made imponant contributions to
the theory of equations and to the theory of algebraic numbers. Due to his belief that mathematics
should deal only with "finite numbers," he vigorously opposed Cantor's introduction of set theory. He is also famous for the quotation "God created the integers, all else is the work of man." 8Marc-Antoine Parseval des Chenes ( 1755-1 836), a French mathematician. He published very little
and he is only known for this identity. He was a royalist who was imprisoned in 1792 for publishing poetry against Napoleon's regime.
Section 34: ORTHONORMAL BASES
299
Now, the Pythagorean Theorem 32.7 guarpntees
2 m m L(x, e;t )eit = L j (x, eh)l 2 k=ll
k=ll
for all n and m. This shows that the sequence {2=�= t (x, eh)eh} is a Cauchy sequence, and thus the series 2:�1 (x, e;, )eh Li e/ (x, ei )e i is norm convergent. Let y = x - 2:�1(x, Since (x - L���(x, eit )eit , e;. ) = 0 for all m > I , 0 for all 11. it follows from the continuity of the inner product that (y, e;.) Consequently, (y, ei) = 0 for each i E I, and so, by our hypothesis, y = 0 or
=
eiJeh ·
x = L; e1(x, e;)e;. (3) ==> (4) Let x
=
=
Lie 1(x, e;)e; .and y = Lj e 1 (y , ej)ej. Since there are
only at most a countable number of non-zero terms in each sum and the two series converge in norm to x and y, respectively, it follows from the joint continuity of the inner product that
(x, y) =
e )
( Lie/
(x, e;)ei, L(y , j )ej
je/ = L L(x , ei)(y, ej )(ei, ej ) ie/ j eJ
= L(i x, e/
ei)(y, ej).
=
(4) ==> (5) The desired identity follows immediately by letting y x. (5) ==> ( 1 ) Suppose that {e;}ie / is not an orthonormal basis. This means that {ei };e / is not a maximal orthonorm�.I set. Thus, there exists a non-zero vector y that is orthogonal toeache; . But then, our hypothesis implies IIY f = L ; e 1 l (y , e;) l 2 0 contrary to the fact that y is a non-zero vector. Therefore, {e; };e1 is an orthonor
mal basis, and the proof of the theorem is finished.
= •
If {e; };e/ is an orthonormal basis in a Hilbert space and x is an arbitrary vector, then thefamily ofscalars { (x, e; )Jiel is referred to as the family ofFourier9 coefficients ofx relative to the orthonormal basis {e;};e/· Definition 34.3.
The Hilbert spaces with countable orthonormal bases are precisely the infinite dimensional separable Hilbert spaces. 9Jean Baptiste Joseph Fourier ( 1768-1830), a French mathematician. He is the founder of Fourier analysis. In his 1807 memoir, On the Propagation of Heat in Solid Bodies, he introduced the idea of expanding functions as trigonometric series. Although this was a controversial revolutionary idea at that time, today it is the basis of modem "Fourier analysis."
Chapter 6: fiLBERT SPACES
300
An infinite-dimensional Hilbert space H is separable if and only if it has a countable orthonormal basis. Moreover, in this case, every orthonormal basis of H is countable. Theorem 34.4.
Proof. Let {e1, e2 , • • • } be a countable orthonormal basis in a Hilbert space H. Then the set of all finite linear combinations of the e, with rational coefficients (call a complex number rational if its real and imaginary parts are both rational) is a countable dense subset of H. Therefore, H is a separable Hilbert space. For the converse, assume that H is a separable Hilbert space and let {x 1 , x2, • • • } be a countable dense subset of H. We claim that there exists a strictly increasing sequence of natural numbers k1 < k2 < k3 · · · such that {xk, , Xk2, • • • } is a linearly independent sequence satisfying
for each n. We shall establish the existence of such a sequence by induction. Start by observing that we can assume x, # 0 for each n. Let k1 = I . For the induction step, assume that we selected natural numbers k1 < k2 < · · · < k, such that the vectors {xk, , Xk2, • • • , Xkn} are linearly independent and
If every vector x; with i > k, lies in the linear span of {xk. , Xk2 , • • • , xkJ, then the set {x 1, x2 , • • • } spans a finite dimensional vector subspace, and so it cannot be dense in the infinite dimensional Hilbert space H , a contradiction. Hence, there exists some vector x; with i > k, which does not lie in the linear span of {Xkp Xk2 , • • • , XkJ• Let
k,+l
=
min { i E lN:
X;
is not in the linear span of {xk, , Xk2 , • • • , Xkn } } •
Clearly,
This completes the induction and the proof of the existence of the sequence {xkn } . Since {x1, x2, • • • } is dense in it follows that the linear span of {xk, } is also dense in H. So, if we apply the Gram-Schmidt orthogonalization process to the sequence {xkJ, we get a complete orthogonal sequence. Normalizing this orthogonal sequence yields a countable orhononnal basis for H.
H,
301
Section 34: ORTHONORMAL BASES
For last part, assume that {b�t b2 , • • . } is a countable orthonormal basis of a _ For each n let Hilbert space and that {e; };e1 is another orthonormal basis of
H
1., =
H.
{i E /: l(e;, b.,)l �}· >
From Parseval's identity 1 = ll bu ii 2 = L; e; l ( bu , e; ) l 2 , t follows that 111 is a finite set. On the other hand, if i E I, then using Parseval 's identity once more, we see that 1 = lle; 11 2 = I:: 1 l(e;, bu) l 2, and so (e;, bn ) i= 0 for some n. Therefore, I = U: 1 111• This shows that I is at most countable-and hence, a countable • s�.
i
In Example 33.13, we saw that the sequence of normalized Legendre polyno mials Qo, Q 1 , Q 2, . . . , where
Q (r) · "
� n + d'' 2 2n n! 2 dx11 • ' 1
1
--
-- (r - 1 )11
is an orthonormal basis of the Hilbert space L 2 ([ - 1 , 1 ]). We now present two more examples of orthonormal bases in some classical Hilbert spaces. The first example exhibits an orthonormal basis in the Hilbert space L 2 ([0, oo)). Example 34.5. Let us denote the differential operator by D, i.e., Df = f' and ok f = f (k) , the kth derivative of f. The Laguerre 1 0 polynomials are defined by the fonnulas
for n =
0, 1, 2, . . . .
Clearly, each Laguerre polynomial L 11 (x) is of degree
define the Laguerre functions
¢o, 2, . . .
>n (x) = We claim that ( >o, >2, 4>3, . . . } is
by
�n. e-i Ln (x), an
n = 0, 1,
2,
11.
Next, we
....
orthononnal basis in the Hilbert space
L z([O, oo)) of
all square integrable functions on the interval [0, oo) equipped with the Lebesgue measure. This claim will be established by steps. STEP I: The sequence (>o, >2, 4>3, . . . } is orthogonal.
Let m <
Then we need to show that
J000 Lm (x)L11(x)e-x dx = O. Since
polynomial of degree m, it suffices to verify that for all k < n we have 11.
L111(x) is a
10Edmond Nicolas Laguerre ( 1 834-1 886), a French mathematician. He worked in geometry and on approximation methods in analysis.
Chapter 6: HILBERT SPACES
302 To establish
() *
, integrate by parts k times to get
Therefore, the Laguerre functions are mutually orthogonal. STEP ll:
The sequence
(
.{¢o, ¢2, ¢3, . . . } is
orthonormal.
Taking into account *) and integrating by parts n times, we get
l e-2 Ln Il .. 1
2
n!
=
_
1
--2 (n!)
1-? -
1 looo 1oo Ln(x)Ln(x)e-.\. dx -(-l)nx" Ln (x)e-x dx 1oo (-l )nxn Dn(x"e-x. )dx 1 looox' e-.\. dx =
o
- ( ! 2 (n!)(n!) = n ) STEP ill:
The
f n
o
1.
sequence {¢o, ¢2 , ¢3, . . . } is
Assume that a function
2
= -2 (n!) (n!) o
(n!)- o
1
(n!)
complete.
2 0, oo)) satisfies f l_ ¢ for each 11, i.e.,
f E L ([
We must show = 0 a.e. To this end, let holds for all = 0, 1 , . . . This implies
. foooxng(x)dx
n
g (x) f(x)e-f =
and note
J;o g(x)xLn(x) dx = 0
= 0, for n = 0, 1 , 2, . . . .
(**)
g
f
To show that = 0 a.e., it suffices to prove that = 0 a.e. To see this, start by considering the complex valued function F(s) =
loooe-.ug(x) dx
define for all complex numbers s with real part Res > 0. Next, fix a complex number so with Reso > 0. Then there exist some neighborhood V of so and some a > 0 such that Res > a holds for all s E V. This implies that there exists some > 0 such that
M ae-sx g(x) '-xe-sx g(x) Mlg(x)l I I I I x E oo). as
=
<
holds for all s E V and all [0, Now, a glance at Theorem 24.5 (and the discussion after its proof) guarantees that F is differentiable at so and that F'(so) = -]000 In other words, F is an analytic function.
xe-sxg(x) dx.
303
Section 34: ORTHONORMAL BASES
Next, observe that for each n, Holder's inequality implies
Ia In particular, for each s Ia
00
lx11g(x)l dx
00
=
!f(x)lx 11 e-� dx
00
)!
x2n e -x dx 2
=
IIJIIJ(2n)!.
and each n, we have
>0
foo
< 11!11 ( Ia
oo s11-lx" g(x)l dx < llfii-J(2n)! s" n!
n!
=
a11.
<s <
1· Since limn �oo a�:1 = 2s, it follows that the series .L�o a11 converges for all 0 But then, Theorem 22.9 (see also Exercise 10 of Section 22) implies that the series L�o ( - 1 )11 (s;;r g(x) = e-sx g (x) defines an integrable function and that
oo s" oo x"g(x) dx e-sx g(x) dx L( -1 )11 F(s) Loo n. Lo o 11=0 holds for all 0 < s < � - Now, a glance at (**) yields F(s) Loo e-s.\ g(."C)dx for 0 < s < 2-I . Since F is an analytic function, the identity theorem for analytic functions implies F (s) 1
=
=
=
0
=
0,
.
=
0
for all complex numbers s with Res > 0; see for instance [2, Theorem 16.25, p. 462]. That is, the Laplace transform of g equals zero and so g = 0 a.e. (see Example 30.12), and the proofs of our claims are finished. • Example 34.6. Consider the weighted Hilbert space Lz(p) with the weight p(:r) = e-�.r• defined over the whole real line ( -oo, oo). The functions 1'
,.2
"' "" ...
"'
,.3 , . . .
' ·"
belong to L2(p) and they are clearly linearly independent. We claim that the linear span of this sequence i� dense in Lz(p). To see this, assume that a function f E L 2(p) satisfies foo f(x).�11 e_ l2·r- dx = 0 c10r each n = 0 , 1 , 2 , ·; . . Now, coosa'der the funct10n g(x) = � f(x)e- �x-. Since the functions lz(x) = f(x)e-lx- and B(x) = e-�x- belong to L 2 (1R), it follows that g = Jze E L 1 (1R) and
_00
·
!-00oo
x" g(x) dx
=
0, n = 0, 1, 2,
..
.
.
We must show that f = 0 a.e. or, equivalently, that g = 0 a.e. To see this, we introduce the complex-valued function G: C --+ G(z)
=
!00
j (x)e- � x2 ixz dx .
-oo
C by
Chapter 6: HILBERT SPACES
304
It should be clear that this (complex) integral is well defined and that (as in the previous example) G is differentiable at each point (and so G is an entire function). At z = 0, we have
for each
n = 0, l , 2, . . .
.
Therefore, G(z) = 0 for each z E C, and consequently =
100f(x)e-ix2e-ls.�: dx 1 00 g(x)e-1sx -oo
-oo
dx
=
0
for all -oo < s < oo. In other words, the "Fourier transfonn" of g is zero, and this guarantees that g = 0 a.e.; see, for instance [15, p. 408] or [26, p. 187]. We now claim that if we apply the Gram-Schmidt orthogonalization process to the sequence of functions 1, x, x2 , . . . , then the members of the resulting complete orthogonal sequence {H,) (aside of scalar factors) are the Hermite 1 1 polynomials which are defined by I ,.2 \ nI1 I .--:;.. n • •, (• · ) - (- '1 )II t:- 'iI ..,.2 u .... \c: - } ' ·
where D denotes the differentiation operator. Clearly, each H,(x) is a polynomial of degree n. Now, if k < n, then-integrating by parts k times-yields
This implies (Hm. H,) =
100 Hm(x)H,(x)e-�';
dx = 0
-oo
for n # m. Finally, a glance at Theorem 32. 1 1 guarantees that (aside of scalar factors) the sequence {H, } is indeed the one obtained by applying the Gram-Schmidt orthogonalization process to the sequence of linearly independent functions I, x, x2 , x3 , . . . . Nonnalizing the sequence of functions {Hn) we get to a complete orthononnal basis for L 2(P ) . •
Next, we shall use Theorem 34.2 to present a concrete realization of an arbitrary Hilbert space. To do this, we need some preliminary discussion. Recall that a linear operator T: -+ Y betwee n two norrned spaces is called norm preserving (or an isometry) if.IITxll = llxll holds for all x Similarly, a linear operator L: H1 -+ H2 between two Hilbert spaces is said to be inner product preserving if (Lx, Ly) = (x, y) holds true for all x, y e H1•
X
E X.
1 1 Charles Hermiie ( 1 822-1901 ), a French maihemaiician. He worked in Ihe Iheory of funciions. He was lhe firs1 1o prove (in 1873) 1ha1 e is a Iranscendemal number.
Section 34: ORTHONORMAL BASES
305
From Theorem 32.6 the following result should be immediate:
linear oper a tor L: H1 H2 between two Hilbert spaces i s norm preserviug if and only if it is inner product preserving. Lemma 34.7. A
�
Now, let Q be an arbitrary nonempty set. If x: Q C is an arbitrary complex valued function (which it will also be denoted as a family {x(q)}qeQ ). then its e2-norm is defined by �
llxl lz =
(L ) q eQ
l x(q) l 2
I
2
complex-valued function x: Q C is said to be square summable if llx112 < This is, of course, equivalent to saying that x(q) #- 0 for at most countably many q and i2f q1, q2, .. . is an enumeration of the set {q E Q: x(q) =f:. 0}, then z=:, jx(q11)1 It should be clear that the collection l2(Q) of all square summable functions with the pointwise algebraic operations and the l2-norm is a normed vector space. In actuality, it can be easily shown that l2( Q) is a Hilbert space.
A
�
oo.
< oo.
The vector space l2 ( Q) of all square summable complex-valued functions defil7ed on a nonempty set Q under the inner product (x, y) qL:eQ x(q)y(q) is a Hilbert space. Lemma 34.8.
=
We now ready to establish that every Hilbert space is linearly isometric to a concrete l2(Q) Hilbert space. are
Every Hilbert space H is linearly isometric to a Hilbert space ofthethelinear formoperator l (Q) . Specifically, if {e; };e/ is an orthonormal basis of H, then L: H lz(l), defil1ed by
Theorem 34.9. z
�
L(x) = {(x, e;)};e/.
is a sw]ective lil1ear isometry.
Clearly, L is linear and by Parseval 's Identity (Theorem 34.2(5)) it is also an isometry. \Ve leave it as an exercise for the reader to show that L is also a surjective linear operator. ProQf.
•
Chapter 6: Hll..BERT SPACES
306
Corollary 34.10.
An infinite dimensional Hilbert space H is separable ifand only if it is linearly isometric to
l2 •
EXERCISES 1. 2.
(ei li e/ and (h ljeJ be two orthonormal bases of a Hilbert space. Show that I and J have the same cardinality. Let (ei l ie/ be an orthonormal basis in a Hilbert space H. If D is a dense subset of H, Let
then show that the cardinality of D is at least as large as that of I. Use this conclusion to provide an alternate proof ofTheorem 34.4 by proving that for an infinite-dimensional Hilbert space
c.
H has a countable orthonormal basis. H is separable. H is linearly isometric to
Let
I be an arbitrary nonempty set, and for each i E
a. b.
3.
H the following statements are equivalent:
l2.
family of functions
4.
l x II
Let
(ei liei = 1.
elements.
5.
(ei lie/
i iet ei
=
X! i l· Show rhar the
is an orthonormal basis for the Hilbert space l2(I).
I: l(x,
k2
be an orthonormal basis in a Hilbert space and let x be a unit vector, i.e,
Show that for each k
E 1N the set (i E
e; )I
:::: t l has at most
M be a closed vector subspace of a Hilbert space H and let (e; l i e/ be an or thonormal basis of M; where M is now considered as a Hilbert space in its own right under the induced operations. If x E H, then show that the unique vector of M closest
Let
to x (which is guaranteed by Theorem
33.6) is the vector
y =
L; e1(x, e; )e;.
(fn
6. Let (e11 l be an orthonormal basis of a separable Hilbert space. For each n, let J,, =
7. 8. 9.
10.
en+ l - en . Show that the vector subspace generated by the sequence
Prove Lemma
34.7.
Prove Lemma 34.8. Complete the details of the proof of Theorem Let
34.9.
L2 [0, f L [0, a2, . . . 1 f{¢n L 2(JJ.) {aJ, f ¢n dJ.L.
(en l be an orthonormal sequence of vectors in
11.
(¢n { ¢n {¢J, ¢2, 2::�1 L,�1
12.
L2[0, f L,�1 (f, L2[0, ¢n)l2 2 11/11 2::� 11(/, . L2(JJ.), J.L f)
the Hilbert space
2rrl we have
pose that for each continuous function f in Show that
l is dense.
(en l is an orthonormal basis.
=
l be an orthonormal sequence of vectors in the Hilbert space 2rrl we have = pose that for eachcontinuous function in 2 Show that l is an orthonormal basis. Let . . . l be an orthonormal basis ofthe Hilbert space where Let
measure. Fix a function
E
coefficients relative to
1. i.e.,
an¢n a11¢n
and let
an = J
2rr ]. Sup
en )en .
2rr]. Sup
is a finite
be its sequence of Fourier
Show that (although the series
need not converge pointwise almost everywhere to
the Fourier series
can be integrated term-by-term in the sense that for every measurable set
E we have
1 f dj.L f 1 tPn dj.L. E
=
n=l
an
E
Section 35: FOURIER ANALYSIS [HINT: If s11
=
Lk=l a11r/J11, then the Cauchy-Schwarz inequality implies
I Lt L d!L -
13.
307
s, d!L
1
2
<
(L
If - Snl d!L
r
� II / - s,II'!L'(E). ]
Establish the following "perturbation" property of orthonormal bases. If orthonormal basis and u;· };e/ is an orthonormal family satisfying
{e; };e/
is an
L lie; - /; 112 < oo,
ie / then {f; };e/ is also an orthonormal basis. 35.
FOURIER ANALYSIS
In this section, we shall study some basic properties of periodic functions. Recall that a function f: ffi. -+ C is said to be periodic if there exists some p > 0 such that f(x + p) = f(x) for each x E ffi.. The positive number p > 0 is called a period of f. In essence, a periodic function is completely determined by its values on the closed interval [0, p]. Any function f: [0, p] -+ C that satisfies f(O) f(p) gives rise to a periodic function via the formula f(x + p) f(x) for each x E ffi.. Notice that if f: ffi. -+ is a periodic function, then the function g: ffi. -+ C, defined by g(x) = f({;x), is a periodic function with period 2rr. From now on, unless otherwise stated, all periodic functions encountered in this section will be assumed to have period 2rr . As a matter of fact, they will be defined only on [0, 2rr] (and they will be tacitly assumed defined on all of ffi. f(x)). That is, a periodic function is any function via the fonnula f(x + 2rr) f: [0, 2rr] -+ C satisfying /(0) = /(2rr). The closed interval [0, 2rr] will always be considered equipped with the Lebesgue measure: we shall write dx instead of d).. ( x). The Hilbert space L2[0, 2rr] is as sumed equipped with the inner product
=
=
C
=
(/, g)
=
["f(x)g(x) dx,
and so it induces the standard Lz-norrn II / 112
=
.j(f, /)
= ([
"
l f(x)l 2 dx
). I
'
A straightforward computation shows that
if if
nn =
=j:. m
m.
Chapter 6: HILBERT SPACES
308
This means that the countable collection of functions e"'x = cosnx + 1 sin nx, n = 0, ±1, ±2, ±3, . . . , is an orthogonal subset of L2[0, 2rr]. Therefore, the collection
I
1
e 11/X:. n - 0, ±1, ±2, ±3, . . . _
.Jirr
l
is an orthonormal set of functions in L2[0, 2rr ]. However, for historical reasons, these functions are normally used in their non-normalized form. We shall establish here that the orthogonal set
{ewx :
11 = 0, ±1, ±2, . . . }
{ Jk, e'n.\· :
is, in fact, a complete orthogonal set-and so n = 0, ±1, ±2, . . } is an orthonormal basis for the Hilbert space L2 [0, 2rr ]. A trigonometric polynomial is any periodic function P of the form P(x) =
1
m
2ao L(a11 +
II== I
.
cos nx + b sin nx)
n
ao, a1, .. . , au
where and b1 , . . . , bu are complex numbers. From 1 sm x it follows that sm nx =
elnX
_ e -11/X
and
-----
2z
COS 11X =
(1)
eu
= cos x +
-----
2
and so every trigonometric polynomial P can also be written in the form m
P(x) =
nx L Cue' n=-m
(2)
e'nx
for appropriate complex coefficients c,. Similarly, using the identities cos nx + 1 sin nx, cos( -nx) = cos nx, and sin( -nx) = -sin nx, we see that every expression P(x) as in (2) can be written as in (1). Thus, we have shown that every trigonometric polynomial P can be written in the following two ways: P(x) =
m 2ao n=lL(a,
1
+
.
m
cos nx + b, sin nx) =
L Cnewx, n=-m
where the coefficients are related by the identities
ao
= 2co,
a, Cn =
- c_n, b, = z(c, - c_n)
Section 35: FOURIER ANALYSIS
309
and taking the inner product of a trigonometric polynomial P with the function Now, kn e' . it follows from the orthogonality properties that 1 /. 2 ±m P(x)e dx, n 0, ±1, ±2 Cn =
11'
21l'
-
-IIIX
=
0
. . . •
.
Next, assume that f L 1 [0. 21l'] (nothing precludes f from being also complex valued). Then the inequality l f(x)e"rx I < lf(x) l coupled with Theorem 22.6 guarantees that f(x)e x E L 1 [0. 21l'] for each integer n. This observation will be used in the next definition. E
'
n
The Fourier coefficients of a function f E L [0, 21l'] are thetermsofthedouble sequence ofcomplexnumbers c-2 · c_ 1 , co. c 1 . c2 defined by Definition 35.1.
1
. . . •
Cn =
for each n = 0. ± 1 , ±2, is theformal series
.
..
.
• . . .
211' /. . f(x)e-"'·t dx 1
21l'
0
The Fourier series of a function f E L 1 [0. 21l'] 00
L Cn e'"x.
n=-oo
An important consequence of the Riemann-Lebesgue lemma (Theorem 25.4) is that the Fourier coefficients of an integrable function converge to zero. If f E L1 [0, 27r], then its Fourier coefficients satisfy
Theorem 35.2.
lim c, = lim c_, = 0. L 1 [0, 21l' ]. Then, by the Riemann-Lebesgue lemma (Theo n-oo
n-oo
Let f rem 25.4). we have 211' 211' /. /. lim f(x) sin nx dx = lim f(x) cos nx dx = 0. Proof.
E
,_ 00 0
,_ 00 0
Chapter 6: HILBERT SPACES
310
So, from the identity
c,
. 2 1 1 f21rf(x)e-mx dx = -- { JTf(x)[cos nx - 1 sin nx ] dx = -2n lo 2n lo 1 f(x)cos nx dx - -1 1 2JTf(x) sin nx dx, = -
12Jr
21Z' 0
21Z' 0
it easily follows that
•
Cn -+ 0 and c_, -+ 0.
Using that the set of functions { e' nx : n = 0, ± 1 , ±2, . . . } is orthogonal and the formulas cosnx = -----
2
and
sin nx =
e'"x - e-t nx 21
-----
we see that the functions
�, cos x , sin x, cos 2x, sin 2x, . . . , cos nx, sin nx, . . . are mutually orthogonal in L2[0, 2n]. With respect to this orthogonal set of func tions, the Fourier series of any function f e L [0, 2n] can also be written as
1
00
"'"""
� Cn e 11 =-oo where the coefficients
I IIX
=
1
2ao + �(a, cosnx + b11 smnx) . 00
"'"""
·
n= l
ao, a1, a2, . . . and b1, b2, . . . are given by
ao = 2co = n� lo{21rf(x) dx, 2 On = Cn + c_n = _!_ { "f(x) cos nx dx, and n lo bn = l (Cn - c_n ) = _!_ {2Jrf(x) sin nx dx. n lo
If we express the Fourier series of a function in terms of the cosine and sine functions, then we shall also refer to the coefficients and as the Fourier coefficients of f. We are now ready to establish a result that will guarantee that the set of orthog onal functions { e'nx : n = 0, ±1, ±2, . . . } is complete in L2[0, 2n].
ao, a1, a2,
.
.
.
b1, b2,
•
.
.
311
Section 35: FOURIER ANALYSIS
Theorem 35.3. A Lebesgue integrable function over
coefficients all vanish is almost everywhere zero. 1
[0, 2rr] whose Fourier
Let f E L [0, 2rr] be a function satisfying !.2rrf(x)e'1 x dx 0 (t) 0 for all n 0, ± 1, ±2, . . . . Assume at the beginning that f is also real valued. We shall establish that f 0 a.e. holds true by considering two cases. First, we shall assume that f is also continuous and then we shall consider f to be an arbitrary integrable function. CASE Assume that f is a continuous function satisfying (t)for all integers n. Assume by way of contradiction that f =I= 0. This implies that there exists some x0 (0, 2rr) satisfying f(xo) =I= 0. Replacing f by -f (if necessary), we can assume that f(x0) 0. Choose some constant C 0 and some 0 < < min{ 1, x0, 2rr - x0} such that f(x) C holds for all x E I (x0 - x0 Now, we need a trigonometric polynomial that is greater than one on the interval (xo- x0 and less than one on the complement Ic [0, 2rr] \I. The reader should verify that the trigonometric polynomial P(x) cos(x - xo) satisfies P(x) for all x E I and 0 < P(x) < 1 on Ic. From (t), it follows that 2 1 f(x)[P(x)]" dx 1 f(x)[P(x)]" dx /.0 rr f(x)[P(x)]" dx 0 for 0, 2, 3, . . . . Therefore, 1 f(x)[P(x)]" d - 1 f(x)[P(x)r dx holds for all n 0, 1, 2, 3, . . . . We claim that the right-hand side of equation (*) is bounded while the left-hand side is unbounded. To see this, first observe that since 0 < P(x) holds for each x E Ic, we have 2 1 f(x)[P(x)]" dx < 1 1/(x)[P(x)]"l dx < /.0 rr lf(x)l dx. Proof.
=
=
=
1:
E
>
>
E,
>
=
+ E)
E + E).
E,
=
=
1+
1 + COSE
.
> 1
I
11 =
+
IC
=
=
1,
I
x =
IC
=
< 1
IC
IC
312
Chapter 6: HaBERT SPACES
Thus, the right-hand side of (*) is bounded. On the other hand, let J be a closed subinterval of I . Since P is strictly greater than one on the open interval I , it follows that there exists some a > I satisfying P (x) > a > I for each x e J . Therefore, for each n we have
[
f(x)[P(x) ]n dx >
[
" f(x)[P(x)] dx > Can /..(1).
This shows that the left-hand side of equation (*) is unbounded, a contradiction. This contradiction establishes that f = 0. CASE II: Assume that f E
Lt [0, 2JT] satisfies (t)for all integers n .
Consider the function F : [0, 2rr] -+
F(x) =
IR defined by
[xf(t)dt .
... u
J02rr
Clearly, F is a continuous function and from F(O) = 0 and F(2rr) = f(t) n dt = 0, it also follows that F is a periodic function. We claim that F ..L e' x holds for all n # 0. To see this, apply the integration by parts formula (see Exercise 12 of Section 3 1 ) to get 0 =
{h
Jo
e-rnxf(x)dx = e-"'xF(x) l 6rr -
= F(O) - F(2rr) - zn =
-z n
{ 21r
Jo
Jo{ 2rr
{h
Jo
(-zn)e-"u:F(x)dx
F(x)e-wx dx
F(x)e-w.r dx.
