Chapter 4. Product Measures

§1. Product Measures

Let (X, S, µ) and (Y, T, ν) be two spaces. In order to define the product of µ and ν, we first define an outer measure on X × Y in terms of µ and ν. For each subset E of X × Y , we define  ∞  X ∞ τ(E) := inf µ(Aj)ν(Bj): Aj ∈ S, Bj ∈ T ∀ j ∈ IN and E ⊆ ∪j=1Aj × Bj . j=1 Theorem 1.1. The set function τ is an outer measure on X × Y .

Proof. Clearly, τ(∅) = 0. Moreover, if E ⊆ F ⊆ X × Y , then τ(E) ≤ τ(F ). It remains ∞ to prove that τ is countably subadditive. Suppose E ⊆ ∪k=1Ek ⊆ X × Y . We wish to P∞ P∞ P∞ show τ(E) ≤ k=1 τ(Ek). This is true if k=1 τ(Ek) = ∞. Suppose k=1 τ(Ek) < ∞. k k Fix ε > 0. For each k ∈ IN, there exist sets Aj ∈ S and Bj ∈ T for j ∈ IN such that ∞ k k Ek ⊆ ∪j=1Aj × Bj and ∞ X ε µ(Ak)ν(Bk) < τ(E ) + . j j k 2k j=1 ∞ ∞ k k Thus, E ⊆ ∪k=1 ∪j=1 Aj × Bj and

∞ ∞ ∞ ∞ X X k k X k X τ(E) ≤ µ(Aj )ν(Bj ) ≤ (τ(Ek) + ε/2 ) = ε + τ(Ek). k=1 j=1 k=1 k=1 P∞ Since ε > 0 could be arbitrary, we conclude that τ(E) ≤ k=1 τ(Ek).

Let A denote the σ-algebra of τ-measurable subsets of X × Y and let

µ × ν = τ|A.

We call µ × ν the of µ and ν. A subset E of X × Y is said to be µ × ν-measurable if and only if it is τ-measurable.

Theorem 1.2. If A ∈ S and B ∈ T , then A × B is τ-measurable and

(µ × ν)(A × B) = µ(A)ν(B).

Proof. In order to show that A × B is τ-measurable, it suffices to prove that for any G ⊆ X × Y with τ(G) < ∞ the following inequality holds:

τ(G) ≥ τ(G ∩ (A × B)) + τ(G ∩ (A × B)c).

1 ∞ Fix ε > 0. There exist sets Aj ∈ S and Bj ∈ T for j ∈ IN such that G ⊆ ∪j=1Aj × Bj P∞ and j=1 µ(Aj)ν(Bj) < τ(G) + ε. It follows that

∞ G ∩ (A × B) ⊆ ∪j=1(Aj ∩ A) × (Bj ∩ B)

and c ∞  c c  G ∩ (A × B) ⊆ ∪j=1 ((Aj ∩ A) × (Bj ∩ B )) ∪ ((Aj ∩ A ) × Bj) .

Consequently,

τ(G ∩ (A × B)) + τ(G ∩ (A × B)c) ∞ ∞ ∞ X X c X c ≤ µ(Aj ∩ A)ν(Bj ∩ B) + µ(Aj ∩ A)ν(Bj ∩ B ) + µ(Aj ∩ A )ν(Bj) j=1 j=1 j=1 ∞ ∞ ∞ X X c X = µ(Aj ∩ A)ν(Bj) + µ(Aj ∩ A )ν(Bj) = µ(Aj)ν(Bj) < τ(G) + ε. j=1 j=1 j=1

Thus, τ(G ∩ (A × B)) + τ(G ∩ (A × B)c) < τ(G) + ε for all ε > 0. Therefore,

τ(G ∩ (A × B)) + τ(G ∩ (A × B)c) ≤ τ(G) ∀ G ⊆ X × Y.

