Decreasing Rearrangement and Lorentz L(p, q) Spaces

Size: px
Start display at page:

Download "Decreasing Rearrangement and Lorentz L(p, q) Spaces"

Transcription

1 Decreasing Rearrangement and Lorentz L(p, q) Spaces Erik Kristiansson December 22 Master Thesis Supervisor: Lech Maligranda Department of Mathematics

2 Abstract We consider the decreasing rearrangement of a measurable function and we prove some fundamental properties like the Hardy-Littlewood inequality. This notion leads us to the Lorentz L(p, q) spaces which are generalizations of the ordinary L p spaces. We will introduce a quasi-norm and also a norm on L(p, q) and examine their metrical and topological properties as well as boundedness of the maximal operator and the Hilbert transform.

3 Acknowledgment I want to take the opportunity to thank my supervisor Lech Maligranda for his excellent supervision, encouragement, and for introducing me to the world of functional analysis. I would also like to thank all the personal, and specially the Ph.D. student Johan Dasht, here at the Department of Mathematics of the Luleå University of Technology for all the help and support. 1

4 Contents 1 Introduction 3 2 Decreasing rearrangement Distribution function Decreasing rearrangement Examples of decreasing rearrangement Properties of decreasing rearrangement Decreasing rearrangement and L p -norms The Hardy-Littlewood inequality Operation ** Lorentz L(p, q) spaces Definition of the Lorentz L(p, q) spaces Normed Lorentz L(p, q) spaces Topological properties of the Lorentz L(p, q) spaces Applications PDEs and decreasing rearrangement The maximal operator The Hilbert transform

5 Chapter 1 Introduction The notion of rearrangement of a function has long been an important tool in classical analysis, playing a key role in many inequalities. Systematically introduced by Hardy and Littlewood it has been used by several authors in real and harmonic analysis, in investigations of singular integrals, function spaces and interpolation (see [1], [12] for instance). The contents of this report is as follows. The decreasing rearrangement f of a measurable function f is a decreasing right continuous function defined on (, ) which is equimeasurable with f. To make a formal definition of f the definition of the distribution function µ f is needed. These definitions can be found in Chapter 2 together with some fundamental properties like the uniqueness of f. In the last section of this chapter some examples of functions and their distribution function and decreasing rearrangement are given. In Chapter 3 some important theorems regarding the decreasing rearrangements and the famous Hardy-Littlewood inequality, f(x)g(x) dµ f (t)g (t) dt are proved. If we, in addition, assume that the measure space is nonatomic this inequality can also generalized to n functions. The last part of Chapter 3 contains some results regarding equality in the Hardy-Littlewood inequality and an explanation of the concept of resonant measure spaces. The decreasing rearrangement will lead us to the Lorentz L(p, q) space which 3

6 is the set of all classes of measurable functions such that the functional pq is finite, where ( ( t 1/p f (t) ) ) 1/q q dt if < p <, < q <, t f pq = sup t 1/p f (t) if < p, q =. t> The Lorentz L(p, q) spaces are investigated in Chapter 4. These spaces are generalizations of ordinary L p spaces and they coincides with L p when q = p. We show that the L(p, q) spaces are linear spaces (vector spaces) but to introduce a norm on L(p, q) is not a trivial task. It turns out that the functional pq will serve as a norm only for 1 q p < or p = q =. To deal with the case when q > p we need to introduce another functional pq which is defined as follows ( ( t 1/p f (t) ) ) 1/q q dt if < p <, < q <, f t pq = sup t 1/p f (t) if < p, q =. t> This functional is subadditive and equivalent to pq for all 1 < p < and 1 q which means that L(p, q) is a normed space with the norm pq. We will also show that in the remaining cases L(p, q) is not normable, that is, we can not on L(p, q) introduce an norm that is equivalent to pq. Next, we examine the topological properties of the Lorentz L(p, q) spaces, first its completeness and then the separability. Finally, the duality of L(p, q) spaces is proved in the last part of Chapter 4. In Chapter 5 we are giving application of rearrangement in PDE s. It is shown that from all the bodies in R 3 with density 1 and fixed volume, it is the ball that has the gravitational field of highest energy. Then, using our earlier investigations, we show that both classical Hardy-Littlewood maximal operator and the Hilbert transform are bounded in Lorentz L(p, q) spaces for 1 < p <. 4

7 Chapter 2 Decreasing rearrangement This chapter is devoted to the decreasing rearrangement. In the first section the distribution function of a measurable function is defined and some basic properties are proved. The next section contains the decreasing rearrangement and which is, in fact, a distribution function with respect to the Lebesgue measure. Then some fundamental properties of rearrangement are proved and we show by a counter-example that the operation * is neither subadditive nor submultiplicative. This fact will be very important when we discuss the Lorentz L(p, q) spaces in Chapter 4. We end this chapter with some examples, where the most important part are the distribution function and decreasing rearrangement of a simple function. We start by making a few assumptions which will remove much of the unnecessary troubles. Let (, Σ, µ) denote a σ-finite measure space. On we let f and g denote two Σ-measurable functions. We will keep these assumption till the end of Chapter 5 and sometimes we write instead of all (, Σ, µ). 2.1 Distribution function We begin with a definition of the distribution function. Definition 2.1. The function µ f : [, ) [, ] defined by µ f (λ) = µ({x : f(x) > λ}), λ (2.1) is called the distribution function of f. 5

8 The distribution function depends only on the absolute value of f and is clearly decreasing. Other useful properties are collected in the following theorem. Theorem 2.1. Let f n be Σ-measurable functions on, n = 1, 2,.... Then (i) µ f is decreasing and continuous from the right. (ii) If f(x) g(x) for x µ-a.e., then µ f (λ) µ g (λ) for any λ. (iii) µ f+g (λ 1 + λ 2 ) µ f (λ 1 ) + µ g (λ 2 ) for any λ 1, λ 2. (iv) µ fg (λ 1 λ 2 ) µ f (λ 1 ) + µ g (λ 2 ) for any λ 1, λ 2. (v) If f(x) lim inf n f n (x) for x µ-a.e., then µ f (λ) lim inf n µ fn (λ) for any λ. Proof. (i) Let λ 1 < λ 2 be arbitrary. Then {x : f(x) > λ 1 } {x : f(x) > λ 2 } so by the monotonicity of the measure we have that µ f (λ 1 ) µ f (λ 2 ), that is, µ f is decreasing. To prove that µ f is continuous from the right, let A λ = {x : f(x) > λ} and let λ be an arbitrary fixed number. The sets A λ increases as λ decreases, and A λ = A λ +1/n. n=1 Hence, by the continuity of the measure, µ f (λ + 1/n) = µ(a λ +1/n) µ(a λ ) = µ f (λ ) as n. Thus µ f is continuous from the right. (ii) If f(x) g(x) for x µ-a.e., then {x : g(x) > λ} {x : f(x) > λ}. Using the monotonicity of the measure, µ f (λ) µ g (λ) for all λ. (iii) Since {x : f(x) + g(x) > λ 1 + λ 2 } {x : f(x) > λ 1 } {x : g(x) > λ 2 } 6

9 it follows that µ f+g (λ 1 + λ 2 ) µ f (λ 1 ) + µ g (λ 2 ). (iv) As in (iii) we have {x : f(x)g(x) > λ 1 λ 2 } {x : f(x) > λ 1 } {x : g(x) > λ 2 } which implies that µ fg (λ 1 λ 2 ) µ f (λ 1 ) + µ g (λ 2 ). (v) Let A n = {x : f n (x) > λ}, n = 1, 2,.... By assumption and the definition of limes inferior, f(x) lim inf n f n(x) = sup m N inf f n(x). n>m This means, that for all x such that f(x) > λ there exists an integer m such that for all integers n > m, f n (x) > λ. Hence A λ m=1 n=m and for any m 1 ( ) µ A n inf µ (A n) sup n m m n=m A n, (2.2) inf µ(a n) = lim inf µ(a n) n m n Furthermore, by (2.2), the monotonicity of µ and since { n=ma n } m=1 is an increasing sequence of sets we finally get ( ) ( ) µ f (λ) = µ(a λ ) µ E n = lim µ E n m lim inf n µ f n (λ). m=1 n=m n=m 2.2 Decreasing rearrangement The next step is to introduce the decreasing rearrangement and its important properties. We start with a definition. 7

10 Definition 2.2. The decreasing rearrangement of f is the function f : [, ) [, ] defined by where we use the convention that inf =. f (t) = inf{λ : µ f (λ) t}, (2.3) If µ f is strictly decreasing, then f is clearly the inverse of µ f. Another very important relationship between the distribution function and the decreasing rearrangement is described in the following theorem. Theorem 2.2. The equation holds, where m is the Lebesgue measure. f (t) = m µf (t), t (2.4) Proof. Since µ f is a decreasing function by Theorem 2.1 (i) it follows that sup{λ : µ f (λ) > t} = m{λ : µ f (λ) > t}. Hence by Definition 2.1 we get f (t) = inf{λ : µ f (λ) t} = sup{λ : µ f (λ) > t} = m{λ : µ f (λ) > t} = m µf. Remark. The previous theorem shows that a decreasing rearrangement is in fact a distribution function with respect to the Lebesgue measure on the semi-axis [, ). Therefore the properties in Theorem 2.1 hold for all decreasing rearrangements. The next theorem establish some basic properties of the decreasing rearrangement. Theorem 2.3. The following properties holds. (i) f (t) > λ if and only if µ f (λ) > t. (ii) f and f are equimeasurable, that is, µ({x : f(x) > λ}) = m({t > : f (t) > λ}) (2.5) for all λ, where m is the Lebesgue measure. 8

11 (iii) If λ and µ f (λ) <, then f (µ f (λ)) λ and f (µ f (λ) + ε) λ for all < ε < f (t). If t and f (t) <, then µ f (f (t)) t and µ f (f (t) ε) t for all ε >. (iv) for < p < ( f p ) (t) = f (t) p (v) If A Σ, then (fχ A ) (t) f (t)χ [,µ(a)) (t) for all t. Proof. (i) First assume that µ f (λ) > t. Then, since µ f is a decreasing function, λ < inf{ν : µ f (ν) t}. Thus f (t) > λ. Now assume that f (t) > λ, that is, inf{ν : µ f (ν) t)} > λ. Thus, since µ f is a decreasing function, µ f (λ) > t. (ii) Let m be the Lebesgue measure on [, ). Then by (i) we get m f (λ) = m({t : f (t) > λ}) = m({t : µ f (λ) > t}) = m([, µ f (λ))) = µ f (λ) (iii) Assume that µ f (λ) <. Since µ f is a decreasing function we get Also, for any arbitrary ε >, f (µ f (λ)) = inf{s : µ f (s) µ f (λ)} λ. f (µ f (λ) + ε) = inf{s : µ f (s) µ f (λ) + ε} λ. Now assume that f (t) <. Then µ f (f (t)) = µ f (inf{λ : µ f (λ) t}) t by the right-continuity of µ f. Furthermore, for any ε >, we have by (ii) that µ f (f (t) ε) = µ({x : f(x) > f (t) ε}) (iv) Since λ < we get = m({s > : f (s) > f (t) ε}) t ( f p ) (t) = inf{λ : µ({x : f(x) p > λ}) t} = inf{ν p : µ({x : f(x) > ν}) t} = f (t) p 9