Therefore,
{ 2rr
Jo
F(x)e-"'x dx
=
0 for
n =
± 1 , ±2, . . . .
Next, let C = 2� J02rr F(x)dx, and consider the continuous function G: [0, 2rr] -+ IR defined by G(x) = F(x) - C.
Since 1 ..L e"':r , it follows that G is perpendicular to er nx for all n -::J: 0. Any easy inspection also shows that G (x) ..L 1 holds. That is, G ..L e' "x holds for all integers
Section 35: FOURIER ANALYSIS
313
n = 0, ±1, ±2, ±3, . . . . But then, by CASE G(x) = 0 for all x E [0, 2rr]. This implies F(x) = for f(t) dt = C for all x E [0, 2rr ]. Since F(O) = 0, it follows that C 0, and so for f(t) dt = 0 for all x E [0, 2rr]. This implies that f = 0 a.e.
I,
=
II
(see Exercise 19 of Section 22), and the proof of CASE is complete. Finally, consider the complex case. That is, assume that f = g + l h is an inte grable function whose Fourier coefficients all vanish. Then, it easily follows that both g and h have their Fourier coefficients all equal to zero. Therefore, by the preceding conclusion, g = h = 0 a.e. This implies f = 0 a.e., and the proof of the theorem is complete. •
The two sets of orthogonalfunctions I. {e"1x: n = 0, ±1, ±2, . . . }, and 2. { �, cos sin x, cos 2x, sin 2x, . . . , cos nx, sin nx, . . . } are both complete in the Hilbert space L2[0, 2n]. Proof. By Theorem 34.2, it suffices to show that if f E L2[0, 2rr] is perpen dicular to e1 11x for each integer 11, then f = 0. So, assume that f E L2[0, 2rr] satisfies f l. e1 11x for each integer n , i.e., assume that the Fourier coefficients of f Corollary 35.4.
x,
are all zero. Since L2[0, 2rr] c L1 [0, 2rr] holds, it follows from Theorem 35.3 that f = 0. Therefore, the set of functions {e111x: n = 0, ±1, ±2, . . . } is a complete orthogonal subset of the Hilbert space L2[0, 2rr ], and the proof is complete. • Parseval's identity for the complete sequence of orthogonal functions {e -111x } takes the following form:
Iff E L2[0, 2rr], then its Fourier series I::-oo cue-1'1.-c is norm convergent to f in L2[0, 2rr] and Corollary 35.5.
I1
-
2rr 0
Proof.
271'
l f(x) l 2 dx =
00
L
n=-oo
lcu l 2 .
From Corollary 35.4, we know that the collection of functions
I .../2iI ie-IIIX.. n
- 0 , ± I , ±2 , . . .
l
is an orthononnal basis for the Hilbert space L2[0, 2rr]. The conclusions now follow from Theorem 34.2. • Corollary 35.6.
Iff belongs to L2[0, 2rr], then its Fourier series
. -111x = a ""'"' Cne 2o + ""'"' � (a11 cos nx + b11 sm nx) � 00
00
11=-00
11=1
Chapter 6: HILBERT SPACES
314
can be illtegrated term-by-term in any closed subinterval [u, v] of[O, 2rr] in the sense that
lv "
f(x)dx
1
= 2ao(v =
a; Let Sn (X) we also know that II f - Sn II then Proof.
1 "
11 f(x)dx
-
1 11 "
-
u)
v l +L oo
n=l "
(a, cos nx + b, sin nx)dx.
+I:Z=1 (a, coskx + b, sinkx). From Theorem 34.2,
� 0. Now, if [u, v] is a closed subinterval of [0, 2rr],
Sn (x)dx
v l � ({ ,
<
l f(x) - S (x)l dx n lf(x) - s.(x)l 2 dx
- il f - sn ll v v - u.
This implies
n 1 -ao(v - u) + 2 k= I
L
vl II
(a, cos nx
+ b, sin nx)dx =
and the desired conclusion follows.
1
" 11
I
I
)' ({ )'
s,(x)dx
! 2 dx
�
lv "
f(x)dx, •
The fact that the Fourier series of a function f E £ 2 [0, 2rr] is norm convergent in the Hilbert space £2 [0, 2rr] should not be interpreted to mean that the partial sums of the Fourier series converge pointwise to f almost everywhere. Pointwise convergence of the. Fourier series is a much more difficult and delicate problem. The study of the convergence of the Fourier series is the subject of inquiry of the field of Fourier analysis with special emphasis on the following two basic problems:
1. 2.
The Convergence Problem: When does the Fourier series converge at a
given point x (or at each point of a given set)? The Representation Problem: If the Fourier series converges at some point x, what is the relationship between the sum of the series and the value of the function f(x) at the point x?
We remark that it is possible for the Fourier series to diverge at every point and, moreover, there exist continuous functions whose Fourier series diverge on an uncountable set. It is even possible to construct examples of continuous peri odic functions whose Fourier series diverge at a countable number of points in the interval [0, 2rr]; see also Example 35.1 1 at the end of this section. The most
Section 35: FOURIER ANALYSIS
315
striking result concerning the pointwise con'{ergence of the Fourier series was es tablished by the Swedish mathematician Lennart Carleson in 1966. He proved that the Fourier series of any function in L2 [0, 2rr ] converges almost everywhere. 1 2 The unpleasant situation regarding the pointwise convergence of the Fourier series can be corrected by considering the sequence of arithmetic means of the partial sums of the Fourier series. Our final objective in this section is to prove that the sequence of arithmetic means of the partial sums of the Fourier series of a periodic function f converges to the value of the function at every point of continuity. To do this, we need some preliminary discussion. For the rest of our discussion, f: [0, 2rr] C is a fixed periodic integrable function which will also be assumed defined on all of via the formula f(x + 2rr) f(x). Its Fourier series is "" • -IIIX = a2"0 + "" � < a cosnx + b, smnx ) � . 11=-CO 11=1 The nth partial sum of the Fourier series is I 1 2ao + L (ak cos kx + bk sin kx) k= J 1 [ ?rrf(t) dt ] + L [-1 ?rrf(t) cos(kt) dt ] cos kx ? - rr Lo k=J rr Lo 2rr 1 [ + L rr f(t ) sin(kt) dt sin kx ] 0 k=l 11 1- 12rrf(t) [-1 + L(cos kx cos kt + sink."( sin kt) ] dt 2 k=J rr o ] 2 i!' 1 1 [ -rr o f(t ) 2- + L cos k(x - t) dt k=I - -rr 12rr f(t)DII(x - t) dt, o where Dll (t) = 21 + L cos kt. k=J --+
1R
=
c,e
:>C
s,(x)
00
,
=
1
-
II
1
11
1
II
1
II
12
0n the convergence and growth of partial sums of Fourier series, Acta Matlzematica 116 ( 1966).
1 35-157.
Chapter 6: HILBERT SPACES
316
D1
is called the Dirichlet kernel of order The trigonometric polynomial Dirichlet kernel can be expressed in a closed form.
The Dirichlet ket·nel satisfies D11(t) = ( . �)t
Lemma 35.7.
n.
The
sin n +
2 sm 21
•
Proof. Notice that I
II
= 21
L ek
r
= 21
e -'
2 n 1
1
L e'k1
Dn(t) = 2 k=lL kt k=-n k=O e-w1 I] ------ - e-"11 [e'<2n+l)1 e11 - 1 2e'�(e'� -e-'�) e-' e'-- 2(e'2 -e-'2) ( �)t 2 � -
cos
+
1
2
II
'
e'
_
(
I ·1 (11+2>
I
11+2J1 -
-
� � ,
sin
n
•
t
•
_
--.,... ,..
+
sin
as desired.
(x)}f.
•
We now consider the sequence {a, of the arithmetic means of the partial and sums of the Fourier series of the function That is, =
( ) - so(x) s,(.x ) s2(.x)
a, .x
·
+
_
+
n+l
+···+
so(.x) �ao s, (x) n - , &
, 10r
_
0
.
1,2.. .
2 2 [ 1 11'f(t)Do(x -t) dt - 1o 11'f(t)Dt t) dt · {21rf(t)D,(x - t) dt ]] 21 1rrr Jo[ -t) dt o2 f(t) n LDk(X k=O 1 11'f(t)K,(x - t) dt,
Using the Dirichlet kernel and some algebraic simplifications, we see that .
a,(x)
=
I
n+1
··· +
=
1 -
=
1 -
rr
1r
1 rr o
+
1
1
+1
0
11
•
1
rr
(x -
+ ··
317
Section 35: FOURIER ANALYSIS
where the expression K,(t) =
1
ll
+1
IIL k=O
Dk(t) =
2(11
1 + 1)
.Stn Lk=O sin (k + -2 )t "2
II
1
1
of order n. Thus, the sequence {a, } of means of is now called the the partial sums of the Fourier series of f are given by the formula Fejer 13 kernel
1 a,(x) = -
121Tf(t)K,(x - t) dt.
Jr 0
It is not difficult to express the Fejer kernel in a close form. Multiplying the numerator and the denominator of K, (t) by 2 sin � and using the trigonometric identity 2 sin(k + !t) sin !t = cos kt -cos(k + 1)t� we get [ sin(n + 1) � ] 2 1 - cos(n + l)t 1 K,(t) 4(n + 1) sin2 � 2(n + 1) sin � This formula shows that the Fejer kernel is a positive even function. If, as usual, we put Kn(2krr) = then K, is a continuous function on all of JR. A simple computation (see Exercise 2 at the end of this section) also shows that = ---
=
"!1 ,
2 -
1 K,(t) dt 1T
Jr 0
=
1.
Now, let us take another look at the nth arithmetic mean of the partial sums of the Fourier series. By the preceding discussion, we have a,(x)
=
flo2rrf(t)K,(x - t)dt .!_ 1xx-+rr f(x rr 1T - 1 f(x u)K,(u) du. -1T .!_
=
rr
1
+ u)K,(u)dtt
rr
+
Jr
Taking into account that K,(-u)
1° f(x - rr
=
K,(u),
+ u)K11(u)du
=
it easily follows that
frrlo f(x - u)Kn(u)du,
13Leopold Fejer ( 1 880-1959), a Hungarian mathematician. He worked mainly on Fourier series and their singularities.
Chapter 6: HILBERT SPACES
318
and consequently, an(X) =
�r rr r Jo
[f
(x
+ u) + f(x - u)
·
2
]
(**)
Kn(u)du.
Next, we introduce a new function a1 defined by the formula )
af (X =
1.
1m
f (x + u) + f (x - u)
11--+0+
2
.
The function a1 is, of course, defined at those points for which the preceding limit exists. For instance, if f is continuous at some point x, then a1(x) = f (x) and if f has a jump discontinuity atx0, then a1(xo) = f<xo+>;t<xo->. We are now ready to state and proof the following remarkable result of L. Fejer concerning the convergence of the sequence of arithmetic means {an }. Theorem 35.8 (Fejer) .
Let f: [0, 2rr] � C be a periodic integrablefunction. Then, the sequence {an} ofarithmetic means of the partial sums of the Fourier series of f satisfies .
hm an(x)
n --+oo
= af(x) =
.
f (x + u) + f (x - u)
hm ------2
u--+0+
at every point x for which a1 (x) exists. Moreover, if f is continuous on some closed subinterval [a, b] of [0, 2rr], then {an} converges uniformly to f on [a , b]. Proof. Assume that a1(x) exists at some point x E [0, 2rr]. Then there exists
some 0 < 8 < rr such that
f (x + u) + f (x - u) 2
- af (X ) < E
for all 0 < u < 8. Notice that the 8 depends upon the E and the point x. f f is con tinuous on [a , b ], then notice that (in view of the uniform continuity of f) we can select the 8 to be independent of x E [a, b]. Now, from (*) and (**), it follows that
I
lan(x) - af(X) I =
<
[
]
{rr f(x + u) + f(x - u) - af(x) Kn(u)du 2 . rr Jo 2 { o f(x + u) + f(x - u) - af(X) Kn(u)du rr Jo 2 2
2
+rr
1rr o
f(x + u) + f(x - u) 2
- af (X ) Kn (U ) d U . ·
319
Section 35: FOURIER ANALYSIS
Now, observe that
2 1.8 rr o
u) + f(x - u) - af(X) K1 (U) dU 8 2 2 < � { EK11 ( t)dt < rrEK11(t)dt =E. rr Jo lo f(x
+
rr
Moreover, since K11(t) < 2(n+ l�sin2 � for each 8 <
x
<
rr ,
it follows that
2 !.7T f(x + tt) + f(x - u) - at (;\: ) K ( ) u 2 rr 8 f.7T f(x + u) + f(x - u) - a1(x) du -< C , < n+ 2 - (11 + )rr sin- o constant C f �.1 2 J;rr l � a ( ) l du is independent ofx. where the Finally, notice that i e "1:hoose some no such that I�I < E, then the pre ceding estimates show that -
11 U d
·
1
.
2
., 8
1
f<x+n> f(x-u>
�
=
-
f
1
x
holds true for all 11 > no (and all x E [a, b] if f is continuous on [a, b]). This completes the proof of the theorem. The following consequence of the preceding theorem plays an important role in applications. •
Let f: [0, 2rr] C be a periodic integrable function. As off converges at some point x and that sume that the Fourier series exists, then . 1 1 ao """""' ( ) = """" �"' c"e - · = 2 + � (a1 cos nx + b1 smnx ) . Corollary 35.9.
---+
af(x)
00
at x
n=-oo
r
00
n= l
Inthenparticular, ifthe Fourier series converges at some point of continuity off, f(x) = L c"e-'1 x = a; + L(a, cos nx + b11 sin nx). Proof. Assume that s holds and that exists. Theorem Since is also convergent to s (see Exercise 1 1 we know that 1 of Section 4), it follows that s = a1(x), and we are done. x
oo
oo
n=-oo
n=l
(x) a (x) ---+ af(x). Sn
af(X)
---+
{a11(x)}
By
35.8, •
Chapter 6: HILBERT SPACES
320
Here is· an example that demonstrates the far-reaching implications of Corol lary 35.9. Example 35.10.
Consider the periodic function f: [0, 21r] � 1R defined by f(x)
=
{
if 0 < X < 1l' 0 if 1r ::: x < 21r 1 if X = 21l'. 1
A direct computation shows that the Fourier coefficients of f are given by
ao
=
an
-
.!. 1r
f 2"f(x)dx
lo
Jo
= .!. [" l dx 1r
=
1,
.!. {21rf(x)cosnxdx = .!_ ["cos 1l' h 1l' h 1
r rr
bn - ...:.. 1 2 1r
Jo
1 f(x)sinnx dx = TC
1o"
x dx = 0, and
n
sinnx dx =
1
- cos ll1l' nTC
Thus, the Fourier coefficients satisfy ao 1, a11 = 0 for each n, b, = 0 for n even. Therefore, the Fourier series of f is given by =
(
b11
1 2 sinx sin3x sin5x sin7x - + - - + -- + -- + -- + · · · 1 3 5 2 1C 7
1
-
=
)
1 - ( -1) 1 llTC
11;
for n odd and
.
This series converges at each x; see Example 9.7. By Corollary 35.9, this Fourier series converges to f(x) at the points x of continuity of f. More precisely, we have 1 2 -+2
1C
(
sin x
sin3x
sin5x
3
5
sin7x 7
- + -- + -- + -- + · · ·
1
) { ;:::
f(x) if x E (0, 1r) U (TC, 2TC),
!
This conclusion is typical and very powerful for applications.
if X = 0, TC, 2TC.
•
Finally, we close the section with an example that guarantees the existence of a continuous periodic function whose Fourier series diverges on a given countable set. Example 35.11. This example demonstrates the existence of a continuous periodic func tion whose Fourier series diverges on a countable set of points. Consider the Banach space C [0, 2TC] of all continuous real-valued functions with the sup norm. Let X be the subspace of C[O, 2TC] consisting of all continuous periodic real-valued functions. Clearly, X is a closed subspace of C[O, 2TC], and hence, it is a Banach space in its own right Next, fix a point x E [0, 2TC] and define the linear functional S11: X -+ 1R by the nth partial sum of the Fourier series of the given function evaluated at the point x. That is, for
Section 35: FOURIER ANALYSIS
each I E X we let Sn(l) =
L qe-,J:..� 11
•
•
k=-11
321
{ 2rr = - J l(t)D1 (x - t)dt. I
o
1f
1
0
It should be clear that this formula defines a bounded linear functional on X whose norm (according to Exercise 3 at the end of the section) is given by the formula liS" II =
I -
la2rr ID"(x - t)i dt.
1f 0
Next, we estimate this norm. Since the Dirichlet kernel D, is periodic, it follows that
{27T
Jo
ID"(x - t)i dt
=
{27T ID11(t)l dt.
Jo
So, using the inequality jsin u I ::: ju I and changing the variable, we get
la2rr j D11 (l)j dt = la2rr sin (n . o
o
>
+
2 sm
!orr jsin(2n
- o
+
t) t
r 2
l)uj
lu i
dt
l sin u j du lu i
2n
=
211
> This implies
I
+
+
dLL =
rr ) +l k 1( .l: k rr k=O [ 1(k+l)rr - L (k J:.rr k=O
!orr jsin(2n. I)u I du o lsm uj la <2"+t )rr jsinuj
=
I )rr
-du lu i
o
]
jsin uj du =
{xt , x2 ,
2u
L (k +2 l)rr .
k=O
and consequently limiiS1111 = oo independently of the point x. XJ, • • • } be a (possibly dense) countable subset of the interval [0, 2rr]. Now, let Let S11,111 be the bounded linear functional (defined as above) on X whose value at I E X is the nth partial sum of the Fourier series qf I evaluated at the point x111• From the above discussion, for each fixed m we have lim
1/-+00 IISu,m
II = 00.
Chapter 6: HILBERT SPACES
322 .
Therefore, by the Principle of Condensation of Singularities (Theorem 28.1 1 ), there exists some f E X (i.e., a continuous periodic real-valued function f on [0. 2rr]) satisfying
L Cke-l kxm
lim sup ISn,m(/)1 = lim sup
fl
k =-n
fl-+00
for each
m.
=
00
This implies that the Fourier series of f diverges at each point Xm
•
•
EXERCISES 1.
Show that sin11 x is a linear combination of {1, sinx, cosx, sin 2x, cos 2x, sin 3x, cos 3x, . . . , sin nx, cos nx }.
2.
Furthermore, show that the coefficients of the cosine terms are zero when n is an odd integer, and the coefficients of the sine terms are zero when n is an even integer. Show that the Dirichlet kernel Dn and the Fejer kernel Kn satisfy I
rr
ltr D11(t)dr = ltr K11(t)dr = I I
-
rr
-1r
.
-1r
3. Show that the norm of the linear functional Sn defined in Example 35 satisfies
2 IISull = ..!_ { 7r1Du(X - t)! dt. rr
4.
lo
Show that the sequence of functions
I ( ;) , { ; ) }
I
2
2
I
2
cosx ,
(rr) 2
I
2
cos2x ,
(;) 2
I
2
cos3x ,
(;) 2
I
2
cos 4x, . . .
I
is an orthonormal basis in L2[0, rr]. Also show that the preceding sequence is an orthogonal sequence of functions in L2[0, 2rr] which is not complete. S. Show that the sequence of functions
j \G)
sin x ,
! G)
sin2x,
G)!
sin 3x ,
! G)
sin4x ,
·
·
·
I
is an orthonormal basis of L2[0, rr]. Also prove that this set of functions is an orthog onal set of functions in L 2 [0, 27r] which is not complete. 6. The original Weierstrass approximation theorem showed that every continuous func tion of period 27r can be uniformly approximated by trigonometric polynomials. Establish this result. 7. Find the Fourier coefficients of the function / (x)
=
I
}
0
if 0 < X <
if
!
-
2
zr.
� x < 2rr.
Section 35: FOURIER ANALYSIS 8.
323
Find the Fourier series of the function s nx if 0 _::: X < 1r f (x) - smx if rr _::: x < 2rr.
=
I
�
9. Show that for each 0 < x < 2rr we have
n= l
10. Show that
holds for all 0 _::: x 11. Show that
00
-.
smnx x = rr - 2 � LJ 11 ·
_:::
00 cos nx rr 2 x-2 = rrx - - + 2 l: --=-2 n2 n=l
= 0 yields :L� 1 ,:� = �� .) 4 oo ( cos2nx rr sin nx ) = -rr+4 L
2rr. (The value x ?
.c
3
3
?
n=l
_
n
11
holds for each 0 < x < 2rr. 12. Consider the "integral" operator T: L2[0, rr] --+ Lz[O, rr] defined by Tf(x) where the kernel K: [0, rr]
x
= !orr K(x, t)f(t)dt,
[0, rr] --+ 1R is given by
�
K (x, t) = LJ n=l
[sin(n + l)x] sinnt 2 11
Show that the norm of the operator T satisfies II T II = rr /2. [HINT: Use the basis described in Exercise 5.]
•
7
CHAPTER
_ __ _ _ _ _ _ __ _ _ _ _
SPECIAL TOPICS IN INTEGRATION
The powerful techniques of measure and integration theory are utilized in many scientific contexts. However, in a number of applications, set functions that are not measures appear naturally. For this reason a study of more general set functions promises to be very fruitful. In this chapter the set functions known as "signed measures" will be investigated. Loosely speaking, a signed measure is an extended real-valued a-additive func tion on a a-algebra of sets. The first section of the chapter deals with the algebraic and lattice structures of signed measures, while the following adheres to compar ison properties of signed measures. The most important comparison properties are those of "absolute continuity" and "singularity." Regarding absolute continu ity, the far-reaching classical Radon-Nikodym theorem is proven here: If a finite signed measure v on a a-algebra is absolutely continuous with respect to a a-finite measure J..L , then there exists a (unique) J..L-integrable function f such that v(A) i f dJ..L This powerful theorem is used to show that L;CJ..L) Lq(J..L) holds for all holds for each 1 After comparing signed measures, our attention is turned to regular Borel mea sures on a Hausdorff locally compact topological space Another classical result, known as the "Riesz Representation Theorem," is proved: If F is a positive linear functional on then there exists a unique regular Borel measure J..L such that F(f) Jf dJ..L holds for all f E As an application of this theorem, the norm dual of will be characterized in terms of regular Borel measures. I;
=
A E I;.
=
< p < oo.
X.
Cc(X),
=
Cc(X).
Cc(X)
325
�hapter 7: SPECIAL TOPICS IN INTEGRATION
326
The last two sections of this book deal with differentiation and integration in Rn . First, the study focuses on the differentiation of Borel signed measures on R". It will be shown that every Borel signed measure is differentiable al most everywhere. This basic theorem will then be used effectively to derive the classical results about ordinary derivatives of functions of bounded variation. Finally, a detailed proof of the familiar "change of variables formula" will be presented.
36. SIGNED MEASURES
:E
X,
Throughout this section, will be a fixed a-algebra of subsets of a set and the measures considered will be assumed defined on :E . Since a variety of measures will be studied, it is a custom (for simplicity) to call the members of :E the measurable subsets of Let JL and v be two measures, and let a > 0. Then the two set functions JL + v �11d a f.l· defined by
X.
(/L + v)(A) (aJL)(A)
JL(A) + v(A), = a JL(A) =
for each A e are obviously measures. That is, the collection of all measures on :E is closed under addition and also by multiplication by nonnegative scalars. Clearly, this collection cannot be a vector space since multiplication by - 1 yields negative-valued s�t functions. An order relation < can be introduced among measures by letting JL < v whenever JL(A) < v(A) holds for each A e :E. The reader should stop and check that < is indeed an order relation. Remarkably, the collection of all measures under this ordering is a lattice. That is, for every pair JL and v of measures, the least upper bound JL v v and the greatest lower bound f..L 1\ v exist. The details are included in the next theorem.
:E
The collection of all measuze s ozz :Eforms a lattice, wherefor each pair ofmeasuz-es azzd v the lattice opez·arions are given by :E and A}, and JL v v(A) + v(A :E and A} JL 1\ v(A) + v(A for each A Moreover, Theorem 36.1.
e
f..L
= sup{JL(B)
\ B): B
= inf{J.L(B)
\ B) : B
e
e
:E.
/L
1\ v + J.l.. v v = /L + v.
B c
B c
327
Section 36: SIGNED MEASURES 'E .
Let and v be a pair of measur�s on For each A 'E define w(A) = sup{J.L(B) + v(A \ B): B E 'E and B c A } . First, we shall verify that w is a measure, and then, that it is the least upper bound of J.L and v. Clearly, w(A) 0 holds for each A E I: and w(0) = 0. It remains to be shown that w is a-additive. To this end, let { An } be a disjoint sequence of 'E and put A = u:I A" . If B E satisfies B c A, then Proof.
J.L
E
>
I:
00
= L[J.L(A/1 n n
=l
B) + v(A n \ (An n B))]
and so, w(A) E:1 w(A11) holds. For the reverse inequality, note that if w(A) = then w(A) = E: 1 w(A11) = is clearly true. Hence, assume w(A) and let E 0; clearly, w(A11) w(A) holds for all Thus, for each theren exists some B11 E with B11 C A11 and J.L(B11) + v(A11 \ B11) w(A11) - E2- . Obviously, {B n} is a disjoint sequence of 'E, and if B = U: 1 Bn c A, then U:1 (A11 \ B11) = A \ B holds. Moreover, �
oo,
oo
< oo,
n.
< oo
00
= L[ J.L (B11)
n=l
n
>
+
v (A11 \
00
>
>
'E
<
00
B11)] n=l L[w(A,) - E2-"]
= L w(A11) - E
holds for each E as required.
n=l
>
0.
That is, w(A)
>
2:::: 1 w(A11) and so w(A) = E:1 w(A,),
Chapter 7: SPECIAL TOPICS IN INTEGRATION
328
It should be clear that J.L < w and v < w both hold. Now, assume that another measure rr satisfies J.L < rr and v < rr . Let A e l:. If B e l: satisfies B c A, then
J1(B) + v(A \ B) < rr(B) + rr(A \ B)
=
rr(B
U (A \ B)) =
rr(A),
and so, w(A) < rr(A) holds for each A e l:. That is, w < rr, and therefore, w is the lest upper bound of J.L and v. That is, w = 11- v v. The proof of the infimum parallels the preceding one and is left for the reader. To see that J.L 1\ v + J.L v v = J.L + v holds, let A e l:. Then for every B e l: with B c A we have
�-t(B) + v(A \ B) + J.L v v(A)
>
J,L(B) + v(A \ B) + J.L(A \ B ) + v(B) = J.L(A) + v(A), 11- 1\ v(A ) + J,L(B) + v( A \ B) 5 J,L(A \ B) + v(B) + J,L(B) + v(A \ B) = J.L(A) + v(A). By taking the inf and sup over all measurable subsets B contained in A, the preceding inequalities yield J.L 1\ v(A) + J.L v v(A) = J,L(A) + v(A), and the proof is complete. • The formula J.L 1\ v + J.L v v = J.L + v is reminiscent of the familiar identity in vector lattices. The next result describes an order completeness property of the lattice of all measures.
/fa sequence {Jln} ofmeasures satisfiesby J.Lt (i.e., J.Ln < 1-Ln+ l foreachn), thenthesetfunction J.L: [0, oo] defined (A) limJ,L11(A) for each A i s a measur e . Mor e over, J.L11 t Jl holds; that is , J.L is the least upper bound of the sequence {J.Ln l· A I: Proof. 0 < J.L(A) < J.L(0) 0
. Theorem 36.2. e
J..l11
l: --+
l:
=
oo for each e and Clearly, = hold. Also, if A c B, then J.L(A) < J,L(B) is obviously true. For the a-additivity, let {A,} be a disjoint sequence of l:. Let A = A11 • In view of J.Lk(A) = .L::1 J.Lk(An) < �-t(A,) for each k, it follows that J,L(A) < .L:: 1 J,L(A,). On the other hand, for each k we have
U:1
.L:: 1
k
k
i=
i=
•
� J,L(A;) = lim � � J.Ln(A;) = lim J.Ln �l n-+oo 11-HXJ l
(U )
and so, L�t J,L(A;) < J,L(A) also holds. Thus, The verification of J.Ln t J.L is straightforward.
k
i=l
A;
J,L(A)
=
< lim J,L,(A) n-+oc
=
J,L(A),
.L::, J,L(An) , as required.
