MathIsimple
π
θ
Δ
ε
RA-2
7-10 hours
Advanced

Measurable Functions & Approximations

Explore the theory of measurable functions, simple function approximations, Egorov's theorem on almost uniform convergence, and Lusin's theorem on almost continuity. These concepts form the foundation for Lebesgue integration.

Learning Objectives
By the end of this course, you will be able to:
1
Define measurable functions and understand their fundamental properties
2
Construct simple function approximations to measurable functions
3
Apply Egorov's theorem to understand the relationship between pointwise and uniform convergence
4
Use Lusin's theorem to approximate measurable functions by continuous functions
5
Understand almost-everywhere convergence and its implications
6
Prove that limits of measurable functions are measurable

Prerequisites

Before starting this course, you should be familiar with:

  • Measure foundations: outer measure, measurable sets, σ-algebras
  • Basic properties of functions: domain, range, composition
  • Pointwise and uniform convergence of sequences of functions
  • Continuous functions and their properties

Core Concepts

Fundamental definitions and properties

Definition 2.1: Measurable Function

Let (X,M)(X, \mathcal{M}) be a measurable space. A function f:XRf: X \to \mathbb{R} is called measurable if for every aRa \in \mathbb{R}, the set

f1((a,))={xX:f(x)>a}f^{-1}((a, \infty)) = \{x \in X : f(x) > a\}

is measurable (i.e., belongs to M\mathcal{M}).

Note: This definition is equivalent to requiring that f1(B)f^{-1}(B) is measurable for all Borel sets BB, or for all open sets BB.

Definition 2.2: Simple Function

A function φ:XR\varphi: X \to \mathbb{R} is called a simple function if it takes only finitely many distinct values. That is, there exist distinct real numbers c1,c2,,cnc_1, c_2, \ldots, c_n and measurable sets E1,E2,,EnE_1, E_2, \ldots, E_n such that

φ(x)=i=1nciχEi(x)\varphi(x) = \sum_{i=1}^n c_i \chi_{E_i}(x)

where χEi\chi_{E_i} is the characteristic function of EiE_i.

Definition 2.3: Almost Everywhere

A property is said to hold almost everywhere (abbreviated a.e.) if it holds except possibly on a set of measure zero.

Two functions ff and gg are said to be equal almost everywhere if {x:f(x)g(x)}\{x : f(x) \neq g(x)\} is a null set.

Theorem 2.1: Properties of Measurable Functions

Let ff and gg be measurable functions. Then:

  1. f+gf + g is measurable
  2. fgfg is measurable
  3. max(f,g)\max(f, g) and min(f,g)\min(f, g) are measurable
  4. f|f| is measurable
  5. If g0g \neq 0 a.e., then f/gf/g is measurable
Proof of Theorem 2.1:

Proof:

(1) For any aa, we have

{x:(f+g)(x)>a}=rQ({x:f(x)>r}{x:g(x)>ar})\{x : (f+g)(x) > a\} = \bigcup_{r \in \mathbb{Q}} (\{x : f(x) > r\} \cap \{x : g(x) > a - r\})

Since ff and gg are measurable, each set in the union is measurable, and a countable union of measurable sets is measurable.

(2) Note that fg=14[(f+g)2(fg)2]fg = \frac{1}{4}[(f+g)^2 - (f-g)^2]. Since squares and differences of measurable functions are measurable, the result follows.

(3) max(f,g)=f+g+fg2\max(f, g) = \frac{f+g+|f-g|}{2} and min(f,g)=f+gfg2\min(f, g) = \frac{f+g-|f-g|}{2}.

(4) f=max(f,f)|f| = \max(f, -f).

(5) {x:(f/g)(x)>a}={x:f(x)>ag(x)}\{x : (f/g)(x) > a\} = \{x : f(x) > ag(x)\}, which can be written as a countable union similar to (1). ∎

Theorem 2.2: Pointwise Limits of Measurable Functions

If {fn}\{f_n\} is a sequence of measurable functions and fnff_n \to f pointwise (or pointwise a.e.), then ff is measurable.