12 where ν = λ (1/p). (v) Since (fχ A )(x) f(x) for all x we have by Theorem 2.1(ii) and Theorem 2.2 that (fχ A ) (t) g (t), t. On the other hand, since µ({x : (fχ A )(x) > λ}) µ(a), we also have that (fχ A ) (t) =, t > µ(a). Combining these two estimates we can conclude that for all t. (fχ A ) (t) f (t)χ [,µ(a)) (t) Remark. Observe that property (ii) does not hold if we remove the strict inequalities, that is µ({x : f(x) λ) = m({t R : f (t) λ}) does not hold in the general case. This can easily be seen by taking Then f (t) 1 which means that f(x) = x 1 + x. µ({x : x/(1 + x) 1}) = = m({t R : 1 1}). An important question is if the decreasing rearrangement is unique and we will answer on this in the next theorem. Theorem 2.4 (Uniqueness). There exists only one right-continuous decreasing function f equimeasurable with f. Hence, the decreasing rearrangement is unique. Proof. Let f 1 (t) and f 2 (t) be two different right-continuous decreasing functions equimeasurable with f. Then there exist a t such that f 1 (t ) f 2 (t ) and we can assume without loss of generality that f 1 (t ) > f 2 (t ). Choose ε > such that f 1 (t ) > f 2 (t ) + ε. Then by the right continuity of f 1 (t) there exist a nonempty interval [t, t 1 ] such that f 1 (t) > f 2 (t ) + ε for all 1

13 t [t, t 1 ]. Using this estimate and the fact that both f 1 (t) and f 2 (t) are decreasing we get that and m({t : f 1 (t) > f 2 (t) + ε}) t 1 m({t : f 2 (t) > f 2 (t) + ε}) t 2 < t 1 which is a contradiction to the equimeasurability with f. Hence, f 1 (t) = f 2 (t), i.e. the decreasing rearrangement is unique. Remark. The operation * is neither subadditive, in the sense that (f + g) (t) f (t) + g (t) is not true in general, nor submultiplicative, that is, (fg) (t) f (t)g (t) does not hold for any t. The next example shows this. Example 2.1. Let A and B be measurable sets such that A B and < µ(a) < µ(b). Put f(x) = χ A (x) and g(x) = χ B (x). The decreasing rearrangements are f (t) = χ [,µ(a)) (t) and g (t) = χ [,µ(b)) (t) which means that 2 if t < µ(a), f (t) + g (t) = 1 if µ(a) t < µ(b), if t µ(b). Moreover, since (f + g)(x) = χ A (x) + χ B (x) it follows that 2 if t < µ(a B), (f + g) (t) = 1 if µ(a B) t < µ(a B), if t µ(a B). Hence, for any t such that µ(b) < t < µ(a B) we have (f + g) (t) = 1 > = f (t) + g (t). Similar calculations shows that * neither is submultiplicative. 11

14 Despite the previous remark and example we have the following theorem. Theorem 2.5. The inequalities (f + g) (t 1 + t 2 ) f (t 1 ) + g (t 2 ) and (fg) (t 1 + t 2 ) f (t 1 )g (t 2 ) hold for all t 1, t 2. In particular, (f + g) (t) f (t/2) + g (t/2) and for all t. (fg) (t) f (t/2)g (t/2) Proof. We start with the first inequality. Let f (t 1 ) + g (t 2 ) <, otherwise there is nothing to prove. Then for λ 1 = f (t 1 ) and λ 2 = g (t 2 ), by Theorem 2.3(iii), µ f (λ 1 ) t 1 and µ g (λ 2 ) t 2 and by Theorem 2.1(iii) µ f+g (λ 1 + λ 2 ) µ f (λ 1 ) + µ g (λ 2 ) t 1 + t 2. Using the definition of the decreasing rearrangement we obtain (f + g) (t 1 + t 2 ) λ 1 + λ 2 = f (t 1 ) + g (t 2 ). The proof of the second inequality is similar. We assume that f (t 1 )g (t 2 ) < and use Theorem 2.1(iv) to get that µ fg (λ 1 λ 2 ) µ f (λ 1 ) + µ g (λ 2 ). Again, using the definition of the decreasing rearrangement, we get (fg) (t 1 + t 2 ) λ 1 λ 2 = f (t 1 )g (t 2 ) which is our desired inequality. The rest of the theorem now follows by taking t 1 = t 2 = t/2. 12

15 2.3 Examples of decreasing rearrangement In this section we take a look on some examples of distribution function and decreasing rearrangement. The first example establish some important relations between a simple function, its distribution function and decreasing rearrangement. Example 2.2 (Decreasing rearrangement of a simple function). Let s be a simple function of the following form s(x) = α j χ Aj (x) where α 1 > α 2 > > α k >, A j = {x : s(x) = α j } and χ A is the characteristic function of the set A. Then µ s (λ) = µ({x : s(x) > λ}) = µ({x : = β j χ Bj (λ) α j χ Aj (x) > λ}) = where β j = j i=1 µ(a i), B j = [α j+1, α j ) for j = 1, 2,..., k and α k+1 = which shows that the distribution function of a simple function is a simple function itself (see Figure 2.1b). We can also find the decreasing rearrangement s (t) = inf{λ : µ s (λ) t} = inf{λ : = α j χ [βj 1,β j )(t) β j χ Bj (x) t} which is also a simple function (see Figure 2.1c). 13

16 s α 1 µ s s α 1 α 2 α 3 β 4 β 3 β 2 α 2 α 3 α 4 β 1 α 4 A 3 A 4 A 1 (a) A 2 x α 4 α 3 α 2 α 1 λ (b) β 1 β 2 β 3 β 4 (c) t Figure 2.1: Simple function, its distribution function and decreasing rearrangement Example 2.3. Let = [, ), Σ = all Lebesgue-measurable subsets of and µ = m, where m denotes the Lebesgue measure on. Define f : [, ) [, ) as { 1 (x 1) 2 if x 2, f(x) = if x > 2. The distribution function can be easily calculated { 2 1 λ if λ 1, m f (λ) = if λ > 1, and the decreasing rearrangement becomes { f 1 t 2 /4 if t 2, (t) = if t > 2. Figure 2.2 shows f, m f and f. Observe that the integral over f, m f and f are all the same, that is, f(x) dx = 2 1 (x 1) 2 dx = 1 2 (1 λ) dλ = 2 1 t 2 /4 dt = 4/3. 14

17 f 2 1 m f f λ 1 2 (a) (b) (c) x t Figure 2.2: The function f, its distribution function and decreasing rearrangement. Example 2.4. Let = [, ) and Σ and µ be as in the previous example. First we define an extended function f : [, ) [, ] as if x =, ln( 1 ) if < x < 1, 1 x f(x) = if 1 x 2, ln( 1 ) if 2 < x < 3, x 2 if x 3. Even if f is infinite over some interval the distribution function and the decreasing rearrangement are still defined and can be calculated and µ f (λ) = 1 + 2e λ, λ if t 1, f (t) = ln( 2 ) if 1 < t < 3, t 1 if t 3. Now, consider the function g : [, ) [, ) defined by g(x) = x for all x [, ). Then µ g (λ) = for all λ, which implies that f (t) = for 15

18 all t. Finally, let h : [, ) [, ) be defined as h(x) = x 1 + x, x. Observing Figure 2.3 we immediately get that { if λ < 1, µ h (λ) = if 1 λ. Thus, the unique decreasing right continuous function equimeasurable with h is the constant function h (t) 1. These examples shows that the distribution function and the decreasing rearrangement can be infinite, even if the original function is finite-valued. h 1 h Figure 2.3: The function h and its decreasing rearrangement h from Example 2.4. Example 2.5. Take = N, Σ = 2 N and µ({n}) = 1. Then any Σ- measurable function f : R is a sequence {a n }, n = 1, 2,... and the distribution function can be rewritten as µ f (λ) = µ({n N : a n > λ}) for λ. As always, the decreasing rearrangement is a function defined for all t but we can interpret it as a sequence if we define that is, a n = f (t) for n 1 t < n, a n = inf{λ : µ f (λ) k 1} = inf{λ : µ({n N : a n > λ}) n 1} Hence {a k } is the sequence { a n } permutated in a decreasing order. 16

19 Chapter 3 Properties of decreasing rearrangement The purpose of this third chapter is to establish some important and useful theorems about decreasing rearrangement which we will need later on. We start with a theorem stating that the L p norm of a function can be calculated from the distribution function or the decreasing rearrangement. This result can be found in the first section. In the next section we introduce and prove the famous Hardy-Littlewood inequality, first for two functions on an arbitrary σ-finite measure space and later for more functions when the measure µ is nonatomic. This inequality leads us to an operation **, which is an average of the decreasing rearrangement and this average turns out to be the largest possible. Section 3 is devoted to this averaging operation **, where we prove some basic properties and establish the very important subadditivity property. Indeed, its subadditivity is the key to the triangle inequality of the norm of Lorentz L(p, q) spaces which will be introduced in Chapter 4. We still keep the assumptions from Chapter 2, that is, (, Σ, µ) is a σ-finite measure space and f and g are Σ-measurable functions on. 3.1 Decreasing rearrangement and L p -norms The last equalities in Example 2.3 in the previous chapter comes from a more general theorem which states that the L p norm of a measurable function and 17

20 its decreasing rearrangement are the same. As we will see later on, this result is extremely important in evaluating the norm of elements in the Lorentz L(p, q) space. The case when p = 1 serves as a lemma and the general case follows directly after that. Lemma 3.1. The equalities f(x) dµ = and hold. sup λ> µ f (λ) dλ = f (t) dt. (3.1) λµ f (λ) = sup tf (t) (3.2) t> Proof. We first prove (3.1) for simple positive functions. Let s be a positive simple function on of the form s(x) = α j χ Aj (x) (3.3) where α 1 > α 2 > > α k > and all A j are pairwise disjoint sets. Then (see Example 2.2) µ s (λ) = β j χ Bj (λ) where β j = j i=1 µ(a i), B k = [α j+1, α j ),,2,...,k, and α k+1 =. Thus, by the definition of the integral of a simple function, Furthermore, µ s (λ) dλ = β j m([α j+1, α j )) = β j (α j α j+1 ). (3.4) β j (α j α j+1 ) = β 1 (α 1 α 2 ) + β 2 (α 2 α 3 ) + + β k α k = α 1 β 1 + α 2 (β 2 β 1 ) + + α k 1 (β k 1 β k 2 ) + α k β k = 18 α j µ(a j )

21 and using (3.4) and again the definition of the integral of a simple function, we get that µ s (λ) dλ = α j µ(a j ) = s(x) dµ. This proves the first part of (3.1) for simple functions. The next equality in (3.1) follows by observing that for a positive simple function s, so s (t) dt = s (t) = α j χ [βj 1,β j )(t) α j (β j β j 1 ) = α j µ(a j ) = s(x) dµ. The general case now follows from Theorem 2.1(v), Theorem 2.2 and the Monotone Convergence Theorem. Now we prove (3.2). If there exist a λ (, ) such that µ(λ ) = then sup λ> λµ f (λ) = and sup t> tf (t) sup tλ =. t> Conversely, if there exist a t (, ) such that f (t ) = then sup t> tf (t) = and sup λ> λµ f (λ) sup λt =. λ> This proves that (3.2) hold for the case when µ f (λ) or f (t) are infinite for a certain λ > or t >, respectively. We can therefore assume that µ f (λ) < for all λ and f (t) < for all t. To complete the proof of (3.2) we start by proving that sup λ> λµ f (λ) sup tf (t). t> Assume that sup t> tf (t) < otherwise there is nothing to prove. Since µ f (λ) < for all λ we get by Theorem 2.3(iii) that sup tf (t) µ f (λ)f (µ f (λ)) µ f (λ)λ t> 19