•
329
Section 36: SIGNED MEASURES
As mentioned before, multiplicationofa measure by - 1 yields a negative-valued set function. For this reason, it is desirable to· consider a -additive set functions that also assume extended negative values. However, if we do this, we run immediately into trouble. Suppose that a set function J..L: I: JR.* satisfies J..L(A) oo and with A n B 0. If J..L is to be additive, then J..L(A B) J..L(A) + J..L(B) J..L(B) must hold, and we face the problem of having to give a meaning to the expression The preceding difficulty can be avoided by excluding from the range of the set function at least one of the infinite values. If this is assumed, then the form does not appear, and the additivity property does not cause problems. To be distinguished from a measure, such a a-additive set function is usually referred to as a signed measure. Its precise definition follows. i s said to be a signed measure if Definition 36.3. A set function J..L : I: it satisfies the following properties: a. J..L assumes at most one of the values oo and -oo, b. J..L(0) 0, and c. J..L is a-additive, that is, if {A,} is a disjoint sequence of members of then /1( u:::l A,) L�I J,.L(A,) holds. IfJ..L is a signed measure, and a disjoint sequence {A,} of satisfies IJ..L( u:, A,)I then (since every permutation of {A,} has the same union as the original sequence) it follows from (c) of the preceding definition that :L: 1 J..L(A,) is re arrangement invariant. Thus, :L: 1 IJ..L(A11) l is convergent in JR. (see Exercise 7 of Section 5). Clearly, every measure is a signed measure. Also, (b) and (c) of Definition 36 together show that every signed measure is finitely additive. The next few results will reveal that signed measures behave in a manner similar to measures. The first one informs us that a signed measure is always subtractive. Theorem 36.4. Let J..L be a signed measure on I:, and let A E I: be such that IJ..L(A)I oo.If B E I: satisfies B c A, then IJ..L(B)I oo and J..L( A \B) J..L( A)- J..L( B). Proof. The identity A (A \ B) B combined with the additivity of J..L imply J..L(A) J..L( A \B) J..L(B). Since J..L( A) is a real number and J..L assumes at most one of the values oo and it follows that both J..L(A \B) and J..L(B) are real numbers. Now. the identity J..L( A \B) J..L(!\) - J..L( B) should be obvious. An immediate and useful conclusion of the preceding theorem is the following: If a set A E :E has a measurable subset of infinite signed measure, then A itself has infinite signed measure. --+
= -oo, = oo - oc
U
==
= =
oo - oo.
oo - oo
--+
JR.*
=
I:,
==
I:
< oo,
1
<
<
=
=
=
U
+
-oo,
=
•
Chapter 7: SPECIAL TOPICS IN INTEGRATION
330
The usual continuity properties of measures are inherited by signed measures. J.L
For a signed measure and a sequence {A,} of �, the fol lowing statements hold: IIff A,A t,!.. A,A and thenJ.L(AdJ.L(An) = J.L(A). is a real numberfor at least one then Theorem 36.5. a. b.
lim
1
1
k,
lim
11-00
J.L(An) = J.L(A).
Proof. Repeat the proof of Theorem 15.4 taking into consideration the con•
clusion of Theorem 36.4.
A measurable set A is said to be a positive set with respect to a signed measure J.L, in symbols A > 0, whenever J.L(E n A) > 0 holds for each E Equivalently, A set A with is a positive set whenever J.L(E) > 0 holds true for all E E C A. Obviously, the empty set is a positive set. Also, it should be clear that any measurable subset of a positive set is likewise a positive set. In addition, any countable union. of positive sets is a positive set. To see this, let {An } be a sequence of positive sets and let A = U:1 An. Now, let B 1 = A 1 , Bn+l = An+l \ U;'= 1 A; for > I , and note that { Bn} is a disjoint sequence such that A = U:1 B11 • Since B11 C An holds for all it follows that each Bn is a positive set. Therefore, for each measurable set E we have J.L(E n A) = J.L( U:1 E n B,) = :L:1 J.L(E n B11) > 0, so that A is a positive set. Similarly, a set A is called a negative set for a signed measure J.L, in symbols As before, measurable A < 0, whenever J.L(A n E) < 0 holds for each E subsets of negative sets are negative, and countable unions of negative sets are likewise negative sets. The next lemma is a basic result for this section and guarantees the existence of nonempty positive sets.
E :E. E �
E�
n
n,
E �.
Let be a signed measure on :E, and let E �with J.L(E) > 0. Then there exists a positive set A suclz that A and > 0.
Lemma 36.6.
J.L
CE
E J.L(A)
Proof. If for every measurable subset B of E we have J.L(B)
> 0, then E is
itself a positive set and there is nothing to prove. Thus, assume that there exists some B with B c E and J.L(B) < 0. By Zorn's lemma (how?) there exists a maximal collection C of mutually dis joint measurable subsets of E such that J.L(C) < 0 holds for each C C. We claim that C is at-most countable. To see this, note first that C = U:1 C,, where C: J.L(C) < II On the other hand, if some Cn is not finite, then it C11 = {C must contain a countable subset of C, say { C 1 , C2, . . } . But then, the measurable
E� E
_1. }.
E
.
Section 36: SIGNED MEASURES
331
set C = U: 1 C; satisfies C c E, and J-L(C) = 2::: 1 J-L(C;) = This implies J-L(E) = J-L(E \ C)+ J-L(C) = which is a contradiction. Thus, each Cu is finite, and so C is at most countable. Hence, the set D = Ucec C belongs to :E , and we claim that A = E \ D is a positive set satisfying J-L(A) > 0. Indeed, since 0 < J-L(E) = J-L(A) + J-L(D) and J-L(D) < 0, it is easy to see that J-L(A) > 0 holds. On the other hand, if J-L(F) 0 holds for some measurable subset F of A, then we can incorporate F into C and violate the maximality property of C. Thus, A is a positive set, and the proof is finished. If is a signed measure on :E and two disjoint sets A and B satisfy A > 0, B < 0, and A B = then the pair (A, B) is referred to as a Hahn decomposition of X with respect to J-L. Loosely speaking, a Hahn decomposition is a splitting of the space into two pieces, where is positive on one of the pieces and negative on the other. Such a splitting always exists and is "essentially" unique, the next theorem shows. -oo.
-oo,
<
11-
•
X,
U
11-
X
as
Let J-L be a signed measure on :E . Then has a Hahn decom position with respect to J-L. That is, there exist a positive set A and a negative setMoreover, B such that = A B and A n B = 0. if (A, B) and (Ah B1) are two Hahn decompositions of with respect to J-L, then a. J-L(A�AI) = J-L(B �B1) = 0, b. J-L(E n A) = J-L(E A 1 ), and Theorem 36.7.
X
X
U
X
n
c. J-L(E n B) = J-L(E n B1) holdfor each E E :E . Proof. We can assume without loss of generality that J-L(E) # holds for each E :E (otherwise, replace by -J-L). > Put a = sup{p.(E): E 0} and note that a > 0. Choose a sequence {Au} of positive sets such that lim J-L(An) = a. Then A = U: 1 Au is a positive set, and since J-L(Au) < J-L(A) < a, it follows that a = J-L(A) Now, we claim that B = \ A is a negative set for J-L. To see this, assume by way of contradiction that there exists some measurable subset C of B with J-L(C) 0. Then, by Lemma 36.5, there exists a positive set E with E c C and J-L(E) > 0. It follows that A U E > 0 and a + J-L(E) = J-L(A) + J-L(E) = J-L(A U E) a < which is impossible. Thus, (A, B) is a Hahn decomposition of with respect to J-L. For the "uniqueness" of the Hahn decomposition, let (A 1 , B 1) be another Hahn decomposition of Since A \ A 1 c A and A \ A 1 = A n A) = A n B 1 C B1, it oo
11-
E
X
< oo.
>
<
oo,
X
X.
Chapter 7: SPECIAL TOPICS IN INTEGRATION
332
follows that tL(A \ A 1 ) > 0 and tL(A \ A 1 ) < 0, that is, tL(A \ A 1 ) = 0. Similarly, tL(A 1 \ A) = 0, and so
tL(A � A J ) = tL((A \ A J ) U (A I \ A)) = tL(A \ A J) + tL(AI \ A) = 0. Now, if E
E
l;, then
tL(E n A) = tL(E n ( (A \ A 1) u (A n A 1) ]) = tL(E n (A \ A 1 )) + tL(E n A n A 1 ) = tL(E n A n A J ) = tL(E n (A1 \ A)) + tL(E n A n A 1 ) = tL(E n [ (A 1 \ A) u (A 1 n A) ]) = tL(E n A J).
By the symmetry of the situation, the corresponding formulas for also true, and the proof is finished.
B
and
B1
are •
If (A, B) is a Hahn decomposition of X with respect to a signed measure JL, then tL and tL restricted to A and B, respectively, are in actuality measures. As we shall see, these two measures determine the structure of JL. -
l;,
Let tL be a signed measure on and let (A, B) be a Hahn decomposition of with respect to JL. Then the three setfunctions
Definition 36.8.
X
tL+(E) = tL�E n A), tL-(E) = -tL(E n B), ltLI(E) = tL(E n A) - tL(E n B) = tL+(E) + tL-(E),
and
l;
forthe each E E are called the positive variation, the negative variation, and total variation ofJL, respectively.
A glance at Theorem 36.7 guarantees that the values of tL+, tL -, and ltL I are independent of the chosen Hahn decomposition. Also, it should be clear that tL+, tL- , and I J.l l are measures on l;. A signed measure tL is said to be a finite signed measure if JL(A) E IR holds for each A E l;. Note that in view of·Theorem 36.4, a signed measure tL is finite if and only if ltL(X)I < oo holds. Similarly the expression "tL is a finite measure" means that 0 < JL(A) < JL(X) < oo holds for all A E l; . Finally, let us say (as usual) that a signed measure is a-finite if there exists a disjoint sequence {An} of l; witb X = U:1 An and tL(An) E IR for each
n.
Section 36: SIGNED MEASURES
333
From Definition 36.8 it follows that holds. This identity is known as the Jordan 1 decomposition of the signed measure decomposition, then in view of the inequalities JJ-/1-+. Moreover, if (A,andB)JJ--is(E)a Hahn < -J.L(B) for each E E :E, it follows that at least (£) < JJ-(A) one of the measures J.L+ and J.L- is a finite measure. Thus, after all, the Jordan decomposition shows that each signed measure is the difference of two measures, at least one of which is a finite measure. Some other useful expressions for the different variations of a signed measure are presented next. Then for every E the Theorem 36.9. Let 11- be a signed measure on. following formulas hold: 1 . JJ- + (£) = sup{JJ-(F): F E :E and F £}. 2. - (E) = sup{ -J.L(F): F E :E and F £}. 3. I JJ - 1 ( £) = sup{:L IJJ-(F;)I: {F;} is finite and disjoim with U F; c £}. Proof. Let (A, B) be a Hahn decomposition of X with respect to JJ-, and let E :E. (1) Clearly, 11- + (£) = JJ-(E n A) < sup{JJ-(F): F E :E and F c £} holds. On the other hand, if F E :E satisfies F c E, then JJ-(F) = JJ-(F A) + JJ-(F n B) < JJ-(F n A) < JJ-(E n A) = JJ-+(E), so that sup{J.L(F): F E :E and F c £} < JJ-+(E). Therefore, the identity in (1) holds. (2) This follows from (1) by observing that 11- - = (-JJ-)+. (3) Let v(E) = sup { L IJJ-(F; )I: { F;} is a finite disjoint collection. of :E with U F; c E } . Then, IJ.LI(E) = JJ- +(E) + 11--(E) = JJ-(E n A) - J.L (E n B) = IJJ-(E n A) I + IJJ-(E n B)l < :E .
E :E
c
11-
c
c :E
E
n
v(E).
1Camille Jordan ( 1 838-1921 ), an eminent French mathematician. He was considered a "universal'' mathematician because he published papers in virtually all areas of mathematics of his time. His main contributions were in group theory.
Chapter 7: SPECIAL TOPICS IN INTEGRATION
334
Now, let { F1 , . . . , F,} be a finite disjoint collection of � such that U�1= 1 F; C E. Then II
L IJ.L( F;)I i=l
=
=
II
II
i=l
i=l
L IJ.L+(F;) - J.L-(F; ) I < L [J.L+( F; ) + J.L-(F;)] n
B IJ.LI( F;) = IJ.LI ( � F;) < IJ.L I(E). 11
This implies v(E) < IJ.LI (E). Thus, IJ.LI(E) = v(E) holds, and the proof of the theorem is complete. • It is easy to establish, but important to observe, the following inequality regard ing the total variation: IJ.L(A) I < IJ.L I (A)
for each A E � -
Summarizing, the different variations of a signed measure J.L satisfy the following identities: a. J.L = J.L+ - J.L- and IJ.L I = J.L+ + J.L-. b. J.L+ 1\ J.L- = 0. c. J.L- = ( -J.L)+. They are, of course, reminiscent of the usual identities of vector lattices. Fur thermore, they suggest that the set of all signed measures forms a vector lattice. Unfortunately, this is not the case. Although the (pointwise) sum of two measures is always defined, the sum of two signed measures may fail to exist for the very simple reason that by adding the signed measures pointwise one might encounter the expression oo - oo. However, if we restrict ourselves to the collection M(�) of all finite signed measures, then we obtain a vector lattice. Note first that a signed measure J.L satisfies J.L E M(�) if and only if IJ.LI(X) < oo. Indeed, if IJ.LI(X) < oo, then clearly, J.L E M(�). On the other hand, if J.L E M(�). then the identity J.L = J.L+ - J.L with either J.L+ or J.L- finite shows that J.L+ and J.L- are both finite, and so IJ.L I (X) = J.L+(X) + J.L-(X) < oo. (Because of the last statement, the finite signed measures are also known as the signed measures of finite total variation.) Clearly, if J.L, v E M(�). then J.L + tt and CXJ.L belong to M(�). where, of course, (J.L + v)(A) = J.L(A) + v(A) and (aJ.L)(A) = CXJ.L(A) for each A E � and a E JR. That is, M(�) is a vector space. Now, if J.L < v means J.L(A) < v(A) for each A E �, then < is an order relation under which M ( �) is a vector lattice. Its lattice
Section 36: SIGNED MEASURES
335
operations are given by J-L v v(A) = sup{,u(B) + v(A \ B): B E :E and B A}, and J-L v(A) = inf{J-L(B) + v(A \ B): B E :E and B A}. The preceding formulas, combined with Theorem 36.9, show that J-L+ = J-L v 0, ,u- = (-,u) v O, and 1 1-L I = J-L v (-,u) in M(:E) are precisely the positive, negative, and total variations of ,u as given by Definition 36.8. The next thing to observe is that l .ull = IJ-L!( X) defines a norm on M(:E). Indeed, (A )I < l.ui( A) < IJ-LI(X) = l .ull for each a. l.ui (X) > 0 holds, and since I J-L E :E, we have 1 1-L I = 0 if and only if J-L = 0. A b. l aJ-LII = la,ui (X) = Ia! l.ui (X) = lal · (IJ-LI(X)) = lal · I J-LI . c. 1 1-L +vii = l .u + vi(X) < (IJ-LI + lvi)(X) = IJ-LI(X) + lvi(X) = l .ull + I vii . ItI J-LisI also true that < lvl holds, then is a lattice norm on M(:E). Indeed, if 1 1-L 11 · 11 I = l.ui (X) < l v i(X) = I vii . Therefore, M(:E) is a nonned vector lattice which is actually a Banach lattice, the next result shows. c
1\
c
·
as
The collection of allfmite signed measures on a a-algebra is a Banach lattice.
Theorem 36.10.
According to the preceding discussion, it remains to be shown that M(:E) is a Banach space. To this end, let {J-L,} be a Cauchy sequence of M(:E). We have to show that there exists some ,u E M(:E) such that lim I J-L, - .u I = o. Let E > 0. Choose some k such that I J-L, - J-Lm I < E for all m, n > k. The inequalities Proof.
1/-+00
show that {,u,(A)} is a Cauchy sequence of real numbers for each A E :E. Let J-L(A) = lim J.t,(A), so that J-L is a real-valued set function. From ( ) it follows that ( ) IJ-L,(A )- J-L(A)I < E for each A E :E and n > k. *
**
336
Chapter 7: SPECIAL TOPICS IN INTEGRATION
Clearly, f.J-(0) = lim f.J-,(0) = 0. For the a-additivity of f.-L , let {A, } be a disjoint sequence of 'E, and let A = U:1 A,. Then p
p
L
[f.-Lk (A ;) - f.-Ln (A ; )]
i=l
<
L l f.-Lk(A;) - f.-Ln(A; )I < lf.-Lk - f.-L,I(A) i=l
< l lf.-Lk - f.-Ln II < E
'
holds for all p and ll > k. Hence, r�=r=l [f.-Lk (A;) - f.-L(A ; )] I � E holds for each p . Now, choose no such that lf.-Lk (A) - 2:f= 1 f.-Lk (A ; )I < E for all p > no, and note that p
p
i=l
i=l
f.-L(A) - L f.-L(A;) < lf.-L(A) - f.-Lk(A)I + f.-Lk (A) - L f.-Lk(A;) p
+ L [f.-Lk (A ;) - f.-L(A; )]
< 3E
i=l
holds for all p > no. Therefore, f.-L(A) = 2::1 f.-L(An), and so f.-L E Finally, combining (**) with Theorem 36.9, we get
M('E).
(f.-Ln - f.-L)+(X ) = sup{f.-Ln(A) - f.-L(A): A E 'E } < E and (f.-Ln
- f.-L)- (X) < E for all n
> k. Hence,
holds for all n > k. That is, lim II f.l-11
- f.-L II
=
•
0, as required.
EXERCISES 1.
Give an example of a signed measure and two Hahn decompositions (A, B) and (A 1 , B 1 ) of X with respect to the signed measure such that A =F A 1 and B =F B I · 2. If 11- is a signed measure, then show that 11-+ A 11-- = 0. 3. If 11- is a signed measure, then show that for each A E :E we have if.LI(A) = sup
{�
if.L(A n ll: I A n } is a disjoint sequence of E with
Q
An
=
}
A .
4. Verify that if 11- and v are two finite signed measures, then the least upper bound 11- v v and the greatest lower bound 11- A v in M(:E) are given by
11- v v(A) 11- A
v(A)
for each A E :E.
=
=
sup{�.t(B) + v(A \ B): B E :E and B c A}. and inf(�.t(B) + v(A \ B): B E :E and B c A }
337
Section 36: SIGNED MEASURES
Let A be the Lebesgue measure on the Lebesgue measurable subsets of 1R. If J.L is the Dirac measure, defined by J.L(A) = 0 if 0 rf. A and J.L(A) = l if 0 E A, describe A v J.L and A 1\ J.L. 6. Show that the collection of all a-finite measures fonns a distributive lattice. That is, show that if J.L, v, and w are three a-finite measures, then
5.
(J.L v v) 1\ w = (J.L 1\ w) v (v 1\ w) and (J.L 1\ v) v w = (J.L v w) 1\ (v v w). [HINT: Every vector lattice is a distributive lattice.] 7. If :E is a a-algebra of subsets of a set X and J.L: :E � 1R* is a signed measure, then show that AJ1.+ n AJ1.- = A IJJ. I· 8. Let J.L and v be two measures on a a-algebra :E with at least one of them finite. Assume also that S is a semiring such that S c :E , X E S, and that the a-algebra generated by S equals :E . Then show that J.L = v on :E if and only if J.L = v on S. [HINT: See Theorem 1 5 . 1 0; see also Exercise l l of Section 20.] 9. Let (X, S, J.L) be a measure space and let f E L 1 (J.L). Then show that v(A) = for each A E
All
v+(A) =
i f dJ.L
defines a finite signed measure on A ll. Also, show that
i t+ dJ.L,
v - (A) =
i !- dJ.L
and l v i(A) =
i Ill dJ.L
hold for each A E All. [HINT: Use Theorem 1 5 . 1 1.] 10. Let v be a signed measure on :E. A function f: X � 1R is said to be v-integrable iff is simultaneously v+- and v- -integrable (in this case, we write J f dv = J f dv+ Jf d v-). Show that a function f is v-integrable if and only if f E L 1 ( Ivi). 11. Show that the Jordan decomposition is unique in the following sense: If v is a signed measure, and J.LI and J.L2 are two measures such that v = J.LI - J.L2 and J.LI 1\ J.L2 = 0, then J.L 1 = v+ and J.L2 = v - . 12. In a vector lattice x11 -1- x means that x11+ 1 ::.: x, for each n and that x is the greatest lower bound of the sequence (x11}. A nonned vector lattice is said to have a-order continuous norm if x, -1- 0 implies lim llx11 ll = 0. a. Show that every L p(J.L) with 1 ::.: p < oo has a-order continuous nonn. b. Show that L00([0, 1]) does not have a-order continuous nonn. c. Let :E be a a -algebra of sets, and let (J.L,} be a sequence of M (:E) such that J.L11 -1- J.L. Show that lim J.L11(A) = J.L(A) holds for all A E :E . d. Show that the Banach lattice M(:E) has a-order continuous nonn. 13. Prove the following additivity property of the Banach lattice M(:E): If J.L, v E M(:E)
disjoint (i.e., I J.LI 1\ jv l = 0), then II J.L + vii = II J.L il + II vii holds. [HINT: In a vector lattice l.t l 1\ IYI = 0 implies j.t Yl = lxl IYI. Reason:
are
+
+
+
+
+
lx yl ?::: l lxl - lyl l = lxl v IYI - lxl 1\ IYI = lxl v IYI = lxl IYI ?::: lx y j.] 14. Let :E be a a-algebra of subsets of a set X and let (J.L11} be a disjoint sequence of MCE). If the sequence of signed measures (J.L11} is order bounded, then show that limiiJ.L,II = 0.
Chapter 7: SPECIAL TOPICS IN INTEGRATION
338
37. COMPARING MEASURES AND THE RADON-NIKODYM THEOREM Again, in this section, :E will be a fixed a-algebra of subsets of a nonempty set X and all set functions will be assumed defined on :E. Two important comparison notions for signed measures will be introduced here. Both notions will be defined in terms of measure properties. The first one is referred to as the "concept of absolute continuity."
sigz z e d measure is said to be absolutely continuous with respect to another signed measuz · e in symbols << or whenever E :E and 0 imply 0. Definition 37 .1. A
A
v J.l, v(A) =
IJ.LI(A) =
v
J.l >> v,
J.l
Some characterizations of absolute continuity in terms of variations are included in the next result.
For a pair of sigz z e d measures az z d the followiz z g state ments are equivalent: << << and << Theorem 37 .2. a.
b. c.
v v+ I vI
J.l
J.l. IJ.ll << I J.L ! .
v-
v,
I J.l l .
(a) => (b) Let A E :E satisfy IJ.L!(A) = 0. If B E I: satisfies B then IJ.LI(B) = 0, and so by our hypothesis v(B) = 0. Thus,
Proof.
v+(A) = sup{v(B): B
E :E
and B
c A}
=
c A,
0.
Similarly, v-(A) = 0, and so, both v+ << IJ.ll and v- << IJ.ll hold. (b) => (c) If IJ.LI(A) = 0, then by hypothesis v+(A) = v-(A) = I vi = v+ + v-, it follows that l v i(A) = 0, and thus I vi << IJ.ll. (c) => (a) It follows from the inequality lv(A)I < l v i (A).
0.
Since •
If a measure v is absolutely continuous with respect to another measure J.l, then it is natural to expect some relationship between the J.,L-measurable and v-measurable sets. The following theorem tells us the precise relationship:
For two measures and on a a-algebra we have the following: << and a subset of X satisfies 0, then 0. fevery<<J.,L-measurable and is a-function finite, thenis alsocv-measurable. holds. In particular, in this case Theorem 37.3. a. b.
lf v l v
J.l J.l
J.l
E
J.l
Ap.
v
J.l*(E) =
A 11
I:
v*(E) =
Section 37: COMPARING MEASURES AND THE RADON-NIKODYM THEOREM 339
Proof. (a) Assume JL*(E) = 0. By Theqrem 15.1 1 , there exists some A E :E
such that E c A and JL(A) = JL*(E). From v << JL it follows that v(A) = 0. Therefore, 0 < v*(E) < v(A) = 0, so that v*(E) = 0. (b) Assume E is JL-measurable such that JL*(E) < oo. By Theorem 15. 1 1 , there exists some A E I: with E c A and JL(A) = JL*(E) < oo. But then JL*(A \ E) = 0, and by (a) we have v*(A \ E) = 0. Therefore, A \ E is v-measurable. The v measurability of E now follows from the identity E = A \ (A \ E). Finally, the preceding conclusion, coupled with the a-finiteness of JL, easily implies that AJ.L c A v holds. • Now, let JL and v be two measures. If v < JL, then it should be clear that v << JL also holds. The converse of the latter is false. For instance, 2JL << JL holds for each measure JL, and 2JL < JL is true only if JL = 0. However, if v << JL holds, then (up to a multiplication factor) JL and v are "locally" comparable in 'the order sense. The exact details follow.
Let be a finite nonzero measure, and let be a a-finite measure.andIfsomeis absolutewith ly continuous with respect to then there exist some such that 0 0 v
Theorem 37.4. v
E >
A E :E
JL,
< JL(A) < oo
EJL(B)
holdsfor all
B E :E
with n.
JL
< v(B)
B C A.
Proof. Let {En} _be a disjoint sequence of I: such that X = U:1 E11 and
Since v is a nonzero measure, there exists some k with JL(E11) < oo for .each v(Ek ) > 0. Choose some E > 0 so that v(Ek ) - EJL(Ek ) = (v - EJL)(Ek ) >
0.
By Lemma 36.6 there exists a measurable subset A of Ek that is a positive set for the signed measure v - EJL satisfying (v - EJL)(A) > 0. Clearly, JL(A) < oo holds. Now observe that JL(A) > 0. Indeed, if JL(A) = 0 holds, then the last inequality yields v(A) > 0, contradicting the absolute continuity of v with respect to JL. On the other hand, the inequality (v - EJL)(B) > 0 for each B E I: with B c A implies E JL(B) < v(B ), and the proof of the theorem is finished. • The opposite notion to absolute continuity is that of singularity. Two signed measures JL and v are said to be singular (or orthogonal), in symbols JL j_ v, if there existtwodisjointsets A andB of:E with AUB = X and I JLI(A) = l vi(B) = 0. If JL is a signed measure, then Theorem 36.7 shows that JL+ and JL- are two singular measures. The singularity concept is characterized in terms of lattice properties as follows.
Chapter 7: SPECIAL TOPICS IN INTEGRATION
340
Two signed measures J.L and v are singular if and only if IJ.L I /\ I v i = o.
Theorem 37.5.
v. Choose two disjoint sets A and B of l; with A u B = X and IJ.LI(A) = lvi(B) = 0. Then Proof. Assume that J.L 0<
..l
IJ.LI A lvi(X) = IJ.L I A lvi(A) + IJ.L I A lvi(B) < I J.LI(A) + lvi(B) = 0,
so that IJ.LI /\ I vi = 0. Conversely, assume that IJ.LI /\ I vi = 0. By Theorem 36.1, for each n there exists some E, E � such that IJ.LI(En) + lvi(E�) < 2- " . Let A, = u:n E;' and then let A = n:1 A, and B = Ac. Note that
n,
holds for all and thus, IJ.LI(A) = 0. Now, obse�e that lvi(A�) = I vi( n�, Ej) < 2-i for all i > implies lvi(A�) = 0 for each n. Therefore, since B = Ac = U: 1 A�, it follows that lvi(B) = 0. Hence, X = A U B and IJ.LI(A) = lvi(B) = 0 hold, proving that J.L ..l v. •
n
The following list of statements presents a number of useful relationships be tween absolute continuity and singularity. The set functions J.L, v, and w are signed measures on a common a -algebra �.
1. If J.L << w and v << w, then IJ.LI + lvl << w. 2. If J.L ..l w and v ..l w, then IJ.LI + I vi ..l w.
3. 4. 5.
If J.L << w and I vi < IJ.LI. then v << w. If J.L ..l w and I vi < IJ.LI. then v ..l w. If v << J.L and v ..l J.L, then v = 0.
Their proofs are straightforward applications of the definitions and Theorems 37.2 and 37.5. For instance, the proof of (5) goes as follows: Since v ..l J.L, there exists some A e � such that lvi(A) = IJ.LI(Ac) = 0. From v << J.L and Theorem 37.2, lvi(Ac) = 0, and so, lvi(X) = lvi(A)+ lvi(Ac) = 0. That is, I vi = 0, so that v = 0. Now, let J.L be a a-finite measure. A classical result of H. Lebesgue asserts that any other a -finite measure v can be det:omposed as a sum of two measures v1 and v2 such that v1 << J.L and v2 ..l J.L. Its proof will be based upon the following simple property of vector lattices: Lemma 37.6. In
a vector lattice
x,
t x mpli
s
i e
Xn 1\
x
y t 1\ y for each y.