Proof of Theorem 2.2:

Proof:

For any aa, we have

{x:f(x)>a}=k=1N=1n=N{x:fn(x)>a+1k}\{x : f(x) > a\} = \bigcup_{k=1}^{\infty} \bigcap_{N=1}^{\infty} \bigcup_{n=N}^{\infty} \left\{x : f_n(x) > a + \frac{1}{k}\right\}

This expresses the limit superior condition: f(x)>af(x) > a if and only if there exists kk such that for all NN, there exists nNn \geq N with fn(x)>a+1/kf_n(x) > a + 1/k.

Since each {x:fn(x)>a+1/k}\{x : f_n(x) > a + 1/k\} is measurable, and measurable sets are closed under countable unions and intersections, the result follows. ∎

Theorem 2.3: Simple Function Approximation

Let ff be a nonnegative measurable function. Then there exists an increasing sequence {φn}\{\varphi_n\} of simple functions such that φnf\varphi_n \to f pointwise.

Proof of Theorem 2.3:

Proof:

For each nn, define

φn(x)=k=0n2n1k2nχEn,k(x)+nχFn(x)\varphi_n(x) = \sum_{k=0}^{n2^n - 1} \frac{k}{2^n} \chi_{E_{n,k}}(x) + n \chi_{F_n}(x)

where

En,k={x:k2nf(x)<k+12n},Fn={x:f(x)n}E_{n,k} = \left\{x : \frac{k}{2^n} \leq f(x) < \frac{k+1}{2^n}\right\}, \quad F_n = \{x : f(x) \geq n\}

Each φn\varphi_n is a simple function, the sequence is increasing, and φnf\varphi_n \to f pointwise. ∎

Theorem 2.4: Egorov's Theorem

Let (X,M,μ)(X, \mathcal{M}, \mu) be a measure space with μ(X)<\mu(X) < \infty. Suppose {fn}\{f_n\} is a sequence of measurable functions and fnff_n \to f pointwise a.e. Then for every ϵ>0\epsilon > 0, there exists a measurable set EE such that μ(E)<ϵ\mu(E) < \epsilon and fnff_n \to f uniformly on XEX \setminus E.

Proof of Theorem 2.4:

Proof:

For n,kNn, k \in \mathbb{N}, define

En,k=m=n{x:fm(x)f(x)1k}E_{n,k} = \bigcup_{m=n}^{\infty} \left\{x : |f_m(x) - f(x)| \geq \frac{1}{k}\right\}

Since fnff_n \to f pointwise a.e., for each kk, we have μ(n=1En,k)=0\mu(\bigcap_{n=1}^{\infty} E_{n,k}) = 0.

By continuity of measure, for each kk, there exists nkn_k such that μ(Enk,k)<ϵ/2k\mu(E_{n_k, k}) < \epsilon/2^k.

Let E=k=1Enk,kE = \bigcup_{k=1}^{\infty} E_{n_k, k}. Then μ(E)<ϵ\mu(E) < \epsilon, and on XEX \setminus E, we have uniform convergence. ∎

Theorem 2.5: Lusin's Theorem

Let ff be a measurable function on Rn\mathbb{R}^n that is finite a.e. on a measurable set EE of finite measure. Then for every ϵ>0\epsilon > 0, there exists a closed set FF with FEF \subseteq E, m(EF)<ϵm(E \setminus F) < \epsilon, and ff is continuous on FF.

Proof of Theorem 2.5:

Proof:

First approximate ff by a simple function φ\varphi such that fφ<ϵ|f - \varphi| < \epsilon except on a set of small measure.

Then approximate each characteristic function in φ\varphi by continuous functions using regularity of measure.

Finally, use Tietze's extension theorem to extend the continuous function to the closed set. The details involve careful approximation and measure theory techniques. ∎

Example 2.1: Characteristic Function is Measurable

Problem: Show that the characteristic function χE\chi_E of a measurable set EE is measurable.

Solution:

For any aRa \in \mathbb{R}, we have

χE1((a,))={Eif a<1if a1\chi_E^{-1}((a, \infty)) = \begin{cases} E & \text{if } a < 1 \\ \emptyset & \text{if } a \geq 1 \end{cases}

Since EE and \emptyset are measurable, χE\chi_E is measurable.

Example 2.2: Constructing Simple Function Approximation

Problem: Construct a simple function approximation for f(x)=x2f(x) = x^2 on [0,4][0, 4].