22 for all λ >. Taking supremum over all λ > on both sides we obtain that sup λ> λµ f (λ) sup tf (t). t> We now prove the other direction, that is, sup t> tf (t) sup λµ f (λ). λ> Assume that sup λ> λµ f (λ) < since otherwise there is nothing to prove. By hypothesis, f (t) < for all t > so by Theorem 2.3 we have for all < ε < f (t) that sup λµ f (λ) (f (t) ε)µ f (f (t) ε) (f (t) ε)t λ> Since this inequality holds for any ε > it follows that sup λµ f (λ) tf (t) λ> and taking the supremum over all t > on both sides we obtain that sup λ> Thus, the equality (3.2) is proved. λµ f (λ) sup tf (t). t> Theorem 3.2. Let < p <. Then f(x) p dµ = p λ p 1 µ f (λ) dλ = and sup λ> Moreover, for p = we have that f (t) p dt (3.5) λµ f (λ) 1/p = sup t 1/p f (t). (3.6) t> ess sup f(x) = inf{λ : µ f (λ) = } = f (). x 2

23 Proof. First, let < p <. Then, since f is a Σ-measurable function, f p is a also a Σ-measurable function for all < p < and it follows therefore by Lemma 3.1 and a change of variable that f(x) p dµ = µ f p(λ) dλ = µ f (λ 1/p ) dλ = µ f (ν) d(ν p ) = p ν p 1 µ f (ν) dν. This gives the first equality in (3.5). The second equality follows directly from the equimeasurability of f p and (f ) p. To prove (3.6) we use Theorem 2.3(iv) and Lemma 3.1 to get that ( sup t> t 1/p f (t) = sup t> = sup λ> (t [f (t)] p ) 1/p = sup (t( f p ) (t)) 1/p = t> ( λµf (λ 1/p ) ) 1/p = sup u> sup λµ f p(λ) λ> (u p µ f (u)) 1/p = sup u [µ f (u)] 1/p. u> ) 1/p Now for the case when p =. By the definition of essential supremum and since the distribution function is positive we have that ess sup f(x) = inf{λ : µ({x : f(x) > λ}) = } = inf{λ : µ f (λ) = } x = inf{λ : µ f (λ) } = f (). This completes the proof of the theorem. 3.2 The Hardy-Littlewood inequality In this section we prove the important Hardy-Littlewood inequality. This inequality says that the integral over the product of two measurable functions always is less than or equal to the integral over the product of the decreasing rearrangement of the functions. To prove this we first need the following lemma, which is the Hardy-Littlewood inequality for the case when one of the functions is the characteristic function. Lemma 3.3. The inequality f(x) dµ holds for all A Σ such that µ(a) a. A 21 a f (t) dt.

24 Proof. If a = the inequality is true by Lemma 3.1. Assume that a <. Then by Lemma 3.1 and Theorem 2.3(v) we obtain f(x) dµ = f(x)χ A (x) dµ = (fχ A ) (t) dt A f (t)χ [,µ(a)) (t) dt = µ(a) f (t) dt a f (t) dt. Using the previous lemma we can now prove the Hardy-Littlewood inequality for simple functions and using the Monotone Convergence Theorem for measurable functions on. Theorem 3.4 (Hardy-Littlewood). The inequality f(x)g(x) dµ f (t)g (t) dt. (3.7) holds. Proof. We prove this inequality for positive simple functions and the general result will follow by Theorem 2.1(v), Theorem 2.2 and the Monotone Convergence Theorem for measurable functions on. Let s be a simple function of the form s(x) = α j χ Aj where α 1 > α 2 > > α k > and all A j Σ are pairwise disjoint. Then we can rewrite s as s(x) = β j χ Bj where B j = j i=1 A i and β j = α j α j+1, α k+1 =. By Example 2.2 and 22

25 Lemma 3.3 we get that ( ) s(x)g(x) dµ = β j χ Bj (x) g(x) dµ = = = µ(bj ) β j g (t) dt = γj α j γ j 1 g (t) dt = s (t)g(t) dt where γ j = j i=1 µ(a j) and γ =. β j g(x) dµ B j (α j α j+1 ) µ(bj ) g (t) dt α j χ [γj 1,γ j )(t)g (t) dt Example 3.1. Let = N, Σ = 2 N and µ({n}) = 1. Then, as in Example 2.5, any Σ-measurable function on is a sequence. Thus, for two sequences {a n } n=1 and {b n } n=1 the Hardy-Littlewood inequality becomes a n b n a nb n where {a n} and {b n} are the sequences { a n } and { b n } permutated in decreasing order. The Hardy-Littlewood inequality arises many interesting questions. One of importance is how to characterize functions which give equality in the estimate 3.7. Another question is if there is any way to prove estimate 3.7 for n functions instead of only two. We will devote the rest of the chapter answering to these two questions and it turns out that both depends on the following result. Theorem 3.5. Let µ be a nonatomic measure and let a be any number satisfying a µ(). (i) Then the equality holds. sup f(x) dµ = µ(a) a A 23 a f (t) dt

26 (ii) If µ() <, then there is a set A Σ such that µ(a) = a and a f(x) dµ = f (t) dt. A (iii) If µ() = and f (a) > lim t f (t), then there is a set A Σ such that µ(a) = a and a f(x) dµ = f (t) dt. A Proof. First we prove (ii) and (iii). (ii) We need to consider two separate cases, according to whether the number a does or does not lie in the range of the distribution function µ f of f. Suppose first that there exist an α > for which µ f (α) = a. In that case, it follows that f (a) = inf{λ : µ f (λ) = a} α and so i.e., µ f (f (a)) = a. Therefore, let Then a = µ f (α) µ f (f (a)) a, A = {x : f(x) > f (a)}. µ fχa (λ) = µ f (max(λ, f (a)) and by the equimeasurability of f and f m f χ [,a) (λ) min(m f (λ), a) = min(µ f (λ), a) = min(µ f (λ), µ f (f (a))) = µ f (max(λ, f (a))) Let ε > be arbitrary and take t = min(µ f (λ), a) ε. Then by Theorem 2.3(iii) that is, f (t )χ [,a) (t ) = f (t ) f (µ f (λ) ε) λ ε > λ, m f χ [,a) (λ) = m { t > : f (t)χ [,a) (t) > λ } t. Since ε was arbitrary, we have m f χ [,a) (λ) = µ f (max(λ, f (a))) = µ fχa (λ) 24

27 for all λ. Hence fχ A and f χ [,a) are equimeasurable which implies that their integrals are the same a f(x) dµ = f (t) dt. A It is also clear that the A increases with a. Next, we consider the case where a is not in the range of µ f. Since µ() < and µ f is decreasing it follows that lim µ f(λ) = a. λ Let λ = f (a) and suppose first that λ >. Since a > (a = is in the range of µ f (λ) for λ µ()) it follows that λ <. Also, since a is not in the range of µ f, we see that µ f (λ ) < a < µ f (λ) for < λ < λ. Thus, if a 1 denotes the left-hand limit µ f (λ ) (which exists because µ f is decreasing), then a µ f (λ ) < a µ f (λ ) = a 1. (3.8) This shows in fact that We claim that f (t) = λ for a t < a 1. (3.9) a 1 = µ({x : f(x) λ }). (3.1) To see this, we need only to express the set on the right as the intersection of a decreasing sequence of sets B n = {x : f(x) > λ 1 }, n = 1, 2,..., n observe that µ() is finite, and then µ({x : f(x) λ }) = lim n µ(b n ) = lim n µ f It follows from (3.8) and (3.1) that the set G = {x : f(x) = λ } ( λ 1 ) = µ f (λ ). n 25

28 has measure a 1 a. Since µ is nonatomic we may select a subset B of G with measure a a. The set A defined as the disjoint union A = {x : f(x) > λ } B has measure µ f (λ ) + (a a ) = a, as desired. Moreover, f(x) dµ = f(x) dµ + f(x) dµ. A { f >λ } Now, by (3.8), the value a does lie in the range of µ f. The first part of the proof shows therefore that the first of the integrals has the value a f (t) dt. Also, since f = λ on B, the second integral has the value λ µ(b) = λ (a a ) = a B a f (t) dt, by (3.9). Combining these results we get that a f(x) dµ = f (t) dt. A In the remaining case λ =, we obtain in place of (3.8) the equality µ({x : f(x) > }) a < a. In that case, we may choose B disjoint from the support of f, with measure µ(b) = a a. Defining A again as A = {x : f(x) > } B and observing that f (t) vanishes for t a, we obtain as before a a f(x) dµ = f (t) dt = f (t) dt. A We observed in the first part of the proof that the sets A increases with a in the range of µ f. For the other values of a it is possible to show that the 26

29 sets B in the second part of the proof may be chosen so they increase with a. This completes the proof of (ii). (iii) The condition f (a) > lim t f (t) f ( ) implies that in the neighborhood of f (a) we have µ f (λ) <, and therefore If we denote and then µ({x : f(x) f (a)}) = m({t > : f (t) f (a)}) a. G = {x : f(x) > f (a)} H = {x : f(x) f (a)}, µ(g) a µ(h). Since the measure µ is nonatomic we can find a set A Σ such that G A H and µ(a) = a. The functions fχ A and f χ [,a) are equimeasurable and (fχ A ) (t) = f (t)χ [,a) (t). Indeed, by Theorem 2.3(v) we have (fχ A ) (t) f (t)χ [,a) (t). Assume that t < a. Then fχ A fχ G > f (a), that is, (fχ A ) (t) f (a). If x / A, then f(x) f (a) (fχ A ) (t) so Thus and B = {x : f(x) > (fχ A ) (t)} = {x A : f(x) > (fχ A ) (t)} = {x : f(x)χ A (x) > (fχ A ) (t)}. µ(b) = µ f ((fχ A ) (t)) = µ fχa ((fχ A ) (t)) t f (t) f (µ f ((fχ A ) (t))) (fχ A ) (t). Equimeasurability of fχ A and f χ [,a) implies the equality of the integrals. This completes the proof of (ii). (i) From Lemma 3.3 we immediately obtain the inequality sup µ(a)=a A f(x) dµ a f (t) dt. Therefore, in the case where f (a) > f ( ) from (iii) we have assertion. If f (a) = f ( ), then for G = {x : f(x) > f (a)} 27

30 and H = {x : f(x) > f (a) ε a } where ε > is an arbitrary small number we have µ(g ) a µ(h ) =. Therefore there exists a set A(ε) = A such that G A H and µ(a) = a. The functions (fχ G ) and f χ [,µ(g )) are equimeasurable. Thus G f(x) dµ = µ(g ) µ(g ) and we have f(x) dµ = f(x) dµ + A G µ(g ) f (t) dt + = µ(g ) f (t) dt + A\G f (t) dt, f(x) dµ A\G ( f ( ) ε ) a ( f ( ) ε a dt ) (a µ(g )) f (t) dt + f ( ) (a µ(g )) ε. On the interval (µ(g ), a) the function f is equal to f ( ) so a f(x) dµ f (t) dt ε. Hence (i) holds. A If we apply the Hardy-Littlewood inequality to the Σ-measurable functions f, g and h we get that f(x)g(x)h(x) dµ (fg) (t)h (t) dt. Thus, it does not immediately hold for more than two functions and if we try to use Theorem 2.5 we see that f(x)g(x)h(x) dµ f (t/2)g (t/2)h (t) dt 28

31 which leads to the following unsatisfactory result f (t/2)g (t/2)h (t) dt f (t)g (t)h (t) dt. Hence we need a stronger result than Theorem 2.5 and it turns out that the next theorem combined with Hardy s lemma (Lemma 3.7) will give us what we want. The only drawback is that we have to restrict considerations to nonatomic measure spaces. Theorem 3.6. Let µ be a nonatomic measure. Then a for any a >. In particular, (fg) (t) dt (fg) (t) dt a f (t)g (t) dt. f (t)g (t) dt. Proof. We begin with the case when a =. Then, by Lemma 3.1 and the Hardy-Littlewood inequality (Theorem 3.4) we immediately get that (fg) (t) dt = f(x)g(x) dµ f (t)g (t) dt. Now, if a is any positive number, we have by Theorem 3.5(i) that a (fg) (t) dt = f(x)g(x) dµ = f(x)g(x)χ A (x) dµ sup µ(a)=a A sup µ(a)=a Since we integrate f(x)g(x)χ A (x) over we can apply the Hardy-Littlewood inequality (Theorem 3.4) and Theorem 2.3(v) to get that f(x)g(x)χ A (x) dµ = sup f (t)(gχ A ) (t) dt sup µ(a)=a This completes the proof. = 29 µ(a)=a a sup µ(a)=a a f (t)g (t) dt f (t)g (t) dt.