Section 37: COMPARING MEASURES AND THE RADON-NIKODYM THEOREM 341
Note first that X11 A y t < x A y h_olds. On the other hand, if x11 A y < z holds for each n and some z, then Theorem 30.1 (5) shows that Proof.
(X A y- .:)+ < (X A y -X11 A y)+ < !x A y -X11 A Yl < X -X11• Hence, X1 <+ x - (x A y - z)+ < x holds for each n. But then X11 t x implies is, x A y is the least upper bound of < z. That (x{x1 AAy soz) that= 0,X11andA so,t x AA y holds, y x y as claimed. 1 y}, •
-
We are now ready to state and prove the Lebesgue decomposition theorem.
Let J.L and v be two a-finite measures on I:. Then there exist llvo l nique measures v1 and v2 such that v = v1 + v2, where v1 J.L and v2 J.L. Proof. Assume first that both J.L and v are finite measures. Then the formula v1(A) = sup{(v A 11J.L)(A): n = 1, 2, ... < v(A) defines a finite measure on (Theorem 36.2). Moreover, if J.L(A) = 0, then clearly (v A nJ.L)(A) = 0 for each n. Thus, v1 (A) = 0, that is, v1 Now, let note that v2 is a finite measure such that v1 + v2 = v. Next, v2observe = v -thatv1,v and A l J .L t v1 implies (J.L + v) A (n + l )J.L = J.L + v A nJ.L t J.L + v1, and so, by Lemma 37.6, V A l )J.L = V A (J.L + v) A (n + )J.L t V A (J.L + VJ ). On the other hand, v A (n )J.L t v1 implies v1 = v A (J.L + v1 ). Thus, 0 < v2 A J.L = (v - vJ) A J.L = V A (J.L + VJ) - VJ = VJ - VJ = 0, and hence, by Theorem 37.5, v2 J.L. For the uniqueness of the decomposition, note first that v1 v2• Indeed, since sets A and B with X = A B v2and J.L, there exist twoThendisjoint measurable J.L implies v1 (B) = 0, so that v1 v2• Now, v2(A) = J.L(B) = 0. v1 assume that another pair of measures cv1 and cv2 satisfies cv1 J.L, cv2 J.L, and 2. Clearly, v1v2 - cvJ.L1 .=Itcvfollows 2 - v2 holds, and on one hand v1 - cv1 J.L, vand= oncv1 thecvother that v1 - cv1 = cv2 - v2 = 0, so that = theandgeneral V2 = VJ For case, choose a disjoint sequence { A11 } of with v(A11) < X = U:, A" . Let v" (A) = v(A n A") and J.L" (A) = J.L(A n A") J.Lfor(A1each 1)
<<
j_
}
I:
<< J.L.
(11 +
1
+1
j_
j_
j_
<<
+
cv2
CV2.
CVJ
oo,
2:.
<<
j_
I:
I: .
j_
U
j_
<<
oo,
Chapter 7: SPECIAL TOPICS IN INTEGRATION
342
n
for each there exists a unique pair of measures (v;', v2) such that v!' << J-L11 , v;' .l J-L11, and Vn = v!' + v2. Now, define 00
v1(A) = L v'; ( A) ll=l
and
00
v2(A) = L v�(A) ll=l
for each A e �. It is a routine matter to verify that v1 << Jl, v2 .l J-L, and v = v1 + v2. The uniqueness of v1 and v2 follows easily from the uniqueness of the decomposition of each v,. • The identity v = v1 + v2 provided by the preceding theorem (where v1 << J-L and v2 _L J-L), is called the Lebesgue decomposition of v with respect to J-L. Now, assume that J-L is a measure and f e L 1 (J-L). Then the formula
for each A e :E defines a (typical) finite signed measure that is absolutely con tinuous with respect to f-L. Sometimes formula (*) is referred to as the indefinite integral of f. A direct verification shows that
hold for each A e :E. . It is natural to ask whether the converse of the preceding statement is true. That is to say, whenever v is absolutely continuous with respect to J1 . does there exist some f e L 1 (J1) such that v is given by (*)? In general, the answer is negative; see Exercise 7 at the end of the section. On the other hand, whenever J-L is a -finite and v is finite, the answer is yes. This result is of great import�ce because of its wide range of applications. It is known as the Radon-Nikodym theorem. It was established first by J. Radon2 (in 1913) for the Euclidean spaces with the Lebesgue measure, and later (in 1930) 0. Nikodym3 extended it to the general case.
Let v be a finite signed measure which is absolutely continuous with respect to a a -finite measure Then there exists a Theorem 37.8 (Radon-Nikodym).
f-L.
Johann Radon (1 887-1956), an Aus1rian ma1hema1ician. He is well known for his con1ribu1ions 1o analysis and especially 10 lhe calculus of variations. 3 0non Martin Nikodym (1 889-1974), a Polish ma1hema1ician. He con1ribu1ed 10 Boolean algebras, measure 1heory, and ma1hematical physics. 2
Section 37: COMPARING MEASURES AND THE RADON-NIKODYM THEOREM 343
unique function f L 1 (J.L) such that = if dJ.L holdsfor all in
v(A)
A E I: .
Proof. (Uniqueness.) This part is straightforward and is left for the reader;
6 at the end of this section.
see Exercise J.L-a.e.)
(The function
f
is, of course, unique
(Existence.) Clearly, v+ and v- are finite measures, both are absolutely contin
J.L,
uous with respect to and v = v+ - v- holds. This shows that we can assume without loss of generality that v is itself a finite measure. By Theorem 37.3, we have :E C AJ.L C A11 • Now, let C
= { g L 1 (p.): g 0 E
>
igdp.
p.-a.e. and
< v*(A) for all A E /1."
g g(x) f(x)}
Observe that C is closed under finite suprema. Indeed, if f, then the two sets E =
{x
E A:
2:
f(x) g(x)}
and
F
= {x
E A:
)·
E C and A E AJJ.,
>
are J.L-measurable (and hence, v-measurable) and disjoint, and E U F = A . Therefore,
i f v gdj.L = £ ! v gdj.L + f. ! v gdj.L = £! dj.L + lgdj.L < v*(E)
f0v g
so that Since
+
v*(F)
=
v*(E U F)
E C. E C, the set C is nonempty. Let
=
a = { Jg dp.: g ) {g,} J g, dJ.L = a.0 sup
EC
v*(A),
< v(X) < oo .
= g1 J gn dJ.L fX Lt(J.L)d J,,xA ! A· fJ,,xA J.L
Choose a sequence of C with lim Define J,, v ···v for each n , and note that {f,, } is a sequence of C with < f,, t and lim f,, = < oo. By Levi's theorem (Theorem 22.8) there exists some such E = t In view of 0 < and that f,, t < v*(A), and the Lebesgue dominated convergence theorem, it follows
a = a. f ffdJ.L JA !n dJ.L
Chapter 7: SPECIAL TOPICS IN INTEGRATION
344
that fA t dJ.L < v* (A) for each A E Ap.; that is, t E C. To finish the proof, we show next that v(A) = fA t d J.L holds for all A E :E. Clearly, w(A) = v*(A) - fA t df.J- for A E Ap. defines a finite measure on Ap. which is absolutely continuous with respect to J.L*. Assume by way of contradiction that w # 0. Then, by Theorem 37 .4, there exists some € > 0 and some B E Ap. such that 0 < J.L*(B) < oo and € J.L*(E) < w(E) for each E E Ap. with E c B. Moreover, the function g = t + € xs > 0 belongs to L 1 (J.L), and since a = Jt dJ.L < Jg dJ.L, the function g does not belong to C. On the other hand, if A E Ap., then
1g dJ.L = 1(t + =
€Xs) dJ.L =
1 t dJ.L + €J.L*(B n A) < 1 t dJ.L + w(B n A)
t df.l- + v * (B n A) - f t d J.L = f t dJ.L + v * (B n A) f 1BnA 1A 1A \ B
< v.-t (A \ B ) + v*(B n A) = v*(A), and therefore g oom��.
E C,
which is a contradiction. The proof of the theorem is now
•
The unique (J.L-a.e.)'function t determined by the Radon-Nikodym theorem is called the Radon-Nikodym derivative of v with respect to Jl., in symbols
dv = t or d v = t dJ.L. dj.L
-
The function t is also referr�d to as the density function of v with respect to J.L. In Section 31 it was shown that if ( X, S, J.L) is a measure space and I < p < oo, then for each g E Lq(f.l-), where * + � = I , the function (**) for t E L p(J.L) defines a continuous linear functional on L p(J.L) such that II Fg II = llg llq· Also, in the same section we proved that the mapping g � Fg from Lq(f.l-) to L;(11-) was a lattice isometry and promised the reader to establish that this lattice isometry is also onto. That is (according to F. Riesz), every member of L ;CJ.L) can be obtained from a function of Lq(J.L) as in (**). With the help of the Radon-Nikodym theorem, we are now in the position to establish this claim.
Let (X, S, f.J-) be a measure space, I < p < oo, and let F be a continuous /ineartunctional on L p(J.L). Then there exists a unique
Theorem 37.9 (F. Riesz).
Section 37: COMPARING MEASURES AND THE RADON-NIKODYM THEOREM 345 l+!
g L(/(J.L), where p
= 1 , such that F(f) = Ifg df.L holdsfor all f Lp(J.L). Moreover, I FII = l gl qProof. g g Step I. Assume that (X, S, J.L) is a finite measure space. v: JR = F(XA) AE E
I
(/
•
E
The uniqueness of istence" of has two steps.
has been discussed before. The proof of the "ex
=
0. Define Ap. -+ by v(A) Ap.- Clearly, v(0) for each Also, if (Au} is a disjoint sequence of measurable sets and then
A = U:1 A11,
II
L XA; - XA i=l
Therefore, by the continuity of
F,
we have
v(A) A= 2::1 v(A11) J.L*(A)=
That is, holds, and hence, v is a finite signed measure on Ap.. Moreover, if E Ap. satisfies 0, then. 0; therefore, 0, so that v is absolutely continuous with respect to By the Radon-Nikodym theorem, there exists a unique function in L 1 such that holds for all E Aw Thus, for every E Ap. we have By linearity, it follows that for every step function Now let
F(O) = v(A) g df.L = J A F(XA) = JxAgdJ.L. l/J.
A = (x E X: g(x)
A
> 0}
v(A) = F(XA) = J.L*. g (J.L) A F(l/J) = J¢gdJ.L
XA =
and
x
B = { E X:
g(x) <
0},
f. l/J n t f = l/J"g+ t fg+ limlll/JnXA - fXAIIp = tPnXAg F Jl/J, g+ dJ.L = F (l/J1 XA) += F(f XA) fg- LI(J.L) F(fX B ) =jg+-Jfg-dJ.L. F(fXA) = jfjg dJ.L. Lp(J.L) fg LI(J.L ) F (f) = J fg dJ.L. g q (J.L). = l g lq- I XEnSgng. = l g(x)l f p(J.L)
and letO < E L p (J.L). Choose a sequence ofstep functions {¢11} with O < Then and 0. By the continuity lim of we have lim < oo, and so, by Levi's theorem Similarly, E and E L I (J.L) and Thus, for every we have and E E To complete the proof of this case, it remains to be shown that EL < n}, andthendefinef,, To this end, let£, {x E X : holds, and moreover, we have Since each /11 is essentially bounded, { , } c L
Chapter 7: SPECIAL TOPICS IN INTEGRATION
346
1 lglq XE dJL 1J.g dJL ..
=
=
( 1 lf,. IP dJL r I
= F( f,. )
< IIFII .
II F I I · (J lgl q XE.. dJL r . I
from which it follows that lg l q XEn d 11-F < II F II < oo for each 11. Levi's theorem now implies that g E Lq(J.L). Step II. The general case. For each A E AJ.L write Lp(A) = {/ E Lp(/1-): f = 0 on Ac}, and note that Lp(A) = {/XA: f E Lp(/1-)}. Also, let
(j
I
C = {A E AJ.I.: Ji.* (A) < oo}. By the preceding case, it should be obvious that for each A unique gA E Lq(A) such that
holds for all f
E
E C
there exists a
L p(/1-). Moreover,
holds. Let a. = sup{ llgAIIq: A E C} < IIFII < oo. Now, observe that if A, B E C satisfy A c B. then gA = gB on A; therefore, lgA I < lgB I. and so, llgAIIq < llgBIIq· It follows that there exists an increasing sequence {A,} of C such that II g A,,lq a < oo. The latter, combined with Levi's theorem, shows that the function g defined by g (x ) = lim gAu (x) for each x E X belongs to Lq(J.L). Clearly, g vanishes outside of the set A = U: 1 A,. Next, we claim that F vanishes on Lp(Ac). Indeed, if F is not zero on Lp(Ac), then in view of the norm denseness of the step functions there exists some B E C with B c Ac, so that F does not vanish on Lp(B). Thus, g8 :;{: 0, and since B n An = 0 for all n , it follows that for each n
t
·
holds. Since lim llgA"IIq = a , it follows that 11gB llq = 0, which is impossible. Hence, F vanishes on Lp(Ac).
Section 37: COMPARING MEASURES AND THE RADON-NIKODYM THEOREM 347
Now, let f E Lp(J..L ). Note first that sinc.e g E Lq(J..L), we have fg E L 1 (J..L ) . Thus, from fgAn 4 fg, IfgA, I < Ifg I , and the Lebesgue dominated con vergence theorem, it follows that lim J fgAn d f.l = J fg d f.l. Similarly, it fol lows that lim II/ XAn - f XA l i p = 0, and so, by the continuity of F we must have lim F(f XA,) = F(f XA). Therefore, F(f)
= =
F(fXA + fXA < ) lim F(fXA,)
n-oo
=
F(fXA) + F(f XA< )
=
j8
lim
t A
n-oo
,
j
=
F( /XA)
dJ..L = tg d f.l ,
as required. The equality I I F II = ll gllq was established in Theorem theorem is now complete.
31. 16. The proof of the •
The conjugate space of L 1 (J.L) will be considered next. If we assume that (X, S, J..L) is a a-finite measure space, then the norm dual of L 1 (J..L) is lattice iso morphic to L00(J.L). For a general measure space this may not be the case (see Exercise 16 of Section 31 ).
Theorem 37.10. Let (X, S, J..L) be a a-finite measure space, and let F be a
continuous linear ftmctional on L 1 (J..L) . Then there exists a unique g such that F(f) =
holdsfor all f
E
Lt(J.L. ). Moreover,
E
L00(J..L )
Jfg df.l
IIFII
L r (J..L) =
=
!l g!l00. /n other words,
L oo(J..L) .
Proof. First consider a finite measure space (X, S, J.L). Then the arguments of
Step I in the proof of Theorem 37.9 also apply here and show that there exists a unique function g E L 1 (J..L) such that F(f) = Jfg dJ..L holds for all f E L 1 (J..L ). Next, we show g E Loo(J..L ) and II F II = 11 8 11oo · Indeed, if E > 0, and A = {x E X: I I F II + E < lg(x)l } , then ( !IF II + E)J.L*(A)
<
i lgl d ig J..L
=
Sgn g df.l
=
F(XA Sgn g)
< II F II · II X ASgn g !I t = II F II · J..L* (A) holds. Thus, J..L * (A) = 0, and so, lg(x)l < II F II + E for almost all x. This shows that g E L00(J.L) and that llgll00 < IIFII + E . Since E > 0 is arbitrary, llgll00 < II F II holds. On the other hand, II F II < llglloo holds trivially, and thus, I I F II = llgll00 •
Chapter 7: SPECIAL TOPICS IN INTEGRATION
348
To extend the theorem to the a-finite case, let {X,} be a disjoint sequence of measurablesatisfying X = U: 1 Xn andJL*(X,) < oo foreach n . By thepreceding case, for each n there exists a unique e L00(J..L ) which vanishes off Xn, with and satisfying dJ..L for each E Lt(J..L). Let !lg,lloo < be the function that equals 8n on each X,. Clearly, is measurable, and since it follows that g E L00(J..L ) . 118lloo < We leave now the details for the reader to verify that dJL holds for all E L 1 (J..L) . •
g, F(fXxn) = Jfg,
I FI I FI .
f
g F(f) = Jfg
g
·
f
As an application of Theorem 37.9 we shall establish that every L p-space with 1 < p < oo is a reflexive Banach space. Recall that a Banach space is said to be reflexive if the natural embedding of the space to its second dual is onto.
Let (X, S, J..L) be a measure space. /f 1 < p < oo, then Lp(J..L) is a reflexive Banach lattice. Proof. Let 4> e L ;*(J..L) . Then for e Lq(J..L ) defines a bounded linear functional on Lq(J..L) , since Theorem 37.11.
lfr(g) = tj>(F�:) g
f l/f(g) = Jgf /(F.r: ) = F�:(f) = Jgf = l/f(g) . =f
By Theorem 37.9, there exists some E Lp(J..L) such that dJ..L holds dJ..L for all e Lq(J..L ). On the other hand, -..tJ> is, the na�ural embedding of L p(J..L) to its s�cond dual L ;*(J..L ) is onto, and hence, Lp(J..L) is a reflexive Banach lattice. •
g
g
EXERCISES 1.
Verify the following properties of signed measures: J.J.
« J.J.. b. v « J.J. and J.J. « w imply v « w. c. If 0 :::: v .:::: J.J., then v « J.J. . d. If J.J. « 0, then J.J. = 0. a.
2. 3.
(l)
Verify statements through (5) following Theorem 37.5. Le J.J. and v be two measures on a a-algebra I: . If v is a finite measure, then show
t
iv l
that the following statements are equ a ent: a.
b.
c.
v
« J.J. holds.
For each sequence {An } of I: with
lim J.J.(An) = 0, we have lim v(An) = 0.
0 there exists some 8
> 0 (depending on E) such that whenever A e I: satisfies J.J.(A ) < 8, then v(A) < E holds.
For each E >
Section 37: COMPARING MEASURES AND THE RADON-NIKODYM THEOREM 349
4.
[fllNT: If (a) does not imply (b), then there exists some E > 0 and a sequence {An} of 'E such that p.(A,) < z-" and v(A,) > €. Let A = n�1 U:::, A;, and note that p.(A) = 0, while v{A) ::: E.] Let p. be a finite measure and let { v,} be a sequence of finite measures (all on 'E) such that v, « p. holds for each n . Furthermore, assume that lim Vn (A) exists in R for each A E 'E. Then show that: a. b.
For each E > 0 there exists some c5 > 0 such that whenever A E :E satisfies p.(A) < 8, then v,(A) < E holds for each n. The set function v: 'E � [0, oo], defined by v(A) = lim Vn{A) for each A E 'E, is a measure such that v « p..
[HINT: Consider :E equipped with the distance d(A, B) = p.(AtlB), where A and B are identified if J.L(A 68) 0. Then {:E, d) is a complete metric space; see Exercise 3 of Section 31. Next, use v, « p. to verify that each v, is a well-defined continuous function on :E . Furthermore, if E > 0, then each C1.: {A E :E: lvn(A) - Vm(A)l ::S E for n, m ::: k} is a closed set, and :E = u�l ck holds. By Theorem 6.18, some ck has an interior point. Use this and the preceding exercise to establish (a). Statement (b) follows from (a).] 5. Let { v,} be a sequence of nonzero finite measures such that lim v,{A) exists in R for each A E 'E. Show that v(A) = lim Vn{A) for A E 'E is a finite measure. [HINT: Consider the finite measure p.{A) = L� l 2-"[vn(A)fvn(X)], and then apply the conclusion of the preceding exercise.] 6. Verify the uniqueness of the Radon-Nikodym derivative by proving the following statement: If (X. S, p.) is a measure space and t E L 1 (p.) satisfies fA t dp. = 0 for all A E S, then t = 0 a.e. 7. This exercise shows that the hypothesis of a-finiteness of p. in the Radon-Nikodym theorem cannot be omitted. Consider X = [0, 1], 'E the a-algebra of all Lebesgue measurable subsets of [0, 1 ], v the Lebesgue measure on 'E and p. the measure defined by p.(C/J) = 0 and p.{A) = oo if A =/= 0. (Incidentally, p. is the largest measure on :E.) Show that: =
=
a. v is a finite measure, p. is not a-finite, and v « p.. b. There is no function t E L 1 (p.) such that v(A) = fAt dp. holds for all A E 'E. 8.
Let p. be a finite signed measure on :E. Show that there exists a unique function t E L 1 (jp.l) such that
holds for all A E :E. 9. Assume that v is a finite measure and p. is a a-finite measure such that v « p.. Let g = dvjdp. E L I (J-L) be the Radon-Nikodym derivative of v with respect to p.. Then show that: a. Iff = {x E X: g(x) > O}, then YnA is a p.-measurablesetforeachv-measurable set A. b. If t E L 1 (v), then tg E L I (J-L) and ft dv = ftg dp. holds.
350
Chapter 7: SPECIAL TOPICS IN INTEGRATION
15.10)
[HINT: Note first that :E c AJl c A v holds, and then (by Theorem that v*(A) = A must hold for each A E Jl; hence, v*(Yc) = For (a) use Theorem 37.3 fA g dJL and the fact that :E is a a-algebra. Now, if¢ = 2::?=1 a; XA; is a v-step function, then
0.
n
II
i=l
i=l
J¢dv = La;v*(A;) = L a;v*(A; n Y) = L a; 1
10.
II
i=l
0
0
' w-a.e.,
holds. [HINT: Use the preceding exercise.] All measures considered here will be assumed defined on a fixed a-algebra :E. a. b. c.
Call two measures JL and v equivalent (in symbols, JL = v) if JL « v and v « JL both hold. Show that = is an equivalence relation among the measures on :E. If JL and v are two equivalent a-finite measures, then show that AJl = A,,. Show that if JL and v are two equivalent finite measures, then
dJL dv · = dv dJL
-
d.
e.
f.
12.
f
g dJL = ¢g dJL,
and so, ¢g E L t (JL). Now, if O � f E L1 (v), then pick a sequence {¢11} of v-step functions such that � ¢n (x) t f (x) for all x E X. To finish the proof, observe that � ¢11g t fg and f¢n8 dv = f¢, dv t ff dv < oo.] Establish the chain rule for Radon-Nikodym derivatives: If w is a a -finite measure and v and JL are two finite measures (all on :E) such that v « JL and JL « w, then v « w and
dv dv dJL = dw dJL · dw
11.
A;nY
-
1 a.e. holds.
If JL and v are two equivalent fin.ite measures, then show that f �---+ f · ��, from L 1 (JL) to L 1 (v), is an onto lattice isometry. Thus, under this identification, L 1 (JL) = L 1 (v) holds. Generalize (d) to equivalent a-finite measures. That is, if JL and v are two equiv alent a -finite measures, then show that the Banach lattices L 1 (J.L) and L 1 ( v) are lattice isometric. Show that if JL and v are two equivalent a-finite measures, then the Banach lattices L p(JL) and L p(v) are lattice isometric for each 1 � p � oo.
Let JL be a a-finite measure, and let AC(JL) be the collection of all finite signed measures that are absolutely continuous with respect to JL; that is,
AC(JL) = { v E M(:E): a. b.
v
« JL}.
Show that AC(JL) is a nonn closed ideal of M(:E) (and hence, AC(JL) with the norm II v ii = jvj(X), is a Banach lattice in its own right). For each f E L 1 (JL), let JL 1 be the finite signed measure defined by JLJ(A) = fA f d JL for each A E :E. Then show that f �---+ JLf is a lattice isometry from L 1 (JL) onto AC(JL).
13. Let :E be a a-algebra of subsets of a set X and JL a measure on :E. Assume also that :E* is a a-algebra of subsets of a set Y and that T: X � Y has the property that r- t (A) E :E for each A E :E*.
Section 37: COMPARING MEASURES AND THE RADON-NIKODYM THEOREM 351 a. b.
1 Show that v(A) = J.L(T - (A)) for each A e �* is a measure on �*. If E L 1 (v), then show that o T E L I (J.L) and
f
f £t dv fxt T dJ.L. =
w
o
c. If J.L is finite and is a a-finite measure on �* such that v « w, then show that there exists a function E L 1 (w) such that
g
fxt T dJ.L = £tg dw o
holds for each f E L 1 (v) .
14. Let (X, S, J.L) be a a -finite measure space, and let g be a measurable function. Show
that if for some 1 < p < oo we have fg E L I (J.L) for all f E Lp(J.L), then g E Lq(J.L), where + = 1 . Also, show by a counterexample that for 1 < p < oo, the a finiteness of J.L cannot be dropped. We can assume that 2: 0 (why?). Then = J for E Lp( ) defines a positive linear functional on L p(J.L). By Theorem 30. 1 0, is continu ous. Now, by Theorems 37.9 and 37.10 there exists some h E Lq(J.L) such that Jfg dJ,L = J fh for each E Lp(J.L). Show that = h a.e. holds.] a measurable function, and 1 < p < 15. Let (X, S, J.L) be a a-finite measure space, oo. Assume that there exists some real number M > 0 such that ¢g E L 1 and J¢g � Mil¢ lip holds for every step function ¢. Then show that:
� �
[HINT:
g
dJ,L
a. b.
f
g
dJ,L g Lq(J.L), where � + � = and ffgdJ,L Mll/llq holds for all f
F(f) fgdJ,L f F g
J.L
(J.L)
1,
E
<
E Lp(J.L).
16. Let 11- be a Borel measure on JR.k and suppose that there exists a constant
c,
c.
c > 0 such
that whenever a Borel set E satisfies A( E) = then J,L(E) Show that J.L coincides with i.e., show that = Let J.L and v be two a-finite measures on a a-algebra � of subsets of a set X such that 17. v « and v # 0. Show that there exist a set E E � and an integer n such that
A, J.L
a. b. 18.
J.L A.
v(E) >
0; and A E � and A C
E imply
�J.L(A) < v(A) < nJ.L(A).
Let J.L be a finite Borel measure on a. b.
A,
=
[ 1, oo) such that
and J.L « J,L(B) = aJ.L(aB) for each a > 1 and each Borel subset B of aB = {ab: b E B}.
dJ.L/dA [dJ.L/dA](x) cjx2
[1, oo), where
If the Radon-Nikodym derivative is a continuous function, then show that > there exists a constant 0 such that = for each 19. Let be a finite Borel measure on (0, oo) such that
J.L
c
x :::. 1.
A,
a. J.L « and b. J.L(aB) = J,L(B) for each a > 0 and each Borel subset B of (0, oo). If the Radon-Nikodym derivative is a continuous function, then show that there exists = for each > 0. a constant > 0 such that
c
[dJ.L/dA](x) cfx
x
Chapter 7: SPECIAL TOPICS IN INTEGRATION
352
38. THE RIESZ REPRESENTATION THEOREM In this section, X will denote a Hausdorff locally compact topological space.
Recall that Cc(X) denotes the vector lattice of all real-valued continuous functions defined on X having compact support. That is, f: X -+ 1R belongs to Cc(X) if and only if f is continuous and vanishes off a compact set. The main purpose of our discussion here is to characterize the positive linear functionals on Cc(X). As usual, a linear functional F on Cc(X) is said to be positive if 0 < F(f) holds whenever 0 < f E Cc(X). Clearly, every continuous function on X is Borel measurable. Thus, if J.L is a Borel measure on X (i.e., J.L is a measure on the a-algebra B generated by the open sets and satisfies J.L(K) < oo for each compact set K), then the formula
f dJ.L, f
F(f) = f
E Cc(X),
defines a positive linear functional on Cc(X). In general, if F is a positive linear functional on Cc(X) and a Borel measure J.L on X satisfies F(f) = J f dJ.L for each f E Cc(X), then J.L is called a representing measure for F (or that J.L represents F). The following basic result (known as the Riesz representation theorem) will be obtained: Cc(X)
Every positive linear functional on i s repr e sented by a unique regular Borel measure. Recall that a Borel measure J.L on is called a regular Borel measure (see X
Definition 18.4) if J.L satisfies the following two extra properties:
B
B
a. J.L( ) = inf{J,L(V): V open and c V } for each Borel set B; and b. J.L(V) = sup{J.L(K): K compact and K c V } for each open set V. Of course, the preceding two conditions are enough to ensure that whenever a Borel set has finite measure, then
B
J.L(
B = sup{J.L(K): K compact and K B) c
)
holds; see Lemma 1 8.5. Let us start by introducing some standard notation. Let V be an open set, and let < I holds for each E X f E Cc(X). The symbol f -< V means that 0 < and Supp f C V. Similarly, for a compact set K and a function f E Cc(X), the I for notation K -< f means that 0 < f < I holds for all x E X, and f all x E K. The notation K -< f -< V simply means that K -< f and f -< V both hold. In this terminology, for instance, Theorem I 0.9 can be stated as follows: If K C U7= 1 V; K V; fJ , . . . , fn E Cc(X) f; -< V; 2:7= 1 f; = 1 K. Now, let F be a positive linear functional on Cc(X). For each open subset V of X , define
f(x)
x
(x) (x) = holds with compact �nd each open, then there exist functions such that for each i and on J.L(V) = sup{F(f): f
-<
V}.