This shows that A × B is τ-measurable. We have (µ × ν)(A × B) ≤ µ(A)ν(B). To prove µ(A)ν(B) ≤ (µ × ν)(A × B), for ∞ j ∈ IN we let Aj and Bj be sets in S and T respectively such that A × B ⊆ ∪j=1Aj × Bj. It follows that ∞ X χA(x)χB(y) ≤ χAj (x)χBj (y), x ∈ X, y ∈ Y. j=1 Integrating both sides of the above inequality first with respect to ν and then with respect to µ, we obtain Z Z µ(A)ν(B) = χA(x)χB(y) dν(y) dµ(x) X Y ∞ ∞ X Z Z X ≤ χAj (x)χBj (y) dν(y) dµ(x) = µ(Aj)ν(Bj). j=1 X Y j=1

Therefore, µ(A)ν(B) ≤ (µ × ν)(A × B).

∞ A subset F of X×Y is said to be a σ-set, if F can be represented as ∪j=1Aj ×Bj, where Aj ∈ S and Bj ∈ T for j ∈ IN. By Theorem 1.2, a σ-set is (µ × ν)-measurable. Suppose

2 that E is a (µ × ν)-measurable subset of X × Y . By the definition of the product measure, for any ε > 0 there exists a σ-set F such that E ⊆ F and (µ × ν)(F ) < (µ × ν)(E) + ε. Let E be a subset of X × Y . For x ∈ X, the x-section of E is the subset of Y given by

Ex := {y ∈ Y :(x, y) ∈ E}.

For y ∈ Y , the y-section of E is the subset of X given by

Ey := {x ∈ X :(x, y) ∈ E}.

Regarding sections of sets, the following identities hold: y y (a) (∪i∈I Ai)x = ∪i∈I (Ai)x and (∪i∈I Ai) = ∪i∈I (Ai) ; y y (b) (∩i∈I Ai)x = ∩i∈I (Ai)x and (∩i∈I Ai) = ∩i∈I (Ai) ; y y y (c) (A \ B)x = Ax \ Bx and (A \ B) = A \ B . In what follows, (X, S, µ) and (Y, T, ν) are assumed to be complete measure spaces.

Theorem 1.3. Let E be a µ × ν-measurable subset of X × Y with (µ × ν)(E) < ∞.

Then for µ-almost every x ∈ X the set Ex is ν-measurable and the function x 7→ ν(Ex) is integrable over X, and Z (µ × ν)(E) = ν(Ex) dµ(x). X Similarly, for ν-almost every y ∈ Y the set Ey is µ-measurable and the function y 7→ µ(Ey) is integrable over Y , and Z (µ × ν)(E) = µ(Ey) dν(y). Y Proof. It suffices to establish the first formula. The proof goes by steps.

Step 1. Assume E = A × B, where A ∈ S and B ∈ T . Then Ex = B if x ∈ A and

Ex = ∅ if x∈ / A. Hence, ν(Ex) = ν(B)χA(x). It follows that

Z Z ν(Ex) dµ(x) = ν(B)χA(x) dµ(x) = µ(A)ν(B) = (µ × ν)(E). X X

∞ Step 2. Suppose that E is a σ-set, that is, E = ∪n=1(An × Bn), where An ∈ S and

Bn ∈ T for n ∈ IN. Let En := An × Bn. We may assume that En (n ∈ IN) are mutually ∞ P∞ disjoint. Then Ex = ∪n=1(En)x is ν-measurable and ν(Ex) = n=1 ν((En)x). Therefore,

∞ ∞ Z X Z X ν(Ex) dµ(x) = ν((En)x) dµ(x) = (µ × ν)(En) = (µ × ν)(E). X n=1 X n=1

3 ∞ Step 3. Suppose that E = ∩n=1En, where each En is a σ-set, En+1 ⊆ En for n ∈ IN and (µ × ν)(E1) < ∞. By Step 2 we have