Solution:

Using the construction in Theorem 2.3 with n=2n = 2, we partition [0,4][0, 4] into intervals of length 1/4=0.251/4 = 0.25:

φ2(x)=k=015k4χ[k/4,(k+1)/4)(x)+2χ[2,4](x)\varphi_2(x) = \sum_{k=0}^{15} \frac{k}{4} \chi_{[k/4, (k+1)/4)}(x) + 2 \chi_{[2, 4]}(x)

This gives a step function approximation that increases as nn increases, converging to f(x)=x2f(x) = x^2.

Corollary 2.1: Measurable Functions are Closed under Algebraic Operations

The collection of measurable functions forms a vector space and an algebra over R\mathbb{R}.

Theorem 2.6: Composition of Measurable Functions

Let f:XRf: X \to \mathbb{R} be measurable and g:RRg: \mathbb{R} \to \mathbb{R} be Borel measurable (i.e., g1(B)g^{-1}(B) is Borel for all Borel sets BB). Then the composition gfg \circ f is measurable.

In particular, if gg is continuous, then gfg \circ f is measurable.

Proof of Theorem 2.6:

Proof:

For any Borel set BB, we have

(gf)1(B)=f1(g1(B))(g \circ f)^{-1}(B) = f^{-1}(g^{-1}(B))

Since gg is Borel measurable, g1(B)g^{-1}(B) is a Borel set. Since ff is measurable, f1(g1(B))f^{-1}(g^{-1}(B)) is measurable.

Therefore, (gf)1(B)(g \circ f)^{-1}(B) is measurable for all Borel sets BB, which implies gfg \circ f is measurable.

If gg is continuous, then gg is Borel measurable (since continuous functions map Borel sets to Borel sets), so the result follows. ∎

Theorem 2.7: Detailed Construction of Simple Function Approximation

Let ff be a nonnegative measurable function. For each nNn \in \mathbb{N}, define the simple function φn\varphi_n by:

φn(x)={k2nif k2nf(x)<k+12n,k=0,1,,n2n1nif f(x)n\varphi_n(x) = \begin{cases} \frac{k}{2^n} & \text{if } \frac{k}{2^n} \leq f(x) < \frac{k+1}{2^n}, \, k = 0, 1, \ldots, n2^n - 1 \\ n & \text{if } f(x) \geq n \end{cases}

Then {φn}\{\varphi_n\} is an increasing sequence of simple functions converging pointwise to ff, and the convergence is uniform on any set where ff is bounded.

Proof of Theorem 2.7:

Proof:

For each nn, the function φn\varphi_n takes only finitely many values (at most n2n+1n2^n + 1 values), so it is simple.

To show the sequence is increasing, fix xx and nn. If f(x)nf(x) \geq n, then φn(x)=n\varphi_n(x) = n and φn+1(x)n\varphi_{n+1}(x) \geq n, so φn+1(x)φn(x)\varphi_{n+1}(x) \geq \varphi_n(x).

If f(x)<nf(x) < n, then φn(x)=k2n\varphi_n(x) = \frac{k}{2^n} for some kk. Since the partition at step n+1n+1 is finer, we have φn+1(x)φn(x)\varphi_{n+1}(x) \geq \varphi_n(x).

For pointwise convergence, fix xx. If f(x)=f(x) = \infty, then φn(x)=n\varphi_n(x) = n \to \infty. If f(x)<f(x) < \infty, then for large nn, we have f(x)<nf(x) < n, and

0f(x)φn(x)<12n0 \leq f(x) - \varphi_n(x) < \frac{1}{2^n}

so φn(x)f(x)\varphi_n(x) \to f(x).

If ff is bounded by MM, then for n>Mn > M, the approximation error is at most 12n\frac{1}{2^n} uniformly, giving uniform convergence. ∎

Theorem 2.8: Supremum and Infimum of Measurable Functions

Let {fn}\{f_n\} be a sequence of measurable functions. Then:

  1. supnfn\sup_n f_n is measurable
  2. infnfn\inf_n f_n is measurable
  3. lim supnfn\limsup_{n \to \infty} f_n is measurable
  4. lim infnfn\liminf_{n \to \infty} f_n is measurable
Proof of Theorem 2.8:

Proof:

(1) For any aa, we have

{x:supnfn(x)>a}=n=1{x:fn(x)>a}\{x : \sup_n f_n(x) > a\} = \bigcup_{n=1}^{\infty} \{x : f_n(x) > a\}

Since each {x:fn(x)>a}\{x : f_n(x) > a\} is measurable, the countable union is measurable.