353
Section 38: THE RlESZ REPRESENTATION THEOREM
Clearly, 0 < JL(V) < oo holds for each open set V . It is easily seen that JL(V) < JL(W) holds for every pair of ope·n sets with V c W (since f -< V implies -< W). This observation allows us to extend the set function JL from the open sets to all subsets A of by defining
f
X
.u(A) = inf{.u(V): V open and A
c
V}.
As expected, the set function .u is an outer measure.
Let be a positive linear functional on Cc(X). Then the set function (previously defined) is an outer measure on X. Proof. 0 X. 0, 0. c B, .u(B) Theorem 38.1.
F
.u
Clearly, < JL(A) < oo holds for every subset A of Since f -< 0 if and only if f = it follows that .u(0) = Also, it should be obvious that if A then .u(A) < holds. To complete the proof, it remains to be shown that .u is a-subadditive. To this end, let {A,} be a sequence of subsets of X . Let A = U:1 A,. If 2::1 .u(A,) = oo, then .u(A) < 2::1 ,u.(A,) is obvious. Hence, assume 2::1 choose an open set Vn such that A, c V, JL(A,) < oo. Let E > 0. For each and .u(V,) < JL(A,) + E2-". Put V = U: 1 Vn . Now, if f -< V holds, then K = Supp f c U:1 V,, and in view ofthe compactness of K there exists some such , E that K C U7,� 1 V, . By Theorem 10.9, there exist functions /1 such that f,, -< V, for n = I, . . . , m and 2:�'= 1 f,, = 1 on K . Clearly, < 2::7= 1 holds, and by the positivity of F it follows that
n,
• . . .
F(f) <
m
111
oo
m fm Cc(X) f, f
oo
L F(f,, ) < L1 JL(V,) < L ,U (Vn) < L JL(A11) + E . n= n=1 11= 1 11=1
Thus, since thelatterholdsforeach f -< V , it follows that .u(V) < 2::1 JL(A,)+E. Now, notice that A c V implies .u(A) < .u(V) < 2::, .u(A,) + E for all E > 0, from which it follows that JL(A) =
Jl
(Q ) � A..
The proof of the theorem is now complete.
<
JL (A ., ). •
The outer measure .u determined by the positive linear functional F is called the outer measure induced by F (or the outer measure associated with F). The next goal is to show that .u is a regular Borel measure on B.
Let be a positive linear functional on Cc(X), and let be set i s J L -measurable outer measure. Then ever y Borel (i. e . itsandinduced , restricted to is a regular Borel measure. Theorem 38.2. JL
F
B
.u B C AJL),
354
Chapter 7: SPECIAL TOPICS IN INTEGRATION
Proof. The proof goes by steps.
Step I. For each compact set K, we have J.L(K) < oo. Let K be a compact set. Choose an open set V with compact closure such that such that V g. K C V. By Theorem 10.9, there exists a function g E Now, if f V holds, then f < g, and so F(f) < F(g). Hence,
Cc(X )
-<
-<
J.L(K) < J.L(V) = sup{F(f): f -< V} < F(g) < oo. Step 11. IfK1 and K2 are disjoint compact sets, then J.L(KI UK2)=J.L(KJ)+J.L(K2)· In view of the a -subadditivity of J.L, it suffices to show JJ. (K 1 ) + J.L(K2) < J.L(K1UK2). Since K1 c K2 holds, there exists (by Lemma 10.7) an open set V1 with compact closure such that K1 C V1 C V1 C K2. Thus, K2 C (VJ)c holds. Let V2 = (VJ)c, and note that V1 and V2 are two disjoint open sets. Now, let E > 0. Choose ar1 open set V such that K1 U K2 c \l and J.L(V) < J.L(KJ UK2)+E; clearly, K1 C VnV1 and K2 C V nV2. Next, selecttwo continuous functions f and g such that f -< V n V1 , g -< V n V2 , J.L(V n \11 ) < F(f) + E, and J,L(V n V2) < F(g) + E. From VI n v2 = 0. it follows that f + g -< v Now, 0
note that
J.L(KJ ) + J.L(K2) < J.L(V n V1) + J.L(V n V2) < F(f) + F(g) + 2E = F(f + g) + 2E < J,L(V) + 2E < J,L(KJ U K2) + 3E, 0. Therefore, J.L(K 1) + J.L(K2) < J.L(K 1 U K2), as claimed. StepIll. Forevei)'Subset A ofX have J.L(A)= inf{J,L(V): V open and A C V}. This is precisely the definition of J.L. Step IV. For every open set V, J.L( V) = sup{J.L(K): K compact and K C V}. To see this, let V be an open set and let r E JR. satisfy r < J.L(V). Choose a continuous function f V with r < F(f), and let K Supp f. Now if W is an open set such that K c W, then f W holds, and so r < F(f) < J.L(W). holds for all E
>
we
-<
Therefore,
=
-<
J.L(V) > J,L(K) = inf{J,L(W): W open and K
c
W } :;:: F(f) > r
holds for every r E IR with r < J.L(V ) : and the desired identity follows.
Step V. Every Borel set is J.L-measurable, that is, B C Aw Note first that if K is a compact set and V is an open set disjoint from K, then J.L(K) + J.L(V) = J.L(K U V) holds. Indeed, by Step IT, for every compact subset K1
Section 38: THE RIESZ REPRESENTATION THEOREM
355
of V we have J.L(K) + J.L(K
I) =
J.L(K
u
K I) < J.L(K
u V)
< J.L(K) + J.L (V) ,
and the claim follows from the identity of Step IV. Since Ap.. is a
•
and the proof is finished.
Now, we come to the main result of this section. Namely: The regular Borel measure, induced by a positive linear functional on Cc(X), is the only regular Borel measure that represents F.
F
For ever y positive li there exists a unique regular Borel measure such
Theorem 38.3 (The Riesz Representation Theorem). Cc(X),
F on nearfu.nctional that
J.L
F(f) = Jf dJ.L holds for ever y f E Cc:(X). Moreover, the representing measure is the regular Borel measure induced F on X. v X Proof. (Uniqueness.) J f dJ.L = J f dv f Cc(X). = by
that
Let J.L and be two regular Borel measures on holds for each E We shall show that J.L(A)
such v(A )
Chapter 7: SPECIAL TOPICS IN INTEGRATION
356
holds for each Borel set A. In view of the regularity properties of J.l and v , it is v(K) holds for each compact set K. enough to establish that J.,L(K) To this end, let K be a compact set. Given E > 0, choose an open set V such that K c V and J.,L�V) J.,L(K) + E. By Theorem 10.9, there exists a function X\', we have E Cc(X) such that K -< -< V . But then, since XK
f
=
x
z) = (x JR."
4Alfred Haar ( 1 885-1933), a Hungarian mathematician. He studied orthogonal systems of functions and partial differential equations and is best remembered for his introduction of the integral that bears his name on Hausdorff locally compact topological groups.
Section 38: THE RIESZ REPRESENTATION THEOREM
363
and fa defined respectively by
fa (x) = f(ax)
and
fa (x) = f(xa)
for each x E G. Note that fa and fa both belong to Cc(G) and that fa = fa holds if G is commutative. A linear functional F: Cc(G) ---+ lR is said to be left invariant (resp. right-invariant) if F(f) = F(fa) (resp. F(f) = F(/0)) holds for all a E G. Here is the fundamental result regarding left-invariant linear functionals.
Theorem 38.8 (Haar). !fG isaHausdorfflocallycompact topological group,
then there exists a unique (aside of scalar factors) non-zero positive linear functional H: Cc(G) ---+ lR which is left-invariant. By the symmetry of the situation, there exists, of course, also a unique non-zero positive linear functional on Cc(G) which is right-invariant. The unique non-zero left-invariant positive linear functional Cc(G) ---+ lR is called the left-Haar integral of G. If G is commutative, then the left- and right-Haar integrals coincide and the common integral is called the Haar integral of G. Now, let G be a Hausdorff locally compact topological group and let H: Cc(G) ---+ lR be its unique left-Haar integral. By the Riesz Representation Theorem 38.3, there exists a unique (aside of scalar factors) regular Borel measure f..LG satisfying
H:
H(f) =
l f(g)dJ..LG(g).
It is easy to prove that the left-Haar measure satisfies f..LG (aA) = J..L G (A) for each Borel set A. The measure /1-G is called the left Haar measure of G. Here are two examples of the Haar integral and the Haar measure on two familiar Hausdorff locally compact topological groups.
1. 2.
Consider the additive Hausdorff locally compact topological group G = JR." . Its Haar integral coincides with the Lebesgue integral and its Haar measure is the Lebesgue measure. Consider G = (0, oo) as a commutative group under the ordinary multipli cation. Then the Haar integral H: G ---+ lR is given by H (f) = dx for for each f E Cc(G) and the Haar measure satisfies J.LG(A) = each Borel subset A of G.
J000 f�x> JA �r
For proofs and details regarding the Haar measure and Haar integral we refer the reader to [14, Chapter 6].
364
Chapter 7: SPECIAL TOPICS IN INTEGRATION
EXERCISES Unless otherwise specified, in the exercises below X will denote a Hausdorff locally compact topological space. 1.
2.
3.
If X is a compact topological space, then show that a continuous linear functional F
IIFII holds. Let X be a compact topological space, and let F and G be two positive linear func tionals on C(X). If F( 1) + G(1) � IIF - Gil, then show that F A G = 0. [HINT: Show that F v G(1) = !I F - G il.] Let co(X) = { f E C(X): V E > 0 3 K compact with lf(x)l < E V ¢ K}. Show on.C(X) is positive if and only if F(1) =
x
that:
a. b.
4.
5.
6.
7.
F be a positive linear functional on Cc(X), and let 11- be the outer measure induced by F on X. Show that if 11-* is the outer measure generated by the measure space Y R then " * hnldc fnr t J Let ( \•• '
"\
( A \ - ,_ ", (A \ ._,
'""'""'' "Uh" .. -• -• "' -'-'-
Let 11- and v be two regular Borel measures on X. Then show that 11- � v holds if and - ' ,_.,,
••
•• ,.-
\• •1
•
••-•
u
&-•
I '-
V.I.
.IL
ff dJJ, 2: f f dv
for each 0 � f E Cc(X). Fix a point x E X, and define F(f) = f(x) for each f E Cc(X). Show that F is a positive linear functional on Cc(X) and then describe the unique regular Borel measure 11- that satisfies F(f) = Jf dJJ, for each f E Cc(X). What is the support of 11-? Let X be a compact Hausdorff topological space. If 11- and v are regular Borel measures, then show that the regular Borel measures 11- v v and 11- A v satisfy only if
a.
b.
8.
co(X) equipped with the sup norm is a Banach lattice. The norm completion of Cc(X) is the Banach lattice co(X).
Supp(JJ- v v) = Supp 11Supp(JJ- " v) c Supp 11-
u Supp v, and n Supp v.
Use (b) to show that if Supp 11- n Supp v = 0, then 11- ...L v holds. Also, give an example for which Supp(JJ- " v) =f: Supp 11- n Supp v.
Characteriz� the positive linear functionals F on Cc(X) that are also lattice homomor phisms, that is, F(f v g) = max{F(/), F(g)} holds for each pair f, g E Cc(X).
· [HINT: Let 11- be the representing regular Borel measure for F. Show that Supp 11contains at most one point.]
9.
If X is an uncountable set, then show that a.
c;(X) is not separable, and
b.
C [O, 1] (with the sup norm) is not a reflexive Banach space.
x E X define the positive linear functional F.\·(f) = f(x). Show that II Fx - Fy II = 2 if x =f: y. For (b) combine (a) with Exercise 8 of Section 27 and [HINT: For every
Exercise 12 of Section 12.]
10.
For a finite signed measure 11- on B, show that the following statements are equivalent: a.
b. c.
J1 belongs to Mb(X).
11-+ and 11- - are both finite regular Borel measures. For each A E B and E > 0, there exist a compact set
K £ A £ V such that IJJ-(B)I
K and an open set V with B £ V \ K.
< E holds for all B E B with
Section 38: THE RIESZ REPRESENTATION THEOREM 11.
A sequence {xn} in a nonned space is said -to converge weakly to some vector x if lim f(x11 ) = f(x) holds for every continuous linear functional f. a. b.
12.
365
Show that a sequence in a nonned space can have at most one weak limit. Let X be a Hausdorff compact topological space. Then show that a sequence {J,,} of C(X) converges weakly to some function f E C(X) if and only if Un } is nonn bounded and lim /nCr) = f(x) holds for each x E X.
[HINT: For (b) use Theorem 28.8 and the Riesz representation theorem.] Let J-L be a regular Borel measure on X, and let f E L 1 (J-L). Show that the finite signed measure v. defined by
v(E) =
Lt
dJ-L
for each Borel set E. is a (finite) regular Borel signed measure. In other words, show that v E Mb(X). [HINT: Use the fact that {x E X : f(x) =f. 0} is a a-finite set.] 13. Generalize part (3) of Theorem 38.5 as follows: If J-L and v are two regular Borel mea sures on a Hausdorff locally compact topological space and one of them is a-finite, then show that J-L 1\ v is also a regular Borel measure. 14. Show that every finite Borel measure on a complete separable metric space is a regular Borel measure. Use this conclusion to present an alternate proof of the fact that the Lebesgue measure is a regular Borel measure. 15. Let X be a Hausdorff compact topological space. If ljJ: X � X is a continuous func tion, then show that there exists a regular Borel measure on X such that
holds for each f E C(X). [HINT: If Lim is a Banach-Mazur limit (see Exercise 7 of Section 29) and w E X is a fixed point, then consider the positive linear functional F: C ( X) � 1R defined by
2 F(f) = £im(f(l/J(w)). /(l/J (w)), f(ljJ3 (w)), . . ).] .
16.
This exercise gives an identification of the order dual c;(X) of Cc(X). Consider the collection M(X) of all fonnal expressions J-L 1 - JJ- 2 with J-L 1 and JJ-2 regular Borel measures. That is,
M (X) = {JJ- 1 - JJ-2: JJ- 1 and JJ-2 are regular Borel measures on X}.
a. b.
Define JJ-1 - JJ-2 = v1 - v2 in M (X) to mean J-L 1 (A) + v2( A) = v1 (A) + J-L2(A) for all A E B. Show that = is an equivalence relation. Denote the collection of all equivalence classes by M(X) again. That is, JJ- 1 - JJ-2 and v1 - v2 are considered to be identical if JJ-1 + v2 = v1 + JJ-2 holds. In M(X) define the algebraic operations
(JJ-1
- JJ-2) + ( v i
- v 2)
=
a ( J-L I
- JJ-2 )
=
(JJ- 1 + v t ) - (JJ-2 + V2).
{
aJJ-1 - aJJ-2 ( -a)JJ-2 - (-a )JJ- 1
if a ::: 0 if a < 0 .
366
Chapter 7: SPECIAL TOPICS IN INTEGRATION Show that these operations are well-defined (i.e., show that they depend only upon the equivalence classes) and that they make M(X) a vector space. c. Define an ordering in M(X) by J.q - J..L2 2: v1 - v2 whenever J..L I(A) + v2(A) > v1 (A) + J..L2(A) holds for each A e B. Show that > is well defined and that it is an order relation on M(X) under which M{X) is a vector lattice. d. Consider the mapping J..L = J..L l - J..L2 �---* Fp. , from M(X) to c;-(X), defined by Fp.(f) = Jf dJ..L 1 - J f dJ..L2 for each f E Cc(X). Show that Fp. is well-defined and that J..L �---* Fp. is a lattice isomorphism (Lemma 38.6 may be helpful here) from M(X) onto c;-(X). That is, show that c;-(X) = M(X) holds.
17. This exercise shows that for a noncompact space X, in general c;(X) is a proper ideal of c;-(X). Let X be a Hausdorff locally compact topological space having a sequence (On} of open sets such that On c On+ I and 011 :f:. On+ l for each n, and with X = U� 1 On. a. Show that if X is cr-compact but not a compact space, then X admits a sequence (0n } of open sets with the above properties. b. Choose XJ E 0: a.nd Xn E 0;; \ On-l for n � 2. Then shO\V that F (f) =
L f(xn ), 00
n=l
for f E Cc:(X),
defines a positive linear functional on Cc(X) that is not continuous. c. Determine the (unique) regular Borel measure J..L on X that represents F. What is the support of J..L?
39. DIFFERENTIATION AND INTEGRATION In this section the differentiability properties of a Borel signed measure on lR.k will be studied. The derivative of a measure will always be taken with respect to the Lebesgue measure. For a differentiation theory of arbitrary set functions in a more general context, the interested reader is referred to Chapter 8 of [34]. The results obtained about differentiation of signed measures will be used to derive the classical properties of functions of bounded variation. Throughout this section, the domain of all signed measures will be the a-algebra B of all Borel sets of lR.k. Also, unless otherwise specified, the expression "al most everywhere" will be synonymous with "almost everywhere with respect to the Lebesgue measure." The main Borel sets considered for the purposes of dif ferentiation will be the open balls. Recall that the open ball with center at x and radius r is the subset of lR.k defined by {y E lR.k: - y II < r}, where 2 Y ll = k Y; ) 1 2 • Occasionally, measures defined on the Borel subsets of an open set V of 1Rk appear. By assigning the value zero to the Borel subsets of vc, it is easy to see that these measures can be assumed defined on all Borel sets oflR.k . Therefore, theorems about Borel measures on lRJ: can be applied equally well to such a situation.
llx -
(Li=l (x; -
!
.
llx
Section 39: DIFFERENTIATION AND INTEGRATION
367
Let J.L be a signed measure on the a-algebra of all Borel sets of 1R.k. For each x E 1Rk and r > 0, we define the following two extended real numbers: L'l;(x) = sup t.,(x)
=
inf
l l ��;; : J.L(B) >..(B)
I, I·
: B is an open ball of radius < r and x E B -
B is an open ball of radius < r and x E B
and
Observe that x is not required to be the center of the open balls B where the sup and inf are taken. Clearly, -00 < L'lr (X) < L'l;(x) < 00 holds for each x E 1Rk and all r > 0. Also, it should be obvious that if 0 < r < then l:l.;(x) < L'l;(x), and fls(x) < l:l.r (x) both hold for each x. Therefore, the limits
s,
D* J.L(X) = lim L'l;(x) r�O
and
D*J.L(x) = lim L'lr(x) r�O
k exist in 1R* , and they satisfy -oo < D * J.L(X) < D * J.L(x) < oo for each x E 1R . The extended real numbers D*J.L(X) and D * J.L(x) are called the lower and upper derivatives of J.L (with respect to >... ) at the point x. If they are real and equal, then this common value is called the derivative of J.L at x.
Definition 39.1.
Let be a signed measure on Band J.L
-
k x E 1R .
oo < D*J.L(x) = D* J.L(X) < oo
If
holds, then i s said to be differentiable at the point The common value is called the derivative of at and is denoted by that is, J.L
J.L
x. DJ.L(x),
x
m i�:� m
Here is a rephrasing of the preceding definition: A real number satisfies = D J.L(X) if for each E > 0, there exists some 8 > 0 such that I - I <E holds for each open ball B with x E B and with a radius less than 8. An alternative definition of the derivative using sequences is this: A signed measure J.L on is differentiable at some point x E 1Rk if and only if there exists holds for every sequence (B11} of open a real number such that lim balls containing x whose radii tend to zero. (The number is, of course, the derivative DJ.L(X) ) The expression "D J.L(x) exists" is synonymous (as usual) with "J.L is differen tiable at x ." It is easy to see that if two signed measures J.L and v are differentiable
m
Bm
i�::�
.
=
m
m
368
Chapter 7: SPECIAL TOPICS IN INTEGRATION
at some point x and their sum J.L + v defines a signed measure, then J.L + v is also differentiable at x and D(J.L + v)(x) = DJ.L(X) + Dv(x) holds. Similarly, if J.L - v defines a signed measure, then D(J.L - v)(x) = DJ.L(X) - Dv(x). Our first objective is to show that a finite signed Borel measure J.L that is ab solutely continuous with respect to the Lebesgue measure is differentiable al most everywhere and that the derivative D J.L coincides with the Radon-Nikodym derivative dJ.L/dA.. This will be the key differentiation result for this section. To prove this theorem, we need some preliminary discussion. The reader can verify easily the following property of the Lebesgue measure: If A is a subset of JR.k and r > 0, ihen the sei r A = irx : x E A} satisfies A.(r A) = ,-k A.( A).
This, combined with the fact that >.. is translation invariant, shows that if B is an open ball with radius r and B * is another open ball with radius ar, then A.(B * ) = ak A.(B) holds. The latter will be used in the proof of the next lemma. Lemma 39.2. Let B 1 ,
•
•
, B, be open balls in IRk . Then there exist painvise
•
disjoint open balls Bk, , . . . , Bk"' among the B1 , . . . , Bn such that
Proof. Let r; be the radius of B;. Rearranging the balls, we can assume without loss of generality that r1 > r2 > · · · > �'n· Put k1 = 1, and let k2 be the smallest integer (if there is any) such that Bk2 is
disjoint from Bk , . Let k3 be the smallest integer such that Bk3 is disjoint from Bk, and , Bn } has been exhausted. This Bk2 • Continue this process until the finite set { B 1 , gives us the open balls Bk, , , Bk"' , and we claim that they satisfy the property of the lemma. To see this, let Ak; be the open ball with the same center as Bk; and with radius three times that of Bk; · Now, note that each B; must intersect some Bki · It follows that B; c Akj must hold for that j. �us, u�=l B; c U7=1 Akj' and so, .
.
as claimed.
•
•
•
.
•
Section 39: DIFFERENTIATION AND INTEGRATION
369
k. �:(x), JR. k a r > { x JR. : �:(x) > a} {r11 } rn oo oo l k {x E JR. : D*JJ,(X) > a} = u n X E JR.k : �;,,(x) > a + m1 1 e JR.k : D* JJ,(x) > a} a e JR. Lemma 39.3. Let be a Borel measure on JR.k , and let A be a Borel set such that JJ,(A) = Then DJJ,(x) = holdsfor almost all points in A. k, A Proof. JR. (x) x. DA({x JJ,(A ) = 0. *JJ,(x)e A :D*D*JJ,JJ,(x) 0}) = > 0. E A({x e A: D* JJ,(X) > E}) = 0 > E = {x e A: D*JJ,(x) > E}, E E > 0. K E V A c V. e K, D *JJ,(x) > E B K , x B c V JJ,(B) > EA(B). B , B11 K. Bk" . . . , Bkm B B11
Now, let JJ- be a signed measure on Ftom the definition of it is easy to see that for each e 1R and 0 the set e is an open set. Thus, if is a sequence of positive real numbers with .J,. 0, then the identity
m=l 11= 1
shows that {x is a Borel set for each will be used in the next key lemma.
This observation
JJ-
0.
0
x
Let J.l be a Borel measure on and let be a Borel set such that < Since JJ- is a measure, 0 < holds for each To establish the lemma it must be shown that 0. For this, for each it is enough to show that and note that is a To this end, fix Let Borel set. Let be an arbitrary compact subset of and an arbitrary open set such that Now, if x then holds, and so, there exists an open ball containing with and In view of the there exists a finite number of such open balls 1 , that compactness of cover Let that satisfy be the disjoint open balls among 1 , , Lemma 39.2. Then •
.
•
A
•
•
Since JJ- is also regular Borel measure (Theorem 38.4), it follows that But then the regularity of implies for every compact subset K of and we are done.
E.
•
A(A(£)K ) ==
0 0,
•
Now, we come to the main differentiation theorem regarding Borel measures.
k that is absolutely Ever y finite si g ned Borel measur e on JR. continuous with respect to the Lebesgue measure (i. e . , << A), is differentiable almost eve1ywhere. coincides (a. e . , of course) with the Radon Moreover, i t s derivative D Nikodym derivative dJJ,/dA; that is, DJJ, = dJJ,/dA holds. Theorem 39.4.
JJ-
JJ-
JJ-
Chapter 7: SPECIAL TOPICS IN INTEGRATION
370
f = djJ-jd). be the Radon-Nikodym derivative provided by Theorem That is, JJ-(E) JE ! d). holds for each Borel set E. Fix a real number r, and let A {x lR.k : f(x) r}. Clearly, A is Lebesgue measurable, and by Theorem ).(A) < there Also, by Theorem exists a Borel set such that A C F and ).(F) ).(A); clearly, \A) k Next, consider the Borel measure v on lR. defined by v(E) lFnE { [f(x) -r] d).(x) l{AnE [f(x) -r] d).(x) for each Borel set E. Note that if B is an open ball, then JJ-(B) -r).(B) la{ [f(x) -r] d).(x) < lan{ A [f(x) -r] d).(x) v(B) Proof. Let
E L 1 ().)
37.8.
=
=
>
E
22.5,
15. 1 1 ,
oo.
F
).(F
=
=
0.
=
=
=
=
JJ-(B) < r + v(B) ).(B) ).(B)
B, from which it follows that D* JJ-(X) � r + D * v(x) for each x lR.k . Since it follows from Lemma that D v(x) holds for almost all x. and consequently, (*) implies D* JJ-(X) < r for almost all x *D JJ-(X) < rThus� for almost all E A Since f (x) < r implies A it follows that the set Er {x lR.k : f(x) < r < D*JJ-(x)} satisfies Er c A Therefore, by the last conclusion, each E,. has Lebesgue mea sure zero. But then {x lR.k : D * JJ-(x) f(x)} C UreQ Er. where Q is the set of all rational numbers, shows that {x lR.k : D* JJ-(X) f(x)} has Lebesgue measure zero. Applying the preceding arguments to -JJ- , and taking into consideration that d(-JJ-)jd). -f and D*(-JJ-) -D*JJ-, we easily infer that the set
holds for each open ball
E
v(Fc) E pc.
=
39.3
0,
x
=
=
E pc, x E
c.
E
c.
E
>
E
=
=
has Lebesgue measure zero. In other words,
>
0
c,
Section 39: DIFFERENTIATION AND INTEGRATION
holds for almost all x E 1Rk . That is, a.e. holds, as desired.
371
DJ.l.(X) exists for almost all x and DJ.1. = f
•
As a first application of the preceding theorem, let us establish a classical result.
If E is a Lebesgue measurable subset of JR., then >..(En(x�E,x+E)) = 1 for almost all x in E, and limE-o+ >..(En(x�E ..t+E)) = 0 for almost all x in Ec. limE-o+
Theorem 39.5. 1. 2.
Proof. Without loss of generality, we can assume that )..(£)
the finite Borel measure J.l. on 1R defined by
J.l.(A) = )..( £ n A) =
< oo.
Consider
J. XE d).. .
Clearly, J.l. << ).. and dJ.1./d).. = XE · By Theorem 39.4, DJ.l. the formulas in ( 1 ) and (2) follow.
= XE a.e. holds, and •
A point x of a Lebesgue measurable subset E of 1R for which ( 1 ) in the pre ceding theorem holds is referred to as a density point of E. In this terminology, Theorem 39.5( 1 ) is usually stated as follows: If E is a Lebesgue measurable subset
of
JR., then almost every point of E is a density point.
Now, let J.l. be a signed Borel measure. If J.l. is singular with respect to the Lebesgue measure ).. , then we shall show that its derivative equals zero almost everywhere. This, coupled with Theorem 39.4, will yield the following important differentiation result.
Every signed Borel measure J.l. on JR.k is differentiable almost everywhere. Moreover, if J.1. = J.l.l + J.l-2 is the Lebesgue decomposition of J.1. (i.e., J.l- 1 << ).. and J.l-2 ..L >..), then Theorem 39.6.
DJ.l-2 = 0 a.e. and
DJ.l. = DJ.l-1
=
dj.l. a.e. d)..