Z ν((E1)x) dµ(x) = (µ × ν)(E1) < ∞. X

Hence, ν((E1)x) < ∞ for µ-almost every x ∈ X. Moreover, (En+1)x ⊆ (En)x for n ∈ IN. ∞ Thus, Ex = ∩n=1(En)x is ν-measurable and

Z Z ν(Ex) dµ(x) = lim ν((En)x) dµ(x) = lim (µ × ν)(En) = (µ × ν)(E). X n→∞ X n→∞

Step 4. Suppose (µ × ν)(E) = 0. There exists a sequence (En)n=1,2,... of σ-sets such ∞ that E ⊆ ∩n=1En, En+1 ⊆ En for n ∈ IN, (µ × ν)(E1) < ∞, and limn→∞(µ × ν)(En) = 0. ∞ ∞ Let F := ∩n=1En. Then Fx = ∩n=1(En)x is ν-measurable for each x ∈ X and

Z Z ν(Fx) dµ(x) = lim ν((En)x) dµ(x) = lim (µ × ν)(En) = 0. X n→∞ X n→∞

Hence, ν(Fx) = 0 for µ-almost every x ∈ X. But Ex ⊆ Fx and (X, S, µ) is complete.

Therefore, for µ-almost every x ∈ X, Ex is ν-measurable and ν(Ex) = 0. Consequently,

Z ν(Ex) dµ(x) = 0 = (µ × ν)(E). X

Step 5. The general case. Since (µ×ν)(E) < ∞, There exists a sequence (En)n=1,2,... ∞ of σ-sets such that E ⊆ F := ∩n=1En,(µ × ν)(F \ E) = 0, En+1 ⊆ En for n ∈ IN, and

(µ × ν)(E1) < ∞. By Step 3 and Step 4, for µ-almost every x ∈ X, Fx and (F \ E)x are

ν-measurable, so Ex = Fx \ (F \ E)x is ν-measurable. Moreover,

Z Z (µ × ν)(F ) = ν(Fx) dµ(x) and (µ × ν)(F \ E) = ν(Fx \ Ex) dµ(x). X X

Subtracting the second equality from the first one, we obtain

Z (µ × ν)(E) = ν(Ex) dµ(x). X

The proof is complete.

4 §2. The Fubini-Tonelli Theorem

Throughout this section, (X, S, µ) and (Y, T, ν) are assumed to be two complete mea- sure spaces. Let f be a function from X × Y to IR. For each fixed x ∈ X, we define

fx(y) := f(x, y), y ∈ Y,

and for each fixed y ∈ Y define

f y(x) := f(x, y), x ∈ X.

Theorem 2.1. Let f : X × Y → IR be a (µ × ν)-integrable function. Then fx is ν- integrable for µ-almost every x ∈ X, f y is µ-integrable for ν-almost every y ∈ Y , the R R y function g : x 7→ Y fx dν is µ-integrable, the function h : y 7→ X f dµ is ν-integrable, and Z Z Z  Z Z  f d(µ × ν) = f(x, y) dν(y) dµ(x) = f(x, y) dµ(x) dν(y). X×Y X Y Y X Proof. The proof goes by steps.

Step 1. If f = χE, where E is (µ × ν)-measurable and (µ × ν)(E) < ∞, then our assertions are true, by Theorem 1.3. Pm Step 2. Our assertions are valid if f = j=1 cjχEj is a simple function with cj ∈ IR and (µ × ν)(Ej) < ∞ for each j = 1, . . . , m.

Step 3. Suppose f is integrable and f ≥ 0. Then there exists a sequence (φn)n=1,2,... of measurable simple functions such that 0 ≤ φn ≤ φn+1 ≤ f for all n ∈ IN and limn→∞ φn(x, y) = f(x, y) for all (x, y) ∈ X × Y . Since f is integrable, each φn is

integrable. By Step 2, (φn)x is ν-integrable for µ-almost every x ∈ X, the function R gn : x 7→ Y φn(x, y) dν(y) is µ-integrable over X, and Z Z gn(x) dµ(x) = φn d(µ × ν). X X×Y By the monotone convergence theorem, for µ-almost every x ∈ X, Z Z lim gn(x) = lim φn(x, y) dν(y) = f(x, y) dν(y) = g(x). n→∞ n→∞ Y Y