32 Lemma 3.7 (Hardy s Lemma). Let f and g be positive Lebesgue measurable functions on (, ) and suppose that a f(t) dt a g(t) dt for all a >. Let h be any positive decreasing function on (, ). Then h(t)f(t) dt h(t)g(t) dt. Proof. We first prove this result for a simple functions. Let s be a simple positive decreasing function on (, ). Then s can be represented as s(t) = α j χ [,aj )(t) where α 1 > α 2 > α k > and a 1 > a 2 > > a k >. Therefore, s(t)f(t) dt = aj α j f(t) dt aj α j g(t) dt = s(t)g(t) dt and the general results follows by the Monotone Convergence Theorem. We can now combine Theorem 3.6 and Lemma 3.7 to generalize the Hardy- Littlewood inequality. Theorem 3.8. Let µ be nonatomic and let f, g and h be Σ-measurable functions on. Then f(x)g(x)h(x) dµ f (t)g (t)h (t) dt Proof. By Theorem 3.4, Theorem 3.6 and Lemma 3.7 with h(t) = f (t) we immediately get that f(x)g(x)h(x) dµ f (t)(gh) (t) dt f (t)g (t)h (t) dt 3

33 Corollary 3.9. Let µ be nonatomic and f 1, f 2,..., f n Σ-measurable functions on, n N. Then n f k (x) dµ k=1 n fk (t) dt. Proof. This follows from Theorem 3.8 and the principle of induction. We will now take a closer look on the question if there is a Σ-measurable function h equimeasurable with g such that f(x)h(x) dµ = f (t)g (t) dt The answer is in general no, and it depends if the measure space is nonatomic or completely atomic where all atoms are of equal measure and if the measure space is finite. To make the terminology less cumbersome we first make a definition which is taken from Bennett and Sharpley [1]. Definition 3.1. A measure space (, Σ, µ) or a measure µ is said to be resonant if it is σ-finite and either nonatomic or a countable union of atoms of equal measure. Theorem 3.1. If µ is resonant, then sup f(x) g(x) dµ = k=1 f (t)g (t) dt (3.11) where the supremum is taken over all Σ-measurable functions g equimeasurable with g, that is, µ g (λ) = µ g (λ), λ. Moreover, if also µ() <, then there is a Σ-measurable function g such that f(x) g(x) dµ = f (t)g (t) dt. (3.12) Proof. We can clearly assume that both f and g are positive. Let µ() <. If µ is completely atomic where all atoms are of equal measure, then g can be constructed by permutating the atoms. More specific, this permutation is the composition of the inverse of the permutation that takes f to f and the permutation that takes g to g. That is, g is g, but where all atoms are permutated so that the atom where g has its largest value is the same as the 31

34 atom where f has its largest value, the atom where g has its second largest value is the same as the atom where f has its second largest value, and so on. Clearly, g and g are equimeasurable which means that we have proved (3.12) for the case when µ is completely atomic where all atoms are of equal measure. Now let µ be nonatomic. Since g is positive we can find a sequence of simple positive functions {g n }, such that g n g. Then, for any arbitrary fixed integer n 1, we can represent g n as g n (x) = δ j χ Dj (x) where D 1 D 2 D k and δ j >, j = 1, 2,..., k. Since µ() < we can apply Theorem 3.5(ii) for each number µ(d j ) and get a sequence of sets E 1, E 2,..., E k such that µ(e j ) = µ(d j ), j = 1, 2,..., k and µ(dj ) f(x) dµ = f (t) dt (3.13) E j Define the simple function g n : [, ) by g n (x) = δ j χ Ek (x). Since the measures of E j and D j are equal for all j = 1, 2,..., k, the decreasing rearrangement of g n and g n are equal, that is gn(t) = g n(t) = δ j χ [,µ(ej )(t). Totally get a sequence of positive simple functions { g n }, such that g n and g n are equimeasurable, n = 1, 2,.... Using Theorem 2.1(v) it also follows that g = lim n g n is equimeasurable with g. Moreover, by (3.13), µ(dj ) f(x) g n (x) dµ = δ j f(x) dµ = δ j f E (t) dt j = f (t)δ j χ [,µ(dj ))(t) dt = 32 f (t)g n(t) dt.

35 Thus, we have proved (3.12) for g n, n = 1, 2,... and by Theorem 2.1(v), Theorem 2.2, the Monotone Convergence Theorem and the fact that g and g are equimeasurable, the general case follows. Now, let µ() =, µ be nonatomic or completely atomic where all atoms are of equal measure and α > be a real number such that α < f (t)g (t) dt. Since f and g are assumed to be positive it will suffice to show that there exists a positive function g equimeasurable with g such that α < f(x) g(x) dµ. Because µ is a σ-finite measure we can find a sequence of pairwise disjoint sets 1, 2,..., such that µ( n ) <, n = 1, 2,... and = n=1 n. Hence, we can therefore find two sequences of positive functions {f n } and {g n } such that f n f, g n g and both f n and g n has support in n, n = 1, 2,.... Since f n(t)g n(t) dt f (t)g (t) dt by Theorem 2.1(v), Theorem 2.2 and the Monotone Convergence Theorem it follows that there exist an integer N 1 such that α < f N(t)g N(t) dt. Then ( N, Σ, µ) is a finite measure space which is either nonatomic or completely atomic with all atoms of equal measure. We can therefore apply the first part of the proof and find a positive function ĥ on N equimeasurable with gχ N such that f(x)ĥ(x) dµ = (fχ N ) (t)(gχ N ) (t) dt. N Since f N fχ N and g N gχ N by the construction of {f n } and {g n } it follows that α < = N fn(t)g N(t) dt f(x)ĥ(x) dµ = 33 (fχ N ) (t)(gχ N ) (t) dt f(x)h(x) dµ

36 where Thus, if we take h(x) = {ĥ(x) if x N, if x / N. g(x) = h(x)χ N (x) + g(x)χ \N (x), x, then g is equimeasurable with g and h g. Hence α < f(x)h(x) dµ f(x) g(x) dµ which completes the proof. 3.3 Operation ** If we apply the Hardy-Littlewood inequality to g = χ A, where A is any measurable set with measure t we get that µ(a) t f(x)g(x) dµ = f(x) dµ f (s) ds = f (s) ds. If we divide both sides with µ(a) = t we get 1 f(x) dµ 1 µ(a) t A A t f (s) ds. The right hand side is a function which is defined for all < t. This function is of great importance for the Lorentz L(p, q) spaces and its application will be clear after Chapter 4. A formal definition is as follows. Definition 3.2. The function f : (, ) [, ] is defined as f (t) = 1 t t f (s) ds. Remark. f is sometimes called the maximal function of f since it is the largest of all average values over f. Even if the value of f (t) at t = is not included in the definition above, the 34

37 limit as t approaches zero from the right is defined for all rearrangements. In fact, lim f (t) = f () = ess sup f(x), t + where the last equality is from Theorem 3.2. Theorem Let f n be Σ-measurable functions on, n = 1, 2,.... Then (i) f is decreasing and continuous on (, ). (ii) f (t) f (t) for all t >. (iii) If f(x) g(x) for x µ-a.e., then f (t) g (t) for all t >. (iv) If {f n } is a sequence such that f n (x) f(x) for x µ-a.e., n = 1, 2,... and lim n f n (x) = f(x) for x µ-a.e., then fn (t) f (t) and lim n fn (t) = f (t) for all t >. Proof. (i) The continuity of f (t) follow directly from the continuity of the integral so the only thing left to prove is that f (t) is decreasing. Let < t 1 < t 2 be arbitrary. Then f (t 2 ) = 1 t2 f (s) dt = 1 t1 f (s) ds + 1 t2 f (s) ds t 2 t 2 t 2 x t 1 1 t1 f (s) ds + 1 f (t 1 )(t 2 t 1 ) t 2 t 2 = 1 t1 ( 1 f (s) ds + 1 ) t 1 f (t 1 ) t 2 t 1 t 2 1 t2 ( 1 f (s) ds + 1 ) t1 f (s) ds t 2 t 1 t 2 = 1 t1 f (s) ds = f (t 1 ). t 1 (ii) Since f is a decreasing function we get that f (t) = 1 t t f (s) ds 1 t t f (t) ds = f (t). 35

38 (iii) By Theorem 2.1(ii) and Theorem 2.2 we have that f (t) g (t) for all t. Moreover, since f and g are positive functions we get for t > that f (t) = 1 t t f (s) ds 1 t t g (s) ds = g (t). (iv) We know by Theorem 2.1(ii) that f n(t) f (t), n = 1, 2,... and by Theorem 2.1(v) that lim n f n(t) = f (t) for t. Since f (t) = 1 t t f (s) ds t > and since fn (t) f (t) by (iii) we get by the Monotone Convergence Theorem that fn (t) f (t) for all t >. Example 3.2. Let A be any measurable set and let f = χ A. Then, clearly, f (t) = χ [,µ(a)) so using the definition of f we get that f (t) = 1 t t ( f (s) ds = min 1, µ(a) ) t for t >. A graph of f can be found in Figure 3.1. f 1 µ(a) t Figure 3.1: Graph of f, where f(x) = χ A (x). Example 3.3. Let s be the simple function in the following form s(x) = α j χ Aj (x) 36

39 where α 1 > α 2 > > α k > and A j = {x : s(x) = α j }. Then, by Example 2.2, we have that s (t) = α j χ [βj 1,β j )(t), where β j = j i=1 µ(a j). Using the definition of s (t) we get that s (t) = 1 t t s (u) du = 1 t t α j χ [βj 1,β j )(t) for all t >. We consider two cases. First, if β m 1 < t β m for m = 1, 2,..., k, then s (t) = 1 t = 1 t t m 1 α j χ [βj 1,β j )(t) and the second one, if t > β k then α j µ(a j ) + α m t ( t m 1 µ(a j ) ) s (t) = 1 t α j µ(a j ), Hence, s (t) = 1 t 1 t [ m 1 α jµ(a j ) + α m (t ] m 1 µ(a j)) [ k ] α jµ(a j ) if β m 1 < t β m for m = 1, 2,..., k, if t > β k Figure 3.3 shows a graph of s when s is a simple function with k = 4. 37

40 α 1 α 2 α 3 α 4 t β 1 β 2 β 3 β 4 Figure 3.2: s and s where s is a simple function. Example 3.4. Let = [, ), Σ = all Lebesgue-measurable subsets of and m the Lebesgue measure. Define f : [, ) as { 1 (x 1) 2 if x 2, f(x) = if x > 2. From Example 2.3 we get for all t that { f 1 t 2 /4 if t 2, (t) = if t > 2, and we can obtain f by calculating the integral in Definition 3.2 { f 1 t 2 /12 if < t 2, (t) = 4/(3t) if t > 2, which can be seen in Figure