-
Proof. Let J.l. be a Borel signed measure on JR.k . By considering separately the
positive and negative parts of J.l., we can assume without loss of generality that J.l. is a Borel measure. Clearly, ·J.l. is a-finite. Let J.l. = J.l. l + J.l-2 be the Lebesgue decomposition of J.1. (see Theorem 37.7), where J.l.l << >.. and J.l-2 ..L A. If B11 = {x E 1Rk : llx II < n}, then the set function defined by v,;(E) = J.l.l (E n 811) for each Borel set E is a finite Borel measure satisfying v11 << )... Thus, by Theorem 39:4, v11 is differentiable for almost all points of 811, and as a routine verification shows, D J.l.l (x) = D Vn (x) holds for almost all x E 811• Since 811 t 1Rk , it follows that J.l- 1 is differentiable almost everywhere.
Chapter 7: SPECIAL TOPICS IN INTEGRATION
372
On the other hand, JL2 ..L ).. implies the existence of a Borel set A with JL2(A) = A(Ac) = 0. But by Lemma 39.3, DJL2(x) = 0 holds for almost all points x of A, and hence (since A(Ac ) = 0), DJL2(x) = 0 for almost all points x ofJR.k . Using Theorem 39.4, we see that
djL DJL(x) = DJLt(X) + DJL2(x) = DJL1(x) = -(x) d).. holds for almost all x
E JR.k , and the proof is finished.
•
Having established the basic differentiation properties of measures, our attention is now turned to the relationship between measures and real valued functions defined on an interval. In the following discussion some of the deepest classical results of ordinary derivatives will be obtained. Recall that a function f: I -+ JR. (where I is an interval) that satisfies f(x) < f(y) whenever x < y is called an increasing function (�Tld if f(x) > f(y) whenever x < y, then f is called decreasing). An increasing or a decreasing function is referred to as a monotone function. Likewise, if x > y implies f (x) > f(y), then f is called a strictly increasing function (and iff(x) < f(y) whenever x > y, then f is called strictly decreasing). A strictly increasing or a strictly decreasing function is known as a strictly monotone function. Every monotone function f has the property that both limits rtx
f(x-) = lim f(t) =
r--..c
lim
f(t)
and
lim f(t) f(x+) = lim f(t) = r-x+ r.t.x
exist in JR. for each x. Moreover, the oscillation of f at any point x is given by
wf(x) = lf(x+) - f(x-)1 < oo. Therefore, the discontinuities of a monotone function are jump discontinuities, and as the next result shows, they are at-most countably many.
Theore� 39.7.
The set ofall discontinuities ofa monotonefunction is at-most
Proof. Let f: I
-+
countable.
JR. be monotone. Replacing f by - f (if necessary), we can assume that f is increasing. Since the interior of the interval can be written as a countable union of open finite intervals, it suffices to show that f has at-most countably many discontinuities in each finite open interval. To this end, let (a, b) be a finite open subinterval of I . If A is the set of all discontinuities of f in (a, b), then A = U: A,, where 1 A, = {x E (a , b): f(x+) - f(x-) > � }. Now, if a < XJ < · · · < Xk < b belong
Section 39: DIFFERENTIATION AND INTEGRATION
373
to some A,, then it is easy to see that
k
- <
n
k
L r= I .
[f(x;+) - f(x;-)] f(b) -f(a) <
< oo
holds, which implies that each A, is finite. Hence, A is at-most countable, and the proof is finished. • An immediate and very useful consequence of the preceding theorem is that a monotone function has a point of continuity in every open interval. In applications this observation is translated as follows: Iff is a monotone function, then for every point there exist sequences and such that x, -1, and with continuous at each and y,. Recall that if a function 1R � 1R is increasing and left continuous (i.e., satisfying limrtx for each , then a set function J.L 1 can be defined on < by the semiring S In Example 1 3 .6 we verified that 1 is indeed a measure. Clearly, every Borel set is J.Lrmeasurable; therefore, the outer measure j restricted to B is a Borel measure. For simplicity, the restriction of J.L j to B will be denoted by J.L 1 again, and it will be called the Borel measure induced by It is a routine matter to verify that On the other hand, for < holds for each open interval the relation
x
{x,} x, f: f(t) = f(x) = {[a, b): a b} J.L J.L f. f(b) - f(a+)
{y,} x, y, t x xl([a, ) b)) = f(b)- f(a). J.L
(a, b). f(b)- f(a) = J.LI ([a, b)) = J.LI ({a}) + J.LI((a, b)) f a f ({a}) = J.L
f
J.LI ((a,ab)) b,=
shows that is continuous at some E 1R if and only if J.L 1 0 holds. The first "differentiation" relation between and 1 is stated in the next impor tant theorem. Keep in mind that the open balls of 1R are precisely the finite open intervals.
Let f: 1R 1R be an incz·easing left continuous function, and Then the Boz·el measure J.L is difef rentiable at x0 ifand letonly.\:0ifbefa irsea/number. differentiable at x0• Moreover, in this case D!-LI(xo) = f'(xo). Proof. f x0 m = f'(x0). - f(xo) m + m - f(x)x -x0 lx -xol 8. (*)
Theorem 39.8.
--+
1
Assume first that is differentiable at E > 0, choose 8 > 0 such that E <
<
E
and let
whenever 0 <
Given
<
374
Chapter 7: SPECIAL TOPICS IN INTEGRATION
If an
open interval (a, b) satisfies x0 E (a, b) and b - a < 8, then it is easy to see from (*) that m - E < f;!a < m + E. This implies JL t ((a , b)) f(b) - f(a +) -m < E --::... -- - m = b-a A((a, b))
for each open interval (a, b) with x0 E (a, b) and b - a < 8. Hence, DJL 1(x0) = m holds. For the converse, assume that m = D JL1(x0) exists in 1R and let E > 0. Choose some 8 > 0 such that whenever x0 E (a, b) and b - a < 8, then JL1((a, b)) --..::... -- - m < E. b-a
Let b > xo satisfy b - xo < 8. Since f has at most countably many discontinu ities (Theorem 39.7). there exists a sequence {ar.} such that o,. < x0, lim a,. = x0 , b - an < 8, and f is continuous at each a,. Thus, JL !((an, b)) = f(b) - f(an+) = f(b) - f(an). and so f(b) - f(an) b - a,
-m < E
for each n . By the left continuity of f, it follows that f(b) - f(xo) <E ----- - m b - x0
for each b > xo with b - x0 < 8. This means that f is differentiable from the right at x0 and thus, it is right continuous at x0• Reason:
.
b.J.xu
hm[f(b) - f(xo)] =
. [
b.J.xu
hm
f(b) - f(xo) (b - xo) b - Xo ·
Now, by the symmetry of the situation, we have
]
= m 0 = 0. ·
f(a) - f(xo)
----- - m < E
a - xo
for all a < xo with xo - a < 8. Thus, .f'(xo) exists, and f'(xo) the theorem is now complete.
= m. The proof of •
By Theorem 39.7 we know that a monotone function has at-most countably many discontinuities. Now, we are ready to prove that a monotone function also has a derivative at almost every point.
Section 39: DIFFERENTIATION AND INTEGRATION
375
Every monetonefunction is differentiable almost everywhere. Proof. ff ff : f * f* (x) f(x-) f(t) f(t). f* (x) < f(x) x. f(x11* } f X11 x t f ; f x (x1 1 ) f(x f(x-) (x-) n) n * * f*(x).) x. w1.(x) w1(x) f(x+) f(x-) f f* f* . f 1. * J;(a) a. f a, * f* (a) = f(a). E - f(a) < m + E < !x - a! < m - E < j. (x)x -a x ! x -a! < (x11 } < + x, < !x11 -a! < 8, f Xn. - f(a) < f(x) - f(a) < f(x + ) - f(a) m -E < f*(x)x -a x -a x-a I. f(xn)- f(a) < m + E. n-oo a Theorem 39.9 (Lebesgue).
Let be a monotone function. Replacing by - (if necessary), we can assume that is increasing. Define a new function JR. --7- JR. by =
= lim tfx
= lim t-x+
Clearly, holds for all Moreover, is increasing and left continu ous. (Indeed, for each x, there exists a sequence and with such that = lim therefore, continuous at each = lim = = Also, a similar argument shows that = = holds for each In particular, this shows that and have precisely the same points of continuity. Now, let J.Lf. be the Borel measure induced by By Theorem 39.6, fJ. is differ entiable almost everywhere, and hence, by Theorem 39.8, is also differentiable almost everywhere. Assume that m = exists at some point Then is continuous at and so Given > 0, choose some 8 > 0 such that holds whenever 0
Now, fix some
Xn
with 0
0
8, and then choose a sequence such that and with continuous at each It follows that
-----
- tm
8.
,;,__ _ _ ___;_ _
x, -
Therefore,
f(x) - f(a) - m < E x-a < !x - a! < 8. f f'(a) * I ..;.___;,_,__
f'(a) J:(a) =m f. f
holds whenever 0 Hence, exists and is differentiable, so is holds. That is, at every point where differentiable almost everywhere.
= Therefore,
is
•
For the rest of the section, only real-valued functions defined on a (finite) closed interval will be considered. To simplify matters (when the circumstances
[a, b]
Chapter 7: SPECIAL TOPICS IN INTEGRATION
376
b] f(x) f(b) x > b. f(x) = f(a) x . . {t0, . , t1 1 } = b] to t11 b f: [a, b]
require it), a function defined on [a, will also be tacitly assumed defined on all of 1R by if < a and if = We start by reviewing some basic properties of functions of bounded variation. Recall that a collection of points P is a partition of an interval [a, if a = < t1 < < = holds. Let -4 1R be a function. Then the (total) variation V1 off over is defined to be · · ·
V1
= sup
{
t. l/(1;)- f(t; )I: P {to, tn} -1
=
[a, b]
a b] } .
is a partition of [ ,
. . . •
f
If V1 < oo holds, then is said to be a function of bounded variation. A function of bounded variation is necessarily a bounded function. [Reason: If < < b, then
a x (1/(x)l - lf(a)l) (1/(x)l - lf(b)l) < l f(x) - f(a)l + lf(b) - f(x)l < VI .] Also, it is a routine matter to verify that if f and are functions of bounded variation .(on [a , b]) and then f af, fg, and Ill are all functions of bounded variation. Thus, if B V [a, b] denotes the collection of all real-valued functions of bounded variation on [a , b], then the latter shows that B V [a, b] is a function space and an algebra of functions. Every monotone function f is of bounded variation, and V1 lf(b) - f(a)l holds. [Indeed, if f is increasing, and a = to < t1 _< < tn = b, then L lf(t;) - f(t;-J)I L [f(t;)- f(t;_J)] = f(b) - f(a).] +
g
+ g,
a E JR.,
=
· · ·
n
n
=
i=l
i=l
Therefore, every function that can be written as a difference of two monotone functions is of bounded variation. In actuality, these are the only types of functions that are of bounded variation. The following discussion will clarify the situation. -4 1R be a function of bounded variation. Then restricted Now, let to any closed subinterval [c, d] of [a, is of bounded variation there. As a matter of fact, if for any closed subinterval [c, d] of [ we denote the (total) variation of f over [c, d] by Var1(c, d), i.e.,
f: [a, b]
b]
f
a b] ,
Var1(c, d) = { II sup
8 1/(t;) - f(t;_J)I: P = {to, . . . , t,} is a partition of [c, d] } ,
Section 39: DIFFERENTIATION AND INTEGRATION
377
then it is easy to verify (and the reader should do so) that Varf(c, d) = Varf(c,
e) + Varf (e , d).
for all a < c < e < d < b. Therefore, we can define a new real function Vt(·) by Vf(a) = 0 and Vt(x) = the variation of f over [a, x] for a < x < b. Note that V1(b) = v1. The function V1(·) is called the variation function of f , and its basic properties are included in the next theorem.
For a function f: [a, b] � 1R of bounded variation, the following state_ments hold: l . The total variation function Vt(·) off is increasing. 2. l f(y) - f(x)l < Vt(Y ) - Vt(x) holds for all a < x < y < b. 3. The function g: [a, b] � 1R defined by g(x) = Vt(x) - f(x) is increasing. 4. The function f is continuous at some xo E [a, b] if and only if its total variation function V1 is continuous at x0• Proof. (1) Assume a < x < y < b. If a = to < t 1 < · · · < t11 = x < y Theorem 39.10.
holds, then
II
L l f(t;) - f(t; _, )l + l f(y) - f(x) l < Vf(y). i=l
Hence, Vt(x) < Vt(x) + l f(y) - f(x)l < Vt(Y) holds. (2) The preceding inequality immediately gives (2). (3) If a < x < y < b holds, then (2) implies
f(y) - f(x) < l f(y) - f(x)l < Vt(Y) - Vt(x), so that g(y) - g(x) = [Vt(Y ) - f(y)] - [Vf(x) - f(x)] > 0. (4) If V1 is continuous at xo, then it should be obvious from (2) that f is also continuous at x0. For the converse, assume that f is continuous at x0 E [a, b ]. We shall establish that v1 is right-continuous at xo, and we shall leave the identical arguments for the left continuity of Vf at xo for the reader. So, assume a < x0 < b and let E > 0. Choose some o > 0 such that x E [a, b] and l x - xol < o imply l f(x) - f(xo)l < E . Next, fix some xo < s < b such that s - x0 < o, and then select a partition P = {a = to < t1 < · · · < t, = s] of [a, s] such that
Vt(S) - E < L l f(t;) - f(t;- J)I . II
i=l
378
Chapter 7: SPECIAL TOPICS IN INTEGRATION
By adding the point x0 to the partition that
P, we can assume without loss of generality
P = {a = to < t1 < · · · < tj = xo < tj+ l < · · · < t, = s}. Now, assume that xo
< x < lj+ l · Then, we have
v,(x) + var,(x, s) = v,(s)
j
fl
< L lf(t; ) - f(t;- dl + E i=l
< L l f(t; ) - f(t;-dl + l f(x) - f(xo) l + l f(x) - f(tj+dl i= l n
+ L l f(t; ) - f(t;-dl + E i=j+2 < Vf (Xo) + if(x) - f(xo)l + Varf(X, s) + E < Vf(Xo) + Varf(X , s) + 2E. This implies 0 ,:5 Vf(X) - Vf(Xo) right-continuous at x0 , as desired.
< 2E
for all XQ
<
X
< lj+ l•
and
SO
Vf is •
An immediate application of the preceding theorem is the following:
Theorem 39.11. If f: [a, b] � R
is of bounded variation, then 1 .. f is the difference of two increasing functions (which, in addition, can be taken to be continuous iff is also continuous), and 2 . f is differentiable almost everywhere.
Proof. For ( 1 ) apply Theorem 39.10 to f(x) = Vf(x) - [V,(x) - f(x)], and
•
for (2) use Theorem 39.9.
Let f: [a, b] � R be an increasing function and consider it defined on all of R by f(x) = f(b) if x > b and f(x) = f(a) if x < a. Then the set function
J.l.f([c,d)) =
f(d-) - f(c-)
defines a measure on the semiring {[c, d): c, d E R and c < d}. Clearly, the measurable sets of J.l. 1 contain the Borel sets of R, and hence, J.L1 is a Borel measure. Also, J.l.f vanishes off [a, b]. That is, Supp Jl 1 c [a, b], and therefore, J.l. 1 is a finite Borel measure. Now, consider a function f: [a, b] � R of bounded variation. Again, f is considered defined on all of R by f(x) = f(b) if x > b and f(x) = f(a) if
Section 39: DIFFERENTIATION AND INTEGRATION
a. By Theorem 39. 1 1 , there exist two increasing functions g and h on [a, b] such that f = g - h on [a, b]. Then the set function
x <
379
defined
for each Borel subset E of 1R defines a finite Borel signed measure. It is easy to see (by Theorem 15. 10) that the value fJ-J(E ) does not depend upon the particular representation of f as a difference of two increasing functions. In addition, note that JJ-f vanishes off [a, b] and that
JJ-J( [c, d)) = f(d-) - f(c-)
MJ({x}) = f(x+) - f(x-).
and
The finite Borel signed measure JJ-1 is referred to as the Lebesgue-Stieltjes5 measure generated by the function of bounded variation f.
Which Lebesgue-Stieltjes measures are absolutely continuous with respect to the Lebesgue measure? The answer is provided by the next theorem. For a continuous function f: [a, b] � 1R ofbounded varia tion, the following statements are equivalent: 1 . fJ-f is absolutely continuous with respect to the Lebesgue measure. 2. For each E > 0, there exists some 8 > 0 such that if(a1, b 1 ), . . . , (an, b11) are disjoint open subintervals of [a, b] satisfying "L�'= 1 (b; - a; ) < 8, then "L�'= 1 l f(b;) - f(a;) l < E holds.
Theorem 39.12.
Proof. (1)
===>
(2) Assume (2) is false. Then there exists some E > 0 so that given any 8 there exist disjoint open subintervals (a 1 , b 1 ), , (a11, b11) of [a, b] such that •
i
=l
Note that the open set 0
and
L l f(b; ) - f(a; )l > E . i
=l
= U�'= 1 (a; , b;) satisfies A.(O) = "L�'= 1 (b; - a;) < 8 and 11 < L l f(b; ) - f(a; ) l = L IJJ-J ((a; , b;))l < L IMJI ((a; , b;)) = IMJI( O). i= l i=l i=l n
E
•
II
II
L(b; - a;) < 8
•
n
Thus, for each n there exists an open set 0, c (a, b) such that A.(011) < 2-u and IMJI(O,) > E. Let A = n�l u�ll 0; , and note that A is a Borel set. Since
5Thomas Jan Stieltjes ( 1 856-1894), a Dutch mathematician. He published extensively in all parts of analysis in his time. He is remembered today mainly for his generalization of the Riemann integral.
380
Chapter 7: SPECIAL TOPICS IN INTEGRATION
it follows that >..( A) = 0. On the other hand,
holds for each n, and so, IILJI(A) = limi/LJI( u�ll tJ;) > E, contrary to 1/LJI << A (recall that IILJI << >.. if and only if /Lf << .>..). Hence, (2) must be true. (2) ==> ( l ) LetA beaBorel setwith>.. ( A) = 0. Wecan supposethat A c (a, b). Let E > 0. Choose some fJ > 0 so that statement (2) is satisfied. Since /Lf is a finite regular Borel signed measure and >..(A) = 0 holds, there exists an open set tJ such that A c tJ c (a, b), 1/LJ(tJ) - /LJ(A)I < E, and .>..(0) < !J. Let tJ = U(a;, b;) be written as an (at tnost countable) union of disjoint open intervals. Then, it is easy to see that t"' VJ 111/(/ '1\1 -
� ' 'j(("" � /-A' ....,· , i
b·'' 1/1
-
-
"("·, )] -< L... �.., l f(J.. "/· )-f'n \-1·) l
J.. · )-J �.., [J.f(VI L... i
..
i
< e:
� �·
Hence, IJ.LJ(A)I < IlLJ(A ) - ILf(tJ) I + IlLf(tJ)I < 2E holds for each E > 0. That is, ILJ(A) = 0, so that IL1 << .>.. , and the proof of the theorem is complete. • Statement (2) of the preceding theorem is taken as the definition of absolute continuity of functions. Definition 39.13. A function f: [a, b] � 1R is said to be absolutely contin uous iffor every E > 0 there exists some fJ > 0 such that whenever (a 1 , b 1 ) , , •
•
•
(an, bn) are disjoint open subinten1als of [a, b], then 1/
1/
i=l
i=l
L(b; - a;) < fJ implies L l f(b; ) - f(a; ) I < E. It should be clear that an absolutely continuous function is necessarily contin uous. The converse is false; see Exercises 7 and 8 at the end of this section. Also, it is easy to verify that if f and g are absolutely continuous functions on [a, b] and a E lR, then f + g, af, fg, and If I are all absolutely continuous functions. Thus, the collection AC [a, b] of all absolutely continuous functions on [a, b] is a function space and an algebra of functions.
For an absolutely continuous function f: [a, b] � lR, the following statements hold: 1 . f is of bounded variation, and hence, AC [a, b] is a vector sublattice of B V [a, b]. Theorem 39.14.
Section 39: DIFFERENTIATION AND INTEGRATION 2.
381
The variationftmction V1(·) is absolutely continuous, and hence, difference of two increasing absolutely continuous functions.
f is the
Proof. (1) For any closed subinterval [c, d] of [a, b], we denote (as usual) by
Var1(c, d) the total variation off over [c, d]. Keep in mind that if c < e < d, then Var1(c, d) = Var1(c, e) + Var1(e, d).
Next, choose some 8 > 0 so that the definition of absolute continuity is satisfied with E = I . Now, fix some natural number n with h-n a < 8, and let a = to < t1 < b�a for each i . Clearly, · · · < tn = b be the partition of [a, b] with t; - t;-1 = Var1 (t; - t , t;) < I holds for each i, and thus, Vt = Varf(a, b) =
II
VartCti- t , t;) < n < oo. L = i l
, t11 } is a partition (2) If [c, d] is a closed subinterval of [a, b] and P = {t0, t1, of [c, d], then write v(c, d, P) = L:;'= t l f(t;) - f(t; 1)1. Note that •
•
•
-
Vt(d) - Vt(c) = Varf(c, d) = sup{v(c, d, P): P is a partition of [c, d]}.
Now, let E > 0. Choose some 8 > 0 for which the definition of absolute continuity of f is satisfied. Assume that (at, bt ), . . . , (an, bn) are pairwise disjoint open subintervals of [a, b] such that L�=1 (b; -a;) < 8. For each I < < n, let P; be an arbitrary partition of [a;, b;]. Each P; subdivides (a; , b;) into a finite number of open subintervals. Clearly, these open subintervals taken together are pairwise disjoint, and the sum of their lengths equals I:�= I (b; - a;) < o. The absolute continuity of f applied to the open subintervals yields
i
It follows that 11
L[Vf(b;) - V1(a;)] < E, i= l
so that V1(x) is absolutely continuous.
•
The final theorem of this section is a classical result. It asserts that for the Lebesgue integration, the fundamental theorem of calculus holds true precisely for the absolutely continuous functions. (As is customary, the Lebesgue measure on IR is denoted by dt .)
Chapter 7: SPECIAL TOPICS IN INTEGRATION
382
�heorem 39.15. A function only iff' E L.t ([a, b]) and
f: [a, b]
--+
f(x) - f(a) =
IR
is absolutely continuous zf and
[ f'(t) dt
holds for each x E [a, b]. Proof. Assume that f is absolutely continuous. By Theorem 39.14, we can suppose that f is increasing. Also, by Theorem 39.12, JL1 << ).. holds, and so, by d). holds for some E L 1([a, b]). the Radon-Nikodym theorem JLt(E) Now, combine Theorems 39.4 and 39.8 to obtain that f' a.e. To get the desired identity let E = [a, x). =
JE g
g
=
g
The converse is straightforward and is left as an exercise for the reader. See also Exercise 3 of Section 37. • ·
EXERCiSES 1.
If Jl is a Borel measure on
IRk , then show that Jl J.. ).. holds if and only if DJ.L(x) = 0
2.
Generalize Theorem
to Lebesgue measurable subsets of
for almost all
if
E
is
points.
3.
x.
39.5
JR.k . That is, show that
a Lebesgue measurable subset of IRk , then almost all points of E are density
Write Br(a) for the open ball with center
at a e JR.k and radius r . If I is a Lebesgue k integrable function on IR , then a point a e IRk is called a Lebesgue point for I if lim )..
\
{ ll(x) - l(a)l dJ..(x) = 0. (Br a )) JB,(a) Show that if 1 is a Lebesgue integrable function on JR.k , then almost all points of JR.k r-+0+
are
Lebesgue points.
[HINT: Let Q be the set of all rational numbers in JR. Apply Theorem 39.4 to conclude
that for each
a e Q there exists a null set Ea such that lim
holds for all
x
¢ Ea. Let E
Lebesgue point.] 4.
Let
\
r-+0+ A(Br X
))
al dJ..(t) ll(x) - al { JB,(x) ll(t) =
= UaeQ Ea, and then show that every point of Ec is a
I: JR. --+
JR. be an increasing, left continuous function. Show directly (i.e., with out using Theorem 38.4) that the Lebesgue-Stieltjes measure Jlf is a regular Borel
measure.
5.
(Fubini) Let
Un) be a sequence of increasing functions defined on [a, b] such that I:::1 ln(x) = l(x) converges in JR. for each x E [a, b]. Then show that I is differ entiable almost everywhere and that I' (x) = 2:�1 I� (x) holds for almost all x . [HINT: Replacing each In by In - ln(a), we can assume that In ::::_ 0 holds for each n. Let sn = It + · · · + 111 , and note that each sn is increasing and sn(x) t l(x) for
Section 39: DIFFERENTIATION AND INTEGRATION
6. 7.
383
each x. By Theorem 39.9, f and all the j;, and s11 are differentiable almost every where. Since sn+l - s11 = f"+ 1 (an increasing function), s:,+ 1 (x) ::: s:, (x) must hold for almost all x; similarly, /'(x) > s:,(x) for almost all x. Now, choose a subsequence {sk, } of {s"} such that :[:�1 [f(x) - Sk, (x)] :=: L� l [/(b) - Skn (b)] < oo. Observe that {f - Skn} is a sequence of increasing functions. Use the preceding arguments to obtain that sL � !' a.e., and then conclude that L�l J,; = f' a.e. holds.] Suppose { f, } is a sequence of increasing functions on [a, b] and that f is an increasing function on [a, b] such that J.1.. J,, t J.1.. 1 . Establish that f' (x) = lim 1,: (x) holds for almost all x. This exercise presents some basic properties of functions of bounded variation on an interval [a, b]. Iff is differentiable at every point and I!' (x) I � M < oo holds for all x E [a, b], then show that f is absolutely continuous (and hence, of bounded variation). b. Show that the function f: [0, 1 ] --+ .lR defined by /(0) = 0 and f(x) =x 2 cos(x -2) for 0 < x < 1 is differentiable at each x, but not of bounded variation (and hence, f is continuous but not absolutely continuous). c. If f is a function of bounded variation and 1/(x) l ::: M > 0 holds for each x E [a , b] , then show that j is a function of bounded variation. d. If a function f: [a, b] � .lR satisfies a Lipschitz condition (i.e., if there exists a constant M > 0 such that 1 /(x) - f(y)l < M ix - yl holds for all x, y E [a, b]), then show that f is absolutely continuous.
a.
8.
This exercise presents an example of a continuous increasing function (and hence, of bounded variation) that is not absolutely continuous. Consider the Cantor set C as constructed in Example 6.15. Recall that C was obtained from (0, 1] by removing certain open intervals by steps. In the first step we removed the open middle third interval. At the nth step there were 2"- 1 closed intervals, all of the same length, and we removed the open middle third interval from each one. of them. Let us denote by I;', . . . , I;,,_, (counted from left to right) the removed open intervals at the nth step. Now, define the function f: [0, 1] � [0, 1] as follows: i. ii. iii.
/(0) = 0; 1 if x E If' for some 1 :=: i :=: 2"- , then f(x) = (2i - 1)/2"; and if x E C with x ¥= 0, then f(x) = sup{f(t): t < x and t E [0, I ] \ C).
Part of the graph of f is shown in Figure 7.1. a. b. c. d. 9. 10.
Show that f is an increasing continuous function from [0, 1] to [0, 1 ] . Show that f'(x) = 0 for almost all x. Show that f is not absolutely continuous. Show that J.1.. 1 j_ A. holds.
Let f: [a, b] � .lR be an absolutely continuous function. Then show that f is a constant function if and only if f'(x) = 0 holds for almost all x. Let f and g be two left continuous functions (on .IR). Show that J.1..1 = J.l..g holds if and only if f - g is a constant function.
Chapter 7: SPECIAL TOPICS IN INTEGRATION
384
y
1
l r r
1.
4
r
..!.
-----e
2
i• �
..!.
4
i.
.1.. .1 .1.. 9 9 3
2 3
2 9
.! 1
X
9
FIGURE 7.1. 11. This exercise presents another characterization of the nann dual of C [a, b]. Start by
letting L denote the collection of all functions of bounded variation on [a, b] that are left continuous and vanish at a.
a. b.
c. d. 12.
13.
Show that L under the usual algebraic operations is a vector space, and that f H- /1-f• from L to Mb([a, b]), is linear, one-to-one, and onto. Define f � g to mean that f - g is an increasing function. (Note that f 2:. g does not imply f � g.) Show that L under � is a partially ordered vector space such that f � g holds in L if and only if /1-f 2:. 11-K in Mb([a. b]). Establish that L with the nann 11/11 = Vl fl is a Banach Janice. Show, with an appropriate interpretation, that C*[a, b] = L .