(2) Similarly, infnfn=supn(fn)\inf_n f_n = -\sup_n (-f_n), so it is measurable.

(3) Note that lim supnfn=infn1supknfk\limsup_{n \to \infty} f_n = \inf_{n \geq 1} \sup_{k \geq n} f_k. By (1) and (2), this is measurable.

(4) Similarly, lim infnfn=supn1infknfk\liminf_{n \to \infty} f_n = \sup_{n \geq 1} \inf_{k \geq n} f_k is measurable. ∎

Example 2.3: Application of Egorov's Theorem: Constructing a Counterexample

Problem: Show that the condition μ(X)<\mu(X) < \infty in Egorov's theorem is necessary by constructing a counterexample on an infinite measure space.

Solution:

Let X=RX = \mathbb{R} with Lebesgue measure, and define

fn(x)=χ[n,n+1](x)={1if x[n,n+1]0otherwisef_n(x) = \chi_{[n, n+1]}(x) = \begin{cases} 1 & \text{if } x \in [n, n+1] \\ 0 & \text{otherwise} \end{cases}

Then fn0f_n \to 0 pointwise everywhere, since for any fixed xx, eventually fn(x)=0f_n(x) = 0.

However, for any set EE with m(E)<m(E) < \infty, the convergence is not uniform on RE\mathbb{R} \setminus E. In fact, for any ϵ>0\epsilon > 0, no matter how large the exceptional set EE is (as long as it has finite measure), there will be points in RE\mathbb{R} \setminus E where fn=1f_n = 1 for arbitrarily large nn.

This shows that Egorov's theorem fails when the measure space is infinite, demonstrating the necessity of the finite measure condition.

Example 2.4: Application of Lusin's Theorem: A Specific Function

Problem: Consider the function f(x)=χQ(x)f(x) = \chi_{\mathbb{Q}}(x) on [0,1][0,1] (the characteristic function of the rationals). Apply Lusin's theorem to approximate this function by a continuous function.

Solution:

The function ff is measurable (since Q\mathbb{Q} is countable and hence measurable) but is discontinuous everywhere.

By Lusin's theorem, for any ϵ>0\epsilon > 0, there exists a closed set FF with m([0,1]F)<ϵm([0,1] \setminus F) < \epsilon such that ff is continuous on FF.

Since Q[0,1]\mathbb{Q} \cap [0,1] has measure zero, we can take F=[0,1](Q[0,1])F = [0,1] \setminus (\mathbb{Q} \cap [0,1]) (the irrationals in [0,1][0,1]). However, this set is not closed. Instead, we can remove a small open neighborhood around each rational point.

More precisely, enumerate Q[0,1]={q1,q2,}\mathbb{Q} \cap [0,1] = \{q_1, q_2, \ldots\}. For each qiq_i, remove an open interval (qiδi,qi+δi)(q_i - \delta_i, q_i + \delta_i) where δi<ϵ\sum \delta_i < \epsilon. The complement is closed, and on this closed set, ff is identically 0, hence continuous.

This demonstrates that even a function that is discontinuous everywhere can be made continuous on a set of nearly full measure.

Example 2.5: Step-by-Step Construction of Simple Function Approximation

Problem: Construct the first three approximations φ1,φ2,φ3\varphi_1, \varphi_2, \varphi_3 for the function f(x)=x2f(x) = x^2 on [0,4][0, 4] using the method in Theorem 2.7.