41 f Figure 3.3: The function f and f (dotted) from Example 3.4. t The last step before we introduce the Lorentz spaces is to derive the subadditivity of **. Theorem 3.12 (Subadditivity). If µ is a resonant measure, then (f + g) (t) f (t) + g (t) for all t >. Hence, operation ** is subadditive. Proof. If we apply Theorem 3.5 to the definition of f we see that f (t) = 1 t f (s) dt = 1 sup f(x) dµ. t t µ(a)=t A Hence, by the triangle inequality and the subadditivity of supremum (f + g) (t) = 1 sup f(x) + g(x) dµ t µ(a)=t A 1 ( ) sup f(x) dµ + g(x) dµ t µ(a)=t A A 1 f(x) dµ + 1 g(x) dµ t t sup µ(a)=t A = f (t) + g (t). sup µ(a)=t A 39

42 Chapter 4 Lorentz L(p, q) spaces In this chapter we introduce Lorentz L(p, q) spaces and prove their most important properties. The first section defines the general L(p, q) spaces and the Lorentz sequence space l(p, q). We show that these two definitions are almost equivalent. After that the linearity of L(p, q) is established we show that to introduce a norm on L(p, q) is not a trivial task. We end the first section with some relations between different Lorentz spaces. The second section deals with the problem of finding a suitable norm for L(p, q) and we show that a certain functional can be seen as a norm for some values of p and q. We also prove when L(p, q) is normable and not normable. The last section in this chapter is devoted to the topological properties of the Lorentz L(p, q) spaces. We prove completeness, density of simple functions and separability, and then we discuss the nature of the space of all bounded linear functionals on L(p, q), that is, the dual space of L(p, q). 4.1 Definition of the Lorentz L(p, q) spaces We start this section with a definition of the Lorentz L(p, q) space and the Lorentz sequence space l(p, q). Definition 4.1. Let (, Σ, µ) be a σ-finite measure space and < p, < q. Then the Lorentz L(p, q) space is the set of all classes of Σ- 4

43 measurable functions f such that the functional f pq <, where ( ( t 1/p f (t) ) ) 1/q q dt if < p <, < q <, t f pq = sup t 1/p f (t) if < p, q =. t> The Lorentz sequence space, denoted by l(p, q), is the set of all sequences a = {a n } c such that the functional a pq <, where ( ) ( ) 1/q n 1/p a q n n 1 if < p, < q <, a pq = n=1 sup n 1/p a n if < p, q =. n 1 and a = {a n} is the sequence { a n } permutated in a decreasing order. Remark. For the Lorentz L(p, q) space, the case when p = and < q < is not of any interest. The reason for this is that f q < implies f = µ-a.e on. This can be seen by using the following argument: Assume that L(, q) is a non-trivial space. Then there exists a nonzero function f L(, q) which means that there exists c > and a set A Σ of positive measure such that f(x) > c for all x A. Then f q q = f (t) q dt t (fχ A ) (t) dt t µ(a) c q dt t = Hence, we have a contradiction which implies that the only element in L(, q) is the zero function. The Lorentz sequence space l(p, q) and L(p, q), when = N, Σ = 2 N and µ({n}) = 1 are equivalent for < p <, < q <. In fact, if we let a n = f (t) for n 1 t < n as in Example 2.5 we get ( f pq = ( t 1/p f (t) ) ) ( 1/q q dt 1/q n = a q n t dt) q/p 1, t n=1 n 1 and since ( ) q/p 1 n n q/p 1 t q/p 1 dt 2n q/p 1 2 n 1 41

44 it follows that the two definitions in Definition 4.1 are equivalent and all results that hold for general L(p, q) will hold also for l(p, q) Observe that the space l(p, q) is not empty when p =. For example, all sequences which only have a finite number of nonzero elements are in l(, q) for all < q. This shows that there is a fundamental difference between L(, q) and l(, q). Lorentz L(p, q) spaces can be seen as generalizations of the ordinary L p spaces. The reason for this is that if we take q = p we obtain L(p, p) = L p for < p. In fact, for < p < we get by the definition of pq that ( p ( f pp = t 1/p f (t) ) ) 1/p p dt/t p ( ) 1/p = t (f (t)) p dt/t ( 1/p = f (t) dt) p. Using Theorem 3.2 we then obtain ( 1/p ( 1/p f pp = f (t) dt) p = f(x) dµ) p. For p = we get immediately, by the fact that f is decreasing and by Theorem 3.2, that f = sup t> f (t) = f () = ess sup f(x) = f x Hence, f pp = f p, which implies that L(p, p) = L p with equality of norms. Example 4.1. Let A be any Σ-measurable set of finite measure. Then ( ) 1/q p χ A pq = µ(a) 1/p q for < p < and < q <. If < p < and q = we get χ A p = µ(a) 1/p. 42

45 The next theorem consider the linearity of L(p, q) and the nature of the functional pq. It turns out that it can be seen as a quasi-norm. Theorem 4.1. The Lorentz L(p, q) space is a linear space and the functional pq is a quasi-norm 1. Proof. To prove that L(p, q) is a linear space we need to show that f + g L(p, q) and αf L(p, q) for any f, g L(p, q) and any scalar α. Let therefore f and g be two arbitrary elements in L(p, q). We start with the case when < p < and < q < and get by Theorem 2.5 that [ ( f + g pq = t 1/p (f + g) (t) ) ] 1/q q dt/t [ ( t 1/p [f (t/2) + g (t/2)] ) ] 1/q q dt/t [ ( = (2u) 1/p [f (u) + g (u)] ) ] 1/q q du/u [ ( [ 2 1/p 2 max u 1/p f (u), u 1/p g (u) ]) ] 1/q q du/u 2 1/p [ = 2 1/p+1 ( f q pq + g q pq ( [u 2 q 1/p f (u) ] q [ + u 1/p g (u) ] ) ] 1/q q du/u ) 1/q 2 1/p+1 [ 2 max( f q pq, g q pq) ] 1/q 2 1/p+1/q+1 ( f pq + g pq ). Moreover, for any scalar α we clearly have that αf pq = α f pq. Now, let < p and q =. Then, again by Theorem 2.5, we obtain 1 A quasi-norm is a functional that is like a norm except that it does only satisfy the triangle inequality with a constant C 1, that is, f + g C( f + g ) where C 1. 43

46 that f + g p = sup t> t 1/p (f + g) (t) sup t 1/p (f (t/2) + g (t/2)) t> = sup(2u) 1/p (f (u) + g (u)) u> 2 1/p (sup u> = 2 1/p ( f p + g p ). Clearly, we also have in this case that u 1/p f (u) + sup u 1/p g (u) u> ) αf pq = α f p Hence, L(p, q) is a linear space for all < p and < q <. Moreover, from our proof above follows that the functional pq is a quasi-norm. From now we will assume that the measure space is resonant. The reason for this is that resonant spaces are simpler than the general ones. The next theorem strengthen the previous result. It states that pq in some cases on p and q is even a norm. This result depends on monotonicity of the function t q/p 1. We prove that if t q/p 1 is a decreasing function then pq is a norm. Theorem 4.2. The functional pq is a norm if and only if either 1 q p or p = q =. Proof. To prove that pq is a norm we need only to show that the triangle inequality is satisfied. The rest follows from the last part of Theorem 4.1. First assume that 1 q p. Then we have q/p 1 which implies that the continuous function t q/p 1 is decreasing and hence (t q/p 1 ) = t q/p 1. We can therefore, since the measure space is resonant, use Theorem 3.1 for any 44

47 f and g in L(p, q) and obtain that [ f + g pq = (t 1/p (f + g) (t)) q dt/t [ = = [ sup [ = sup ] 1/q ] 1/q t q/p 1 (f + g) (t) q dt ] 1/q f(x) + g(x) q h(x) dµ ] f(x) h(x) 1/q + g(x) h(x) 1/q q 1/q dµ where the supremum is taken over all Σ-measurable functions h equimeasurable with t q/p 1. We can now use the Minkowski inequality and the subadditivity of the supremum to get [ ] f + g pq = sup f(x) h(x) 1/q + g(x) h(x) 1/q q 1/q dµ ( [ 1/q [ ] ) 1/q sup f(x) q h(x) dµ] + g(x) q h(x) dµ [ 1/q sup f(x) q h(x) dµ] + sup [ 1/q g(x) q h(x) dµ]. Finally, we use Theorem 3.1 again to get back to the old representation [ 1/q [ ] 1/q f + g pq sup f(x) q h(x) dµ] + sup g(x) q h(x) dµ [ 1/q [ ] 1/q = t q/p 1 f (t) dt] q + t q/p 1 g (t) q dt = f pq + g pq. To complete the proof we need to show that pq is not a norm for the remaining cases, that is, the triangle inequality doesn t hold. We start with the case when µ is nonatomic and we split it into the following parts. 1 1 p < q 2 < p < 1, < q 45

48 3 < q < 1 p < 1. We start with the case when 1 p < q <. Take and f(x) = (1 + ε)χ [,a+h) (x) + χ [a+h,a+2h) (x) g(x) = χ [,h) (x) + (1 + ε)χ [h,+a+2h) (x), where a, h, ε >. Clearly, f = g = f and since (f + g) (t) = (2 + 2ε)χ [,a) (t) + (2 + ε)χ [a,a+2h) (x) it follows that we can evaluate the norms of f, g and f + g by f q pq = g q pq = p [ (1 + ε) q (a + h) q/p + (a + 2h) q/p (a + h) q/p] q and f + g q pq = p q [ (2 + 2ε) q a q/p + (2 + ε) q ( (a + 2h) q/p a q/p)], respectively. Assume that the triangle inequality holds, that is, Then the inequality f + g pq f pq + g pq. (2 + 2ε) q a q/p + (2 + ε) q ( (a + 2h q/p a q/p) 2 q ( (1 + ε) q (a + h) q/p + (a + 2h) q/p (a + h) q/p) holds and it can be written as (a + 2h) q/p (a + h) q/p (1 + ε)q (1 + ε/2) q (1 + ε/2) q 1 Taking ε + we obtain that If we define a function f as ( (a + h) q/p a q/p). (a + 2h) q/p + a q/p 2(a + h) q/p. (4.1) f(x) = x 46 t q/p 1 dt,

49 then we can rewrite inequality (4.1) as f(a + 2h) + f(a) 2f(a + h) which implies, together with the fact that f is continuous, that it is a concave function. By the concavity of f, the derivative f (x) = x q/p 1 must be a decreasing function, that is, it must be that q p. This contradicts q > p, and we conclude that the triangle does not hold. For q =, take measurable sets A B such that 1 < and µ(b \ A) µ(a). If we let and then and Thus, ( ) 1/p µ(b) a µ(a) f(x) = aχ A (x) + χ B\A (x) g(x) = χ A + aχ B\A, f (t) = aχ [,µ(a)) + χ [µ(a),µ(b)) g (t) = aχ [,µ(b\a)) (t) + χ [µ(b\a),µ(b)) (t). f p, = max ( aµ(a) 1/p, µ(b) 1/p) = µ(b) 1/p, g p, = max ( aµ(b \ A) 1/p, µ(b) 1/p) = µ(b) 1/p and since f(x) + g(x) = (a + 1)χ B, Then f + g p = (a + 1)µ(B) 1/p. f + g p = (a + 1)µ(B) 1/p > 2µ(B) 1/p = f p + g p, which shows that the triangle inequality does not hold. Hence, the proof of the first case is complete. 2. Let < p, q < and A, B be two measurable sets such that A B =. Then ( ) 1/q p χ A + χ B pq = (µ(a) + µ(b)) 1/p q 47

MATHS 730 FC Lecture Notes March 5, Introduction

MATHS 730 FC Lecture Notes March 5, Introduction 1 INTRODUCTION MATHS 730 FC Lecture Notes March 5, 2014 1 Introduction Definition. If A, B are sets and there exists a bijection A B, they have the same cardinality, which we write as A, #A. If there exists

More information

(U) =, if 0 U, 1 U, (U) = X, if 0 U, and 1 U. (U) = E, if 0 U, but 1 U. (U) = X \ E if 0 U, but 1 U. n=1 A n, then A M.