If f: [a, b] � IR is an increasing function, then show that f' E L;pa, b]) and that 1: f'(x)dx ::: f(b) - f(a) holds. Give an example for which f'(x) dx < f(b) - f(a) holds. [HINT: Let gn(X) = n[f(x + �)- f(x)] for each n and x E [a. b]. Note that gn � !' a.e. holds. Now, apply Fatou's lemma to the sequence {gn).] If f: [a, b] � IR is an absolutely continuous function, then show that
fa
Vt
= 1hl f'(x) l dx
holds. 14. For a continuously differentiable function f: [a, b] properties: a.
�
IR, establish the following
The signed measure 11- f is absolutely continuous with respect to the Lebesgue measure and d11-f fdA. = f' a.e.
Section 40: THE CHANGE OF VARIABLES FORMULA
385
b. If g: [a, b] -+ 1R is Riemann integrable. then gf' is also Riemann integrable and
JgdJkt = lhg(x)f'(x)dx.
15. For each n, consider the increasing continuous function j., (x) = If f: 1R -+ 1R is
16.
{ -:1) n(x
17.
1R -+ 1R defined by
if X > 0, if 1 - II! < \: < if X < - * .
+I
a continuous function, then show that
a. f is J1- f, -integrable for each n, and b. lim Jf dJk J,, = f(l). Let f: 1R -+ 1R be a bounded function and let E = {x E JR.:
/11:
1
•
1'
J\r) exists in JR.}.
If >..( £) = 0, then show that >..(f(E)) = 0. This exercise presents an example of a continuous function f: 1R -+ 1R which is nowhere differentiable. Consider the function ¢: [0, 2] -+ 1R defined by ¢ (."C) = ."C if 0 < x ,::: 1 and ¢(x) = 2 - x if < x .::: 2. Extend ¢ to all of 1R (periodically) so that ¢(x) = ¢(x + 2) holds for all x E JR. Now define the function f: 1R -+ 1R by
1
f(x) =
)1 1 ( 3 1L=0 00
4
¢(4nx).
Show that f is a continuous nowhere differentiable function. (Compare this conclusion with Exercise 28 of Section 9.) 40. THE CHANGE OF VARIABLES FORMULA The objective of this section is to establish the familiar formula known as the "change of variables formula." To do this, we need some preliminary discussion. For the rest of the section, V and W will be two fixed open sets of some Euclidean space JR.k . Also, 11·11 will always denote the Euclidean norm on JR.k. We remind the reader about the representation of a linear mapping on JR.k by a matrix. Let T : JR.k -+ JR.k be a linear mapping, and let {e 1 , , ek } be the standard k basis of JR. (that is, the ith coordinate of ej is 1 if i = j and 0 if i # j). For each j, there exist constants a1j, . . . , akj (uniquely determined) suoh that . L..Jki = J aije;. If we define the matrix T(ej ) = "\' •
A=
GJJ GJ2 a21 a22
G Jk a2k
•
•
386
Chapter 7: SPECIAL TOPICS IN INTEGRATION
then T(x) = Ax holds for each x e 1Rk. The matrix A is called the matrix representation of the linear mapping T (with respect to the standard basis). Conversely, any k x k matrix A = [aij] defines a linear operator T: 1Rk � 1Rk by the formula T(x) = Ax for 1 each x e 1Rk. Also, it is easy to see that if M = k ·max{ laij 1 : i, j = , . . . , k}, then II Ax II < M llx II holds so that each matrix defines a continuous linear operator. The determinant ofT is simply the determinant of the matrix A ; that is, det T = det A . A function T : V � 1Rk is said to be differentiable at some point0a of (the open set) V if there exists a linear operator A : 1Rk � 1Rk and some r > such that T(x) = T (a) + A(x - a) + o(x - a) holds for all x e V with ll x - a II V to 1Rk such that
< r. Here (as usual) o(x - a) is a function from
1.
un
x-+-a
o(x - a)
ll x - a ll
_ -
0
.
The linear operator A is denoted by T '(a) and is called the derivative ofT at the point a. It should be clear that if T is differentiable at a, then T must be continuous at a. A function T: V � 1Rk is said to be differentiable ifT'(x) exists at each point x e V. Moreover, if T = (T1 , T,) is differentiable, then it follows that each partial derivative •
8T·
1
ax;
•
•
,
h--+-0
(x) = lim
T · (x 1
+ he;) - T·(x) h 1
exists at every point x e V . In addition, the matrix representation of T'(x) is
determinant is called the This matrix is called the Jacobiall6 matrix, and Jacobian. The Jacobian determinant is denoted by lr(x); that is, lr(x) = det T'(x) = det[(8T;/8xj )(x)].
its
6Carl Gustav Jacob Jacobi ( 1 804-1851 }, a Gennan mathematician. He was a prolific writer who contributed to many diverse mathematical disciplines-including number theory, mathematical physics, mechanics, and the history of mathematics.
Section 40: THE CHANGE OF VARIABLES FORMULA
387
A function T: V --+ IRk is said to be C 1.-differentiable if T is differentiable and all of its partial derivatives are continuous functions on V . Observe that if T is C 1 -differentiable, then the Jacobian JT(·) is a continuous real valued function on V . It is important to know that C 1 -differentiable mappings carry null sets onto null sets. The details follow.
Lemma 40.1. Let V c IRk be open and let T: V --+ IRk be a C 1 -differentiable function. If A c V satisfies A.(A) = 0, then A.(T(A)) = 0 also holds. Proof. Assume first that there exists some M > 0 such that l(aTifaxj)(x)l < M holds for each x E V and i, j = 1 , . . . , k. Let C = kM + 1 . If a E V , then by the differentiability of T there exists some open ball Ba c V with center at a such that
II T(x) - T(a)ll < kMIIx - a ll + llx - a ll = C llx - all holds for each x E Ba. Let E > 0. Choose an open set 0 with A c 0 c V and A.(O) < E, and then select a sequence {K11} of compact sets such that 0 = U: 1 K11 • Since T is continuous, each T(K11) is compact, and the identity T(O) = U: 1 T(K11) shows that T ( 0) is a Borel set. Now, let K be an arbitrary compact subset of T(O). For each y E K fix some z E 0 with y = T(z). By the above, there exists an open ball By c 0 with center at z such that
II T(x) - Y ll < C ll x - zll holds for each x E By. Let B; be the open ball with center at y, and radius C times · , Yu E K that of By; clearly, T( By ) c B;. Since K is compact, there exist y 1 , such that K c U;'= 1 B;;. By Lemma 39.2, there exist pairwise disjoint balls Bj, . . . , B,�, among the B>�I , . . . , BY*II such that •
Note that the corresponding B 1 , balls. Therefore,
A(K) < A <
( Q B;)
�
•
•
•
, Bm
(3C)'
(3C)kA.(0) < (3C)kE
•
•
are necessarily pairwise disjoint open
t, }._(8
f) =
(3C)'A
(Q B ) 1
388
Chapter 7: SPECIAL TOPICS IN INTEGRATION
holds for each compact subset K ofT (0). Since T (0) is a Borel set, the regularity of ). implies A.(T(O)) < (3C)kE, and hence, A.(T(A)) < A.(T(O)) < (3C)kE for each E > 0. That is, A.(T(A)) = 0. For the general case, let B 1 , B2 , be an enumeration of the open balls with "rational" centers and rational radii whose closures lie entirely in V. Clearly, V = U: 1 Bn. Since each B, is compact and T is C 1 -differentiable, it follows easily that T : B n --+ IRk satisfies the hypotheses of the preceding case. Thus, A.(T(AnBn )) = 0 holds for each n, and consequently, we have A.(T(A)) = A.( U:1 T(A n Bn )) = 0, as required. • •
•
•
The notion of a diffeomorphism plays an important role for this section. Its definition follows.
Definition 40.2. Let V and W be two open sets ofiRk . Aftmction T: V --+ W is said to be a diffeomorphism if a. b. c. d.
T Is one-to-one and onto, T is C 1 -differentiable, lr(x) i= 0 holdsforall x E V, and T is a homeomorphism (from V onto W).
We remark that (a), (b), and (c) are enough to ensure that T is an open map ping and thus, a homeomorphism. However, property (d) has been added to the definition of a diffeomorphism in order to emphasize its importance. Also, it should be noted that the inverse mapping T - l : W --+ V is necessarily a diffeo morphism. For details about this and other properties of differentiable functions see [2]. The next theorem tells us that a diffeomorphism preserves the Borel and the Lebesgue measurable sets.
Let T : V --+ W be a diffeommplzism between two open sub sets of IRk , and let E be a subset of V. Then: a. T(E) is a Borel set if and only if E is a Borel set. b. T(£) is Lebesgue measurable if and only if E is Lebesgue measurable. Theorem 40.3.
Proof. (a) Let Bv and Bw denote the Borel sets of V and W, respectively.
Clearly, T(Bv) = {T(£): E E Bv } is a a-algebra of subsets of W containing the open sets of W; hence, Bw c T(Bv ). By the symmetry of the situation, B1, c T (Bw) holds, and from this· we get T (Bv) = Bw. The conclusion now follows easily from the last identity. (b) Let E be Lebesgue measurable. We can suppose A.(£) < oo (why?). Choose a Borel set B with E C B c V and A.(B \ £) = 0. By Lemma 40.1, ).(T (B \ £)) 0,
-l
=
Section 40: THE CHANGE OF VARIABLES FORMULA
so that T(B \ E) is Lebesgue measurable. set, and hence, the relation T(E)
=
389
�ow from (a) the set T(B) is a Borel
T(B) \ T(B \ E)
shows that T (E) is Lebesgue measurable. The converse should be obvious from the symmetry of the situation. • Now, assume that T : V ---+ W is a diffeomorphism. From the preceding theorem it is easy to see that the set function J.L(E)
=
A(T(E))
defined for each Lebesgue measurable subset E of V is a measure. On the other hand, it follows from Lemma 40.1 that J.L << A holds. The proof of the "change of variables formula" rests upon the fact that the Radon-Nikodym derivative dJ.L/dA satisfies (dJ.LfdA)(x) = llr(x)l for each x E V . To establish this identity, we need to know how a linear operator alters the volume of the unit cube of lR.k . Lemma 40.4.
matrix), then
If A : lR.k
---+
lR.k is a linear operator (which we identify with a
A(A(E))
=
I det A I · A( E)
holds for all Lebesgue measurable subsets E oflR.k . Proof. If A is not invertible, then A maps lR.k onto a linear. subspace oflower di mension, and hence, onto a set of measure zero (why?). Hence, A(A(E)) = ldet A I · A(E) = 0 holds trivially. Now, assume that A is invertible; clearly, A is a diffeomorphism. Let J.L(E) = A(A(E)) for each Borel set E . Then J.L is a Borel measure, and by Theorem 15.10 it is enough to establish J.L = ldet A I · A on the Borel sets. Start by observing that J.L is a translation invariant Borel measure on lR.k , and hence, by Lemma 18.7, there exists a constant C such that J.L = C A holds. In particular, if U = [0, 1 ] x · · · x [0, 1], then C = J.L(U). To complete the proof, we must show that C = ldet A I . To this end, note first that if A = A 1 o A2 and C , C 1 , and C2 are the constants associated with A, A t . and A2, respectively, then C
=
A(A(U))
=
A(A1 (Az(U)))
=
C J A(Az(U))
=
C1 C2A(U)
=
C 1 C2.
390
Chapter 7: SPECIAL TOPICS IN INTEGRATION
If we denote the matrix representing A by A again, then the matrix A can be written as a product of matrices A = A I A 2 An, where each A; is either · · ·
1. 2. 3.
a matrix obtained from the (k x k) identity matrix by multiplying one of its rows by a nonzero constant, a matrix obtained from the identity matrix by interchanging two of its rows, or a matrix obtained from the identity matrix by adding to one of its rows a nonzero multiple of some other row of the identity.
(The three types of matrices described previously are known as elementary ma trices.) Thus, since C = CIC2 C, [where C; = A.(A;(U))] and det A = det AI det A 2 det An, in order to complete the proof it suffices to verify the formula for the above three types of matrices. · · ·
·
· · ·
Type 1 .
Without loss of generality we can assume that the matrix A is obtained from the identity by mu!t!p!ying its first row by a nonzero number a . Note that det A = Ia I. Here U is mapped onto [0, a] x [0, 1 ] x x [0, I] if a > 0 and [a, 0] x [0, 1 ] x x [0, 1] if a < 0. In either case, C = A.(A(U)) = Ia I = ldet A I .
Type 2 .
matrix
· · ·
· · ·
In this case, det A = -1. It i s easy to see that U is mapped under A onto U . Hence, C = A.(A(U)) = A.(U) = 1 = ldet A I .
Type 3.
It is easy to see that this case is reduced to the matrix A obtained from the identity matrix by adding to its first row the second one. Clearly, det A = 1 . Now, observe that if x = (xi, x2, . . . , xk ), then Ax = (xi + x2, x2, X3, , xk). 2 In particular, when k = 2 the unit square in is mapped under A onto the parallelogram P as shown in Figure 7 .2. Obviously, A.(P) = 1 . On the other hand, if we consider the Lebesgue measure on JR.k (where k > 2) as the product measure of the Lebesgue measure on JR.2 and the Lebesgue measure on JR./.:-2 , then U is mapped under A onto the set P x Ut. where U1 is the (k - 2)-dimensional unit cube. By Theorem 26.2, A.(A(U)) = )�(P) A.(UI) = 1 = ldet AI, and the proof of • the lemma is complete.
JR.
·
0
XI
0
FIGURE 7.2. The Transformation of the Unit Square
.
•
.
391
Section 40: THE CHANGE OF VARIABLES FORMULA
Now, we are in the position to establish the identity
df.J./d'A = l iT I·
Theorem 40.5. Let T: V � W be a diffeomorphism between two open sub sets of lR.k . If J.J.(E) = 'A(T (E)) for each Borel set E of V, then J.1. is a Borel measure on V whose derivative satisfies
D J.J.(x) = lly(x) l for all x E V . Proof. Let a E V . Replacing T by S(x) = T(a + x) - T(a), we can assume without loss of generality that a = 0 and T (0) = 0. We shall prove first the result for the case when the Jacobian matrix equals the k
Jy(O) = 1 .
x
k identity matrix I, and so
Let E > 0. Fix 0 < a < i such that 1 - E < ( 1 - 2a )k < ( 1 and then select some o > 0 so that
II T(x) - x ll
<
+ 2ai
< 1 + E,
a llxll
holds for each x E V with llxll < o. (Since T(x) = T'(x)x + o (x) = x + o(x) such a o > 0 always exists.) Now, let B c V be an open ball containing zero with center at x0 and radius r < �. Let B 1 and B2 be the open balls both with center at xo and radii ( 1 - 2a)r and ( 1 + 2a )r, respective! y. We claim that
Indeed, note first that if x E B, then llx II < llx - xo II
IIT (x) - xoll
<
II T(x ) - x ll
+ llx - xo ll
<
+ llxo II
a llx ll
+r
<
<
2r
<
o, and so
(1 + 2a)r
for each x E B, which shows that T (B) c B2. Now, write
Bt = [Bt n T(B)] U [Bt \ T (B)] , and note that the two sets of the union are disjoint. Since T is a diffeomorphism, T(B) is an open set, and hence, B1 n T(B) is also open. Since
II T(xo ) - xoll
<
allxoll
<
2ar
<
( 1 - 2a)r,
it follows that T(x0) E B1 n T(B), and so B1 n T(B) # (/). Also, if llx - xoll = r holds, then llx II < 2r < o implies
r = llx - xoll
<
llx - T(x) ll
+ IIT(x) - xoll
<
2ar
+ II T(x) - xoll,
Chapter 7: SPECIAL TOPICS IN INTEGRATION
392
and so II T (x) - xo II > ( 1 - 2a )r. This shows that no boundary point of B is mapped under T into B 1 • In other words, B 1 \ T (B) = B 1 \ T (B) holds. Since T is continuous, T(B) is compact, and hence, B 1 \ T(B) is also an open set. But then, it follows that B 1 \ T(B) = (/) (see Exercise 1 at the end of this section), and so, B 1 = B 1 n T(B) c T(B) holds. Now, if c is the Lebesgue measure of the open ball with center at zero and radius one, then (1 - E)A.(B)
= c(l - E)rk < c( l
- 2a/rk = A.(B 1 ) <
J.L(B) < A.(B2) = c(l + 2a/rk < ( 1 + E)crk = ( 1 + E)A.(B).
=
A.(T(B))
Thus, 1 E < J.L(B)/A.(B) < 1 + E holds for all open balls B c V containing zero and having radius less than 8/2. 'That is, D J.L(O) = 1 holds. For the general case, let A be the Jacobian matrix at zero. Since det A = lr (O ) "# 0, the matrix A is invertible. Then S (x) = A -I (T (x)) defines a diffeomorphism between V and A -I (W) whose Jacobian matrix at zero equals I . By the preceding case, it follows that the Borel measure v ( E) = A.(S(E)) satisfies Dv(O) = 1 . Now, by Lemma 40.4, we have -
v(E) = ldet A -I I
·
A.(T(E)) = ldet Al-1 • J.L(E).
This implies DJ.L(O) = ldet AI = llr(O)I and the proof is finished. A particular case of the "change
•
of variables fonnula" is stated next.
Let T: V � W be a diffeomorphism between two open sub sets of JR.k such that A.(W) < oo. Then for each Lebesgue measurable subset E ofV we have Theorem 40.6.
A.(T (E)) =
L llr I dA..
A.(T(E)) for each Lebesgue measurable subset E of V. Then, by Lemma 40.1 , J.L << A. holds, and hence, by the Radon-Nikodym theorem J.L(E) = JE (dJ.L/dA.) dA. holds for each Lebesgue measurable set E. Now, com bine Theorem 39.4 and the preceding theorem to obtain dJ.LfdA. = D J.L = llrl Proof. Let J.L(E) =
393
Section 40: THE CHANGE OF VARIABLES FORMULA
Therefore, A.(T(E)) =
l llr I dA..
•
holds for each Lebesgue measurable subset of V. We now come to the main result of this section. E
Let T: V � W be a k diffeomorphism between two open sets of JR. . Then for every function f E L 1 (W), theftmctiorz (f T) · llr I belongs to L 1 (V) and
Theorem 40.7 (The Change of Variables Formula). o
fw f dA. = fv (f
holds.
o
T ) · llrl dA.
< oo.
Assume first that A.(W) Let f = X£, where E is a Lebesgue measurable subset of W. By Theorem we have Proof.
40.6,
1 [ f dA. = A.(£) = A.(T(T - (£))) = [
lw
=
fv (X£
o
T) · llrl dA. =
lr-1(£) llr I dA. (f o T ) · llr l dA..
fv
Thus, the conclusion is valid for characteristic functions of Lebesgue measurable subsets of W . It follows that the formula is true for step functions, and then by a simple continuity argument for each f L 1 (W). Now, if A.(W) = then let W, = {x E<W : llxll n} and Vn = r- 1 (W,). Clearly, A.(W,) for each n, and so, if f L 1 (W), then by the preceding case (and Levi's Theorem 22.8) [ f dA. = lim [ f dA. = lim 1 (f T ) · llr I dA. = 1 (f T) llr I dA. , lw lw, and the conclusion follows. The formula appearing in the preceding theorem is referred to as the change of variables formula and is commonly written as follows: E
oo,
< oo
<
0
,_,.oo v,
"-�'oo
o
E
v
o
·
•
{ lr(V)
f(y)dy =
1 f(T(x)) · llr(x)l dx. v
Chapter 7: SPECIAL TOPICS IN INTEGRATION
394
The following version of Theorem 40.7 is most often used in applications, and its proof follows immediately from the preceding theorem.
Let T: A B be a a mapping between nvo Lebesgue mea surable subsets of 1R.k. Assume that there exist two open sets V c A and W is a diffeomorphism, and 'A(A \ V) W c B such that T(V) W, T: V 'A(B \ W) 0. Thenfor each f E L 1 (B), the function (f T) · liT I (defined a.e. on A) belongs to L 1 (A) and Theorem 40.8.
-+
-+
=
=
=
o
f. f d). i(f =
holds.
0
T) . liT I d).
EXERCISES 1.
Show that an open ball in a Banach space is a connected set. That is, show that if B is an open ball in a Banach space such that B = 0 1 U 0 holds with both 01 and 0 2 2 open and disjoint, then either 01 = 0 or 0 = (/;. 2 [lllNT: If a E 01 and b E 02, then let a = inf{t E [0, I]: ta + (1 r)b E 01 }, and note that c = aa + (I a)b E B. To obtain a contradiction, show that c ¢. 0 1 and c ¢. 02.] 1 Let T: V -+ 1Rk be C -differentiable. Show that the mapping x � T' (x) from V into L(JRk, 1Rk) is a continuous function. Show that the Lebesgue measure on 1R2 is "rotation" invariant. (Polar Coordinates) Let E = {(r, 0) E JR2: r � 0 and 0 � 0 � 2rr}. The transfor mation T : E -+ 1R2 defined by T(r, 0) = (r cos O, r sinO), or as it is usually written -
-
2. 3.
4.
x = r cos 0
and
y = r sin 0,
is called the polar coordinate transformation on 1R2, shown graphically in Figure
7.3.
a. Show that /..(E \ E0) = 0. b. If A = {(x, 0): x � 0} , then show that A is a closed subset of 1R2 whose (two dimensional) Lebesgue measure is zero. c. Show that T: E0 -+ JR2 \ A is a diffeomorphism whose Jacobian determinant satisfies lT(r, 0) = r for each (r, 0) E E0• d. Show that ifG is a Lebesgue measurable subset of E with /..(G \ G0) = 0, then T(G) is a Lebesgue measurable subset of JR2. Moreover, show that if f E L1 (T(G)), then
[
lT holds.
f d/.. =
1·la[ f(r
cos 0, r sin O)r dr dO
Section 40: THE CHANGE OF VARIABLES FORMULA
(}
395
y
2rr
; . . 'E
.
:
(} .
'
.
.
. .. . . . . · cr-. e)· .. . . . :· .
.
T -------....
·
.
:
.. ..
;" � . .
r
.
. r
:· . . .
FIGURE
(x,y)
y
X
X
7.3. The Polar Coordinate Transformation
5. This exercise uses polar coordinates (introduced in the preceding exercise) to present an alternate proof of Euler's formula J000 e-x 2 dx = .J7i/2.
For each r > 0, let Cr = {(x, y) E JR2: x2 + y2 Sr = [0, r] X [0, r]. Show that Cr c Sr c c,..Ji· b. If f(x, y) = e-<x2+y2), then show that a.
1
C,
f dA. <
r
Js,
f dA. �
1
C, .fi
� r2,
.�.:
> 0, y > 0}
and
f dA.,
where >.. is the two-dimensional Lebesgue measure. c. Use the change of variables to polar coordinates and Fubini 's theorem to show that
L f dJ. f' { =
,-r' I dt dO
=
: (I - ,-'').
d. Use (b) to establish that
. and then let 6.
r
�
oo
to obtain the desired formula.
In JR4, "double" polar coordinates are defined by x = r cos(},
7.
y = r sin(},
z = p cos ¢,
w = p sin ¢.
State the change of variables formula for this transformation, and use it to show that the "volume" of the open ball in 1R4 with center at zero and radius a is �.rr2a4. (Cylindrical Coordinates) Let E = {(r, (}, z) E JR3: r :=:: 0, 0 � (} � 2rr, z E JR.}. The transformation T: E � 1R3 defined by T (r, (}, z) = (r cos 0, r sin(}, z) or as it is usually written X = r
COS(},
y
= r sin 0,
Z
= Z,
is called the cylindrical coordinate transformation, shown graphically in Figure 7.4.
Chapter 7: SPECIAL TOPICS IN INTEGRATION
396
z
z (r,
(x, y, z)
(), z)
l
2rr
()
FIGURE 7.4.
y
The Cylindrical Coordinate Transformation
a. Show that A.(£ \ £0) = 0. b. If A = {(x, 0, z) E 1R.3: x � 0, z E JR.}, then show that A is a closed subset of1R.3 whose (three-dimensional) Lebesgue measure is zero. c. Show that T: £0 -+ 1R.3 \ A is a diffeomorphism whose Jacobian detenninant satisfies Jr(r, () , z) = r for each (r, (), z) E £0• d. ShowthatifG isaLebesguemeasurable subsetof E with A.(G \ G0) = 0, then T(G) is a Lebesgue measurable subset of 1R.3. Moreover, show that if f E L 1 (T(G)), then
[
lr
f dA. =
holds. 8.
jj.laf
f(r cos(), r sin(), z) r dr d() dz
(Spherical Coordinates) Let
E = {(r, (), ¢) E 1R.3 : r � 0, 0 ,:5 () The transfonnation T : E
-+
,:5 2:rr,
0 ,:5 ¢
,:5 :rr } .
1R.3 defined by
T(r, (), ¢) = (r cos () sin ¢, r sin () sin ¢, r cos ¢), or as it is usually written x = r cos() sin¢,
y
= r sin() sin¢,
z = r cos ¢,
is called the spherical coordinate transformation, shown graphically in Figure
7.5.
a. Show that A.( £ \ £0) = 0. b. If A = { (x, 0, z): x � 0 and z E JR.}, then show that A is a closed subset of 1R.3 whose (three-dimensional) Lebesgue measure is zero. c. Show that T: £0 -+ 1R.3 \ A is a diffeomorphism whose Jacobian detenninant satisfies lr(r, (), ¢) = -r2 sin¢.
397
Section 40: THE CHANGE OF VARIABLES FORMULA
z c/>
(x, y, z) r
()
y
FIGURE 7.5. The Spherical Coordinate Transformation d.
Show that if G is a Lebesgue measurable subset of E with A.(G \ G0) = 0, then T(G) is a measurable subset of 1R.3. In addition, show that if f E L 1 (T(G)), then
{
}T(G)
holds.
f dA. =
!!Jc{
f(rcos () sin cp , r sin O sincp, r coscp)r2 sin cp dr dO d(jJ
BIB LIOGR APH Y I . C. D. Aliprantis and K. C. Border, Infinite Dimensional Analysis, Studies in Economic Theory, #4, Springer-Verlag, New York & Heidelberg, 1994. 2. T. M. Apostol, Mathematical Analysis, Second Edition, Addison-Wesley, Reading, MA, 1974. 3. G. Birkhoff and S. MacLane, A Survey ofModern Algebra, Third Edition, Macmillan, New York, 1965. 4. E. Borel, Le�ous sur Ia T!Leorie des Fonctions, Gauthier-Villars, Paris, 1898. (Third Edition, 1928.) 5. P. J. Daniell, A general form of integral, Annals of Matlzematics 19 (1917), 279-294. 6. M. M. Day, Normed Lineaz· Spaces, T hird Edition, Springer-Verlag, Heidelberg & New York, 1973. 7. 1. Dieudone, Foundations ofModern Analysis, Academic Press, New York & London, 1969. '8. A. A. Fraenkel, Abstract SetTlzeory, Fourth Edition, North-Holland, Amsterdam, 1976. 9. M. Frantz, On Sierpinski's nonmeasurable set, Fundamenta Mathematicae 139 (1991), 17-22. 10. B. R. Gelbaum and J. M. H. Olmsted, Theorems and Counterexamples in Mathematics, Springer-Verlag, Heidelberg & New York, 1990. 1 1 . C. Goffman, Real Functions, Prindle, Weber & Schmidt, New York, 1953. 12. C. Goffman and G. Pedrick, First Course in Functional A11alysis, Prentice-Hall, Englewood Cliffs, NJ, 1965. · 13. P. R. Halmos, Naive Set Theory, Springer-Verlag, Heidelberg & New York, 1974. 14. P. R. Halmos, Measure Theozy, Van Nostrand, New York, 1950. 15. E. Hewitt and K. Stromberg, Real and Abstract Analysis, Third Edition, Springer Verlag, Heidelberg & New York, 1975. ·16. T. Jech, Set Theory, Academic Press, New York, 1978. ·17. I. Kaplansky, Set Theozy and Metric Spaces, Allyn and Bacon, Boston, 1972. 18. 1. L. Kelley, General Topology, Graduate Texts in Mathematics, #27, Springer-Verlag, Heidelberg & New York, 1975. 19. A. N. Kolmogorov and S. V. Fomin, Measure, Lebesgue Integrals, and Hilbert Space, Academic Press, New York & London, 1961. ·20. K. Kuratowski and A. Mostowski, Set Theozy, North-Holland, Amsterdam, 1976. 21. H. Lebesgue, Integrale, longueur, aire, Annali Mat. Pura Appl., Ser. 3, 7 (1902), 23 1-359.