Note that 0 ≤ gn ≤ gn+1 for all n ∈ IN. By using the monotone convergence theorem again we obtain Z Z Z Z g(x) dµ(x) = lim gn(x) dµ(x) = lim φn d(µ × ν) = fd(µ × ν). X n→∞ X n→∞ X×Y X×Y

5 The assertion on h can be proved analogously. Step 4. In general, suppose that f is a (µ × ν)-integrable function on X × Y . We have f = f + − f −, where f + := max{f, 0} and f − := max{−f, 0}. Then f + and f − are integrable. By Step 3, our assertions are valid for f + and f −, and thereby valid for f.

Theorem 2.2. Let (X, S, µ) and (Y, T, ν) be two σ-finite complete measure spaces, and let f : X × Y → IR be a (µ × ν)-measurable function. If one of the iterated integrals R R R R X Y |f| dν dµ or Y X |f| dµ dν is finite, then the function f is integrable and Z Z Z  Z Z  f d(µ × ν) = f(x, y) dν(y) dµ(x) = f(x, y) dµ(x) dν(y). X×Y X Y Y X

Proof. It suffices to consider the case f ≥ 0. Since (X, S, µ) and (Y, T, ν) are σ-finite measure spaces, the product measure µ × ν is also σ-finite. There exists a sequence

(φn)n=1,2,... of (µ × ν)-integrable functions such that 0 ≤ φn ≤ φn+1 ≤ f for all n ∈ IN and limn→∞ φn(x, y) = f(x, y) for all (x, y) ∈ X × Y . By Theorem 2.1, for each n ∈ IN we have Z Z Z  φn d(µ × ν) = φn(x, y) dν(y) dµ(x). X×Y X Y By the monotone convergence theorem we obtain Z Z Z Z Z Z f d(µ × ν) = lim φn d(µ × ν) = lim φn dν dµ = f dν dµ. X×Y n→∞ X×Y n→∞ X Y X Y R R Thus, if the integral X Y f dν dµ is finite, then f is a (µ × ν)-integrable function. Hence, the desired result follows from Theorem 2.1.

Let us consider the following example. Suppose that f is the function defined on [0, 1] × [0, 1] given by

( 2 2 x −y if (x, y) 6= (0, 0), f(x, y) = (x2+y2)2 0 if (x, y) = (0, 0).

Then we have

Z 1 Z 1 Z 1h −x ix=1 Z 1 dy π f(x, y) dx dy = 2 2 dy = − 2 = − 0 0 0 x + y x=0 0 1 + y 4 and Z 1 Z 1 Z 1h y iy=1 Z 1 dx π f(x, y) dy dx = 2 2 dy = 2 = . 0 0 0 x + y y=0 0 1 + x 4

6 §3. The on IRn

n Let λ be the Lebesgue measure on IR. For n ∈ IN, the Lebesgue measure λn on IR is defined by λ1 := λ and λn := λ × λn−1 for n > 1.

In what follows, n is fixed and λn is abbreviated as λ. Let Λ denote the collection of all Lebesgue measurable subsets of IRn.