Solution:

Step 1: n=1n = 1

We partition [0,1)[0, 1) into intervals of length 12\frac{1}{2}: [0,12),[12,1)[0, \frac{1}{2}), [\frac{1}{2}, 1). For x[0,1)x \in [0, 1), we have f(x)=x2<1f(x) = x^2 < 1, so:

φ1(x)={0if 0x2<12, i.e., 0x<1212if 12x2<1, i.e., 12x<11if x21, i.e., x1\varphi_1(x) = \begin{cases} 0 & \text{if } 0 \leq x^2 < \frac{1}{2}, \text{ i.e., } 0 \leq x < \frac{1}{\sqrt{2}} \\ \frac{1}{2} & \text{if } \frac{1}{2} \leq x^2 < 1, \text{ i.e., } \frac{1}{\sqrt{2}} \leq x < 1 \\ 1 & \text{if } x^2 \geq 1, \text{ i.e., } x \geq 1 \end{cases}

Step 2: n=2n = 2

We partition [0,2)[0, 2) into intervals of length 14\frac{1}{4}: [0,14),[14,12),,[74,2)[0, \frac{1}{4}), [\frac{1}{4}, \frac{1}{2}), \ldots, [\frac{7}{4}, 2). This gives a finer approximation:

φ2(x)={k4if k4x2<k+14,k=0,1,,72if x22, i.e., x2\varphi_2(x) = \begin{cases} \frac{k}{4} & \text{if } \frac{k}{4} \leq x^2 < \frac{k+1}{4}, \, k = 0, 1, \ldots, 7 \\ 2 & \text{if } x^2 \geq 2, \text{ i.e., } x \geq \sqrt{2} \end{cases}

Step 3: n=3n = 3

We partition [0,3)[0, 3) into intervals of length 18\frac{1}{8}, giving an even finer approximation with 24 steps in [0,3)[0, 3) and the constant value 3 for x3x \geq \sqrt{3}.

As nn increases, φn\varphi_n becomes a better approximation to f(x)=x2f(x) = x^2, with the error bounded by 12n\frac{1}{2^n} on [0,n)[0, n).

Example 2.6: A Measurable Function That Is Not Continuous

Problem: Construct a measurable function on [0,1][0,1] that is not continuous at any point.

Solution:

Consider the function f(x)=χQ[0,1](x)f(x) = \chi_{\mathbb{Q} \cap [0,1]}(x), the characteristic function of the rationals in [0,1][0,1].

This function is measurable because Q[0,1]\mathbb{Q} \cap [0,1] is countable and hence measurable.

However, ff is discontinuous at every point. To see this, fix any x0[0,1]x_0 \in [0,1]. If x0Qx_0 \in \mathbb{Q}, then f(x0)=1f(x_0) = 1, but in any neighborhood of x0x_0, there are irrationals where ff takes the value 0. If x0Qx_0 \notin \mathbb{Q}, then f(x0)=0f(x_0) = 0, but in any neighborhood of x0x_0, there are rationals where ff takes the value 1.

This example shows that measurability does not imply continuity, even though Lusin's theorem tells us that measurable functions are "almost continuous" in a measure-theoretic sense.

Example 2.7: Pointwise Convergence That Is Not Uniform

Problem: Give an example of a sequence of measurable functions that converges pointwise almost everywhere but not uniformly, even on sets of finite measure.

Solution:

On [0,1][0,1] with Lebesgue measure, define

fn(x)={nif x(0,1n]0otherwisef_n(x) = \begin{cases} n & \text{if } x \in (0, \frac{1}{n}] \\ 0 & \text{otherwise} \end{cases}

Then fn0f_n \to 0 pointwise almost everywhere. Indeed, for any x>0x > 0, we have fn(x)=0f_n(x) = 0 for all n>1/xn > 1/x. The only point where convergence fails is x=0x = 0, which is a null set.

However, the convergence is not uniform on [0,1][0,1]. For any nn, we have fn(12n)=nf_n(\frac{1}{2n}) = n, so supx[0,1]fn(x)0=n\sup_{x \in [0,1]} |f_n(x) - 0| = n, which does not tend to 0.

This example illustrates why Egorov's theorem is remarkable: it tells us that even though uniform convergence may fail on the whole set, we can always find a large subset (all but a set of small measure) where uniform convergence holds.

Corollary 2.2: Limits of Measurable Functions

If {fn}\{f_n\} is a sequence of measurable functions and fnff_n \to f pointwise (or pointwise a.e.), then ff is measurable.

This follows immediately from Theorem 2.8, since f=lim supfn=lim inffnf = \limsup f_n = \liminf f_n when the limit exists.

Corollary 2.3: Equivalence of Measurable and Borel Functions

A function ff is measurable if and only if f1(B)f^{-1}(B) is measurable for all Borel sets BB.