(U) =, if 0 U, 1 U, (U) = X, if 0 U, and 1 U. (U) = E, if 0 U, but 1 U. (U) = X \ E if 0 U, but 1 U. n=1 A n, then A M. 1. Abstract Integration The main reference for this section is Rudin s Real and Complex Analysis. The purpose of developing an abstract theory of integration is to emphasize the difference between the

More information

Chapter 4. Measure Theory. 1. Measure Spaces

Chapter 4. Measure Theory. 1. Measure Spaces Chapter 4. Measure Theory 1. Measure Spaces Let X be a nonempty set. A collection S of subsets of X is said to be an algebra on X if S has the following properties: 1. X S; 2. if A S, then A c S; 3. if

More information

Chapter 8. General Countably Additive Set Functions. 8.1 Hahn Decomposition Theorem

Chapter 8. General Countably Additive Set Functions. 8.1 Hahn Decomposition Theorem Chapter 8 General Countably dditive Set Functions In Theorem 5.2.2 the reader saw that if f : X R is integrable on the measure space (X,, µ) then we can define a countably additive set function ν on by

More information

Real Analysis Notes. Thomas Goller

Real Analysis Notes. Thomas Goller Real Analysis Notes Thomas Goller September 4, 2011 Contents 1 Abstract Measure Spaces 2 1.1 Basic Definitions........................... 2 1.2 Measurable Functions........................ 2 1.3 Integration..............................

More information

MATH MEASURE THEORY AND FOURIER ANALYSIS. Contents

MATH MEASURE THEORY AND FOURIER ANALYSIS. Contents MATH 3969 - MEASURE THEORY AND FOURIER ANALYSIS ANDREW TULLOCH Contents 1. Measure Theory 2 1.1. Properties of Measures 3 1.2. Constructing σ-algebras and measures 3 1.3. Properties of the Lebesgue measure

More information

Chapter 5. Measurable Functions

Chapter 5. Measurable Functions Chapter 5. Measurable Functions 1. Measurable Functions Let X be a nonempty set, and let S be a σ-algebra of subsets of X. Then (X, S) is a measurable space. A subset E of X is said to be measurable if

More information

Integration on Measure Spaces

Integration on Measure Spaces Chapter 3 Integration on Measure Spaces In this chapter we introduce the general notion of a measure on a space X, define the class of measurable functions, and define the integral, first on a class of

More information

Examples of Dual Spaces from Measure Theory

Examples of Dual Spaces from Measure Theory Chapter 9 Examples of Dual Spaces from Measure Theory We have seen that L (, A, µ) is a Banach space for any measure space (, A, µ). We will extend that concept in the following section to identify an

More information

5 Measure theory II. (or. lim. Prove the proposition. 5. For fixed F A and φ M define the restriction of φ on F by writing.

5 Measure theory II. (or. lim. Prove the proposition. 5. For fixed F A and φ M define the restriction of φ on F by writing. 5 Measure theory II 1. Charges (signed measures). Let (Ω, A) be a σ -algebra. A map φ: A R is called a charge, (or signed measure or σ -additive set function) if φ = φ(a j ) (5.1) A j for any disjoint

More information

CHAPTER 6. Differentiation

CHAPTER 6. Differentiation CHPTER 6 Differentiation The generalization from elementary calculus of differentiation in measure theory is less obvious than that of integration, and the methods of treating it are somewhat involved.

More information

Defining the Integral

Defining the Integral Defining the Integral In these notes we provide a careful definition of the Lebesgue integral and we prove each of the three main convergence theorems. For the duration of these notes, let (, M, µ) be

More information

Chapter 6. Integration. 1. Integrals of Nonnegative Functions. a j µ(e j ) (ca j )µ(e j ) = c X. and ψ =

Chapter 6. Integration. 1. Integrals of Nonnegative Functions. a j µ(e j ) (ca j )µ(e j ) = c X. and ψ = Chapter 6. Integration 1. Integrals of Nonnegative Functions Let (, S, µ) be a measure space. We denote by L + the set of all measurable functions from to [0, ]. Let φ be a simple function in L +. Suppose

More information

l(y j ) = 0 for all y j (1)

l(y j ) = 0 for all y j (1) Problem 1. The closed linear span of a subset {y j } of a normed vector space is defined as the intersection of all closed subspaces containing all y j and thus the smallest such subspace. 1 Show that

More information

THEOREMS, ETC., FOR MATH 515

THEOREMS, ETC., FOR MATH 515 THEOREMS, ETC., FOR MATH 515 Proposition 1 (=comment on page 17). If A is an algebra, then any finite union or finite intersection of sets in A is also in A. Proposition 2 (=Proposition 1.1). For every

More information

REAL AND COMPLEX ANALYSIS

REAL AND COMPLEX ANALYSIS REAL AND COMPLE ANALYSIS Third Edition Walter Rudin Professor of Mathematics University of Wisconsin, Madison Version 1.1 No rights reserved. Any part of this work can be reproduced or transmitted in any

More information

MATH6081A Homework 8. In addition, when 1 < p 2 the above inequality can be refined using Lorentz spaces: f

MATH6081A Homework 8. In addition, when 1 < p 2 the above inequality can be refined using Lorentz spaces: f MATH68A Homework 8. Prove the Hausdorff-Young inequality, namely f f L L p p for all f L p (R n and all p 2. In addition, when < p 2 the above inequality can be refined using Lorentz spaces: f L p,p f

More information

HILBERT SPACES AND THE RADON-NIKODYM THEOREM. where the bar in the first equation denotes complex conjugation. In either case, for any x V define

HILBERT SPACES AND THE RADON-NIKODYM THEOREM. where the bar in the first equation denotes complex conjugation. In either case, for any x V define HILBERT SPACES AND THE RADON-NIKODYM THEOREM STEVEN P. LALLEY 1. DEFINITIONS Definition 1. A real inner product space is a real vector space V together with a symmetric, bilinear, positive-definite mapping,

More information

1/12/05: sec 3.1 and my article: How good is the Lebesgue measure?, Math. Intelligencer 11(2) (1989),

1/12/05: sec 3.1 and my article: How good is the Lebesgue measure?, Math. Intelligencer 11(2) (1989), Real Analysis 2, Math 651, Spring 2005 April 26, 2005 1 Real Analysis 2, Math 651, Spring 2005 Krzysztof Chris Ciesielski 1/12/05: sec 3.1 and my article: How good is the Lebesgue measure?, Math. Intelligencer

More information

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure? MA 645-4A (Real Analysis), Dr. Chernov Homework assignment 1 (Due ). Show that the open disk x 2 + y 2 < 1 is a countable union of planar elementary sets. Show that the closed disk x 2 + y 2 1 is a countable

More information

MATH41011/MATH61011: FOURIER SERIES AND LEBESGUE INTEGRATION. Extra Reading Material for Level 4 and Level 6

MATH41011/MATH61011: FOURIER SERIES AND LEBESGUE INTEGRATION. Extra Reading Material for Level 4 and Level 6 MATH41011/MATH61011: FOURIER SERIES AND LEBESGUE INTEGRATION Extra Reading Material for Level 4 and Level 6 Part A: Construction of Lebesgue Measure The first part the extra material consists of the construction

More information

Functional Analysis Winter 2018/2019

Functional Analysis Winter 2018/2019 Functional Analysis Winter 2018/2019 Peer Christian Kunstmann Karlsruher Institut für Technologie (KIT) Institut für Analysis Englerstr. 2, 76131 Karlsruhe e-mail: peer.kunstmann@kit.edu These lecture

More information

Measure Theory on Topological Spaces. Course: Prof. Tony Dorlas 2010 Typset: Cathal Ormond

Measure Theory on Topological Spaces. Course: Prof. Tony Dorlas 2010 Typset: Cathal Ormond Measure Theory on Topological Spaces Course: Prof. Tony Dorlas 2010 Typset: Cathal Ormond May 22, 2011 Contents 1 Introduction 2 1.1 The Riemann Integral........................................ 2 1.2 Measurable..............................................

More information

(1) Consider the space S consisting of all continuous real-valued functions on the closed interval [0, 1]. For f, g S, define

(1) Consider the space S consisting of all continuous real-valued functions on the closed interval [0, 1]. For f, g S, define Homework, Real Analysis I, Fall, 2010. (1) Consider the space S consisting of all continuous real-valued functions on the closed interval [0, 1]. For f, g S, define ρ(f, g) = 1 0 f(x) g(x) dx. Show that

More information

2 (Bonus). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

2 (Bonus). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure? MA 645-4A (Real Analysis), Dr. Chernov Homework assignment 1 (Due 9/5). Prove that every countable set A is measurable and µ(a) = 0. 2 (Bonus). Let A consist of points (x, y) such that either x or y is

More information

Lebesgue measure and integration

Lebesgue measure and integration Chapter 4 Lebesgue measure and integration If you look back at what you have learned in your earlier mathematics courses, you will definitely recall a lot about area and volume from the simple formulas

More information

Summary of Real Analysis by Royden

Summary of Real Analysis by Royden Summary of Real Analysis by Royden Dan Hathaway May 2010 This document is a summary of the theorems and definitions and theorems from Part 1 of the book Real Analysis by Royden. In some areas, such as

More information

CHAPTER I THE RIESZ REPRESENTATION THEOREM

CHAPTER I THE RIESZ REPRESENTATION THEOREM CHAPTER I THE RIESZ REPRESENTATION THEOREM We begin our study by identifying certain special kinds of linear functionals on certain special vector spaces of functions. We describe these linear functionals

More information

Mathematical Methods for Physics and Engineering

Mathematical Methods for Physics and Engineering Mathematical Methods for Physics and Engineering Lecture notes for PDEs Sergei V. Shabanov Department of Mathematics, University of Florida, Gainesville, FL 32611 USA CHAPTER 1 The integration theory

More information

Measure and integration

Measure and integration Chapter 5 Measure and integration In calculus you have learned how to calculate the size of different kinds of sets: the length of a curve, the area of a region or a surface, the volume or mass of a solid.