399
400
BIBLIOGRAPHY
22. J. R. Munkres, Topology: a Fh-st Course, Prentice-Hall, Englewood Cliffs, NJ, 1 975. 23. F. Riesz and B. Sz.-Nagy, Fuuctional Analysis, L. Boron translator, F. Ungar, New York, 1955. 24. H. L. Royden, Real Analysis, Third Edition, Macmillan, New York & London, 1 988. 25. W. Rudin, Principles of Mathematical Analysis, Third Edition, McGraw-Hill, New York, 1 976. 26. W. Rudin, Real and Complex Analysis, Third Edition, McGraw-Hill, New York, 1987. 27. W. Sierpinski, Sur un probleme concernant les ensembles mesurables superficiellement, Fundameuta Mathematicae 1 ( 1 920), 1 12-1 1 5 . 28. G. F. Simmons, lutroduction to Topology and Modern Analysis, McGraw-Hill, Macmil lan, New York, 1 963. 29. G. Strang, Linear Algebm and its Applications, Second Edition, Academic Press, New York, 1 980. 30. A. E. Taylor, General Theozy of Functions and Integration, Blaisdell, Waltham, MA, 1965. 3 1 . A. E. Taylor and D. C. Lay, Introduction to Functional Analysis, Second Edition, Wiley, l'le\V York, 1980. 32. A. Torchinsky, Real Variables, Addison Wesley, New York, 1988. 33. R. L. Wheeden and A. Zygmund, Measure and Integral, Marcel Dekker, Inc., New York & Basel, 1 977. 34. A. C. Zaanen, Integration, North-Holland, Amsterdam, 1967.
LIS T OF S YMBOLS (xi ); e / . 6 A1 , 6
Ac , 4
B(Q), 40 C(X), 66 Cb(X), 68 Fa -set, 6 1 Fa . G&, 61 G&-set, 61 Loo. 261
e2(Q}, 305 e p . 264 0. 2
L p (�J.), 255 A, Ap.. 104 IN, 6 n;eJ A; , 6 12 nie/ A ; , 3
�o .
Ui e/ A; , 3 XA. 126
C,
13
E, 2
A0 , 59
fJ,
*.
107 x-. 244
a A, 38 ±oo, 29 U, 162 a(:F), 150 d(x, A), 8 1 f: A -+ 8, 5 g 0 J, 5 Xn t X, 23
A, 59
401
I NDEX absolute value, 16 absolute value of vector, 67, 243 absolutely continuous function, 380 absolutely continuous measure, 338 absolutely summable series, 223 accumulation point, 37, 59 addition of real numbers, 14 additive function, 138 additivity of set functions (1'
98
finite, 99 adjoint of operator, 242 adjoint operator, 297
of continuity, 16 axiom of choice, 7 axiom of continuity, 16 axioms field, 14 order, 1 5 Baire space, 43
Baire 's category theorem, 44 Baire, Rene L. ( 1874-1932), 43 balanced set, 224 ball closed, 36
algebra of functions, 89
open, 35
algebra of sets, 95, 150 0'-, 97
unit, 220 Banach lattice, 246
algebraic difference of sets, 143 algebraic number, 13
Banach space, 218
almost everywhere relation, 120
Banach, Stefan ( 1892-1945), 217, 218,
reflexive, 239
alternating series, 33 angle between vectors, 282 antisymmetry, 7, 66 Archimedean property, 17 Archimedes (287-2 12 BC), 17, 180
229, 232, 237 Banach-Mazur limit, 241 Banach-Steinhaus theorem, 229 base, 65 for topology, 65
area function, 186, 187 Arzela, Cesare ( 1 846-1912), 75, 188
basis
Ascoli, Guido ( 1887-1957), 75 Ascoli-Arzela theorem, 75
Bernoulli's inequality, 14 Bernoulli, Jacob (1 654-1705), 14
associative laws, 15
Bernstein polynomials, 254
atom, 120
Bernstein, Felix ( 1 878-1956), 12
averaging operator, 253 axiom
Bernstein, Sergei N. ( 1880-1968), 254
completeness. 16
Bessel, Friedrich W. (1784-1846), 283
orthonormal, 298
Bessel's inequality, 284
403
404
INDEX
binary relation, 7 Balzano, Bernard (178 1-1848), 25, 29 Borel measurable function, 152 Borel measure, 136 induced by a monotone function, 373 Borel sets, 97 Borel signed measure, 362 Borel, Emile (187 1-1958), 5 1 bound essential, 260 lower, 16 upper, 16 boundary of set, 38, 59 boundary point, 38, 59 bounded from above set, 16 bounded from below set, 1 6 bounded metric space, 39 bounded operator, 225 bounded set, 1 6 C 1 -differentiable function, 387 C00-function, 203 cancellation law of addition, 1 9 of division, 20 of multiplication, 20 Cantor set, 41, 140 €-, 141 Cantor, Georg (1 845-1918), 12, 4 1 Caratheodory extension of measure, I l l Caratheodory function, 156 Caratheodory, Constantin ( 1873-1950), 103, 107' 1 1 1 ' 156 cardinal infinite, 13 cardinal number, 1 2 cardinality of the continuum, 1 3 Cartesian product, 6 Cartesius ( 1596-1650), 6 Cauchy sequence, 26, 39 in measure, 149 uniform, 71 Cauchy, Augustin L. (1789-1875), 26, 39, 278
Cauchy-Schwarz inequality, 278 Cavalieri's principle, 214 Cavalieri, Bonaventura (1589-1647), 214 cell, 186 chain, 8 chain rule, 350 change of variable, 176 change of variables formula, 393 characteristic function, 126 choice function, 6 circled set, 224 clopen set, 65 closed ball, 36 closed graph theorem, 234 closed mapping, 52 closed set, 35, 58 closed unit ball, 220 closure of set, 36, 59 closure point, 36, 59 cluster poipt, 23 co-meager set, 48 commutative group, 362 commutative laws, 1 5 compact metric space, 49 compact operator, 226 compact set, 49, 62 compact topological space, 62 complement orthogonal, 287, 291 complement of set, 4 complemented subspace, 235 complete metric space, 39 complete orthogonal set, 293 complete orthonormal set, 298 completeness axiom, 1 6 completion of a metric space, 46 complex inner product space, 276 complexification of a real vector space, 277 component of a point, 48 composition of functions, 5 condensation of singularities principle, 230 conjugate exponents, 264
INDEX
connected metric space, 47 connected set, 47 connected topological space, 65 constant sequence, 23 continuous function, 38, 60 at a point, 60 continuum, 13 continuum hypothesis, 13 generalized, 1 3 contraction, 55 convergence pointwise, 68 uniform, 69 convergence in measure, 146 convergent sequence, 22, 37, 60 in mean, 202 in measure, 146 weakly, 365 convergent series in a Banach space, 223 convex set, 224 countable set, 9 counting measure, 99 cover, 48 measurable of a set, 1 17 of set, 48 open, 48 pointwise finite, 56 cylindrical coordinate transformation, 395 cylindrical coordinates, 395 Daniell, P. J. (1889-1946), 166 Darboux, Jean-Gastin ( 1842-19 17), 181 De Morgan's laws, 4 De Morgan, Augustin ( 1806-1871 ), 4 decomposition Hahn, 331 Jordan, 333 Lebesgue, 342 decreasing function, 372 decreasing sequence, 22 decreasing sequence of functions, 70 dense subset, 37, 59 density function, 344
405
d�nsity point, 371 derivative, 386 lower of measure, 367 of measure, 367 Radon-Nikodym, 344 upper of measure, 367 derived set, 37, 59 Descartes, Rene (1596-1650), 6 determinant, 386 Jacobian, 386 diameter of a set, 39 diffeomorphism, 388 difference algebraic of sets, 143 set, 2 symmetric, 3 difference of two numbers, 20 differentiable function, 386 differentiable measure, 367 differential operator, 225, 227 Dini 's theorem, 70 Dini, Ulisse (1845-1918), 70 Dirac's measure, 99 Dirac, Paul A. M. (1902-1984), 99 Dirichlet kernel, 3 1 6 Dirichlet, Johann P. G. L. (1805-1859), 71, 316 discrete distance, 34 discrete metric space, 34 discrete topology, 58 disjoint sets, 2 distance, 34 discrete, 34 Euclidean, 34 , uniform, 70 distance function, 8 1 distance of a point from a set, 8 1 distributive law, 15 distributive laws, 3, 4 division, 20 division law for fractions, 20 domain of function, 5 double series, 30
406 dual order, 244
INDEX finer partition, 178 finite additivity, 99
Dynkin system, 152
finite intersection property, 55
Dynkin, Eugene B. (1924- ), 152
finite measure space, 1 1 3 finite set, 9
E -Cantor set, 141
finite signed measure, 332
Egorov, Dimitry F. (1 869-1931}, 124
finite subset property, 272
Egorov's theorem, 125 element
finitely additive measure, 102 first category set, 43, 59
first, 10
first element, 10
least, 10
Fischer, Ernst G. (1875-1954}, 258 Fisher, Ernst S. (1 875-1954), 258
element of a set, 2
fixed point, 55
elementary matrix, 390
Fourier coefficients, 299, 3 1 0
empty set, 2
Fourier coefficients of a function, 309
enumeration, 9
Fourier series of a function, 309
equality of sets, 2
Fourier, Jean B. J. (1768-1830), 299
equality of two functions, 5
Fresnel integrals, 199, 215
equicontinuous set, 75
Fresnel, Augustin J. ( 1788-1 827}, 199
equivalence class, 7
Fubini's theorem, 2 1 2
equivalence relation, 7
Fubini, Guido (1 879- 1943), 212, 382
equivalent distances, 39
function, 5 C 1 -differentiable, 387
maximal, 8
equivalent norms, 220 equivalent sets, 9
C00-, 203
essential bound, 260
absolutely continuous, 380
essential supremum, 260
additive, 138
essentially bounded function, 260
area, 186, 187
Euclid (ca 365-300 BC}, 34
Borel measurable, 152
Euclidean distance, 34
Caratheodory, 156
Euclidean norm, 218
characteristic, 126
Eudoxus of Cnidus (ca 400-347 BC), 17
choice, 6
even polynomial, 90
continuous, 38, 60
eventually true property, 22
continuous at a point, 60 contraction, 55
Fa -set, 6 1
decreasing, 372
family
density, 344
pairwise disjoint, 4
differentiable, 386
family of generators for u -algebra, 1 5 1
distance, 8 1
family of sets, 3
essentially bounded, 260
Fatou, Pierre J. L. (1 878-1929}, 171
Gamma, 200
Fejer kernel, 317
increasing, 372
Fejer, Leopold (1880-1959}, 317, 3 1 8
indicator, 126
field axioms, 14
injective, 5
INDEX
407
integrable, 166
greatest lower bound, 1 6
integrable over a set, 173 inverse, 5
group, 362 commutative, 362
isometry, 46
topological, 362
jointly measurable, 1 5 5 Lebesgue integrable, 166 measurable, 121, 152 measurable in a variable, 155 monotone, 372
Holder's inequality, 256
Holder, Otto Ludwig ( 1 859-1937), 256
Haar, Alfred ( 1885-1933), 362 Hahn decomposition, 331
one-to-one, 5 onto, 5
Hahn-Banach theorem, 237
periodic, 307
harmonic series, 33
Riemann integrable, 180, 187 separately measurable, 155 set, 98 simple, 127 square summable, 305 step, 127 strictly decreasing, 372 strictly increasing, 372 strictly monotone, 372 surjective, 5 uniformly continuous, 44
Hahn, Hans ( 1 879-1934), 217, 237, 33 1 Hausdorff, Felix ( 1 868-1 942), 60
Heine, Heinrich E. ( 1821-1881 ), 5 1 Heine-Borel theorem, 5 1 Hermite polynomial, 304 Hermite, Charles ( 1822-190 I), 304 Hilbert cube, 296 Hilbert space, 288 Hilbert, David (1 862-1943), 275 homeomorphic metric spaces, 39 homeomorphic topological spaces, 63 homeomorphism, 63
upper, 162 variation, 377 weight, 289 function measurable in a variable, 155
ideal, 247 identity
function of bounded variation, 376
Parseval, 298 image of a point, 5
function space, 68
image of set,
function with compact support, 85, 201 functional
improper Riemann integral, 190 increasing function, 372 increasing sequence, 22
left-invariant, 363 limit, 241
5
of functions, 70
right-invariant, 363
indefinite integral, 342 index set of family, 3
Gamma function, 200
indicator function, 126
generalized continuum hypothesis, 13 generating sequence for upper function,
indiscrete topology, 58 induced measure, 106
162 Gram, Jorgen P. ( 1850-1916), 284 Gram-Schmidt orthogonalization process, 284 graph of operator, 234
induced topology, 58 inequality Bernoulli's, 14 Bessel, 284 Cauchy-Schwarz, 278
408 (cont.) inequality Holder's, 256 Minkowski, 256 triangle, 16, 34, 2 1 8 infimum, 1 6 infinite cardinal, 1 3 infinite set, 9 infinity minus, 29 plus, 29 initial values, 233 injective function, 5 inner product, 276 inner product preserving operator, 304 inner product space complex, 276 inner, 276 input, 5 integrable function, 166 over a set, 173 Riemann, 180 integral, 127, 163, 166 Haar, 363 improper Riemann, 190 indefinite, 342 iterated, 2 1 1 Lebesgue, 1 27, 132, 163, 166 lower Riemann, 179, 186 of step function, 127 Riemann, 180, 187 upper Riemann, 179, 187 integral operator, 226 integral test, 33 interior of set, 35, 59 interior point, 35, 59 intersection, 2 intersection of sets, 3 interval, 134 inverse function, 5 inverse image of set, 5 inverse of real number, 20 isolated point, 56 isometric metric spaces, 46
INDEX isometry, 46 lattice, 249 linear, 239 isomorphic normed vector lattices, 249 iterated integral, 2 1 1 Jacobi, Carl G. J. ( 1 804-185 1), 386 Jacobian, 386 Jacobian matrix, 386 jointly measurable function, 155 Jordan dec�mposition of a signed measure,
333 Jordan, Camille (1838-1921), 333 kernel Dirichlet, 3 1 6 Fejer, 3 17
kernel of functional, 240 kernel of operator, 226 Kolmogorov, Andrey N. ( 1 903-1987), 269 Korovkin, Pavel P. ( 1 9 1 3-1985), 250, 254 Kronecker, Leopold ( 1 823-1891 ), 33, 298
L p-norm, 255 L p-space, 255 L'Hopital, Guillaume F. A. (1661-1704), 203 Laguerre functions, 30 I Laguerre, Nicolas E. ( 1834-1886), 301 Laplace transform, 250 Laplace, P. S. M. ( 1 749-1827), 250 lattice homomorphism, 253 lattice isometry, 249 lattice norm, 246 law distributive, 1 5 parallelogram, 279 laws associative, 15 commutative, 1 5 least element, 1 0 least upper bound, 1 6 Lebesgue decomposition, 342
INDEX
Lebesgue integrable function, 166 Lebesgue integral, 127, 132, 163, 166 Lebesgue integral of step function, 127 Lebesgue measure, 10 I. 1 1 3, 133, 134 Lebesgue number for open cover, 50 Lebesgue point, 382 Lebesgue, Henri L. ( 1875-1941 ), 50, 1 6 1 , 172, 1 84, 202, 341 , 375
Lebesgue-Stieltjes measure, 379 left Haar measure, 363 left translate of function, 362 left-Haar integral, 363 left-invariant functional, 363 Legendre polynomials, 295 Legendre, Adrien-Marie ( 1752-1833), 295
Leibniz, Gottfried W. ( 1646-1 7 1 6), 185 Levi, Beppo ( 1 875-1961 ), 170 limit Banach-Mazur, 241 pointwise, 68 limit functional, 241 limit inferior, 24 limit of sequence, 22 limit point of sequence, 23 limit superior, 24 LindelOf, Ernst L. ( 1870-1940), 48 linear functional, 235 order bounded, 244 positive, 244 linear isometry, 239 linear operator, 224 linearity of the integral, 127 Lipschitz condition, 145, 383 Lipschitz, Rudolf ( 1 832-1903), 1 45 little L p -spaces, 264 locally compact topological space, 84 locally finite measure, 272 lower bound, 1 6 greatest, 1 6 lower derivative of measure, 367 lower Riemann integral, 1 79, 186 lower sum, 178
rp apping closed, 52 sublinear, 236 mathematical induction, I 0 matrix elementary, 390 Jacobian, 386 positive definite, 287 symmetric, 287 matrix representation of operator, 386 maximal element, 8 Mazur, Stanislaw ( 1 905-1981 ), 241 meager set, 43, 59 measurable cover of set, 1 17 measurable function, 1 2 1 , 152 measurable set, I 03, 1 12, 1 5 1 measurable space, 1 5 1 measure, 98 absolutely continuous, 338 Borel, 136 counting, 99 differentiable, 367 Dirac, 99 finitely additive, I 02 Haar, 363 induced, I 06 Lebesgue, 1 0 1 , 1 13, 133, 134 Lebesgue-Stieltjes, 379 locally finite, 272 orthogonal, 339 outer, 103 outer generated by a measure, I 07 product, 204 regular Borel, 136, 352 representing a linear functional, 352 signed, 329 singular, 339 translation invariant, 137 measure space, 99 cr -finite, 1 1 3 finite, 1 1 3 nonatomic, 120 with the finite subset property, 272
410
member of a set, 2 membership symbol, 2 mesh of partition, 178 menic, 34 induced by a norm, 218 uniform, 70 menic space, 34 bounded, 39 compact, 49 complete, 39 connected, 47 discrete, 34 separable, 54 totally bounded, 53 menic spaces homeomorphic, 39 Minkowski inequality, 256 Minkowski, Hermann (1864--1909), 256 minus infinity, 29 monotone class, 158 monotone function, 372 monotone sequence, 22 monotone sequence of functions, 70 monotonicity of measure, 99 monotonici ty of the integral, 129 multiplication law for fractions, 20 multiplication of real numbers, 14 multiplication rule of signs, 20 natural embedding of normed space, 239 negative number, 15 negative of real number, 19 negative part of vector, 67, 243 negative set, 330 negative variation of a signed measure, 332 neighborhood of point, 59 Newton, Isaac (1642-1727), 185 Nikodym, Otton M. (1889-1974), 342 nonatomic measure space, 120 norm, 217 Lp. 255 e2, 305
INDEX
a-order continuous, 337 Euclidean, 218 in an inner product space, 278 lattice, 246 of operator, 225 sup, 69, 219 uniform, 69 norm bounded set, 218 norm dual of Banach space, 238 norm induced by an inner product, 278 norm preserving operator, 304 normal topological space, 80 normed vector lattice, 246 with a-order continuous norm, 337 normed vector space, 218 nowhere dense set, 41, 59 null set, 104 number algebraic, 13 cardinal, 12 Lebesgue, 50 negative, 15 positive, 15 odd polynomial, 90 one-to-one function, 5 onto function, 5 open mapping, 52 open ball, 35, 366 open cover, 48 open mapping, 52 open mapping theorem, 232 open set, 35, 58 open unit ball, 220 operator, 224 adjoint, 297 averaging, 253 bounded, 225 compact, 226 differential, 225, 227 inner product preserving, 304 integral, 226 Laplace, 250
INDEX linear, 224 norm preserving, 304 positive, 249 projection, 235 rank-one, 242 symmetric, 287 unbounded, 225 operator norm, 225 order partial, 7 order axioms, 1 5 order bounded linear functional, 244 order bounded set, 244 order complete vector lattice, 252 order continuity of the integral, 129, 164 order dual, 244 ordered pair, 6 ordered vector space, 66, 242 ordinate set of function, 2 1 5 orthogonal complement, 287, 291 orthogonal measures, 339 orthogonal set, 282 complete, 293 orthogonal vectors, 282 orthonormal basis, 298 orthonormal set, 282, 298 complete, 298 oscillation, 372 oscillation of function, 61 outer measure; 103 induced by a positive functional, 353 outer measure generated by a measure, 107 output, 5 p-adic representation of number, 32 pairwise disjoint family, 4 parallelogram law, 279 Parseval's identity, 298 Parseval, Marc-Antoine C. (1755-1836), 298 partial derivative, 193 partial order, 7 partially ordered set, 8
partially ordered vector space, 242 partition, 186 finer, 178 partition of interval, 178, 376 partition of set, 7 partition of unity, 86 perfectly normal space, 87 period of function, 307 periodic function, 307 permutation of lN, 30 plus infinity, 29 point accumulation, 37, 59 boundary, 38, 59 closure, 36, 59 cluster, 23 density, 371 fixed, 55 interior, 35, 59 isolated, 56 Lebesgue, 382 limit, 23 point of a set, 2 pointwise bounded set, 228 pointwise convergence, 68 pointwise finite cover, 56 pointwise limit, 68 polar coordinate transformation, 394 polar coordinates, 394 polynomial even, 90 Hermite, 304 odd, 90 positive cone, 242 positive definite matrix, 287 positive linear functional, 244 positive number, 15 positive operator, 249 positive part of vector, 67, 243 positive set, 330 positive variation of a signed measure, 332 positive vector, 67, 242 power set, 4
411
INDEX
412 principle
Riemann integral, 1 80, 187
condensation of singularities, 230
improper, 190
mathematical induction, 10
lower, 179, 186
uniform boundedness, 229
upper, 179, 187
well ordering, 1 0 product Cartesian, 6 . inner, 276
Riemann sum, 1 8 1 Riemann's criterion, 180 Riemann, Bernhard G. F. ( 1826-1 866), 177, 180, 202
product a -algebra, 154
Riesz representation theorem, 355
product measure, 204
Riesz space, 242
product semiring, 154, 204
Riesz, Frigyes (l 880-1956), 242, 244, 258, 292, 344, 355
projection, 235 proper subset, 2
right translate of function, 362
property
right-invariant functional, 363
Archimedean, 1 7 eventually L-uc, 22
ring cr ,
150
finite intersection, 55
ring of sets, 96, 149
topological, 55
root of real number, 19
Pythagoras of Samos (ca 580-500 BC), 282 Pythagorean theorem, 282
a -additivity, 98 a -algebra product, 154 a -algebra generated by a collection, 97
quotient topology, 65
a -algebra generated by a family, 150 a-algebra of sets, 97
Radon, Johann ( l 887-1956), 342
a-compact set, 357
Radon-Nikodym derivative, 344.
a -finite measure space, 1 1 3
Radon-Nikodym theorem, 342
a -finite set, 1 1 6, 175
range of function, 5
a-finite signed measure, 332
rank-one operator, 242
a-ring, 150
rearrangement invariant series, 30
a-set, 94
reciprocal of real number, 20
Schmidt, Erhard (1 876-1959), 284
rectangle, 204
Schroder, Ernst ( 1841-1 902), 12
reflexive Banach space, 239
Schroder-Bernstein theorem, 12
reflexivity, 7, 66
Schwarz, Hermann A. ( 1843-1921 ), 256,
regular Borel measure, 136, 352
278
regular Borel signed measure, 362
second dual of Banach space, 239
relation, 7
section of set, 154
a.e., 120
segment, 9
binary, 7
selection of points, 1 8 1
equivalence, 7
semiring
relative topology, 58
product, 154, 204
representing measure, 352
restriction, 97
restriction semiring, 97
semiring of sets, 94
Riemann integrable function, 180, 187
separable metric space, 54
INDEX separately measurable function, 155 separation by a continuous function, 80 by open sets, 80 separation of points by a family, 87 sequence, 6 Cauchy, 26, 39 constant, 23 convergent, 22, 37, 60 convergent in mean, 202 convergent to ±oo, 29 decreasing, 22 increasing, 22 monotone, 22 of averages, 27 sequence convergent, 22, 37 to ±oo, 29 sequence of averages, 27 sequence of functions decreasing, 70 increasing, 70 monotone, 70 series absolutely summable, 223 alternating, 33 convergent in Banach space, 223 double, 30 Fourier, 309 harmonic, 33 rearrangement invariant, 30 unconditionally convergent, 33 uniformly convergent, 71 set, 2 Fa . 61 Go. 6 1 €-Cantor, 141 CT, 94 cr-compact, 357 cr-finite, 1 16, 175 atom, 120 balanced, 224 Borel, 97 bounded, 16 bounded from above, 16
. bounded from below, 1 6 Cantor, 41, 140 circled, 224 clopen, 65 closed, 35, 58 co-meager, 48 compact, 49, 62 connected, 47 convex, 224 countable, 9 dense, 37, 59 derived, 37, 59 empty, 2 equality, 2 equicontinuous, 75 finite, 9 index, 3 infinite, 9 meager, 43, 59 measurable, 103, 1 12, 151 negative, 330 norm bounded, 218 nowhere dense, 41, 59 null, 104 of first category, 43 open, 35, 58 order bounded, 244 ordinate, 2 1 5 orthogonal, 282 orthonormal, 282, 298 partially ordered, 8 pointwise bounded, 228 positive, 330 power, 4 symmetric, 224 totally ordered, 8 uncountable, 9 void, 2 set difference, 2 set function, 98 cr-subadditive, 103 set inclusion, 2 set intersection, 2, 3 set of first category, 59
413
414
INDEX
set union, 2, 3 sets equivalent, 9 signed measure, 329 cr-finite, 332 Borel, 362 finite, 332 regular Borel, 362 simple function, 127 :E, 158
singleton, 2 singular measures, 339 space Loo(J.L), 261 Baire, 43 Banach, 218 function, 68 Hausdorff, 60 Hilbert, 288 measurable, 1 5 1 measure, 99 metric, 34 Riesz, 242 topological, 58 spherical coordinate transformation, 396 spherical coordinates, 396 square summable function, 305 standard representation of simple function, 127
Steinhaus, Hugo (1887-1972), 144, 229 step function, 127 Stieltjes, Thomas J. (1856-1894), 379 Stone, Marshall H. (1903-1989), 88, 89 Stone-Weierstrass theorem, 88, 89 strictly decreasing function, 372 strictly increasing function, 372 strictly monotone function, 372 subcover of set, 48 sublinear mapping, 236 subsequence, 6 subset proper, 2 subset of set, 2
subtraction, 20 sum Riemann, 1 8 1 sum of arbitrary numbers, 283 sup norm, 69, 219 support of function, 85, 201 support of measure, 263 supremum, 16 essential, 260 surjective function, 5 symmetric difference, 3 symmetric matrix, 287 symmetric operator, 287 symmetric set, 224 symmetry, 7 term sequence, 6 Tietze's extension theorem, 84 Tietze, Heinrich F. F. (1880-1964), 84 Toeplitz, Otto (1881-1940), 33 Tonelli, Leonida (1 885-1946), 213 topological group, 362 topological property, 55 topological space, 58 compact, 62 connected, 65 locally compact, 84 normal, 80 perfectly normal, 87 topology, 57 discrete, 58 indiscrete, 58 induced, 58 quotient, 65 relative, 58 total variation of a function, 376 total variation of a signed measure, 332 totally bounded metric space, 53 totally ordered set, 8 transformation cylindrical coordinate, 395 polar coordinate, 394 spherical coordinate, 396 transitivity, 7, 66
INDEX translate of function, 362 translation invariant measure, 137 translation of a set, 137 triangle inequality, 16, 34, 2 1 8 trigonometric polynomial, 308 unbounded operator, 225 unconditionally convergent series, 33 uncountable set, 9 unifonn boundedness principle, 229 unifonn convergence, 69 of series, 7 1 unifonn distance, 70 unifonn metric, 70 unifonn nonn, 69 unifonnly Cauchy sequence, 7 1 unifonnly continuous function, 44 union, 2 union of sets, 3 upper bound, 8, 16 least, 16 upper derivative of measure, 367 upper function, 1 62 upper Riemann integral, 179, 187 upper sum, 178 Uryson's lemma, 8 1 , 85 Uryson, Pavel S. ( 1898-1924), 8 1 , 85
variation function, 377 vector positive, 67, 242 vector lattice, 67, 242 nonned, 246 order complete, 252 vector space nonned, 218 ordered, 66 vector sublattice, 247 Vitali, Giuseppe ( 1 875-1932), 1 17, 184 void set, 2 weakly convergent sequence, 365 Weierstrass M-test, 7 1 Weierstrass, Karl T. W. ( 1 8 1 5-1897), 25, 29, 71, 88-90 weight function, 289 well ordering principle, 10 x-section of set, 207 y-section of set, 207 zero product rule, 20 Zorn's lemma, 8 Zorn, Max (1906-1993), 8
415