This is because the Borel σ-algebra is generated by intervals of the form (a,)(a, \infty), and measurability can be checked on a generating family.

Remark 2.2: Historical Context of Egorov and Lusin Theorems

Egorov's theorem (1911) and Lusin's theorem (1912) were fundamental results in the early development of measure theory and integration theory.

Egorov's theorem was proved by Dmitri Egorov, a Russian mathematician who was a student of Lebesgue. It bridges the gap between pointwise convergence (which is natural but weak) and uniform convergence (which is strong but often too restrictive). The theorem shows that on finite measure spaces, these two notions are "almost" equivalent.

Lusin's theorem was proved by Nikolai Luzin (also spelled Lusin), another Russian mathematician. It demonstrates a remarkable fact: measurable functions, which can be very irregular, are "almost continuous" in a precise measure-theoretic sense. This result connects measure theory with topology and is crucial for many applications.

Together, these theorems form part of "Littlewood's three principles," which summarize the approximation properties in measure theory: measurable sets are nearly open, measurable functions are nearly continuous, and pointwise convergence is nearly uniform.

Remark 2.3: Connection to Topology

There is a deep connection between measurability and topology:

  • Continuous functions are measurable: Since continuous functions map open sets to open sets, and open sets are Borel, continuous functions are Borel measurable and hence measurable.
  • Lusin's theorem: Shows that measurable functions can be approximated by continuous functions, establishing a bridge between measure theory and topology.
  • Borel σ-algebra: The smallest σ-algebra containing all open sets. All Borel sets are measurable, but there exist non-Borel measurable sets (requiring the Axiom of Choice to construct).
  • Regularity: The regularity properties of measure (approximation by open/closed sets) connect measure theory with the topology of the underlying space.

These connections are essential for understanding how measure theory generalizes and extends classical analysis.

Remark 2.4: Applications in Probability Theory

Measurable functions play a central role in probability theory:

  • Random variables: In probability theory, random variables are defined as measurable functions from a probability space to R\mathbb{R}. The measurability condition ensures that events like {X>a}\{X > a\} have well-defined probabilities.
  • Almost sure convergence: The concept of "almost everywhere" in measure theory corresponds to "almost surely" in probability theory. Egorov's theorem has probabilistic analogs.
  • Simple function approximation: The approximation of measurable functions by simple functions is crucial for defining the expectation (integral) of random variables.
  • Convergence in measure: Various modes of convergence (pointwise, uniform, in measure) have probabilistic interpretations (almost sure convergence, convergence in probability, etc.).

The theory of measurable functions provides the rigorous foundation for modern probability theory and stochastic processes.

Example 2.8: Constructing Measurable Functions from Simple Functions

Problem: Show that any simple function can be written as a linear combination of characteristic functions of disjoint measurable sets.

Solution:

Let φ=i=1nciχEi\varphi = \sum_{i=1}^n c_i \chi_{E_i} be a simple function. If the sets EiE_i are not disjoint, we can refine them.

For example, if n=2n = 2 and E1E_1 and E2E_2 overlap, define:

F1=E1E2,F2=E1E2,F3=E2E1F_1 = E_1 \setminus E_2, \quad F_2 = E_1 \cap E_2, \quad F_3 = E_2 \setminus E_1

Then φ=c1χF1+(c1+c2)χF2+c2χF3\varphi = c_1 \chi_{F_1} + (c_1 + c_2) \chi_{F_2} + c_2 \chi_{F_3}, where the FiF_i are disjoint.

This canonical form is useful for defining the integral of simple functions.

Example 2.9: Egorov's Theorem: Detailed Application

Problem: Let fn(x)=xnf_n(x) = x^n on [0,1][0,1]. Apply Egorov's theorem to show that the convergence fn0f_n \to 0 (pointwise on [0,1)[0,1)) can be made uniform on a large subset.

Solution:

We have fn(x)0f_n(x) \to 0 for x[0,1)x \in [0, 1) and fn(1)=1f_n(1) = 1 for all nn.

For any ϵ>0\epsilon > 0, Egorov's theorem guarantees a set EE with m(E)<ϵm(E) < \epsilon such that fn0f_n \to 0 uniformly on [0,1]E[0,1] \setminus E.