More information

Notes on the Lebesgue Integral by Francis J. Narcowich November, 2013

Notes on the Lebesgue Integral by Francis J. Narcowich November, 2013 Notes on the Lebesgue Integral by Francis J. Narcowich November, 203 Introduction In the definition of the Riemann integral of a function f(x), the x-axis is partitioned and the integral is defined in

More information

Herz (cf. [H], and also [BS]) proved that the reverse inequality is also true, that is,

Herz (cf. [H], and also [BS]) proved that the reverse inequality is also true, that is, REARRANGEMENT OF HARDY-LITTLEWOOD MAXIMAL FUNCTIONS IN LORENTZ SPACES. Jesús Bastero*, Mario Milman and Francisco J. Ruiz** Abstract. For the classical Hardy-Littlewood maximal function M f, a well known

More information

ON A MAXIMAL OPERATOR IN REARRANGEMENT INVARIANT BANACH FUNCTION SPACES ON METRIC SPACES

ON A MAXIMAL OPERATOR IN REARRANGEMENT INVARIANT BANACH FUNCTION SPACES ON METRIC SPACES Vasile Alecsandri University of Bacău Faculty of Sciences Scientific Studies and Research Series Mathematics and Informatics Vol. 27207), No., 49-60 ON A MAXIMAL OPRATOR IN RARRANGMNT INVARIANT BANACH

More information

Math 4121 Spring 2012 Weaver. Measure Theory. 1. σ-algebras

Math 4121 Spring 2012 Weaver. Measure Theory. 1. σ-algebras Math 4121 Spring 2012 Weaver Measure Theory 1. σ-algebras A measure is a function which gauges the size of subsets of a given set. In general we do not ask that a measure evaluate the size of every subset,

More information

Analysis Comprehensive Exam Questions Fall 2008

Analysis Comprehensive Exam Questions Fall 2008 Analysis Comprehensive xam Questions Fall 28. (a) Let R be measurable with finite Lebesgue measure. Suppose that {f n } n N is a bounded sequence in L 2 () and there exists a function f such that f n (x)

More information

NAME: Mathematics 205A, Fall 2008, Final Examination. Answer Key

NAME: Mathematics 205A, Fall 2008, Final Examination. Answer Key NAME: Mathematics 205A, Fall 2008, Final Examination Answer Key 1 1. [25 points] Let X be a set with 2 or more elements. Show that there are topologies U and V on X such that the identity map J : (X, U)

More information

An introduction to some aspects of functional analysis

An introduction to some aspects of functional analysis An introduction to some aspects of functional analysis Stephen Semmes Rice University Abstract These informal notes deal with some very basic objects in functional analysis, including norms and seminorms

More information

ABSTRACT INTEGRATION CHAPTER ONE

ABSTRACT INTEGRATION CHAPTER ONE CHAPTER ONE ABSTRACT INTEGRATION Version 1.1 No rights reserved. Any part of this work can be reproduced or transmitted in any form or by any means. Suggestions and errors are invited and can be mailed

More information

1.1. MEASURES AND INTEGRALS

1.1. MEASURES AND INTEGRALS CHAPTER 1: MEASURE THEORY In this chapter we define the notion of measure µ on a space, construct integrals on this space, and establish their basic properties under limits. The measure µ(e) will be defined

More information

Lebesgue measure on R is just one of many important measures in mathematics. In these notes we introduce the general framework for measures.

Lebesgue measure on R is just one of many important measures in mathematics. In these notes we introduce the general framework for measures. Measures In General Lebesgue measure on R is just one of many important measures in mathematics. In these notes we introduce the general framework for measures. Definition: σ-algebra Let X be a set. A

More information

On maxitive integration

On maxitive integration On maxitive integration Marco E. G. V. Cattaneo Department of Mathematics, University of Hull m.cattaneo@hull.ac.uk Abstract A functional is said to be maxitive if it commutes with the (pointwise supremum

More information

4 Integration 4.1 Integration of non-negative simple functions

4 Integration 4.1 Integration of non-negative simple functions 4 Integration 4.1 Integration of non-negative simple functions Throughout we are in a measure space (X, F, µ). Definition Let s be a non-negative F-measurable simple function so that s a i χ Ai, with disjoint

More information

02. Measure and integral. 1. Borel-measurable functions and pointwise limits

02. Measure and integral. 1. Borel-measurable functions and pointwise limits (October 3, 2017) 02. Measure and integral Paul Garrett garrett@math.umn.edu http://www.math.umn.edu/ garrett/ [This document is http://www.math.umn.edu/ garrett/m/real/notes 2017-18/02 measure and integral.pdf]

More information

A VERY BRIEF REVIEW OF MEASURE THEORY

A VERY BRIEF REVIEW OF MEASURE THEORY A VERY BRIEF REVIEW OF MEASURE THEORY A brief philosophical discussion. Measure theory, as much as any branch of mathematics, is an area where it is important to be acquainted with the basic notions and

More information

3 Measurable Functions

3 Measurable Functions 3 Measurable Functions Notation A pair (X, F) where F is a σ-field of subsets of X is a measurable space. If µ is a measure on F then (X, F, µ) is a measure space. If µ(x) < then (X, F, µ) is a probability

More information

212a1214Daniell s integration theory.

212a1214Daniell s integration theory. 212a1214 Daniell s integration theory. October 30, 2014 Daniell s idea was to take the axiomatic properties of the integral as the starting point and develop integration for broader and broader classes

More information

Lebesgue Measure. Dung Le 1

Lebesgue Measure. Dung Le 1 Lebesgue Measure Dung Le 1 1 Introduction How do we measure the size of a set in IR? Let s start with the simplest ones: intervals. Obviously, the natural candidate for a measure of an interval is its

More information

Recall that if X is a compact metric space, C(X), the space of continuous (real-valued) functions on X, is a Banach space with the norm

Recall that if X is a compact metric space, C(X), the space of continuous (real-valued) functions on X, is a Banach space with the norm Chapter 13 Radon Measures Recall that if X is a compact metric space, C(X), the space of continuous (real-valued) functions on X, is a Banach space with the norm (13.1) f = sup x X f(x). We want to identify

More information

Notes on the Lebesgue Integral by Francis J. Narcowich Septemmber, 2014

Notes on the Lebesgue Integral by Francis J. Narcowich Septemmber, 2014 1 Introduction Notes on the Lebesgue Integral by Francis J. Narcowich Septemmber, 2014 In the definition of the Riemann integral of a function f(x), the x-axis is partitioned and the integral is defined

More information

( f ^ M _ M 0 )dµ (5.1)

( f ^ M _ M 0 )dµ (5.1) 47 5. LEBESGUE INTEGRAL: GENERAL CASE Although the Lebesgue integral defined in the previous chapter is in many ways much better behaved than the Riemann integral, it shares its restriction to bounded

More information

Lebesgue-Radon-Nikodym Theorem

Lebesgue-Radon-Nikodym Theorem Lebesgue-Radon-Nikodym Theorem Matt Rosenzweig 1 Lebesgue-Radon-Nikodym Theorem In what follows, (, A) will denote a measurable space. We begin with a review of signed measures. 1.1 Signed Measures Definition

More information

A List of Problems in Real Analysis

A List of Problems in Real Analysis A List of Problems in Real Analysis W.Yessen & T.Ma December 3, 218 This document was first created by Will Yessen, who was a graduate student at UCI. Timmy Ma, who was also a graduate student at UCI,

More information

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9 MAT 570 REAL ANALYSIS LECTURE NOTES PROFESSOR: JOHN QUIGG SEMESTER: FALL 204 Contents. Sets 2 2. Functions 5 3. Countability 7 4. Axiom of choice 8 5. Equivalence relations 9 6. Real numbers 9 7. Extended

More information

Problem Set 2: Solutions Math 201A: Fall 2016

Problem Set 2: Solutions Math 201A: Fall 2016 Problem Set 2: s Math 201A: Fall 2016 Problem 1. (a) Prove that a closed subset of a complete metric space is complete. (b) Prove that a closed subset of a compact metric space is compact. (c) Prove that

More information

2 Lebesgue integration

2 Lebesgue integration 2 Lebesgue integration 1. Let (, A, µ) be a measure space. We will always assume that µ is complete, otherwise we first take its completion. The example to have in mind is the Lebesgue measure on R n,

More information

Measures. Chapter Some prerequisites. 1.2 Introduction

Measures. Chapter Some prerequisites. 1.2 Introduction Lecture notes Course Analysis for PhD students Uppsala University, Spring 2018 Rostyslav Kozhan Chapter 1 Measures 1.1 Some prerequisites I will follow closely the textbook Real analysis: Modern Techniques

More information

MATH 426, TOPOLOGY. p 1.

MATH 426, TOPOLOGY. p 1. MATH 426, TOPOLOGY THE p-norms In this document we assume an extended real line, where is an element greater than all real numbers; the interval notation [1, ] will be used to mean [1, ) { }. 1. THE p

More information

Class Notes for Math 921/922: Real Analysis, Instructor Mikil Foss

Class Notes for Math 921/922: Real Analysis, Instructor Mikil Foss University of Nebraska - Lincoln DigitalCommons@University of Nebraska - Lincoln Math Department: Class Notes and Learning Materials Mathematics, Department of 200 Class Notes for Math 92/922: Real Analysis,

More information

MA359 Measure Theory

MA359 Measure Theory A359 easure Theory Thomas Reddington Usman Qureshi April 8, 204 Contents Real Line 3. Cantor set.................................................. 5 2 General easures 2 2. Product spaces...............................................

More information

L p Spaces and Convexity

L p Spaces and Convexity L p Spaces and Convexity These notes largely follow the treatments in Royden, Real Analysis, and Rudin, Real & Complex Analysis. 1. Convex functions Let I R be an interval. For I open, we say a function

More information

MEASURE AND INTEGRATION. Dietmar A. Salamon ETH Zürich

MEASURE AND INTEGRATION. Dietmar A. Salamon ETH Zürich MEASURE AND INTEGRATION Dietmar A. Salamon ETH Zürich 9 September 2016 ii Preface This book is based on notes for the lecture course Measure and Integration held at ETH Zürich in the spring semester 2014.

More information

Introduction to Real Analysis Alternative Chapter 1

Introduction to Real Analysis Alternative Chapter 1 Christopher Heil Introduction to Real Analysis Alternative Chapter 1 A Primer on Norms and Banach Spaces Last Updated: March 10, 2018 c 2018 by Christopher Heil Chapter 1 A Primer on Norms and Banach Spaces

More information

3. (a) What is a simple function? What is an integrable function? How is f dµ defined? Define it first

3. (a) What is a simple function? What is an integrable function? How is f dµ defined? Define it first Math 632/6321: Theory of Functions of a Real Variable Sample Preinary Exam Questions 1. Let (, M, µ) be a measure space. (a) Prove that if µ() < and if 1 p < q

More information

consists of two disjoint copies of X n, each scaled down by 1,

consists of two disjoint copies of X n, each scaled down by 1, Homework 4 Solutions, Real Analysis I, Fall, 200. (4) Let be a topological space and M be a σ-algebra on which contains all Borel sets. Let m, µ be two positive measures on M. Assume there is a constant

More information

Some topics in analysis related to Banach algebras, 2

Some topics in analysis related to Banach algebras, 2 Some topics in analysis related to Banach algebras, 2 Stephen Semmes Rice University... Abstract Contents I Preliminaries 3 1 A few basic inequalities 3 2 q-semimetrics 4 3 q-absolute value functions 7

More information

1 Topology Definition of a topology Basis (Base) of a topology The subspace topology & the product topology on X Y 3

1 Topology Definition of a topology Basis (Base) of a topology The subspace topology & the product topology on X Y 3 Index Page 1 Topology 2 1.1 Definition of a topology 2 1.2 Basis (Base) of a topology 2 1.3 The subspace topology & the product topology on X Y 3 1.4 Basic topology concepts: limit points, closed sets,

More information

+ 2x sin x. f(b i ) f(a i ) < ɛ. i=1. i=1

+ 2x sin x. f(b i ) f(a i ) < ɛ. i=1. i=1 Appendix To understand weak derivatives and distributional derivatives in the simplest context of functions of a single variable, we describe without proof some results from real analysis (see [7] and

More information

Measure Theory & Integration

Measure Theory & Integration Measure Theory & Integration Lecture Notes, Math 320/520 Fall, 2004 D.H. Sattinger Department of Mathematics Yale University Contents 1 Preliminaries 1 2 Measures 3 2.1 Area and Measure........................