Suppose −∞ < aj < bj < ∞ for j = 1, . . . , n. Then the product (a1, b1)×· · ·×(an, bn) is called an open cell in IRn. Clearly, an open cell is Lebesgue measurable. An open subset of IRn can be represented as a countable union of open cells in IRn. Hence, an in IRn is Lebesgue measurable. Consequently, a closed set in IRn is Lebesgue measurable. Theorem 3.1. Let E be a Lebesgue measurable subset of IRn. Then for any ε > 0 there exists an open set U ⊇ E and a closed set F ⊆ E such that λ(U \E) < ε and λ(E \F ) < ε. Proof. (a) First, consider the case λ(E) < ∞. For given ε > 0, there exists a sequence n (Am)m=1,2,... of subsets of IR with the following properties: (1) each Am can be repre- sented as Am1 × · · · × Amn, where Am1, ··· ,Amn are Lebesgue measurable subsets of IR; ∞ P∞ (2) E ⊆ A := ∪m=1Am; (3) m=1 λ(Am) < λ(E)+ε/2. For each Am, we can find an open n m+1 ∞ subset Gm of IR such that Gm ⊇ Am and λ(Gm) < λ(Am) + ε/2 . Let U := ∪m=1Gm. Then U is an open set, U ⊇ E, and ∞ X λ(U \ E) ≤ λ(U \ A) + λ(A \ E) ≤ λ(Gm \ Am) + ε/2 < ε. m=1 ∞ Next, consider the case λ(E) = ∞. We express E as a disjoint union ∪j=1Ej, where each Ej is Lebesgue measurable and λ(Ej) < ∞. Let ε > 0 be given. For each j ∈ IN, j ∞ there exists an open set Uj ⊇ Ej such that λ(Uj \ Ej) < ε/2 . Let U := ∪j=1Uj. Then U is an open set, U ⊇ E, and ∞ X λ(U \ E) ≤ λ(Uj \ Ej) < ε. j=1 For any ε > 0, there is an open set U such that U ⊇ Ec and λ(U \ Ec) < ε. Let F := U c. Then F is a closed set and F ⊆ E. Moreover, E \ F = E ∩ F c = E ∩ U = U \ Ec. Hence, λ(E \ F ) = λ(U \ Ec) < ε.

Let B be the σ-algebra generated by open subsets of IRn. A member of B is called a Borel set. Since any open subset of IRn is Lebesgue measurable, B ⊆ Λ. In other n words, any Borel set in IR is Lebesgue measurable. Recall that a Gδ set is a countable

intersection of open sets and an Fσ set is a countable union of closed sets. Evidently, Gδ sets and Fσ sets are Borel sets.

7 Theorem 3.2. The following statements are equivalent for a subset E of IRn: (a) E is Lebesgue measurable;

(b) E = V \ N1, where V is a Gδ set and λ(N1) = 0;

(c) E = H ∪ N2, where H is an Fσ set and λ(N2) = 0.

Proof. Clearly, either (b) or (c) implies (a). By Theorem 3.1, for j = 1, 2,..., there exist

an open set Uj ⊇ E and a closed set Kj ⊆ E such that

λ(E \ Kj) < 1/j and λ(Uj \ E) < 1/j.

∞ ∞ Let H := ∪j=1Kj and V := ∩j=1Uj. Then V is a Gδ set and H is an Fσ set. Moreover, λ(V \ E) = 0 and λ(E \ H) = 0.

For E ⊆ IRn and t ∈ IRn, define E + t := {x + t : x ∈ E}.

Theorem 3.3. Let E ⊆ IRn and t ∈ IRn. Then λ∗(E + t) = λ∗(E). Consequently, if E is Lebesgue measurable, then E + t is also Lebegue measurable.

Proof. Suppose A is a product of measurable sets, that is, A = A1 × · · · × An, where n A1,...,An are Lebesgue measurable subsets of IR. Then for t = (t1, . . . , tn) ∈ IR we have

n n ∗ ∗  Y ∗ Y ∗ ∗ λ (A + t) = λ (A1 + t1) × · · · × (An + tn) = λ (Aj + tj) = λ (Aj) = λ (A). j=1 j=1

n Let E be a subset of IR . For given ε > 0, there exists a sequence (Am)m=1,2,... of ∞ P∞ ∗ products of measurable sets such that E ⊆ ∪m=1Am and m=1 λ(Am) ≤ λ (E) + ε. We ∞ have E + t ⊆ ∪m=1(Am + t). It follows that

∞ ∞ ∗ X X ∗ λ (E + t) ≤ λ(Am + t) = λ(Am) ≤ λ (E) + ε. m=1 m=1

Since ε > 0 could be arbitrary, we have λ∗(E + t) ≤ λ∗(E). A similar argument gives λ∗(E) = λ∗(E + t − t) ≤ λ∗(E + t). This shows λ∗(E + t) = λ∗(E).