In this case, we can take E=[1δ,1]E = [1-\delta, 1] for small δ\delta. On [0,1δ][0, 1-\delta], we have fn(x)=xn(1δ)n|f_n(x)| = x^n \leq (1-\delta)^n, which tends to 0 uniformly.

This demonstrates how Egorov's theorem works: we exclude a small neighborhood of the problematic point (where convergence fails) to get uniform convergence on the remainder.

Example 2.10: Lusin's Theorem: Step Function Example

Problem: Consider the step function f(x)=i=1nciχ[ai,bi)(x)f(x) = \sum_{i=1}^n c_i \chi_{[a_i, b_i)}(x) on [0,1][0,1]. Show how Lusin's theorem applies.

Solution:

The function ff is measurable (as a simple function) but has jump discontinuities at the points aia_i and bib_i.

By Lusin's theorem, for any ϵ>0\epsilon > 0, there exists a closed set FF with m([0,1]F)<ϵm([0,1] \setminus F) < \epsilon such that ff is continuous on FF.

In this case, we can take FF to be [0,1][0,1] minus small open neighborhoods around each discontinuity point. On the resulting closed set, ff is constant on each connected component, hence continuous.

This shows that even functions with many discontinuities can be made continuous on a set of nearly full measure.

Lemma 2.1: Approximation by Continuous Functions

Let ff be a measurable function on Rn\mathbb{R}^n that is finite a.e. on a measurable set EE of finite measure.

Then for any ϵ>0\epsilon > 0, there exists a continuous function gg on Rn\mathbb{R}^n such that m({xE:f(x)g(x)})<ϵm(\{x \in E : f(x) \neq g(x)\}) < \epsilon.

This is a consequence of Lusin's theorem combined with Tietze's extension theorem.

Proof of Lemma 2.1:

Proof:

By Lusin's theorem, there exists a closed set FF with FEF \subseteq E, m(EF)<ϵ/2m(E \setminus F) < \epsilon/2, and ff is continuous on FF.

By Tietze's extension theorem, the continuous function fFf|_F can be extended to a continuous function gg on all of Rn\mathbb{R}^n.

Then m({xE:f(x)g(x)})m(EF)<ϵm(\{x \in E : f(x) \neq g(x)\}) \leq m(E \setminus F) < \epsilon. ∎

Theorem 2.9: Density of Simple Functions in Lp

For 1p<1 \leq p < \infty, the simple functions are dense in LpL^p. That is, for any fLpf \in L^p and ϵ>0\epsilon > 0, there exists a simple function φ\varphi such that fφp<ϵ\|f - \varphi\|_p < \epsilon.

Proof of Theorem 2.9:

Proof:

First approximate ff by a bounded function fN=fχ{fN}f_N = f \chi_{\{|f| \leq N\}}. For large NN, ffNp<ϵ/2\|f - f_N\|_p < \epsilon/2.

Then approximate fNf_N by a simple function using the construction in Theorem 2.7. The approximation can be made arbitrarily close in LpL^p norm. ∎

Corollary 2.4: Density of Continuous Functions

For 1p<1 \leq p < \infty, continuous functions with compact support are dense in Lp(Rn)L^p(\mathbb{R}^n).

This follows from the density of simple functions and Lemma 2.1, which allows approximation of simple functions by continuous functions.

Example 2.11: Measurable Function That Is Not Borel Measurable

Problem: Show that there exist measurable functions that are not Borel measurable.

Solution:

Let CC be the Cantor set and VV be a non-Borel measurable subset of CC (such sets exist by the Axiom of Choice).

Define f=χVf = \chi_V. Since VV is measurable (as a subset of a null set), ff is measurable.

However, f1({1})=Vf^{-1}(\{1\}) = V is not Borel, so ff is not Borel measurable.

This shows that the class of measurable functions is strictly larger than the class of Borel measurable functions.

Remark 2.5: Convergence in Measure

In addition to pointwise and uniform convergence, there is another important mode of convergence:

Convergence in measure: A sequence {fn}\{f_n\} converges in measure to ff if for every ϵ>0\epsilon > 0,

limnm({x:fn(x)f(x)ϵ})=0\lim_{n \to \infty} m(\{x : |f_n(x) - f(x)| \geq \epsilon\}) = 0

Convergence in measure is weaker than pointwise convergence a.e. but stronger than convergence in Lp. It plays an important role in probability theory (where it corresponds to convergence in probability).