More information

3 Integration and Expectation

3 Integration and Expectation 3 Integration and Expectation 3.1 Construction of the Lebesgue Integral Let (, F, µ) be a measure space (not necessarily a probability space). Our objective will be to define the Lebesgue integral R fdµ

More information

Exercise 1. Show that the Radon-Nikodym theorem for a finite measure implies the theorem for a σ-finite measure.

Exercise 1. Show that the Radon-Nikodym theorem for a finite measure implies the theorem for a σ-finite measure. Real Variables, Fall 2014 Problem set 8 Solution suggestions xercise 1. Show that the Radon-Nikodym theorem for a finite measure implies the theorem for a σ-finite measure. nswer: ssume that the Radon-Nikodym

More information

Notes on Measure, Probability and Stochastic Processes. João Lopes Dias

Notes on Measure, Probability and Stochastic Processes. João Lopes Dias Notes on Measure, Probability and Stochastic Processes João Lopes Dias Departamento de Matemática, ISEG, Universidade de Lisboa, Rua do Quelhas 6, 1200-781 Lisboa, Portugal E-mail address: jldias@iseg.ulisboa.pt

More information

ANALYSIS QUALIFYING EXAM FALL 2017: SOLUTIONS. 1 cos(nx) lim. n 2 x 2. g n (x) = 1 cos(nx) n 2 x 2. x 2.

ANALYSIS QUALIFYING EXAM FALL 2017: SOLUTIONS. 1 cos(nx) lim. n 2 x 2. g n (x) = 1 cos(nx) n 2 x 2. x 2. ANALYSIS QUALIFYING EXAM FALL 27: SOLUTIONS Problem. Determine, with justification, the it cos(nx) n 2 x 2 dx. Solution. For an integer n >, define g n : (, ) R by Also define g : (, ) R by g(x) = g n

More information

CHAPTER V DUAL SPACES

CHAPTER V DUAL SPACES CHAPTER V DUAL SPACES DEFINITION Let (X, T ) be a (real) locally convex topological vector space. By the dual space X, or (X, T ), of X we mean the set of all continuous linear functionals on X. By the

More information

Chapter IV Integration Theory

Chapter IV Integration Theory Chapter IV Integration Theory This chapter is devoted to the developement of integration theory. The main motivation is to extend the Riemann integral calculus to larger types of functions, thus leading

More information

x n x or x = T -limx n, if every open neighbourhood U of x contains x n for all but finitely many values of n

x n x or x = T -limx n, if every open neighbourhood U of x contains x n for all but finitely many values of n Note: This document was written by Prof. Richard Haydon, a previous lecturer for this course. 0. Preliminaries This preliminary chapter is an attempt to bring together important results from earlier courses

More information

Chapter 1. Measure Spaces. 1.1 Algebras and σ algebras of sets Notation and preliminaries

Chapter 1. Measure Spaces. 1.1 Algebras and σ algebras of sets Notation and preliminaries Chapter 1 Measure Spaces 1.1 Algebras and σ algebras of sets 1.1.1 Notation and preliminaries We shall denote by X a nonempty set, by P(X) the set of all parts (i.e., subsets) of X, and by the empty set.

More information

Metric Spaces and Topology

Metric Spaces and Topology Chapter 2 Metric Spaces and Topology From an engineering perspective, the most important way to construct a topology on a set is to define the topology in terms of a metric on the set. This approach underlies

More information

Construction of a general measure structure

Construction of a general measure structure Chapter 4 Construction of a general measure structure We turn to the development of general measure theory. The ingredients are a set describing the universe of points, a class of measurable subsets along

More information

Maths 212: Homework Solutions

Maths 212: Homework Solutions Maths 212: Homework Solutions 1. The definition of A ensures that x π for all x A, so π is an upper bound of A. To show it is the least upper bound, suppose x < π and consider two cases. If x < 1, then

More information

The Caratheodory Construction of Measures

The Caratheodory Construction of Measures Chapter 5 The Caratheodory Construction of Measures Recall how our construction of Lebesgue measure in Chapter 2 proceeded from an initial notion of the size of a very restricted class of subsets of R,

More information

FUNCTIONAL ANALYSIS CHRISTIAN REMLING

FUNCTIONAL ANALYSIS CHRISTIAN REMLING FUNCTIONAL ANALYSIS CHRISTIAN REMLING Contents 1. Metric and topological spaces 2 2. Banach spaces 12 3. Consequences of Baire s Theorem 30 4. Dual spaces and weak topologies 34 5. Hilbert spaces 50 6.

More information

L p Functions. Given a measure space (X, µ) and a real number p [1, ), recall that the L p -norm of a measurable function f : X R is defined by

L p Functions. Given a measure space (X, µ) and a real number p [1, ), recall that the L p -norm of a measurable function f : X R is defined by L p Functions Given a measure space (, µ) and a real number p [, ), recall that the L p -norm of a measurable function f : R is defined by f p = ( ) /p f p dµ Note that the L p -norm of a function f may

More information

G1CMIN Measure and Integration

G1CMIN Measure and Integration G1CMIN Measure and Integration 2003-4 Prof. J.K. Langley May 13, 2004 1 Introduction Books: W. Rudin, Real and Complex Analysis ; H.L. Royden, Real Analysis (QA331). Lecturer: Prof. J.K. Langley (jkl@maths,

More information

Real Analysis Math 131AH Rudin, Chapter #1. Dominique Abdi

Real Analysis Math 131AH Rudin, Chapter #1. Dominique Abdi Real Analysis Math 3AH Rudin, Chapter # Dominique Abdi.. If r is rational (r 0) and x is irrational, prove that r + x and rx are irrational. Solution. Assume the contrary, that r+x and rx are rational.

More information

6.2 Fubini s Theorem. (µ ν)(c) = f C (x) dµ(x). (6.2) Proof. Note that (X Y, A B, µ ν) must be σ-finite as well, so that.

6.2 Fubini s Theorem. (µ ν)(c) = f C (x) dµ(x). (6.2) Proof. Note that (X Y, A B, µ ν) must be σ-finite as well, so that. 6.2 Fubini s Theorem Theorem 6.2.1. (Fubini s theorem - first form) Let (, A, µ) and (, B, ν) be complete σ-finite measure spaces. Let C = A B. Then for each µ ν- measurable set C C the section x C is

More information

MAT 571 REAL ANALYSIS II LECTURE NOTES. Contents. 2. Product measures Iterated integrals Complete products Differentiation 17

MAT 571 REAL ANALYSIS II LECTURE NOTES. Contents. 2. Product measures Iterated integrals Complete products Differentiation 17 MAT 57 REAL ANALSIS II LECTURE NOTES PROFESSOR: JOHN QUIGG SEMESTER: SPRING 205 Contents. Convergence in measure 2. Product measures 3 3. Iterated integrals 4 4. Complete products 9 5. Signed measures

More information

Banach Spaces II: Elementary Banach Space Theory

Banach Spaces II: Elementary Banach Space Theory BS II c Gabriel Nagy Banach Spaces II: Elementary Banach Space Theory Notes from the Functional Analysis Course (Fall 07 - Spring 08) In this section we introduce Banach spaces and examine some of their

More information

Reminder Notes for the Course on Measures on Topological Spaces

Reminder Notes for the Course on Measures on Topological Spaces Reminder Notes for the Course on Measures on Topological Spaces T. C. Dorlas Dublin Institute for Advanced Studies School of Theoretical Physics 10 Burlington Road, Dublin 4, Ireland. Email: dorlas@stp.dias.ie

More information

Exercise 1. Let f be a nonnegative measurable function. Show that. where ϕ is taken over all simple functions with ϕ f. k 1.

Exercise 1. Let f be a nonnegative measurable function. Show that. where ϕ is taken over all simple functions with ϕ f. k 1. Real Variables, Fall 2014 Problem set 3 Solution suggestions xercise 1. Let f be a nonnegative measurable function. Show that f = sup ϕ, where ϕ is taken over all simple functions with ϕ f. For each n

More information

CHAPTER VIII HILBERT SPACES

CHAPTER VIII HILBERT SPACES CHAPTER VIII HILBERT SPACES DEFINITION Let X and Y be two complex vector spaces. A map T : X Y is called a conjugate-linear transformation if it is a reallinear transformation from X into Y, and if T (λx)

More information

THEOREMS, ETC., FOR MATH 516

THEOREMS, ETC., FOR MATH 516 THEOREMS, ETC., FOR MATH 516 Results labeled Theorem Ea.b.c (or Proposition Ea.b.c, etc.) refer to Theorem c from section a.b of Evans book (Partial Differential Equations). Proposition 1 (=Proposition

More information

+ = x. ± if x > 0. if x < 0, x

+ = x. ± if x > 0. if x < 0, x 2 Set Functions Notation Let R R {, + } and R + {x R : x 0} {+ }. Here + and are symbols satisfying obvious conditions: for any real number x R : < x < +, (± + (± x + (± (± + x ±, (± (± + and (± (, ± if

More information

Notions such as convergent sequence and Cauchy sequence make sense for any metric space. Convergent Sequences are Cauchy

Notions such as convergent sequence and Cauchy sequence make sense for any metric space. Convergent Sequences are Cauchy Banach Spaces These notes provide an introduction to Banach spaces, which are complete normed vector spaces. For the purposes of these notes, all vector spaces are assumed to be over the real numbers.

More information

Lebesgue Measure on R n

Lebesgue Measure on R n CHAPTER 2 Lebesgue Measure on R n Our goal is to construct a notion of the volume, or Lebesgue measure, of rather general subsets of R n that reduces to the usual volume of elementary geometrical sets

More information

A note on some approximation theorems in measure theory

A note on some approximation theorems in measure theory A note on some approximation theorems in measure theory S. Kesavan and M. T. Nair Department of Mathematics, Indian Institute of Technology, Madras, Chennai - 600 06 email: kesh@iitm.ac.in and mtnair@iitm.ac.in

More information

Measurable functions are approximately nice, even if look terrible.

Measurable functions are approximately nice, even if look terrible. Tel Aviv University, 2015 Functions of real variables 74 7 Approximation 7a A terrible integrable function........... 74 7b Approximation of sets................ 76 7c Approximation of functions............

More information

Real Analysis Chapter 3 Solutions Jonathan Conder. ν(f n ) = lim

Real Analysis Chapter 3 Solutions Jonathan Conder. ν(f n ) = lim . Suppose ( n ) n is an increasing sequence in M. For each n N define F n : n \ n (with 0 : ). Clearly ν( n n ) ν( nf n ) ν(f n ) lim n If ( n ) n is a decreasing sequence in M and ν( )

More information

Lebesgue Integration on R n

Lebesgue Integration on R n Lebesgue Integration on R n The treatment here is based loosely on that of Jones, Lebesgue Integration on Euclidean Space We give an overview from the perspective of a user of the theory Riemann integration

More information

Three hours THE UNIVERSITY OF MANCHESTER. 24th January

Three hours THE UNIVERSITY OF MANCHESTER. 24th January Three hours MATH41011 THE UNIVERSITY OF MANCHESTER FOURIER ANALYSIS AND LEBESGUE INTEGRATION 24th January 2013 9.45 12.45 Answer ALL SIX questions in Section A (25 marks in total). Answer THREE of the

More information

Compendium and Solutions to exercises TMA4225 Foundation of analysis

Compendium and Solutions to exercises TMA4225 Foundation of analysis Compendium and Solutions to exercises TMA4225 Foundation of analysis Ruben Spaans December 6, 2010 1 Introduction This compendium contains a lexicon over definitions and exercises with solutions. Throughout

More information