Now suppose E is Lebesgue measurable. By Theorem 3.2, there exists a Gδ set ∗ V ⊇ E such that λ (V \ E) = 0. Clearly, V + t is a Gδ set and V + t ⊇ E + t. Moreover, λ∗((V + t) \ (E + t)) = λ∗(V \ E) = 0. Hence, E + t is Lebesgue measurable.

8 §4. The Lebesgue Integral on IRn

Theorem 4.1. If f is a Lebesgue integrable function on IRn, then for given ε > 0 there exist finitely many open cells I1,...,Im and real numbers c1, . . . , cm such that Z m X f − cjχIj dλ < ε. n IR j=1 Proof. Since f is Lebesgue integrable, for ε > 0 we can find finitely many measurable sets

E1,...,Es with finite measure and real numbers c1, . . . , cs such that the simple function Ps φ := arχE satisfies r=1 r Z ε f − φ dλ < . IRn 2 Ps Let M := r=1 |ar|. In light of the proof of Theorem 3.1, for each Er there exists a ∞ countable family of open cells Iri (i ∈ IN) such that Er ⊆ ∪i=1Iri and ∞ X ε λ(E ) ≤ λ(I ) < λ(E ) + . r ri r 2M i=1

Consequently, for each r ∈ {1, . . . , s}, there exists some kr ∈ IN such that

kr Z X ε χ − χ dλ < . Er Iri n 2M IR i=1

Ps Pkr R Let g := r=1 i=1 arχIri . Then we have IRn |f − g| dλ < ε.

Theorem 4.2. If f : IRn → IR is a Lebesgue integrable function, then for given ε > 0 there exists a continuous function g : IRn → IR of compact support such that Z

f − g dλ < ε. IRn

Proof. By Theorem 4.1, it suffices to establish the theorem for the case f = χI , where

I = (a1, b1) × · · · × (an, bn) is an open cell. Suppose −∞ < a < b < ∞. For m = 1, 2,..., (a,b) (a,b) let gm be the function on IR given by gm (t) = 1 for t ∈ [a + 1/m, b − 1/m], gm(t) = 0 for t ∈ IR \ (a, b),

(a,b) (a,b) gm (t) = m(t − a) for a ≤ t ≤ a + 1/m and gm (t) = m(b − t) for b − 1/m ≤ t ≤ b.

(a,b) Then gm is a continuous function supported on [a, b]. For m = 1, 2,..., define

n Y (ak,bk) n gm(x) := gm (xk), x = (x1, . . . , xn) ∈ IR . k=1

9 n Each gm is a continuous function on IR with compact support. It is easily seen that Z lim |f − gm| dλ = 0. m→∞ IRn R Hence, for given ε > 0, there exists some N ∈ IN such that IRn |f − gN | dλ < ε.

Theorem 4.3. If f : IRn → IR is a Lebesgue integrable function, then for t ∈ IRn, Z Z f(· − t) dλ = f dλ. IRn IRn

Proof. By Theorem 3.3, our assertion is valid if f is an integrable simple function. Con- sider the case f ≥ 0. In this case, there exists a sequence (φn)n=1,2,... of measurable simple functions such that 0 ≤ φn ≤ φn+1 ≤ f for all n ∈ IN and limn→∞ φn(x) = f(x) for all n n x ∈ IR . It follows that limn→∞ φn(x − t) = f(x − t) for all x ∈ IR . By the monotone convergence theorem we have Z Z Z Z f(· − t) dλ = lim φn(· − t) dλ = lim φn dλ = f dλ. IRn n→∞ IRn n→∞ IRn IRn

The general case is established by considering f = f + − f −.

10