Egorov's theorem shows that on finite measure spaces, pointwise convergence a.e. implies convergence in measure, and the converse holds along a subsequence.

Remark 2.6: Measurable Functions and Integration

The theory of measurable functions is essential for Lebesgue integration:

  • Definition of integral: The Lebesgue integral is first defined for simple functions, then extended to nonnegative measurable functions via approximation, and finally to general measurable functions.
  • Convergence theorems: The fact that limits of measurable functions are measurable ensures that convergence theorems (monotone, dominated) can be applied.
  • Lp spaces: The space Lp consists of measurable functions (modulo functions that are zero a.e.) with finite Lp norm.
  • Almost everywhere equality: Two functions that are equal a.e. have the same integral, which is why we work with equivalence classes in Lp spaces.

Without the theory of measurable functions, we could not have a satisfactory theory of integration for the wide class of functions that Lebesgue integration handles.

Remark 2.1: Key Insights

Key takeaways:

  • Measurable functions generalize continuous functions and are closed under limits
  • Simple functions provide a bridge between measure theory and integration
  • Egorov's theorem shows that pointwise convergence can be made uniform on large subsets
  • Lusin's theorem demonstrates that measurable functions are "almost continuous"
  • Almost-everywhere properties are sufficient for most purposes in analysis

Practice Quiz

Measurable Functions & Approximations
10
Questions
0
Correct
0%
Accuracy
1
A function f:RnRf: \mathbb{R}^n \to \mathbb{R} is measurable if and only if:
Easy
Not attempted
2
A simple function is:
Easy
Not attempted
3
Egorov's theorem states that:
Medium
Not attempted
4
Lusin's theorem states that:
Medium
Not attempted
5
If fnff_n \to f pointwise a.e. and each fnf_n is measurable, then:
Medium
Not attempted
6
A function ff is measurable if f1(B)f^{-1}(B) is measurable for all:
Hard
Not attempted
7
The simple function approximation theorem states that:
Medium
Not attempted
8
Almost everywhere (a.e.) means:
Easy
Not attempted
9
If ff and gg are measurable functions, then which of the following is NOT necessarily measurable?
Hard
Not attempted
10
Littlewood's three principles state that:
Hard
Not attempted

Frequently Asked Questions

Why do we need measurable functions instead of just continuous functions?

Measurable functions form a much larger class than continuous functions and are closed under pointwise limits. Many important functions in analysis (like step functions, indicator functions, and limits of sequences) are measurable but not continuous. The theory of Lebesgue integration requires measurable functions.

What is the relationship between measurable functions and simple functions?

Simple functions are a special class of measurable functions that take only finitely many values. They serve as building blocks: any nonnegative measurable function can be approximated by an increasing sequence of simple functions, which is crucial for defining the Lebesgue integral.

How does Egorov's theorem relate to uniform convergence?

Egorov's theorem shows that on a set of finite measure, pointwise convergence almost everywhere can be 'upgraded' to uniform convergence on a large subset (all but a set of arbitrarily small measure). This bridges the gap between pointwise and uniform convergence in measure theory.

What does Lusin's theorem tell us about measurable functions?

Lusin's theorem shows that measurable functions are 'almost continuous' in a precise sense: given any measurable function on a set of finite measure, we can find a subset of nearly full measure where the function is continuous. This is remarkable since measurable functions can be very irregular.

Are all continuous functions measurable?

Yes, all continuous functions are measurable. This follows because the preimage of an open set under a continuous function is open, and open sets are measurable. However, not all measurable functions are continuous.

What does 'almost everywhere' mean in practice?

A property holds almost everywhere if it fails only on a set of measure zero. In practice, this means the property holds 'almost always' and exceptions are negligible from a measure-theoretic perspective. For example, two functions that differ only on a countable set are equal almost everywhere.

Why is the measurability of limits important?

The fact that pointwise limits of measurable functions are measurable ensures that the class of measurable functions is closed under limit operations. This is essential for integration theory, as we often need to take limits of sequences of functions.